04.01.2013 Views

Abstracts Poster Abstracts - Dr. Falk Pharma GmbH

Abstracts Poster Abstracts - Dr. Falk Pharma GmbH

Abstracts Poster Abstracts - Dr. Falk Pharma GmbH

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Falk</strong> Workshop<br />

Immunology and<br />

Liver Disease<br />

October 15 – 16, 2009<br />

Maritim Airport Hotel<br />

Hannover<br />

<strong>Abstracts</strong><br />

<strong>Poster</strong> <strong>Poster</strong> <strong>Abstracts</strong>


FALK FOUNDATION e.V.<br />

www.falkfoundation.org<br />

Leinenweberstr. 5<br />

79108 Freiburg<br />

Germany<br />

©2009 <strong>Falk</strong> Foundation e.V.<br />

All rights reserved.


<strong>Abstracts</strong> of Invited Lectures<br />

<strong>Poster</strong> <strong>Abstracts</strong><br />

<strong>Falk</strong> Workshop<br />

IMMUNOLOGY AND LIVER DISEASE<br />

Hannover (Germany)<br />

October 15 – 16, 2009<br />

Scientific Organization:<br />

M.E. Gershwin, Davis (USA)<br />

A.W. Lohse, Hamburg (Germany)<br />

M.P. Manns, Hannover (Germany)<br />

D. Vergani, London (Great Britain)


CONTENTS<br />

Session I<br />

Liver immunology I<br />

Session II<br />

Chair:<br />

V. Desmet, Leuven<br />

S.W. Schalm, Rotterdam<br />

Keynote lecture<br />

HCV – 20 years after discovery (No abstract)<br />

M. Houghton, San Francisco<br />

Page<br />

Immune responses mediated by endothelial cells<br />

P.A. Knolle, Bonn 17<br />

NKT cells within the liver<br />

M.G. Swain, Calgary 18<br />

Regulatory T cells induced within the liver<br />

J. Herkel, Hamburg 19<br />

Suicidal emperipolesis mediates deletion of T cells<br />

undergoing primary activation in the liver<br />

P. Bertolino, Sydney 20 – 21<br />

Liver immunology II<br />

Chair:<br />

E. Schrumpf, Oslo<br />

H.C. Thomas, London<br />

Role of serotonin and iNOS in LCMV-induced hepatitis<br />

K.S. Lang, Düsseldorf 25<br />

Lymphocyte homing to the liver<br />

D.H. Adams, Birmingham 26<br />

Monitoring immunosuppression after transplantation<br />

S. Meuer, C. Sommerer, M. Zeier, T. Giese, Heidelberg 27<br />

3


Session III<br />

4<br />

Cancer immunity and cancer immunotherapy<br />

A. Knuth, Zurich 28<br />

Pathology of immune-mediated liver injury<br />

H.P. Dienes, U. <strong>Dr</strong>ebber, Cologne 29 – 30<br />

Autoimmunity I<br />

Session IV<br />

Chair:<br />

G. Ramadori, Göttingen<br />

H. Ring-Larsen, Copenhagen<br />

Adaptive immunity in autoimmune hepatitis<br />

D. Vergani, G. Mieli-Vergani, London 33 – 34<br />

Old and new animal models of AIH<br />

E. Jaeckel, Hannover 35<br />

Cytochrome P450 2D6 as a model autoantigen<br />

U. Christen, Frankfurt 36 – 37<br />

Spontaneous chronic liver inflammation in a new transgenic<br />

mouse model of autoimmune hepatitis: Impact and regulation<br />

of autoreactive CD8 + T cells<br />

M. Zierden, Cologne 38<br />

Induction of tolerance in a mouse model of inflammatory<br />

liver disease<br />

G. Tiegs, Hamburg 39<br />

Autoimmunity II<br />

Chair:<br />

G. Gerken, Essen<br />

A.W. Lohse, Hamburg<br />

The role of the liver for the induction of oral tolerance in<br />

presence of naïve and antigen-experienced T cells<br />

N. Kruse, K. Neumann, A. Schrage, K. Eulenburg,<br />

M. Zeitz, A. Hamann, K. Klugewitz, Berlin 43<br />

Therapeutic options to treat AIH in 2009<br />

C.P. Strassburg, Hannover 44 – 48


Session V<br />

New insights into autoimmune cholangitis by animal models<br />

M. Trauner, Graz 49 – 50<br />

Etiological and molecular issues in primary biliary cirrhosis<br />

M.E. Gershwin, Davis 51 – 52<br />

Immunology of viral hepatitis I<br />

Session VI<br />

Chair:<br />

A. Alberti, Padua<br />

C. Trautwein, Aachen<br />

Occult hepatitis B virus infection: Diagnosis and<br />

significance<br />

W.H. Gerlich, Giessen 55<br />

Innate and adaptive immune responses in HBV infection<br />

M.K. Maini, London 56<br />

Hepatitis Delta: Immunopathogenesis and clinical challenges<br />

H. Wedemeyer, J. Grabowski, Hannover 57<br />

Clinical spectrum and therapeutic options in HDV infections<br />

M. Rizzetto, Torino 58<br />

Immunology of viral hepatitis II<br />

Chair:<br />

J. Schölmerich, Regensburg<br />

C. Trépo, Lyon<br />

Apoptosis in immune-mediated liver diseases<br />

A. Kahraman, G. Gerken, A. Canbay, Essen 61<br />

Immune response in HCV infection<br />

C. Ferrari, Parma 62<br />

Immunotherapy of hepatocellular carcinoma<br />

T.F. Greten, Hannover 63 – 64<br />

List of Chairpersons, Speakers and Scientific Organizers 65 – 68<br />

5


<strong>Poster</strong> <strong>Abstracts</strong><br />

NAFLD and NASH<br />

1. Nonalcoholic fatty liver disease in autopsy of children and adolescents<br />

F. Yüksel, D. Ergün, I. Yüksel, S. Kara, E. Samdanci, N. Celik (Ankara, TR)<br />

2. Non-invasive evaluation of liver fibrosis in patients with NAFLD (comparison<br />

between transient elastography and BARD score)<br />

A. Tudora, O.P. Ciof, E. <strong>Dr</strong>agan, D. Rill, I. Sporea, M. Scurtu-Martin<br />

(Timisoara, RO)<br />

3. Assessment of liver fibrosis by using transient elastography (TE) in patients with<br />

nonalcoholic fatty liver disease (NAFLD)<br />

O.P. Ciof, A. Tudora, E. <strong>Dr</strong>agan, D. Rill, I. Sporea, M. Scurtu-Martin<br />

(Timisoara, RO)<br />

4. HFE gene mutations in patients with biopsy-proven NAFLD<br />

E. Stachowska, J. Raszeja-Wyszomirska, J. Suchy, M. Lawniczak, I. Zawada<br />

(Szczecin, PL)<br />

5. Validation of BARD scoring system among Polish patients with NAFLD<br />

J. Raszeja-Wyszomirska, M. Lawniczak, B. Szymanik, P. Milkiewicz, M. Hartleb<br />

(Szczecin, Katowice, PL)<br />

6. High rate of undetected arterial hypertension in patients with NASH<br />

M. Demir, S. Schulte, S. Ubben, S.M. Lang, T. Goeser, U. Töx, H.-M. Steffen<br />

(Cologne, D)<br />

7. Fatty liver disease in children and adolescents<br />

V. Tzaneva, S. Galcheva, D. Baleva (Varna, BG)<br />

8. Effects of dietary chromium picolinate in the treatment of the experimental<br />

nonalcoholic steatohepatitis<br />

N. Kuzu, I.H. Bahcecioglu, A. Gencaslan, K. Metin, B. Üstündag, I.H. Özercan,<br />

K. Sahin (Izmir, Elazig, TR)<br />

9. In what manner does the metabolic syndrome modify the viral and alcoholic<br />

hepatic cirrhosis course?<br />

N.N. Silivontchik, E.I. Adamenko (Minsk, WR)<br />

10. Assessment of liver fibrosis by TE (FibroScan ® ) in patients with metabolic<br />

syndrome<br />

M. Scurtu-Martin, A. Tudora, O.P. Ciof, E. <strong>Dr</strong>agan, D. Rill, I. Sporea<br />

(Timisoara, RO)<br />

7


8<br />

Hepatitis A and B<br />

11. Investigation on cytokines at viral hepatitis A<br />

A. Petrov, N. Vatev (Plovdiv, BG)<br />

12. Blocking Bim-mediated attrition of T cells in patients with chronic HBV infection<br />

A. Schurich, P. Khanna, A. Ross Lopes, G. Nebbia, G.M. Dusheiko, M.K. Maini<br />

(London, GB)<br />

13. Chronic hepatitis HBV in children treated with interferon-α – Prognostic value of<br />

serum interleukin 6 and interleukin 12<br />

M. Gora-Gebka, A. Liberek, G. Sikorska-Wisniewska, J. Golebiewski,<br />

B. Kaminska (Gdansk, PL)<br />

14. High frequency of HBV-specific T cell responses in delta hepatitis<br />

J. Grabowski, P.V. Suneetha, J. Jaroszewicz, V. Schlaphoff, B. Bremer,<br />

M.P. Manns, M. Cornberg, H. Wedemeyer (Hannover, D; Bialystok, PL)<br />

15. Hepatitis B prevalence and flare in Asian patients with rheumatoid arthritis<br />

treated with disease modifying anti-rheumatic drug (DMARD) therapy<br />

O. Bebb, V.J. Appleby, P. Southern, P. Helliwell, S. Moreea (Bradford, GB)<br />

16. Interferon-α treatment fails to induce activation of interferon responsive genes in<br />

HBV chronically infected human chimeric uPA/SCID mice<br />

M. Lütgehetmann, T. Volz, A.W. Lohse, J.H. Bockmann, J. Petersen, M. Dandri<br />

(Hamburg, D)<br />

17. Relationship between liver fibrosis and topographic distribution of alpha-smooth<br />

muscle actin staining in chronic hepatitis B<br />

O. Kosseva, E. Pophristova, D.G. Adjarov, D. Jelev, Z. Krastev (Sofia, BG)<br />

18. Is any role of hepatic steatosis in chronic virus B hepatitis?<br />

C. Mihai, C. Cijevschi, G. Stefanescu, V.L. <strong>Dr</strong>ug, M. <strong>Dr</strong>agna, L. Graur, B. Mihai<br />

(Iasi, RO)<br />

19. The effects of chronic B and C hepatitis infections on bone mineral density<br />

C. Serban, L.M. Susan, A. Pacurari, L. Copaceanu, C. Banciu, G. Savoiu,<br />

C. Borza, R. Mateescu, I. Romosan (Timisoara, RO)<br />

20. TGF-β-driven hepatic progenitor cell expansion and ductular reaction<br />

contributes to HBV, but not schistosomiasis, associated liver fibrogenesis<br />

H.-L. Weng, L. Ciuclan, Y. Liu, C. Zhu, E. Wiercinska, I. Ilkavets, J. Dzieran,<br />

T. Huang, C. Meyer, R. Gebhardt, R. Heuchel, P. ten Dijke, J.-L. Chen,<br />

M.V. Singer, P.R. Mertens, S. Dooley (Mannheim, Magdeburg, Leipzig, D;<br />

Leiden, NL; Ningbo, Jiaxing, RC; Uppsala, Huddinge, S)<br />

21. Is there a link between the viral load value and liver stiffness for the<br />

asymptomatic virus B carriers?<br />

A. Tudora, O.P. Ciof, M. Scurtu-Martin, I. Sporea, E. <strong>Dr</strong>agan, D. Rill<br />

(Timisoara, RO)


Hepatitis C<br />

22. Evaluating antiinflammatory and antiviral effect due to the treatment with statins<br />

in patients with chronic virus C hepatitis<br />

O. Andreescu, L. Nedelcu, C. Scarneciu, D. Paul, I. Pantea (Brasov, RO)<br />

23. Variants of the innate immune response genes in patients with chronic<br />

HCV-infection<br />

A.O. Romanov, T.V. Belyaeva, E.V. Esaulenko (St. Petersburg, R)<br />

24. Expression and effect of the NK cell receptor 2B4 (CD244) on virus-specific<br />

CD8+ T cells – Another player in the control of viral hepatitis?<br />

V. Schlaphoff, J. Jaroszewicz, S.V. Pothakamuri, J. Grabowski, K. Stegmann,<br />

M.P. Manns, H. Wedemeyer, M. Cornberg (Hannover, D)<br />

25. Ethanol impairs MHC class I-restricted antigen presentation on HCV-infected<br />

liver cells<br />

N.A. Osna, R. White, S. Weinman, T. Donohue (Omaha, Kansas City, USA)<br />

26. Impairment of interferon-gamma (IFN-γ) signaling in hepatocytes by HCV<br />

proteins and ethanol<br />

L. Austin, R. White, K. Kharbanda, M. Beard, N.A. Osna<br />

(Omaha, USA; Adelaide, AUS)<br />

27. The platelet-derived chemokine CXCL4 (PF4) is a mediator of HCV-induced<br />

and experimental liver fibrosis<br />

M. Moreno Zaldivar, K. Pauels, M.-L. Berres, P. Schmitz, A. Kowalska,<br />

R. Weiskirchen, C. Trautwein, H.E. Wasmuth (Aachen, D)<br />

28. Interleukin 17 pathway is suppressed in chronic hepatitis C<br />

K. Gutkowski, T. Kacperek-Hartleb, J. Musialik, G. Boryczka, M. Kajor,<br />

M. Hartleb, E. Kaminska-Kiszka (Katowice, Rzeszow, PL)<br />

29. Allelic expression of rs2569190/C-159T CD14 gene variants in PBMC and liver<br />

samples from chronic hepatitis C patients<br />

R. Bregadze, G. Ramadori, S. Mihm (Göttingen, D)<br />

30. The content of sphingolipids in the liver is associated with the severity of chronic<br />

hepatitis C<br />

M. Dabrowska, A. Panasiuk, P. Zabielski, J. Gorski, R. Flisiak (Bialystok, PL)<br />

31. Ribavirin-induced anaemia during hepatitis C treatment is related to baseline<br />

haemoglobin leven and drug erythrocyte concentration<br />

M. Delle Monache, M. Carbone, L. Nosotti, L.R. Conti, I. Lenci, F. De Leonardis,<br />

M. Angelico, L. Baiocchi (Rome, I)<br />

32. Sylibin-vitamin E phospholipid complex reduces increased ferritin levels during<br />

antiviral treatment for chronic hepatitis C<br />

R. Cecere, L. Baiocchi, I. Lenci, M. Carbone, F. De Leonardis,<br />

M. Delle Monache (Rome, I)<br />

9


33. Thyroid function in patients with chronic hepatitis C virus infection under<br />

interferon therapy<br />

M. Azmy, Y. Ahmad, A. AbouNar, M. El-Hamamsy (Cairo, ET)<br />

34. The prevalence and impact of serum autoantibodies among chronic hepatitis C<br />

genotype 4 (HCV) patients attending a national center of HCV treatment<br />

program in Egypt<br />

M. Sharaf-Eldin, H. El-Batae (Tanta, ET)<br />

35. Absence of a correlation between endotoxin receptor CD14 rs2569190/C-159T<br />

polymorphism and liver disease progression in chronic hepatitis C<br />

E. Askar, M. Odenthal, H.P. Dienes, G. Ramadori, S. Mihm<br />

(Göttingen, Cologne, D)<br />

36. The metabolic control and antiviral treatment response in diabetic patients with<br />

chronic hepatitis C<br />

B. Mihai, C. Mihai, C. Cijevschi, V.L. <strong>Dr</strong>ug, M. <strong>Dr</strong>agna, L. Graur (Iasi, RO)<br />

37. The correlation between erythropoietin and pro- and anti-inflammatory cytokines<br />

and the decreasing of the viral C hepatitis RNA level under the treatment with<br />

statins<br />

R.G. Mihaila, R. Mihaila, O. Fratila, E.C. Rezi, A.V. Zaharie, L. Mocanu,<br />

M. Deac, A. Olteanu, L. <strong>Dr</strong>aghila, M. Boitan (Sibiu, Oradea, RO)<br />

38. Increased CD14 transcription initiation in T allele carriers of the rs2569190/<br />

C-159T promoter polymorphism in healthy individuals and chronic hepatitis C<br />

patients<br />

J. Mertens, G. Ramadori, S. Mihm (Göttingen, D)<br />

10<br />

Autoimmune hepatitis<br />

39. Familial occurrence of autoimmune hepatitis<br />

M. Wozniak, M. Woynarowski, J. Socha (Warsaw, PL)<br />

40. Liver sinusoidal endothelial cells induce anti-inflammatory CD4 + T cells<br />

suppressing murine autoimmune hepatitis<br />

N. Kruse, K. Neumann, K. Derkow, A.A. Kühl, A. Hamann, K. Klugewitz<br />

(Berlin, D)<br />

41. Frequency of autoimmune disorders and therapy in patients with autoimmune<br />

hepatitis<br />

A. Genunche-Dumitrescu, P. Mitrut, D. Badea, M. Badea (Craiova, RO)<br />

42. The prediction of the recrudescence risk in autoimmune hepatitis (AIH)<br />

P. Mitrut, A. Genunche-Dumitrescu, A.O. Mitrut, L. Nita, D. Badea<br />

(Craiova, RO)


43. Immune disruptions at the mycotoxin fumonisin-induced autoimmune hepatitis<br />

in young and old mice<br />

E.A. Martinova (Moscow, R)<br />

44. Emerging importance of autoimmune hepatitis in Bulgaria, an endemic area for<br />

viral hepatitis<br />

K. Kalinova, L.M. Pekova, P. Chakarova, P. Stefanova, K. Georgiev<br />

(Stara Zagora, BG)<br />

Cholestatic liver diseases (PBC/PSC)<br />

45. Non-invasive assessment of liver fibrosis by transient elastography in patients<br />

with cholestatic liver diseases (PBC and PSC)<br />

D. Rill, A. Tudora, O.P. Ciof, E. <strong>Dr</strong>agan, I. Sporea (Timisoara, RO)<br />

46. A meta-analysis regarding the effects of ursodeoxycholic acid in patients with<br />

primary sclerosing cholangitis<br />

E.C. Rezi, R.G. Mihaila, M. Deac (Sibiu, RO)<br />

47. The long-term budesonide-UDCA combined therapy in patients with primary<br />

biliary cirrhosis<br />

A. Genunche-Dumitrescu, P. Mitrut, M. Badea, D. Badea (Craiova, RO)<br />

Malignant and benign neoplasm<br />

48. Value of transient elastography (FibroScan ® ) in the evaluation of patients with<br />

hepatocellular carcinoma (HCC)<br />

O.P. Ciof, A. Tudora, E. <strong>Dr</strong>agan, D. Rill, M. Martin (Timisoara, RO)<br />

49. Constitutive active gp130 in the liver and its connection to chronic inflammation<br />

and carcinoma development<br />

A. Schütt, J. Scheller, S. Rose-John (Kiel, D)<br />

50. Increased intrahepatic infiltration with CD4+CD25+FOXP3+ regulatory T cells<br />

correlates with increased tissue levels of TGF-beta and IL-10 in patients with<br />

primary or metastatic liver cancers<br />

I.M. Manolova, M.V. Gulubova, D. Kyurkchiev, J. Ananiev, I.P. Altunkova,<br />

A. Julianov (Stara Zagora, Sofia, BG)<br />

51. TGF-beta occurrence and macrophage infiltration in liver metastasis<br />

J. Ananiev, M.V. Gulubova, I.M. Manolova, A. Matev (Stara Zagora, Sofia, BG)<br />

52. TGF-beta inhibits dendritic cell infiltration in liver metastasis<br />

M.V. Gulubova, I.M. Manolova, J. Ananiev, Y. Yovchev (Stara Zagora, BG)<br />

53. Focal nodular hyperplasia of the liver within our experience<br />

H. Cichoz-Lach, B. Prozorow-Król, J. Swatek, K. Celinski, M. Slomka, L. Buk,<br />

E. Korobowicz, E. Lis (Lublin, PL)<br />

11


12<br />

Basic immunology<br />

54. Liver sinusoidal endothelial cells induce TGF-β dependent conversion of<br />

CD4+Foxp3+ regulatory T cells from conventional CD4+CD25- T cells<br />

A. Carambia, S. Huber, S. Lüth, D. Schwinge, C. Frenzel, C. Schramm,<br />

A.W. Lohse, J. Herkel (Hamburg, D)<br />

55. Alcohol-metabolizing enzyme gene polymorphisms in alcohol liver cirrhosis<br />

among Polish men<br />

H. Cichoz-Lach, E. Lis, K. Celinski, M. Slomka (Lublin, PL)<br />

56. Conversion of naïve T cells into regulatory T cells after tolerance induction in<br />

the murine liver<br />

B. Claaß, A. Erhardt, G. Tiegs (Hamburg, D)<br />

57. Impact of IL-6-transsignaling on liver damage and regeneration after CCI4<br />

J. Gewiese, C. <strong>Dr</strong>ucker, J. Scheller, S. Rose-John (Kiel, D)<br />

58. CXCL10 plays a key role in the development of hepatic fibrosis<br />

E. Hintermann, M. Bayer, A. Luster, U. Christen<br />

(Frankfurt, D; Charlestown, USA)<br />

59. Epitope mapping of human cytochrome P450 2D6 (CYP2D6) in patients with<br />

autoimmune hepatitis and in the CYP2D6 mouse model<br />

E. Hintermann, M. Holdener, M. Bayer, M.P. Manns, U. Christen<br />

(Frankfurt, Hannover, D)<br />

60. The role of the hepatic asialoglycoprotein receptor in inflammatory-induced liver<br />

injury and accumulation of intrahepatic lymphocytes<br />

B.L. McVicker, D.J. Tuma, G.M. Thiele, C.A. Casey, N.A. Osna (Omaha, USA)<br />

61. Common variable immunodeficiency-associated granulomatous disease and<br />

nodular regenerative hyperplasia of the liver<br />

H. Varnholt, T. Rubinas (Chapel Hill, USA)<br />

62. Aggravation of liver damage and loss of immunological tolerance in CXCR3deficient<br />

mice in the murine model of concanavalin A-induced liver injury<br />

C. Wegscheid, A. Erhardt, U. Panzer, G. Tiegs (Hamburg, D)<br />

63. Metallothioneins (MT I and MT II) and their gene expression in some chronic<br />

liver diseases<br />

M. Sharaf-Eldin, I. Saber, U. Negm, H. Hamouda, A.A. Hameed (Tanta, ET)<br />

64. Endothelial chemokine presentation as therapeutic target in T-cell mediated<br />

hepatitis<br />

K. Neumann, K. Wechsung, N. Kruse, M. Schumann, A.A. Kühl, A. Hamann,<br />

K. Klugewitz (Berlin, D)


Hepatic vascular disorders, drugs and toxicity, epidemiology and<br />

endoscopy posters<br />

65. Hemostatic disturbances in patients with portal thrombosis<br />

E. Lukina, E. Sysoeva, E. Jakovleva, E. Kitsenko, G. Sukhanova, S. Vasijev<br />

(Moscow, R)<br />

66. False iron deficiency in patients with prehepatic portal hypertension<br />

E. Lukina, E. Sysoeva, A. Levina, Y. Mamukova, E. Voronkova, E. Kitsenko<br />

(Moscow, R)<br />

67. Diagnostics improvement for liver chronic diseases in Turkmenistan<br />

O.B. Nepesova (Ashgabat, XTU)<br />

68. Epidemiology of liver diseases in a pediatric population. Snapshot data from<br />

selected centers in Poland<br />

M. Woynarowski, M. Wozniak, W. Chlebcewicz-Szuba, T. Chmurska Motyka,<br />

A. Gorczyca, B. Korczowski, A. Krzywicka, D.M. Lebensztejn, A. Liberek,<br />

E. Majda-Stanislawska, M. Rokitka, E. Strawinska, S. Wiecek, I. Zaleska,<br />

A. Zdanowska-Ruskan, J. Socha (Warsaw, PL)<br />

69. Alcohol withdrawal treatment audit March 2009, Northern General Hospital,<br />

Sheffield, UK<br />

A. Al-Joudeh, H. Morton, S. Heikal, K. Basu (Sheffield, GB)<br />

70. Effect of different degrees of hepatic dysfunction on the pharmacokinetics of<br />

telmisartan<br />

M. El-Hamamsy, S. Hegazy, W. Awara (Cairo, ET)<br />

71. Risk perception and use of herbal remedies in patients with liver/biliary tract<br />

disorders: An Italian study<br />

M. Marignani, S. Gallina, M. Di Fonzo, I. Deli, P. Begini, F. Attilia, E. Gigante,<br />

M. Epifani, S. Angeletti, G. Delle Fave (Rome, I)<br />

72. Hepatotoxicity induced by antituberculosis drugs<br />

M. González, E. Aravena, C. Peña, E. Astrosa, J. Lyon, G. Serrano, P. Escobar<br />

(Santiago de Chile, RCH)<br />

73. Treatment with Ursofalk ® of bile reflux gastritis<br />

D. Pestroiu, M. Ilie, D. Brooks, C. Vladut (Bucharest, RO)<br />

74. Incidence and therapy of colonic perforations: How feasible and effective is<br />

endoscopic closure?<br />

K. Mönkemüller, H. Neumann, P. Malfertheiner, H.-U. Schulz, L.C. Fry<br />

(Bottrop, Magdeburg, D)<br />

13


Session I<br />

Liver immunology I<br />

15


Immune responses mediated by endothelial cells<br />

Percy A. Knolle, MD<br />

Institute of Molecular Medicine, Bonn, Germany<br />

The liver is an immune organ favouring induction of immune tolerance rather than<br />

T cell immunity. Both, organ-resident cell populations as well as the unique hepatic<br />

microenvironment contribute to skewing of immune responses towards induction of<br />

tolerance.<br />

Endothelial cells of the liver, in particular liver sinusoidal endothelial cells (LSEC)<br />

constitute one these organ-resident cell populations with important immune functions.<br />

LSEC play a central role in local immune regulation in the liver, as they scavenge<br />

most efficiently soluble circulating antigens for cross-presentation on MHC I<br />

molecules to CD8 T cells. Cross-presentation to naïve CD8 T cells by LSEC in vitro<br />

and in vivo leads to induction of a tolerant T cell phenotype characterized by lack of<br />

cytokine expression and cytotoxic effector function upon restimulation. Mechanistically,<br />

mutually tolerogenic signaling through the co-inhibitory molecules B7H1<br />

and PD1 on LSEC and naïve CD8 T cells, respectively, is required for tolerance<br />

induction. Moreover, LSEC function as veto cells to impede antigen-presentation by<br />

dendritic cells to naïve T cells while they reside in the liver and thereby further skew<br />

local hepatic immune responses towards tolerance induction. Importantly, LSEC play<br />

an important role in the effector phase of CD8 T cell immune responses in the liver.<br />

Antigens released from virus-infected hepatocytes are cross-presented by LSEC to<br />

effector CD8 T cells and initiate virus-specific immunity even in the absence of direct<br />

antigen-recognition on hepatocytes, thus demonstrating that cross-presentation is<br />

important also during the execution phase of the immune response in the target<br />

tissue.<br />

Given their strategic localization in the hepatic sinusoid separating hepatocytes from<br />

circulating lymphocytes LSEC play an important role both in skewing of immune<br />

responses towards tolerance induction and immune surveillance during viral<br />

infection.<br />

17


NKT cells within the liver<br />

M.G. Swain, MD, MSc, FRCP<br />

Professor of Medicine, University of Calgary, Calgary, AB, Canada<br />

NKT cells are distinct innate immune T cells which play key roles in the pathogenesis<br />

of various immune-mediated liver diseases. NKT cells are traditionally defined as<br />

cells that co-express the T cell receptor (TCR) and certain natural killer cell surface<br />

markers (e.g. NK1.1 in some mouse strains). Based on CD1d restriction, they are<br />

broadly classified as invariant (iNKT) and non-invariant NKT cells. Both NKT cell<br />

types are abundant within the liver and, mainly through the study of animal models of<br />

immune-mediated liver injury, the role of NKT cells in the context of liver inflammation<br />

is becoming better understood. iNKT cells are the NKT cell subtype which has been<br />

best characterized and therefore this presentation will focus mainly on iNKT cells.<br />

NKT cells recognize both host and microbial glycolipid antigens, presented in the<br />

context of the CD1d molecule, via their TCR. Upon activation NKT cells rapidly<br />

release large quantities of T helper type 1 (Th1; e.g. IFNγ, TNFα), Th2 (e.g. IL-4,<br />

IL-10), and Th17 (e.g. IL-17, IL-21) type cytokines with diverse immunoregulatory<br />

and immunomodulatory effects. Moreover, through their release of these cytokines<br />

NKT cells can drive subsequent downstream immune responses in tissues such as<br />

the liver, including the upregulation of adhesion molecule expression and increased<br />

chemokine production. By regulating immune cell adhesion pathways and chemokine<br />

secretion, NKT cells can direct both proinflammatory (e.g. Th1 cell) and anti-inflammatory<br />

(e.g. regulatory T cell) cell infiltration into the liver. Through this mechanism<br />

NKT cells orchestrate a balance between pro- and anti-inflammatory responses<br />

within the liver during immune mediated liver damage. In this presentation I will<br />

highlight the diverse roles played by iNKT cells as master regulators of immunemediated<br />

liver injury.<br />

18


Regulatory T cells induced within the liver<br />

Johannes Herkel<br />

Department of Medicine I, University Medical Centre Hamburg-Eppendorf, Hamburg,<br />

Germany<br />

CD4 + CD25 + Foxp3 + regulatory T cells (Treg) cells are important mediators of immune<br />

tolerance to self-antigens, and their inactivation may cause autoimmune disease.<br />

Therefore, immunotherapy for autoimmune diseases may aim at restoring or<br />

expanding autoantigen-specific Treg activity. We reported recently that Tregmediated<br />

suppression of autoimmune disease could be achieved in vivo, by taking<br />

advantage of the liver’s ability to promote immune tolerance (Lüth et al. J Clin Invest<br />

2008; 118: 3403–3410). Ectopic expression of the neural autoantigen myelin basic<br />

protein (MBP) in the liver, which was accomplished constitutively in liver-specific<br />

MBP-transgenic mice, or transiently after gene transfer to liver cells in vivo, induced<br />

protection from autoimmune neuroinflammation in a mouse model for human multiple<br />

sclerosis. Protection from autoimmunity was mediated by MBP-specific<br />

CD4 + CD25 + Foxp3 + Treg cells, as demonstrated by adoptive transfer to nontransgenic<br />

mice. The protective Treg were MBP-specific, since they efficiently<br />

suppressed conventional CD4 + CD25 - T cells after antigen-specific stimulation with<br />

MBP in vitro. The generation of these MBP-specific Treg cells in vivo depended on<br />

expression of MBP in the liver, but not in the skin, and occurred by TGFβ-dependent<br />

peripheral conversion from conventional CD4 T cells to CD4 + CD25 + Foxp3 + Treg. To<br />

further address the mechanism of Treg generation in the liver, we analyzed which<br />

liver cell type could mediate the conversion to Treg, and found that the liver<br />

sinusoidal endothelial cells (LSEC) seem to be efficient converters. Indeed, the rate<br />

of TGFβ-dependent conversion of non-Treg into Treg induced by LSEC was<br />

significantly greater than that induced by dendritic cells. These findings indicated that<br />

autoantigen expression in the liver may generate autoantigen-specific Treg cells and<br />

that hepatic Treg generation seems to be an important component of hepatic<br />

tolerance.<br />

19


Suicidal emperipolesis mediates deletion of T cells undergoing<br />

primary activation in the liver<br />

Patrick Bertolino<br />

Centenary Institute of Cancer Medicine and Cell Biology, Sydney, Australia<br />

E-Mail: p.bertolino@centenary.usyd.edu.au<br />

Despite being a non-lymphoid organ, the liver displays immunological properties<br />

distinct from other solid organs and is associated with the induction of T cell<br />

tolerance. This property has been demonstrated in several clinical settings including<br />

transplantation, and hepatotropic viral infections, such as those induced by hepatitis<br />

B and C viruses (HBV and HCV).<br />

Many models have been proposed to explain the “liver tolerance effect”, but the<br />

molecular and cellular mechanism/s mediating this phenomenon remain unknown.<br />

We have previously shown that the liver is the only non-lymphoid organ able to retain<br />

and activate naïve CD8 + T cells independently of lymphoid tissues in an antigenspecific<br />

manner. These findings, confirmed by other groups, have opened new<br />

possibilities to explain the remarkable property of the liver to induce antigen-specific<br />

tolerance in transplantation and following infection by hepatotropic viruses such as<br />

HCV and HBV.<br />

To investigate the mechanisms by which the liver induces antigen-specific tolerance,<br />

we have studied the fate of T cell activated intrahepatically in a model in which most<br />

T cells are retained and activated in this organ. Naïve transgenic Des-TCR CD8 +<br />

T cells specific for the MHC class I molecule H-2K b plus self peptide were adoptively<br />

transferred into mice expressing the antigen ubiquitously (C57BL/6 or C57BL/6x C3H<br />

F1) or into syngeneic mice with no antigen expression (B10.BR). Des-TCR cells were<br />

either radio-labelled with 51 Cr or labelled with the fluorochrome CFSE before transfer.<br />

Within 1 hour, most T cells were retained in the liver of C57BL/6 mice and<br />

upregulated the early activation marker CD69 within 30 min after adoptive transfer.<br />

Despite accumulation of potentially autoreactive T cells, no clinical signs of dermatitis<br />

(rash), colitis (diarrhea) or serological hepatitis (ALT levels) were observed in<br />

C57BL/6 mice injected with CD8 + Des-TCR T cells. Flow cytometry and histology<br />

confirmed that lower numbers of CFSE-labelled Des-TCR T cells were detected in<br />

the liver of C57BL/6 mice within 22 hours after adoptive transfer suggesting that<br />

T cells were physically eliminated in the recipient animals in an antigen-specific<br />

manner.<br />

To determine whether the fallen T cell numbers recovered from the liver was due to<br />

migration to other sites, donor cells were radio-labelled prior to transfer, and organs<br />

harvested at 1 and 22h. Consistent with CFSE data, up to 70% of the total<br />

radioactivity recovered from B6 and F1 animals was contained within the liver at 1h.<br />

Surprisingly, in contrast to the CFSE studies, radioactivity distribution remained<br />

almost unchanged at 22h suggesting that the donor cell content was still retained<br />

within the liver, despite failure to recover viable transferred autoreactive cells. Few<br />

apoptotic cells were detected in the livers of recipient B6 mice during the first 22h<br />

and neither macrophages nor exposure of phosphatidylserine on the T cell surface<br />

following intrahepatic activation were critical for rapid T cell clearance, as inhibition of<br />

20


these pathways failed to prevent deletion. This deletion was not mediated by NK cells<br />

or allogeneic host T cells and was not due to cell phagocytosis by Kupffer cells.<br />

Instead, T cells actively invaded hepatocytes and were found into Lamp-1 coated<br />

vesicles suggesting that T cells are degraded into lysosomes.<br />

In summay, we identify a novel mechanism of peripheral deletion in which the<br />

majority of naïve autoreactive CD8 + T cells is rapidly eliminated in the liver following<br />

intrahepatic activation. T cells actively invade hepatocytes, enter endosomal/lysosomal<br />

compartments where they are degraded by a non-apoptotic mechanism.<br />

Invasion of a cell into another cell is known as emperipolesis, a long-observed<br />

phenomenon for which a physiological role has not been previously demonstrated.<br />

We propose that T cell suicidal emperipolesis is a novel mechanism of autoreactive T<br />

cell deletion, a process critical for the maintenance of tolerance.<br />

21


Session II<br />

Liver immunology II<br />

23


Role of serotonin and iNOS in LCMV-induced hepatitis<br />

<strong>Dr</strong>. K.S. Lang<br />

Humboldt Research Group, Department of Gastroenterology, Hepatology and<br />

Infectiology, Heinrich Heine University Düsseldorf, Germany<br />

More than 500 million people worldwide are persistently infected with hepatitis B<br />

(HBV) or hepatitis C virus (HCV). Although both viruses are poorly cytopathic,<br />

persistence of either virus carries a risk of chronic liver inflammation potentially<br />

resulting in liver-steatosis, liver cirrhosis, end stage liver failure or hepatocellular<br />

carcinoma. Virus-specific T cells are a major determinant of the outcome of hepatitis<br />

since these cells contribute to the early control of chronic hepatitis viruses, however<br />

they also mediate immunopathology during persistent virus infection.<br />

We have analyzed the role of the platelet-derived vasoactive serotonin during virusinduced<br />

CD8 + T-cell dependent immunopathological hepatitis in mice infected with<br />

the noncytopathic lymphocytic choriomeningitis virus (LCMV). Following virus<br />

infection, platelets were recruited to the liver and their activation correlated with<br />

severely reduced sinusoidal microcirculation, delayed virus elimination and increased<br />

immunopathological liver cell damage. Lack of platelet-derived serotonin in serotonin<br />

deficient mice normalized hepatic microcirculatory dysfunction accelerated virus<br />

clearance in the liver and reduced CD8 + T-cell dependent liver cell damage. In<br />

keeping with these observations, serotonin treatment of infected mice delayed entry<br />

of activated CD8 + T-cells into the liver, delayed virus control and aggravated<br />

immunopathological hepatitis.<br />

Next we analyzed the role of dendritic cells in virus infected livers. We found that<br />

dendritic cells were recruited into the liver during infection with lymphocytic<br />

choriomeningitis virus. Liver-infiltrating CD11c + cells expressed high levels of MHC II<br />

and iNOS. Interestingly, in the liver, iNOS expression in dendritic cells enhanced<br />

T cell mediated IFN-γ production. Conversely, IFN-γ production by T cells enhanced<br />

iNOS expression. Reduced IFN-γ production led to enhanced infection of<br />

hepatocytes followed by severe immunopathology.<br />

In conclusion, Serotonin and iNOS influence CD8 T cell function in the liver, which<br />

influences T cell mediated immunopathology during virus infection.<br />

25


Lymphocyte homing to the liver<br />

David H. Adams<br />

Centre for Liver Research, 5th Floor IBR, The University of Birmingham Medical<br />

School, Wolfson <strong>Dr</strong>ive, Edgbaston, Birmingham, B15 2TT, UK<br />

In chronic liver disease leucocyte recruitment from blood drives inflammation, fibrosis<br />

and cirrhosis. The liver has a dual blood supply from the portal vein and hepatic<br />

artery the terminal branches of which supply the parenchyma via hepatic sinusoids<br />

lined by endothelial cells intimately associated with pericytes called hepatic stellate<br />

cells. Blood percolates slowly through the sinusoids before re-entering the circulation<br />

via hepatic veins. Sinusoidal endothelium is morphologically and functionally distinct<br />

from endothelium in other tissues, lacking both tight junctions and a basement<br />

membrane. These unique features suggest that lymphocyte recruitment via the<br />

sinusoids will involve distinct mechanisms. I will review how specific combinations of<br />

chemokines and adhesive events regulate lymphocyte recruitment in the context of<br />

this unique microanatomical environment. The effect of low flow rates in sinusoids on<br />

the molecular requirements for capture and the consequences of the discontinuous<br />

nature of sinusoidal endothelium on the dynamics of transendothelial migration will<br />

be discussed. I will also present evidence to show how all of these processes are<br />

modulated by paracrine signals from adjacent cells in the sinusoid including epithelial<br />

cells (hepatocytes), fibroblasts (stellate cells) and haematopoietic cells (Kupffer<br />

cells). Under basal conditions these paracrine signals maintain sinusoidal endothelial<br />

phenotype and function and in response to inflammation prime endothelium to recruit<br />

effector cells thereby driving the development and maintenance of chronic<br />

inflammation.<br />

26


Monitoring immunosuppression after transplantation<br />

Stefan Meuer, Claudia Sommerer, Martin Zeier and Thomas Giese<br />

University Hospital Heidelberg, Germany<br />

With the introduction of calcineurin inhibitors (CNI) long-term allograft function has<br />

significantly improved. The problem of limited therapeutic margins and the toxicity of<br />

CNI remain unsolved. The recently reported quantitative assessment of inhibition of<br />

NFAT-regulated gene expression 2 hr after Cyclosporine A (CsA) intake represents a<br />

novel approach to evaluate the biological effectiveness of CsA therapy and provides<br />

means to enable individualized immunosuppressive regimens.<br />

In more than 200 patients carrying kidney and in 55 patients with heart allografts we<br />

compared the degree of inhibition of IL-2, IFN-γ and GM-CSF gene expression with<br />

the peak blood concentration of CsA. Functional immunosuppression as assessed by<br />

qRT-PCR varies considerably among CsA treated individuals with stable graft<br />

function. Given the relatively constant level of inhibition over a broad range of drug<br />

concentrations, we felt that a considerable group of patients with unnecessarily high<br />

CsA doses might benefit from a reduced dosing of the drug without compromising the<br />

efficacy of the immunosuppressive therapy. Importantly, we could demonstrate that<br />

patients with infectious episodes during the period of observation had a significant<br />

stronger NFAT-inhibition than patients without complications, indicating that this<br />

pharmacodynamic assay can better predict the risk of infection due to excessive<br />

immunosuppression of transplanted patients in comparison to the conventional<br />

C2-monitoring of CsA. In several patients the doses of CsA could be safely reduced<br />

without significantly changing the level of immunosuppression.<br />

In conclusion, patients treated with CsA might benefit from a reduced dosage of the<br />

drug, if they respond to the drug with a strong inhibition of NFAT-regulated gene<br />

expression.<br />

27


Cancer immunity and cancer immunotherapy<br />

Alexander Knuth<br />

Department of Oncology, University Hospital of Zurich, Zurich, Switzerland<br />

Nearly 35 years ago a new strategy was introduced for the discovery of human<br />

cancer antigens called ‘autologous typing’. Cancer reactive serum antibodies were<br />

tested for their specificity in reacting with autologous tumor cells in vitro. The concept<br />

of autologous typing was then extended to the study of T-lymphocyte clones (CTL)<br />

with testing their specificity for autologous cancer cells in vitro. Combined with<br />

molecular cloning techniques, CTL clones were employed to identify the first human<br />

cancer specific antigen family, the MAGE gene family in a patient MZ-2 from the<br />

University Hospital in Mainz, Germany. MAGE became the prototype of what now is<br />

known as the cancer testis family of antigens (CT-Ag). CT antigens have a unique<br />

expression pattern, i.e. predominantly in cancer cells of different tissue origins and in<br />

germ cells but not in any other normal tissues. The molecular function of these<br />

intracellular antigens is largely unknown, although 10% of the genes encoded on the<br />

X-chromosome appear to be CT antigens. Their immunogenicity in vivo is indicated<br />

by spontaneous and often strong immune responses in cancer patients. The CT<br />

antigen NY-ESO-1 is the most immunogenic human cancer antigen known to date,<br />

often leading to integrated immune responses with NY-ESO-1 specific T cell<br />

responses as well as antibodies. The presence of NY-ESO-1 specific antibodies<br />

indicates the presence of disease and of specific T cells in approximately 50% of<br />

patients. CT-antigen specific immune interventions may induce T cells and antibodies<br />

or may augment pre-existent CT-antigen specific immunity. The cellular and<br />

molecular networks shaping these immune responses are studied with respect to<br />

regulatory T cell subsets and inhibitory receptors like CTLA-4 and other agents to<br />

influence immune reactions in cancer patients. In combination with new antigen<br />

formulations and delivery technologies, adjuvants and regulators of specific immune<br />

responses, cancer immunotherapy will become a valid modality in cancer treatment.<br />

28


Pathology of immune-mediated liver injury<br />

H.P. Dienes, U. <strong>Dr</strong>ebber<br />

Institute of Pathology, University of Cologne, Cologne, Germany<br />

The liver is a special lymphoid organ with its own defense mechanisms and is prone<br />

to chronic viral and autoimmune diseases. All constituent liver cells are involved in<br />

immune mechanisms against foreign toxins and agents: hepatocytes, Kupffer cells,<br />

endothelial cells, hepatic stellate cells (HSC) and liver resident lymphocytes as NK,<br />

NKT and dendritic cells. The first line of defense is build up by innate immune system<br />

where as the second line and more antigen specific immune barrier is the adaptive<br />

immune response. Even in drug injury of the liver the immune system is involved. In<br />

the following contribution I just concentrate on some major issues of various<br />

etiologies and the corresponding immune reaction with underscoring the histopathology<br />

of immune mediated lesions in the liver.<br />

In drug induced liver injury mediated by idiosyncratic mechanisms the innate immune<br />

system is activated and can lead to hepatocellular necrosis especially in zone 3.<br />

There is an activation of NK and NKT cells and liver cell damage is induced mainly<br />

by the Fas/FasL pathway, a mechanism that has been investigated for the side<br />

effects in acetaminophen and dihydralazine damage.<br />

In viral hepatitis the pathway of liver injury is mainly immune mediated since almost<br />

none of the hepatitis viruses have turned out to be cytopathic by itself.<br />

In hepatitis A the viral agents are orally transmitted and invade the liver by the portal<br />

vein blood. The predominant liver lesions are therefore in the portal and periportal<br />

area with cytotoxic lymphocytes being the main effectors cells. In HBV infection<br />

about 5 to 10 percent of the individuals run a chronic course and the severity of liver<br />

damage is completely dependent on the presence of an efficient immune response to<br />

the virus with CDS antigen specific cytotoxic lymphocytes. In hepatitis C virus<br />

infection the percentage of a chronic course is much higher than in HBV infection.<br />

Several mechanisms have been explored, that could show, that the viral components<br />

interfere with intracellular signals of interferon production but also at the level of<br />

cellular immune response. We could show that the percentage of inhibitory Tregs is<br />

higher in chronic hepatitis C than in chronic hepatitis B. The fraction of T-cells<br />

displaying PD1, FoxpS and the number of dendritic cells with IDO are significant<br />

larger in hepatitis C than in hepatitis B.<br />

In hepatitis D there are almost identical histopathology and immune mechanisms as<br />

in hepatitis B, but they are on average more severe than in hepatitis B. Also HEV is<br />

regarded as a non cytopathic viral agent and the immune mechanisms and also<br />

histopathology are similar to hepatitis A. In EBV infection liver cell injury seems not to<br />

be due to direct antigen specific CTL attack since hepatocytes are not permissive for<br />

EBV replication and hepatocytes do not harbour viral proteins. Instead the injury is<br />

induced by so called 'bystander reaction' from activated lymphocytes in the sinusoids<br />

of the liver. They seem to be involved Kupffer cell activated T-cell, NK and NKT cells<br />

with various kinds of cytotoxic cytokines and chemokines. So an infection with EBV<br />

mechanisms different from antigen specific immune response seems to be effective.<br />

29


Autoimmune hepatitis is paradigm of immune mediated liver injury and has a special<br />

but not pathognomonic histopathology. Necroinflammatory activity starts from the<br />

portal and periportal area and moves on to the other zones of the lobule. By histopathology<br />

autoimmune hepatitis is a rather progressive disease with cell necrosis<br />

and apoptosis linked to the presence of effector lymphocytes that show the<br />

phenomenon of emperipolesis. Periportal hepatocytes undergo resetting and the<br />

presence of active interface hepatitis is a salient feature of autoimmune hepatitis.<br />

In conclusion the manifestation of immune mediated liver injury is mainly determined<br />

by the dominant mechanism if it is innate immune response, adaptive immune<br />

answer or 'bystander reaction'. Histopathology reflects different patterns of the<br />

etiologic agent and pathways of defense mechanisms.<br />

30


Session III<br />

Autoimmunity I<br />

31


Adaptive immunity in autoimmune hepatitis<br />

Diego Vergani, MD, PhD, FRCPath, FRCP; Giorgina Mieli-Vergani, MD, PhD,<br />

FRCPCH, FRCP<br />

Institute of Liver Studies, King’s College London School of Medicine at King’s College<br />

Hospital, Denmark Hill, London SE5 9RS, UK<br />

The histological lesion of interface hepatitis, with its striking infiltrate of lymphocytes,<br />

plasma cells, and monocytes/macrophages was the first to suggest an autoaggressive<br />

cellular immune attack in the pathogenesis of autoimmune hepatitis (AIH).<br />

Immunohistochemical studies, focused on the phenotype of the inflammatory cells<br />

infiltrating the liver parenchyma in AIH, have revealed a predominance of alpha/beta<br />

T-cells. Amongst these cells, a majority was found to be positive for the CD4<br />

helper/inducer phenotype and a sizeable minority for the CD8 cytotoxic/suppressor<br />

phenotype. Lymphocytes of non-T-cell lineage included NK cells, monocytes/macrophages<br />

and B-lymphocytes.<br />

Tr<br />

APC<br />

APC<br />

Peptide<br />

Co-stimuli<br />

TGF-β<br />

Class II<br />

Th0<br />

IL-4<br />

IL-12<br />

Th2<br />

Liver<br />

cell<br />

Class II<br />

Th1<br />

IL-4<br />

Class I<br />

IL-2<br />

IL-10<br />

IFN-γ<br />

IFN-γ<br />

IL-13<br />

IL-6<br />

CTL Tc<br />

T Tc C<br />

B<br />

M<br />

IL-1<br />

NK<br />

P<br />

Th17 Ts<br />

Liver damage is believed to initiate with the presentation of a self-antigenic peptide to<br />

the T-cell receptor (TCR) of an uncommitted T helper (TH0) lymphocyte by professional<br />

antigen presenting cells (APCs), such as macrophages, dendritic cells and<br />

B-lymphocytes, in the presence of co-stimulating signals, induced by the interaction<br />

of CD28 on TH0 and CD80 on APC. Once exposed to the antigen embraced by an<br />

HLA class II molecule on the APC, the TH0 lymphocytes become activated and,<br />

according to the presence in the milieu of interleukin 12 (IL-12) or IL-4, they<br />

differentiate into TH1 or TH2 cells and initiate a series of immune reactions<br />

determined by the cytokines they produce. TH1 cells predominantly secrete IL-2 and<br />

IFN-gamma, the latter being the main orchestrator of tissue damage because of its<br />

ability to stimulate cytotoxic T-cells (CTL), to enhance the expression of HLA class I<br />

molecules on APCs and of HLA class II molecules on hepatocytes and activate<br />

TNF-α<br />

IL-17<br />

IL-17<br />

33


monocytes/macrophages, which in turn release IL-1 and TNF-alpha. The expression<br />

of HLA class II on hepatocytes, that normally do not express it, enables them to<br />

present the autoantigen to TH1 lymphocytes and hence the perpetuation of the<br />

autoimmune process. The function of TH1 cells is counterbalanced by that of TH2<br />

cells, arising in the presence of IL-4 and mainly producing IL-4, IL-10 and IL-13.<br />

These cytokines induce also the maturation of B-cells into plasma cells, with<br />

consequent production of autoantibodies. As regulatory cells in AIH are numerically<br />

and functionally deficient, effector responses are perpetuated with ensuing persistent<br />

liver cell destruction by the direct action of CTL, cytokines released by TH1 cells and<br />

monocytes/macrophages, complement activation or engagement of natural killer<br />

(NK) cells by the autoantibodies bound to the hepatocyte surface. The role of the<br />

recently described TH17 cells, which arise in the presence of transforming growth<br />

factor beta (TGF-beta) and IL-6, is under investigation.<br />

34


Old and new animal models of AIH<br />

E. Jaeckel, M.D.<br />

Department of Gastroenterology, Hepatology, Endocrinology, Diabetology, Hannover<br />

Medical School, Hannover, Germany<br />

Autoimmune hepatitis (AIH) is difficult to study in humans because of the late<br />

diagnosis, the heterogenic genetic background and the lack of tissue specimens.<br />

Therefore, animal models have been used for more than 90 years to study liverspecific<br />

immune regulation. Early studies of experimental AIH established the<br />

importance of T cells for causing the liver injury and how regulatory pathways were<br />

detected. Studies in transgenic models demonstrated that T cells can be tolerant to<br />

liver antigens by ignorance, deletion, anergy, and receptor downregulation. They<br />

furthermore defined functions in antigen presentation for hepatocytes and sinusoidal<br />

endothelial cells. Studies in cytokine-induced models demonstrated common effector<br />

cascades resulting from T cell activation and the interaction of innate and adaptive<br />

immune response. In the past approaches for breaking tolerance against liver<br />

antigens in animal models were just able to induce transient and self-limiting immune<br />

responses. Here we have developed a strategy to break tolerance of a negatively<br />

selected, wild-type T cell repertoire by combining a self limited infection providing a<br />

danger signal together with modified liver self-antigens into mice. To this end we<br />

generated replication-deficient adenoviral constructs expressing common<br />

autoantigens of human AIH showing homology up to 80% with the murine proteins.<br />

Following an acute phase of liver destruction with increasing levels of transaminases<br />

two weeks after intravenous injection of the virus, we could show chronic evolving<br />

hepatic autoimmune reactions after twelve weeks. Chronic liver disease was<br />

confirmed by consistent leukocyte infiltrates and antigen-specific autoantibodies<br />

within the sera. Immunofluorescent staining of liver sections revealed that CD4 T<br />

cells are more abundant than CD8 T cells in these liver infiltrates. Control<br />

experiments with expression vectors devoid of viral components did not result in any<br />

hepatic reaction after the acute viral infection had been cleared. Furthermore chronic<br />

AIH was not induced in FVB mice. However we detected autoimmunity within the<br />

liver of non-obese diabetic (NOD) mice which received viral constructs expressing<br />

FTCD, an important auto-antigen in AIH type II. Surprisingly the more important AIH<br />

type II auto-antigen CYP2D6 as well as the type III soluble liver antigen (SLA) did not<br />

result in any chronic hepatic immune reaction. This strain and antigen-specificity of<br />

murine AIH supports the notion that autoimmunity develops in genetically<br />

predisposed individuals. To identify the leukocyte populations involved in hepatitis<br />

within our model we isolated liver-specific leukocytes of hepatitis-bearing mice and<br />

analysed these by flow cytometry as well as by a new technique of microimmunology<br />

called iterative Slide Based Cytometry (iSBC). Taken together, this murine model of<br />

chronic AIH will be significant in the study of liver-specific autoimmunity particularly<br />

with regard to the liver-specific immune regulation by different T cell populations.<br />

Moreover the novel model might open new therapeutic options to treat the disease.<br />

35


Cytochrome P450 2D6 as a model autoantigen<br />

Urs Christen, PhD<br />

Klinikum der Johann Wolfgang Goethe-Universität Frankfurt, <strong>Pharma</strong>zentrum<br />

Frankfurt/ZAFES, Theodor-Stern Kai 10, Frankfurt am Main, Germany<br />

The etiology of autoimmune hepatitis (AIH) is poorly understood although the major<br />

autoantigen, cytochrome P450 2D6 (CYP2D6), has been identified and immunodominant<br />

epitopes mapped. One reason might be the lack of an appropriate animal<br />

model that reflects the pathogenic processes occurring in patients suffering from AIH.<br />

Often experimental hepatitis was reported to be only transient and many current<br />

models for autoimmune liver disease depend on a rather complex disease induction<br />

protocol. Therefore, we generated a mouse model for human AIH using the natural<br />

human autoantigen CYP2D6 as a triggering molecule for autoimmunity. We infected<br />

wildtype FVB mice with an adenovirus expressing the human CYP2D6 (Ad-2D6) to<br />

break self-tolerance to the mouse CYP2D6 homologues. Ad-2D6-infected mice<br />

showed several persistent features characteristic for liver damage associated with<br />

AIH. These features included massive hepatic fibrosis, ‘fused’ liver lobules,<br />

disorganized architecture, cellular infiltrations and focal to confluent necrosis.<br />

Further, all Ad-2D6-infected mice generated high titers of anti-CYP2D6 antibodies.<br />

Epitope mapping revealed that such anti-CYP2D6 antibodies predominantly<br />

recognized the identical immunodominant linear epitope WDPAQPPRD that is<br />

recognized by LKM-1 antibodies from AIH patients. In contrast, mice infected with a<br />

control adenovirus (Ad-C) did neither develop liver damage nor generated anti-<br />

CYP2D6 antibodies. Interestingly, Ad-2D6-infection of transgenic mice expressing<br />

the human CYP2D6 in the liver (CYP2D6 mice) resulted in a delayed kinetics and a<br />

reduced severity of liver damage. Antibody formation was only moderately reduced in<br />

transgenic CYP2D6 mice compared to wildtype FVB mice. Further, anti-CYP2D6<br />

antibodies generated in CYP2D6 mice did react to the major B-cell epitope<br />

WDPAQPPRD indicating that B cell tolerance induction does not account for the<br />

observed differences in disease kinetics and severity between wildtype and CYP2D6transgenic<br />

mice.<br />

We analyzed the T cell mediated immune response by stimulating lymphocytes with<br />

61 overlapping 20-mer peptides covering the entire human CYP2D6. T cells<br />

produced IFNγ after stimulation with peptides only if isolated from Ad-2D6-infected<br />

mice. The frequency of CYP2D6-specific T cells was more then three times higher in<br />

lymphocytes isolated from the liver then from the spleen indicating either an organ<br />

specific proliferation or a selective retention process in the liver. Importantly, the<br />

frequency of CYP2D6-specific CD4 as well as CD8 cells was dramatically decreased<br />

in transgenic CYP2D6 mice indicating the presence of a strong T cell tolerance to<br />

human CYP2D6 established in transgenic CYP2D6 expressing the identical target<br />

antigen compared to wildtype FVB mice expressing the mouse homologues only.<br />

T cell epitope mapping revealed that CYP2D6-specific T cells reacted to human<br />

CYP2D6 peptides with intermediate homology to the mouse homologues but not to<br />

those with high homology or identity. Our data indicate that molecular mimicry rather<br />

than molecular identity breaks tolerance on both the B cell and T cells level<br />

subsequently causing severe and persistent autoimmune liver damage.<br />

36


In summary, the CYP2D6 mouse model for autoimmune hepatitis uses a true human<br />

autoantigen, which may be important for establishing a model that comes close to the<br />

immunopathogenesis of autoimmune liver disease as observed in patients with AIH.<br />

The CYP2D6 model should therefore provide a platform to investigate mechanisms<br />

involved in the immunopathogenesis of autoimmune mediated chronic hepatic injury<br />

as seen in human AIH and to evaluate possible ways of therapeutic interference.<br />

37


Spontaneous chronic liver inflammation in a new transgenic<br />

mouse model of autoimmune hepatitis: Impact and regulation<br />

of autoreactive CD8 + T cells<br />

<strong>Dr</strong>. Mario Zierden<br />

Institute for Pathology, University Hospital of Cologne, Kerpener Str. 62,<br />

50924 Cologne, Germany<br />

Autoimmune hepatitis (AIH) is a chronic inflammatory liver disease of unknown<br />

etiology. Autoreactive T cells are believed to initiate and orchestrate the destruction<br />

of hepatocytes but the pathogenesis of AIH remains poorly understood.<br />

To investigate liver-specific CD8 + T-cell responses as well as the impact and<br />

regulation of autoreactive CD8 + T cells in the pathogenesis of AIH, we have<br />

generated transgenic mice that express the influenza virus hemagglutinin (HA) model<br />

autoantigen under the control of albumin regulatory elements exclusively in the liver<br />

(Alb-HA mice), affording presentation of the neoself-antigen by hepatocytes and by<br />

professional APCs. Adoptive transfer of HA-specific naïve CD8 + T cells into Alb-HA<br />

mice leads to an activation of autoreactive CD8 + T cells and elicits transient hepatitis.<br />

Alb-HA/CL4-TCR double transgenic mice, in which the majority of CD8 + T cells<br />

express a HA-specific TCR, spontaneously develop persistent autoimmune-mediated<br />

hepatitis characterized by moderate liver inflammation and elevated alanine<br />

aminotransferase level. Liver infiltrates are dominated by HA-specific CD8 + T cells<br />

that display diminished TCR expression, no particular signs of apoptosis induction,<br />

reduced proliferative capabilities and only little effector function upon restimulation<br />

with HA peptide, compatible with the state of T-cell anergy. Moreover, liver infiltrates<br />

are enriched for CD4 + /CD25 + /FoxP3 + regulatory T cells that are potent suppressors<br />

of CD8 + T-cell activation in vitro, suggesting that regulatory CD4 + T cells contribute to<br />

the suppression of a broad HA-specific autoimmune response.<br />

Our results show that even if many autoreactive liver-specific CD8 + T cells are<br />

constantly generated that elicit chronic hepatitis, peripheral tolerance mechanisms<br />

like the induction of T-cell anergy and the accumulation of regulatory CD4 + T cells<br />

protect the liver under normal conditions from fatal injury.<br />

38


Induction of tolerance in a mouse model of inflammatory liver<br />

disease<br />

Gisa Tiegs<br />

Division of Experimental Immunology and Hepatology, University Medical Center<br />

Hamburg-Eppendorf, Hamburg<br />

We have recently observed in a model of immune-mediated liver injury inducible by<br />

the T cell mitogenic lectin concanavalin A (ConA) that C57BL/6 mice developed long<br />

term tolerance towards a second ConA injection within 8 days after the first<br />

application which lasted for at least 6 weeks. Immune-mediated liver injury in this<br />

model depends on activation of CD4 + T cells, NKT cells, and Kupffer cells and is<br />

mediated by T cell derived IFNγ and Kupffer cell derived TNFα whereas IL-10 is<br />

protective. The immunopathology of the ConA model shares certain features of<br />

autoimmune hepatitis, i.e. relevance of CD4 + T cells, genetic prevalence regarding<br />

haplotype-specificity and susceptibility, good responsiveness to immunosuppressive<br />

drugs, and immunosuppression (tolerance) in state of remission.<br />

Induction of tolerance in the liver was characterized by significantly reduced serum<br />

transaminase activities and improvement of liver histopathology as well as by an antiinflammatory<br />

cytokine milieu, i.e. down-modulation of IFNγ, TNFα, IL-6, IL-17 and<br />

IL-2 production and a concomitant increase of IL-10 release. CD4 + CD25 + Tregs as<br />

well as Kupffer cells contributed to IL-10 release in tolerant mice, whereby the<br />

Kupffer cells might have been differentiated into alternatively activated macrophages.<br />

Experiments using IL-10 -/- mice argued for a critical role of IL-10 for induction of<br />

tolerance. The tolerogenic effect was reversed upon in vivo depletion of<br />

CD4 + CD25 + FoxP3 + Tregs in DEREG (“depletion of regulatory T cell“) mice prior to restimulation,<br />

and CD4 + CD25 + FoxP3 + Tregs from ConA-tolerant mice produced more<br />

IL-10 and displayed a higher immunosuppressive potential in vitro and following<br />

adoptive transfer in vivo compared to those from non-pretreated animals.<br />

Frequencies of PD-1 + CCR5 + CXCR3 + Treg cells where higher and those of CD62L +<br />

Tregs were lower in adoptively transferred Tregs from tolerant mice compared to<br />

naïve Tregs from control animals. Tolerance was not inducible in PD-1 -/- or CCR5 -/- or<br />

CXCR3 -/- mice. Interestingly, Tregs from ConA pretreated CCR5 -/- and CXCR3 -/- mice<br />

still retained their immunosuppressive activity in an in vitro suppression assay. These<br />

results suggest that ConA tolerance is mediated by PD-1 expressing Tregs trafficking<br />

into the liver in dependence of the IFNγ-inducible chemokine receptors CCR5 and<br />

CXCR3. Hence, blockade of receptors by low molecular weight antagonists might<br />

have detrimental consequences with respect to regulation of an inflammatory<br />

response in the liver.<br />

39


Session IV<br />

Autoimmunity II<br />

41


The role of the liver for the induction of oral tolerance in<br />

presence of naïve and antigen-experienced T cells<br />

Nils Kruse, Katrin Neumann, Arnhild Schrage, Katharina Eulenburg, Martin Zeitz,<br />

Alf Hamann and Katja Klugewitz<br />

Medizinische Klinik I und Research Center Immunosciences, Charité Campus<br />

Benjamin Franklin, Berlin<br />

Experimentelle Rheumatologie, Charité Campus Mitte, Berlin<br />

Within the liver the local immune balance is shifted to tolerance rather than immunity,<br />

although the organ can also build up a sufficient immune reaction against for<br />

example viral hepatitis. Under physiological conditions, the portal blood flow<br />

transports various food-derived or bacterial antigens from the intestine into the liver.<br />

The observation that oral antigen uptake can induce antigen-specific unresponsiveness,<br />

described already many years ago as “oral tolerance”, has raised the question<br />

whether the liver might also participate in the induction of oral tolerance. The finding<br />

that a portocaval shunt can abrogate oral tolerance in an animal model has enforced<br />

this assumption.<br />

In the healthy liver several cell populations, such as liver dendritic cells, Kupffer cells,<br />

Ito cells and liver sinusoidal endothelial cells (LSEC) constitutively express MHCII,<br />

suggesting the potential to function as antigen-presenting cells and, thus, contribute<br />

to the induction of antigen-specific tolerance. In particular LSEC are believed to shift<br />

the hepatic immune balance towards systemic unresponsiveness: LSEC can crosspresent<br />

MHCI antigens, thereby inducing anergy in naïve CD8 + T cells and oral<br />

tolerance.<br />

Oral antigens are predominantly MHCII-restricted raising the question as to the<br />

intrahepatic modulation of CD4 + T cells upon antigen feeding. Following oral antigen<br />

application, naïve CD4 + T cells undergo apoptosis or acquire a regulatory phenotype<br />

within the liver. However, under physiological conditions only very few naïve CD4 + T<br />

cells enter the liver and, secondly, mesenteric lymph nodes have been demonstrated<br />

to be obligatory for the induction of oral tolerance creating doubts whether tolerance<br />

is locally imposed. To exclude that CD4 + T cells modulated elsewhere are recruited<br />

into the liver, liver-restricted antigen presentation models are required. In a bone<br />

marrow transplantation model that ensures MHCII-restricted antigen presentation<br />

exclusively by LSEC local antigen presentation was not sufficient to establish oral<br />

tolerance in the presence of naïve CD4 + T cells whereas Th1 effector cells that are<br />

predominantly recruited into the liver became deleted.<br />

In conclusion, all these findings point to a role of the liver for oral tolerance by<br />

modulation of CD8 + and antigen-experienced CD4 + T cells populations whereas<br />

under healthy conditions the mesenteric lymph nodes or at least professional liver<br />

APC seem to be mandatory for the (tolerogenic) priming of naïve CD4 + T cells.<br />

43


Therapeutic options to treat AIH in 2009<br />

Christian P. Strassburg<br />

Clinic for Gastroenterology, Hepatology and Endocrinology, Hannover Medical<br />

School, Hannover, Germany<br />

The indication for treatment of AIH is based on inflammatory activity and not so much<br />

on the presence of cirrhosis. In the absence of inflammatory activity immunosuppressive<br />

treatment has only limited effects. An indication for treatment is present<br />

when aminotransferase levels are elvated and thus hepatic inflammation has to be<br />

assumed. Usually IgG is also elevated. Liver biopsy should be performed in every<br />

patient who is considered for therapy and the presence of interface hepatitis<br />

underscores the decision for treatment. Biopsies will also help for the determination<br />

of differential diagnoses and for grading and staging<br />

Independent of clinically or immunoserologically defined type of AIH, treatment is<br />

usually implemented with predniso(lo)ne alone or in combination with azathioprine.<br />

Both strategies are equally effective and this strategy is based on trials that were<br />

performed more tha 2 decades ago and has since then essentially remained<br />

unchanged. The use of prednisone or its metabolite prednisolone is equally effective<br />

since chronic liver disease does not seem to have an effect on the synthesis of<br />

prednisolone from prednisone. Important is the exact differentiation between virus<br />

infection and autoimmune hepatitis. Treatment of replicative viral hepatitis with<br />

corticosteroids must be prevented as well as administration of interferon in AIH, in<br />

which it can lead to dramatic disease exacerbation.<br />

Therapy is usually administered over the course of 2 years. The decision between<br />

monotherapy and combination therapy is guided by principle considerations: Long<br />

term steroid therapy leads to cushingoid side effects. Especially cosmetic side effects<br />

decrease patient compliance considerably. Serious complications such as steroid<br />

diabetes, osteopenia, aseptic bone necrosis, psychiatric symptoms, and<br />

hypertension and cataract formation also have to be anticipated in long term<br />

treatment. Side effects are present in 44% of patients after 12 months and in 80% of<br />

patients after 24 months of treatment. However, predniso(lo)ne montherapy is<br />

possible in pregnant patients. Azathioprine, on the other hand, leads to a decreased<br />

dose of prednisone. It bears a theoretical risk of teratogenicity. In addition, abdominal<br />

discomforts, nausea, cholestatic hepatitis, rashes and leukopenia can be<br />

encountered. These side effects are seen in 10% of patients receiving a dose of 50<br />

mg per day. The side effect profile suggests that there is a need for less side effect<br />

prone future strategies for maintenance therapy.<br />

The consequent administration of immunosuppression is essential since most cases<br />

of relapse are the result of erratic changes of medication and/or dose. Dose<br />

reduction is aimed at finding the maintenance dose. Since histology lags 3 to 6<br />

months behind the normalization of serum parameters therapy has to be continued<br />

beyond the normalization of aminotransferase levels. Usually, maintenance doses of<br />

prednisone range between 10 to 2.5 mg. After 12–24 months of therapy<br />

predniso(lo)ne can be tapered over the course of 4–6 weeks to test whether a<br />

sustained remission has been achieved. Tapering regimens should be attempted<br />

with great caution and only after obtaining a liver biopsy which demonstrates a<br />

complete resolution of inflammatory activity. Relapse of AIH and the risk of<br />

progression to fibrosis is almost universal when immunosuppression is tapered in the<br />

presence of residual histological inflammation.<br />

44


Outcomes can be classified into four categories: remission, relapse, treatment failure<br />

and stabilization.<br />

Remission is a complete normalization of all inflammatory parameters including<br />

histology. This is achieved in 65% of patients after 24 months of treatment.<br />

Remission can be sustained with azathioprine monotherapy of 2 mg/kg bodyweight.<br />

This prevents cushingoid side effects. However, side effects such as arthralgia<br />

(53%), myalgia (14%), lymphopenia (57%) and myelosuppression (6%) have been<br />

observed.<br />

Relapse is characterized by a renewed increase of aminotransferase levels and the<br />

re-occurrence of clinical symptoms. Relapse is present in 50% of patients within<br />

6 months of treatment withdrawal and in 80% after 3 years. Relapse is associated<br />

with progression to cirrhosis in 38% and liver failure in 14%. Occurrence of a relapse<br />

calls for re-initiation of standard therapy and perhaps a long term maintenance dose<br />

with predniso(lo)ne or azathioprine monotherapy.<br />

Treatment failure characterizes a progression of clinical, serological and histological<br />

parameters during standard therapy. This is seen in about 10% of patients. In these<br />

cases the diagnosis of AIH has to be carefully reconsidered to exclude other<br />

etiologies of chronic hepatitis. In these patients experimental regimens can be<br />

administered or ultimately liver transplantation becomes necessary.<br />

Stabilization is the achievement of a partial remission. Since 90% of patients reach<br />

remission within 3 years, the benefit of standard therapy has to be re-evaluated in<br />

this subgroup of patients. Ultimately, liver transplantation provides a definitive<br />

treatment option.<br />

If standard treatment fails or drug intolerance occurs, alternative therapies such as<br />

cyclosporine, tacrolimus, cyclophosphamide, mycophenolat mofetil, rapamycin,<br />

UDCA, and budendoside can be considered. The efficacy of these options has not<br />

yet been definitively decided.<br />

Deflazacort has been proposed as an alternative corticosteroid for immunosuppression<br />

with fewer side effects than conventional glucocorticoids. In a recent<br />

study, 15 patients with AIH type I were treated with Deflazacort, which previously had<br />

been treated with prednisone with or without azathioprine until biochemical remission<br />

was obtained. Remission was sustained during 2 years of follow-up. However, the<br />

long-term role of second generation corticosteroids to sustain remission in AIH<br />

patients with reduced treatment related side effects requires further controlled<br />

studies.<br />

Cyclosporine A (CyA) is a lipophylic cyclic peptide of 11 residues produced by<br />

Tolypocladium inflatum that acts on calcium dependent signaling and inhibits T cell<br />

function via the interleukin 2 gene. Of all alternative agents, the greatest experience<br />

to date has been with CyA. In these studies CyA was successfully used for the<br />

treatment of AIH and was well tolerated. The principal difficulty in advocating<br />

widespread use of the drug as first-line therapy relates to its toxicity profile,<br />

particularly with long-term use (increased risk of hypertension, renal insufficiency,<br />

hyperlipidemia, hirsutism, infection, and malignancy)<br />

45


Tacrolimus is a macrolide lactone compound with immunosuppressive capabilities<br />

exceeding those of cyA. The mechanism of action is similar to that of CyA but it binds<br />

to a different immunophilin. The application of Tacrolimus in 21 patients treated for<br />

1 year lead to an improvement of aminotransferase and bilirubin levels with a minor<br />

increase in serum BUN and creatinine levels. Although Tacrolimus represents a<br />

promising immunosuppressive candidate drug, larger randomized trials are required<br />

to assess its role in the therapy of AIH.<br />

Mycophenolate has attracted attention as a transplant immunosuppressant with an<br />

important role in the steroid free immunosuppressive therapy of patients transplanted<br />

for chronic hepatitis C infection. Mycophenolate is a noncompetitive inhibitor of<br />

inosine monophosphate dehydrogenase, which blocks the rate-limiting enzymatic<br />

step in de novo purine synthesis. Mycophenolate has a selective action on<br />

lymphocyte activation, with marked reduction of both T and B lymphocyte<br />

proliferation. In a recent pilot study 7 patients with AIH type 1 who either did not<br />

tolerate azathioprine or did not respond to standard therapy with a complete<br />

normalization of aminotransferase levels were treated with Mycophenolate in addition<br />

to steroids. In 5 out of 7 patients normalization of aminotransferase levels was<br />

achieved within 3 months. These preliminary data suggest that Mycophenolate may<br />

represent another promising treatment strategy of AIH. However, a recent<br />

retrospective study on mycophenolate showed that many patients do not reach<br />

remission and is less optimistic regarding the role of mycophenolate.<br />

The induction of remission with 1–1.5 mg per kg per day of cyclophosphamide in<br />

combination with steroids has been reported. However the dependency of continued<br />

application of cyclophosphamide with its potentially severe hematological side effects<br />

renders it a highly experimental treatment option.<br />

Ursodeoxycholic acid is a hydrophilic bile acid with putative immunomodulatory<br />

capabilities. It is presumed to alter HLA class I antigen expression on cellular<br />

surfaces and to suppress immunoglobulin production. Uncontrolled trials have shown<br />

a reduction in histological abnormalities, clinical and biochemical improvement but<br />

not a reduction of fibrosis in 4 patients with AIH type 1. However, its role in AIH<br />

therapy or in combination with immunosuppressive therapy is still unclear.<br />

Budesonide is a synthetic steroid with high first-pass metabolism in the liver, which<br />

should limit systemic side effects compared to conventional steroids. Recently the<br />

first multicenter randomized trial testing its efficacy in inducing and maintaining<br />

remission in non-cirrhotic patients with AIH was completed. This study was<br />

undertaken based on previous reports of budesonide in AIH. In a study treating<br />

13 AIH patients with budesonide over a period of 9 months the drug was well<br />

tolerated and aminotransferase levels were normalized. Our own experiences have<br />

confirmed that budesonide is effective but does not offer an advantage over<br />

conventional steroids when cirrhosis and porto-systemic shunts are present and that<br />

induction of remission can be reached in treatment naïve patients with AIH. However,<br />

another study studying difficult to treat AIH patients with previous therapy showed<br />

that budesonide therapy was associated with a low frequency of remission and high<br />

occurrence of side effects. The main advantage of budesonide for the future<br />

treatment of autoimmune hepatitis may be to replace prednisone in long-term<br />

maintenance therapy to reduce steroid side effects. This was now shown in the<br />

46


aforementioned large trial with 207 patients from 30 centers in Europe and Israel.<br />

Budesonide in combination with azathioprine was able to induce complete remission<br />

and was tested against prednisone in combination with azathioprine in a 6 month<br />

treatment schedule followed by an open label phase in which prednisone patients<br />

were switched to budesonide. In this study a stringent definition of remission was<br />

employed that required a complete normalization of aminotransferase activities in<br />

addition to the absence of steroid-specific side effects. Complete responses were<br />

reached in the budesonide arm at 12 months in 60.2% (prednisone 49.4%) and<br />

biochemical remission in 68% (prednisone 50.6%). In patients who were switched<br />

from prednisone to budesonide steroid-specific side effects were reduced from<br />

40.2% to 18.4%. These data show in one of the largest randomized placebo<br />

controlled trials that budesonide is effective, leads to less side effects and represents<br />

an alternative for the future of treatment of AIH.<br />

Liver Transplantation in AIH<br />

In approximately 10% of AIH patients liver transplantation remains as the only life<br />

saving option. The indication for liver transplantation in AIH is similar to that in other<br />

chronic liver diseases and includes clinical deterioration, development of cirrhosis,<br />

bleeding esophageal varices and coagulation abnormalities despite adequate<br />

immunosuppressive therapy. There is no single indicator or predictor for the<br />

necessity of liver transplantation. Candidates for liver transplantation are usually<br />

patients who do not reach remission within 4 years of continuous therapy. Indicators<br />

of a high mortality associated with liver failure are histological evidence of<br />

multilobular necrosis and progressive hyperbilirubinemia. In Europe, 4% of liver<br />

transplants are for AIH. The long term results of liver transplantation for AIH are<br />

excellent. The 5-year survival is up to 92% and well within the range of other<br />

indications for liver transplantation.<br />

The potential of AIH to recur after liver transplantation has been a matter of<br />

controversial debate but are now generally accepted. It is a major reason for graft<br />

dysfunction and loss after OLT. The first case of recurrent AIH after liver transplantation<br />

was reported in 1984, and was based upon serum biochemistry, biopsy<br />

findings and steroid reduction. Studies published during the past years indicate that<br />

the rate of recurrence of AIH ranges between 10–35%, and that the risk of AIH<br />

recurrence is perhaps as high as 68% after 5 years of follow-up. It is important to<br />

consider the criteria upon which the diagnosis of recurrent AIH is based. When<br />

transaminitis is chosen as a practical selection parameter many patients with mild<br />

histological evidence of recurrent AIH may be missed. It is therefore suggested that<br />

all patients with suspected recurrence of autoimmune hepatitis receive a liver biopsy,<br />

biochemical analyses of aminotransferases as well as a determination of immunoglobulins<br />

and autoantibody titers. Significant risk factors for the recurrence of AIH<br />

have not yet been identified although it appeared that the presence of fulminant<br />

hepatic failure before transplantation protected against the development of recurrent<br />

disease. An attractive risk factor for the development of recurrent AIH is the presence<br />

of specific HLA antigens that may predispose toward a more severe immunoreactivity.<br />

In two studies recurrence of AIH appeared to occur more frequently in HLA<br />

DR3 positive patients receiving HLA DR3 negative grafts. However, this association<br />

was not confirmed in all studies. Interestingly, there have not been conclusive data to<br />

support the hypothesis that a specific immunosuppressive regimen represents a risk<br />

47


factor for the development of recurrent AIH. However, data indicate, that patients<br />

transplanted for AIH require continued steroids in 64% versus 17% of patients<br />

receiving liver transplants for other conditions. Based on these results and other<br />

studies it would appear that maintenance of steroid medication in AIH patients is<br />

indicated to prevent not only cellular rejection but also graft threatening recurrence of<br />

AIH but this is controversial. In addition to AIH recurrence the development of de<br />

novo autoimmune hepatitis after liver transplantation has been reported.<br />

48


New insights into autoimmune cholangitis by animal models<br />

Michael Trauner, MD<br />

Division of Gastroenterology and Hepatology, Laboratory of Experimental and<br />

Molecular Hepatology, Department of Internal Medicine, Medical University of Graz,<br />

8036 Graz, Austria; E-Mail: michael.trauner@meduni-graz.at<br />

Improving our understanding of the pathogenesis of chronic immune-mediated<br />

cholangiopathies such primary biliary cirrhosis (PBC) and primary sclerosing<br />

cholangitis (PSC), as well as the development of novel diagnostic, prognostic and<br />

therapeutic tools for these disorders critically depends on easily reproducible animal<br />

models. Recently, several spontaneous mouse models for PBC (not requiring<br />

previous manipulations for breakdown of immunotolerance to pyruvate dehydrogenase<br />

(PDC)-E2 protein) have been reported including NOD.c3c4 and NOD.c3c4derived<br />

mice, IL-2Rα -/- mice, dominant negative TGF-β receptor II mice and Ae2a,b -/mice.<br />

Interestingly, a common feature of all “spontaneous” PBC mouse models<br />

appears to be a relative reduction of circulating naturally T regs, indicating that<br />

disturbed T reg function may play a key role in the pathogenesis of autoimmune<br />

diseases such as PBC through loss of self tolerance. Ae2a,b -/- mice show<br />

immunologic, and, at least in part, biochemical and hepatobiliary morphological<br />

features resembling PBC in humans. The animals show a significantly reduced<br />

CD4 + /CD8 + ratio as a consequence of CD8 expansion and a reduced number of<br />

natural T regs. Moreover, most mice spontaneously developed increased serum IgG,<br />

IgM, and AMA These finding suggest that Ae2 deficiency in mice alters intracellular<br />

pH homeostasis in immunocytes and cholangiocytes consequently leading to<br />

characteristic immunologic and hepatobiliary changes also observed in PBC patients.<br />

Notably genetic AE2 (SLC4A2) variants may determine disease susceptibility,<br />

progression and/or response to therapy in PBC patients. Several studies found<br />

reduced expression and function of AE2 (a Cl - /HCO3 exchanger) in PBC, which may<br />

contribute to reduced bile flow and cholestasis. Moreover, decreased AE2 expression<br />

in salivary and lacrimal glands could explain the frequently associated sicca<br />

syndrome in these patients and pointed towards a more generalized “glandular<br />

failure” in PBC. UDCA restores impaired AE2 function in patients with PBC and<br />

recent findings suggest that a synergistic effects with glucocorticoids.<br />

To date, no animal model has been developed that exhibits all of the attributes of<br />

PSC. Rodent models induced by bacterial cell components or colitis may help to<br />

explain the strong association between PSC and inflammatory bowel disease (IBD).<br />

Other models include direct injury to biliary epithelia, peribiliary vascular endothelia or<br />

portal venous endothelia. Mice with targeted disruption of the Mdr2 (Abcb4) gene<br />

encoding a canalicular phospholipid flippase (Mdr2 -/- mice) spontaneously develop<br />

sclerosing cholangitis with macroscopic and microscopic features of human PSC.<br />

Bile duct injury in these mice is linked to defective biliary phospholipid secretion<br />

resulting in an increased concentration of free non-micellar bile acids which<br />

subsequently cause bile duct epithelial cell (cholangiocyte) injury, pericholangitis,<br />

periductal fibrosis with ductular proliferation and finally sclerosing cholangitis.<br />

In analogy to the Mdr2 -/- mouse model of sclerosing cholangitis, MDR3/ABCB4<br />

(the human orthologue of rodent Mdr2/Abcb4) defects could play a role in the pathogenesis<br />

of various cholangiopathies in humans. MDR3 variants could play a role as<br />

49


modifyer gene in the pathogenesis of various cholangiopathies such as PSC, PBC<br />

and adulthood idiopathic ductopenia/biliary fibrosis.<br />

Another example for a transport defect associated with sclerosing cholangitis is cystic<br />

fibrosis (CF) due to mutations of the cystic fibrosis transmembrane conductance<br />

regulator (CFTR/ABCC7) gene resulting in impaired hydration and alkalinization of<br />

bile, inspissated bile and bile duct injury with sclerosing cholangitis in 4 to 18% of<br />

adult CF patients. Cftr -/- mice develop progressive liver disease with hepatosteatosis,<br />

focal cholangitis, inspissated secretions, bile duct proliferation and progression to<br />

focal biliary cirrhosis after 1 year of age. Interestingly, induction of colitis in Cftr -/- mice<br />

aggravates bile duct injury. Colonic inflammation may impair cftr expression via<br />

impairment of PPARα expression, which has also been implicated in the transcriptional<br />

regulation of Mdr2. Also, patients with IBD who are (heterozygous) carriers of<br />

CFTR mutations may be at increased risk of developing PSC. Recent studies<br />

however suggest that CFTR variants may protect (instead of predispose) to<br />

sclerosing cholangitis via modulation of pathogen internalization resulting in reduced<br />

for inflammation and damage to cholangiocytes.<br />

Xenobiotics and drugs may also lead to bile duct injury and biliary fibrosis via direct<br />

toixc and indirect immune-mediated injury. Toxic drug metabolites excreted into bile<br />

may cause a drug-induced vanishing bile duct syndrome and eventually biliary<br />

cirrhosis. Feeding of 3,5-diethoxycarbonyl-1,4-dihydrocollidine (DDC) leads to<br />

increased biliary porphyrin secretion, induction of VCAM, osteopontin, and TNF-α<br />

expression, a pronounced pericholangitis with significantly increased number of<br />

CD11b-positive cells, ductular reaction and activation of periductal myofibroblasts<br />

leading to a biliary fibrosis. Lithocholic acid (LCA) causes bile duct injury, since LCA<br />

per se is highly hydrophobic and toxic. As such, LCA leads to segmental bile duct<br />

obstruction, destructive cholangitis, periductal fibrosis, all features of sclerosing<br />

cholangitis.<br />

In summary, major progress has been made with the identification of spontanous<br />

cholangiopathy models. These novel models should enhance our understanding of<br />

the pathogenesis of PBC and PSC and will hopefully result in improved treatment of<br />

these disorders.<br />

50


Etiological and molecular issues in primary biliary cirrhosis<br />

M. Eric Gershwin, MD<br />

Division of Rheumatology, Allergy and Clinical Immunology, University of California<br />

at Davis School of Medicine, 451 Health Sciences <strong>Dr</strong>ive, Suite 6510, Davis, CA<br />

95616, USA; Telephone: 530-752-2884; Fax: 530-752-4669<br />

E-Mail: megershwin@ucdavis.edu<br />

There have been significant advances both in humans and experimental models that<br />

relate to the etiopathogenesis of primary biliary cirrhosis. Many of these advances<br />

are based on the rigorous definition of the antimitochondrial response, the serologic<br />

signature of PBC. First, it is well established that AMA are directed against members<br />

of the 2-oxoacid dehydrogenase complexes (2-OADC), among which the major<br />

epitopes are within the lipoylated domains of the E2 subunit of the pyruvate<br />

dehydrogenase complex (PDC-E2). Second, autoreactive CD4 + and CD8 + T cells<br />

can be detected in PBC peripheral blood, regardless of the AMA status, and the<br />

infiltration of autoreactive T cells in the liver and periductular spaces is one of the<br />

most prominent immune features. Autoreactive T cells of both subtypes recognize<br />

PDC-E2 sequences overlapping with the AMA epitopes. An increase in cytotoxic<br />

T cell precursors in the blood in the early stages of the disease compared to the<br />

advanced ones and a 10-fold increase of specific liver CD8 + T cells compared to<br />

peripheral blood have been demonstrated. Third, additional data on the immunobiology<br />

components of PBC autoimmunity has been recently obtained in<br />

CD4 + CD25 high natural regulatory T cells which appear to be numerically reduced in<br />

PBC. PBC bile duct cells manifest unique features during apoptosis while co-culture<br />

experiments do not support a direct role for these cells in determining their immunemediated<br />

injury. Apoptotic cells are phagocytosed by BECs and consequently are an<br />

exogenous source of autoantigens in cholangiocytes, possibly through anti-CD16. As<br />

a result, the impact of putative changes in apoptosis and autophagy specific to BEC<br />

remains to be fully determined in PBC. Fifth, the innate immune compartment has<br />

been recently investigated in PBC with promising results. PBC monocytes manifest<br />

an increased response to pathogen associated stimuli, as indicated by higher levels<br />

of pro-inflammatory cytokines. Further, the hyper-IgM associated with PBC is<br />

secondary to an aberrant innate immune response, potentially induced by stimulation<br />

of toll like receptor 9 by bacterial CpG-B.<br />

The female preponderance may hold an important key to PBC etiology. X-linked<br />

genes determine gender-related characteristics at different levels while also<br />

regulating the immune function, particulalry to maintain tolerance. Major X chromosome<br />

defects such as those leading to Turner’s syndrome or premature ovarian<br />

failure are commonly characterized by autoimmune comorbidities (particularly thyroid<br />

disease) and, less frequently, cholestasis. Our group first determined a significantly<br />

higher frequency of monosomy of the X chromosome in peripheral leukocytes<br />

(particularly those of the adaptive immune response, i.e. T and B cells) in women<br />

PBC compared to age-matched control women. Monosomy frequency correlated with<br />

age in all three groups, as expected but monosomic cells were not microchimeric<br />

cells. We further demonstrated that the X loss in PBC affected was not random but<br />

affected more frequently one parentally-inherited chromosome.<br />

51


Several key animal models of autoimmune cholangitis have now been described.<br />

First, a genomic variant of the non obese diabetic (NOD) mouse (NOD.c3c4) has<br />

been observed to manifest autoimmune cholestasis with AMA and ANA positivities in<br />

50–60% and 80–90%, respectively. Liver histology demonstrated portal lymphocyte<br />

infiltration with chronic non-suppurative cholangitis and PBC-like granulomas.<br />

Second, a dominant negative form of trasforming growth factor β (TGFβ) receptor II<br />

(dnTGFβRII) mouse develop serum AMA in 100% of mice. The TGFβ receptor II<br />

regulates lymphocyte activation and the appearance of PBC in this model suggests a<br />

specific condition of T cells with impaired TGFβ signaling in the presence or absence<br />

of B cells is involved. Third, the knockout of interleukin 2 receptor α leads to a murine<br />

phenotype with 100% serum AMA positivity, 80% serum ANA positivity, and portal<br />

lymphocyte infiltration and vanishing bile ducts. This model is of particular interest<br />

based on the report of autoimmune cholangitis in a pediatric case of IL2Rα<br />

deficiency. Fourth, Ae2a,b also develop autoimmune phenomenon and a PBC-like<br />

disease. Finally, immunization of mice with chemical xenobiotics has also been<br />

shown to lead to a PBC-like disease.<br />

These data and observations will be put in the context of the key mechanisms that<br />

lead to autoimmune cholangitis in patients.<br />

52


Session V<br />

Immunology of viral hepatitis I<br />

53


Occult hepatitis B virus infection: Diagnosis and significance<br />

Wolfram H. Gerlich<br />

Institute of Medical Virology, Justus-Liebig University Giessen, Germany<br />

An expert group defined March 2008 at Taormina "Occult hepatitis B virus infection"<br />

(OBI) as: "Presence of HBV DNA in the liver (with detectable or undetectable HBV<br />

DNA in the serum) of individuals testing HBsAg negative with currently available<br />

assays. When detectable, the amount of HBV DNA in the serum is usually very low<br />

(< 200 IU/ml)". This definition implies OBI-like phases even in HBsAg positive cases,<br />

because the early “window phase” of acute HBV infection may last several weeks<br />

before HBsAg becomes detectable. We identified two blood donors whose donations<br />

were HBV DNA negative (< 12.5 IU/ml) but transmitted HBV to the recipient. Both<br />

developed later HBsAg acute hepatitis. However, OBI without HBsAg positive phase<br />

occurs also. We followed an acute OBI which showed a peak viremia of 15,000 IU/ml<br />

HBV DNA and sub-borderline HBsAg-levels suggesting a ratio of virions to subviral<br />

particles of 1:10 whereas "normal" cases show at peak viremia 1:1000 or less. In that<br />

case a rapid immune response had down-regulated the viral gene expression.<br />

It is known since long that anti-HBc positive persons without HBsAg may reactivate<br />

HBV infection if they experience immunosuppression and that livers from HBsAg<br />

negative donors may transmit HBV to transplant recipients. Studies from Italy show<br />

that the majority of anti-HBc positive healthy individuals have HBV DNA in the liver.<br />

But one third with HBV DNA had no anti-HBc. On that basis one may estimate that<br />

ca. 7% of the German population may have OBI.<br />

Controversial data exist on the impact of persistent OBI on development of cirrhosis<br />

or HCC and on the course and therapy of HCV- or HIV-coinfection. Consensus exists<br />

that OBI may reactivate under severe immunosuppressiuon and should be searched<br />

for before starting this therapy. Furthermore, blood donors with OBI may transmit<br />

HBV. We studied 5 blood donors with OBI and 55 of their recipients. In 22 cases,<br />

transmission was probably but the recipients remained healthy. However, in 3 cases<br />

who were immunosuppressed at the time of transfusion, fatal fulminant hepatitis B<br />

developed. OBI leads often to selection of HBsAg escape mutations. Inversely,<br />

infection of vaccinated individuals favors development of OBI which may be difficult<br />

to recognise in blood donors. Thanks to HB vaccination, the problem of persistent<br />

overt HBV infection may eventually fade away but the problem of persistent OBI will<br />

possibly remain for longer.<br />

55


Innate and adaptive immune responses in HBV infection<br />

Mala K. Maini<br />

Division of Infection and Immunity, University College of London, London, UK<br />

The hepatotropic hepatitis B virus (HBV) is non-cytopathic; liver disease resulting<br />

from this infection is therefore thought to be immune-mediated. In order to develop<br />

immunotherapeutic strategies for improving the treatment of HBV, there is a pressing<br />

need to dissect out the immune components contributing to viral control versus<br />

disease pathogenesis.<br />

Defects in many aspects of the coordinated innate and adaptive immune response<br />

have been described in patients failing to control HBV infection, but one of the most<br />

profound and critical is depletion of the virus-specific CD8 T cell response. We have<br />

identified Bim-mediated apoptosis as one of the processes accounting for the failure<br />

of HBV-specific CD8 T cells to persist in the face of high antigen load [Lopes et al.,<br />

JCI 2008]. We are investigating whether T cell tolerance through induction of Bim<br />

may be imposed by inappropriate co-stimulation during intrahepatic antigen<br />

presentation and whether blocking this pathway may reconstitute more effective<br />

responses in vivo. We have also characterized signaling defects attributable to the<br />

liver microenvironment (local depletion of arginine), that bias the effector function of<br />

CD8 T cells in chronic HBV infection [Das et al., JEM 2008].<br />

Data from humans and HBV transgenic mice have pointed to the non-virus-specific<br />

lymphocytes infiltrating the liver in the presence of uncontrolled HBV replication as a<br />

major contributor to the ensuing damage. We have identified a pathway whereby NK<br />

cells, which account for 30–40% of intrahepatic lymphocytes, can mediate<br />

hepatocyte death and have shown that this pathway can be stimulated by cytokines<br />

induced during flares of HBV-related liver disease [Dunn et al., JEM 2007]. We have<br />

defined a role for CXCR6/CXCL16 chemokine/chemokine receptor interactions in the<br />

homing of NK cells with pathogenic potential to the HBV-infected liver. These NK<br />

cells have an impaired capacity to produce the antiviral cytokine IFN-γ, which can be<br />

reversed by blocking IL-10 signals. We find that the immunosuppressive cytokine<br />

IL-10 is induced in the early stages of acute HBV infection [Dunn et al., Gastro in<br />

press] and also in flares of chronic infection, and that it is capable of suppressing<br />

both innate and adaptive immune responses.<br />

In summary, features of the intrahepatic cytokine and nutrient microenvironment,<br />

together with aberrant co-regulatory signals, combine with the prolonged high HBV<br />

load to deviate innate and adaptive immune responses towards a pathogenic rather<br />

than antiviral role.<br />

56


Hepatitis Delta: Immunopathogenesis and clinical challenges<br />

Heiner Wedemeyer, Jan Grabowski<br />

Department of Gastroenterology, Hepatology and Endocrinology, Hannover Medical<br />

School, Carl-Neuberg Str. 1, 30625 Hannover, Germany, Tel.: +49 511 532-6814,<br />

Tel.: +49 511 532-8662<br />

Website: www.mh-hannover.de/ag-wedemeyer.html<br />

E-Mail: wedemeyer.heiner@mh-hannover.de<br />

Hepatitis Delta is caused by infection with the hepatitis D virus (HDV) and is<br />

considered as the most severe form of viral hepatitis in humans. Hepatitis Delta<br />

occurs only in HBsAg-positive individuals as HDV is a defective RNA virus which<br />

requires the hepatitis B virus (HBV) surface antigen (HBsAg) for complete replication<br />

and transmission. At least seven different HDV genotypes have been described with<br />

specific geographic distributions and distinct clinical courses. HDV infection affects<br />

mainly immigrants in Central Europe with about one third of patients born in states of<br />

the former Soviet Union and another third born in Turkey. HDV/HBV coinfection can<br />

be associated with complex and dynamic viral dominance patterns. While HDV is<br />

frequently the dominating virus not only in HBV/HDV coinfection but also in<br />

HBV/HCV/HDV triple-infected patients, fluctuating courses HDV and HBV viremia<br />

can be observed in other patients. Chronic HDV infection leads to more severe liver<br />

disease than HBV mono-infection with accelerated fibrosis progression, earlier<br />

hepatic decompensation and an increased risk for the development of hepatocellular<br />

carcinoma. However and in contrast to HCV infection, hepatic decompensation rather<br />

than development of liver cancer is the first clinical endpoint that develops during the<br />

course of infection.<br />

Few studies have investigated immune responses against HDV and HBV in humans.<br />

Our findings suggest that HDV-specific T cell responses are detectable in the<br />

majority of patients, however, so far we were not able to find a clear-cut association<br />

with the stage or grade of liver disease. Of note, HBV-specific can be detected<br />

similarly at a high frequency. So far, only interferon alpha treatment has proven<br />

antiviral activity against HDV in humans and has been linked to improved long-term<br />

outcome. Recent studies on the use of pegylated interferon showed a sustained<br />

virological response concerning HDV-RNA in about one quarter of patients.<br />

HDV-specific immune responses might be associated with the response to treatment.<br />

Novel alternative treatment options including prenylation inhibitors are still awaiting<br />

clinical development for delta hepatitis.<br />

57


Clinical spectrum and therapeutic options in HDV infections<br />

M. Rizzetto<br />

University of Torino, Department of Gastroenterology, Torino, Italy<br />

Clinical experience has confirmed that chronic hepatitis D (CDH) most usually runs a<br />

severe and progressive course; as HBV is inhibited by the HBV, the diagnostic<br />

harbinger to an underlying HDV disease is the discrepancy between a florid HBsAg<br />

positive disease and lack of HBV-DNA in serum.<br />

HDV is divided into eight genotypes differing as much as 40% in nucleotide<br />

sequence. Genotype I is the most frequent world-wide and has variable<br />

pathogenicity. Genotypes II and IV were found in East Asia causing relatively mild<br />

disease. Genotype III was associated with fulminant hepatitis in South America.<br />

Since the 1990s the circulation of HDV has declined significantly in Europe,<br />

decreasing in Italy in chronic HBsAg disease from a rate of 24.6% in 1983 to 8.3% in<br />

1997. The clinical scenario of hepatitis D has also changed; while the majority of<br />

hepatitis D collected in Italy in the 1980s had a chronic active hepatitis, by the end of<br />

the 1990s the proportion of cirrhosis residual to burnt-out inflammation had increased<br />

to 70%.<br />

However, hepatitis D has not disappeared in Europe but is now being restored<br />

through immigration; the reservoir of HDV in Europe is currently sustained by the<br />

residual ageing domestic pool that survived the brunt of the hepatitis D epidemic in<br />

the 1970–1980s and by a population of young patients with recent HDV infections<br />

migrating to Europe from areas where HDV remains endemic.<br />

α-Interferon was first used in CHD 25 years ago and still remains the only available<br />

treatment. Results, however, have been limited.<br />

Peg-IFN may be more beneficial; in a study of CHD treated with Peg-IFN-α2b<br />

1.5 µg/kg weekly for 12 months, 6 of 14 (43%) achieved a sustained clearance of<br />

serum HDV-RNA.<br />

Famciclovir, Lamivudine and Adefovir, inactive against HDV but active against the<br />

helper HBV, were used in monotherapy or in combination with IFN, but they were not<br />

efficacious.<br />

The problem confronting HDV therapy is that there is no specific viral function to<br />

target; HDV depends on the HBsAg and not on HBV replication, therefore its<br />

synthesis is not influenced by the level of HBV-DNA in serum. A novel strategy<br />

involves disruption of prenylation to prevent the binding of the HD-Ag to HBsAg;<br />

in-vitro inhibition of HDAg prenylation by the prenylation inhibitor BZA-5B and in-vivo<br />

by the inhibitors FTI – 277 and FTI – 2153 were effective at clearing HD viremia in<br />

mice.<br />

58


Session VI<br />

Immunology of viral hepatitis II<br />

59


Apoptosis in immune-mediated liver diseases<br />

Alisan Kahraman, Guido Gerken and Ali Canbay<br />

Division of Gastroenterology and Hepatology, University Hospital Essen, Germany<br />

Autoimmune diseases of the liver are chronic inflammatory processes leading to<br />

immune-mediated injury of hepato- and cholangiocytes. Cell death by apoptosis is a<br />

prominent feature in a variety of liver diseases. It is likely that apoptosis is the initial<br />

cellular response to liver and biliary injury and may thus initiate several cellular and<br />

cytokine cascades that culminate in tissue injury with subsequent fibrosis and finally,<br />

result in cirrhosis. Obviously, this cascade of events within the apoptotic machinery is<br />

of particular importance. However, Fas ligand and granzyme B significantly<br />

contribute to apoptosis observed in autoimmune liver diseases. This overview is<br />

focused on the role of apoptosis in immune-mediated liver diseases, especially in<br />

primary biliary cirrhosis (PBC), primary sclerosing cholangitis (PSC), and<br />

autoimmune hepatitis (AIH), respectively. Recently, also soluble forms of major<br />

histocompatibility complex (MHC) class I-related chains A and closely related B (MIC<br />

A and B) were reported to be increased in the sera of patients with autoimmune liver<br />

diseases. MIC A and B are cell surface glycoproteins that function as indicators for<br />

cellular stress by displaying peptides derived from proteins degraded in the cytosol<br />

on the cell surface and thus facilitate the recognition of intracellular antigens by<br />

circulating cytotoxic NK cells leading to enhanced cell death. Nowadays rationalbased<br />

strategies are being developed to suppress apoptotic cell death as a novel<br />

therapeutic option for the treatment of these liver diseases.<br />

Keywords: apoptosis, autoantibodies, cholestatic liver disease, stress-induced<br />

ligands<br />

Short title: Apoptosis in autoimmune liver disease<br />

Abbreviations: AIH, autoimmune hepatitis; AMA, antimitochondrial antibody; ALT,<br />

alanine aminotransferase; AST, aspartate aminotransferase; MIC A/B, major histocompatibility<br />

complex class I-related chains A/B; PBC, primary biliary cirrhosis; PSC,<br />

primary sclerosing cholangitis<br />

Correspondence to:<br />

Ali Canbay, M.D.<br />

Associate Professor of Medicine<br />

Division of Gastroenterology and Hepatology<br />

University Hospital Essen<br />

Hufelandstr. 55<br />

45122 Essen<br />

Germany<br />

Tel: +49 201 723-84713<br />

Fax: +49 201 723-5719<br />

E-Mail: ali.canbay@uni-due.de<br />

61


Immune response in HCV infection<br />

Carlo Ferrari<br />

Unità Operativa di Malattie Infettive ed Epatologia, Laboratorio di Immunopatologia<br />

Virale, Azienda Ospedaliero-Universitaria di Parma, Italy<br />

HCV is able to induce efficiently innate immune response genes in the liver of<br />

HCV-infected chimpanzees in the first weeks of infection and is sensitive to IFNs in<br />

vitro. Yet, HCV seems to ignore early innate defense mechanisms, as it replicates<br />

almost immediately after penetration into target cells, suggesting that the innate<br />

immunity does not significantly contribute to control virus infection. This is the likely<br />

consequence of multiple complementary strategies that HCV has developed to<br />

interfere with the antiviral effect of the different components of the innate immune<br />

system, including NK cells and the intracellular signaling cascade which leads to<br />

type I IFN expression following recognition of specific HCV associated motifs and to<br />

IFN-stimulated gene expression following cellular exposure to IFN-α and IFN-β. As a<br />

result of this interference, the physiologic priming of adaptive HCV-specific<br />

responses may be inadequately promoted and supported by the innate immunity.<br />

Evasion from innate responses and subsequent spread of the virus may set in motion<br />

a series of additional mechanisms of escape from immune surveillance. They include<br />

T cell exhaustion due to the rapid viral spread in the infected host causing an early<br />

and persistent exposure of T cells to high virus and antigen loads, direct functional T<br />

cell inhibition by HCV proteins, emergence of T and B cell escape mutations, hyperactivity<br />

of suppressive regulatory pathways (PD-1/PD-L1, CD4 + CD25 + FoxP3 positive<br />

cells, regulatory cytokines). These mechanisms can all contribute to delay and impair<br />

priming and maturation virus-specific adaptive responses at the early stages of<br />

infection when an efficient helper function of the CD4 T cell subset seems to be<br />

crucial for subsequent control of infection, allowing HCV-specific CD8 cells to<br />

differentiate into functionally efficient antiviral effector and memory cells. In contrast,<br />

inadequate CD4 help probably precludes appropriate maturation of CD8 antiviral<br />

effector and memory functions, favoring persistence of the virus. This leads in turn to<br />

a progressively more severe impairment of the overall antiviral T cell function, which<br />

represents the common feature of chronic HCV infection. In this scenario, the role of<br />

the neutralizing antibody response in HCV pathogenesis is still largely undefined,<br />

because neutralizing antibodies have been detected at high titers not only in<br />

resolving acute hepatitis, but also in patients with a long-lasting condition of chronic<br />

HCV infection. In chronic patients these antibodies can fail to neutralize co-existing<br />

autologous HCV glycoprotein, while good neutralizing activity can be observed<br />

against earlier viral strains. Sequential emergence of viral mutations able to abrogate<br />

B cell recognition of the viral envelope glycoprotein sequence may explain why high<br />

titers of neutralizing antibodies are detectable in chronic infections but are unable to<br />

control the coexisting viral strains.<br />

Thus, available data indicate that HCV is able to acquire an early survival advantage<br />

over the immune system by virtue of its rapid start of replication and its ability to<br />

inhibit innate responses. The subsequent spread of the virus can then set in motion<br />

additional mechanisms of evasion from immune control allowing HCV to outpace the<br />

protective immune response and to maintain a persistent advantage over antiviral<br />

immunity.<br />

62


Immunotherapy of hepatocellular carcinoma<br />

Tim F. Greten<br />

Department of Gastroenterology, Hepatology and Endocrinology, Medical School<br />

Hannover, Germany<br />

No systemic cytotoxic chemotherapy has proven to be safe and efficient for the<br />

treatment of patients with hepatocellular carcinoma (HCC) (1). Therefore, new<br />

treatment options are urgently needed. Recently, the multi-tyrosine kinase inhibitor<br />

sorafenib was shown to enhance overall survival in patients with advanced disease<br />

(2). However, the median survival for patients with HCC remains below 12 months. In<br />

addition, there is also accumulating evidence that immunotherapy might become a<br />

potent therapeutic option for patients with HCC (3) for several reasons: First, it has<br />

been shown that patient’s survival directly depends on the type and number of tumor<br />

infiltrating immune cells (4) (5) (6), indicating that immune responses have a direct<br />

effect on the clinical course of the disease. Second, HCC has been shown to be the<br />

first cancer, against which a potent vaccine is available. Vaccination against hepatitis<br />

B infection has been shown to effectively prevent the development of HCC (7). Third,<br />

local-ablative treatment options, which cause physical destruction of tumors, are<br />

associated with the release of high antigen load, which could induce anti-tumor<br />

immune responses. In addition, more than 80% of all patients with HCC have<br />

significant liver cirrhosis. Cohort studies indicate that HCC has become the major<br />

cause of liver-related death in patients with compensated cirrhosis (8). Therefore, this<br />

patient population might represent an ideal group for a “preventive cancer vaccine”.<br />

Interestingly we have been able to demonstrate that HCC patients mount<br />

spontaneous tumor-specific immune responses (9) indicating that immunesuppressor<br />

mechanisms might counter-balance these immune responses. Indeed,<br />

we have been able to detect different types of immune suppressor mechanisms in<br />

HCC patients such as regulatory T cells and myeloid derived suppressor cells (10)<br />

(11). Recently we have performed a clinical trial with the aim to specifically deplete<br />

immune suppressor cells such as regulatory T cells which could potentially augment<br />

the effect of either spontaneous or vaccination induced T cell responses in HCC<br />

patients. The results from these studies will be presented.<br />

References:<br />

1. Lopez, P.M., A. Villanueva, and J.M. Llovet. 2006. Systematic review: evidencebased<br />

management of hepatocellular carcinoma – an updated analysis of<br />

randomized controlled trials. Aliment <strong>Pharma</strong>col Ther 23: 1535–1547.<br />

2. Llovet, J.M., S. Ricci, V. Mazzaferro, P. Hilgard, E. Gane, J.F. Blanc,<br />

A.C. de Oliveira, A. Santoro, J.L. Raoul, A. Forner, M. Schwartz, C. Porta,<br />

S. Zeuzem, L. Bolondi, T.F. Greten, P.R. Galle, J.F. Seitz, I. Borbath,<br />

D. Häussinger, T. Giannaris, M. Shan, M. Moscovici, D. Voliotis, and J. Bruix.<br />

2008. Sorafenib in advanced hepatocellular carcinoma. N Engl J Med 359:<br />

378–390.<br />

63


3. Greten, T.F., M.P. Manns, and F. Korangy. 2006. Immunotherapy of hepatocellular<br />

carcinoma. J Hepatol 45: 868–878.<br />

4. Wada, Y., O. Nakashima, R. Kutami, O. Yamamoto, and M. Kojiro. 1998.<br />

Clinicopathological study on hepatocellular carcinoma with lymphocytic infiltration.<br />

Hepatology (Baltimore, Md) 27: 407–414.<br />

5. Unitt, E., A. Marshall, W. Gelson, S.M. Rushbrook, S. Davies, S.L. Vowler,<br />

L.S. Morris, N. Coleman, and G.J. Alexander. 2006. Tumour lymphocytic<br />

infiltrate and recurrence of hepatocellular carcinoma following liver transplanttation.<br />

J Hepatol 45: 246–253.<br />

6. Gao, Q., S.J. Qiu, J. Fan, J. Zhou, X.Y. Wang, Y.S. Xiao, Y. Xu, Y.W. Li, and<br />

Z.Y. Tang. 2007. Intratumoral balance of regulatory and cytotoxic T cells is<br />

associated with prognosis of hepatocellular carcinoma after resection. J Clin<br />

Oncol 25: 2586–2593.<br />

7. Chang, M.H., C.J. Chen, M.S. Lai, H.M. Hsu, T.C. Wu, M.S. Kong, D.C. Liang,<br />

W.Y. Shau, and D.S. Chen. 1997. Universal hepatitis B vaccination in Taiwan<br />

and the incidence of hepatocellular carcinoma in children. Taiwan Childhood<br />

Hepatoma Study Group. N Engl J Med 336: 1855–1859.<br />

8. Fattovich, G., T. Stroffolini, I. Zagni, and F. Donato. 2004. Hepatocellular<br />

carcinoma in cirrhosis: incidence and risk factors. Gastroenterology 127:<br />

S35–50.<br />

9. Korangy, F., L.A. Ormandy, J.S. Bleck, J. Klempnauer, L. Wilkens, M.P. Manns,<br />

and T.F. Greten. 2004. Spontaneous Tumor-Specific Humoral and Cellular<br />

Immune Responses to NY-ESO-1 in Hepatocellular Carcinoma. Clin Cancer<br />

Res 10: 4332–4341.<br />

10. Ormandy, L.A., T. Hillemann, H. Wedemeyer, M.P. Manns, T.F. Greten, and<br />

F. Korangy. 2005. Increased populations of regulatory T cells in peripheral<br />

blood of patients with hepatocellular carcinoma. Cancer Res 65: 2457–2464.<br />

11. Hoechst, B., L.A. Ormandy, M. Ballmaier, F. Lehner, C. Kruger, M.P. Manns,<br />

T.F. Greten, and F. Korangy. 2008. A new population of myeloid-derived<br />

suppressor cells in hepatocellular carcinoma patients induces<br />

CD4 + CD25 + Foxp3 + T cells. Gastroenterology 135: 234–243.<br />

64


List of Chairpersons, Speakers and Scientific Organizers<br />

Prof. <strong>Dr</strong>. D.H. Adams<br />

Queen Elizabeth Hospital<br />

Institute of Clinical Sciences<br />

Liver Research Laboratories<br />

Birmingham B15 2TH<br />

Great Britain<br />

Prof. <strong>Dr</strong>. A. Alberti<br />

Università di Padova<br />

Clinica Medica 2<br />

Patologia Medica<br />

Via Giustiniani 2<br />

35128 Padova<br />

Italy<br />

<strong>Dr</strong>. P. Bertolino<br />

Centenary Institute of<br />

Cancer Medicine and Cell Biology<br />

Locked Bag No. 6<br />

Newtown, Sydney, NSW 2042<br />

Australia<br />

C. Büttner, M.D.<br />

Assistant Professor of Medicine<br />

Mount Sinai School of Medicine<br />

Endocrinology, Diabetes<br />

& Bone Disease<br />

One Gustave L. Levy Place<br />

New York, NY 10029-6574<br />

USA<br />

S.H. Caldwell, M.D.<br />

Associate Professor of Medicine<br />

University of Virginia<br />

School of Medicine<br />

GI & Hepatology Division<br />

Hospital West, P.O. Box 800708<br />

Charlottesville, VA 22908<br />

USA<br />

<strong>Dr</strong>. A. Canbay<br />

Gastroenterologie/Hepatologie<br />

Universitätsklinikum Essen<br />

Hufelandstr. 55<br />

45147 Essen<br />

Germany<br />

<strong>Dr</strong>. U. Christen<br />

Allgemeine <strong>Pharma</strong>kologie<br />

Klinikum der Johann Wolfgang<br />

Goethe-Universität Frankfurt<br />

Theodor-Stern-Kai 7<br />

60596 Frankfurt<br />

Germany<br />

Prof. <strong>Dr</strong>. V.J. Desmet<br />

Catholic University of Leuven<br />

University Hospital St. Rafael<br />

Department of Pathology<br />

Minderbroederstraat 12<br />

3000 Leuven<br />

Belgium<br />

A.M. Diehl, M.D.<br />

Professor of Medicine<br />

Duke University Medical Center<br />

Gastroenterology Division<br />

Box 3256, Snyderman/GSRB-1<br />

595 LaSalle Street<br />

Durham, NC 27710<br />

USA<br />

Prof. <strong>Dr</strong>. H.P. Dienes<br />

Pathologie<br />

Klinikum der Universität zu Köln<br />

Kerpener Str. 62<br />

50924 Köln<br />

Germany<br />

Prof. <strong>Dr</strong>. C. Ferrari<br />

Università di Parma<br />

Azienda Ospedaliera di Parma<br />

Laboratorio di Immunopatologia V<br />

Via Gramsci 14<br />

43100 Parma<br />

Italy<br />

Prof. <strong>Dr</strong>. G. Gerken<br />

Gastroenterologie/Hepatologie<br />

Universitätsklinikum Essen<br />

Hufelandstr. 55<br />

45147 Essen<br />

Germany<br />

65


Prof. <strong>Dr</strong>. W.H. Gerlich<br />

Medizinische Mikrobiologie<br />

Universitätsklinikum<br />

Gießen und Marburg<br />

Frankfurter Str. 107<br />

35392 Gießen<br />

Germany<br />

M.E. Gershwin, M.D.<br />

Professor of Medicine<br />

University of California<br />

School of Medicine<br />

Rheumatology & Allergy<br />

One Shields Ave., Tupper Hall<br />

Davis, CA 95616<br />

USA<br />

Prof. <strong>Dr</strong>. T.F. Greten<br />

Klinik für Gastroenterologie,<br />

Hepatologie und Endokrinologie<br />

Medizinische Hochschule Hannover<br />

Carl-Neuberg-Str. 1<br />

30625 Hannover<br />

Germany<br />

PD <strong>Dr</strong>. J. Herkel<br />

Medizinische Klinik I<br />

Universitätsklinikum Eppendorf<br />

Martinistr. 52<br />

20251 Hamburg<br />

Germany<br />

Prof. <strong>Dr</strong>. H. Holzmann<br />

Institut für Virologie<br />

Medizinische Universität Wien<br />

Kinderspitalgasse 15<br />

1095 Wien<br />

Austria<br />

M. Houghton, Ph.D.<br />

Epiphany Biosciences, Inc.<br />

Suite 2800<br />

One California Street<br />

San Francisco, CA 94111<br />

USA<br />

66<br />

<strong>Dr</strong>. E. Jaeckel<br />

Klinik für Gastroenterologie,<br />

Hepatologie und Endokrinologie<br />

Medizinische Hochschule Hannover<br />

Carl-Neuberg-Str. 1<br />

30625 Hannover<br />

Germany<br />

PD <strong>Dr</strong>. K. Klugewitz<br />

Gastroenterologie<br />

Charité Universitätsmedizin<br />

Campus Benjamin Franklin (CBF)<br />

Hindenburgdamm 30<br />

12203 Berlin<br />

Germany<br />

Prof. <strong>Dr</strong>. P.A. Knolle<br />

Molekulare Medizin<br />

Universitätsklinikum Bonn<br />

Sigmund-Freud-Str. 25<br />

53127 Bonn<br />

Germany<br />

Prof. <strong>Dr</strong>. A. Knuth<br />

Medizinische Poliklinik<br />

Rämistr. 100<br />

8091 Zürich<br />

Switzerland<br />

<strong>Dr</strong>. K. Lang<br />

Gastroenterologie/Hepatologie<br />

Universitätsklinikum Düsseldorf<br />

Moorenstr. 5<br />

40225 Düsseldorf<br />

Germany<br />

Prof. <strong>Dr</strong>. A.W. Lohse<br />

Medizinische Klinik I<br />

Universitätsklinikum Eppendorf<br />

Martinistr. 52<br />

20251 Hamburg<br />

Germany<br />

<strong>Dr</strong>. M.K. Maini<br />

University College London<br />

Department of Immunology<br />

& Molecular Pathology<br />

46 Cleveland Street<br />

London W1T 4JF<br />

Great Britain


<strong>Dr</strong>. A. Mallat<br />

Hôpital Henri Mondor<br />

51, Av. Mal de Lattre de Tassigny<br />

94010 Creteil<br />

France<br />

Prof. <strong>Dr</strong>. M.P. Manns<br />

Klinik für Gastroenterologie,<br />

Hepatologie und Endokrinologie<br />

Medizinische Hochschule Hannover<br />

Carl-Neuberg-Str. 1<br />

30625 Hannover<br />

Germany<br />

Prof. <strong>Dr</strong>. S. Matern<br />

Höfchensweg 113<br />

52066 Aachen<br />

Germany<br />

Prof. <strong>Dr</strong>. S. Meuer<br />

Immunologie<br />

Universitätsklinikum Heidelberg<br />

Im Neuenheimer Feld 305<br />

69120 Heidelberg<br />

Germany<br />

Prof. <strong>Dr</strong>. <strong>Dr</strong>. <strong>Dr</strong>.<br />

K.-H. Meyer zum Büschenfelde<br />

Trabener Str. 8<br />

14193 Berlin<br />

Germany<br />

Prof. <strong>Dr</strong>. S. Mihm<br />

Gastroenterologie<br />

Universitätskliniken Göttingen<br />

Robert-Koch-Str. 40<br />

37075 Göttingen<br />

Germany<br />

<strong>Dr</strong>. E. Powell<br />

Royal Prince Charles Hospital<br />

137 Central Avenue<br />

St. Lucia, Brisbane, QLD 4067<br />

Australia<br />

Prof. <strong>Dr</strong>. <strong>Dr</strong>. h.c. G. Ramadori<br />

Gastroenterologie<br />

Universitätskliniken Göttingen<br />

Robert-Koch-Str. 40<br />

37075 Göttingen<br />

Germany<br />

<strong>Dr</strong>. H. Ring-Larsen<br />

University of Copenhagen<br />

Faculty of <strong>Pharma</strong>ceutical Scienc<br />

Department of <strong>Pharma</strong>cology<br />

& <strong>Pharma</strong>cotherapy<br />

Universitetsparken 2<br />

2100 Copenhagen<br />

Denmark<br />

Prof. <strong>Dr</strong>. M. Rizzetto<br />

Azienda Ospedaliera<br />

S. Giovanni Battista di Torino<br />

Divisione di Gastroenterologia<br />

Corso Bramante 88<br />

10126 Torino<br />

Italy<br />

A.J. Sanyal, M.D.<br />

Assistant Professor of Medicine<br />

Medical College of Virginia<br />

Gastroenterology & Hepatology<br />

Box 980711<br />

Richmond, VA 23298<br />

USA<br />

Prof. <strong>Dr</strong>. S.W. Schalm<br />

LiverDoc B.V.<br />

Veerkade 8d<br />

3016 DE Rotterdam<br />

The Netherlands<br />

<strong>Dr</strong>. J. Schattenberg<br />

Innere Medizin I<br />

Klinikum der Universität<br />

Langenbeckstr. 1<br />

55131 Mainz<br />

Germany<br />

Prof. <strong>Dr</strong>. J. Schölmerich<br />

Klinik für Innere Medizin I<br />

Klinikum der Universität Regensburg<br />

93042 Regensburg<br />

Germany<br />

67


Prof. <strong>Dr</strong>. E. Schrumpf<br />

University of Oslo<br />

Rikshospitalet<br />

Section of Gastroenterology<br />

& Hepatology<br />

0027 Oslo<br />

Norway<br />

Prof. <strong>Dr</strong>. C.P. Strassburg<br />

Klinik für Gastroenterologie,<br />

Hepatologie und Endokrinologie<br />

Medizinische Hochschule Hannover<br />

Carl-Neuberg-Str. 1<br />

30625 Hannover<br />

Germany<br />

<strong>Dr</strong>. M.G. Swain<br />

University of Calgary<br />

Health Sciences Centre<br />

3330 Hospital <strong>Dr</strong>ive N.W.<br />

Calgary, AB T2N 4N1<br />

Canada<br />

Prof. <strong>Dr</strong>. H.C. Thomas<br />

St. Mary's Hospital<br />

Medical School<br />

Department of Medicine<br />

Praed Street<br />

London W2 1NY<br />

Great Britain<br />

Prof. <strong>Dr</strong>. G. Tiegs<br />

Experimentelle Immunologie<br />

Universitätsklinikum Eppendorf<br />

Martinistr. 52<br />

20251 Hamburg<br />

Germany<br />

Prof. <strong>Dr</strong>. H. Tilg<br />

Bezirkskrankenhaus<br />

Hall i.T.<br />

Interne Abteilung<br />

Milserstr. 10<br />

6060 Hall/Tirol<br />

Austria<br />

68<br />

Prof. <strong>Dr</strong>. M. Trauner<br />

Medizinische Universität Graz<br />

Klinische Abteilung für<br />

Gastroenterologie & Hepatologie<br />

Auenbruggerplatz 15<br />

8036 Graz<br />

Austria<br />

Prof. <strong>Dr</strong>. C. Trautwein<br />

Medizinische Klinik III<br />

Universitätsklinikum Aachen<br />

Pauwelsstr. 30<br />

52074 Aachen<br />

Germany<br />

Prof. <strong>Dr</strong>. C. Trépo<br />

Inserm U. 271<br />

151, Cours Albert Thomas<br />

69003 Lyon<br />

France<br />

Prof. <strong>Dr</strong>. D. Vergani<br />

King's College Hospital<br />

Institute of Liver Studies<br />

Denmark Hill<br />

London SE5 9RS<br />

Great Britain<br />

PD <strong>Dr</strong>. H. Wedemeyer<br />

Klinik für Gastroenterologie,<br />

Hepatologie und Endokrinologie<br />

Medizinische Hochschule Hannover<br />

Carl-Neuberg-Str. 1<br />

30625 Hannover<br />

Germany<br />

<strong>Dr</strong>. F.T. Wunderlich<br />

Genetisches Institut<br />

Universität Köln<br />

Zülpicher Straße 47<br />

50674 Köln<br />

Germany<br />

<strong>Dr</strong>. M. Zierden<br />

Pathologie<br />

Klinikum der Universität zu Köln<br />

Kerpener Str. 62<br />

50924 Köln<br />

Germany


POSTER ABSTRACTS<br />

<strong>Poster</strong> Numbers<br />

NAFLD and NASH 1 – 10<br />

Hepatitis A and B 11 – 21<br />

Hepatitis C 22 – 38<br />

Autoimmune hepatitis 39 – 44<br />

Cholestatic liver diseases (PBC/PSC) 45 – 47<br />

Malignant and benign neoplasm 48 – 53<br />

Basic immunology 54 – 64<br />

Hepatic vascular disorders, drugs and toxicity,<br />

epidemiology and endoscopy posters 65 – 74<br />

Author Index to <strong>Poster</strong> <strong>Abstracts</strong>


Nonalcoholic fatty liver disease in autopsy of children and<br />

adolescents<br />

Fadime Yüksel 1 , Dilhan Ergün 2 , İlhami Yüksel 3 , Sema Kara 4 , Emine Şamdancı 5 ,<br />

Nurullah Çelik 6<br />

1<br />

Pediatric Specialist, Ankara Forensic Group, Ankara, Turkey<br />

2<br />

Pathology Specialist, Ankara Forensic Group, Ankara, Turkey<br />

3<br />

Department of Gastroenterology, Dışkapı Yıldırım Beyazıt Education and Research<br />

Hospital, Ankara, Turkey<br />

4<br />

Department of Neonatology, Fatih University School of Medicine, Ankara, Turkey<br />

5<br />

Department of Pathology, İnönü University School of Medicine, Malatya, Turkey<br />

6<br />

Department of Endocrinology, Fatih University School of Medicine, Anakara, Turkey<br />

Objective: Nonalcoholic fatty liver disease (NAFLD) is probably the most common<br />

cause of liver disease in the childhood and adolescent age groups. Although there<br />

are few population-based data available and there are limitations in the populationbased<br />

approach as previously noted, the available data suggest an overall<br />

prevalence of at least 3% for suspected fatty liver disease among children and<br />

adolescents. Liver biopsy cannot be used as a screening tool in population studies;<br />

however, autopsy series are another means of estimating prevalence based on a<br />

histological diagnosis. We conducted our study to determine the prevalence of fatty<br />

liver by histology in autopsy of children and adolescents.<br />

Material and methods: We included all children aged 2 through 20 years at the time<br />

of death who had an autopsy performed by the medical examiner between 2006 and<br />

2008. We included a minimum age of 2 years, because this is the youngest age<br />

reported for biopsy-proven steatohepatitis. Missing clinical data or the absence of a<br />

liver slide for review and identification of a factor that may influence liver histology<br />

(alcohol or drug use) were exclusion criteria. Fatty liver was defined as 5% of<br />

hepatocytes containing macrovesicular fat. Steatohepatitis were subclassified as<br />

Type 1, Type 2 or overlap.<br />

Results: 405 children aged 2 to 20 years had an autopsy performed by the Ankara<br />

Forensic Group medical examiner. Of these, 10 subjects were excluded because of<br />

insufficient data: no available liver tissue (n = 1) or missing clinical data (n = 2).<br />

7 subjects were excluded because of factors that may influence liver histology;<br />

alcohol-use (n = 2), positive drug toxicology (n = 5). The remaining 395 subjects were<br />

included to study.<br />

Motor vehicles were involved in 4 of all deaths in cases with NAFLD. Other accidents<br />

(drowning, electric shock and falls) accounted for an additional 4 deaths. Remaining<br />

2 cases have died because of other reasons. The prevalence of fatty liver was found<br />

5.1% in children and adolescents. Fatty liver was present in 9 boys and 11 girls. Fatty<br />

liver prevalence increased with age ranging from 3/20 for ages 2 to 4 years up to<br />

13/20 for ages 15 to 20 years. Children and adolescents with fatty liver were<br />

significantly older (p < 0.05). There was no significant difference between sexes.<br />

1


BMI in subject with fatty liver was higher than in subjects with nonfatty liver, but this<br />

finding was not statistically significant (p > 0.05). Overweight children accounted for<br />

50% of all of the cases of fatty liver. Simple steatosis was detected in 7 subjects.<br />

Steatohepatitis was characterized as Type 1 in 5 subjects, Type 2 in 7 subjects, and<br />

as overlap in 1 subject.<br />

Discussion: The prevalence of fatty liver was found 5.1% in children aged 2 to<br />

20 years in Ankara, Turkey. These findings showed that fatty liver is an important<br />

health problem in our country. Our findings showed that 50% of fatty liver were<br />

overweight children, consistent with the literature. The risk factors identified may be<br />

useful in the development of protocols designed to screen at-risk children and<br />

adolescents. Following studies about NAFLD in the children and adolescent, will<br />

contribute to our knowledge for both diagnosis and treatment.


Non-invasive evaluation of liver fibrosis in patients with NAFLD<br />

(comparison between transient elastography and BARD score)<br />

Adriana Tudora, O. Ciof, E. <strong>Dr</strong>agan, Diana Rill, Sporea Ioan, Scurtu-Martin Mihaela<br />

Department of Gastroenterology and Hepatology, University of Medicine and<br />

<strong>Pharma</strong>cy “Victor Babes”, Timisoara, Romania<br />

Introduction: Assessment of the degree of fibrosis, in patients with liver steatosis<br />

(LS) is important for the clinical outcome of the patient, and the progression to<br />

cirrhosis. Therefore, non-invasive tests were developed in order to provide useful<br />

information to discriminate NAFLD from NASH, quantifying the degree of liver<br />

fibrosis. LSMs using transient elastograpy (TE) has intrinsic limitation of subcutaneous<br />

adipose tissue making the method difficult to perform in obese subjects. Seric<br />

markers such as BARD score share the benefits of non-invasive scores, due to the<br />

fact that it was developed in obese subjects and to its simplicity. We aimed to<br />

analyze the correlation between assessment of LF in patient with NAFLD by TE<br />

(FibroScan ® , Echosense, Paris) and BARD score.<br />

Methods: We evaluated 226 patients with NAFLD: 62 women (27.4%) and 164 men<br />

(72.6%), mean age 44.3 ± 3 years. In all patients LSMs by TE was performed and<br />

Bard Score was calculated (Body mass index + AST/ALT ratio + Diabetes mellitus).<br />

Liver biopsy (LB) was carried out in 64% (144) patients.<br />

Results: Mean BMI was 29.9 kg/m 2 (13% patients had normal BMI), mean AAR was<br />

0.93 and diabetes mellitus was present in 27% patients (resulting a mean BARD<br />

score of 2.5 ± 0.7). Mean fibrosis score at LB (METAVIR) was 1.4. Mean value of<br />

LSMs was 7.1 ± 1.2 kPa (vs. 4.8 ± 1.3 kPa in normal subjects). In 77 (36.2%) cases,<br />

LS was ≤ 5.9 kPa (equivalent to F = 0); in 45 (21.1%) patients, LS varied between<br />

6 and 6.8 kPa (equivalent to F = 1); and in the 91 (42.7%) remaining patients, LSMs<br />

were ≥ 6.9 kPa (equivalent to significant fibrosis). In these patients mean BARD<br />

score was 2.9 ± 1.1 vs. 1.2 ± 0.2 (p < 0.001-ES) in F0F1patients. Patients with F ≥ 2<br />

at LB (n = 86, 38%) had a mean BARD score of 3.1 ± 0.8 compared to those with<br />

F < 2 (p < 0.001-ES).<br />

Discussion/Conclusion: In our study, both TE and BARD score were statistically<br />

significant correlated with advanced fibrosis (F ≥ 2).<br />

2


3<br />

Assessment of liver fibrosis by using transient elastography<br />

(TE) in patients with nonalcoholic fatty liver disease (NAFLD)<br />

Octavian Ciof, Adriana Tudora, E. <strong>Dr</strong>agan, Diana Rill, I. Sporea, M. Scurtu-Martin<br />

Department of Gastroenterology and Hepatology, University of Medicine and<br />

<strong>Pharma</strong>cy “Victor Babes”, Timisoara, Romania<br />

Introduction: Determination of the extent of liver fibrosis in NAFLD subjects is of<br />

great clinical importance due to progression to cirrhosis. Until recently, liver biopsy<br />

(LB) was the only method for evaluating liver fibrosis in these patients. Transient<br />

elastography (FibroScan ® , Echosense, Paris) is a new technique that allows rapid,<br />

non-invasive measurement of mean tissue stiffness, which was evaluated in<br />

estimating liver fibrosis in patients with chronic hepatitis C. We carried out a study to<br />

determine the value of liver stiffness measurements (LSMs) in patients with NAFLD.<br />

Methods: Out of 3459 TE evaluations for hepatopathies of different etiologies<br />

performed in the Department of Gastroenterology and Hepatology Timisoara<br />

between 2008–2009 NAFLD was present in 226 patients: 62 women (27.4%) and<br />

164 men (72.6%), mean age 44.3 ± 3 years. Liver biopsy was performed in<br />

140 patients (62%).<br />

Results: Mean value of LS values was 7.1 ± 1.2 kPa (compared to 4.8 ± 1.3 kPa in<br />

normal subjects) (p < 0.0001-ES). Valid measurements could not be obtained in 7%<br />

(13) cases. In the 213 remaining patients, the following values were obtained: in<br />

77 (36.2%) cases, LS was ≤ 5.9 kPa (equivalent to F = 0); in 45 (21.1%) patients, LS<br />

varied between 6 and 6.8 kPa (equivalent to F = 1); and in the 91 (42.7%) remaining<br />

patients LS was ≥ 6.9 kPa (equivalent to significant fibrosis). When referred to<br />

morphological assessment, mean LF (METAVIR score) was 1.4 ± 1.2, and<br />

furthermore, we noticed a stepwise increase in liver stiffness with increasing<br />

histological severity of hepatic fibrosis (p < 0.0001). The AUROC were 0.78, 0.81,<br />

0.89, and 0.94, for F1, F2, F3 and F4, respectively.<br />

Discussion/Conclusion: TE using FibroScan ® allows a rapid and non-invasive<br />

estimation of the stage of fibrosis in NAFLD patients, especially NASH patients. The<br />

results of our study showed a significant positive correlation between liver stiffness<br />

and the severity of liver fibrosis in patients with NAFLD.


HFE gene mutations in patients with biopsy-proven NAFLD<br />

E. Stachowska 1 , J. Raszeja-Wyszomirska 2 , J. Suchy 3 , M. Ławniczak 4 , I. Zawada 4<br />

1<br />

Department of Biochemistry and Human Nutrition, Pomeranian Medical University,<br />

Szczecin, Poland<br />

2<br />

Liver Unit, Pomeranian Medical University, Szczecin, Poland<br />

3<br />

International Hereditary Cancer Center, Department of Genetics and Pathology,<br />

Pomeranian Medical University, Szczecin, Poland<br />

4<br />

Department of Gastroenterology, Pomeranian Medical University, Szczecin, Poland<br />

Background: The role of iron deposition in the pathogenesis of NAFLD has raised<br />

general interest. Primary hepatic iron overload is associated with clinical features of<br />

insulin resistance. Iron can directly cause lipid peroxidation, leading to activation of<br />

stellate cells and increase collagen production. The results of prevalence of HFE<br />

gene mutations in NAFLD patients are still scanty and controversial.<br />

Aim of the study: To evaluate the frequency of HFE gene mutations in patients with<br />

biopsy-proven NAFLD and to compare it to the frequency in general population and<br />

in patients with ALD.<br />

Materials and methods: 50 consecutive Caucasian patients with biopsy-proven<br />

NAFLD were included into the study. The protocol was approved by local Bioethics<br />

Board.<br />

To evaluate C282Y and H63D HFE gene mutations PCR-RFLP was performed,<br />

according to methodology previously described. The control groups comprised of<br />

119 patients with ALD and 1517 DNA samples obtained from either cord bloods or<br />

healthy subjects from family medicine practices.<br />

Results: The frequency of HFE gene mutations in all analyzed groups was<br />

comparable and statistically not significant. These data are summarized in table 1.<br />

Mutation of Healthy controls ALD<br />

NAFLD<br />

HFE gene<br />

(n = 1517)<br />

(n = 119)<br />

(n = 50)<br />

C282Y Homo 2 (0.13%) 1 (0.63%) 1 (2%)<br />

C282Y Hetero 117 (7.8%) 8 (5%) 1 (2%)<br />

H63D Homo 38 (2.5%) 8 (5%) 2 (4%)<br />

H63D Hetero 380 (25%) 40 (25%) 11 (22%)<br />

Discussion/Conclusion: The results of our study support the hypothesis that HFE<br />

gene mutations do not play a critical role in the development of NAFLD. However<br />

these findings have to be verified in bigger cohorts of patients.<br />

This paper was supported by grant from State Committee for Scientific Research, in<br />

years 2006–2009, No N 402 099 31/3037.<br />

4


5<br />

Validation of BARD scoring system among Polish patients with<br />

NAFLD<br />

J. Raszeja-Wyszomirska 1 , M. Ławniczak 2 , B. Szymanik 3 , P. Milkiewicz 1 , M. Hartleb 4<br />

1 Liver Unit, Pomeranian Medical University, Szczecin, Poland<br />

2 Department of Gastroenterology, Pomeranian Medical University, Szczecin, Poland<br />

3 West-Pomeranian University of Technology, Electric Faculty, Department of<br />

Theoretic Electrotechnics and Computer Science, Szczecin, Poland<br />

4 Department of Gastroenterology and Hepatology, Silesian Medical University,<br />

Katowice, Poland<br />

Introduction: Nonalcoholic fatty liver disease (NAFLD) includes a wide spectrum of<br />

liver diseases, ranging from pure steatosis to nonalcoholic steatohepatitis (NASH),<br />

and eventually to liver cirrhosis with its complications. Identifying patients with more<br />

advanced fibrosis at diagnosis is crucial for their prognosis and possible therapeutic<br />

intervention. Harrison et al. (Gut 2008; 57: 1441–7) proposed a novel easy to<br />

conduct in everyday clinical practice scoring system perfectly identifying NAFLD<br />

patients without significant fibrosis (BARD score).<br />

Aim of the study: To validate clinical utility of BARD scoring system among Polish<br />

patients with NAFLD.<br />

Methods: Group of 101 Caucasians with biopsy-proven NAFLD were included into<br />

the study. Liver biopsies were assessed according to Scheurer score (Clin Liver Dis<br />

2002; 6: 335–47) or Histological Scoring System for Nonalcoholic Fatty Liver<br />

Disease. BARD scoring system was assessed according to Harrison et al.:<br />

BMI ≥ 28 = 1 point, AST/ALT ratio (AAR) ≥ 0.8 = 2 points, type 2 diabetes<br />

mellitus = 1 point. Statistic analyses were performed with Chi2 and ANOVA, using<br />

SPSS 15.0.<br />

Results: None of BARD component showed strong association with advanced<br />

fibrosis, however, summarized BARD score of 2–4 points was associated with F3 or<br />

F4 stages of fibrosis with an odds ratio of 15.5 (95% Cl: 3.18–75.5) and negative<br />

predictive value of 97%.<br />

Discussion/Conclusion: Our results show that BARD scoring system holds a<br />

universal value in non-invasive diagnosis of advanced fibrosis in NAFLD patients.<br />

This paper was supported by grant from State Committee for Scientific Research, in<br />

years 2006–2009, No N 402 099 31/3037.


High rate of undetected arterial hypertension in patients with<br />

NASH<br />

M. Demir, S. Schulte, S. Ubben, S.M. Lang, T. Goeser, U. Töx, H.-M. Steffen<br />

Department of Gastroenterology and Hepatology, University of Cologne, Cologne,<br />

Germany<br />

Introduction: Non-alcoholic fatty liver disease (NAFLD) is an emerging metabolicrelated<br />

disorder characterized by fatty infiltration of the liver in the absence of alcohol<br />

consumption. NAFLD ranges from simple steatosis to non-alcoholic steatohepatitis<br />

(NASH), which might progress to end-stage liver disease. This progression is related<br />

to insulin resistance, which is strongly linked to the metabolic syndrome consisting of<br />

central obesity, dyslipidemia, and hypertension. Each of these abnormalities carries a<br />

risk for cardiovascular disease. The aim of this study was to determine the<br />

prevalence of arterial hypertension and to identify the rate of undetected<br />

hypertension in patients with NAFLD and NASH.<br />

Methods: Patient records from 2006 to 2008 of the in- and outpatient department of<br />

the Clinic of Gastroenterology and Hepatology at the Abdominal Center of the<br />

University of Cologne with biopsy proven NAFLD (n = 16) or NASH (n = 30) were<br />

retrospectively reviewed for the prevalence of arterial hypertension (blood pressure<br />

≥ RR140/90 mmHg). Patients with a history of hypertension or concomitant therapy<br />

with blood pressure lowering agents were categorized as known hypertensives.<br />

Patients with chronic hepatitis C virus infection served as a control group (n = 122).<br />

Results: The number of known hypertensives was 13/122 (10.7%) in the HCV group,<br />

3/16 (18.7%) in the NAFLD group and 5/30 (16.7%) in the NASH group. The number<br />

of previously undetected hypertensives was 25/122 (20.5%) in the HCV group,<br />

2/16 (12.5%) in the NAFLD group and 9/30 (30%) in the NASH group. The<br />

prevalence of arterial hypertension was 31.1% in the HCV group, 31.2% in the<br />

NAFLD group and 46.7% in the NASH group.<br />

Discussion/Conclusion: We observed a high rate of undetected and untreated<br />

arterial hypertension in patients with NASH. Since the incidence of NAFLD/NASH is<br />

expected to rise in Western countries gastroenterologists and hepatologists should<br />

be encouraged to take the opportunity to measure blood pressure routinely and<br />

recommend adjustment of antihypertensive therapy during specialized care for this<br />

high risk population.<br />

6


7<br />

Fatty liver disease in children and adolescents<br />

V. Tzaneva, S. Galcheva, D. Baleva<br />

Department of Pediatric Endocrinology, Department of Radiology 1 , University<br />

Hospital “St. Marine”, Varna, Bulgaria<br />

Introduction: Fatty liver disease (FLD) is a liver condition with increasing frequency.<br />

FLD is associated with diabetes, insulin resistance, hypercholesterolemia, hypertriglyceridemia<br />

and obesity. Until recently it is an important childhood liver disease,<br />

especially because childhood obesity is much more common. Aim was to determine<br />

the prevalence of FLD among children with diabetes mellitus (DM) type 1 and with<br />

obesity, its clinical features and treatment.<br />

Methods: In 1992–2008, 15 children were diagnosed with FLD among these with<br />

DM (258) and 44 children among these with obesity (135). The diagnosis was<br />

determined by elevated serum aminotransferases and abnormal hepatic sonogram.<br />

Test for Wilson’s disease and chronic hepatitis B and C were made as well as a lipid<br />

profile (patients with FLD). In 35 children a therapy was made: vitamin E 200–400<br />

mg/23 h, for 3–6 months or ursodeoxycholic acid (10 mg/kg/24 h) for 3 months.<br />

Results: Among 258 peditatric patients with DM type 1, 15 children were diagnosed<br />

with FLD (10 boys and 5 girls, from 7 to 18 years, median range 13.81 years), it was<br />

5.81%. Among 135 obese children, 44 were with FLD (23 boys and 21 girls, from 4 to<br />

18 years, median range 7.45 years). It was 32.59%. Among obesity grade IV<br />

(19 patients) 10 were with FLD (52.63%). All patients had hepatomegaly and only 2<br />

had splenomegaly. All children had abnormal sonogram suggestive of fatty<br />

infiltration. Abnormal aminotransferases (steatohepatitis) were found in 2.7% among<br />

DM type 1. Abnormal aminotransferases (steatohepatitis) were found in 6.67%<br />

among obese children DM type 1 steatohepatitis. Evaluated children were negative<br />

for Wilson’s disease and viral hepatitis. Lipid profile was abnormal in 28 cases:<br />

15 with hypercholesterolemia and 13 with hypertriglyceridemia. The effect of<br />

treatment was controlled clinically, biochemically, ultrasonographically after<br />

3 months. The effect was better when the control of DM was better and when there<br />

was a reduction in the weight of the obese children.<br />

Discussion/Conclusion: FLD occurs in children with DM type 1 and children with<br />

obesity. It is not a rare condition. The essential basis of the treatment is the adequate<br />

control of DM and body weight.


Effects of dietary chromium picolinate in the treatment of the<br />

experimental nonalcoholic steatohepatitis<br />

N. Kuzu 1 , İ.H. Bahçecioğlu 2 , A. Gençaslan 3 , K. Metin 4 , B. Üstündağ 4 , İ.H. Özercan 5 ,<br />

K. Şahin 6<br />

1<br />

Ege University, Faculty of Medicine, Division of Gastroenterology, Izmir, Turkey<br />

2<br />

Firat University, Faculty of Medicine, Division of Gastroenterology, Elazig, Turkey<br />

3<br />

Firat University, Faculty of Medicine, Department of Internal Medicine, Elazig,<br />

Turkey<br />

4<br />

Firat University, Faculty of Medicine, Department of Biochemistry, Elazig, Turkey<br />

5<br />

Firat University, Faculty of Medicine, Department of Pathology, Elazig, Turkey<br />

6<br />

Firat University, Faculty of Veterinary, Department of Nutrition, Elazig, Turkey<br />

Introduction: The development of nonalcoholic steatohepatitis (NASH) is strongly<br />

associated with the metabolic syndrome and insulin resistance. Cr-picolinate<br />

supplementation facilitates normal protein, fat, and carbohydrate metabolism. Cr-picolinate<br />

is often used as dietary supplement to improve insulin sensitivity and to<br />

correct dyslipidemia. The aim of this study was to investigate to the therapeutic effect<br />

of dietary Cr-picolinate in the experimental nonalcoholic steatohepatitis induced by<br />

high fat diet (HFD).<br />

Methods: Twenty-seven male Sprague-Dawley rats were divided into three equal<br />

groups randomly. First group received only standard rat diet (control group), while<br />

groups 2, and 3 were given HFD ad libitum. HFD was purchased from Dyets<br />

(Benthlehem, P.A.). After the 12 weeks, Cr-picolinate (200 microg Cr/kg BW) was<br />

administered by orogastric gavages daily for 4 weeks to rats in group 3. After<br />

16 weeks all rats killed.<br />

Serum biochemistry, TNF-alpha, plasma and liver tissue malondialdehyde (MDA),<br />

glutathione levels were analyzed. Insulin resistance was assessed by Homa-R<br />

method. Histopathological signs of steatosis, inflammation, and fibrosis were scored<br />

semi-quantitatively and were modified from Brunt’s criteria. Immunohistochemical<br />

staining was done for CYP2E1 and a sMA expression.<br />

Results: Mean body weight, oxidative stress, CYP 2E1 expression and serum<br />

TNF-alpha levels were increased in rats with steatohepatitis. Cr-picolinate was<br />

improved insulin sensivity.<br />

Cr-picolinate was effective in reducing hepatic steatosis and inflammation. Balloning<br />

degeneration and a sMA expression were decreased by supplemantation of Cr-picolinate.<br />

CYP 2E1 expression, liver malondialdehyde level and serum TNF-alpha level were<br />

decreased by supplementation of Cri-picolinate (p < 0.05). Plasma cholesterol and<br />

triglyceride concentrations were lower in the rats with supplementaton of Cr-picolinate.<br />

Discussion/Conclusion: These results indicate that Cr-picolinate attenuates<br />

steatohepatitis inducued by HFD in in the rats. This effect was mediated in part by<br />

reducing body wieght and improving insulin sensivity.<br />

Key words: steatohepatitis, high-fat diet, chromium picolinate<br />

8


9<br />

In what manner does the metabolic syndrome modify the viral<br />

and alcoholic hepatic cirrhosis course?<br />

N.N. Silivontchik, E.I. Adamenko<br />

Educational Establishment “Belarusian Medical Academy of Post-Graduate<br />

Education”, Minsk, Belarus<br />

Introduction: Metabolic syndrome (MS) is an independent cause of liver disease<br />

and the problem has an increasing interest.<br />

Aim: Clinical peculiarities of viral (HCV) and alcoholic hepatic cirrhosis (HC)<br />

associated with MS were analyzed.<br />

Methods: Clinical study was carried out in 623 hospitalized patients with viral and<br />

alcoholic HC including 41 with MS (6.1% had Child A, 51.1% – B, 42.3% – C). HC<br />

was proved by biopsy or simply derived from clinical and laboratory data. MS was<br />

diagnosed according to ATP (2001). All patients suffer from diabetes (28 received<br />

oral hypoglycemic agents, 13 insulin).<br />

Results: MS patients were older than others (54.41 ± 10.36 versus 50.28 ± 11.94<br />

years, p = 0.029), had longer history of HC (Me = 16 versus 12 months, p = 0.045),<br />

and didn’t differ in sex. Contrary to expectations HC associated with MS didn’t<br />

exceed HC without MS in the severity Child’s score (Me = 9 versus 10 points,<br />

p = 0.06) and the majority of characteristics (only cholelithiasis was reliably more<br />

frequent in MS patients: 26.3 versus 10.3%, p < 0.001). Mortality rate (hepatic and<br />

nonhepatic) was equal (χ 2 = 0.338, p = 0.501).<br />

Discussion/Conclusion: MS isn’t a risk factor of alcoholic and HCV HC aggravation<br />

and clinical course worsening. The explanation may be: 1) satisfactory MS patients’<br />

nutrition status, 2) more strict treatment control because of diabetes, 3) decreasing<br />

cardiovascular complication risk.


10<br />

Assessment of liver fibrosis by TE (FibroScan ® ) in patients with<br />

metabolic syndrome<br />

Scurtu-Martin Mihaela, Adriana Tudora, O. Ciof, E. <strong>Dr</strong>agan, Diana Rill, Sporea Ioan<br />

Department of Gastroenterology and Hepatology, University of Medicine and<br />

<strong>Pharma</strong>cy “Victor Babes”, Timisoara, Romania<br />

Introduction: The aim of this study was to investigate the accuracy of liver stiffness<br />

measurements (LSMs) by transient elastography (TE) (FibroScan ® , Echosense,<br />

Paris, France) as a noninvasive test, in the diagnosis of hepatic fibrosis, and to<br />

determine whether it can replace liver biopsy in patients with metabolic syndrome<br />

without other cause of liver disease.<br />

Methods: 3459 TE evaluations were performed between 2007–2009 in the<br />

Department of Gastroenterology and Hepatology, Timisoara. MS (defined using<br />

NCEP definition as the presence of at least 3 of the following criteria: waist<br />

circumference > 102 cm in men (M) and 88 in women (W), triglycerides<br />

> 1.69 mmol/l, HDL-cholesterol < 1.04 mmol/l (M) and 1.29 (W), blood pressure<br />

> 130/85 mmHg, fasting glucose > 6.1 mmol/l) was present in 10.5% (n = 363 cases,<br />

99 women, 27,4% and 264 men, 72,6%, mean age 44.3 ± 3 years). All subjects had<br />

a complete medical examination and laboratory tests additionally to LSM performed<br />

the same day by a single operator unaware of clinical and biological data.<br />

Results: Valid measurements could not be obtained in 7% (25) cases. In the<br />

338 remaining patients, mean LSMs in patients with MS values were 7.1 ± 1.2 kPa<br />

significantly higher compared to normal subjects (4.8 ± 1.3 kPa), (p < 0.0001). The<br />

following values were obtained: in 77 (36.2%) cases LS was ≤ 5.9 kPa (equivalent to<br />

F = 0); in 45 (21.1%) patients, LS varied between 6 and 6.8 kPa (equivalent to F = 1);<br />

and in the 91 (42.7%) remaining patients, LS was ≥ 6.9 kPa (equivalent to significant<br />

fibrosis). LSM was also higher in obese (BMI > 30 kg/m 2 ) compared to overweight or<br />

normal weight subjects (p = 0.003).<br />

Discussion/Conclusion: MS was positively associated with liver stiffness and liver<br />

fibrosis, thus LSMs using TE proved to be a reliable non-invasive method in<br />

evaluating LF in these patients.


11<br />

Investigation on cytokines at viral hepatitis A<br />

A. Petrov, N. Vatev<br />

Medical University Plovdiv, Department of Infectious Diseases, Parasitology and<br />

Tropical Medicine, University Hospital “S. George”, Clinic of Infectious Diseases,<br />

Plovdiv, Bulgaria<br />

Introduction: The harm mechanisms of the hepatocytes in viral hepatitis A are not<br />

fully cleared. The cytolysis is given mainly to immunopathogenic answer of the<br />

contaminated cells, than to the direct effect of the virus. Non-specific immune<br />

response is fulfilled largely by NK cells and cytokines. They cooperate on the<br />

communication between the cells and reveal various and important role in the liver in<br />

this manner. The cytokines act against virus infections in two main ways: direct – by<br />

inhibition of viral replication and indirectly – by determination of the dominate model<br />

of human's immune answer.<br />

Methods: We investigated Th1 cytokines: IL-1beta, IL-2, TNF-alpha, IL-6 and<br />

IFN-gamma of 83 patients with viral hepatitis A. In 29 patients surveys were done as<br />

of heat of the disease, so as well at early reconvalescence. The cytokines were<br />

determined by the enzyme immunoassay.<br />

Results: Our surveys showed increased amounts of the investigated cytokines in<br />

heat to the viral hepatitis A: IL-1beta – 3.35 pg/ml, IL-2 – 8.54 pg/ml, TNF-alpha –<br />

13.92, IL-6 – 1.92 pg/ml. The average values were significant demoted in the<br />

beginning to the reconvalescence: 1.68 pg/ml, 8.54 pg/ml, 6.63 and 1.42 pg/ml<br />

respectively. Only with three patients the serum level of IFN-gamma were increasing<br />

considerably. Nonparametric correlation test of Spearman showed correlation of<br />

values of the cytokines by clinical heaviness of viral hepatitis A.<br />

Discussion/Conclusion: Proinflammatory cytokines – IL-1beta, IL-2, TNF-alpha,<br />

IL-6 and IFN-gamma at viral hepatitis A are investigated in Bulgaria for the first time.<br />

Positive correlation is discovered between Тh1 cytokines values and the heaviness of<br />

the hepatitis which is contribution to the pathogenesis of the disease.


12<br />

Blocking Bim-mediated attrition of T cells in patients with<br />

chronic HBV infection<br />

Anna Schurich 1 , Pooja Khanna 1,2 , A. Ross Lopes 1 , Gaia Nebbia 1 ,<br />

Geoffrey Dusheiko 2 , Mala K. Maini 1,2<br />

1 Division of Infection and Immunity, UCL, London, UK<br />

2 Centre for Hepatology, Royal Free Campus, University College London Medical<br />

School, London, UK<br />

Introduction: We recently identified a role for Bim-mediated apoptosis in the<br />

profound attrition of HBV-specific CD8 T-cells characteristic of patients with chronic<br />

HBV infection (CHB). Interruption of this pathway rescued functional, multispecific<br />

CD8 responses directly ex-vivo (Lopes et al., JCI 2008). We postulate that<br />

intrahepatic antigen-presentation, shown to induce T cell expression of Bim (Holz et<br />

al., Gastro 2008), imposes this mechanism of T cell tolerance through an<br />

unfavourable balance of positive and negative co-regulation. Here we explore<br />

therapeutic options for reversing this pro-apoptotic phenotype to reconstitute effective<br />

antiviral responses in-vivo.<br />

Methods: HBV-specific CD8 were identified by ICS for IFN-γ following stimulation<br />

with peptides covering HLA-A2 restricted epitopes or spanning the HBV genome.<br />

Intracellular Bim levels were quantified by flow-cytometry before and after blocking<br />

upstream and downstream targets.<br />

Results: Cross-sectional comparison of CHB patients with/without antiviral therapy<br />

showed no differences in intracellular Bim levels in HBV-specific CD8. A longitudinal<br />

sub-study monitoring patients before and during antiviral therapy showed no<br />

reduction in Bim expression during the short-lived recovery of HBV-specific CD8<br />

upon suppression of viral load. In-vitro cyclosporin A treatment of CD8 blocked Bim<br />

expression and we are now investigating the impact of manipulating co-inhibitory<br />

(e.g. CTLA-4) or co-stimulatory (e.g. 41BB) signals on the pro-apoptotic tendency of<br />

HBV-specific CD8 in CHB.<br />

Discussion/Conclusion: HBV-specific CD8 T-cell responses reconstituted with<br />

potent antiviral therapy in CHB are transient and remain highly susceptible to<br />

apoptosis. We propose that specific measures to block Bim-mediated attrition are<br />

necessary to recover durable off-treatment antiviral responses.


13<br />

Chronic hepatitis HBV in children treated with interferon-α –<br />

Prognostic value of serum interleukin 6 and interleukin 12<br />

Gora-Gebka Magdalena, Liberek Anna, Sikorska-Wisniewska Grazyna,<br />

Gołebiewski Jacek, Kaminska Barbara<br />

Department of Paediatrics, Paediatric Gastroenterology, Hepatology and Nutrition,<br />

Medical University of Gdańsk, Poland<br />

Introduction: The course of HBV infection and the outcome of IFN-α of patients with<br />

chronic hepatitis B are determined by the antiviral immune response of the host.<br />

Serum cytokine profile could enable the selection a group of patients prone to<br />

positive response to IFN-α therapy. It appears especially important in pediatric<br />

patients with immature immunological system in whom exogenous IFN-α may cause<br />

distinct, serious side effects.<br />

The aim of the study was 1) to investigate the correlation between IL-6 and IL-12<br />

serum levels and biochemical and histopathological changes in children with chronic<br />

hepatitis B, 2) to evaluate the predictive value of the pre-treatment serum levels of<br />

these cytokines for response to treatment with IFN-α.<br />

Methods: Serum levels of IL-6, IL-12 (heterodimer p70) and IL-12 (heterodimer p70<br />

and p40 subunit) were determined by specific ELISA in 57 children (aged 5.9–17.1<br />

years) with chronic hepatitis B on the first day of IFN-α therapy.<br />

Results: Serum IL-6 levels before IFN-α treatment were 0.5–1.7 pg/ml (median<br />

0.9 pg/ml). Serum levels of IL-12 (p70) were 13.9–81.1 pg/ml (median 22.9 pg/ml)<br />

and IL-12 (p70 and p40) were 89.9–451.7 pg/ml (median 292.8 pg/ml) and they<br />

showed no correlation with biochemical and histopathological changes. There was no<br />

significant difference between the pre-treatment serum levels of these cytokines in<br />

the group of patients who seroconverted to anti-HBe (36%) and those who did not<br />

respond to IFN-α therapy (68%).<br />

Conclusion: Serum levels of IL-6 and IL-12 do not correlate with biochemical and<br />

histopathological markers of hepatitis and seem to be of poor prognostic value for<br />

positive response to the IFN-α therapy in children with chronic hepatitis B.


14<br />

High frequency of HBV-specific T cell responses in delta<br />

hepatitis<br />

J. Grabowski 1 , P.V. Suneetha 1 , J. Jaroszewicz 1,2 , V. Schlaphoff 1 , B. Bremer 1 ,<br />

M.P. Manns 1 , M. Cornberg 1 , H. Wedemeyer 1<br />

1<br />

Department of Gastroenterology, Hepatology and Endocrinology, Hannover Medical<br />

School, Hannover, Germany<br />

2<br />

Department of Infectious Diseases and Hepatology, Medical University in Bialystok,<br />

Bialystok, Poland<br />

Introduction: Infection with hepatitis D virus (HDV) can cause severe acute liver<br />

disease and rapid progression of chronic hepatitis. However, not all patients develop<br />

liver cirrhosis and immune correlates associated with different outcomes are poorly<br />

defined. The aim of this study was to investigate both HDV-specific and HBV-specific<br />

T cell responses as well as in-vitro and in-vivo cytokine levels in correlation with<br />

virological and clinical parameters.<br />

Methods: In this study, 23 individuals with HBV/HDV infection were included. Ten<br />

HBV monoinfected patients and ten healthy individuals served as controls.<br />

PBMC were stimulated by overlapping peptides spanning the entire HDV genome,<br />

HBV surface, HBcore and HBV polymerase proteins, respectively.<br />

T-cell-proliferation was investigated by 3 H-Thymidin-incorporation. Production of<br />

interleukin-2, interleukin-10, interferon-gamma and interferon-gamma-inducedprotein-10<br />

was evaluated by CBA in cell culture supernatants, serum samples of<br />

HDV patients additionally were tested for interleukin-17A and interleukin-21.<br />

Results: HDV-specific proliferative responses were detected less often than HBVspecific<br />

T-cell-responses in patients with delta hepatitis. Only 2 patients displayed<br />

both HBV- and HDV-specific responses. HDV-specific T-cell-responses were<br />

targeted mainly against amino acids 95–181 while HBcore-specific responses<br />

dominated. HDV patients with HBV-specific T-cell-responses were significantly<br />

younger than individuals with HDV-specific immune responses or nonresponder<br />

patients and had higher HBsAg levels. Furthermore, HBV-DNA-levels were<br />

significantly lower in patients with HDV-specific responses than in patients with<br />

HBV-responses.<br />

Discussion/Conclusion: HDV patients show a rather high frequency of HBVspecific<br />

T-cell-responses despite chronic HBV/HDV infection which may correlate<br />

with distinct virological characteristics. Ongoing work is investigating possible<br />

mechanisms how HDV-coinfection interferes with HBV-specific immunity.


15<br />

Hepatitis B prevalence and flare in Asian patients with<br />

rheumatoid arthritis treated with disease modifying antirheumatic<br />

drug (DMARD) therapy<br />

Owen Bebb, Victoria Appleby, Paul Southern, Philip Helliwell, Sulleman Moreea<br />

Bradford Royal Infirmary, Bradford, UK<br />

Introduction: The prevalence of hepatitis B (HBV) in our large South Asian<br />

population is around 3–4%. HBV can flare during immunosuppressant therapy,<br />

including the use of DMARD therapy. We aimed to analyse the characteristics of and<br />

determine the incidence and outcome of HBV flare in our Asian patients with<br />

rheumatoid arthritis (RA) on DMARD therapy.<br />

Methods: Our RA database and results system were analysed to obtain the<br />

following: demographics, ethnicity, treatment with DMARD therapy, HBV serology/virology<br />

and biochemical results<br />

Results: There were 2861 patients on the RA database: 934 males (32.7%) mean<br />

age 55.6 y; 1927 females (67.3%) mean age 56.0 y. 505 (17.7%) were of Asian<br />

origin – 114 M (20.5% of Asians) mean age 46.3 y; 391 F (79.5% of Asians), mean<br />

age 48.6 y. 103 of all Asian patients (24.3%) had a HBsAg test. 2/103 patients (1 M<br />

[51], 1 F [52]) were HBV positive, prevalence rate 1.94%.<br />

There were 424 Asians on DMARD therapy (84.0% of all Asians and 14.8% of the<br />

population) consisting of 89 M mean age 44.9 y and 335 F mean age 47.6 y (ratio<br />

M:F = 1:4). 206/424 (48.6%) had an ALT flare: 51 M mean age 48.7 y, 155 F mean<br />

age 48.1 y (ratio 1:3). 103 (24.3%) 26 M (mean age 42.1 y)and 77 F mean age<br />

43.8 y) were tested for HBV and 2 patients (1 M [51], 1 F [52]) had a HBV flare. They<br />

were both eAg negative patients with viral breakthrough. They were treated successfully<br />

with lamivudine with no mortality. In the 305 patients (62 M mean age 53.3 y,<br />

243 F mean age 52.7 y) on DMARD therapy without HBV status, 105 (34.4%) had a<br />

rise in ALT during treatment.<br />

Discussion/Conclusion: Our figures suggest a prevalence rate of 1.94% HBV in the<br />

local Asian population. This study has highlighted the ad-hoc HBV testing in our<br />

population. We suggest serological testing for all RA patients on DMARD therapy<br />

before the start of treatment but particularly in high risk patients originating from<br />

South Asia.


16<br />

Interferon-α treatment fails to induce activation of interferon<br />

responsive genes in HBV chronically infected human chimeric<br />

uPA/SCID mice<br />

Marc Lütgehetmann 1 , Tassilo Volz 1 , Ansgar Lohse 1 , Jan-Hendrick Bockmann 1 ,<br />

Jörg Petersen 2 and Maura Dandri 1<br />

1<br />

Department of Internal Medicine, University Medical Hospital Hamburg-Eppendorf,<br />

Germany<br />

2<br />

IFI, Institute for Interdisciplinary Medicine at Asclepius Clinic St. Georg, Hamburg,<br />

Germany<br />

Introduction: Interactions between HBV and the interferon system of the host are<br />

still poorly defined. Experimental infection of chimps showed that the interferon<br />

system is not induced during acute HBV infection, suggesting that HBV may have<br />

evolved mechanisms to avoid activation of the IFN signaling cascade.<br />

Aim of the study was to investigate whether HBV can directly interfere with the<br />

hepatocellular IFN-α response in chronically infected human hepatocytes.<br />

Methods: We used primary human hepatocytes isolated from the same donor to<br />

repopulate the livers of uPA/SCID mice and analysed changes in the expression<br />

levels of IFN responsive genes (IRGs) using primers specific for human transcripts<br />

and not cross-reacting with murine genes. Chimeric mice were then infected with<br />

HBV-positive mouse serum (genotype D) and after establishment of infection<br />

(median viremia 3 x 10 7 copies/ml) mice received 1350 IU/g IFN-α daily (n = 5) or<br />

PBS (n = 4) for 5 days. To determine responsiveness of transplanted human<br />

hepatocytes to IFN treatment, uninfected human chimeric mice were treated for<br />

5 days either with IFN-α (n = 4) or with PBS (n = 7). All animals were sacrified 8 h<br />

after the last injection.<br />

Results: HBV infection of human hepatocytes in chimeric animals did not increase<br />

expression levels of human IRGs like OAS and MXA, though significant activation of<br />

these genes could be demonstrated in the livers of IFN-treated uninfected chimeric<br />

mice (median 3-fold OAS and 10-fold MXA). Treatment with IFN-α induced only a<br />

moderate reduction of HBV-DNA titres (median 0.75 log) as well as intrahepatic<br />

rcDNA and pgRNA amounts compared to untreated control animals. Strikingly, IFN<br />

treatment was not able to increase expression levels of OAS and MXA in HBV-infected<br />

hepatocytes repopulating the uPA mouse livers, while expression levels of<br />

IFNA2 receptor remained unchanged regardless of infection status in treated mice.<br />

Discussion/Conclusion: These results suggest that HBV can specifically hinder the<br />

activation of the IFN signaling in vivo, in chronically infected primary human<br />

hepatocytes.


17<br />

Relationship between liver fibrosis and topographic<br />

distribution of alpha-smooth muscle actin staining in chronic<br />

hepatitis B<br />

O. Kosseva, E. Pophristova, D. Adjarov, D. Jelev, Z. Krastev<br />

University Hospital St. Ivan Rilski, Sofia, Bulgaria<br />

Introduction: Observations have led to the description of distinct patterns of fibrosis,<br />

related to the underlying disorders causing fibrosis. It can therefore be expected that<br />

in this different patterns of fibrogenic evolution, different myofibroblast like subpopulation<br />

participate and distinct topographic localization of activated myofibroblasts<br />

preferentially contribute to the formation of fibrotic tissue.<br />

Aim: To evaluate the relationship between liver fibrotic stage in patients with chronic<br />

viral hepatitis B and the topographic distribution of alpha-smooth muscle actin (alpha-<br />

SMA)-positive myofibroblasts, divided in four compartments: portal/septal and in<br />

lobular zones 1, 2 and 3.<br />

Methods: Liver biopsies from 50 adult patients with HBV chronic hepatitis were<br />

assessed according to Metavir and immunohistochemical staining for alpha-SMA<br />

was performed on paraffin-embedded tissue sections. The number of a smooth<br />

muscle-positive cells in the lobules was scored using the system developed by<br />

Schmitt-Graff et al. Portal/septal expression of alpha-SMA was estimated by score<br />

proposed by us:<br />

0 alpha-SMA with expression in smooth cells within portal blood vessels only.<br />

1 A few alpha-SMA-positive cells beside vascular smooth muscle cells.<br />

2 alpha-SMA-positive expansion of most portal areas with short and thin septa of<br />

alpha-SMA-positive staining.<br />

3 One or few positive septa for alpha-SMA.<br />

4 alpha-SMA staining forms numerous septa.<br />

5 Pseudonodules embraced by alpha-SMA positive staining.<br />

Statistical analyses was done by Mann-Whitney and Spearman correlation test.<br />

Results: Strongest positive correlation was found between fibrotic stage and<br />

portal/septal expression of alpha-SMA (r = 0.757; p < 0.01), modest correlation was<br />

detected with periportal myofibroblast score (r = 0.427), while alpha-SMA staining in<br />

zone 2 showed week relationship (r = 0.256; p < 0.05). Expression of alpha-SMA in<br />

the pericentral part of the lobule did not shown any correlation with fibrotic stage.<br />

Discussion/Conclusion: It should be emphasized that counting the total number of<br />

myofibroblast, without taking in to account their topographic localization could blur<br />

the estimation of fibrogenic potential.


Is any role of hepatic steatosis in chronic virus B hepatitis?<br />

18<br />

Cătălina Mihai 1 , Cristina Cijevschi 1 , Gabriela Ştefănescu 1 , Vasile <strong>Dr</strong>ug 1 ,<br />

Mihaela <strong>Dr</strong>agna 1 , Lidia Graur 2 , Bogdan Mihai 2<br />

1<br />

Institute of Gastroenterology and Hepatology, University of Medicine and <strong>Pharma</strong>cy<br />

“Gr. T. Popa”, Iasi, Romania<br />

2<br />

Clinical Centre of Diabetes and Metabolic Diseases, University of Medicine and<br />

<strong>Pharma</strong>cy “Gr. T. Popa”, Iasi, Romania<br />

Introduction: It is well-known the influence of hepatic steatosis over fibrosis and the<br />

response to antiviral treatment in chronic hepatitis C, but limited data are available in<br />

chronic hepatitis B. The aim of this study was to evaluate the hepatic steatosis in<br />

patients with chronic hepatitis B and to correlate the steatosis with clinical, biological,<br />

virusological and histological factors.<br />

Methods: We studied 75 patients with chronic hepatitis B who were evaluated for<br />

antiviral treatment by liver biopsy between 1st January 2006 and 31st December<br />

2008. In group A there were no hepatic steatosis and in group B the hepatic steatosis<br />

was over 5%. We studied in these two groups the following parameters: age, sex,<br />

body mass index (BMI), glycaemia, lipidic profile, liver enzymes, the presence/absence<br />

of HBe antigen, co infection with D virus, viremia, the degree of necroinflammation<br />

and fibrosis.<br />

Results: There were 52 patients in group A (69%) and 23 (31%) in group B. The<br />

parameters which were correlated with the presence of hepatic steatosis (p < 0.05)<br />

were: age (mean value 48.6 in group B and 36.5 in group A), BMI (29.7 kg/m 2 in<br />

group B and 24.1 kg/m 2 in group A), glycaemia (109.6 mg% compared to 88.5 mg%),<br />

triglycerides (137.2 mg% compared to 87.4 mg%), necro-inflammatory score (> 8 –<br />

Metavir score – at 56.5% patients in group B and 36.5% in group A) and fibrosis<br />

(F3–F4 – Metavir score – at 39.13% patients in group B and 15.38% in group A). In<br />

group B were more male patients, more elevated liver enzymes and cholesterol, but<br />

without statistical significance. There were no differences between the two groups<br />

regarding the presence of HBe antigen (38.46% in group A and 43.47% in group B),<br />

the presence of virus D (7.69% in group A vs. 8.69% in group B), the level of viremia<br />

(mean value 767,000 U/ml in group A vs. 643,000 U/ml in group B).<br />

Discussion/Conclusion: In chronic hepatitis B the steatosis is present in 1/3 patients,<br />

related with metabolic and non viral factors. It seems to be a positive correlation<br />

between hepatic steatosis and severity of liver injury (necroinflammation and<br />

fibrosis). The complex therapeutic approach in these patients (metabolic and antiviral<br />

treatment) may improve the outcome of chronic hepatitis B.


19<br />

The effects of chronic B and C hepatitis infections on bone<br />

mineral density<br />

Corina Serban, Lelia Susan, Alina Pacurari, Loredana Copaceanu, Christian Banciu,<br />

Germaine Savoiu, Claudia Borza, Rodica Mateescu, Ioan Romosan<br />

UMF Timisoara, Romania<br />

Introduction: The importance of osteoporosis as a complication of liver disease is<br />

well known. The aim of this study was to evaluate bone mineral density in patients<br />

with chronic B and C hepatitis, hospitalised in one year period in The IVth Medical<br />

Clinic of University of Medicine and <strong>Pharma</strong>cy “Victor Babes”, Timisoara.<br />

Methods: Clinical data, hepatitis B and C status, and markers of bone metabolism<br />

were determined in 120 patients (39 men and 81 women; mean age 52 years). Bone<br />

mineral density was assessed in all patients at the lumbar spine (LS; L1–L4) and<br />

femoral neck (FN) region using X-ray absorbtiometry. Osteoporosis and osteopenia<br />

were defined by WHO criteria. The values were expressed as the T and Z score.<br />

Results: 86 of the patients were with hepatitis C (71.66%) and 34 of them with<br />

hepatitis B (28.34%). 43 patients were with osteoporosis (35.83%), 55 of them<br />

presented osteopenia (45.83%) and 22 of the patients were with normal BMD<br />

(18.33%). Mean T-score value was lower in patients with chronic hepatitis C (lumbar<br />

spine T = -1.15 ± 1.12, femoral neck T = -1.62 ± 1.01) compared with chronic<br />

hepatitis B (lumbar spine T = -0.75 ± 1.31, and femoral neck T = -1.35 ± 1.03). The<br />

values of serum and urine biochemical markers were normally in both groups.<br />

Discussion/Conclusion: This study reveals a high prevalence of osteoporosis in<br />

chronic hepatitis B or C patients. These secondary effects of chronic viral hepatitis<br />

should be further investigated.


20<br />

TGF-β-driven hepatic progenitor cell expansion and ductular<br />

reaction contributes to HBV, but not schistosomiasis,<br />

associated liver fibrogenesis<br />

Hong-Lei Weng 1 , Loredana Ciuclan 1 , Yan Liu 1 , Cheng Zhu 2 , Eliza Wiercinska 3 ,<br />

Iryna Ilkavets 1 , Johanna Dzieran 1 , Tong Huang 4 , Christoph Meyer 1 , Rolf Gebhardt 5 ,<br />

Rainer Heuchel 6,7 , Peter ten Dijke 3 , Jia-Lin Chen 8 , Manfred V. Singer 1 ,<br />

Peter R. Mertens 2 , Steven Dooley 1<br />

1 Molecular Hepatology – Alcohol-dependent Diseases, II. Medical Clinic Faculty of<br />

Medicine Mannheim, University of Heidelberg, Mannheim, Germany<br />

2 Department of Nephrology and Hypertension, Otto-von-Guericke-University,<br />

Magdeburg, Germany<br />

3 Department of Molecular Cell Biology and Center for Biomedical Genetics, Leiden<br />

University Medical Center, RC Leiden, The Netherlands<br />

4 Department of Cardiac Vascular Medicine, Affiliated Hospital, Medical School,<br />

Ningbo University, Ningbo, China<br />

5 Institute for Biochemistry, University of Leipzig, Leipzig, Germany<br />

6 Ludwig Institute for Cancer Research, Uppsala University, Uppsala, Sweden<br />

7 Department of Clinical Science, Intervention and Technology, Karolinska University<br />

Hospital, Huddinge, Sweden<br />

8 Department of Pathology, First Hospital of Jianxing, College of Jiaxing, Jiaxing,<br />

China<br />

Introduction: In HCV- and NASH-associated liver diseases, hepatic progenitor cells<br />

(HPCs) represent the major source of hepatocyte regeneration. Further, HPCs<br />

promote a periportal ductular reaction (DR) and contribute to periportal fibrogenesis.<br />

However, the factor(s) that mediate HPC activation are presently unknown. We<br />

hypothesize that TGF-β may drive HPC activation/DR in patients with chronic HBV<br />

infection.<br />

Methods: 110 biopsied liver tissues with chronic HBV infection were stained with<br />

cytokeratin-7 to quantify HPCs and DR, and with p21 to assess hepatocyte<br />

replication arrest.<br />

Results: Cytokeratin-7 showed significant correlation with inflammatory grades<br />

(r = 0.47, p < 0.001), fibrotic stages (r = 0.53, p < 0.001) and phopsho-Smad2<br />

staining, a marker of active TGF-β signaling (r = 0.24, p < 0.01). p21 positive<br />

hepatocytes were only found in 10 advanced fibrotic liver tissues, indicating that<br />

hepatocyte proliferation arrest is not required for HPC activation/DR in this disease.<br />

9-month IFN-γ treatment, which attenuated TGF-β signaling, significantly decreased<br />

HPCs in 13/18 patients with chronic HBV infection. Confocol microscopy analysis<br />

revealed co-localisation of phospho-Smad2/3 and HPC marker, indicating activated<br />

TGF-β signalling in these cells. Compared with wild type mice, M(2)-isozyme of<br />

pyruvate kinase, a marker for oval cells (HPCs in rodents) increased in TGF-β<br />

transgenic mice. The numbers of oval cells and p21 positive hepatocytes were<br />

significantly increased in Smad7 exon-1 deleted mice after 4-weeks of CCl4 treatment<br />

and were decreased in Smad7 transgenic mice after 2-week of bile duct ligation<br />

compared to wild types. In contrast to HBV patients, HPC activation was not<br />

associated with fibrotic degree in 42 Schistosoma japonicum infected patients.


Discussion/Conclusion: TGF-β signaling plays a crucial role in HPC activation/DR,<br />

which contributes to fibrogenesis in patients with chronic hepatitis B, but not<br />

schistosomiasis, suggesting etiology dependent differences in the mechanism of liver<br />

damage.


21<br />

Is there a link between the viral load value and liver stiffness<br />

for the asymptomatic virus B carriers?<br />

Adriana Tudora, O. Ciof, Mihaela Scurtu-Martin, I. Sporea, E. <strong>Dr</strong>agan, Diana Rill<br />

Department of Gastroenterology and Hepatology, University of Medicine and<br />

<strong>Pharma</strong>cy “Victor Babes”, Timisoara, Romania<br />

Introduction: To evaluate the role of transient elastography (FibroScan ® Echosense,<br />

Paris), as a noninvasive method to exclude significant liver fibrosis in nonreplicative<br />

virus B carriers compared to those presenting detectable viral load yet below<br />

10,000 copies/ml.<br />

Methods: 687 consecutive patients with viral B infection were evaluated (261 women<br />

– 38%, and 426 men – 62%), average age 43.4 ± 14.2 years. Out of this group,<br />

138 were defined as asymptomatic carriers based on normal ASAT, ALAT levels and<br />

viral load


22<br />

Evaluating antiinflammatory and antiviral effect due to the<br />

treatment with statins in patients with chronic virus C hepatitis<br />

Oana Andreescu 1 , Laurentiu Nedelcu 1 , Camelia Scarneciu 1 , Daniela Paul 2 ,<br />

Ileana Pantea 1<br />

1 Faculty of Medicine, University of Brasov, Romania<br />

2 Clinic Emergency Hospital of Brasov, Romania<br />

Introduction: Lately has been shown that statins can have more beneficial role than<br />

lowering lipid levels. Our study aimed to reveal antiinflammatory as well as antiviral<br />

effect due to the treatment with statins.<br />

Methods: The study was performed on 47 patients (31 women and 16 men aged<br />

between 40–82 years) diagnosed with chronic hepatitis with C virus who were nonresponders<br />

or without antiviral treatment at the beginning of the study. Patients were<br />

divided into 2 groups that received as treatment lovastatin and fluvastatin separately.<br />

All patients underwent clinical examination, viremia level and biochemical tests<br />

including a series of cytokines (IL-6, IL-8, IL-10, TNF-α, erythropoietin) at the<br />

beginning of the study and after 1 month.<br />

Results: Both groups registered especially lower level of TNF-α (56% cases that<br />

received lovastatin and 65% cases that received fluvastatin). The majority of the<br />

study group associated also lower levels of viremia (70% cases that received<br />

lovastatin and 83% cases that received fluvastatin)<br />

Discussion/Conclusion: Our study has demonstrated that the statins could have<br />

both antiinflammatory and antiviral effect which seems to be more important in case<br />

of administrating fluvastatin.


23<br />

Variants of the innate immune response genes in patients with<br />

chronic HCV-infection<br />

A.O. Romanov, T.V. Belyaeva, E.V. Esaulenko<br />

Pavlov State Medical University, Saint-Petersburg, Russia<br />

Introduction: The 2-5 oligoadenylate synthetase (OAS)-ribonuclease L (RNase L)<br />

pathway is a critical component of the immune response to viruses that constitutively<br />

presents in human cell. Double-stranded RNA structure within nontranslated regions<br />

HCV RNA activates OAS resulting in the activation of RNase L that cleaves viral<br />

RNA. The A allele at single nucleotide polymorphism (SNP) rs10774671 of the OAS1<br />

gene has been reported to be associated with low enzyme activity. Furthermore,<br />

1385G/A RNaseL variant showed a 3-fold decrease in RNase L activity. The aim of<br />

this study was to evaluate whether these functional polymorphisms are associated<br />

with the natural outcome of hepatitis C virus infection.<br />

Methods: Genotyping was performed in 76 patients with hepatitis C virus-induced<br />

cirrhosis, in 88 non-cirrhotic patients infected with HCV and in 122 healthy controls.<br />

The A/G OAS1 SNP rs10774671 was determined by PCR-RFLP assay using AluI<br />

restriction enzyme. The polymorphism +1385G/A in the RNaseL gene was detected<br />

by PCR-RFLP assay using a Msp20I site introduced into the forward primer next to<br />

the G to A transition.<br />

Results: There were no significant differences in the distributions of genotype and<br />

allele frequencies of A/G OAS1 SNP rs10774671 or +1385G/A RNASEL gene<br />

polymorphism between HCV patients with or without cirrhosis and between patient<br />

groups and healthy controls. Nevertheless, composite genotypes rs10774671 (AA or<br />

AG)/+1385(AG or AA) RNaseL had increased risk for cirrhosis among HCV-infected<br />

patients (OR 5.51; 95% CI 1.74–17.17).<br />

Discussion/Conclusion: The results of the study suggest that combinations of the<br />

variant genotypes of A/G OAS1 SNP rs10774671 and +1385 G/A RNaseL might<br />

affect the development of cirrhosis in persons with chronic HCV infection.


24<br />

Expression and effect of the NK cell receptor 2B4 (CD244) on<br />

virus-specific CD8+ T cells – Another player in the control of<br />

viral hepatitis?<br />

V. Schlaphoff, J. Jaroszewicz, S.V. Pothakamuri, J. Grabowski, K. Stegmann,<br />

M.P. Manns, H. Wedemeyer, M. Cornberg<br />

Department of Gastroenterology, Hepatology and Endocrinology, Hannover Medical<br />

School, Carl-Neuberg-Str. 1, 30625 Hannover, Germany<br />

This study aimed to investigate expression patterns and function of the costimulatory<br />

NK-cell receptor 2B4 on CD8+ T-cells and to elucidate a possible role in viral<br />

hepatitis.<br />

Expression of 2B4 was analyzed on lymphocytes and virus-specific CD8+ T-cells in<br />

healthy individuals by FACS. 2B4 was found on CD8+ T-cells (mean 35.6%, range<br />

11.9–52.5%), CD56+ NK-cells (100%) and on monocytes (100%), but was only rarely<br />

present on CD4+ T-cells (mean 0.92%, range 0.18–2.1%). A clear activation and<br />

memory phenotype could be seen identifying 2B4+CD8+ T-cells as effector or<br />

effector memory cells. Isolated 2B4+CD8+ cells showed a decreased survival<br />

in-vitro, whereas they displayed strong degranulation upon CD3/28 stimulation in<br />

contrast to 2B4-CD8+ cells.<br />

Expression of other inhibitory molecules PD-1 or CTLA-4 correlated with expression<br />

of 2B4. Similarly, CD161-, CD38- or CD57-positive CD8+ T-cells uniformly expressed<br />

2B4. In contrast, not all 2B4+CD8+ T-cells showed expression of these markers.<br />

2B4 expression was further investigated on PBMCs from patients with chronic HCV<br />

(n = 60) or HBV (n = 26), acute HCV or HBV infection (n = 6 and n = 10) and patients<br />

with AIH, PBC or PSC (n = 15). 2B4+CD8+ T-cells ex-vivo frequency was higher in<br />

patients with chronic HCV, HBV or AIH as compared with healthy individuals (46.9%,<br />

52.8% and 50.6% vs. 35.6%, p > 0.02).<br />

Analyzing virus-specific T-cells for 2B4, CMV- and EBV-specific CD8+ T-cells<br />

showed high expression of 2B4, while Influenza-A-specific CD8+ T-cells were mostly<br />

negative for 2B4. HCV- and HBV-specific CD8+ T-cells were mainly 2B4-positive<br />

during acute infection, while in chronic HCV different patterns of 2B4 on virus-specific<br />

CD8+ T-cells were observed.<br />

In conclusion, 2B4 expression correlates with an end-stage effector type T-cell and<br />

distinct early activation phenotype. Higher frequencies of 2B4+CD8+ T-cells during<br />

acute and persistent viral infection suggest a regulatory role during antigen exposure.<br />

We propose that besides PD-1, CTLA-4 and other receptors the costimulatory<br />

molecule 2B4 is involved in the immune control of hepatitis C and B.


25<br />

Ethanol impairs MHC class I-restricted antigen presentation on<br />

HCV-infected liver cells<br />

Natalia Osna 1 , Ronda White 1 , Steven Weinman 2 , Terrence Donohue 1<br />

1 University of Nebraska Medical Center, Omaha, NE, USA<br />

2 University of Kansas Medical Center, Kansas City, MO, USA<br />

Introduction: Elimination of HCV-infected hepatocytes by cytotoxic T-cells is<br />

restricted to efficient MHC class I-restricted antigen presentation of viral peptides<br />

generated by proteasome. Previously, we have shown that in cultured hepatoma<br />

cells, the combination of HCV core protein and ethanol suppresses proteasome<br />

activity (Gastroenterology, 2008). This study seeks to determine the level ethanolmediated<br />

proteasome suppression in liver cells of HCV core-expressing mice and its<br />

link to impaired antigen presentation in hepatocytes.<br />

Methods: Hepatocytes isolated from livers of HCV core positive C57Bl/6 mice were<br />

exposed overnight to 50 mM ethanol in the presence or absence of IFN-γ and then<br />

C-extended SIINFEKL peptide was delivered to cytoplasm. After processing by<br />

proteasome, SIINFEKL-H2Kb complex was presented on hepatocytes and visualized<br />

with specific antibody by flow cytometry<br />

Results: Ethanol exposure suppressed SIINFEKL-H2Kb presentation on HCV core<br />

positive hepatocytes and these effects were reversed in the presence of N-acetyl<br />

cysteine (NAC), indicating the involvement of oxidative stress. High level of oxidative<br />

stress in HCV-expressing ethanol-exposed hepatocytes was confirmed by 5-fold<br />

induction of ROS. Furthermore, SIINFEKL-H2Kb presentation was abrogated by<br />

treatment with the proteasome inhibitor, Mg132, demonstrating the critical role of this<br />

enzyme for SIINFEKL processing. Proteasome activity was also suppressed in<br />

hepatocytes by ethanol exposure in ethanol metabolism-dependent manner.<br />

Discussion/Conclusion: We conclude that the combination of HCV core protein and<br />

ethanol creates high oxidative stress, leading to the reduction of proteasome activity<br />

and inability of the enzyme to efficiently generate peptides for MHC class I-restricted<br />

antigen presentation in hepatocytes


26<br />

Impairment of interferon-gamma (IFN-γ) signaling in hepatocytes<br />

by HCV proteins and ethanol<br />

Lee Austin 1 , Ronda White 1 , Kusum Kharbanda 1 , Michael Beard 2 , Natalia Osna 1<br />

1 University of Nebraska Medical Center, Omaha, NE, USA<br />

2 University of Adelaide, Adelaide, Australia<br />

Background: In HCV patients, alcohol consumption aggravates the course of<br />

hepatitis and enhances the viral load. IFN-γ signaling in liver cells regulates the main<br />

stages in antigen presentation, where HCV-infected hepatocytes serve as the targets<br />

for immune cells. This study is aimed to determine IFN-γ signaling in HCV-expressing<br />

liver cells exposed to ethanol.<br />

Methods: C5B cells are replicon Huh7-based cells that express full genome of HCV<br />

proteins and an ethanol metabolizing enzyme, CYP2E1. After cell exposure to<br />

ethanol, STAT1 phoshorylation was induced by IFN-γ and measured differentially in<br />

cells, which express HCV proteins, NS5B or core, and in non- or low-expressing cells<br />

by Phoshoflow. In addition, the ratio of pSTAT1/STAT1 was measured by Western<br />

blot in lysates of IFN-γ-exposed C5B and CYP2E1 + parental Huh7 cells.<br />

Results: IFN-γ-induced pSTAT1 was lower in HCV + C5B cells than in Huh7 cells.<br />

Ethanol exposure further suppressed pSTAT1. Importantly, STAT1 phoshorylation<br />

was depended on expression of HCV protein expression in C5B cells: It was<br />

abrogated in HCV-expressing C5B cells, while in non- or low-expressing C5B cells<br />

the level of pSTAT1 was decreased by ethanol treatment. However, these effects of<br />

ethanol were reversed in the presence of ethanol metabolism inhibitor, 4 methyl<br />

pyrazole or N-acetyl cystein, indicating that ethanol metabolism-induced oxidative<br />

stress was responsible for the pSTAT1 reduction.<br />

Discussion/Conclusion: We conclude that HCV protein expression and ethanol<br />

metabolism both suppress IFN-γ signaling in hepatocytes, providing an implication<br />

for the impaired antigen presentation in liver cells of HCV + alcohol drinkers.


27<br />

The platelet-derived chemokine CXCL4 (PF4) is a mediator of<br />

HCV-induced and experimental liver fibrosis<br />

Mirko Moreno Zaldivar, Katrin Pauels, Marie-Luise Berres, Petra Schmitz,<br />

M. Anna Kowalska*, Ralf Weiskirchen, Christian Trautwein, Hermann E. Wasmuth<br />

Medical Department III, University Hospital of Aachen, Aachen, Germany<br />

*University of Pennsylvania School of Medicine, Philadelphia, PA, USA<br />

Background: Liver fibrosis is a major cause of mortality and morbidity worldwide.<br />

Platelets are involved in liver damage but the underlying molecular mechanisms<br />

remain elusive. We here investigate the platelet chemokine CXCL4 (platelet factor 4)<br />

as a molecular mediator of fibrotic liver damage.<br />

Methods: Serum concentrations and intrahepatic mRNA of CXCL4 were measured<br />

in patients with hepatitis C. Platelet infiltration in the liver was visualized by electron<br />

microscopy in humans. Cxcl4 -/- and wild-type mice were subjected to two models of<br />

chronic liver injury (CCl4 and TAA). The fibrotic phenotype of these mice was<br />

analysed by histological, biochemical and molecular analyses. Intrahepatic infiltration<br />

of immune cells was investigated by FACS and immunohistochemistry. Isolated<br />

myofibroblasts were stimulated with recombinant Cxcl4.<br />

Results: Patients with advanced HCV-induced fibrosis had increased PF4 serum<br />

levels and intrahepatic CXCL4 mRNA concentrations. CCl4 and TAA treatment led to<br />

an increase of hepatic Cxcl4 levels and activation of platelets in the peripheral blood<br />

in mice. Accordingly, genetic deletion of Cxcl4 in mice significantly reduced<br />

histological and biochemical damage in vivo. Amelioration of liver damage in Cxcl4 -/mice<br />

was functionally associated with changed expression of fibrosis-related genes<br />

(Timp-1, Mmp9, Tgf-β, IL-10). Functionally, Cxcl4 -/- mice had a strongly decreased<br />

infiltration of neutrophils (Ly6G) and CD8 + T-cells into the liver. In vitro, recombinant<br />

murine Cxcl4 stimulates the proliferation of isolated hepatic myofibroblasts.<br />

Conclusions: These results functionally underscore the important role of the platelet<br />

derived chemokine CXCL4 in chronic liver damage and imply a new pathway for antifibrotic<br />

therapies.


28<br />

Interleukin 17 pathway is suppressed in chronic hepatitis C<br />

Gutkowski Krzysztof 1 , Kacperek-Hartleb Teresa 1 , Musialik Joanna 1 ,<br />

Boryczka Grzegorz 1 , Kajor Maciej 2 , Hartleb Marek 1 , Kamińska-Kiszka Ewelina 3<br />

1<br />

Department of Gastroenterology and Hepatology, Medical University of Silesia,<br />

Katowice, Poland<br />

2<br />

Department of Pathomorphology, Medical University of Silesia, Katowice, Poland<br />

3<br />

Department of Biotechnology, University of Rzeszow, Poland<br />

Introduction: Immune dysregulations in hepatitis C are still unclear, especially<br />

regarding inflammatory pathways. Recent studies showed that TGF-β alone<br />

promoted the generation of anti-inflammatory regulatory T lymphocytes (T-reg) from<br />

naive T CD4 + cells, whereas TGF-β in the presence of IL-6 promoted Th17 cells.<br />

Materials and methods: We studied IL-17 serum levels and its activators (IL-6,<br />

TGF-β) in 36 chronic hepatitis C patients and 33 age and gender-matched healthy<br />

subjects. The results in HCV patients were related to liver histology. Cytokine levels<br />

were assessed by enzyme-linked immunosorbent assays (ELISA).<br />

Results:<br />

Cytokines<br />

Healthy subjects<br />

(n = 33)<br />

Hepatitis C patients<br />

(n = 36)<br />

Mean ± SD Range Mean ± SD Range p<br />

IL-17 [pg/ml] 16.5 ± 6.6 9.6–32.6 12.5 ± 12.2 0.0–48.2 0.0214<br />

IL-6 [pg/ml] 1.7 ± 1.2 0.5–6.6 16.7 ± 7.2 8.4–39.7 0.3732<br />

TGF-β [ng/ml] 30.9 ± 7.4 16.8–46.2 47.9 ± 23.1 18.2–116.0 0.0008<br />

Mann-Whitney test<br />

Serum levels of IL-6 did not differ between healthy subjects and HCV patients but<br />

positively correlated in hepatitis C cohort with stage of fibrosis (r = 0.45; p = 0.005).<br />

Conclusions:<br />

1. Chronic hepatitis C down-regulates Th17 response.<br />

2. TGF-β level is increased in HCV patients and may be responsible for suppressing<br />

of IL-17 production by Th17 cells.


29<br />

Allelic expression of rs2569190/C-159T CD14 gene variants in<br />

PBMC and liver samples from chronic hepatitis C patients<br />

Rusudan Bregadze, Giuliano Ramadori, Sabine Mihm<br />

Department of Gastroenterology and Endocrinology, University Medical Center<br />

Göttingen, Germany<br />

Introduction: The single nucleotide polymorphism (SNP) rs2569190 within the<br />

endotoxin receptor CD14 gene has been described to be associated with the<br />

progression of alcoholic liver disease. This association was suggested to be due to<br />

sensitization of the liver for gut-derived endogenous LPS via an enhanced T allele<br />

mediated hepatic CD14 expression. In parallel to association studies on this SNP<br />

and disease progression in chronic hepatitis C, this study aimed at investigating<br />

allelic expression of CD14 transcripts in PBMC and liver samples from chronically<br />

hepatitis C virus (HCV) infected patients.<br />

Methods: Genotyping was performed by allelic discrimination in 5’-nuclease<br />

reactions. Allelic expression was investigated by allele specific transcript quantification<br />

(ASTQ) using a reporter SNP within the transcript to discriminate mRNAs<br />

originating from rs2569190 C or T allele gene variants in peripheral blood<br />

mononuclear cell (PBMC) and liver biopsy samples from rs2569190- and reporter<br />

SNP-heterozygous hepatitis C patients (n = 3). Samples from rs2569190homozygous<br />

(but reporter SNP-heterozygous) patients served as controls (n = 5).<br />

Results: In 2 of 3 PBMC samples a preferential expression of the T allele gene<br />

variant less than 2-fold could be seen. In liver biopsy specimens a slight overexpression<br />

of C and T allele gene variants could be observed in 2 samples, respecttively.<br />

Among the 5 homozygous controls, however, allelic imbalances in gene<br />

expression to a similar extent were found.<br />

Discussion/Conclusion: ASTQ data on SNP rs2569190 thus support the absence<br />

of an association between rs2569190 genotype and liver fibrosis progression in<br />

patients with chronic hepatitis C.


30<br />

The content of sphingolipids in the liver is associated with the<br />

severity of chronic hepatitis C<br />

M. Dabrowska 1 , A. Panasiuk 1 , P. Zabielski 2 , J. Górski 2 , R. Flisiak 1<br />

1<br />

Department of Infectious Diseases and Hepatology, Medical University of Bialystok,<br />

Bialystok, Poland<br />

2<br />

Department of Physiology, Medical University of Bialystok, Bialystok, Poland<br />

Background and aims: Sphinganine, ceramide and sphingosine are known as<br />

proapoptotic factors, whereas sphingosine-1-phosphate augments cell proliferation<br />

and suppresses apoptosis. The aim of study was to assess the association between<br />

liver sphingolipids content and the activity of chronic HCV infection.<br />

Methods: The study was performed in 38 HCV (+) treatment naïve patients<br />

(11 female and 27 male, aged from 19 to 62 years). Plasma HCV RNA was<br />

evaluated using RT-PCR. The inflammatory activity and fibrosis were graded<br />

according to Scheuer histological classification. Sphinganine, ceramide, sphingosine<br />

and sphingosine-1-phosphate contents as well as neutral and acid ceramidases, and<br />

neutral and acid sphingomyelinases activities were measured in liver tissues. The<br />

significance of differences was calculated by non-parametric U Mann-Whitney and<br />

Kruskal-Wallis test. For correlation analysis, the Spearman nonparametric correlation<br />

was used.<br />

Results: We observed the significant increase in sphinganine content which was<br />

associated with severity of portal/periportal inflammation (14.7 ± 9.0 vs. 21.0 ± 10.0<br />

pmol/mg of protein, p = 0.015) and fibrosis advancement (17.0 ± 10.0 vs. 19.5 ± 6.9<br />

pmol/mg of protein, p < 0.05). Moreover, the decrease in sphingosine-1-phosphate<br />

content was related to inflammatory activity. The positive correlation between<br />

sphinganine and sphingosine and biochemical markers of liver injury (for ALT:<br />

r = 0.52 and r = 0.41 respectively, p < 0.01) were noticed. Additionally, the positive<br />

correlation between spinganine and sphingosine content was observed (r = 0.71,<br />

p < 0.001). Plasma HCV RNA was negatively correlated with ceramide (r = -0.59,<br />

p = 0.02).<br />

Conclusions: Sphinganine and sphingosine could play an important function in<br />

immune mediated liver injury and fibrosis progression during chronic HCV infection.


31<br />

Ribavirin-induced anaemia during hepatitis C treatment is<br />

related to baseline haemoglobin level and drug erythrocyte<br />

concentration<br />

M. Delle Monache 1 , M. Carbone 2 , L. Nosotti 3 , L.R. Conti 4 , I. Lenci 2 , F. De Leonardis 2 ,<br />

M. Angelico 2 and L. Baiocchi 2<br />

1<br />

Hepatology Service, ASL RMG, Palestrina, Rome, Italy<br />

2<br />

Hepatology Unit, Department of Internal Medicine, University of Rome “Tor Vergata”,<br />

Rome, Italy<br />

3<br />

Department of Preventive Medicine S. Gallicano Hospital (IRCCS), Rome, Italy<br />

4<br />

Hepatology Service, S. Spirito Hospital, Rome, Italy<br />

Introduction: Ribavirin-induced haemolytic anaemia is a frequent adverse event<br />

during treatment of chronic hepatitis C but the mechanisms are not fully understood.<br />

Few human studies have evaluated the factors influencing haemolysis in patients<br />

undergoing standard antiviral therapy for hepatitis C, providing conflicting results.<br />

Plasma and erythrocyte ribavirin levels were seldom assessed, with inconclusive<br />

findings possibly due to the small number and the non-homogeneity of the patients<br />

studied. We evaluated plasma and erythrocyte ribavirin levels and other factors<br />

potentially associated with haemolytic anaemia in patients undergoing treatment with<br />

Peginterferon alfa 2a (Peg-IFN) plus ribavirin for hepatitis C.<br />

Methods: Thirty patients (12/18; mean age 44.6 ± 13; 26 with genotype 1 or 4)<br />

undergoing a standard treatment schedule for hepatitis C (Peg-IFN 180 µg weekly +<br />

ribavirin 1000 or 1200 mg daily according to body weight) were included in the study.<br />

Plasma and erythrocyte ribavirin levels were evaluated between week 6 and 8 of<br />

treatment (when a plateau in the drug levels has been consistently reported).<br />

Ribavirin concentration was measured by high performance liquid chromatography<br />

after solid phase extraction and employing 3-methyl-cytidine as internal standard.<br />

Results: The mean daily ribavirin administered dose (14.9 ± 2.2 mg/kg) was<br />

unrelated to the plasma (7.7 ± 4.6 µm/l) and erythrocyte (1592 ± 826 µm/l) ribavirin<br />

concentration, nor to the extent of haemoglobin drop during treatment, which<br />

averaged 3.4 ± 1.4 g/dl. Plasma ribavirin levels were significantly correlated with drug<br />

erythrocyte concentrations (r = 0.4; p = 0.03). High haemoglobin baseline levels and<br />

high erythrocyte ribavirin concentrations were both independently associated with a<br />

greater extent of haemolysis (p = 0.001 and p = 0.01, respectively) at multiple<br />

regression analysis.<br />

Discussion/Conclusion: The extent of ribavirin-induced anaemia during antiviral<br />

treatment for hepatitis C is correlated to the baseline haemoglobin level and to the<br />

drug erythrocyte concentrations rather than to the administered dose.


32<br />

Sylibin-vitamin E phospholipid complex reduces increased<br />

ferritin levels during antiviral treatment for chronic hepatis C<br />

R. Cecere 1 , L. Baiocchi 2 , I. Lenci 2 , M. Carbone 2 , F. De Leonardis 2 , and<br />

M. Delle Monache 1<br />

1 Hepatology Service ASL RMG, Palestrina, Rome, Italy<br />

2 Hepatology Unit, Department of Internal Medicine, University “Tor Vergata”, Rome,<br />

Italy<br />

Introduction: A common finding during antiviral therapy for chronic hepatitis C is an<br />

increase in serum ferritin levels, likely determined by ribavirin-induced haemolysis or<br />

necro-inflammatory response to Peg-IFN. A persistent hepatic iron overload, in this<br />

setting, may increase liver injury and progression toward cirrhosis. Aim of this study<br />

is to investigate the role of iron overload during antiviral therapy for chronic hepatitis<br />

C and the effect of sylibin-vitamin E complex on this.<br />

Methods: Fifty-five consecutive patients underwent treatment with Peg-IFN plus<br />

ribavirin. Patients with ferritin > 500 µg/l during antiviral therapy were treated with<br />

sylibin-vitamin E phospholipid complex (SVEP) 188 mg daily for 3 months.<br />

Afterwards patients with persistently high ferritin level, transferrin saturation > 45%<br />

and raised ALT were treated with desferroxamine 30 mg/kg, iv daily, 5 days a week<br />

plus 250 mg vitamin C for additional 2 months. Biochemical and virologic data were<br />

analyzed before, during and after therapy.<br />

Results: The clinical data of 55 patients (M/F 36/19, mean age 50 ± 10) were<br />

reviewed. Ten patients (18%) had high basal level of ferritin (> 300 µg/l). During<br />

antiviral therapy a consistent increase in serum ferritin was present in thirty out of<br />

55 patients (54.5%), while haemoglobin level significantly decreased (-17%, p < 0.05)<br />

in the whole sample. In all these subjects SVEP was administered. There was a<br />

significant decrease in serum ferritin from beginning to end of treatment with SVEP<br />

( -25.6%, p < 0.05). Only four patients maintained ferritin levels > 500 µg/l, transferrin<br />

saturation > 45%, associated with a flare of ALT (x 4–6 ULN). HCV-RNA was<br />

undetectable in these patients during ALT flare. In these patients desferroxamine<br />

was added. Two patients did not respond to treatment: they were male, infected by<br />

GT3 HCV and heterozygote for H63D mutation of HFE gene.<br />

Discussion/Conclusion: Increase of ferritin level during antiviral therapy for HCV is<br />

a frequent finding and may be a concomitant expression of ribavirin-induced<br />

haemolysis and necro-inflammatory activity, or response to Peg-IFN. Therapy with<br />

SVEP is associated with a significant reduction in serum ferritin. When levels are<br />

very high (≥ 500), associated with ALT flare, and transferrin saturation > 45%, liver<br />

injury by iron overload should be considered, as well as association with chelating<br />

therapy. In this case, screening for HFE gene mutations is to advice.


33<br />

Thyroid function in patients with chronic hepatitis C virus<br />

infection under interferon therapy<br />

Muhammad Azmy*, Youssef Ahmad*, Abedelraouf AbouNar**, and<br />

Manal El-Hamamsy***<br />

Departments of Internal Medicine*, Clinical Pathology** and Clinical <strong>Pharma</strong>cy***,<br />

Faculty of Medicine, Al-Azhar University and Faculty of <strong>Pharma</strong>cy, Ain Shams<br />

University, Egypt<br />

Introduction: The aim of the present study was to investigate the effect of interferon<br />

therapy on the thyroid function on Egyptian patients with HCV infection.<br />

Methods: The study included 60 HCV infected patients with normal baseline levels<br />

of TSH. The patients received a subcutaneous pegylated interferon alfa-2b weekly in<br />

addition to oral ribavirin. Before the start of interferon therapy, serum TSH,<br />

thyroglobulin-Ab (TG-Ab) and antiperoxidase antibodies (TPO-Ab) were measured.<br />

Three months after interferon therapy serum levels of TSH were performed to all<br />

patients, patients with abnormal TSH were suspected to the measurements of FT3,<br />

FT4, TPO-Ab, TG-Ab and thyroid stimulating immunoglobulin levels (TSI).<br />

Results: After 3 months of interferon therapy, 48 patients (80%) had normal TSH<br />

and 12 patients (20%) had abnormal TSH. Out of 12 patients had abnormal TSH,<br />

10 patients (16.6%) had high serum levels of TSH (hypothyroidisms), while the<br />

remaining 2 patients (3.4%) had low serum levels of TSH (hyperthyroidism). All<br />

patients with abnormal TSH had significant higher levels of TG-Ab, TPO-Ab and STI<br />

(in cases with hyperthyroidism only) than the patients with normal levels of TSH. The<br />

levels of TPO-Ab of patients with abnormal TSH were above the normal reference<br />

range. In our study 40 patients (66.6%) were responder to interferon therapy (-ve<br />

PCR for HCV), while 20 patients (33.3%) were non-responder.<br />

Discussion/Conclusion: The incidence of thyroid dysfunction during pegylated<br />

interferon therapy in patients with HCV is 20%; hypothyroids was more common than<br />

hyperthyroidism. The patients most at risk for thyroid dysfunctions are female sex<br />

and people with preexisting TPO-Abs. The numbers of non-responder patients were<br />

significantly higher in patients with abnormal TSH than those with normal TSH.<br />

Patients with HCV infection under pegylated interferon and ribavirin therapy should<br />

be screened for thyroid dysfunction before and during treatment.


34<br />

The prevalence and impact of serum autoantibodies among<br />

chronic hepatitis C genotype 4 (HCV) patients attending a<br />

national center of HCV treatment program in Egypt<br />

Mohamed Sharaf-Eldin, and Hassan El-Batae<br />

Tropical Medicine and Infectious Diseases Department, Faculty of Medicine, Tanta<br />

University, Egypt<br />

Introduction: Non-organ-specific autoantibodies (NOSAs) are commonly detected in<br />

HCV. The aim was to determine the prevalence of anti-nuclear (ANA), anti-smooth<br />

muscle (ASMA) and anti-liver kidney microsomes type 1 (anti-LKM1) antibodies in<br />

HCV and assess their impact on the response to antiviral therapy.<br />

Methods: One thousand and sixty two HCV patients were reviewed retrospectively<br />

for autoantibodies and response to treatment. All patients received combined antiviral<br />

therapy in the form of pegylated interferon-alpha 2 (a or b) plus ribavirin (1000–1200<br />

mg/day) for 48 weeks.<br />

Results: Non-organ-specific autoantibodies (NOSAs) were observed in 308 patients<br />

(29%): ANA in 195 (18.36%), ASMA in 142 (13.37%) and anti-LKM1 was the rarest<br />

occurring in 16 (1.51%). Concomitant positivity for ANA and ASMA was observed in<br />

57 of these 308 cases (5.37%). The presence of NOSAs was associated with higher<br />

aspartate transaminase (AST), alanine transaminase (ALT) and γ-globulin. In<br />

contrast no differences were observed regarding age, gender and viral load. End<br />

treatment response in HCV patients with NOSAs was 59.47% while it was 64.57% in<br />

patients negative to NOSAs, and this difference was non significant.<br />

Discussion/Conclusion: ANA was the commonest while anti-LKM1 was the rarest<br />

among HCV patients. NOSAs were associated with more necroinflammatory grades.<br />

Autoantibodies did not predict any change in response to HCV treatment.


35<br />

Absence of a correlation between endotoxin receptor CD14<br />

rs2569190/C-159T polymorphism and liver disease progression<br />

in chronic hepatitis C<br />

Eva Askar 1 , Margarete Odenthal 2 , Hans-Peter Dienes 2 , Giuliano Ramadori 1 ,<br />

Sabine Mihm 1<br />

1<br />

Department of Gastroenterology and Endocrinology, University Medical Center<br />

Göttingen, Germany<br />

2<br />

Institute of Pathology, University of Cologne, Germany<br />

Introduction: Chronic hepatitis C is characterized by a mostly mild hepatic inflammatory<br />

activity, which holds, however, a significant risk to proceed into liver cirrhosis<br />

and hepatocellular carcinoma. The liver is constantly exposed to gut-derived bacterial<br />

lipopolysaccharides (endotoxin), which are suggested to be cofactors in alcoholic<br />

liver disease (ALD) through exacerbating ongoing injury. Recently, the T allele of a<br />

single nucleotide polymorphism (SNP) in the endotoxin receptor CD14 gene has<br />

been reported to be related to susceptibility to ALD-related cirrhosis. This study<br />

aimed at analyzing the correlation between rs2569190/CD14 C-159T and disease<br />

progression in chronic hepatitis C.<br />

Methods: Liver biopsy specimens from a total of 349 chronic hepatitis C patients,<br />

(201 men, mean age 45.8 ± 13.5 years), provided by the German Network of<br />

Competence for Hepatitis (Hep-Net), were evaluated with respect to necroinflammatory<br />

activity (grading) and architectural changes (staging). Samples of genomic<br />

DNA were genotyped for the respective SNP by 5’-nuclease assays using fluorescent<br />

dye-labelled allele-specific probes.<br />

Results: The genotype distribution (CC:CT:TT) of 109:170:70, respectively, followed<br />

Hardy-Weinberg equilibrium (p = 0.828), and was close to that given for Caucasians<br />

in public data bases. Demographic analysis revealed no significant relationship<br />

between genotypes and gender or age. Severe hepatitis activity was found in 14.9%<br />

of patients, and cirrhosis in 6.3% of them. Histopathological analysis showed no<br />

relationship between rs2569190/C-159T CD14 genotypes and hepatitis activity<br />

(p = 0.513) or fibrosis progression (p = 0.928).<br />

Discussion/Conclusion: Association data argue against a possible effect of CD14<br />

rs2569190/C-159T polymorphism on promoting hepatitis activity or fibrosis<br />

progression in chronic hepatitis C.


36<br />

The metabolic control and antiviral treatment response in<br />

diabetic patients with chronic hepatitis C<br />

Bogdan Mihai 1 , Catalina Mihai 2 , Cristina Cijevschi 2 , Vasile <strong>Dr</strong>ug 2 , Mihaela <strong>Dr</strong>agna 2 ,<br />

Lidia Graur 1<br />

1 Clinical Centre of Diabetes and Metabolic Diseases, University of Medicine and<br />

<strong>Pharma</strong>cy “Gr. T. Popa”, Iasi, Romania<br />

2 Institute of Gastroenterology and Hepatology, University of Medicine and <strong>Pharma</strong>cy<br />

“Gr. T. Popa”, Iasi, Romania<br />

Introduction: Diabetes mellitus (DM) represents a poor prognostic factor for antiviral<br />

therapy response in chronic hepatitis C patients. The aim was to study the diabetic<br />

factors which can influence treatment response in patients with chronic hepatitis C<br />

and DM.<br />

Methods: We studied 50 patients with DM who underwent antiviral therapy with<br />

pegylated interferon and ribavirin for chronic hepatitis C in the Institute of<br />

Gastroenterology and Hepatology Iasi, between 1st January 2007 and 31st<br />

December 2008. We analyzed the following parameters: type of DM, glycated<br />

hemoglobin (HbA1c) determined at the beginning of antiviral therapy, lasting of DM,<br />

macro- and microangiopathic complications. The “cut off” value of HbA1c was<br />

considered 7%. A sustained response was defined as the absence of detectable<br />

hepatitis C virus RNA 6 months after treatment was stopped.<br />

Results: There were 29 females and 21 men, aged between 21 and 60 years (mean<br />

age 38.7 years). 20 patients had Type 1 DM and 30 Type 2 DM. The global rate of<br />

response was 60%. In Type 1 DM were 35% non-responders, and in Type 2 –<br />

43.33%. Abnormal HbA1c was found in 23.33% responders and in 80% nonresponders.<br />

Macroangiopathic complications were found in 30% responders and<br />

40% non-responders. Microangiopathic complications were found in 26.66%<br />

responders and 35% non-responders. There were 11 patients (36.66%) with more<br />

than 10 years of DM in responders and 9 patients (45%) in non-responders. HbA1c<br />

was the only statistical significant factor (p < 0.05) between the responders vs. nonresponders.<br />

Discussion/Conclusion: HbA1c is an important predictive factor to antiviral therapy<br />

response in diabetic patients with chronic hepatitis C. There are more nonresponders<br />

patients with Type 2 DM comparative with Type 1, probably due to<br />

insulin-resistance. All patients with DM and chronic hepatitis C need a good glycemic<br />

control before starting antiviral therapy.


37<br />

The correlation between erythropoietin and pro- and antiinflammatory<br />

cytokines and the decreasing of the viral C<br />

hepatitis RNA level under the treatment with statins<br />

R. Mihăilă 1 , R. Mihăilă 2 , O. Frăţilă 3 , E.C. Rezi 4 , A. Zaharie 4 , L. Mocanu 4 , M. Deac 2 ,<br />

A. Olteanu 4 , L. <strong>Dr</strong>aghila 4 , M. Boitan 2<br />

1 Public Health Authority, Sibiu, Romania<br />

2 Lucian Blaga University Sibiu, Medical Faculty, Sibiu, Romania<br />

3 University of Oradea, Medical Faculty, Oradea, Romania<br />

4 Emergency Clinical County Hospital, Sibiu, Romania<br />

Introduction: Some statins inhibit the RNA replication and the assemblage of the<br />

replicative hepatitis C virus complex.<br />

Aim: Our aim was to study if there is any correlation between erythropoietin, pro- and<br />

anti-inflammatory cytokines and the answer of the C virus viral load at the treatment<br />

with fluvastatin.<br />

Methods: We have studied a group of 37 consecutive patients with viral C chronic<br />

hepatitis, who were not under treatment with interferon, to whom we have determined<br />

the serum levels of IL-6, IL-8, TNF-alpha, IL-10, erythropeietin, the liver biochemical<br />

tests and the viral load (determined by real time PCR), before and after a month of<br />

treatment with fluvastatin 40 mg/day. We have compared the results of the biological<br />

parameters with the initial and final viral load and with its variations and we have<br />

statistically analyzed the results (average, standard deviation, Pearson’s test and t<br />

student test).<br />

Results: The medium age of the studied group was 49.26 ± 10 years. The gender<br />

repartition was 25 women and 12 men. In the whole studied group, the viral load<br />

decreased, in average with 1,592,050 ± 3,654,745 UI/ml (p = 0.016). There was a<br />

direct linear correlation between the final IL-6 level and the initial viral load<br />

(r = 0.254). The viral load variation correlated directly with the initial level of IL-6<br />

(r = 0.250) and inversely with the initial level of erythropoietin (r = -0.280). The final<br />

viral load correlated directly with the initial level of erythropoietin (r = 0.437) and<br />

inversely with the initial level of IL-8 (r = -0.286). Also, the initial IL-6 level correlated<br />

directly with the initial AST level (r = 0.424) and inversely with the initial cholesterol<br />

level (r = -0.288). There was an inverse linear correlation between the initial IL-8 level<br />

and the initial level of creatinine (r = -0.322).<br />

Discussion/Conclusion: The initial level of IL-6, a pro-inflammatory cytokine, is<br />

directly correlating with the AST level and inversely correlating with the initial level of<br />

cholesterol and it is a favoring prognostic factor for the decreasing of the viral load<br />

during the treatment with fluvastatin. The higher initial IL-8 level is the lower final viral<br />

load is. The higher initial erythropoietin level is the lower is the decreasing viral load<br />

under the treatment with fluvastatin.<br />

Acknowledgments: This study belongs to a complex research grant (179/2006),<br />

financed by the Research and Education Ministry from Romania, through the<br />

Academy of Medical Science and VIASAN, to whom we are deeply gratefully.


38<br />

Increased CD14 transcription initiation in T allele carriers of the<br />

rs2569190/C-159T promoter polymorphism in healthy individuals<br />

and chronic hepatitis C patients<br />

Jasmin Mertens, Giuliano Ramadori, Sabine Mihm<br />

Department of Gastroenterology and Endocrinology, University Medical Center<br />

Göttingen, Germany<br />

Introduction: Lipopolysaccharide, a main component of the outer cell wall of Gramnegative<br />

bacteria, is recognized by an endotoxin receptor complex of Toll-like<br />

receptor 4, MD2 and CD14. The single nucleotide polymorphism (SNP)<br />

rs2569190/C-159T in the CD14 gene in relation to liver disease progression has<br />

been investigated in various association studies. Furthermore, in vitro experiments<br />

have shown an enhanced promoter activity for T allele reporter gene constructs. This<br />

study aimed at analysing a possible relationship between this SNP and CD14<br />

transcription initiation in healthy individuals and in chronic hepatitis C patients in a<br />

genomic context in vivo.<br />

Methods: Peripheral blood mononuclear cells (PBMC) from healthy (n = 5) and HCV<br />

infected (n = 5) heterozygous individuals were subjected to haplotype-specific<br />

chromatin immunoprecipitation (HaploChIP) using an antibody directed against<br />

serine 5 phosphorylated RNA polymerase II and an IgG-mock control. With this<br />

specifically captured material (output), non-immunoprecipitated material (input) and<br />

the mock control quantitative restriction fragment length polymorphism (RFLP)<br />

analyses were performed.<br />

Results: The input and mock material was always found to contain about the same<br />

amount of C and T allele fragments, while the output material from healthy individuals<br />

and hepatitis C patients contained about twice the number (T:C ratio = 1.65–2.83) or<br />

even more T allele fragments (T:C ratio = 3.08–5.62), respectively.<br />

Discussion/Conclusion: Our data argue for an effect of the rs2569190/C-159T SNP<br />

on CD14 transcription initiation, i.e. on a preferential binding of serine 5 phosphorylated<br />

RNA polymerase II to T allele fragments, in PBMC from healthy and HCV<br />

infected individuals. These data are quite concordant with in vitro findings by others.<br />

A possible impact on the ongoing CD14 mRNA synthesis, however, is going to be<br />

determined.


Familial occurrence of autoimmune hepatitis<br />

39<br />

Woźniak Małgorzata, Woynarowski Marek, Socha Jerzy<br />

Department of Gastroenterology Hepatology and Immunology, Children’s Health<br />

Memorial Institute, Warsaw, Poland<br />

Introduction: It is known that presence of DR3 or DR4 and several autoimmune<br />

diseases are risk factors for autoimmune hepatitis (AIH). The aim of our study was to<br />

detect the families with more than one child affected with AIH.<br />

Methods: We retrospectively reviewed case records of 189 children with AIH treated<br />

at our institution between 1989 and 2006.<br />

Results: Three pairs of siblings (9–14 years of age) were found. The time window<br />

between the diagnosis varied from 4 months to 9 years. Jaundice was the first<br />

symptom of the disease in one child. The remaining children did not present any<br />

characteristic symptoms of liver disease and the diagnosis was based on typical<br />

laboratory and pathological abnormalities. At diagnosis all children presented with<br />

moderate or severe liver fibrosis. All children received standard prednisone and<br />

azathioprine therapy. The treatment was successfully discontinued in one child. Four<br />

subject continue the therapy and one child died due to liver function decompensation.<br />

Discussion/Conclusion: More than one child with AIH was found in 1.6% of the<br />

families affected with this disease. Our data show that the presentation, clinical<br />

course and treatment outcome of AIH in siblings are the same as in isolated cases of<br />

the disease. The presence of AIH in siblings may justify the implementation of<br />

screening liver function tests in the families of AIH patients.


40<br />

Liver sinusoidal endothelial cells induce anti-inflammatory<br />

CD4 + T cells suppressing murine autoimmune hepatitis<br />

Nils Kruse 1 , Katrin Neumann 1 , Katja Derkow 3 , Anja Kühl 2 , Alf Hamann 4 , and<br />

Katja Klugewitz 1<br />

1<br />

Medizinische Klinik I, Charité Campus Benjamin Franklin, Berlin, Germany<br />

2<br />

Research Center Immunosciences, Charité Campus Benjamin Franklin, Berlin,<br />

Germany<br />

3<br />

Medizinische Klinik m. S. Hepatologie und Gastroenterologie, Charité Campus<br />

Virchow-Klinikum, Berlin, Germany<br />

4<br />

Experimentelle Rheumatologie, Charité Campus Mitte, Berlin, Germany<br />

Introduction: Cellular mechanisms that maintain the intrahepatic immune balance<br />

are crucial for inflammatory liver diseases. For naive CD8 + T cells, liver sinusoidal<br />

endothelial cells (LSEC) have been shown to act as non-professional antigen<br />

presenting cells and thus induce tolerance by inhibition of cytotoxicity. In this study<br />

we investigated consequences of CD4 + T cell priming by LSEC.<br />

Methods: We studied the cytokine expression of LSEC primed CD4 + T cells (TLSEC)<br />

and determined the stability of their phenotype in vivo by adoptive transfer into<br />

congenic mice and immunogenic antigen application. Investigating suppressive<br />

capacities of TLSEC we performed an in vitro suppression assay. The ability of TLSEC to<br />

influence proinflammatory reactions in vivo was analyzed in a model of T cellmediated<br />

autoimmune hepatitis. Hepatic inflammation was monitored by ALT levels<br />

and histologic analyses.<br />

Results: We demonstrated that LSEC induce proliferation of naive CD4 + T cells in<br />

vitro. LSEC primed CD4 + T cells (TLSEC) acquired a CD45RB low phenotype without<br />

effector cytokine expression. This phenotype remained stable in vivo. Interestingly,<br />

TLSEC negative for CD25 and Foxp3, suppressed the proliferation of naive CD4 + T<br />

cells in vitro. They did neither support a DTH reaction nor a hepatic inflammation and<br />

were even able to suppress hepatitis.<br />

Discussion/Conclusion: Priming of naive CD4 + T cells by LSEC leads to a<br />

suppressive phenotype here referred to as TLSEC. Thus liver sinusoidal endothelia<br />

may directly contribute to limiting hepatic inflammation and supporting local tolerance<br />

towards exogenous antigens or endogenous self-antigens.


41<br />

Frequency of autoimmune disorders and therapy in patients<br />

with autoimmune hepatitis<br />

Amelia Genunche-Dumitrescu, P. Mitrut, Daniela Badea, M. Badea<br />

University of Medicine and <strong>Pharma</strong>cy, Clinical Hospital of Emergency, Craiova,<br />

Romania<br />

Introduction: The aim was to assess the frequency of concurrent autoimmune<br />

disorders (CAD) in patients with AIH types 1 and 2 and to evaluate the clinical and<br />

biological response of treatment.<br />

Methods: We studied 88 patients (62 females/26 males, mean ages 43.2 years) with<br />

autoimmune hepatitis (AIH). The diagnosis of AIH was based on international criteria,<br />

including biochemical tests, autoantibodies and liver tissue morphology.<br />

Results: AIH type I was present in 52 patients and type II in 36 patients. CAD was<br />

present at 32 patients with AIH type I and at 24 patients with type II. The incidence of<br />

CAD in AIH type I was: autoimmune thyroiditis (15.62%), rheumatoid arthritis<br />

(18.75%), type I diabetes mellitus (37.5%), Sjogren’s syndrome (9.37%), vitiligo<br />

(12.5%), polyarteritis nodosa (6.26%). At AIH type II was observed: thyroiditis<br />

(20.84%), rheumatoid arthritis (16.67%), type I diabetes mellitus (33.33%), vitiligo<br />

(8.33%), vasculitis (12.5%), colitis ulcerosa (8.33%). Type I diabetes was diagnosed<br />

before AIH type I in 4 cases and in 8 cases it developed after diagnosis was<br />

established. In AIH type II all the cases of diabetes was diagnosed after AIH.<br />

Most patients with AIH type I were treated by only corticoids (40 mg/day with a<br />

reduction of 10 mg every 5 days) and 23 patients follow up associated treatment:<br />

corticoids and azathioprine. After diabetes was diagnosed the steroid doses were<br />

lowered and subsequently stopped.<br />

After 12 months, 42 patients had normal liver biochemistry, 6 had partial response<br />

and 4 no response. In AIH type II, 32 patients had therapeutic success and 4 had<br />

partial response. In AIH type I, after treatment, 3 patients developed mild anemia,<br />

4 developed osteoporosis and 3 severe leukopenia. In AIH type II there were no<br />

serious side-effects.<br />

Discussion/Conclusion: The frequency of autoimmune disorders in AIH was<br />

relatively high. Azathioprine and prednisone associated therapy was effective and<br />

safe.


42<br />

The prediction of the recrudescence risk in autoimmune<br />

hepatitis (AIH)<br />

P. Mitrut, A. Genunche-Dumitrescu, A.O. Mitrut, L. Nita*, D. Badea<br />

University of Medicine and <strong>Pharma</strong>cy, Craiova, Romania<br />

*Emergency District Hospital, Craiova, Romania<br />

Introduction: There are not yet established initial factors to predict the treatment<br />

response in autoimmune hepatitis (AIH). Immunosuppressive therapy with<br />

corticosteroids, often in combination with azathioprine remains the “golden standard”<br />

for induction and maintains remissions in AIH.<br />

Aim: To evaluate these predictive factors for therapeutic response in AIH patients.<br />

Methods: A prospective study was carried out, including 156 patients who were<br />

diagnosed with AIH in our Medical Clinic beginning with January 1999 to January<br />

2009. To apply this diagnosis we have used the standard score system of<br />

International Group of Study AIH. In our protocol study there were mentioned<br />

demographical and clinical data, cholestatic tests, virological and immunological<br />

tests, hepatic biopsy with histological results, ERCP exams or RMN and therapeutic<br />

response.<br />

Results: According International AIH Group criteria they were 72 (46.1%) patients<br />

with “definite” (> 15 points) and 84 (53.9%) with “probable” diagnosis of AIH (10–15<br />

points) pretreatment. We noted a complete response in 92 patients (58.9%) and<br />

recrudescence in 64 patients (41.1%). Recrudescence risk was correlated with<br />

therapy duration, age at diagnosis and immunity profile. A long lasting period of<br />

therapy, at least 4 years, was associated with a low risk of recrudescence. The<br />

presence of anti-LKM1 and ANCA antibodies was correlated with frequent<br />

recrudescence and the presence of ANA and SMA antibodies was correlated with<br />

long lasting remissions. In overlap syndromes (AIH/PBC or AIH/PSC) the response<br />

of immunosuppressive therapy was insufficient in 48% cases.<br />

Conclusions: The absence of response is not infrequent in AIH. Age at diagnosis<br />

< 30 years, a low therapy duration and the presence of anti-LKM1 and ANCA<br />

antibodies were the parameters independently associated with the absence of<br />

complete response to combined therapy.


43<br />

Immune disruptions at the mycotoxin fumonisin-induced<br />

autoimmune hepatitis in young and old mice<br />

Elena A. Martinova<br />

Institute of Nutrition, Russian Academy of Medical Sciences, Moscow, Russian<br />

Federation<br />

Introduction: Previously, we have shown our data on an experimental autoimmune<br />

hepatitis induced in mice fed by mycotoxin Fumonisin B1 (FB1) produced by fungi<br />

Fusarium moniliforme. FB1 contaminates corn all round the world and it is the<br />

carcinogen for some animals and people. Our new results elucidate the immune<br />

disruptions in liver of old and young mice exposed to FB1.<br />

Methods: FB1 (0.5 mM–20 mM) was fed to mice C57Bl/6 (10 weeks old and<br />

18 months old) in the different regimes. Animal were killed by guillotine. Removed<br />

liver was separated and one part was frozen. Samples were stained and<br />

photographed. Liver cells were stained with monoclonal antibodies (mAb) followed by<br />

flow cytometry.<br />

Results: FB1 causes the autoimmune reaction in liver tissue that was confirmed by<br />

staining to Ig+ cells and nuclear antigens. Reaction of old mice versus young animals<br />

to FB1 was much more marked. Liver tissues of old mice were filled by B-cells<br />

surrounding by cytotoxic lymphocytes. There was clear difference in expression of<br />

such proteins as kinase mTOR, receptors GLUT1, GLUT2, GLUT4, IRβ, heat shock<br />

proteins 25, 60, 70, 90 as well as TNF-R2 and markers of activation, on liver<br />

lymphocytes and hepatocytes from old and young mice. The energy potential of liver<br />

cells was disrupted more significant in ageing mice. Activation of lymphocytes in liver<br />

may be regulated by simple sphingolipids (sphingosine, sphinganine, C2-ceramide)<br />

that cross-talks with signalling initiated by FB1.<br />

Discussion/Conclusion: Animal model of autoimmune hepatitis induced by<br />

mycotoxin Fumonisin B1 discloses the mechanism of immune disruptions and shows<br />

the pathways to regulate the activation of immune cells in liver.


44<br />

Emerging importance of autoimmune hepatitis in children in<br />

Bulgaria, an endemic area for viral hepatitis<br />

Krasimira Kalinova PhD, L. Pekova PhD, P. Chakarova PhD, P. Stefanova PhD,<br />

K. Georgiev<br />

University Hospital Stara Zagora, University Hospital Plovdiv, Bulgaria<br />

Introduction: Untreated autoimmune hepatitis (AIH) can develop into liver cirrhosis,<br />

with potentially fatal outcomes. Viral hepatitis in children was endemic in Bulgaria<br />

before universal hepatitis B vaccination, but AIH has rarely been reported in<br />

Bulgarian children. We performed this retrospective study to characterize the clinical<br />

features of AIH in Bulgarian children.<br />

Methods: We enrolled children with AIH, based on the revised scoring system of the<br />

International Autoimmune Hepatitis Group (IAIHG) from 81 children hospitalized with<br />

hepatitis from January 2000 to April 2009. Other etiologies of hepatitis were<br />

excluded.<br />

Results: There were three definite and six probable AIH cases. The incidence of AIH<br />

among children hospitalized with hepatitis was 2.3%. Ten children had other<br />

autoimmune diseases, including systemic lupus erythematosus (SLE) (6), discoid<br />

lupus erythematosus (DLE) (1), and autoimmune polyendocrinopathy syndrome type<br />

1 (1). Another had biliary atresia, and AIH developed after cadaveric liver<br />

transplantation. Antinuclear antibodies ranged from 1:160–1:2560. Peak alanine<br />

aminotransferase (ALT) values were 346 ± 188 U/l (mean ± SD). Jaundice occurred<br />

in four patients. Liver histology in the three definite AIH patients showed chronic<br />

hepatitis with predominantly lymphoplasmacytic infiltrates in portal areas, with<br />

prominent interface activity. Treatment included prednisolone, azathioprine, and/or<br />

cyclosporine. All patients survived. ALT fell to < 60 U/l after treatment. Hepatitis<br />

relapse occurred in one patient.<br />

Discussion/Conclusion: AIH in Bulgarian children is commoner than previously<br />

thought. It is associated with other autoimmune diseases and may occur before,<br />

simultaneously with, or after other autoimmune diseases. Children with liver<br />

transplants are also at risk of AIH.


45<br />

Non-invasive assessment of liver fibrosis by transient elastography<br />

in patients with cholestatic liver diseases (PBC and PSC)<br />

Diana Rill, Adriana Tudora, O. Ciof, E. <strong>Dr</strong>agan, Ioan Sporea<br />

Department of Gastroenterology and Hepatology, University of Medicine and<br />

<strong>Pharma</strong>cy “Victor Babes”, Timisoara, Romania<br />

Introduction: Liver stiffness measurement (LSMs) by means of transient elastograpy<br />

(TE) (FibroScan ® , Echosense, France) has been shown to be correlated to liver<br />

fibrosis in patients with liver diseases of different etiologies, including cholestatic<br />

diseases. Non-invasive assessment of liver stiffness measurement (LSMs) is a<br />

reliable tool in the detection and follow up of significant fibrosis in these patients. We<br />

aimed to study role of TE evaluation in patients with cholestatic liver diseases a well<br />

as correlation to fibrosis stages.<br />

Methods: Out of 3459 TE evaluations carried out in the Department of Gastroenterology<br />

and Hepatology, Timisoara, 54 were performed for cholestatic liver<br />

diseases – 44 for primary biliary cirrhosis (PBC) and 10 for primary sclerosing<br />

cholangitis (PSC). Liver biopsy was performed in 39 (72%) patients. 31 (56.9%)<br />

patients were had the diagnosis of cirrhosis (esophageal varices or other<br />

complications).<br />

Results: Overall, mean LSMs was 17.3 ± 11.9 kPa (ranging from 2.8 to 70.1 kPa). In<br />

patients in pre-cirrhotic stage the LSMs average was 7.6 ± 3.7 kPa, and in cirrhotic<br />

patients 27.3 ± 20.2 kPa (p = 0.00014 ES) LSM correlated to both fibrosis (r = 0.8,<br />

p < 0.0001) and histological (0.70, p < 0.0001) stages, correlations that persisted<br />

when PBC and PSC patients were analyzed separately. Comparing the patients with<br />

significant fibrosis (F ≥ 2) (n = 23) in the pre-cirrhosis stage with those having mild<br />

fibrosis (F < 2), we noticed significant statistical differences (p < 0.0001 ES). Optimal<br />

cut-off value for significant fibrosis (F ≥ 2) was 13.5 kPa and for F < 2 – 6.3 kPa.<br />

Discussion/Conclusion: TE is a simple and reliable non-invasive method for<br />

assessing billiary fibrosis suitable in excluding advanced fibrosis using a threshold of<br />

13.5 kPa.


46<br />

A meta-analysis regarding the effects of ursodeoxycholic acid<br />

in patients with primary sclerosing cholangitis<br />

E.C. Rezi 1 , R. Mihaila 2 , M. Deac 2<br />

1 Second Medical Department, Emergency Clinical County Hospital, Sibiu, Romania<br />

2 Lucian Blaga University, Medical Faculty, Sibiu, Romania<br />

Introduction: Primary sclerosing cholangitis (PSC) is a chronic cholestatic liver<br />

disease characterized by inflammation and progressive bile duct fibrosis. Therapy<br />

with ursodeoxycholic acid (UDCA) has been reported to be associated with<br />

improvements in both clinical symptoms and abnormal serum biochemical liver tests<br />

in patients with primary sclerosing cholangitis.<br />

Methods: In order to evaluate the effects of UDCA on patients with primary<br />

sclerosing cholangitis, we have performed a meta-analysis study. In a Medline<br />

search we identified 12 clinical trials which studied the efficacy and safety of<br />

treatment with UDCA in PSC, with a particular analyse of the biochemical liver tests<br />

(serum alkaline phosphatase [AP], gamma-glutamyl transpeptidase [GGT], alanine<br />

aminotransferases [ALAT], aspartate aminotransferases [ASAT], total bilirubin level<br />

[Bt]) and of the fibrosis score and histological aspects.<br />

Results: A total number of 282 patients were admitted in these studies. A high<br />

percentage (57%) of the patients with PSC also presented inflammatory bowel<br />

diseases. The average age of the patients was 21 ± 16.46 years. The average<br />

treatment period was 1.68 ± 0.79 years. The medium dosage of UDCA which was<br />

used was 16.18 ± 5.29 mg/kg per day. Three studies used high-dosage of UDCA<br />

(20–30 mg/kg per day). The ASAT and ALAT levels decreased significantly in<br />

72.72% (p = 0.001), respectively 81.81% of the patients (p = 0.02). AP decreased at<br />

81% (p = 0.01) and GGT decreased at 72% of the patients (p = 0.0055). The bilirubin<br />

level decreased in 36% of the patients (p = 0.01). Two studies found a significant<br />

improvement in hepatic inflammation, fibrosis and histological stage of the disease.<br />

No side effects were observed during the period of treatment and follow-up.<br />

Discussion/Conclusion: The treatment with standard-dosage (8–15 mg/kg/day) or<br />

high-dosage (20–30 mg/kg/day) of UDCA generally improves liver chemistries and<br />

sometimes liver histology in patients with PSC. Because UDCA treatment improves<br />

but does not cure cholestatic liver diseases, permanent treatment seems to be<br />

necessary. Such prolonged treatment with UDCA may be recommended because,<br />

until now, no side effects have been reported.


47<br />

The long-term budesonide-UDCA combined therapy in patients<br />

with primary biliary cirrhosis<br />

Amelia Genunche-Dumitrescu, P. Mitrut, M. Badea, D. Badea<br />

University of Medicine and <strong>Pharma</strong>cy, Clinical Hospital of Emercency, Craiova,<br />

Romania<br />

Introduction: Aim was to evaluate the efficacy and safety of two years combined<br />

therapy (UDCA plus budesonide) and compared with UDCA alone therapy, in<br />

patients with primary biliary cirrhosis (PBC).<br />

Methods: We studied 33 patients with PBC, at stages I to III. A comparative study<br />

was performed on two groups: A group composed 20 patients which received UDCA<br />

(10–15 mg/kg/day) and B group (13 patients) treated with UDCA and budesonide<br />

(6 mg daily divided in 3 doses). We evaluated liver histology, serum levels of<br />

aminotransferase, Bb, AP at 6, 12 and 24 months.<br />

Results: In A group, clinical symptoms significant improved in 25% of cases after<br />

6 month, in 60.7% after 12 months and 89.28% after 24 months. The mean value of<br />

serum bilirubin concentration was reduced from 6.7 ± 2.5 mg%, at baseline, to<br />

2.8 ± 1.3 mg% at 6 months and to 1.7 ± 0.7 mg% at 12 months. Aminotransferase<br />

values were reduced more quickly comparative with bilirubin and AP levels: with<br />

44.6% at 6 months and 63.2% at 12 months. In B group, aminotransferase values<br />

reduced more slowly, but significant decrease AP activity after one year (p = 0.001).<br />

Inflammatory activity was significantly reduced in the combined therapy (6 cases,<br />

46.15%) and in 4 cases (20%) with monotherapy. Fibrosis decreased in group B in<br />

5 cases, but in A group only in one case. After 24 months, histological stage of<br />

disease improved only in B group (3 cases). In A group, we observed side-effect at<br />

one patient (diarrhea) and in B group 2 patients presented hyperglycemia, 2 mild<br />

hirsutism and 4 osteoporosis. Most of the side effects appeared in patients with stage<br />

III PBC and only in two patients we reduced the budesonide dose.<br />

Discussion/Conclusion: UDCA combined with budesonide improved liver histology<br />

and liver enzymes, where as the effect of UDCA mono-therapy was mainly on liver<br />

function tests.


48<br />

Value of transient elastography (FibroScan ® ) in the evaluation<br />

of patients with hepatocellular carcinoma (HCC)<br />

1. Ciof Octavian Paul, MD, Resident Gastroenterology, Department of<br />

Gastroenterology and Hepatology, Emergency Clinical County Hospital Timisoara;<br />

E-Mail: ciof_o_p@yahoo.com; Phone: +40747025062; Address: Aleea Sanatatii<br />

Nr 18, Bl 1, Sc A, Ap 9, City: Timisoara, Romania<br />

2. Adriana Tudora, MD, Department of Gastroenterology and Hepatology,<br />

Emergency Clinical County Hospital Timisoara; E-Mail:<br />

adrianatudora@yahoo.com: Phone: +40721523258, Address: Str. Tapia Nr 3,<br />

City: Timisoara, Romania<br />

3. Emil <strong>Dr</strong>agan, MD, Resident Internal Medicine, E-Mail: emil_dragan@yahoo.com,<br />

Phone: +40724386118, Address: Str. Paul Iorgovici, Nr 6, City: Timisoara,<br />

Romania<br />

4. Rill Diana, MD, Resident Internal Medicine, E-Mail: diana.rill@yahoo.com, Phone:<br />

+40721410775, Address: Str Electronicii Nr 26, City: Timisoara, Romania<br />

5. Martin Mihaela, MD, E-Mail: miha_ella@yahoo.com, Phone: +40745518371,<br />

Address: Str. Magura Nr 6, City: Timisoara, Romania<br />

Introduction: Aim: Main risk factors in the occurrence of hepatocellular carcinoma<br />

(HCC) are represented by liver cirrhosis (LC) of different ethiologies (HBV, HCV or<br />

alcoholic, hemochromatosis, or unknown ethiologies). Thus, the follow up of risk<br />

group patients for developing HCC is mandatory. Transient elastography (TE)<br />

represents a novel, user-friendly, non-invasive method, appropriate to use in<br />

outpatients method for the evaluation of liver fibrosis in patients with chronic liver<br />

diseases of different etiologies. We aimed to establish the value of TE in the<br />

evaluation of patients with LC and HCC.<br />

Methods: Out of 1668 patients with LC, mean age 58.42 ± 11.57 years (January<br />

2000–December 2006). 104 patient with chronic liver diseases (chronic hepatitis or<br />

LC) and HCC were evaluated by TE.<br />

Results: Liver stiffness measurements (LSMs) in patients with chronic liver diseases<br />

ranged between 4.10 and 75 kPa (mean LSM value for patients with chronic hepatitis<br />

was 9.76 ±4.77 kPa and for patients with LC 34.97 ± 20.25 kPa). Mean LSMs values<br />

for patients with LC and HCC was 36.65 ± 20.49 kPa (AUROC 0.94 ± 0.02, ranging<br />

between 0.89–0.98). Considering a cut-off value of LSMs for LC of 14.5 kPa, our<br />

sensibility was 90% and specificity 85%, with a PPV 92% and NPV 78%. No<br />

statistical differences were noticed between LSMs in patients with LC with or without<br />

HCC (p = 0.1123-NS).<br />

Discussion/Conclusion: Despite some other studies in literature, in our study we<br />

could not evidence no significant predictive value of LSMs by TE in which concerns<br />

the differentiation between LC and HCC.


49<br />

Constitutive active gp130 in the liver and its connection to<br />

chronic inflammation and carcinoma development<br />

Antje Schütt, Jürgen Scheller, Stefan Rose-John<br />

Department of Biochemistry, Christian-Albrechts-University Kiel, Kiel, Germany<br />

Corresponding author: aschuett@biochemie.uni-kiel.de<br />

Introduction: Glycoprotein130 (gp130) is the common signaling receptor subunit for<br />

the cytokines of the IL-6 family. Activation of gp130 leads to activation of the STAT3<br />

and MAP/AKT signaling pathway. By replacing the extracellular part of gp130 with<br />

the leucin-zipper of the c-jun protein, we have recently constructed a dimerized<br />

version of gp130 (L-gp130), which leads to constitutive activation of gp130 signaling.<br />

We have shown earlier in mice that massive gp130 lead to adenoma formation in the<br />

liver. These data were recently confirmed in patients with liver adenomas where<br />

somatic activating mutations in gp130 have been detected.<br />

Methods: We generated a bidirectional expression plasmid in which L-gp130 and<br />

luciferase is placed under the transcriptional control of a tetracycline responsive<br />

promotor element (TRE). In the presence of a tetracycline transactivator protein<br />

(tTA), L-gp130 and luciferase expression can be regulated by the tetracycline<br />

analogue doxycycline. Based on this system we plan to analyze chronic liver<br />

inflammation and carcinoma development in the liver.<br />

Results: L-gp130 and luciferase were shown to be tightly regulated by doxycycline in<br />

human hepatoma cells (HepG2). The expression L-gp130/luciferase expression<br />

plasmid was used to generate transgenic mice, in which expression of L-gp130 and<br />

luciferase can be regulated in tissues, which express the tTA protein, in a cellautonomous<br />

manner.<br />

Discussion/Conclusion: First results on the cellular expression of L-gp130 and on<br />

the generation of liver specific transgenic mice will be presented and the advantages<br />

of an animal model of cell-autonomous gp130 activation will be discussed.


50<br />

Increased intrahepatic infiltration with CD4+CD25+FOXP3+<br />

regulatory T cells correlates with increased tissue levels of<br />

TGF-beta and IL-10 in patients with primary or metastatic liver<br />

cancers<br />

Manolova I 1 , Gulubova M 2 , Kyurkchiev D 3 , Ananiev J 2 , Altunkova I 3 , Julianov A 4<br />

1 Laboratory of Clinical Immunology, University Hospital, 2 Department of General and<br />

Clinical Pathology,<br />

4 Department of General Surgery, Medical Faculty, Trakia<br />

University, Armeiska str. 11, Stara Zagora 6000, Bulgaria; 3 Laboratory of Clinical<br />

Immunology, University Hospital “St Ivan Rilski”, Sofia, Bulgaria<br />

E-Mail: imanolova@mf.uni-sz.bg<br />

Introduction: CD4+CD25+FOXP3+ regulatory T cells (Treg) have been<br />

characterized as a critical population of immunosuppressive cells that maintain<br />

peripheral tolerance to self antigen and also inhibit anti-tumor immune response. The<br />

suppression mechanism also requires interleukin 10 (IL-10) and transforming growth<br />

factor-beta 1 (TGF-beta1) in the local environment. In the present study in patients<br />

with primary or metastatic livers cancers, we examine the relationship between<br />

intrahepatic Treg and the levels of immunosuppressive cytokines IL-10 and<br />

TGF-beta1.<br />

Methods: 17 patients (11 males and 6 females) undergone liver resection for a<br />

primary or metastatic tumor and 6 patients (2 males and 4 females) undergone<br />

surgery for primary gastrointestinal tumors were included in the study. Liver samples<br />

were obtained after surgical resection of metastatic or liver tissue for curative or<br />

diagnostic purpose. Intrahepatic Treg were identified and characterized as<br />

CD4+CD25+FOXP3+ using flow cytometry. Tissue levels of TGF-beta1 and IL-10<br />

were measured by enzyme-linked immunosorbent assay in the supernatants of tumor<br />

homogenate and in normal liver tissue.<br />

Results: Flow cytometry revealed increased proportions of Treg in metastatic liver<br />

compared to normal liver (1.35 ± 1.17 vs. 14.39 ± 10.96%, p = 0.001). TGF-beta1<br />

was detected in supernatants from homogenized normal liver tissue (mean<br />

22.5 ng/100 mg wet tissue, n = 6) and from tumor tissue (45.04 ng/100 mg wet<br />

tissue, n = 17, p = 0.107). We found a trend towards increased IL-10 levels in<br />

metastatic tissue (mean = 115.4 pg/100 mg wet tissue, n = 17) in comparison with<br />

normal liver tissue (mean = 34.6 pg/100 mg wet tissue, n = 6, p = 0.093). There was<br />

a significant, positive correlation between the proportions of CD4+CD25+FOXP3+<br />

Treg and the levels of IL-10 (r = 0.49, p = 0.028) and TGF-beta1 (r = 0.602,<br />

p = 0.005).<br />

Conclusion: Collectively, these data demonstrate the immunosuppressive environment<br />

in the liver bearing primary or metastatic tumors.


51<br />

TGF-beta occurrence and macrophage infiltration in liver<br />

metastasis<br />

Ananiev J. 1 , Gulubova M. 1 , Manolova M. 2 , Matev A. 3<br />

1 Department of General and Clinical Pathology, 2 Laboratory of Clinical Immunology,<br />

University Hospital,<br />

3 Department of General Surgery, Medical Faculty, Trakia<br />

University, Armeiska str. 11, Stara Zagora 6000, Bulgaria; 3 Laboratory of Clinical<br />

Immunology, University Hospital “St Ivan Rilski”, Sofia, Bulgaria<br />

E-Mail: operation@abv.bg<br />

Introduction: TGF-beta1 signaling pathway is implicated in the regulation of cancer.<br />

TGF-beta1 has immune suppressive function. It regulates tumor environment<br />

together with some other cytokines and so modulates tumor cell viability and spread.<br />

The aim of present study was to evaluate the occurrence of TGF-beta1 pathway<br />

components in liver metastasis and to assess its relation with tumor associated<br />

macrophages (TAMs) infiltration.<br />

Methods: 48 patients (30 males and 18 females) with liver metastasis from colorectal<br />

and gastric cancer were investigated. The immunohistochemistry was done with anti-<br />

TGF-beta1, anti-SMAD4, anti-SMAD7, anti-TGF-beta RII, and anti-CD68 antibodies.<br />

Results: Elevated TGF-beta1 expression was detected in tumor cytoplasm in<br />

47.82% of metastasis. SMAD4 was observed mainly in tumor cell cytoplasm in<br />

91.3% of the samples. SMAD7 was observed also in tumor cell cytoplasm in 65.22%<br />

of metastasis. TGF-beta RII was found on tumor cell membranes in 65.22% of<br />

tumors.<br />

Statistically significant correlation was observed between TGF-beta1 expression and<br />

decreased CD68 numbers in tumor stroma (χ 2 = 7.0, p = 0.008) and in tumor border<br />

(χ 2 = 4.33, p = 0.037).<br />

There was a tendency for correlation between low numbers of CD68-positive TAMs<br />

with high expression of TGF-beta RII (χ 2 = 3.582, p = 0.058) and with high<br />

expression of SMAD7 in tumor cells (χ 2 = 3.549, p = 0.06).<br />

Discussion/Conclusion: In liver metastasis TGF-beta1 is located mainly in tumor<br />

cell cytoplasm. Its receptor TGF-beta RII marked with variable intensity the tumor cell<br />

membranes and SMAD4 and SMAD7 had mainly cytoplasmic expression. Elevated<br />

TGF-beta1, TGF-beta RII and SMAD7 expression correlated with decreased TAMs<br />

infiltration.


52<br />

TGF-beta inhibits dendritic cell infiltration in liver metastasis<br />

Gulubova M. 1 , Manolova I. 2 , Ananiev J. 1 , Yovchev Y. 3<br />

1 2<br />

Department of General and Clinical Pathology, Laboratory of Clinical Immunology,<br />

3<br />

University Hospital, Department of General Surgery, Medical Faculty, Trakia<br />

University, Armeiska str. 11, Stara Zagora 6000, Bulgaria<br />

E-Mail: mgulubova@hotmail.com<br />

Introduction: TGF-beta can utilize various programs to promote cancer metastasis<br />

through its effects on the tumor microenvironment, enhanced invasive properties and<br />

inhibition of immune cell function. The aim of the study is to assess TGF-beta1, its<br />

receptor TGF-beta RII and the signaling proteins Smad4 and Smad7 expression in<br />

liver metastasis and their relation to dendritic cell infiltration.<br />

Methods: We investigated 46 samples from liver metastasis from primary colorectal<br />

and gastric cancer. The patients were 28 males and 18 females with median age of<br />

64 years, range from 38 to 86 years. Paraffin sections (7 µm thick) were processed<br />

for immunohistochemistry with anti-TGF-beta1, anti-TGF-beta RII, anti-Smad4, anti-<br />

Smad7, anti-CD1a and anti-CD83 antibodies.<br />

Results: TGF-beta1 expression in tumor cytoplasm correlated with the low infiltration<br />

with CD1a (χ 2 = 8.27, p = 0.004) and CD83 (χ 2 = 28.83, p = 0.000) in tumor border<br />

and in tumor stroma (for CD1a χ 2 = 3.18, p = 0.074 and for CD83 χ 2 = 17.84,<br />

p = 0.000). TGF-beta1 and Smad4 were observed also in some stellate cells at<br />

metastasis border.<br />

The low CD1a and CD83 infiltration in tumor border correlated with higher SMAD4<br />

expression in tumor cells (χ 2 = 3.062, p = 0.08, χ 2 = 0.23, p = 0.045 respectively).<br />

The higher membranous expression of TGF-beta RII in tumor cells correlated with<br />

low numbers of CD1a and CD83 cells in tumor border (χ 2 = 7.659, p = 0.006 and<br />

χ 2 = 11.016, p = 0.001 respectively).<br />

Tissue levels of TGF-beta1 were measured by enzyme-linked immunosorbent assay<br />

in the supernatant of metastasis homogenate and in normal liver tissue. Normal<br />

hepatic TGF-beta1 was detected at 22.5 ng/100 mg wet tissue, whereas in<br />

metastatic tissue TGF-beta1 was detected at 45.04 ng/100 mg wet tissue.<br />

Discussion/Conclusion: Our data show that TGF-beta1, its receptor and the<br />

signaling Smad proteins in tumor metastasis correlated with inhibition of immune cell<br />

function.


Focal nodular hyperplasia of the liver within our experience<br />

53<br />

Halina Cichoż-Lach 1 , Beata Prozorow-Król 1 , Jarosław Swatek 3 , Krzysztof Celiński 1 ,<br />

Maria Słomka 1 , Leszek Buk 2 , Elżbieta Korobowicz 3 , Emilia Lis 1<br />

1 Department of Gastroenterology, Medical University of Lublin, Poland<br />

2 1st Department of Radiology, Medical University of Lublin, Poland<br />

3 Department of Clinical Pathomorphology, Medical University of Lublin, Poland<br />

Introduction: Focal nodular hyperplasia (FNH) is a tumor-like lesion of the liver. The<br />

purpose of this study was to find out if there were any hints from the medical histories<br />

or additional investigations which would suggest FNH later confirmed by liver biopsy.<br />

Methods: The clinical data of 28 patients with FNH hospitalized in the Department of<br />

Gastroenterology, MU of Lublin in the period 1st January, 2002–31st October, 2007<br />

were analyzed retrospectively.<br />

Results: Liver biopsy diagnosed FNH in 28 patients (92.86% women) of total 667<br />

hospitalized in the Gastroenterology Department, Medical University of Lublin for<br />

focal lesions in the liver. In the group of 26 women 16 patients were on oral<br />

contraceptives, 5 were taking hypotensives, men did not receive any medication. All<br />

28 showed normal values of ALT, AST and AFP while each having a single focal<br />

lesion 24–121 mm in diameter found in USG. Their medical histories revealed that<br />

the FNH was asymptomatic and incidentally detected in USG. A central scar had<br />

been detected in 9 of 28 patients on USG and in 14 of 23 patients on CT.<br />

Discussion/Conclusion: In our study FNH was relatively rare pathology and<br />

constituted 4.20% of all cases with focal liver lesion. FNH prevailed in women<br />

(particularly who were oral contraceptives users), did not differentiate in laboratory<br />

results. It manifested as a single focus being incidentally detected by imaging<br />

techniques with a central scar visualized in 32–60% of patients depending on the<br />

techniques applied. Scar absence accounts for the lack of features specific of FNH<br />

and then biopsy is the best option to diagnose the lesion.


54<br />

Liver sinusoidal endothelial cells induce TGF-β dependent<br />

conversion of CD4+Foxp3+ regulatory T cells from conventional<br />

CD4+CD25- T cells<br />

Antonella Carambia, Samuel Huber, Stefan Lüth, Dorothee Schwinge,<br />

Christian Frenzel, Christoph Schramm, Ansgar W. Lohse, Johannes Herkel<br />

Department of Medicine I, University Medical Centre Hamburg-Eppendorf, Germany<br />

Introduction: Recently, it has been demonstrated that ectopic expression of an<br />

autoantigen in the liver can induce antigen-specific CD4+CD25+Foxp3+ regulatory T<br />

cell (Treg) differentiation and protection from autoimmune disease (Lüth et al. 2008.<br />

J Clin Invest. 2008; 118: 3403–10). However, the underlying cellular and molecular<br />

mechanisms remain undefined.<br />

Methods: To test the capacity of liver sinusoidal endothelial cells (LSEC) to induce<br />

Treg from CD4+CD25- conventional T cells, we cultured LSEC (1 x 10 5 ) or splenic<br />

dendritic cells (DC, 5 x 10 4 ) of C57BL/6 mice together with CD4+CD25- T cells<br />

(5 x 10 5 ) and solute antibody to CD3 (1 µg/ml) in the presence or absence of<br />

exogenous TGF-β (0–2 ng/ml).<br />

Results: Upon stimulation by DC, conversion of CD4+CD25- T cells to Treg was<br />

negligible (< 1%); T cell stimulation by LSEC, in contrast, induced significant levels of<br />

Foxp3 expressing Treg (up to 8%). Addition of exogenous TGF-β led to a dosedependent<br />

increase of Treg upon T cell stimulation by LSEC (up to 35%), but<br />

considerably less by DC (up to 7%). These findings were confirmed when T cells<br />

were primed antigen-specifically in the presence of myelin basic protein (MBP;<br />

5 ng/ml) and exogenous TGF-β (2 ng/ml). Of note, cultures of LSEC or DC together<br />

with CD4+CD25- T cells from transgenic mice with impaired TGF-β signaling showed<br />

greatly diminished conversion rates (LSEC-stimulated < 8,5%, DC-stimulated<br />

< 0,6%).<br />

Discussion/Conclusion: In conclusion, we identified LSEC as potent inducers of<br />

TGF-β dependent Treg conversion, indicating a major role of LSEC in the intrahepatic<br />

generation of Treg.


55<br />

Alcohol-metabolizing enzyme gene polymorphisms in alcohol<br />

liver cirrhosis among Polish men<br />

Halina Cichoż-Lach, Emilia Lis, Krzysztof Celiński, Maria Słomka<br />

Department of Gastroenterology Medical University of Lublin, Poland<br />

Introduction: Alcohol liver disease occurs in 77% individuals who abuse alcohol<br />

consumption. Genetic polymorphism of enzymes involved in alcohol metabolism<br />

plays a relevant role in etiopathogenesis of alcohol liver disease. The aim of the<br />

study was to find in the Polish population the ADH1B, ADH1C and ALDH2<br />

genotypes, which are likely to be responsible for higher susceptibility to alcohol liver<br />

cirrhosis.<br />

Methods: The ADH1B, ADH1C and ALDH2 genotype and alleles frequencies were<br />

examined in 202 men: 77 with alcoholic liver cirrhosis, 64 alcoholics without damage<br />

to gastrointestinal organs and 61 nondrinkers (a control group). Genotyping of the<br />

ADH, ALDH2 was performed using PCR-RFLP method on white cell DNA.<br />

Results: The genotype ADH1C*1/*1 and allele ADH1C*1, ADH1B*1 were found to<br />

be significantly more frequent in alcohol abusers compared to non-drinkers.<br />

Frequency of ADH1C*1 allele in alcohol liver cirrhosis group was 62.8%, and<br />

ADH1C*1/*1 genotype was observed in 45.4% and was significantly higher than in<br />

the controls. The differences between of the group of patients who abuse alcohol<br />

were not statistically significant. In the group of nondrinkers ADH1B*2 and ADH1C*2<br />

alleles were more frequent in comparison to the alcohol liver cirrhosis patients and<br />

alcohol addicts. All examined patients were ALDH2*1/*1 homozygotic.<br />

Discussion/Conclusion: In the Polish population examined ADH1C*1, ADH1B*1<br />

allele and ADH1C*1/*1 genotype favor developing alcoholism and alcohol liver<br />

cirrhosis. However ADH1B*2 and ADH1C*2 allele are likely to protect against them.<br />

Genetic polymorphism of ALDH2 shows no correlation with alcohol addiction or<br />

alcohol cirrhosis, Polish is monomorphic ALDH2*1.


56<br />

Conversion of naive T cells into regulatory T cells after<br />

tolerance induction in the murine liver<br />

Benjamin Claaß, Annette Erhardt, Gisa Tiegs<br />

Experimental Immunology and Hepatology, Center of Internal Medicine, University<br />

Medical Center Hamburg-Eppendorf, 20246 Hamburg, Germany<br />

Introduction: Recently, we have demonstrated that sublethal doses of concanavalin<br />

A (ConA) induce tolerance towards ConA hepatitis in mice within one week.<br />

Tolerance is mediated by IL-10-producing regulatory T cells (Tregs). In vivo and in<br />

vitro data suggested that highly suppressive ConA-primed Foxp3+ Tregs might<br />

convert naive CD4+ T cells into Foxp3+ Tcells with immunosuppressive function.<br />

Hence, this study was intended to identify the relevance of “infectious tolerance” as<br />

one mechanism of ConA tolerance.<br />

Methods: Splenic CD4+ T cells from DEREG (“depletion of regulatory T cells”) mice<br />

were co-cultivated with splenic Tregs from saline- or ConA-pretreated C57BL/6 wt<br />

mice. Moreover, liver non-parenchymal cells (NPCs) or whole splenocytes containing<br />

antigen-presenting cells (APCs) were isolated from DEREG mice. Cells were<br />

stimulated with anti-CD3 mAb for 72hrs. Foxp3 expression correlating with GFP<br />

expression and expression of surface markers with suppressive/regulatory function<br />

were quantified by FACS analysis.<br />

Results: Neither Tregs from ConA- nor from saline-treated mice could induce Foxp3<br />

expression in the co-culture system. However, a transient FoxP3/GFP expression<br />

was detectable in cultures of whole splenocytes and liver NPCs, respectively,<br />

suggesting the requirement of APCs in Treg conversion. PD1, CCR5, and CXCR3<br />

expression were upregulated on Tregs from ConA-tolerant mice, whereas CD45RB<br />

and CD62L were downregulated.<br />

Discussion/Conclusion: In conclusion, different migration patterns and surface<br />

expression profiles seem to characterize the highly suppressive ConA-primed Tregs.<br />

Moreover, conversion of naive responder cells into Foxp3+ T cells by Tregs alone<br />

might be of minor importance for controlling ConA-induced inflammation; however,<br />

differentiation of IL-10 secreting FoxP3-negative Tr1 cells might be induced by Tregs<br />

in an infectious manner. Further studies will identify the mechanism (expansion/proliferation<br />

or conversion) of the transiently increased FoxP3 expression upon<br />

ConA challenge in vivo and the potential of liver-specific APCs, e.g. LSECs and<br />

hepatocytes to convert naive intrahepatic T cells into Tregs.


57<br />

Impact of IL-6-transsignaling on liver damage and regeneration<br />

after CCl4<br />

Jessica Gewiese, Claudia <strong>Dr</strong>ucker, Jürgen Scheller, and Stefan Rose-John<br />

Institute of Biochemistry, Medical Faculty, Christian-Albrechts-University Kiel,<br />

Germany<br />

Corresponding author: jgewiese@biochem.uni-kiel.de<br />

Introduction: The role of interleukin-6 (IL-6) in liver regeneration has been<br />

documented. There are two different pathways described for IL-6. In addition to the<br />

classical IL-6 pathway in which IL-6 binds to a membrane bound receptor (IL-6R),<br />

there also exists IL-6 transsignaling via a soluble IL-6R (sIL-6R). sIL-6R binds to IL-6<br />

and activates cells which do not express membrane bound IL-6R.<br />

Methods: To investigate IL-6 transsignaling during of CCl4 liver damage, we blocked<br />

IL-6 transsignaling with soluble gp130Fc protein (sgp130Fc).<br />

Results: We show that blockade of IL-6 transsignaling enhanced CCl4 liver damage.<br />

sgp130Fc pre-treated mice exhibited higher levels of alanine aminotransferase (ALT)<br />

and aspartate aminotransferase (AST). Higher uric acid and potassium concentations<br />

in the sera of these mice were indications for liver cell damage. These differences<br />

neither depended on the bioactivation of CCl4 via Cyp450 2E1, nor on the infiltration<br />

of neutrophils. Liver regeneration is induced by IL-6 transsignaling. Quantification of<br />

regeneration via BrdU shows that regeneration rates are slower when IL-6 transsignaling<br />

is blocked.<br />

Discussion/Conclusion: We focused on IL-6 transsignaling and its impact on liver<br />

damage and regeneration. We showed that IL-6 transsignaling is important since<br />

blockade enhanced liver damage. In view of the fact that a neutralizing IL-6R<br />

antibody that globally blocks IL-6 signaling is already used in the clinic, we will further<br />

investigate whether specific blockade of transsignaling in the liver is superior to<br />

global IL-6 blockade.


58<br />

CXCL10 plays a key role in the development of hepatic fibrosis<br />

Edith Hintermann 1 , Monika Bayer 1 , Andrew Luster 2 , and Urs Christen 1<br />

1 <strong>Pharma</strong>zentrum Frankfurt/ZAFES, Klinikum der Johann Wolfgang Goethe-<br />

Universität, Frankfurt am Main, Germany<br />

2 Massachusetts General Hospital, Harvard Medical School, Charlestown, MA, USA<br />

Introduction: Hepatic fibrosis in chronic liver inflammation as occurring during<br />

chronic hepatitis C infection or autoimmune hepatitis is one of the major concerns<br />

associated with chronic liver disease. Hepatic stellate cells (HSC) are the main<br />

drivers of hepatic fibrosis generating extracellular matrix proteins, such as collagen I.<br />

Here we asked how HSC and infiltrating lymphocytes are activated and recruited to<br />

the site of chronic inflammation.<br />

Methods: We treated wildtype C57Bl/6 and CXCL10 KO mice with carbon tetrachloride<br />

(CCl4) for up to 8 weeks, isolated RNA from whole liver homogenates and<br />

purified HSC. Hepatic fibrosis was assessed by Sirius Red staining. Analysis of HSC<br />

activation was by staining of liver section and isolated HSCs for alpha-smooth<br />

muscle actin (aSMA).<br />

Results: We found a significant elevation in the expression of the chemokine<br />

CXCL10 RNA in HSC. Thus, we used CXCL10 KO mice to investigate the role of<br />

CXCL10 in hepatic fibrosis. Interestingly, CXCL10 KO mice had much lower numbers<br />

of infiltrating lymphocytes in the liver and significantly reduced hepatic fibrosis.<br />

Similarly, blockade of CXCL10 with a neutralizing antibody resulted in diminished<br />

fibrosis and hepatitis. Mechanistically we found that HSC are activated by CCl4<br />

treatment as demonstrated by staining for aSMA. Activated HSC isolated from CCl4treated<br />

mice express CXCR3 and migrated towards CXCL10 in a Transwell migration<br />

assay. Further, activated HSC responded to CXCL10 stimulation by actin<br />

polymerization.<br />

Discussion/Conclusion: In summary, CXCL10 plays an important role in the<br />

development of hepatic fibrosis influencing HSC migration, lymphocyte attraction and<br />

inflammation.


59<br />

Epitope mapping of human cytochrome P450 2D6 (CYP2D6) in<br />

patients with autoimmune hepatitis and in the CYP2D6 mouse<br />

model<br />

Edith Hintermann 1 , Martin Holdener 1 , Monika Bayer 1 , Michael Manns 2 , and<br />

Urs Christen 1<br />

1 <strong>Pharma</strong>zentrum Frankfurt/ZAFES, Klinikum der Johann Wolfgang Goethe-<br />

Universität, Frankfurt am Main, Germany<br />

2 Medizinische Hochschule Hannover, Hannover, Germany<br />

Introduction: Cytochrome P450 2D6 (CYP2D6) is the major autoantigen recognized<br />

by LKM-1 antibodies of autoimmune hepatitis (AIH) type 2 patients. We recently<br />

developed a novel mouse model for AIH in which mice are infected with an<br />

adenovirus expressing the human CYP2D6 (Ad-2D6). Mice develop chronic hepatitis<br />

with massive lymphocyte infiltration and subcapsular/periportal fibrosis. Just like AIH<br />

patients, Ad-2D6-infected mice generate high-titer anti-CYP2D6 antibodies.<br />

Methods: Epitope mapping was performed using SPOTs membranes containing<br />

162 peptides (15 aa length, 3 aa offset) covering the entire human CYP2D6<br />

molecule.<br />

Results: Here we report the CYP2D6-epitopes recognized by sera of AIH patients<br />

and Ad-CYP2D6-infected mice and CYP2D6 epitope spreading over time. We found<br />

considerable variation in epitope recognition in between individual AIH patients.<br />

However, the immunodominant epitope (aa262-270: WDPAQPPRD) was found in<br />

most patients as well as in all Ad-2D6 infected mice. This immunodomiant epitope<br />

was already present at time of AIH-diagnosis and was recognized throughout the<br />

whole timeframe analyzed. There is limited intramolecular spreading over time in<br />

both AIH patients and Ad-2D6-infected mice. However, we have identified some<br />

novel epitopes that are only recognized at specific times.<br />

Discussion/Conclusion: Such extraordinary epitopes might provide a basis for<br />

finding possible environmental initiation factors, such as pathogens with structural<br />

similarities conferring molecular mimicry to human CYP2D6. Indeed we found<br />

sequence homology of such CYP2D6 epitopes to human pathogens such as<br />

Hepatitis C virus, Herpes simplex virus, and Legionella pneumophila.


60<br />

The role of the hepatic asialoglycoprotein receptor in inflammatory-induced<br />

liver injury and accumulation of intrahepatic<br />

lymphocytes<br />

Benita L. McVicker, Dean J. Tuma, Geoffrey M. Thiele, Carol A. Casey, and<br />

Natalia A. Osna<br />

University of Nebraska Medical Center and Department of Veterans Affairs, Omaha,<br />

NE, USA<br />

Introduction: We have shown that alcohol impairs hepatic protein trafficking, and<br />

that receptor-mediated endocytosis via the abundant hepatocyte asialoglycoprotein<br />

(ASGP) receptor is especially affected. To further clarify the role of ASGP receptors<br />

in liver disease, we hypothesize that the receptor is involved in T lymphocyte binding<br />

to hepatocytes resulting in the triggering of T-cell clearance mechanisms.<br />

Methods: To examine our hypothesis, we injected wild-type (WT) and ASGP<br />

receptor deficient (RD) mice with T cell-activating anti-CD3 monoclonal antibodies<br />

(which increases T-lymphocyte activation and homing to the liver) followed by<br />

assessment of lymphocyte accumulation and liver injury.<br />

Results: Our results indicate that following anti-CD3 treatment in WT mice, we<br />

observed a two-fold elevation of AST and TNF-alpha expression in the liver within<br />

24 hours which was maintained up to 3 days post injection. Caspase 3 activity was<br />

only slightly increased after 24 hours, but was significantly higher (65% increase,<br />

p < 0.05) two days after anti-CD3 injection in the WT mice. In RD mice, all of the<br />

parameters of liver injury (AST, TNF and caspase activation) were enhanced at least<br />

two-fold compared to WT animals. Additionally, the number of CD8 + /CD45R +<br />

intrahepatic T cells (the proposed cell type involved in hepatocyte-lymphocyte<br />

interactions) were found to be enhanced in the RD animals. Specifically, FACS<br />

analysis demonstrated that following anti-CD3 administration, the percentage of<br />

CD8 + /CD45R + lymphocytes in the liver rose from 3% to 34% in WT animals to 49% in<br />

RD mice.<br />

Discussion/Conclusion: These data support our hypothesis that functional ASGP<br />

receptors play an important role in the regulation of CD8 + T cells in the liver.<br />

Therefore, altered ASGP receptor function in an injured liver can contribute to liver<br />

lymphocyte accumulation and thus may be related to the damaging effects<br />

associated with T cells in the development and/or progression of alcoholic liver<br />

disease.


61<br />

Common variable immunodeficiency-associated granulomatous<br />

disease and nodular regenerative hyperplasia of the liver<br />

H. Varnholt, T. Rubinas<br />

Department of Pathology, University of North Carolina, Chapel Hill, USA<br />

Introduction: Common variable immunodeficiency (CVID) is a primary immunodeficiency<br />

disorder characterized by reduced serum immunoglobulins and heterogeneous<br />

clinical features. Histopathologic findings of liver biopsies obtained for<br />

elevated liver enzymes or hepatosplenomegaly remain ill defined.<br />

Methods: A database search at the University of North Carolina from 1996–2009<br />

yielded two patients with known CVID who had undergone liver biopsies for elevated<br />

liver enzymes.<br />

Results: Patient A is a 16 year-old Caucasian male who was first diagnosed with<br />

CVID one year ago when he had fatigue and underwent splenectomy for splenomegaly<br />

(940 g). Current labs: AST 83 U/l, ALT 101 U/l, Alk Phos 330 U/l, IgA 5 mg/dl,<br />

IgG normal. Liver histology shows numerous nonnecrotizing granulomata with<br />

negative AFB stains and mild lobular lymphocytic inflammation without plasma cells<br />

or fibrosis. Patient B is a 15 month-old African-American female infant with recurrent<br />

respiratory infections, agammaglobulinemia and slightly elevated ALT with normal<br />

other liver function tests. Liver biopsies demonstrate small hyperplastic nodules<br />

without fibrous septae consistent with nodular regenerative hyperplasia (NRH).<br />

Discussion/Conclusions: These cases highlight two common hepatic findings in<br />

CVID: Granulomatous disease and NRH. Although sarcoidosis has been observed<br />

as a cause of multi-organ granulomatous disease, we postulate that in this case it<br />

may have been due to T-cell deficiencies or imbalanced production of cytokines<br />

leading to abnormal sequestration of antigens. NRH, seen in > 80% of cases, may be<br />

caused by variations in blood flow to different areas of hepatic parenchyma or<br />

circulating immune complexes or light chain deposits in the sinusoidal walls.


62<br />

Aggravation of liver damage and loss of immunological<br />

tolerance in CXCR3-deficient mice in the murine model of<br />

concanavalin A-induced liver injury<br />

Claudia Wegscheid 1 , Annette Erhardt 1 , Ulf Panzer 2 , Gisa Tiegs 1<br />

1 Experimental Immunology and Hepatology, Center of Internal Medicine, University<br />

Medical Center Hamburg-Eppendorf, 20246 Hamburg, Germany<br />

2 Medical Clinic III, Center of Internal Medicine, University Medical Center Hamburg-<br />

Eppendorf, 20246 Hamburg, Germany<br />

Introduction: Injection of the plant lectin concanavalin A (ConA) induces an acute<br />

immune-mediated liver injury in mice. Animals developed tolerance against ConA<br />

rechallenge within one week. Tolerance is mediated by IL-10 producing<br />

CD4+CD25+FoxP3+ regulatory T cells (Tregs). The chemokine receptor CXCR3,<br />

which is expressed by both polarized Th1 cells and Tregs, and its ligands CXCL9-11<br />

have been shown to be involved in several autoimmune diseases. Hence, ConA<br />

hepatitis and tolerance were investigated in CXCR3 KO mice. The results might be<br />

useful to evaluate novel therapeutic approaches in human liver disorders.<br />

Methods: C57/BL6 wildtype and CXCR3-deficient mice were injected with ConA or<br />

saline, respectively, and restimulated with ConA after 8 days. Plasma transaminase<br />

activities were measured and expression of IL-2, IL-10, IL-6, IL-17, TNF-alpha and<br />

IFN-gamma were determined by ELISA and RT-PCR 8 hours after ConA<br />

rechallenge. FACS analyses were performed to characterize the intrahepatic cell<br />

composition and migration of lymphocytes upon ConA challenge.<br />

Results: The expression of CXCL9-11 was significantly upregulated in wt mice upon<br />

ConA challenge correlating with increased IFN-gamma levels. Interestingly, CXCR3deficient<br />

mice exhibited aggravated liver damage and failed to develop tolerance<br />

against ConA indicated by elevated plasma transaminase activity and a proinflammatory<br />

cytokine profile. The frequency of NKT cells was increased in the liver<br />

of CXCR3-deficient mice, whereas the number of Tregs was decreased.<br />

Discussion/Conclusion: Aggravation of liver damage and loss of tolerance in<br />

CXCR3-deficient mice suggests a protective role of CXCR3 during immune-mediated<br />

liver diseases. In contrast CXCR3-deficient mice developed developed less severe<br />

nephritis in a Th1-mediated nephritis model and administration of a blocking anti-<br />

CXCR3 antibody induced prolonged survival of cardiac and islet allografts. Further<br />

studies are intended to identify the role of CXCR3 for migration, conversion and<br />

immunosuppressive capacity of Tregs in immune-mediated liver disorders, since<br />

recruitment of CXCR3-positive Tregs seems to be liver-specific.


63<br />

Metallothioneins (MT I and MT II) and their gene expression in<br />

some chronic liver diseases<br />

Mohamed Sharaf-Eldin*, Saber Ismail*, Usama Negm*, Hala Hamouda**, and<br />

Amal Abdel Hameed***<br />

Departments of Tropical Medicine*, Medical Biochemistery**, and Clinical<br />

Pathology***, Faculty of Medicine, Tanta University, Egypt<br />

Introduction: Much attention was paid to the association between metallothioneins<br />

and chronic liver diseases including hepatocellular carcinoma (HCC).<br />

The aim of the work was to study the role of metallothioneins in chronic liver diseases<br />

and to clarify the role of metallothioneins I and II mRNA expression in hepatocellular<br />

carcinoma.<br />

Methods: The study was carried out on 45 patients with liver diseases (15 patients<br />

with chronic hepatitis, 15 patients with liver cirrhosis and 15 patients with HCC) as<br />

well as 15 healthy individuals as a control group. All patients and controls were<br />

subjected to estimation of metallothioneins levels; also their tissue levels were<br />

estimated in all patient groups. Metallothioneins (MT I and MT II) mRNA expression<br />

by RT-PCR were done for all cases.<br />

Results: The results of the present study showed a significant decrease in serum<br />

and tissue levels of metallothioneins in patients groups. More changes were<br />

demonstrated in HCC patients. Concerning the PCR results of MT genes expression,<br />

there was a significant decrease in MT I and MT II mRNA expression in HCC patients<br />

when compared to the other groups. They also decreased in patients with liver<br />

cirrhosis when compared to the control group and patients with chronic hepatitis. In<br />

contrast their expression does not show significant difference in chronic hepatitis<br />

when compared to the control group.<br />

Discussion/Conclusion: It could be concluded that metallothioneins levels may be<br />

used as a non-invasive biochemical markers for early detection of the progression of<br />

chronic liver diseases. Moreover, the progressive decrease in MT I and MT II gene<br />

expression may play an important role in carcinogenesis of HCC.


64<br />

Endothelial chemokine presentation as therapeutic target in<br />

T-cell mediated hepatitis<br />

Katrin Neumann 1 , Katja Wechsung 1 , Nils Kruse 1 , Michael Schumann 1 , Anja Kühl 2 ,<br />

Alf Hamann 3 , and Katja Klugewitz 1<br />

1 Medizinische Klinik I, Charité, Universitätsmedizin Berlin, Germany<br />

2 Medizinische Klinik I, Research Center ImmunoSciences (RCIS), Charité,<br />

Universitätsmedizin Berlin, Germany<br />

3 Experimentelle Rheumatologie, Charité, Universitätsmedizin Berlin, Germany<br />

Introduction: Transmigration across liver endothelium maintains the homeostatic<br />

balance of intrahepatic T cells and is a central process in liver inflammation. This<br />

study focuses on chemokine-dependent interactions between liver sinusoidal endothelial<br />

cells (LSEC) and lymphocytes and the influence of these processes in T-cell<br />

transmigration into hepatic parenchyma during autoimmune hepatitis.<br />

Methods: Transmigration analysis of CD4 + T cells across LSEC was performed by<br />

the transwell-assay system. Confocal laser scanning microscopy was used for<br />

visualizing chemokine uptake by LSEC. For induction of T-cell mediated hepatitis<br />

Concanavalin A (ConA) was transferred into mice and chlorpromazine was<br />

administered during hepatitis.<br />

Results: LSEC pre-incubated with CXCL12 actively took up the chemokine and<br />

presented it to CD4 + T cells leading to enhanced transmigration of these cells across<br />

the endothelium. Pre-incubation of LSEC with the clathrin-pathway inhibitor<br />

chlorpromazine blocked CXCL12 uptake and significantly reduced chemokine-driven<br />

CD4 + T cell transmigration. Administration of chlorpromazine during ConA-induced<br />

hepatitis reduced T-cell infiltration into hepatic parenchyma counteracting<br />

development of necrosis. Assessing the mode of chemokine uptake, we found that<br />

inhibition of CXCR4 by AMD3100 resulted in the inability of LSEC to take up CXCL12<br />

and subsequently reduced chemotaxis.<br />

Discussion/Conclusion: Our data provide a model for actively enhanced CXCL12driven<br />

transmigration of CD4 + T cells through LSEC layers, where the chemokines<br />

are taken up and transcytosed by chemokine receptors and clathrin-coated vesicles.<br />

Chlorpromazine, an already approved therapeutic compound, inhibits the process in<br />

vitro and in vivo and might offer new approaches for the inhibition of lymphocyte<br />

infiltration into an inflamed liver.


Hemostatic disturbances in patients with portal thrombosis<br />

65<br />

E. Lukina, E. Sysoeva, E. Jakovleva, E. Kitsenko*, G. Sukhanova, S. Vasijev<br />

National Research Center for Hematology, *National Research Center for Surgery,<br />

Moscow, Russia<br />

Introduction: The prehepatic portal hypertension (PPH) due to portal thrombosis<br />

was believed to be a rare condition (about 10–20% of all cases of portal hypertension).<br />

However, during the last 5 years we observed more than 100 patients with<br />

portal thrombosis, who developed PPH.<br />

The aim of the investigation: To study hemostasis in patients with prehepatic portal<br />

hypertension (PPH) due to portal thrombosis.<br />

Materials and methods: We studied 58 patients (median age 42.8 years) with PPH<br />

due to portal thrombosis confirmed by Doppler sonography. The period from the first<br />

manifestation of portal hypertension (splenomegaly, varicose dilatation of esophageal<br />

veins) to examination in our Center varied from 6 to 480 months (M = 92 months).<br />

26 patients (mean age = 45 years) had blood picture compatible with diagnosis of<br />

chronic myeloproliferative disorders (MPD). Other 32 patients (mean age = 41 years)<br />

had normal pattern of bone marrow, normal blood picture or cytopenias. We studied<br />

the basic laboratory parameters concerning the hemostatic status and some<br />

additional parameters: von Willebrand factor, Lupus anticoagulant (LA) homocysteine,<br />

endothelin, soluble thrombomodulin (ELISA).<br />

Results: The majority of patients (pts) had normal rates of fibrinogen, antithrombin III<br />

and protein C. Prothrombin level was decreased in 62% of patients. The level of von<br />

Willebrand factor was increased in 55% of patients, median = 175 (normal rate<br />

< 150). LA was found in 46% of patients. The antiphospholipid antibodies (APL-Abs)<br />

to membranes phospholipids and β2-glycoprotein I were studied in 37 patients. APL-<br />

Abs were found in 22 patients LA+ and in 15 patients LA-, most frequently in patients<br />

with cytopenias. Hyperhomocysteinemia (mild or moderate) was revealed in 54% of<br />

patients. Concentration of soluble thrombomodulin was decreased in 70% of<br />

patients.<br />

Discussion/Conclusion: Thus, patients with PPH due to portal thrombosis had<br />

signs of endothelial damage. The presence of LA and APL-Abs was the typical<br />

feature of PPH in patients with blood cytopenias.


66<br />

False iron deficiency in patients with prehepatic portal<br />

hypertension<br />

E. Lukina, E. Sysoeva, A. Levina, J. Mamukova, E. Voronkova, E. Kitsenko*<br />

National Research Center for Haematology, Moscow, Russia<br />

*National Research Center for Surgery, Moscow, Russia<br />

Background: Patients with prehepatic portal hypertension caused by portal<br />

thrombosis often have recurrent bleedings from esophageal and gastric varices.<br />

Аnemia in these patients is considered to be posthemorrhagic iron deficiency<br />

anemia. Therefore, patients often receive iron treatment.<br />

The aim of the investigation: To study the iron metabolism in patients with<br />

prehepatic portal hypertension due to portal thrombosis.<br />

Materials: We studied 60 patients (22 male and 38 female, median age 43 years)<br />

with prehepatic portal hypertension (PPH) caused by portal thrombosis. 37 (62%)<br />

patients had bleedings from esophageal and gastric varices. The period from the first<br />

manifestation of portal hypertension (splenomegaly, varicose dilatation of esophageal<br />

veins) to examination in our Center varied from 4 to 480 months (median =<br />

72 months). 32 patients had hypo- or normochromic anemia, 28 patients had normal<br />

hemoglobin level.<br />

Methods: We investigated the serum indices of iron metabolism (iron concentration,<br />

total iron binding capacity, transferrin iron saturation, ferritin, transferrin). Hepcidin<br />

level and urine iron excretion were examined in part of the patient group. The control<br />

group consisted of 35 healthy volunteers.<br />

Results: All patients had low or normal levels of serum iron (median = 9.9 mmol/l),<br />

ferritin (m = 39 mkg/l, normal = 40–200) and transferrin iron saturation (TIS,<br />

m = 17.4%, normal = 25–35%). At the same time, serum transferrin level was low or<br />

normal in 64% of patients (median 2.65 g/l, normal 2.65 ± 0.05) and serum indices of<br />

soluble transferrin receptors was not increased in majority of patients. Positive<br />

desferal test revealed increased urine iron excretion in 33 from 35 examined patients.<br />

We also did not find any differences in iron metabolism indices in patients with and<br />

without anemia.<br />

Conclusion: Iron deficiency in patients with portal thrombosis and PPH may<br />

represent the failure of iron metabolism regulation in liver cells caused by ischemic<br />

injury of the liver.<br />

Correspondence to:<br />

Prof. Elena Lukina, National Research Center for Haematology, Noviy Zykovsky pr.<br />

4, 125167, Moscow, Russia, Tel/Fax: 7 (095) 612-0923, E-Mail: lukina@blood.ru


67<br />

Diagnostics improvement for liver chronic diseases in<br />

Turkmenistan<br />

O.B. Nepesova<br />

Turkmen State Medical Institute, Internal Medicine, Ashgabat, Turkmenistan<br />

The study aims to improve diagnostics for liver chronic diseases in Turkmenistan.<br />

Materials and methods: 184 patients at age of 17–77 (averagely 32.5 years) with<br />

various diffuse liver diseases were examined, covering 82 males and 102 females.<br />

All the patients have undergone a liver ultrasonic examination and blood biochemical<br />

indexes as follows have been identified: a transaminase activity, bilirubin and<br />

cholesterol level, a number of lymphocytes and serum protein fractions. The virus<br />

hepatitis markers as follows have also been identified with an immune-enzyme<br />

method. The PCR examinations for DNA HBV and RNA HCV have been held at<br />

serum-positive patients.<br />

Results: The immunological markers of B, C, and D virus hepatitis were found out at<br />

130 patients (70.7%), and no immunological markers were found at 54 patients<br />

(29.3%). Laboratory tests demonstrated a HCV monoinfection at 78 patients (60.0%),<br />

a HBV monoinfection at 18 patients (13.8%) and a mix-virus infection at 34 patients<br />

(26.2%) in seropositive group of patients under observation: B+C hepatitis viruses<br />

were found out at 8 persons, B+C+D at 2 ones, B+D at 24 ones. 1 genotype of the<br />

present virus was found out at 62 patients with a chronic C virus hepatitis, which is<br />

specific for more severe treatment and requires a longer therapy.<br />

Conclusion: Up-to-date laboratory study methods allow to timely and efficiently<br />

select an antivirus therapy required for patient, to monitor efficiency and to forecast a<br />

clinical course.


68<br />

Epidemiology of liver diseases in a pediatric population.<br />

Snapshot data from selected centers in Poland<br />

Woynarowski Marek, Woźniak Małgorzata, Chlebcewicz-Szuba Walentyna,<br />

Chmurska Motyka Teresa, Gorczyca Anna, Korczowski Bartosz, Krzywicka Anna,<br />

Lebensztejn Dariusz, Liberek Anna, Majda-Stanisławska Ewa, Rokitka Maria,<br />

Strawińska Elżbieta, Więcek Sabina, Zaleska Izabella, Zdanowska-Ruskań Anna,<br />

Socha Jerzy<br />

Polish Autoimmune Hepatitis Study Group (PEGAZ), Warsaw, Poland<br />

Introduction: There are no large cross-sectional studies on epidemiology of liver<br />

diseases in pediatric population in Poland. The collaborative group for pediatric liver<br />

diseases (PEGAZ) was established to describe the pediatric hepatology problems in<br />

Poland.<br />

Methods: The questionnaire survey was sent to the hospital involved in liver disease<br />

therapy in children. 13 centers responded to the questionnaire.<br />

Results: Six responding centers are infectious disease hospitals. 6 centers defined<br />

themselves as pediatric hospitals and one site is a gastroenterology, hepatology and<br />

liver transplant center. All centers provide medical care on inpatient and outpatient<br />

basis for children aged 0–18 years. The number of patients treated at every site<br />

varies from 900 to 7000 per year (total around 36,000). Liver disease was present in<br />

3054 (8%). Viral liver disease was present in 1449 subjects (HBV – 1033, HCV – 356<br />

and HBV +HCV in 60). Nonviral liver disease was present in 1605 children (Gilbert<br />

syndrome – 596, liver transplant recipients – 230, cholestasis – 201, autoimmune<br />

hepatitis – 187, Wilson disease – 117, A-1-ATD – 113, cholelithiasis – 83 and other<br />

diseases in 78 children). Combined viral and nonviral liver disease was present in<br />

98 children. The highest rate of hepatotropic viruses coinfection was observed in liver<br />

transplant recipients (17%), autoimmune hepatitis (13.4%) and A-1-ATD (8.8%). The<br />

lowest rate of viral infections was noted in patients with Gilbert syndrome (1.5%).<br />

Discussion/Conclusion: Our data show that the viral infections (especially HBV<br />

infection) remain the first cause of liver problems in Polish pediatric population.


69<br />

Alcohol withdrawal treatment Audit March 2009, Northern<br />

General Hospital, Sheffield, UK<br />

<strong>Dr</strong>. A. Al-Joudeh, <strong>Dr</strong>. H. Morton, <strong>Dr</strong>. S. Heikal, <strong>Dr</strong>. K. Basu<br />

Sheffield University Teaching Hospitals Foundation NHS Trust, Sheffield, UK<br />

Introduction: In the UK there are approximately 208,000 hospital admissions<br />

(2007/8) with primary or secondary diagnosis of alcohol misuse. 13.3/100,000 deaths<br />

in the UK are alcohol related.<br />

Objectives: To evaluate our current practice in treating alcohol withdrawal with<br />

reference to the local guidelines. Review guidelines and promote the role of<br />

psychiatry liaison team.<br />

Methods: This was a retrospective audit of case notes and prescription charts<br />

between (01/08/2008–31/12/2008). Data collected with a performa (designed against<br />

the Sheffield Teaching Hospitals guidelines).<br />

Inclusion criteria: Alcohol excess > 10 units/day in acute medical admissions.<br />

Results: 183 cases identified, 100 available and 51 patients fulfilled the inclusion<br />

criteria. 69% were male, median age 44 years. 90% of the patients had alcohol<br />

intake recorded. 84% had benzodiazepines prescribed for alcohol withdrawal.<br />

Chlordiazepoxide was prescribed in 53%, this was appropriate in 37%. Lorazepam<br />

was prescribed in 31%, this was appropriate in 50%. 22% had evidence of<br />

hallucinations, 27% of them had haloperidol prescribed. 27% had fits, only 29% of<br />

them had lorazepam prescribed correctly. 45% had Pabrinex prescribed, 44% of<br />

them had Pabrinex prescribed according to guidelines. 61% had Thiamine<br />

prescribed. 75% had vitamin B Co Strong prescribed. 33% of the patients were<br />

referred to psychiatry, 40% of them were followed by alcohol dependence services<br />

after discharge.<br />

Discussion/Conclusion: Unsatisfactory results in the management of alcohol withdrawal<br />

syndrome.<br />

Recommendations: Staff to receive teaching on the subject. Clarify the use and<br />

doses of vitamins and increase the dose of benzodiazepine in the new guidelines. To<br />

formulate a pathway streamlines the management of alcohol withdrawal syndrome in<br />

hospital and follows it up by alcohol dependence services.<br />

Updated local guidelines have taken with the recommendations.


70<br />

Effect of different degrees of hepatic dysfunction on the<br />

pharmacokinetics of telmisartan<br />

Manal El-Hamamsy 1 , Sahar Hegazy 2 , Wageh Awara 3<br />

Faculties of <strong>Pharma</strong>cy, Ain Shams University 1 and Tanta University 2,3 , Egypt<br />

Introduction: Telmisartan is a nonpeptide angiotensin II receptor antagonist, used<br />

for the treatment of hypertension.<br />

Objective: To describe the pharmacokinetics and safety of telmisartan in healthy<br />

volunteers and in subjects with different degree of hepatic impairment.<br />

Methods: Prospective study. Single oral dose of telmisartan 80 mg was given to<br />

10 healthy subjects served as a control group (group 1), and 20 subjects with hepatic<br />

impairment; divided into two groups; group 2 (10 patients): patients with mild<br />

hepatospleenomegaly and group 3 (10 patients): patients with cirrhosis and ascites.<br />

Plasma samples were collected and analyzed with high performance liquid<br />

chromatography (HPLC).<br />

Results: The pharmacokinetic profile of telmisartan was characterized by rapid<br />

absorption kinetics and a slow terminal elimination phase with mean half life of<br />

25 hours.<br />

The maximum plasma concentration and area under the telmisartan plasma<br />

concentration-time curve (AUC) increased in all hepatically impaired subjects<br />

compared with healthy volunteers. There was a significant increase in the Cmax and<br />

AUC of the cirrhotic patients and non significant increase in the patients with mild<br />

hepatosplenomegaly compared with the healthy volunteers. There was no change in<br />

the half life of telmisartan in all the patients and no adverse events were noticed in<br />

the patients during the study indicating the good tolerability of the drug in<br />

hypertensive patients with hepatic dysfunction.<br />

Discussion/Conclusion: Lower doses of telmisartan should be considered in<br />

cirrhotic patients as well as in the patients with mild hepatitis.


71<br />

Risk perception and use of herbal remedies in patients with<br />

liver/biliary tract disorders: An Italian study<br />

Massimo Marignani, Sara Gallina, Michela Di Fonzo, Ilaria Deli, Paola Begini,<br />

Fabio Attilia, Elia Gigante, Marcella Epifani, Stefano Angeletti, Gianfranco Delle Fave<br />

Digestive and Liver Disease Department, Second Medical Faculty University<br />

“Sapienza” at S. Andrea Hospital, Rome, Italy<br />

Introduction: Use of herbal remedies (HR) has increased in the general population,<br />

particularly among patients with various chronic diseases/conditions. <strong>Pharma</strong>covigilance<br />

and regulations regarding HR are still incomplete, and HR-related adverse<br />

reactions are increasingly reported. Nevertheless, studies assessing prevalence of<br />

HR use among patients with liver/biliary tract disorders are limited and no data are<br />

available in Italy. The aims of the study were to assess the prevalence and the risk<br />

perception of HR use in the population attending our outpatient Liver/Biliary Tract<br />

Disorders Clinic. Therefore, to asses the main clinical and demographic characteristics<br />

of patients using HR.<br />

Methods: From October 2007 to April 2008, 231 consecutive patients (119 M, 112 F)<br />

were interviewed, using an ad hoc developed questionnaire. Face to face<br />

questionnaire addressed the following: body mass index (BMI), presence of chronic<br />

conditions, use of conventional therapy, HR-use and perceptions regarding potential<br />

harmful HR-drug interactions. Data were expressed as mean (± SD) or number/total,<br />

and evaluated by student-t and Fisher tests as appropriate. Multivariate logistic<br />

regression (MLR) was also performed.<br />

Results: Prevalence of HR use was 35.5% and 72% of patients using HR had never<br />

considered possible, potentially harmful HR-drug interactions. Therefore 67% of<br />

users used HR in addition to conventional therapy. At MLR use of HR was more<br />

common in women (p = 0.01), and in patients practicing regular sports activity<br />

(p = 0.03). Users were more affected by chronic conditions than non users<br />

(p = 0.0002).<br />

Discussion/Conclusion: More than a third of patients attending Liver/Biliary<br />

Disorders Clinic use HR. Misconceptions about the risk of HR use are widespread<br />

among them. This issue should be specifically considered and addressed with<br />

patients and considered in their management.


72<br />

Hepatotoxicity induced by antituberculosis drugs<br />

M. González, E. Aravena, C. Peña, E. Astrosa, J. Lyon, G. Serrano, P. Escobar<br />

San Borja Arriarán Hospital, Finis Terrae University Santiago, Chile<br />

Introduction: Tuberculosis treatment requires combination of several medicines for<br />

long time, which facilitates emergence of adverse reactions. Liver is involved in drugs<br />

biotransformation, so it develops hepatotoxicity. Its presence demands therapy<br />

suspension and it has mortality, though it is necessary to review the clinical<br />

presentation profile, liver chemistry test pattern, evolution and risk factors associated.<br />

Methods: We reviewed patients’ clinical histories with hepatotoxicity induced by<br />

antituberculosis drugs registered at the Metropolitan Central Health Service of<br />

Santiago of Chile between January 1998 and April 2009. We consigned onset time,<br />

liver chemistry tests alteration pattern, epidemiologic data and their evolution.<br />

Results: In 2676 TBC cases, we identified 53 patients who developed liver toxicity<br />

by drugs (2%). The average age is 50.3±18.4 years (interval 22–81). Gender<br />

proportion: 29 men (57%) and 24 women. Antecedents: 2 VIH patients, 6 cirrhotic<br />

patients, 12 were using multiple drugs. Out of 38 patients consigned, just 8<br />

consumed alcohol. In 34 cases, 29 debuted during the daily dose and 5 during the<br />

twice weekly phase. Liver test pattern: 14 hepatocellular damage (47%), 9 cholestatic<br />

(30%) and 7 mixed (23%). The majority improved with halting and adoptation of a<br />

new anti-TBC schedule, except two patients (3.8%) who developed acute liver<br />

failure, just one survived requiring a liver transplant. The one who died had cirrhosis.<br />

Discussion/Conclusion: In hepatotoxicity induced by antituberculosis drugs, it<br />

emphasizes the antecedents of concomitant ingest of multiple medicines, alcohol<br />

consume and cirrhosis presence. Most of cases present at the beginning of<br />

treatment, highlighting a hepatocellular damage pattern. Patients can develop acute<br />

liver failure.


Treatment with Ursofalk ® of bile reflux gastritis<br />

<strong>Dr</strong>. Dorina Pestroiu, <strong>Dr</strong>. Madalina Ilie, <strong>Dr</strong>. Dana Brooks, <strong>Dr</strong>. Cosmina Vladut<br />

Clinical Emergency Hospital Bucharest, Romania<br />

73<br />

Introduction: Bile acids refluated in the stomach produce cell destruction,<br />

immunologic modifications or/and inflammatory effects.<br />

Recent studies have shown that administration of UDCA increases its concentration<br />

in the gastric reflux inducing the decrease of the cell destruction effects; it has antiinflammatory<br />

and immune-modulator effect leading to the reduction of the symptoms.<br />

Also, UDCA has a role in normalizing the intestinal motility, decreasing the<br />

duodenogastric and gastroesophageal reflux.<br />

Methods: 45 patients (12 female and 33 male) were studied, with average age of 47,<br />

from which 15 had stomach resection, 12 had cholecystectomy, the rest having the<br />

disease in motility disturbances.<br />

There were included in the study patients whose Helicobacter pylori infection and<br />

AINS consumption were excluded and to whom a gastric mucosa biopsy was<br />

performed which showed mucosa edema with moderate lymphoplasmocyte<br />

infiltration.<br />

UDCA was administrated in dose of 10 mg/kg, for 14 days a month, 3 consecutive<br />

months, in unique dose in the evening.<br />

Results: The patients were examined after 2 weeks and the reduction of clinical<br />

symptoms was observed at 50% of the patients.<br />

At 3 months after beginning the treatment were performed: clinical exam, upper<br />

digestive endoscopy with biopsies taken. At 75% of the patients, the remission of<br />

clinical, endoscopic and histological manifestations was obtained.<br />

Discussion/Conclusion: Adding UDCA in bile reflux gastritis provides a marked<br />

improvement of symptoms after 3 months of treatment with 10 mg/kg, for 14 days a<br />

month.


74<br />

Incidence and therapy of colonic perforations: How feasible<br />

and effective is endoscopic closure?<br />

Klaus Mönkemüller, MD, PhD, FASGE 1,2 , Helmut Neumann, MD,<br />

Peter Malfertheiner, MD, PhD, H.-U. Schulz, MD, PhD, Lucia C. Fry, MD 1,2<br />

1 Department of Internal Medicine, Gastroenterology, Hepatology and Infectious<br />

Diseases, Marienhospital, Bottrop, Germany<br />

2 Divsion of Gastroenterology, Hepatology and Infectious Diseases, Otto-von-<br />

Guericke-University, Magdeburg, Germany<br />

Introduction: Colon perforation is one of the most dreaded complications of<br />

colonoscopy. Traditionally, patients with colon perforation have been treated<br />

surgically. Although there are several case reports documenting the usefulness of<br />

endoscopic closure of colon perforations, there are few current data evaluating the<br />

current management of colon perforations.<br />

The aim of the study was to assess the incidence of colon perforations and the utility<br />

of immediate endoscopic closure of the defect in a tertiary endoscopy unit.<br />

Methods: All patients who underwent colonoscopy at our institution from June 2001<br />

to December 2008 were identified. Data were obtained by searching our<br />

prospectively collected electronic database (Medos, Langensebold, Germany).<br />

Results: During the study period, a total of 8601 colonoscopies were performed<br />

(2472 therapeutic interventions, 28.7%). A total of 12 iatrogenic colonic perforations<br />

occurred, yielding an incidence of 0.14%. Five (41.7%) occurred during a diagnostic<br />

colonoscopy (perforation rate of 0.08%). The remaining 7 perforations (58.3%)<br />

occurred as the result of a therapeutic intervention, perforation rate, 0.28% (p < 0.02,<br />

as compared to perforations occurring during a diagnostic colonoscopy). These<br />

included: polypectomy, n = 4, dilation of stenosis, n = 2, decompression tube<br />

placement, n = 1. Successful endoscopic closure of the perforation site was possible<br />

in 5 patients (42%). Seven patients were treated surgically (large defects, n = 3,<br />

difficult scope position, n = 2, stool contamination, n = 2).<br />

Discussion/Conclusion: The perforation rate of colonoscopy was 0.14%.<br />

Immediate endoscopic closure of the perforation was possible in 42% of patients.<br />

Patients with defects larger than 2 cm and those located in angulated areas or the<br />

presence of stool contamination should undergo immediate laparotomy.


Author Index to <strong>Poster</strong> <strong>Abstracts</strong><br />

(Name - <strong>Poster</strong> Number)<br />

AbouNar, A. 33<br />

Adamenko, E.I. 9<br />

Adjarov, D.G. 17<br />

Ahmad, Y. 33<br />

Al-Joudeh, A. 69<br />

Altunkova, I.P. 50<br />

Ananiev, J. 50, 51, 52<br />

Andreescu, O. 22<br />

Angeletti, S. 71<br />

Angelico, M. 31<br />

Appleby, V.J. 15<br />

Aravena, E. 72<br />

Askar, E. 35<br />

Astrosa, E. 72<br />

Attilia, F. 71<br />

Austin, L. 26<br />

Awara, W. 70<br />

Azmy, M. 33<br />

Badea, D. 41, 42, 47<br />

Badea, M. 41, 47<br />

Bahcecioglu, I.H. 8<br />

Baiocchi, L. 31, 32<br />

Baleva, D. 7<br />

Banciu, C. 19<br />

Basu, K. 69<br />

Bayer, M. 58, 59<br />

Beard, M. 26<br />

Bebb, O. 15<br />

Begini, P. 71<br />

Belyaeva, T.V. 23<br />

Berres, M.-L. 27<br />

Bockmann, J.H. 16<br />

Boitan, M. 37<br />

Boryczka, G. 28<br />

Borza, C. 19<br />

Bregadze, R. 29<br />

Bremer, B. 14<br />

Brooks, D. 73<br />

Buk, L. 53<br />

Carambia, A. 54<br />

Carbone, M. 31, 32<br />

Casey, C.A. 60<br />

Cecere, R. 32<br />

Celik, N. 1<br />

Celinski, K. 53, 55<br />

Chakarova, P. 44<br />

Chen, J.L. 20<br />

Chlebcewicz-Szuba, W. 68<br />

Chmurska Motyka, T. 68<br />

Christen, U. 58, 59<br />

Cichoz-Lach, H. 53, 55<br />

Cijevschi Prelipcean, C. 18, 36<br />

Ciof, O.P. 2, 3, 10, 21,<br />

45, 48<br />

Ciuclan, L. 20<br />

Claaß, B. 56<br />

Conti, L.R. 31<br />

Copaceanu, L. 19<br />

Cornberg, M. 14, 24<br />

Dabrowska, M. 30<br />

Dandri, M. 16<br />

De Leonardis, F. 31, 32<br />

Deac, M. 37, 46<br />

Deli, I. 71<br />

Delle Fave, G. 71<br />

Delle Monache, M. 31, 32<br />

Demir, M. 6<br />

Derkow, K. 40<br />

Di Fonzo, M. 71<br />

Dienes, H.-P. 35<br />

Donohue, T. 25<br />

Dooley, S. 20<br />

<strong>Dr</strong>agan, E. 2, 3, 10, 21,<br />

45, 48<br />

<strong>Dr</strong>aghila, L. 37<br />

<strong>Dr</strong>agna, M. 18, 36<br />

<strong>Dr</strong>ucker, C. 57<br />

<strong>Dr</strong>ug, V.L. 18, 36<br />

Dusheiko, G.M. 12<br />

Dzieran, J. 20<br />

El-Batae, H. 34<br />

El-Hamamsy, M. 33, 70<br />

Epifani, M. 71<br />

Ergün, D. 1<br />

Erhardt, A. 56, 62<br />

Esaulenko, E.V. 23<br />

Escobar, P. 72<br />

Flisiak, R. 30<br />

Fratila, O. 37


Frenzel, C. 54<br />

Fry, L.C. 74<br />

Galcheva, S. 7<br />

Gallina, S. 71<br />

Gebhardt, R. 20<br />

Gencaslan, A. 8<br />

Genunche-Dumitrescu, A. 41, 42, 47<br />

Georgiev, K. 44<br />

Gewiese, J. 57<br />

Gigante, E. 71<br />

Goeser, T. 6<br />

Golebiewski, J. 13<br />

González, M. 72<br />

Gora-Gebka, M. 13<br />

Gorczyca, A. 68<br />

Gorski, J. 30<br />

Grabowski, J. 14, 24<br />

Graur, L. 18, 36<br />

Gulubova, M.V. 50, 51, 52<br />

Gutkowski, K. 28<br />

Hamann, A. 40, 64<br />

Hameed, A.A. 63<br />

Hamouda, H. 63<br />

Hartleb, M. 5, 28<br />

Hegazy, S. 70<br />

Heikal, S. 69<br />

Helliwell, P. 15<br />

Herkel, J. 54<br />

Heuchel, R. 20<br />

Hintermann, E. 58, 59<br />

Holdener, M. 59<br />

Huang, T. 20<br />

Huber, S. 54<br />

Ilie, M. 73<br />

Ilkavets, I. 20<br />

Jakovleva, E. 65<br />

Jaroszewicz, J. 14<br />

Jaroszewicz, J. 24<br />

Jelev, D. 17<br />

Julianov, A. 50<br />

Kacperek-Hartleb, T. 28<br />

Kajor, M. 28<br />

Kalinova, K. 44<br />

Kaminska, B. 13<br />

Kaminska-Kiszka, E. 28<br />

Kara, S. 1<br />

Khanna, P. 12<br />

Kharbanda, K. 26<br />

Kitsenko, E. 65, 66<br />

Klugewitz, K. 40, 64<br />

Korczowski, B. 68<br />

Korobowicz, E. 53<br />

Kosseva, O. 17<br />

Kowalska, A. 27<br />

Krastev, Z. 17<br />

Kruse, N. 40, 64<br />

Krzywicka, A. 68<br />

Kühl, A.A. 40, 64<br />

Kuzu, N. 8<br />

Kyurkchiev, D. 50<br />

Lang, S.M. 6<br />

Lawniczak, M. 4, 5<br />

Lebensztejn, D.M. 68<br />

Lenci, I. 31, 32<br />

Levina, A. 66<br />

Liberek, A. 13, 68<br />

Lis, E. 53, 55<br />

Liu, Y. 20<br />

Lohse, A.W. 16, 54<br />

Lukina, E. 65, 66<br />

Luster, A. 58<br />

Lütgehetmann, M. 16<br />

Lüth, S. 54<br />

Lyon, J. 72<br />

Maini, M. 12<br />

Majda-Stanislawska, E. 68<br />

Malfertheiner, P. 74<br />

Mamukova, Y. 66<br />

Manns, M.P. 14, 24, 59<br />

Manolova, I.M. 50, 51, 52<br />

Marignani, M. 71<br />

Martin, M. 48<br />

Martinova, E.A. 43<br />

Mateescu, R. 19<br />

Matev, A. 51<br />

McVicker, B.L. 60<br />

Mertens, J. 38<br />

Mertens, P.R. 20<br />

Metin, K. 8<br />

Meyer, C. 20<br />

Mihai, B. 18, 36<br />

Mihai, C. 18, 36<br />

Mihaila, R. 37<br />

Mihaila, R.G. 37, 46<br />

Mihm, S. 29, 35, 38


Milkiewicz, P. 5<br />

Mitrut, A.O. 42<br />

Mitrut, P. 41, 42, 47<br />

Mocanu, L. 37<br />

Mönkemüller, K. 74<br />

Moreea, S. 15<br />

Moreno Zaldivar, M. 27<br />

Morton, H. 69<br />

Musialik, J. 28<br />

Nebbia, G. 12<br />

Nedelcu, L. 22<br />

Negm, U. 63<br />

Nepesova, O.B. 67<br />

Neumann, H. 74<br />

Neumann, K. 40, 64<br />

Nita, L. 42<br />

Nosotti, L. 31<br />

Odenthal, M. 35<br />

Olteanu, A. 37<br />

Osna, N.A. 25, 26, 60<br />

Özercan, I.H. 8<br />

Pacurari, A. 19<br />

Panasiuk, A. 30<br />

Pantea, I. 22<br />

Panzer, U. 62<br />

Pauels, K. 27<br />

Paul, D. 22<br />

Pekova, L.M. 44<br />

Peña, C. 72<br />

Pestroiu, D. 73<br />

Petersen, J. 16<br />

Petrov, A. 11<br />

Pophristova, E. 17<br />

Pothakamuri, S.V. 24<br />

Prozorow-Król, B. 53<br />

Ramadori, G. 29, 35, 38<br />

Raszeja-Wyszomirska, J. 4, 5<br />

Rezi, E.C. 37, 46<br />

Rill, D. 2, 3, 10, 21,<br />

45, 48<br />

Rokitka, M. 68<br />

Romanov, A.O. 23<br />

Romosan, I. 19<br />

Rose-John, S. 49, 57<br />

Ross Lopes, A. 12<br />

Rubinas, T. 61<br />

Saber, I. 63<br />

Sahin, K. 8<br />

Samdanci, E. 1<br />

Savoiu, G. 19<br />

Scarneciu, C. 22<br />

Scheller, J. 49, 57<br />

Schlaphoff, V. 14, 24<br />

Schmitz, P. 27<br />

Schramm, C. 54<br />

Schulte, S. 6<br />

Schulz, H.-U. 74<br />

Schumann, M. 64<br />

Schurich, A. 12<br />

Schütt, A. 49<br />

Schwinge, D. 54<br />

Scurtu-Martin, M. 2, 3, 10, 21<br />

Serban, C. 19<br />

Serrano, G. 72<br />

Sharaf-Eldin, M. 34, 63<br />

Sikorska-Wisniewska, G. 13<br />

Silivontchik, N.N. 9<br />

Singer, M.V. 20<br />

Slomka, M. 53, 55<br />

Socha, J. 39, 68<br />

Southern, P. 15<br />

Sporea, I. 2, 3, 10, 21,<br />

45<br />

Stachowska, E. 4<br />

Stefanescu, G. 18<br />

Stefanova, P. 44<br />

Steffen, H.-M. 6<br />

Stegmann, K. 24<br />

Strawinska, E. 68<br />

Suchy, J. 4<br />

Sukhanova, G. 65<br />

Suneetha, P.V. 14<br />

Susan, L.M. 19<br />

Swatek, J. 53<br />

Sysoeva, E. 65, 66<br />

Szymanik, B. 5<br />

ten Dijke, P. 20<br />

Thiele, G.M. 60<br />

Tiegs, G. 56, 62<br />

Töx, U. 6<br />

Trautwein, C. 27<br />

Tudora, A. 2, 3, 10, 21,<br />

45, 48<br />

Tuma, D.J. 60<br />

Tzaneva, V. 7


Ubben, S. 6<br />

Üstündag, B. 8<br />

Varnholt, H. 61<br />

Vasijev, S. 65<br />

Vatev, N. 11<br />

Vladut, C. 73<br />

Volz, T. 16<br />

Voronkova, E. 66<br />

Wasmuth, H.E. 27<br />

Wechsung, K. 64<br />

Wedemeyer, H. 14, 24<br />

Wegscheid, C. 62<br />

Weinman, S. 25<br />

Weiskirchen, R. 27<br />

Weng, H.-L. 20<br />

White, R. 25, 26<br />

Wiecek, S. 68<br />

Wiercinska, E. 20<br />

Woynarowski, M. 39, 68<br />

Wozniak, M. 39, 68<br />

Yovchev, Y. 52<br />

Yüksel, F. 1<br />

Yüksel, I. 1<br />

Zabielski, P. 30<br />

Zaharie, A.V. 37<br />

Zaleska, I. 68<br />

Zawada, I. 4<br />

Zdanowska-Ruskan, A. 68<br />

Zhu, C. 20


Innovative <strong>Dr</strong>ugs<br />

for bowel and liver diseases<br />

Modern formulations and specially designed delivery systems<br />

ensure targeted release of the active drug<br />

Scientific Dialogue<br />

in the interest of therapeutic progress<br />

<strong>Falk</strong> Symposia and Workshops<br />

morethan 200, attended by some 100,000 participants from over 100 countries since 1967<br />

Continuing medical education seminars<br />

over 14,000, attended by morethan one million physicians and patients in Germany alone<br />

Comprehensive literature service for healthcare professionals and patients<br />

with more than 200 publications<br />

www.falkfoundation.org www.drfalkpharma.com<br />

Leinenweberstr.5 79108 Freiburg Germany Tel +49 (0)761/1514-0 Fax +49 (0)761/1514-321 Mail zentrale@drfalkpharma.de

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!