IMMUNE EVASION STRATEGIES OF VIRUS

Master Thesis

Masters Program & Immunity

Kevin Budding, 0470201

October 2010

Supervisor: Dr. Debbie van Baarle

About the front cover:

The picture is a fluorescence microscope image of a GFP (Green fluorescent protein) labeled HCV protein, which associates to lipid droplets (red). The lipid mechanism is important for HCV and the association gives an indication of that importance. Picture is courtesy of Mr. Torsten Schaller, Immunology & Molecular Pathology department, University College London, 15th April 2010

Reference: http://www.grad.ucl.ac.uk/comp/2006-2007/research/gallery/index.pht?entryID=126

1 OUTLINE

1. Introduction 3

2. 4

2.1 4

2.2 Hepatitis C pathogenesis 6

2.3 Genomic organization and protein function 6

2.4 Viral life cycle 9

3. The innate immune response against HCV and viral evasion 12

3.1 Pattern recognition and NS3/4A 12

3.2 Interferon-β signaling and viral evasion 14

3.3 Interferon stimulation genes and viral evasion 14

3.4 Natural killer cells and HCV 15

4. The adaptive immune response against HCV and viral evasion 17

4.1 The humoral immune response and viral evasion 17

4.2 The cellular immune response and dendritic cell dysfunction 17

4.3 failure 20

4.4 Regulator T cells 21

4.5 T cell homing capacity 21

4.6 Immune evasion through viral mutations 22

5. Therapies and 24

5.1 Current and future therapies 24

5.2 development 26

6. References 28

2 1. INTRODUCTION

In this thesis I will discuss the “tug of war” between hepatitis C virus (HCV) and our innate and adaptive . A chronic HCV infection is the leading causative agent for liver cirrhosis, and currently over 200 million people are infected worldwide. According to epidemiological studies, HCV infects around three till four million people per year. I will start with elaborating on the epidemiology and pathogenesis of the virus, followed by the viral characteristics of HCV, taxonomy, replication cycle, structure, and genome.

The focus of this paper will be on the aspect of viral persistence. In order to understand the immune evasive strategies of HCV I have divided the immunological response in two different components, the innate and adaptive component. I have tried to integrate the type of immune response with the immune evasion strategies of HCV. In the therapy section of the thesis I will elaborate on the currently used treatments, their mechanisms of action, the possible potential for vaccine development, and the reaction of the virus on the various therapeutic compounds.

With this thesis I have tried to give a coherent, comprehensive insight into the clinical relevance of an HCV infection, the viral characteristics, the immune responses against HCV and HCVs strategies to evade these, and initiate a persistent infection.

Kevin Budding, October 2010

3 2. HEPATITIS C VIRUS

2.1 EPIDEMIOLOGY

The hepatitis C virus (HCV) was discovered in 1989 via molecular cloning instead of virus purification1. The first signs of a different form of hepatitis (known as non-A, non-B hepatitis) were already present around the 1970s2. Data up to 2005 indicate that the seroprevalence (number of persons who test positive for a specific disease based on blood serum specimens) of HCV is 2%, accounting for 123 million people in the world, making it the major causative agent of chronic liver diseases worldwide3. Although the prevalence of HCV infection is rather high worldwide, there is geographical spreading

Figure 1. Worldwide prevalence of HCV3.

Seen in the seroprevalence of the virus (see figure 1). Although the figure depicted above shows quite a clear figure of HCV seroprevalence throughout the world, with the virus mostly present in both Africa and Asia and less in the developed world. It is very difficult to obtain the data and to interpret this. This is due to the fact that most studies are conducted within non-representative populations, such as blood donors or patients who already suffer from chronic liver disease. However, mathematical models have been used in developed countries to measure the seroprevalence of HCV. Via these models it is shown that in the USA there is an increasing

4 seroprevalence of HCV over the past decades, 0-44 per 100000 before 1965, increasing to 100-200 around the 1980s4. An Australian model based upon new HCV also shows a steady increase in new HCV cases over the past decades5. Although there seems to be a decline in newly acquired HCV infection in the 1990s, al mathematical models used for the prediction of HCV seroprevalence support the idea that this seroprevalence will rise in the coming decades.

Transmission of HCV occurs via blood-to-blood contact. Important factors that contribute to this transmission are blood transfusions from unscreened donors, drug injections, and non-sterile therapeutical procedures. The most important factor contributing to HCV transmission in developed countries is the usage of injection drugs. However, in developing countries non-sterile usage of syringes in a therapeutic setting is the primary source of HCV transmission6. According to studies conducted in the USA the majority of HCV injections, 68%, is caused by injection drug usage7. An important difference between the transmission of HCV and other viruses spread via blood-to-blood contact is that fewer sharing partners are necessary for a successful transmission of the virus. Also, indirect drug sharing, usage of the same cooker and rinse water, can contribute to HCV transmission8. Non-sterile injections are the primary source of HCV transmission in the developing world. Often sterile syringes are not available, or the injections are supplied by medical staff that lacks the necessary training and/or education. Another factor contributing to the higher transmission rate in a therapeutic setting is that often medication is injected into the patient, whereas in the western world it would be supplied orally9. A terrible example of HCV transmission via therapeutic injections is the kala-azar (Leishmaniasis, a parasitic disease) treatment program in India. The HCV infection rate among multiple injected patients was up to 31.1%10. Another example is the parenteral therapy for schistosomiasis in Egypt. It is believed that the high seroprevalence in Egypt is due to contamination during this therapy program10. In order to tackle this contribution to HCV transmission in 3rd world countries the World Health Organization (WHO) has set up the SIGN program (Safe Injection Global Network). Within this network, governments, international health agencies, corporations, and individuals collaborate for safer therapeutic injections3. The last important factor contributing to HCV transmission is blood transfusion. In the western world the contribution of this factor can almost be neglected due to strict rules and regulations in the all-volunteer donor system. Blood of the donors is tested for multiple blood-borne viruses, including HCV. Figures in the USA show a drop of HCV infections acquired after blood transfusion from 91% to 4%11. However, in most of the countries in the developing world there are no centralized rules and regulations, and no nation-wide blood-testing program is available. According to figures from the WHO, 43% of the donated blood in third world countries is not screened for HCV infection3. It has always been believed that In contrast to other blood-born viruses sexual transmission of HCV hardly occurs12. A

5 recent study by van Laar et al. that focused on acute HCV in HIV-infected men shows that that there is a substantial increase in acute HCV infection in men who have sex with men (MSM) who denied injecting drug use. These findings give new insights into the discussion whether or not sexual contact plays a significant role on HCV transmission13.

2.2 HEPATITIS C PATHOGENESIS

A persistent HCV infection can lead to serious clinical symptoms. When faced with a HCV infection, the majority of the patients will develop chronic liver diseases (50% - 80)14. Steatosis, increased fat content of the liver, is the first clinical symptom during a HCV infection. When developing steatosis the patient will be able to react less effectively to antiviral therapy, and is prone to develop liver fibrosis15. Both liver fibrosis and cirrhosis are common liver disorders following HCV infection. Different risk factors influence the development of liver fibrosis, which can eventually lead to liver cirrhosis. These risk factors are age, sex, alcohol intake, diabetes, and co-infection with other viruses. The morbidity and mortality of chronic HCV infection is related to progressive fibrosis and the development of cirrhosis16. Patients who suffer from chronic HCV infection are prone to develop different types of liver cancers such as hepatocellular carcinoma (HCC). HCC is the fifth most common cancer in the world17, and chronic infection with HCV accounts for 20% of all HCCs worldwide18 Also, chronic HCV infection seems to be a risk factor for the development of cholangiocarcinomas, carcinomas that originate from the extra-hepatic biliary tree and the intrahepatic bile ducts19.

2.3 GENOMIC ORGANIZATION AND PROTEIN FUNCTION

The hepatitis C virus is an enveloped single stranded positive RNA (+ ssRNA) virus and it belongs to the family of Flaviviridae. The genome length is 9.7 kb20. There are six different genotypes. Each genotype is further divided into different subtypes (a, b, c, etc.). The different genotypes are differentiated according to worldwide distribution and sequence variation21. The genome consists of a single open reading frame (ORF) with 2 untranslated regions (UTR) on both sides. These regions are important for viral protein and RNA synthesis and the coordination of these two processes22. Since the genome does not have a 5’cap the internal ribosomal entry site (IRES) in the 5’UTR is essential for translation. The IRES of HCV binds to the 40s ribosomal subunit and induces a conformational change within the subunit making it mRNA bound like. Due to this conformational change eukaryotic initiation factor (eIF) 3 and other essential factors for translation are recruited and the complete complex can than be coupled to the active 80s ribosomal subunit, initiating translation of the ORF23 (see figure 2).

6 The ORF codes for a single polyprotein of 3,000 amino acids. The polyprotein is both co- and post-translationally processed by cellular and virally encoded proteases leading to the formation of both nonstructural and structural proteins, seven and three respectively, ten in total24, 25. The structural proteins are the conserved core protein and the glycosylated E1 and E2 envelope proteins. The proteins are cleaved via a cellular signal peptidase. The mature core protein is essential in viral nucleocapsid formation and it can be separated in two domains, the N-Terminal D1 and the C- Terminal D226. The D1 domain consists of a high amount of positively charged amino acids, and has homology to capsid proteins of other Flaviviridae. Domain D2 is essential for the folding of D1, necessary for membrane characteristics of the core27. Both E1 and E2 glycoproteins are type-1 transmembrane proteins with an N-Terminal ectodomain and a C-Terminal transmembrane domain and they form heterodimers. The proteins are of high importance in various steps of the viral replication cycle (detailed explanation in section 2.3), since these proteins are on the surface of the viral particle they will be in first contact with the potential target for infection.

In between the loci of structural and non-structural proteins lies the small p7 protein. This is a small integral membrane protein and it seems to function as an ion channel28. The polyprotein is further cleaved into the following nonstructural (NS) proteins: NS2, NS3, NS4A, NS4B, NS5A, and NS5B. Cleavage between p7 and NS2 is also conducted by the cellular signal peptidase. Cleavage of the other NS proteins is carried out by two viral enzymes, the NS2 autoprotease, which cleaves NS2 from NS3, and a NS3-4A serine protease, which cleaves the remaining NS proteins within the ORF24 (see figure 2).

All nonstructural proteins are necessary for the intracellular processes of the viral lifecycle. NS2 is a cysteine protease involved in the cleaving of the NS2 – NS3 junction. Besides that it acts as a protease, there is also evidence that it is an essential assembly cofactor during viral particle formation29. The NS3 protein has a broad variety of functions. It has two active domains, an N- Terminal serine protease domain and a C-Terminal RNA helicase/NTPase domain. The serine protease domain requires the NS4A protein to anchor onto the cellular membrane of the host30. Other factors are also necessary for proper folding and activation of the serine protease domain. The functional NS3 protein is required to cleave the other nonstructural proteins in the ORF. The function of the C-Terminal RNA helicase is not exactly known. The general thought is that it is involved in the initiation of RNA synthesis on the genomic RNA of HCV. Not much research has been conducted into the function of NS4B, however some researchers proposed that it is a component of a membrane associated cytoplasmic HCV replication complex31. The specific function of NS5A remains rather obscure since many different host proteins interact

7 Figure 2. Schematic overview of the 3,000 amino acid polyprotein of HCV. On top the ORF and both the 5’and 3’UTR are visible. After translation the polyprotein of HCV is generated (middle). The cleavage sites as well as the different structural and nonstructural proteins are also depicted in the middle. The bottom illustration shows the function of the various proteins22.

with this viral protein32, 33. Within the host cell, NS5A is phosphorylated on multiple serine residues. Among the different cellular kinases responsible for hyperphosphorylation of NS5A are AKT and cAMP-dependent protein kinase A-α34. The functional and structural characteristics of the, up to now discovered, three functional domains of NS5A remain unclear, although they might influence the efficiency of RNA replication via NS5B35. NS5B is the most important gene involved in the viral replication cycle. It encodes the RNA-dependent RNA polymerase (RdRP). The protein synthesizes RNA from an RNA template. Experimental evidence indicates that this synthesis of RNA can occur de novo. This has been shown in vitro, and it is postulated that this also occurs in vitro36. Since NS5B is a polymerase it shows a high structural homology to other polymerases. It consists of a palm-like structure, where the palm contains the active site and the thumb and finger structure of the enzyme provide stabilization and modulation of contact with the RNA template37. The total enzyme contains a “tunnel-like” structure, due to interaction of the finger and thumb domain, where through the single stranded RNA template is brought to the active site on the palm structure of the enzyme. NTPs, building blocks necessary for transcription are brought to the active site of the enzyme via another “tunnel-like” structure. Final fine-tuning of the transcription process is regulated by a beta- hairpin loop that is present in the thumb structure of the enzyme38, 38.

8 2.4 VIRAL LIFE CYCLE

The viral life cycle of HCV consists of five major steps; viral entry, uncoating of the viral genome, translation of the viral proteins, genomic replication, and the assembly and release of the newly synthesized virions. In order for the virus to infect a host-cell it needs to attach itself to this cell. For the attachment at least two components are necessary and extracellular host-cell component and a receptor on the virus itself. For HCV various factors on the host cell seem to attribute to a successful viral entry. The first factor identified was CD81. The peculiar thing about CD81 is that it is present on a broad variety of host-cell types. It is not solely present on hepatic cell types. However, experiments have shown that HCV viral entry is reduced on liver cells that lack CD81, or on liver carcinoma cells with downregulated expression of CD81. The viral glycoprotein E2 is able to bind to CD8139. Before the binding of E2 to CD81 various attachment molecules on the cell surface such as glycosamineglycans (heparine sulphate)40, and LDLR (low density lipoprotein receptor)41 mediate this initial interaction. A recent study in showed that CD81 independent viral entry is also possible in vitro42. Besides CD81 other interactions are necessary for successful viral entry. Research has shown that the human scavenger receptor class B type I (SR-BI) might mediate the HCV entry process43. After the interactions of E2 with LDLR, SR-BI, and CD81, further steps need to be taken by the viral particle. Interaction between CD81 and tight junction protein claudin-1 (CLDN-1) is essential for viral entry44. Interactions between CD81 and CLDN-1 have also been observed, which suggests the formation of a co-receptor complex between CD81 and CLDN-145. A second tight junction protein has also been identified as a key player in HCV entry in vitro, Occludin (OCLN). OCLN can interact with E2, and suppression of OCLN with siRNA inhibits HCV entry in vitro46. Binding to the tight junction proteins induces receptor-mediated endocytosis. When the virus particle is inside the endosome it becomes acidified due to the influx of H+. This results in the fusion of the viral glycoproteins and the release of the nucelocapsid. The genomic RNA is then released and translation can begin47. It is possible that new membrane proteins will be identified essential in viral entry, but up till now most of these studies are conducted in vitro. The great challenge will be to proof the entry model in an in vivo experimental setting.

As mentioned in paragraph 2.3 the genomic RNA of HCV does not contain a 5’-7meG-cap, which is necessary for eukaryotic translation. This 5’-cap is necessary because the cap recruits eIF4F (eukaryotic initiation factor 4F) to the mRNA, which results in the binding of the 40S ribosomal subunit to the 5’-end of the mRNA and translation can commence. HCV uses the IRES domain in the 5’NTR to recruit the 40S ribosomal subunit. The IRES domain consists of several subunits that are each necessary for proper initiation of the translation process. These subunits are SLII, SLIII, and SLIV and the first codons of the ORF48 (see figure 3). Research has shown that the IRES bound to the 40S

9

Figure 3. A) Functional domains within the IRES structure in the 5’UTR of the HCV genome47 B) Schematic representation of the replication complex of HCV. After the recruitment of the various proteins on the membrane of the ER invagination occurs resulting in a protected environment for the RC47.

ribosomal subunit and eIF3 induces conformational changes in 40S resulting in the proper recruitment of 80S and the initiation of translation without the presence of a 5’-cap49.

After the translation of the 10 viral proteins from the ORF, viral RNA replication can commence. Since HCV is a positive strand RNA virus, replication occurs on intracellular membranes (as is the case for all +RNA viruses)50. The exact replication complex (RC) is not known but the overall thought is that it includes the viral genome, the non-structural proteins, and some cellular cofactors. The model that is proposed nowadays shows invagination of the RC on the endoplasmatic reticulum (ER). Nucleotides necessary for RNA replication can enter the RC, whereas proteases or nucleases cannot. This will protect the RC from degradation, which would occur if the RC was presented to the cytosol51. It seems likely that the positive RNA genome of the virus serves as the template to synthesize negative strand RNA. The two RNA strains remain paired with each other followed by copying of this complex. This results in the formation of multiple viral (+)RNA genomes52. As mentioned in paragraph 2.3 the workhorse of the HCV proteins is NS5B, the catalytic subunit of the RC with RNA dependent RNA polymerase activity. This enzyme lacks proof reading function, which leads to the formation of genomic copies with a high genetic variance53. According to recent studies many cellular cofactors regulate HCV RNA replication in vitro. Host proteins that interact with the non-structural proteins of HCV, but also components of the RNAi pathway and miRNAs seem to play a role in the efficiency of RNA replication54. Other factors that influence RNA replication efficiency in a negative way are nucleotide starvation, temperature shifts, and reactive oxidative species55.

10 For the release of viral particles, an important role is served for lipoproteins. Within released virus particles the human Apos (apolipoprotein) is found, and experiments with ApoE inhibition have showed a reduction in released HCV particles. Also, HCV assembly seems to occur on lipid droplets and lipids also seem to influence viral infectivity. According to recent studies, viral particles that are associated with cholesterol play a major role in viral infectivity. Research showed that viral particles enriched with cholesterol had a higher infectivity rate compared to the control viruses56. The blockade of fatty acid synthase (essential in the process of fatty acid metabolism) down regulates the expression of CLDN-1, mentioned above as crucial for viral entry, and thus reduces viral entry, and tempers the viral infection rate57. So far, there is no over-all model created for viral assembly and release. Also, the specific roles for the non-structural proteins are not completely understood. It has been shown that mutations in NS5A, NS2, and NS3 do not affect the RNA replication machinery, but do have a significant effect the production of new viruses. It seems logical to assume that there must be a specific role in the assembly and release process for these non-structural proteins. Their exact roles, however, are under heavy debate. More research needs to be conducted to unravel the specific functions of the proteins in the viral life cycle. The most up-to-date proposed model is shown in figure 447.

Figure 4. Model for HCV viral assembly and release as proposed by Shimotohno; core protein and non- structural protein complex (NLC) interact on lipid droplet (LD). Association with various lipoproteins determines the density of the virions, low density correlates with higher infectivity. Microsomal triacylglycerol transfer protein (MTP) and triglyceride are important in ApoB/ApoE formation47.

11 3. THE INNATE IMMUNE RESPONSE AGAINST HCV AND VIRAL IMMUNE EVASION

The immune response in humans against HCV consists of two parts; the non-specific, innate, and the specific, adaptive, component. Different parts of both components of the immune response will be discussed in more detail in the following paragraphs.

Several components of the non-specific immune system are thought to play an essential role in the clearance of HCV. Since this system is named non-specific, the following host defence mechanisms can be used against a broad spectrum of pathogens. A very important component is the interferon system. Already for a long time it is known that mice with target deletions in their type I interferon receptor are not able to fight viral infections, although their adaptive immune system is in perfect condition58. The type 1 interferon system makes use of two kinds of interferon, IFN-α and IFN-β. Production of IFN in a cell leads to a stimulus for neighbouring cells to produce antiviral proteins. This will eventually lead to the slow-down of viral production and thus infection, buying the adaptive immune system time to develop a specific immune response59. The production of IFN-β needs activation by viral proteins in the cell. Several sentinel molecules are active, and will be elaborated on in more detail in the following paragraphs. The viral evasion strategies will be discussed together with the different levels of the immune response in order to give a comprehensive overview of both the immune response and the viral immune evasion.

3.1 PATTERN RECOGNITON AND NS3/4A

One of the most important aspects of the non-specific immune system is that it makes use of pattern recognition. Several pathogen-associated molecular patterns (PAMPs) are recognized by specific pattern recognition receptors (PRR). Examples of HCV PAMPs are the single stranded RNA, and double stranded RNA. The PRRs involved in HCV recognition are Toll-like receptors, NOD-like receptors and RIG-like helicases60. The TLRs are a family of 11 receptors in human. The major part of these receptors is present at the plasma membrane of the cell, to detect the extracellular PAMPs. The TLR signalling cascade involves many adapter proteins, Myd88, TRIF, TRAM, and TIRAP, eventually leading to the activation of transcription factors and the production of IFN-β. In humans, TLR3, TLR7, and TLR9 are present intracellulary, for example on endosomes, to scan for PAMPs. dsRNA is recognized by TLR3, ssRNA by TLR7, and the non-methylated CpG motifs by TLR961. Recently, a TLR3 independent dsRNA recognition system has been postulated. The two helicases, retinoic acid inducible gene-I (RIG-1) and melanoma differentiation associated gene 5 (MDA5), have two CARD signalling domains that move into there activated state upon the binding of dsRNA to the helicase domain62. Both TLR3 and RIG1 are thought to be most important in the recognition of

12 intracellular HCV. The signalling cascade followed by endosomal dsRNA recognition by TLR3 starts with the recruitment of TRIF. Autophosphorylation takes place on both TRIF proteins, leading to the recruitment of TBK-1 and IKKε, both kinases that are thought to be essential in the phosphorylation, and thus activation of interferon regulatory factor 3 (IRF-3), which translocates to the nucleus upon phosphorylation and acts as a transcription factor for IFN-β63. Another transcription enhancer that becomes activated via TLR3 is NFκB. Activation of NFκB follows after the recruitment of TRAF6 ( receptor (TNFR) associated factor 6)64. Also, an important role is played by receptor interacting factor RIP1, which in its turn, is modulated in activity by RIP365. RIG1, as mentioned above, acts completely separate from TLR3 on the same PAMP, dsRNA, however, RIG1 recognizes cytoplasmic dsRNA. When the helicase domain of RIG-I/MDA-5 binds dsRNA, the CARD domains can freely interact with the CARD domain adapter protein CARDIF. CARDIF needs to be attached to the outer membrane of the mitochondrion; otherwise it cannot recruit the proteins necessary for the downstream signalling cascade66. The exact signalling cascade and the proteins involved remain unknown. Both TRAF3/6 seem to play a role to activate IRF-3 and NFκB respectively. The last RIG-1/MDA5 activated pathway results in the activation and translocation of the AP1 complex, which consists of C-jun and ATF2. Together with NFκB this is also called the enhanceosome. This enhanceosome facilitates chromatin remodelling by HATs, leading to the uncovering of the promoter, and the production of IFN-β.

The non-structural protein NS3/4A of HCV is able to cleave CARDIF from the mitochondrial membrane; thereby inhibiting RIG1 mediated IFN-β synthesis. Secondly, the protein is also capable to cleave the TRIF adapter protein. Due to this cleavage TRIF is unable to recruit TBK-1, which in turn

Figure 5. INF-β production is mediated via two cannot independent pathways: TLR3 mediated (depicted on the left side), and RIG-1 mediated, depicted on the right side. Translocation of transcription factors IRF-3, ATF2/c-JUN leads to initiation of INF-β transcription. NS3/4A is capable to cleave TRIF, in the TLR3 mediated pathway, and CARDIF, in the RIG-1 mediated pathway, thereby inhibiting IFN-β synthesis61.

13 phosphorylate IRF-3, which as a consequence cannot migrate through the nuclear membrane to act as a transcription factor. Concluding, this protein is capable to shut down IFN-β synthesis, both RIG-1 and TLR3 mediated (see figure 5).

3.2 INTERFERON-Β SIGNALLING AND VIRAL EVASION

The newly synthesized IFN-β can act in a paracrine or autocrine manner. In both cases, IFN-β binds to a transmembrane type I interferon-receptor. Upon binding, a Jak-STAT pathway becomes activated. The two intracellular parts of the interferon-receptor are Janus kinases and phosphorylate each other. Due to this phosphorylation step STAT2 is recruited and associates with IRF-9. STAT1 is also recruited and due to the phosphorylation of both STAT1 and STAT2 the two proteins heterodimerize via SH2 binding domains, resulting in the formation of the interferon stimulated gene factor (ISGF)361. A negative regulator of this process is SOCS3 (suppressor of cytokine signalling 3)67. When ISGF-3 translocates to the nucleus it can bind to the IFN-stimulated response element (ISRE), initiating the transcription of interferon stimulated genes (ISGs). One of these genes codes for the transcription factor IRF-7, which acts as a positive feedback system, due to the enhancing, together with IRF-3 (mentioned in the previous paragraph), of IFN-α transcription68.

Besides interfering with the signalling cascades that result in IFN-β production, HCV is also capable to interfere with the IFN-β signalling, upon IFN-β receptor stimulation. The first anti-IFN signalling action is undertaken by the core protein of HCV. This protein is thought to interfere directly at the point of Jak-STAT pathway. Whether it does so via direct degradation of STAT1, up-regulation of PIAS (protein inhibitor of activated STAT), or up-regulation of SOCS3 (suppressor of cytokine signalling), remains unknown. Experimental evidence has been gathered in favour of the various mechanism described above61. Recently, also the mechanism of epigenetic silencing by HCV, discovered with the usage of (partially) IFN-resistant infected cells, has been proposed as a tool used by HCV to become “IFN-signalling resistant”69.

3.3 INTERFERON STIMULATION GENES AND VIRAL EVASION

Other ISGs are ADAR (RNA-specific adenosine deaminase) 1, protein kinase R (PKR), 2’- 5’oligoadenylate synthetases (2-5 OAS), and P5670. Three of these proteins, PKR, ADAR 1, and OAS, are always expressed at a certain base level in every cell. Interferon signalling enhances the production and activity of these proteins. PKR phosphorylates the alpha subunit of eIF2. Since this factor is essential to initiate translation, this will result in the blockade of both viral and cellular translation71. ADAR 1 is an enzyme that interferes with the synthesis of the building blocks of RNA. It deaminates adenosine on target dsRNAs resulting in less stable inosine – uracil base pairs compared

14 to adenosine – uracil base pairs. Due to this, mutations accumulate in the viral genome leading to the production of non-functional viral particles59. 2-5 OAS activates RNaseL via the formation of short oligoadenylates. This RNaseL degrades viral, but also cellular, RNA. This inhibits translation of the viral proteins72. The last antiviral protein is P56. This protein also interferes with the translation initiation complex by binding to eIF3. As mentioned in chapter 2, this is an essential protein to initiate translation. This process will thus be blocked resulting in the inhibition of viral RNA translation. As explained above, the interferon-stimulated genes (ISG) are important anti-viral defence mechanisms. Several proteins of HCV are, besides being essential for replication and viral assembly, capable to interfere with these protein functions. The first example is the ISG PKR. Both the structural protein E2 as well as the non-structural protein NS5A can inhibit the function of PKR. The structural protein E2 exists in the cytoplasm of the host cell as unglycosylated. This form of E2 shows a high homology with the PKR-eIF2α binding domain. This domain is the site where PKR binds and phosphorylates eIF2, resulting in a shutdown of the transcription machinery. Due to the competition of unglycosylated E2 for this binding domain with eIF2, less eIF2 becomes phosphorylated, and transcription can still take place. The exact binding sequence differs between the various genotypes of HCV, making E2 able to block a broad variety PKR- eIF2α binding sites73. NS5A is believed to bind with its IFN sensitivity-determining region (ISDR) to the central part of PKR, thereby inactivating the functional protein. However, the findings on these interactions remain highly speculative since different models in different studies generate different experimental results61. NS5A also interferes with another ISG, namely the 2-5 OAS. It can bind to 2-5 OAS, with 2 different amino acid sequences on OAS involved. This binding is, however, ISDR independent, giving the virus another option to interfere with IFN-stimulated antiviral proteins74. Besides interfering directly with the protein, it is shown that more virulent genomic strains (HCV genotype 1 vs 2 or 3) have less RNase L cleavage sites, making them less susceptible for RNAase L / 2-5 OAS activity. The protein P56 also seems to be affected to a certain degree by HCV. NS3/4A, and also other (non) structural proteins, interfere with the signalling from IFN to the P56 promoter. The different IFN-signalling pathways and the immune evasion proteins of HCV are illustrated in figure 6 (page 16).

3.4 NATURAL KILLER CELLS AND HCV

Another component of the innate immune system is the natural killer cell (NK cell). NK cells are very important in the primary immune response against viral infections. Cells that do not express MHC class I on their cell surface are recognized by NK cells, and immediate cell lysis induced by the NK cell will follow. Besides MHC class I, many other inhibiting and activating receptors that mediate the binding between NK cells and MHC class I have been discovered; making this system more

15 sophisticated than it was thought to be. In HCV infection NK cells are thought to be necessary for optimal priming and cytolytic function of virus specific T cells due to their production of IFN-γ. NK cells can also induce apoptosis of infected hepatocytes46.

Since NK cells are thought to play an essential role in the primary anti-viral response, it seems logical to conclude that HCV must possess characteristics to evade the function of NK cells75. The binding of viral E2 to CD81 can inhibit the NK cell activity. This interaction inhibits NK cell activation but, in time, will activate B and T cells. This could be regarded as a strategy to establish a high viral load and viral turnover before the specific immune response will be initiated76. NK cells also have an immune-regulatory role, via different cytokines, and interfering with NK cells could lead to a delayed specific immune response. NK cells are primary activated when MHC class I presentation is abundant. The core protein of HCV can upregulate MHC I expression in hepatocytes, which results in inactive NK cells. These immune evasion strategies aimed specifically on NK cells have only recently been discovered and need more extensive research in order to be understood completely77.

Figure 6. IFN-β binds to the interferon receptor and signaling via Jak-STAT commences. The downstream signaling cascades are shown, eventually leading to the production of ISGs. The different proteins of HCV that interfere with these pathways at multiple levels, and the proteins that directly affect the function of the ISGs are also shown61.

16 4. THE ADAPTIVE IMMUNE RESPONSE AGAINST HCV AND VIRAL EVASION

When the innate immune response against HCV is unable to completely clear the virus, via the measurements described above, the slower but more specific adaptive immune response starts with the generation of a specific immune response against HCV. In the following paragraphs both the cellular mechanisms involved in this specific immune response, as well as the counter measurements taken by the virus to evade this specific immune response, will be described in more detail.

4.1 THE HUMORAL IMMUNE RESPONSE AND VIRAL EVASION

The specific immune response against HCV consists of both a humoral component and a cellular component. The humoral component of the specific immune response makes use of viral specific antibodies produced by B-cells. These antibodies bind to the antigens upon the virus, which marks them for destruction by, for example, macrophages. During acute HCV infection and chronic HCV infection neutralizing antibody levels can be detected. These antibodies often attack the envelope glycoproteins E1 and E2, and there is one region that seems especially important, the hyper-variable region 1 (HVR-1). The epitopes around and in this region have a very important function in various steps of the viral entry process, such as the binding of the virus to CD81 and membrane fusion. Besides the high mutation rate of the virus, which I will discuss in more detail, other factors negatively influence the effect of neutralizing antibodies70. The glycoproteins of HCV interact with host high-density lipoproteins and the HCV receptor scavenger receptor B1. These interactions can lead to a conformational change, making it impossible for the antibody to bind the specific epitope. Also steric hindrance could negatively influence the binding capacity for the antibody on the epitope78. There are also specific glycans present on the envelope protein E2 that, beside their function in cell entry, also function as protection for the virus against antibodies.

4.2 THE CELLULAR IMMUNE RESPONSE AND DENDRITIC CELL DYSFUNCTION

The cellular immune response is activated mainly by Type 1 helper cells (Th1). They play a crucial role in starting various cellular immune responses, such as cytotoxic T lymphocytes (CTL), and macrophages. Upon cellular uptake of HCV, antigens of HCV are presented in the liver. These antigens are taken up myeloid dendritic cells (DCs). After the migration of the DCs to a draining lymph node, they mature, and express HLA and various co-stimulatory molecules, such as CD80 and CD86. Via these stimulating molecules, naïve helper T cells become activated (binding of CD28 on Th1 cells to CD80 and CD86). Upon activation Th cells express the ligand CD40. The mature DC binds this ligand via the CD40 receptor that is present on the surface of the mature DC. Together with the expression of CD40 ligand various cytokines, such as tumour necrosis factor α (TNF- α). Sequentially,

17 DCs secrete IL-12, a cytokine that promotes the maturation of Th1 cells from the naïve Th cells. When Th1 cells are fully maturated they express IL-2 and IFN-γ. These cytokines induce the activation and the proliferation of other cells, such as CTLs. The naïve CTL expresses CD28 on its surface and via the CD28 - CD80/CD86 interaction and MHC class I on the DC the viral specific antigen is presented. The HCV specific CTL migrates to the liver recognizes the presented antigens via MHC class 1 on infected hepatocytes, and mediates specific killing of these cells.

Besides the activation of CTLs via dendritic cells and Th cells, CTLs can also become activated in a direct manner. In the process of direct activation of CTLs, cross presentation plays an essential role. Dendritic cells can take up proteins via endocytic mechanisms. Two different mechanisms lead to presentation of the (partial) protein on MHC class I. Either via the transfer of the protein from the phagosome to the cytosol, followed by hydrolization, transport to the associated MHC 1 molecule in the endoplasmatic reticulum or phagosome, and extracellular presentation to patrolling CTLS, or via cleavage of the protein by endosomal proteases, followed by the binding to MHC 1 within the endosome, and extracellular presentation79. Up to now, twenty different CTL epitopes on the HCV polyprotein have been identified80. The virus-specific CTL mediated killing is thought to play an essential role in viral eradication; this is due to the fact that CTLs inhibit viral replication81. As stated above, CTL mediated killing involves the recognition of peptides (8 – 11 amino acids in length) presented by MHC class 1 molecules. Upon recognition of the specific peptide, the CTL releases perforin. Perforin is capable to form a pore in the cellular membrane of the infected cell. Through this pore, granzyme can migrate into the cytoplasm, and induce apoptosis. Besides the induction of apoptosis via a granzyme mediated pathway, CTL can also directly induce apoptosis via the expression of Fas-ligand (Fas-L) or TNF-α. Studies have shown that the expression of Fas is upregulated in HCV infected hepatocytes. Besides the upregulation of Fas, also CTLs that have infiltrated the liver express FasL. The presence of both Fas and FasL in the liver has resulted in the conclusion that this system is important to induce apoptosis in HCV infected liver cells. The binding of Fas to FasL results in the formation of the death-inducing signalling complex. This complex activates caspase-8. Active caspase-8 can start 2 independent signalling pathways. In the first pathway the molecule Bid is cleaved, and cytochrome C is released from the mitochondria into the cytoplasm. The presence of cytochrome C results in the cleavage of various caspase molecules, the final one being caspase-7, resulting in cell death. The second pathway is the presence of active caspase-8 itself in the cytoplasm, resulting in the cleavage of the previous mentioned caspase molecules and the induction of apoptosis82. The most important aspect of the apoptosis induction via Fas-L or TNF-α is that it happens without cell-to-cell contact81. An overview of the different adaptive immune responses is shown in figure 7.

18 Figure 7. All components involved in the adaptive immune response are shown in this picture. Neutralizing antibodies can attack extracellular viral particles. They are also capable to block viral entry. Th1 helper cells (CD4+) activate cytotoxic T lymphocytes (CD8+) via DCs and antigen presentation. CTLs use Perforin, Fas/FasL, or TNF-α to induce apoptosis. The secretion of IFN-γ results in non-cytotoxic clearance of the virus86.

The exact mechanisms underlying the adaptive immune evasion capacities of HCV are poorly understood. This is due to the fact that the chronic HCV infection characteristics differ from one patient to another. If we look chronological to the events that lead to a specific immune response it al starts with the antigen-presenting DCs. Impairments in the interactions between the DCs and T cells can lead to a non-functional selective immune response. Many in vitro studies have been conducted, but gave conflicting results52. One of the studies states that HCV alters dendritic cell function. DCs that were cultured in presence of HCV peptides were less potent in stimulating an antigen specific immune response. This results in less functional Th1 cells and CTLs. A possible mechanism underlying this phenomenon could be the alteration of the cytokine secretion profile83. The alteration in cytokine secretion can have massive effects on the immune system, since cytokines are the primary signalling molecules involved in maturation and proliferation of all the cells of the immune system. The proteins of HCV also seem to play a crucial role in interfering with DC function. In a study by Krishnadas et al. researchers have shown that the core protein, NS3, NS4, NS5, and the fused polyprotein that consists of core, NS3, and NS4, can impair DC function on multiple levels. The proteins seem capable to regulate the expression of co-stimulatory and antigen presentation molecules, reduce IL-12 secretion, induce the expression of FasL to mediate DC apoptosis, interfere with allo-stimulatory capacity, inhibit toll-like receptor signaling and inhibit nuclear translocation of NFkappaB. Both the HCV core protein and E2 can inhibit DC maturation84. The recognition of the core proteins and NS3 seems to involve TLR2. Research has shown that DCs that are activated by ligands of TLRs derived from viruses have reduced cross-presentation ability, which plays an important role in the initiation of an immune response. This ability to evade DC signaling is seen in other viruses and

19 bacteria. Evidence that supports this hypothesis is found recently. From studies it was concluded that TLR2 expression on DCs is reduced in HCV patients and that TLR2 stimulated DCs are less potent in stimulating T cells to proliferate85. The activation of NK cells by DCs involves both IFN-α signaling and the expression of major histocompatibility complex class-I related Chain A and B (MICA/B). Both signals seem to be reduced in patients suffering from chronic HCV, leading to the activation of less NK cells. The structural protein E2 shows high affinity to bind two DC specific receptors, DC-SIGN and L-SIGN. The binding of E2 to these receptors could interfere with intracellular pathways in DCs81. It is still under heavy debate whether the findings described above are essential for HCV persistence or that they are consequences of an already persisted HCV infection. Another discrepancy lies in the fact that the above findings are not present in every patient. For example, it has been reported that DCs that do express the core protein and NS3 showed normal cytokine production and that they were perfectly capable to stimulate T cells. Further research needs to be conducted to clarify the exact mechanisms underlying disrupted DC functioning and impaired T cell or NK cell activation by DCs81.

4.3 T-CELL FAILURE

As is the case with defective antigen presentation, or DC signaling, the mechanisms involved in T-cell failure are poorly understood. One of the factors that could lead to impaired T cell function is the chronic antigenic stimulation. This phenomenon can be observed in other chronic viral infections. The constant presentation of antigen results in chronic T cell activation, but also loss of function. This often starts with less IL-2 production, followed by the decline of cytotoxicity, TNF-α, and IFN-γ production respectively. In HCV specific CD4+ and CD8+ T cells T cell failure is expressed by less IFN-γ production. It is also observed that less IL-2 activated T cell mediated killing precedes the development of chronic hepatitis C77. When a T cell becomes dysfunctional it starts to express the protein-programmed death 1 (PD-1). This protein needs its ligand PD-L1, and upon binding, this protein will activate an intracellular pathway that will result in T cell apoptosis22. Due to the fact that these processes affect the CD4+ T helper cells, they will also have an effect on the proliferation and maturation of cytotoxic CD8+ T cells. Recently a human cytokine has also been found as a mediator of HCV-specific T cell failure. Studies have shown that IL-10 levels are often elevated in chronic HCV infection. IL-10 has multiple effects on the immune response. It inhibits IFN-α production, and it can downregulate the effector T cell responses86.

20 4.4 REGULATOR T CELLS

Regulator T cells (Tregs) are a subpopulation of T cells and they are often elevated during chronic HCV infection. They are identified by the constant expression of factor forkhead box P3 (FoxP3), CD25+ and CD24+, and glucocorticoid-induced TNFR-α family-related protein (GITR). These cells can inhibit both proliferation and/or cytokine production of responding effector T cells (Teffs). Via inhibiting signals, Tregs are thought to play an important role in balancing the immune response to infection, and act as an extra mechanism to evoke auto immunity and an excessive immune response, which could lead to extensive tissue damage87. The exact mechanism of Teffs regulation is not known, but it probably involves, cytokine signaling via IL-10 or TGF-β, or by acting as an IL-2 sink, depleting IL-2 for Teffs88. The inhibitory signals can be delivered to the target cell via direct cell-to-cell contact. This pathway involves the cellular proteins CTLA-4, GITR, and PD-1. Signaling over distance can occur via TGF-β or IL-10. Tregs express the T cell receptor complex, as all other T cells, and Tregs are thought to be antigen specific during HCV infection. Research into protein derived peptides from HCV and the antigen specificity of Tregs showed that only a few peptides, with an overlapping region, were able to stimulate Tregs. This indicates that there could be a dominant region on the core protein of HCV that is capable to activate Tregs, and thus suppress the immune response89. Not much research has been conducted into the co-stimulatory signaling molecules but the PD-1 system may be for a part responsible for the inhibitory function of Tregs during chronic HCV infection88, 90.

The mechanism behind the generation of Tregs is currently thought to be primarily a DC- dependent mechanism. It is known for a long time that DCs are capable to trigger the proliferation of Tregs. Different factors influence the proliferation of Tregs, IL-10 production, an immature DC phenotype and less expression of co-stimulatory DC molecules (such as CD80), to name a few. This information in combination with the fact that HCV interferes with the functionality of DCs on multiple levels, can lead to the conclusion that HCV suppresses the CD4+ and CD8+ mediated immune response via a DC dependent upregulation of Tregs. Whether there is a DC independent pathway to upregulate Tregs remains to be discovered.

4.5 T CELL HOMING CAPACTITY

An experiment with HCV infected chimpanzees conducted by Thimme et al. showed that chimps that progressed to the state of viral persistence, without any temporary viral control, did not have any virus specific CD8+ T cell response. The researchers hypothesized that this could be attributed to the fact that the T cells lacked homing capacity to the site of infection, in this case, the liver91. Again, the results in studies that compared the presence of virus specific CD8+ T cells in the liver and the blood were conflicting, studies in patients showed that sometimes the CD8+ T cell levels in the liver where

21 even higher that in controls92. It could still be possible that there is a lack of T cell homing capacity in certain patients, but up till now, it does not seem likely that HCV makes use of an immune evasive strategy that focuses on T cell homing in general86.

4.6 IMMUNE EVASION THROUGH VIRAL MUTATIONS

One of the most sophisticated immune evasion strategies is that of viral mutation. As described in the viral life cycle chapter, HCV is a RNA virus that makes use of an RNA-dependent RNA polymerase without a proofreading function. This in combination with the extremely high replication rate of the virus, 1012 virions are created each day, gives rise to a quasi species of HCV in every single patient. When this extreme high replication rate was first discovered, it soon led to the idea that this replication rate could facilitate viral persistence due to the escape from cellular immunity. Once the immune system has generated antibodies against a specific viral antigen, for example, the high mutation rate in combination with the selective pressure will lead to Darwinian selection. If a mutation occurs that is beneficial to the virus, it will adopt this mutation in newly synthesized virions, thereby evading the antibody. Two kinds of mutation occur during viral replication. Silent or synonymous mutations are mutations that have no impact on amino acid sequence of the viral genome. Non-synonymous mutations can lead to changes in amino acid sequence, can change the tertiary structure of a protein, and can be lethal for the virus93. The virus cannot mutate endlessly, and indeed, some parts of the genome are highly conserved. The most conserved structure is the non-coding 5’ end of the viral genome, which shows over 90% homology with other viral strains94. The most variable region of the genome is the part that encodes the envelope proteins E1 and E2. Within these proteins lay hyper variable regions, HVR 1 and HVR2, regions that are very mutation prone95. The highest mutation rate of the virus occurs shortly during after infection, during the acute phase of infection. When the infection continues, the mutation rate declines. The high mutation rate during the acute phase of infection is mainly driven by the selective pressure mentioned above. Besides antibodies also T cells exert this selective pressure. It is essential for HCV to constantly mutate its genome in order to diverse the cellular and humoral immune response. The altered peptide ligands that arise may even downregulate the T-cell response against the wild-type peptides96. Just as important is the fact that these altered peptides cannot induce successful T cell responses. The virus is in constant search to optimize its genome, without killing itself, and with a mean frequency of 1.4 x 103 to 1.9 x 103 substitutions per nucleotide per year97, it makes use of very capable system to hide itself from the human immune system.

22 5. THERPAPIES AND VACCINATION

In the following chapter the current used therapies that are in use after HCV infection will be discussed. I will also elaborate on potential vaccine development. Besides that I will also highlight the difficulties faced when using therapies, mutations of the virus that will interfere with a successful therapeutic outcome.

5.1 CURRENT AND FUTURE THERAPIES

When faced with an acute or chronic HCV infection, patients often are not able to generate a successful immune response. I have discussed that multiple causes are responsible for this lack of immune response. Therapies used nowadays focus on enhancing the immune response, by inducing a boost in T cell responses. Also, the enhancement of interferon levels could make cells more susceptible for interferon stimuli and thus facilitate viral clearance in the body (the effects of interferon are described in chapter 3). The current treatment for chronic hepatitis C infection involves the usage of peginterferon alfa-2a/-b with ribavirin for 48 weeks, or 24 weeks when the patient is infected with HCV genotype 1 or 2/3. Peginterferon α is stabilized version of interferon α with a longer half-life time. Ribavirin is a purine-analog with many different antiviral effects98. The exact mechanisms involved in the treatment of chronic HCV with the peginterferon–ribavirin combination is not yet fully understood, which leads to several disadvantages. The main disadvantage of this therapeutic approach is that severe side effects have been recorded in patients. This is a great motivation to search for new, more specific, and more potent anti-viral therapies99. At the moment the only standardized interferon therapy against HCV is treatment with pegylated interferon α. Researchers are looking for novel interferons that are either more specific, or are more convenient in dosage adjustments. A potent new candidate for future therapeutic usage is the interferon Albinterferon. This combination of interferon with albumin has a longer half-life than the current used interferon. This new interferon is now in clinical trials and shows the same virologic response, with a higher patient tolerance. Another new interferon, which is currently tested, is Peginterferon-λ. This type III interferon binds to a more specific receptor than the Peginterferon α. In phase I trials it showed less hematologic side effects compared to peginterfron α100, 101.

The main disadvantage of the usage of interferon type medication is that it is not very virus specific. Research has been conducted into HCV specific medication, mainly focussing on components that target the viral replication cycle. These components can be divided into IRES inhibitors, protease inhibitors, transcription inhibitors, and virus assembly inhibitors. The first subcategory has not led to much success99. The IRES inhibitors are designed to block the initiation of

23 viral translation. IRES inhibitors are anti-sense oligonucleotides, siRNA, or ribozymes. In studies the oligonucleotides seemed not very promising, due to toxic side effects as well as poor antiviral activity102, 103. More research needs to be conducted into IRES inhibitors before they can be tested in a therapeutic setting. The most potent HCV specific compound that interferes with the viral replication cycle of HCV is the protease inhibitor that inhibits the NS3-4A protease. As was described in chapter 2, his protein is essential for the virus since it cuts the polypeptide that contains the six non-structural proteins. Various inhibitors have been, or are being, tested in clinical trials, but with mixed results. Telaprevir showed a reduction of HCV RNA in patients that did not respond to the standard therapy. However several mutations in the viral protease were found that diminish the therapeutic effect of Telaprevir104, 105. Boceprecir is another NS3-4A serine protease inhibitor. It binds to the active site of the protease, interfering with its function. Also in these clinical trials mutations within the viral genome occurred. Also, the mutations found in the genome of HCV after patient treatment with Telaprevir overlap with those found in Borceprecir. It does not seem likely that a combination of these two compounds will eventually lead to a boost in antiviral activity106, 107.

Another class of viral specific therapeutic compounds are the polymerase inhibitors. These inhibitors can further be divided in two subclasses: nucleoside and non-nucleoside polymerase inhibitors. Polymerase inhibitors are already being used clinically in other viral infections such as HIV and . The nucleoside inhibitors are nucleotide analogous that are capable to bind to the active site of the polymerase after the incorporation in the newly synthesized RNA chain108. The non- nucleoside inhibitors also bind to specific sites of the enzyme, forcing it to go a conformational change. Due to this change the enzyme cannot function properly. The combination of different non- nucleoside inhibitors as a therapeutic measure seems very promising, since different non-nucleoside inhibitors bind to various sites of the active enzyme, making it more difficult for the virus to escape these interfering molecules by mutagenic variation. This does not account for the nucleoside inhibitors since all different compounds bind to the active site after the incorporation in the newly synthesized RNA. Several (non-) nucleoside protease inhibitors are being tested in clinical trials but with mixed results, due to genetic variation99.

Another class of HCV drugs are targeted at host cell factors. Drugs that target the cholesterol metabolism, since the viral life cycle of HCV is highly integrated with this mechanism, but also drugs that target the host compounds involved in viral entry are currently being designed56, 109. The key components in viral entrance, as stated in chapter two, are CD81 and the scavenger receptor class B type 1. Antibodies against these proteins are tested in vitro but much research needs to be conducted before these components can move into clinical trials. Several lipid metabolism inhibitors

24 are tested in vitro and in vivo. The main disadvantage with these kinds of compounds is that it is very difficult to incorporate specificity for virus-infected cells99.

The understanding of the viral life cycle has led to many different opportunities for viral specific compounds to battle chronic HCV infection. As with HIV, patient designed therapy (HAART) could provide the solution. So far the combination of several compounds (protease inhibitor together with a polymerase inhibitor) has shown promising results in clinical trials. However, the high viral replication rate in combination with the lack of proof reading on the RNA polymerase leads to the accumulation of many different mutations that leads to drug resistance. Besides the fact that HCV needs to evade the immune system in order to establish a persistent infection, he virus also has to battle our therapeutic interventions. So far, the virus does an extreme good job in evading these interventions, making more research into various compounds extremely important to tackle the devastating effects of HCV infection98.

5.2 VACCINE DEVELOPMENT

Up to now the development of against HCV has not been very successful. Until recently, humanized mouse models have not been available for testing. During chronic HCV infection the major problem is the presence of impaired T cells. The therapeutic vaccine should therefore introduce a broad CD4+ T cell response that can in turn activate cytotoxic T lymphocytes. The different therapeutic vaccines that are currently being tested are epitope vaccines, vector vaccines, recombinant protein vaccines, and DNA vaccines21. The epitope vaccines that are being tested in clinical trials contain epitopes of the core protein, NS3 and NS4 proteins. Together with different adjuvants these vaccines are potent to induce a Th1 and cytotoxic T cell response, but to a limited level110. Modified Virus of Ankara (MVA) is an attenuated poxvirus strain that is being used as a vector vaccine for HCV111. The HCV vectors introduced in this vector vaccine are NS3, NS4, and NS5B. These vectors have shown to induce IFN-γ-secreting CD4+ T cells and viral specific CD8+ T cells. In the first this vaccine lowered viral load in 6 out of 15 patients. The vaccine was well tolerated among the 15 candidates. A phase II clinical trial is currently being designed to treat patients with a combination of this vaccine together with Peginterferon and ribavirin112. The recombinant protein vaccines are tested with E1/E2 glycoproteins, NS3, and core protein that need an adjuvant to boost the T cell response21. The last group of vaccines is the DNA vaccines. One DNA vaccine has been designed around the conserved regions of NS3 and NS4A. The CD8+ T cells primed after vaccination showed eliminating capacity to NS3 and NS4A expression in mice. In patients, this DNA vaccine in combination with IFN α based therapy showed promising results113.

25 All these vaccines are therapeutic, and based upon the findings in both mice, and phase I clinical trial studies. Results seem to be interesting, but further research is eminent. The development of therapeutic vaccines in combination with the current and future treatment options described in the previous paragraph can contribute to the pallet of available therapies that can be used to cure patients that suffer from chronic HCV infection. However, these vaccines face the same problems as the therapies that are currently used, viral escape. All vaccines are targeted on specific viral epitopes. And although most of these epitopes are highly conserved, selective pressure, and the accumulation of evasion specific mutations, remains an enormous hurdle for therapeutic intervention.

26 6. REFERENCES

1. Choo, Q. L. et al. Isolation of a cDNA clone derived from a blood-borne non-A, non-B viral hepatitis genome. Science 244, 359-362 (1989).

2. Feinstone, S. M., Kapikian, A. Z., Purcell, R. H., Alter, H. J. & Holland, P. V. Transfusion-associated hepatitis not due to viral hepatitis type A or B. N. Engl. J. Med. 292, 767-770 (1975).

3. Shepard, C. W., Finelli, L. & Alter, M. J. Global epidemiology of hepatitis C virus infection. Lancet Infect. Dis. 5, 558-567 (2005).

4. Armstrong, G. L., Alter, M. J., McQuillan, G. M. & Margolis, H. S. The past incidence of hepatitis C virus infection: implications for the future burden of chronic liver disease in the United States. Hepatology 31, 777- 782 (2000).

5. Law, M. G. et al. Modelling hepatitis C virus incidence, prevalence and long-term sequelae in Australia, 2001. Int. J. Epidemiol. 32, 717-724 (2003).

6. Wasley, A. & Alter, M. J. Epidemiology of hepatitis C: geographic differences and temporal trends. Semin. Liver Dis. 20, 1-16 (2000).

7. Alter, H. Discovery of non-A, non-B hepatitis and identification of its etiology. Am. J. Med. 107, 16S-20S (1999).

8. Thorpe, L. E. et al. Risk of hepatitis C virus infection among young adult injection drug users who share injection equipment. Am. J. Epidemiol. 155, 645-653 (2002).

9. Hauri, A. M., Armstrong, G. L. & Hutin, Y. J. The global burden of disease attributable to contaminated injections given in health care settings. Int. J. STD AIDS 15, 7-16 (2004).

10. Singh, S., Dwivedi, S. N., Sood, R. & Wali, J. P. Hepatitis B, C and human immunodeficiency virus infections in multiply-injected kala-azar patients in Delhi. Scand. J. Infect. Dis. 32, 3-6 (2000).

11. Seeff, L. B. Natural history of chronic hepatitis C. Hepatology 36, S35-46 (2002).

12. Terrault, N. A. Sexual activity as a risk factor for hepatitis C. Hepatology 36, S99-105 (2002).

13. van de Laar, T. J., Matthews, G. V., Prins, M. & Danta, M. Acute hepatitis C in HIV-infected men who have sex with men: an emerging sexually transmitted infection. AIDS 24, 1799-1812 (2010).

14. Seeff, L. B., Wright, E. C., Zimmerman, H. J. & McCollum, R. W. Va cooperative study of post-transfusion hepatitis, 1969-1974: incidence and characteristics of hepatitis and responsible risk factors. Am. J. Med. Sci. 270, 355-362 (1975).

15. Negro, F. Hepatitis C virus-induced steatosis: an overview. Dig. Dis. 28, 294-299 (2010).

16. Poynard, T. & Afdhal, N. H. Perspectives on fibrosis progression in hepatitis C: an a la carte approach to risk factors and staging of fibrosis. Antivir Ther. 15, 281-291 (2010).

17. Sherman, M. Epidemiology of hepatocellular carcinoma. Oncology 78 Suppl 1, 7-10 (2010).

18. Beasley, R. P., Hwang, L. Y., Lin, C. C. & Chien, C. S. Hepatocellular carcinoma and hepatitis B virus. A prospective study of 22 707 men in Taiwan. Lancet 2, 1129-1133 (1981).

19. Torbenson, M., Yeh, M. M. & Abraham, S. C. Bile duct dysplasia in the setting of chronic hepatitis C and alcohol cirrhosis. Am. J. Surg. Pathol. 31, 1410-1413 (2007).

27 20. Robertson, B. et al. Classification, nomenclature, and database development for hepatitis C virus (HCV) and related viruses: proposals for standardization. International Committee on Virus Taxonomy. Arch. Virol. 143, 2493-2503 (1998).

21. Yu, C. I. & Chiang, B. L. A new insight into hepatitis C vaccine development. J. Biomed. Biotechnol. 2010, 548280 (2010).

22. Rehermann, B. Hepatitis C virus versus innate and adaptive immune responses: a tale of coevolution and coexistence. J. Clin. Invest. 119, 1745-1754 (2009).

23. Otto, G. A. & Puglisi, J. D. The pathway of HCV IRES-mediated translation initiation. Cell 119, 369-380 (2004).

24. Lindenbach, B. D. & Rice, C. M. Unravelling hepatitis C virus replication from genome to function. Nature 436, 933-938 (2005).

25. Dubuisson, J. Hepatitis C virus proteins. World J. Gastroenterol. 13, 2406-2415 (2007).

26. Boulant, S., Vanbelle, C., Ebel, C., Penin, F. & Lavergne, J. P. Hepatitis C virus core protein is a dimeric alpha- helical protein exhibiting membrane protein features. J. Virol. 79, 11353-11365 (2005).

27. McLauchlan, J. Properties of the hepatitis C virus core protein: a structural protein that modulates cellular processes. J. Viral Hepat. 7, 2-14 (2000).

28. Griffin, S. D. et al. The p7 protein of hepatitis C virus forms an ion channel that is blocked by the antiviral drug, Amantadine. FEBS Lett. 535, 34-38 (2003).

29. Jirasko, V. et al. Structural and functional characterization of nonstructural protein 2 for its role in hepatitis C virus assembly. J. Biol. Chem. 283, 28546-28562 (2008).

30. Wolk, B. et al. Subcellular localization, stability, and trans-cleavage competence of the hepatitis C virus NS3- NS4A complex expressed in tetracycline-regulated cell lines. J. Virol. 74, 2293-2304 (2000).

31. Hugle, T. et al. The hepatitis C virus nonstructural protein 4B is an integral endoplasmic reticulum membrane protein. Virology 284, 70-81 (2001).

32. Macdonald, A. & Harris, M. Hepatitis C virus NS5A: tales of a promiscuous protein. J. Gen. Virol. 85, 2485- 2502 (2004).

33. Tellinghuisen, T. L. & Rice, C. M. Interaction between hepatitis C virus proteins and host cell factors. Curr. Opin. Microbiol. 5, 419-427 (2002).

34. Reed, K. E., Xu, J. & Rice, C. M. Phosphorylation of the hepatitis C virus NS5A protein in vitro and in vivo: properties of the NS5A-associated kinase. J. Virol. 71, 7187-7197 (1997).

35. Appel, N., Pietschmann, T. & Bartenschlager, R. Mutational analysis of hepatitis C virus nonstructural protein 5A: potential role of differential phosphorylation in RNA replication and identification of a genetically flexible domain. J. Virol. 79, 3187-3194 (2005).

36. Bartenschlager, R., Frese, M. & Pietschmann, T. Novel insights into hepatitis C virus replication and persistence. Adv. Virus Res. 63, 71-180 (2004).

37. Lesburg, C. A. et al. Crystal structure of the RNA-dependent RNA polymerase from hepatitis C virus reveals a fully encircled active site. Nat. Struct. Biol. 6, 937-943 (1999).

38. Appel, N., Schaller, T., Penin, F. & Bartenschlager, R. From structure to function: new insights into hepatitis C virus RNA replication. J. Biol. Chem. 281, 9833-9836 (2006).

39. Pileri, P. et al. Binding of hepatitis C virus to CD81. Science 282, 938-941 (1998).

28 40. Barth, H. et al. Cellular binding of hepatitis C virus envelope glycoprotein E2 requires cell surface heparan sulfate. J. Biol. Chem. 278, 41003-41012 (2003).

41. Molina, S. et al. The low-density lipoprotein receptor plays a role in the infection of primary human hepatocytes by hepatitis C virus. J. Hepatol. 46, 411-419 (2007).

42. Timpe, J. M. et al. Hepatitis C virus cell-cell transmission in hepatoma cells in the presence of neutralizing antibodies. Hepatology 47, 17-24 (2008).

43. Scarselli, E. et al. The human scavenger receptor class B type I is a novel candidate receptor for the hepatitis C virus. EMBO J. 21, 5017-5025 (2002).

44. Evans, M. J. et al. Claudin-1 is a hepatitis C virus co-receptor required for a late step in entry. Nature 446, 801-805 (2007).

45. Harris, H. J. et al. CD81 and claudin 1 coreceptor association: role in hepatitis C virus entry. J. Virol. 82, 5007-5020 (2008).

46. Liu, S. et al. Tight junction proteins claudin-1 and occludin control hepatitis C virus entry and are downregulated during infection to prevent superinfection. J. Virol. 83, 2011-2014 (2009).

47. Tang, H. & Grise, H. Cellular and molecular biology of HCV infection and hepatitis. Clin. Sci. (Lond) 117, 49- 65 (2009).

48. Lukavsky, P. J. Structure and function of HCV IRES domains. Virus Res. 139, 166-171 (2009).

49. Spahn, C. M. et al. Hepatitis C virus IRES RNA-induced changes in the conformation of the 40s ribosomal subunit. Science 291, 1959-1962 (2001).

50. Ahlquist, P., Noueiry, A. O., Lee, W. M., Kushner, D. B. & Dye, B. T. Host factors in positive-strand RNA virus genome replication. J. Virol. 77, 8181-8186 (2003).

51. Quinkert, D., Bartenschlager, R. & Lohmann, V. Quantitative analysis of the hepatitis C virus replication complex. J. Virol. 79, 13594-13605 (2005).

52. Sklan, E. H., Charuworn, P., Pang, P. S. & Glenn, J. S. Mechanisms of HCV survival in the host. Nat. Rev. Gastroenterol. Hepatol. 6, 217-227 (2009).

53. Martell, M. et al. Hepatitis C virus (HCV) circulates as a population of different but closely related genomes: quasispecies nature of HCV genome distribution. J. Virol. 66, 3225-3229 (1992).

54. Randall, G. et al. Cellular cofactors affecting hepatitis C virus infection and replication. Proc. Natl. Acad. Sci. U. S. A. 104, 12884-12889 (2007).

55. Choi, J. et al. Reactive oxygen species suppress hepatitis C virus RNA replication in human hepatoma cells. Hepatology 39, 81-89 (2004).

56. Aizaki, H. et al. Critical role of virion-associated cholesterol and sphingolipid in hepatitis C virus infection. J. Virol. 82, 5715-5724 (2008).

57. Yang, W. et al. Fatty acid synthase is up-regulated during hepatitis C virus infection and regulates hepatitis C virus entry and production. Hepatology 48, 1396-1403 (2008).

58. Weber, F., Kochs, G. & Haller, O. Inverse interference: how viruses fight the interferon system. Viral Immunol. 17, 498-515 (2004).

59. Samuel, C. E. Antiviral actions of interferons. Clin. Microbiol. Rev. 14, 778-809, table of contents (2001).

60. Saito, T. & Gale, M.,Jr. Principles of intracellular viral recognition. Curr. Opin. Immunol. 19, 17-23 (2007).

29 61. Bode, J. G., Brenndorfer, E. D. & Haussinger, D. Subversion of innate host antiviral strategies by the hepatitis C virus. Arch. Biochem. Biophys. 462, 254-265 (2007).

62. Yoneyama, M. et al. Shared and unique functions of the DExD/H-box helicases RIG-I, MDA5, and LGP2 in antiviral innate immunity. J. Immunol. 175, 2851-2858 (2005).

63. Fitzgerald, K. A. et al. IKKepsilon and TBK1 are essential components of the IRF3 signaling pathway. Nat. Immunol. 4, 491-496 (2003).

64. Sato, S. et al. Toll/IL-1 receptor domain-containing adaptor inducing IFN-beta (TRIF) associates with TNF receptor-associated factor 6 and TANK-binding kinase 1, and activates two distinct transcription factors, NF- kappa B and IFN-regulatory factor-3, in the Toll-like receptor signaling. J. Immunol. 171, 4304-4310 (2003).

65. Meylan, E. et al. RIP1 is an essential mediator of Toll-like receptor 3-induced NF-kappa B activation. Nat. Immunol. 5, 503-507 (2004).

66. Seth, R. B., Sun, L., Ea, C. K. & Chen, Z. J. Identification and characterization of MAVS, a mitochondrial antiviral signaling protein that activates NF-kappaB and IRF 3. Cell 122, 669-682 (2005).

67. Kubo, M., Hanada, T. & Yoshimura, A. Suppressors of cytokine signaling and immunity. Nat. Immunol. 4, 1169-1176 (2003).

68. Au, W. C., Moore, P. A., LaFleur, D. W., Tombal, B. & Pitha, P. M. Characterization of the interferon regulatory factor-7 and its potential role in the transcription activation of interferon A genes. J. Biol. Chem. 273, 29210-29217 (1998).

69. Naka, K. et al. Epigenetic silencing of interferon-inducible genes is implicated in interferon resistance of hepatitis C virus replicon-harboring cells. J. Hepatol. 44, 869-878 (2006).

70. Thimme, R., Lohmann, V. & Weber, F. A target on the move: innate and adaptive immune escape strategies of hepatitis C virus. Antiviral Res. 69, 129-141 (2006).

71. Williams, B. R. PKR; a sentinel kinase for cellular stress. Oncogene 18, 6112-6120 (1999).

72. Zhou, A. et al. Interferon action and apoptosis are defective in mice devoid of 2',5'-oligoadenylate- dependent RNase L. EMBO J. 16, 6355-6363 (1997).

73. Taylor, D. R., Shi, S. T., Romano, P. R., Barber, G. N. & Lai, M. M. Inhibition of the interferon-inducible protein kinase PKR by HCV E2 protein. Science 285, 107-110 (1999).

74. Taguchi, T. et al. Hepatitis C virus NS5A protein interacts with 2',5'-oligoadenylate synthetase and inhibits antiviral activity of IFN in an IFN sensitivity-determining region-independent manner. J. Gen. Virol. 85, 959-969 (2004).

75. Xu, X. N., Screaton, G. R. & McMichael, A. J. Virus infections: escape, resistance, and counterattack. Immunity 15, 867-870 (2001).

76. Crotta, S. et al. Inhibition of natural killer cells through engagement of CD81 by the major hepatitis C virus envelope protein. J. Exp. Med. 195, 35-41 (2002).

77. Golden-Mason, L. & Rosen, H. R. Natural killer cells: primary target for hepatitis C virus immune evasion strategies? Liver Transpl. 12, 363-372 (2006).

78. Dreux, M. et al. High density lipoprotein inhibits hepatitis C virus-neutralizing antibodies by stimulating cell entry via activation of the scavenger receptor BI. J. Biol. Chem. 281, 18285-18295 (2006).

79. Rock, K. L. & Shen, L. Cross-presentation: underlying mechanisms and role in immune surveillance. Immunol. Rev. 207, 166-183 (2005).

30 80. Hakamada, T. et al. Identification of novel hepatitis C virus-specific cytotoxic T lymphocyte epitopes by ELISpot assay using peptides with human leukocyte antigen-A*2402-binding motifs. J. Gen. Virol. 85, 1521-1531 (2004).

81. Hiroishi, K., Ito, T. & Imawari, M. Immune responses in hepatitis C virus infection and mechanisms of hepatitis C virus persistence. J. Gastroenterol. Hepatol. 23, 1473-1482 (2008).

82. Kanto, T. & Hayashi, N. Immunopathogenesis of hepatitis C virus infection: multifaceted strategies subverting innate and adaptive immunity. Intern. Med. 45, 183-191 (2006).

83. Della Bella, S. et al. Decrease and dysfunction of dendritic cells correlate with impaired hepatitis C virus- specific CD4+ T-cell proliferation in patients with hepatitis C virus infection. Immunology 121, 283-292 (2007).

84. Krishnadas, D. K., Ahn, J. S., Han, J., Kumar, R. & Agrawal, B. Immunomodulation by hepatitis C virus-derived proteins: targeting human dendritic cells by multiple mechanisms. Int. Immunol. 22, 491-502 (2010).

85. Yakushijin, T. et al. Reduced expression and functional impairment of Toll-like receptor 2 on dendritic cells in chronic hepatitis C virus infection. Hepatol. Res. 34, 156-162 (2006).

86. Thimme, R., Neumann-Haefelin, C., Boettler, T. & Blum, H. E. Adaptive immune responses to hepatitis C virus: from viral immunobiology to a vaccine. Biol. Chem. (2008).

87. Jager, A. & Kuchroo, V. K. Effector and regulatory T-cell subsets in autoimmunity and tissue inflammation. Scand. J. Immunol. 72, 173-184 (2010).

88. Radziewicz, H., Dunham, R. M. & Grakoui, A. PD-1 tempers Tregs in chronic HCV infection. J. Clin. Invest. 119, 450-453 (2009).

89. Dolganiuc, A. & Szabo, G. T cells with regulatory activity in hepatitis C virus infection: what we know and what we don't. J. Leukoc. Biol. 84, 614-622 (2008).

90. Urbani, S. et al. PD-1 expression in acute hepatitis C virus (HCV) infection is associated with HCV-specific CD8 exhaustion. J. Virol. 80, 11398-11403 (2006).

91. Thimme, R. et al. Determinants of viral clearance and persistence during acute hepatitis C virus infection. J. Exp. Med. 194, 1395-1406 (2001).

92. Penna, A. et al. Dysfunction and functional restoration of HCV-specific CD8 responses in chronic hepatitis C virus infection. Hepatology 45, 588-601 (2007).

93. Le Guillou-Guillemette, H. et al. Genetic diversity of the hepatitis C virus: impact and issues in the antiviral therapy. World J. Gastroenterol. 13, 2416-2426 (2007).

94. Bukh, J., Purcell, R. H. & Miller, R. H. Sequence analysis of the 5' noncoding region of hepatitis C virus. Proc. Natl. Acad. Sci. U. S. A. 89, 4942-4946 (1992).

95. Simmonds, P. et al. Identification of genotypes of hepatitis C virus by sequence comparisons in the core, E1 and NS-5 regions. J. Gen. Virol. 75 ( Pt 5), 1053-1061 (1994).

96. Chang, K. M. et al. Immunological significance of cytotoxic T lymphocyte epitope variants in patients chronically infected by the hepatitis C virus. J. Clin. Invest. 100, 2376-2385 (1997).

97. Ogata, N., Alter, H. J., Miller, R. H. & Purcell, R. H. Nucleotide sequence and mutation rate of the H strain of hepatitis C virus. Proc. Natl. Acad. Sci. U. S. A. 88, 3392-3396 (1991).

98. Lemon, S. M. et al. Development of novel therapies for hepatitis C. Antiviral Res. 86, 79-92 (2010).

99. Jang, J. Y. & Chung, R. T. New treatments for chronic hepatitis C. Korean J. Hepatol. 16, 263-277 (2010).

31 100. Zeuzem, S. et al. Albinterferon Alfa-2b was not inferior to pegylated interferon-alpha in a randomized trial of patients with chronic hepatitis C virus genotype 1. Gastroenterology 139, 1257-1266 (2010).

101. Nelson, D. R. et al. Albinterferon Alfa-2b was not inferior to pegylated interferon-alpha in a randomized trial of patients with chronic hepatitis C virus genotype 2 or 3. Gastroenterology 139, 1267-1276 (2010).

102. Fukushi, S. et al. The sequence element of the internal ribosome entry site and a 25-kilodalton cellular protein contribute to efficient internal initiation of translation of hepatitis C virus RNA. J. Virol. 71, 1662-1666 (1997).

103. McHutchison, J. G. et al. A phase I trial of an antisense inhibitor of hepatitis C virus (ISIS 14803), administered to chronic hepatitis C patients. J. Hepatol. 44, 88-96 (2006).

104. Hezode, C. et al. Telaprevir and peginterferon with or without ribavirin for chronic HCV infection. N. Engl. J. Med. 360, 1839-1850 (2009).

105. McHutchison, J. G. et al. Telaprevir with peginterferon and ribavirin for chronic HCV genotype 1 infection. N. Engl. J. Med. 360, 1827-1838 (2009).

106. Sarrazin, C. et al. SCH 503034, a novel hepatitis C virus protease inhibitor, plus pegylated interferon alpha- 2b for genotype 1 nonresponders. Gastroenterology 132, 1270-1278 (2007).

107. Kwo, P. Y. et al. Efficacy of boceprevir, an NS3 protease inhibitor, in combination with peginterferon alfa- 2b and ribavirin in treatment-naive patients with genotype 1 hepatitis C infection (SPRINT-1): an open-label, randomised, multicentre phase 2 trial. Lancet 376, 705-716 (2010).

108. Manns, M. P. et al. The way forward in HCV treatment--finding the right path. Nat. Rev. Drug Discov. 6, 991-1000 (2007).

109. Meuleman, P. et al. Anti-CD81 antibodies can prevent a hepatitis C virus infection in vivo. Hepatology 48, 1761-1768 (2008).

110. Klade, C. S. et al. Therapeutic vaccination of chronic hepatitis C nonresponder patients with the IC41. Gastroenterology 134, 1385-1395 (2008).

111. Blanchard, T. J., Alcami, A., Andrea, P. & Smith, G. L. Modified vaccinia virus Ankara undergoes limited replication in human cells and lacks several immunomodulatory proteins: implications for use as a human vaccine. J. Gen. Virol. 79 ( Pt 5), 1159-1167 (1998).

112. Fournillier, A. et al. An accelerated vaccine schedule with a poly-antigenic hepatitis C virus MVA-based candidate vaccine induces potent, long lasting and in vivo cross-reactive T cell responses. Vaccine 25, 7339- 7353 (2007).

113. Alvarez-Lajonchere, L. et al. Immunogenicity of CIGB-230, a therapeutic DNA vaccine preparation, in HCV- chronically infected individuals in a Phase I clinical trial. J. Viral Hepat. 16, 156-167 (2009).

32