IDENTIFICATION OF PUTATIVE FUNCTIONAL MOTIFS IN VIRAL PROTEINS ESSENTIAL FOR HUMAN (HCMV) DNA REPLICATION

Heng Giap Woon

Master of Science (Research)

2008 ABSTRACT Human cytomegalovirus (HCMV) is a ubiquitous that causes significant morbidity and mortality in immunocompromised individuals. Although there are prophylactic treatments available, all current antiviral drugs ultimately target the DNA polymerase, resulting in the increasing emergence of antiviral resistant strains in the clinical setting. There is a fundamental need for understanding the role of other essential genes in DNA replication as a foundation for developing new antiviral treatments that are safe and which utilize a mechanism of action different to existing therapies. In this study we looked at six HCMV replication genes encoding for the DNA polymerase accessory protein (UL44), single stranded DNA binding protein (UL57), primase (UL70), helicase (UL105), primase-helicase associated protein (UL102), and the putative initiator protein (UL84) in order to increase our understanding of their role in DNA replication. The aim of this project was to identify variation within these genes as well as to predict putative domains and motifs in order to ultimately express and study the functional properties of the HCMV primase (UL70) through the use of recombinant mutants. Sequencing of these genes revealed a high degree of conservation between the isolates with amino acid sequence identity of >97% for all genes. Using ScanProsite software from the Expert Protein Analysis System (ExPASy) proteomics server, we have mapped putative motifs throughout these HCMV replication genes. In particular, highly conserved putative N- linked glycosylation sites were identified in UL105 that were also conserved across 33 homologues as well as several unique motifs including casein kinase II phosphorylation sites (CKII) in UL105 and UL84, a microbodies signal motif in UL57 and an integrin binding site in the UL102 helicase-primase associated protein. Our investigations have also elucidated motif-rich regions of the UL44 DNA polymerase accessory protein, mapped functionally important domains of the UL105 helicase and identified cysteine motifs that have implications for folding of the UL70 primase. Taken together, these findings provide insights to regions of these HCMV replication proteins that are important for post-translation modification, activation and overall function, and this information can be utilized to target further research into these proteins and advance the development of novel antiviral agents that target these processes.

I ACKNOWLEDGEMENTS I would like to thank my supervisor, Dr Gillian Scott, for her guidance and encouragement throughout the project, and to my co-supervisor, Professor Bill Rawlinson, for his advice and for giving me the opportunity to work in research. I would also like to extend my thanks to Professor Andrew Lloyd for his advice and guidance.

I would also like to thank my colleagues at the Virology Research Lab for their company, friendship, encouragement, advice and humor. In particular, thanks to Sharon and Min for the countless conversations over lunch.

Thanks to the people at the UNSW Ramaciotti Centre for their technical aid in this project.

I would like to give a special thanks to Professor Patrick Tam for his insights and advice on science research in general and to my parents who have supported me throughout.

Finally I would like to give a very special thanks to Ying, who kept me company on the long nights in the lab and while writing this thesis.

II TABLE OF CONTENTS

ABSTRACT………………………………………………………………………….....I

ACKNOWLEDGEMENTS………………………………………………………...... II

TABLE OF CONTENTS…………………………………………………………….III

LIST OF TABLES…………………………………………………………………....VI

LIST OF FIGURES……………………………………………………………….....VII

COMMONLY USED ABBREVIATIONS………………………………………..VIII

1 INTRODUCTION...... 1

1.1 BIOLOGY...... 1 1.1.1 The ...... 1 1.1.2 The ...... 2

1.2 HCMV EPIDEMIOLOGY...... 3 1.2.1 Transmission ...... 4 1.2.2 Congenital infection...... 5 1.2.3 Immunocompromised infection and disease outcomes...... 6 1.2.4 Antiviral treatment...... 6 1.2.5 Antiviral resistance...... 8

1.3 HCMV GENOME STRUCTURE AND ORGANIZATION...... 13 1.3.1 Open reading frame organization ...... 13

1.4 CMV GROWTH CYCLE ...... 14 1.4.1 Cell attachment and viral entry...... 15 1.4.2 HCMV replication genes...... 16 1.4.3 DNA replication...... 24 1.4.4 Capsid assembly, maturation and egress...... 25 1.4.5 Latency and reactivation...... 26

1.5 AIMS AND OBJECTIVES ...... 26

III 2 REAGENTS AND EQUIPMENT ...... 28

2.1 VIRAL STOCKS ...... 28

2.2 DNA EXTRACTION ...... 28 2.2.1 Buffers...... 28 2.2.2 Chenicals...... 28 2.2.3 Kits...... 29

2.3 POLYMERASE CHAIN REACTION (PCR) ...... 29 2.3.1 Buffers...... 29 2.3.2 Enzymes...... 29 2.3.3 Equipment ...... 30 2.3.4 Reagents...... 30

2.4 ELECTROPHORESIS ...... 30 2.4.1 Buffers...... 30 2.4.2 DNA Markers ...... 31 2.4.3 Equipment ...... 31 2.4.4 Reagents...... 31

2.5 DNA SEQUENCING ...... 32 2.5.1 Equipment ...... 32 2.5.2 Reagents...... 32

2.6 RESTRICTION ENZYMES ...... 32 2.6.1 Buffers...... 32 2.6.2 Enzymes...... 33 2.6.3 Reagents...... 33

2.7 CLONING REAGENTS ...... 33 2.7.1 Antibiotics ...... 33 2.7.2 Cells ...... 34 2.7.3 Equipment ...... 34 2.7.4 Kits...... 35 2.7.5 Media ...... 36 2.7.6 Reagents...... 36 2.7.7 Vectors ...... 36

2.8 PROTEIN EXPRESSION AND DETECTION...... 37 2.8.1 Antibodies...... 37

IV 2.8.2 Buffers...... 37 2.8.3 Equipment ...... 38 2.8.4 Kits...... 38 2.8.5 Protein Markers...... 38 2.8.6 Reagents...... 39 2.8.7 Solutions...... 39

3 IDENTIFICATION OF PUTATIVE FUNCTIONAL MOTIFS IN VIRAL PROTEINS ESSENTIAL FOR HUMAN CYTOMEGALOVIRUS DNA REPLICATION...... 40

3.1 INTRODUCTION ...... 40

3.2 METHODS ...... 41 3.2.1 Viral stocks and DNA extraction...... 41 3.2.2 Polymerase chain reaction (PCR) amplification ...... 41 3.2.3 DNA sequencing and analysis...... 44 3.2.4 Phylogenetic analysis ...... 44 3.2.5 Genbank Accession Numbers...... 44

3.3 RESULTS ...... 45 3.3.1 Strain variation amongst HCMV isolates...... 45 3.3.2 Prediction of potential motifs...... 45 3.3.3 Strain variation and identification of protein motifs in UL44 ...... 46 3.3.4 Strain variation and identification of protein motifs in UL57 ...... 51 3.3.5 Strain variation and identification of protein motifs in UL70 ...... 60 3.3.6 Strain variation and identification of protein motifs in UL102 ...... 67 3.3.7 Strain variation and identification of protein motifs in UL105 ...... 73 3.3.8 Strain variation and identification of protein motifs in UL84 ...... 80

3.4 DISCUSSION...... 85

4 HCMV PRIMASE EXPRESSION...... 93

4.1 INTRODUCTION ...... 93

4.2 METHODS ...... 95 4.2.1 Cloning of UL70 into high-copy pBluescript II primase vector...... 95 4.2.2 Construction of baculovirus entry clone for HCMV primase expression 99

V 4.2.3 In vitro UL70 primase expression using TNT Quick Coupled Transcription/Translation ...... 102

4.3 RESULTS ...... 108 4.3.1 UL70 was cloned into pBluescript II high-copy vector ...... 108 4.3.2 Generating Baculovirus entry clones via BP recombination...... 110 4.3.3 UL70 was cloned into pCITE expression vector...... 114 4.3.4 UL70 in vitro expression using the TNT system...... 115

4.4 DISCUSSION...... 117

5 CONCLUDING DISCUSSION AND FUTURE DIRECTIONS...... 119

6 REFERENCES...... 121

VI LIST OF TABLES Table 1 Classification of Human herpesviruses and pathological outcomes...... 2 Table 2 Sequenced cytomegalovirus (CMV) genomes...... 3 Table 3 Epidemiological seroprevalence of HCMV infection...... 4 Table 4 Mutations of UL97 and UL27 antiviral resistance ...... 11 Table 5 Mutations of UL54 associated with antiviral resistance ...... 12 Table 6 Essential DNA replication proteins in HCMV and HSV-1...... 40 Table 7 Primers used for PCR amplification of HCMV DNA fragments ...... 43 Table 8 Sequence variation within HCMV replication genes...... 45 Table 9 Summary of the putative motifs identified in the HCMV replication genes..... 46 Table 10 Nucleotide primers used for UL70/pBluescript cloning ...... 96 Table 11 Nucleotides primers used for UL70/pDONR cloning...... 100 Table 12 Nucleotide primers used for UL70/pCITE cloning...... 103

VII LIST OF FIGURES Figure 1 Arrangement of the HCMV genome...... 13 Figure 2 CMV growth cycle...... 15 Figure 3 Relative positions of genes investigated in this study ...... 27 Figure 4 HCMV strain alignments of UL44 ...... 50 Figure 5 HCMV strain alignments of UL57 ...... 58 Figure 6 Putative zinc finger domain of UL57...... 59 Figure 7 HCMV strain alignments of UL70 ...... 65 Figure 8 Conserved domains between UL70 homologues...... 66 Figure 9 HCMV strain alignments for UL102 ...... 72 Figure 10 HCMV strain alignments for UL105 ...... 78 Figure 11 Putative motifs conserved across UL105 homologues ...... 79 Figure 12 HCMV strain alignments for UL84 ...... 83 Figure 13 Overlapping putative casein kinase-2 (CKII) phosphorylation sites of UL8484 Figure 14 Isolate variation in the HCMV primase-helicase complex...... 91 Figure 15 Insert to vector ratio formula...... 96 Figure 16 Overview of cloning UL70 into pBluescript...... 98 Figure 17 pDONR/Zeocin Vector ...... 100 Figure 18 converting femtomoles to nanograms...... 101 Figure 19 Overview of UL70/pDONR recombination...... 102 Figure 20 Overview of cloning UL70 into pCITE expression vector ...... 105 Figure 21 Overview of UL70 in vitro protein expression...... 107 Figure 22 pBluescript UL70a...... 108 Figure 23 pBluescript UL70b...... 109 Figure 24 Screening of entry clone BPR4 ...... 111 Figure 25 The BPR4 truncation removes key motifs from UL70...... 111 Figure 26 Colony PCR screening for entry clone BPR5 ...... 112 Figure 27 The BPR5 truncation of UL70...... 113 Figure 28 The BPR8 truncation removes the putative zinc finger from UL70...... 113 Figure 29 UL70/pCITE expression vector...... 114 Figure 30 Expression of UL70 in TNT system ...... 115 Figure 31 Time course expression of UL70 with TNT system...... 116 Figure 32 Sequence analysis of BPR4 and BPR8 ...... 118

VIII COMMONLY USED ABBREVIATIONS

AIDS Acquired immunodeficiency syndrome

ANGIS Australian National Genomic Information Service

ATP Adenosine triphosphate bp Base pairs

CDV Cidofovir

CMV Cytomegalovirus

DNA Deoxyribonucleic acid dNTP Deoxyribonucleoside triphosphate

EBV Epstein-Barr virus

FOS Foscarnet

GCV Ganciclovir

HCMV Human cytomegalovirus

HHV Human Herpesvirus

HIV Human immunodeficiency virus

HSV-1/2 type 1/2

IE Immediate early gene kbp Kilo-base pairs

MIE Major immediate early gene mRNA Messenger RNA

NLS Nuclear localization signal

IX OBP Origin binding protein

ORF Open reading frame

PCR Polymerase chain reaction

PEG Polyethylene glycol

RNA Ribonucleic acid

UL Long unique genome region

US Short unique genome region vGCV Valganciclovir

VZV Varicella-zoster virus

X 1 Introduction

1.1 Biology

1.1.1 The herpesviridae The herpesviridae family consists of a large group of host-specific animal that are ubiquitous to the vertebrate species and in at least one invertebrate (Davison, 2002). Herpesviruses are defined by their virion morphology, which consists of the core, capsid, tegument, and envelope (Arvin, 2007). The core consists of a single, linear double- stranded DNA genome of 125–240 kilo-base pairs (kbp) packaged within an icosahedral capsid that is imbedded in an inner proteinaceous tegument matrix and an outer glycoprotein-rich lipid membrane (Davison, 2002). Initial phylogenetic analysis of fish and mammalian herpesviruses dated the divergence of the Alpha-, Beta- and subfamilies to approximately 180-200 million years ago (McGeoch, 1995), but the inclusion of avian and reptilian herpesviruses have pushed back the common ancestor of the herpesvirinae subfamilies to about 400 million years ago (McGeoch, 2005). Classification into Alpha-, Beta- and Gammaherpesvirinae subfamilies were based on biological properties such as host range and ability to establish latent infection in certain cell types. The include four genera (, , , and ) as well as the reptilian herpesviruses and are characterized by their relatively rapid replication and ability to maintain latent infections in sensory ganglia as well as productively infect mucous membranes and skin tissue (Arvin, 2007). The include the Cytomegalovirus, , and lineages and are characterized by their slower replication rate in vitro relative to other herpesviruses (Arvin, 2007). The Gammaherpesvirinae are divided into two genera, the Lymphocryptoviridae and Rhadinoviridae, which primarily infect mammals and are known to induce lymphoproliferation and tumors in endothelial cells (Arvin, 2007). In addition, there is a genus () that is unattached to any subfamily as well as a large number of species not assigned to genera (Arvin, 2007). Currently, there are eight herpesviruses known to infect humans with pathological consequences (Table 1).

1 The herpesviridae subfamilies share many homologous ‘core genes’ which were presumably inherited from a common ancestor and are associated with DNA replication, processing and packaging, capsid assembly and egress, as well as a host of tegument proteins and surface glycoproteins (Davison, 2002).

Table 1 Classification of Human herpesviruses and pathological outcomes

Common Name Genome Subfamily Pathological Outcomes (Classification) size (kbp)* Herpes Simplex Virus 1 (HHV-1) 152 α Predominantly oro-facial lesions, keratoconjunctivitis Herpes Simplex Virus 2 (HHV-2) 155 α Predominantly genital lesions (HHV-3) 125 α Chickenpox, shingles Epstein-Barr Virus (HHV-4) 184 γ Burkitt’s and Hodgkin’s lymphoma infectious mononucleosis, tumors Human Cytomegalovirus (HHV-5) ~230 β Vasculature and end-organ disease, congenital abnormalities, retinitis, (HHV-6A/B) 159/162 β Exanthem subitum (HHV6A) Human Herpesvirus 7 (HHV-7) ~149 β Exanthem subitum Human Herpesvirus 8 (HHV-8) ~141 γ Kaposi’s sarcoma *in kilobase pairs (Arvin, 2007; Davison, 2002)

1.1.2 The Cytomegaloviruses The cytomegalovirus genus infects a wide range of mammalian species including rodents (Rawlinson, 1996; Vink, 2000), primates (Davison, 2003; Hansen, 2003), and humans (Chee, 1990; Dolan, 2004; Dunn, 2003; Murphy, 2003) with several sequenced genomes available on public databases (Table 2). The human cytomegalovirus (HCMV) is the prototype of the betaherpesvirinae subfamily and is the most complex of the eight human herpesvirus species (Davison, 2003). The first HCMV strains were isolated 50 years ago (Craig, 1957; Rowe, 1956; Smith, 1956) while the genetic content of the virus has been available for almost two decades (Chee, 1990). However, the mechanism of HCMV DNA replication and its regulation is not completely understood, with the function of several genes inferred from studies on the herpes simplex virus (HSV). In addition, current anti-viral treatments exclusively target only one aspect of the CMV replication cycle, by competitive inhibition of the DNA polymerase, and with the advent

2 of anti-viral resistance strains, the efficacy of treatment is greatly reduced. Hence other aspects of the HCMV replication machinery need to be studied in greater detail as a basis of identifying a broader range of potential anti-viral targets in the future.

Table 2 Sequenced cytomegalovirus (CMV) genomes

Common Name Strain Accession size (kbp) Reference Human CMV AD169 X17403 229 Chee et al. (1990) Towne AY315197 231 Dunn et al. (2003) Merlin AY446894 236 Dolan et al. (2004) Toledo AC146905 227 Murphy et al. (2003) Chimpanzee CMV - AF480884 241 Davison et al. (2003) Rhesus CMV 68-1 AY186194 221 Hansen et al. (2003) Mouse CMV Smith U68299 230 Rawlinson et al. (1996) Rat CMV Maastricht AF232689 230 Vink et al. (2000)

1.2 HCMV Epidemiology

Human cytomegalovirus (HCMV) is a ubiquitous pathogen that is acquired early in life in most populations and generally increases in seroprevalence with age. However, the patterns of acquisition vary greatly based on geographic and socioeconomic backgrounds such that in developing nations, acquisition of HCMV is nearly universal in early childhood (Kubo, 1991; Miles, 2007; Prabhakar, 1992). HCMV infection is endemic in the human population (Table 3) and is not influenced by seasonal variations (Malm, 2007; Pass, 2005). The seroprevalence of HCMV infection within the adult population is very common ranging from 60% in developed countries, to 100% in developing countries (Munro, 2005). In particular, countries in Africa (Adjei, 2006; Pultoo, 2001), Asia (Kangro, 1994; Kothari, 2002; Liu, 1990; Shen, 1992; Urwijitaroon, 1993; Wong, 2000), and the Eurasian border (Hizel, 1999) tend to have a higher incidence of HCMV infection with less significant differences between age groups than the European countries (Alanen, 2005; Andersson-Ellström, 1995; de Ory, 2004; Hecker, 2004; Kangro, 1994; Knowles, 2005; Natali, 1997), Australia (Seale, 2006), New Zealand (Beresford, 1988), and the United States (Staras, 2007).

3 Table 3 Epidemiological seroprevalence of HCMV infection Location Population Age Seropositivity Reference Study (years) (%) Europe Germany Blood donor 18+ 46 Hecker et al. (2004) Parma, Italy General 2-54 72 Natali et al. (1997) Finland Women 16-45 56 Alanen et al. (2005) Madrid, Spain Women 2-40 66 de Ory et al. (2004) United Kingdom General 12-21 19 Kangro et al. (1993) Ankara, Turkey Women 15-49 99 Hizel et al. (1999) Ireland Women 15-46 30 Knowles et al. (2005) Sweden Women 16 45 Andersson-Ellström et al. (1995) Asia Chengdu, China General 4-7 60 Liu et al. (1990) Taipei, Taiwan General 4-12 58 Shen et al. (1992) Delhi, India Blood donor 95 Kothari et al. (2002) Thailand Blood donor 17-59 93 Urwijitaroon et al. (1993) Hong Kong General 12-21 80 Kangro et al. (1993) Nepal General 1+ 100 Kubo et al. (1991) Singapore Pregnant women ≤30 77 Wong et al. (2000) North America USA General 6+ 59 Staras et al. (2006) Cuernavaca, Mexico Women 13-44 92 Echániz-Avilés et al. (1993) South America Jamaica General 15-25 90+ Prabhakar et al. (1992) Santiago, Chile General <30 60 Abarca et al. (1997) Buenos Aires, Argentina Socioeconomic* <15 46 Damilano et al. (1992) Africa Ghana Blood donor 20-69 93 Adjei et al. (2006) Gambia General 1 85 Miles et al. (2007) Mauritius Blood donor 18-60 94 Pultoo et al. (2001) Australia General 1-59 57 Seale et al. (2006) New Zealand Blood donor 56-65 65 Beresford et al. (1988) *Population study based on middle-socioeconomic classes

1.2.1 Transmission HCMV infection is acquired through direct contact with body fluids from an infected person and is present in saliva, urine, breast milk, blood products, semen, cervical and vaginal secretions, as well as allograft tissues (Arvin, 2007). In a hospital environment, allograft recipients are among the highest at risk to HCMV infection, with over 75% of

4 solid organ transplant patients contracting or reactivating latent CMV after transplantation (Fishman, 1998). Transfusion associated HCMV infections were first described 40 years ago (Kääriäinen, 1966) and is attributed to the virus’s ability to establish latency in cells of the myeloid lineage (Cervia, 2007; Miller, 1991). In a child- care environment, horizontal transmission of HCMV is common amongst young children, parents and other care-givers from the community (Adler, 1988; Adler, 1989; Bale, 1999; de Mello, 1996; Kiss, 2002; Pass, 1990; Pass, 1984) while in young adult populations, sexual activity is the major mode of transmission for CMV (Handsfield, 1985). The presence of infectious virus is common in salivary glands and cervical secretions (Britt, 1996; Shen, 1994). HCMV has also been cultured from semen despite being urine, blood, and saliva negative for infectious virus (Biggar, 1983; Lang, 1975). In addition, artificial insemination therapy has been suggested as another route of transmission (Mansat, 1997; Prior, 1994), with HCMV detected in 0.02% to 34% of donor semen depending on the population study (Mansat, 1997; McGowan, 1983; Shen, 1994a; Tjiam, 1987; Yang, 1995).

1.2.2 Congenital infection HCMV is the leading cause of congenital viral infection, occurring in 0.15-3.0% of newborns worldwide (Hassan, 2007; Malm, 2007). Infection can occur via viral transmission through the placenta, during delivery via cervical secretions and blood or from the mother via breast milk (Malm, 2007). The risk for viral transmission is higher in primary infected mothers than in mothers with reactivated disease (Gaytant, 2002). Primary CMV infections are reported in 1-4% of seronegative women during pregnancy and the risk for viral transmission to the fetus is 30-40% (Stagno, 1986; Stagno, 1982) while reactivation of a CMV infection during pregnancy is reported in 10-30% of seropositive women but the risk of transmitting the virus is only 1-3% (Stagno, 1982). Approximately 30% of those infected will develop symptomatic disease (Ahlfors, 1999), although only 10-15% of children with congenital CMV are symptomatic during the neonatal period (Malm, 2007). Hearing loss is the most common sequela of congenital CMV infection, occurring in 10–15% of infected children (Dahle, 2000; Dollard, 2007; Pass, 2005; Ross, 2006) while other neurological deficits such as mental retardation, autism, learning disabilities, cerebral palsy, epilepsy, visual impairment, microcephaly, encephalitis, seizures, and upper motor neuron disorders may also develop (Gandhi,

5 2004; Ross, 2005). The most severely affected infants have a mortality rate of about 30% with death occurring as a result of hepatic dysfunction, bleeding, disseminated intravascular coagulation or secondary bacterial infections (Malm, 2007).

1.2.3 Immunocompromised infection and disease outcomes HCMV infection often remains asymptomatic within an immunocompetent host, whereas it may cause significant morbidity and mortality in immunocompromised individuals such as transplant patients and HIV/AIDS patients (Arvin, 2007). Over 75% of solid organ transplant patients are newly infected or reactivate latent CMV after transplantation (Fishman, 1998), often within the first 3 months when immunosuppression is most intense (Dummer, 1983). CMV disease manifests in the vast majority of transplant recipients as a viral syndrome that includes fever, malaise, muscle pain or headache (Steininger, 2007) and has been associated with diminished graft survival and allograft rejection (Grattan, 1989; Pouteil-Noble, 1993; Ricart, 2005). In HIV-infected patients, retinitis is the single most common manifestation of CMV disease, accounting for 85% of all cases (Yust, 2004). CMV infection has also been linked to the onset of diabetes mellitus in kidney transplant recipients (Hjelmesaeth, 2004) as well as accelerated cardiac allograft vasculopathy in heart transplant recipients (Lemström, 1995).

1.2.4 Antiviral treatment Currently, only five compounds have been licensed to treat established HCMV infections: ganciclovir (GCV), its oral pro-drug, valganciclovir (vGCV), Foscarnet (FOS), cidofovir (CDV) and fomivirsen (Mercorelli, 2007). GCV is a nucleoside analog commonly used to treat CMV disease and in particular, CMV retinitis in the immunocompromised (De Clercq, 2001). GCV requires intracellular phosphorylation, firstly by the UL97 viral kinase and then by cellular kinases to create a triphosphate active form that acts as a competitive inhibitor of deoxyguanosine triphosphate, incorporating itself into the viral DNA via the DNA polymerase during elongation to disrupt DNA synthesis (Abdel-Haq, 2006). Valganciclovir is an inactive pro-drug form of GCV that is rapidly converted to active GCV in the plasma (Abdel-Haq, 2006) and has a ten-fold higher bioavailability than GCV (Cocohoba, 2002; Pescovitz, 2000),

6 which is normally administered intravenously. Hematologic toxicities such as neutropenia, anemia, and thrombocytopenia are among the most frequently reported adverse side effects (Abdel-Haq, 2006; Mercorelli, 2007) while nausea, diarrhea, vomiting, fever, and abdominal pain have also been reported during GCV and valganciclovir treatment (Cocohoba, 2002). CDV is an acyclic nucleoside phosphonate with a mechanism of action similar to that of other nucleoside analogs, but unlike GCV, is not dependent on activation by a viral encoded enzyme (Mercorelli, 2007). Instead, host kinases convert CDV to the active diphosphoryl form, and CDV disphosphate then acts as a competitive inhibitor of the viral DNA polymerase, causing premature chain termination in viral DNA synthesis (Biron, 2006). Although CDV is approved for use to treat CMV retinitis in AIDS patients, severe nephrotoxicity is a major limitation of its use (Biron, 2006; Ho, 2000; Mercorelli, 2007) and thus remains a second-line therapy option (Biron, 2006). In addition, it has low oral bioavailability and is only available as an intravenous treatment (Mercorelli, 2007). Neutropenia is another toxicity associated with CDV, and CDV was shown to be both carcinogenic and teratogenic in preclinical toxicological studies (Biron, 2006). FOS is a pyrophosphate analogue that inhibits the viral DNA polymerase by binding to the pyrophosphate binding site and blocking cleavage of pyrophosphate from the terminal nucleoside triphosphate added to the growing DNA chain (Biron, 2006; Wagstaff, 1994). FOS is also considered a second- line therapy due to its nephrotoxic potential (Biron, 2006; Mercorelli, 2007), but is the preferred drug for patients who are failing GCV therapy due to viral resistance, or those who cannot be treated with GCV due to dose-limiting neutropenia or leucopenia (Razonable, 2004). Fomivirsen is a 21-nucleotide anti-sense RNA (5’-GCG TTT GCT CTT CTT CTT GCG-3‘) (De Clercq, 2001), specifically targeted against the major immediate-early (MIE) transactivator gene of CMV (Geary, 2002). As it is administered only as an intraocular injection, it is an end-organ treatment that does not affect systemic CMV (Arvin, 2007). The most commonly reported adverse effect associated with fomivirsen is ocular inflammation, which occurs in 25% patients (Group., 2002), with other less common side-effects including abnormal vision, cataracts, bleeding in and around the eye, reduced color vision, eye pain, retinal detachment, stomach pain, low blood count, weakness, dehydration, cough, flu-like symptoms, and chest pain (Group., 2002).

7 In addition to these licensed drugs, several anti-CMV drugs are also currently in clinical development. Maribavir (MBV), (1-(β-L-ribofuranosyl)-2-isopropylamino-5, 6- dichlorobenzimidazole), is a riboside analog with a novel mechanism of action that is being developed to inhibit viral DNA synthesis in CMV and EBV (Biron, 2002; Zacny, 1999). In CMV-infected cells, Maribavir targets the HCMV protein kinase gene (UL97) that has been shown to play a role in viral nucleocapsid egress from the nucleus, thereby reducing the yield of infectious CMV (Evers, 2004; Krosky, 2003a). Its mechanism of action is not fully understood although it has been hypothesized that inhibition of viral DNA synthesis is a consequence of preventing the phosphorylation of the DNA polymerase accessory protein (pUL44)by the protein kinase (Krosky, 2003; Marschall, 2003). In addition, laboratory-generated resistant mutations have been mapped to the UL27 gene, a protein of unknown function (Chou, 2004; Komazin, 2003). Maribavir pre-clinically shows advantages over existing anti-CMV drugs in its in vitro potency, bioavailability, safety profile in acute, chronic and genetic toxicology testing, and the lack of cross-resistance with GCV, FOS, and CDV resistant strains (Biron, 2006; Drew, 2006).

BAY 38-4766 represents a novel class of non-nucleoside antiviral agents that are highly selective inhibitors of CMV in vitro (Reefschlaeger, 2001). The compound demonstrates antiviral activity similar to that of GCV (Weber, 2001) and showed a favorable safety profile in healthy male volunteers at single oral doses up to 2000 mg (Biron, 2006). The mechanism of action is largely unknown although resistance mutations have been mapped to the UL89 and UL56 genes (Krosky, 1998; Reefschlaeger, 2001), which encode the two terminase subunit proteins involved in DNA cleavage and packaging (Bogner, 2001). It is also active against strains that are resistant to GCV (McSharry, 2001) and does not have cross-resistance with other current antiviral drugs (Reefschlaeger, 2001).

1.2.5 Antiviral resistance Resistance to antiviral treatment mainly occurs under selective pressure from lengthy periods of antiviral treatment, although other risk factors including host immune competence and levels of ongoing viral replication are also involved (Chou, 2001). In addition, currently approved drugs for treatment share a similar mechanism of action by

8 targeting the DNA polymerase, hence increasing the risk of mutations that confer resistance to multiple anti-viral drugs. Antiviral resistant CMV strains may emerge between 6 weeks to 2 months of treatment depending on the patient group (Boivin, 2001; Eckle, 2002; Limaye, 2000; Springer, 2005) although low levels of resistant genotypes are present in patients even before antiviral treatment commences (Emery, 2000). Antiviral resistance can be determined by genotypic or phenotypic assays. Phenotypic assays measure the concentration of an antiviral agent necessary to inhibit viral replication while genotypic assays, screen for key resistance mutations in the UL97 and UL54 genes (Springer, 2005). These two gene products encode for the CMV protein kinase and DNA polymerase respectively. Point mutations or deletions of portions of UL97 can lead to GCV resistance (Baldanti, 1995; Baldanti, 1998; Chou, 1995; Chou, 1997; Chou, 2000a; Chou, 2005; Chou, 2002; Faizi-Khan, 1998; Hanson, 1995; Hantz, 2005; Ijichi, 2002; Lurain, 1994; Marfori, 2007; Sullivan, 1992; Wolf, 1995) (Table 4), while UL54 mutations, depending on their locations, can confer resistance to one or more anti-viral drugs (Chou, 2003; Chou, 1997; Chou, 1998; Chou, 2007a; Cihlar, 1998; Cihlar, 1998a; Ducancelle, 2006; Gilbert, 2005a; Marfori, 2007; Mousavi-Jazi, 2001; Mousavi-Jazi, 2003; Scott, 2007; Springer, 2005; Weinberg, 2003) (Table 5).

UL97 GCV resistance mutations are clustered at codons 460, 520, and 590-607 (Chou, 1999), but the degree of resistance varies depending on the mutation or combination of mutations present. For example, the most commonly encountered mutations in UL97, A594V, L595S, and M460V, confer a 5-10 fold increase resistance to GCV while C592G and A594T confer only a 2-3 fold increase resistance to GCV (Chou, 1999; Chou, 2005). Maribavir resistance mutations have also been mapped to the UL97 gene (Biron, 2002; Chou, 2007), although the affected codons are clustered upstream of the known GCV resistance mutations and share no cross-resistance with GCV (Chou, 2007). In addition to GCV, UL54 mutations are known to confer varying degrees of resistance to FOS and/or CDV as well. Among the most frequent mutations associated with resistances to these drugs are V715M, V781I, and L802M, which confer resistance to FOS (Cihlar, 1998), F412C, L501I, and P522S, which confer resistance to GCV and CDV (Chou, 1997; Cihlar, 1998), and A809V, which confer resistance to GCV and FOS (Chou, 1998). In addition, an increasing number of GCV and FOS cross-resistant strains have been observed in several laboratory and clinical isolates (Biron, 2006; Mercorelli, 2007) while there have also been cases of mutations that confer resistance to 9 all three anti-CMV drugs (Chou, 2000; Scott, 2007). The issue of cross-resistance arising from the above mentioned factors poses a need to develop drugs with a novel mechanism of action to use with patients not responding to current treatment or in conjunction with existing anti-viral drugs to reduce the incidence of resistance.

10 Table 4 Mutations of UL97 and UL27 antiviral resistance Gene Mutation Domain Resistance* Reference UL97 V353A --- MBV Biron 2002, Chou 2007 L397R --- MBV Biron 2002 T409M --- MBV Biron 2002, Chou 2007 H411 --- MBV Chou 2008 M460V VI GCV Chou 1995 M460I VI GCV Lurain 1994 H520Q --- GCV Hanson 1995 590-93 deletion IX GCV Sullivan 1992 A591V IX GCV Chou 2002 591-607 deletion IX GCV Chou 2002 C592G IX GCV Chou 2005 A594V IX GCV Chou 1995 A594P IX GCV Ijichi 2002 A594T IX GCV Chou 2002 L595S IX GCV Chou 1995 L595F IX GCV Wolf 1995 L595W IX GCV Chou 2002 L595 deletion IX GCV Baldanti 1995 595-603 deletion IX GCV Chou 2000a E596G --- GCV Chou 2002 K599T --- GCV Faizi-Khan 1998 601 deletion --- GCV Hantz 2005 601-03 deletion --- GCV Marfori 2007 C603W/R/S --- GCV Chou 1997 C607Y --- GCV Baldanti 1998 C607F --- GCV Chou 2002 UL27 L355P --- MBV Komazin 2003 R233S --- MBV Chou 2004 W362R --- MBV Chou 2004 A406V/415Stop --- MBV Chou 2004 *GCV = Gancyclovir, MBV = Maribavir

11 Table 5 Mutations of UL54 associated with antiviral resistance Gene Mutation Domain Resistance* Reference UL54 D301N EXOI GCV Chou 2003 N408D IV GCV, CDV Cihlar 1998 N408K IV GCV, CDV Scott 2007 N410K IV GCV, CDV Chou 2003 F412C/V IV GCV, CDV Chou 1997, Cihlar 1998 D413E or A IV GCV, CDV Chou 2003 D413A IV GCV, CDV Marfori 2007 T419M IV FOS Mousavi 2003 N495K δ-C FOS Ducancelle 2006 L501I δ-C GCV, CDV Cihlar 1998 T503I δ-C GCV, CDV Chou 2003 K513N or E δ-C GCV, CDV Cihlar 1998a L516R δ-C GCV, CDV Chou 2003 P522S δ-C GCV, CDV Cihlar 1998 L545S δ-C GCV, CDV Cihlar 1998 Q578H δ-C FOS Mousavi 2003 D588N or E δ-C FOS Mousavi 2001, Cihlar 1998 A692S --- FOS Chou 2003 T700A II FOS Cihlar 1998 V715M II FOS Cihlar 1998 E756K --- GCV, CDV, FOS Mousavi 2001, Chou 2003 E756D or Q --- FOS Chou 2003, Weinberg 2003 L773V VI FOS Mousavi 2003 V781I VI GCV, FOS Cihlar 1998, Mousavi 2001 V787L VI FOS Weinberg 2003 L802M III FOS Cihlar 1998, Chou 1997 K805Q III CDV Cihlar 1998 A809V III GCV, FOS Chou 1998 V812L III GCV, CDV, FOS Chou 1997, Cihlar 1998a T813S III GCV, FOS Chou 2007a T821I III GCV, CDV, FOS Cihlar 1998 P829S III FOS Gilbert 2005a A834P III GCV, CDV, FOS Scott 2007 T838A III FOS Springer 2005 G841A III GCV, FOS Chou 2007a 981-82 deletion V GCV, CDV, FOS Chou 2000 A987G V GCV, CDV Cihlar 1998 *GCV = Ganciclovir, CDV = Cidofovir, FOS = Foscarnet

12 1.3 HCMV genome structure and organization

Human cytomegalovirus has the largest genome of all the herpesviruses with a linear double stranded DNA genome of 230-240kbp. It is also the only known betaherpesvirus to have a class E genome structure that consists of a long unique (UL) and short unique (US) sequence capable of inverting to give four sequence isomers (Gibson, 1999). Each unique region is flanked by inverted repeat sequences (TRL and IRL, TRS and IRS), such that the overall structure of the HCMV genome is TRL-UL-IRL-IRS-US-TRS (Mocarski, 2001) (Figure 1). Large scale sequencing analysis has revealed the existence of seven conserved sequence blocks present in all herpesviruses, but arranged in different orders (Chee, 1990). These conserved regions were found to have a functional role in DNA replication, DNA repair, nucleotide metabolism, and virion structure (Mocarski, 2001).

Figure 1 Arrangement of the HCMV genome

1.3.1 Open reading frame organization Initial analysis of the fully sequenced AD169 laboratory strain predicted the existence of 208 open reading frames (ORFs) (Chee, 1990), including 33 ORFs that have substantial sequence similarity to HSV-1, varicella zoster virus (VZV), and Epstein- Barr virus (EBV) (Chee, 1990), as well as 41 ORFs that are dispensable for replication (Spaete, 1987). Several studies have proposed their models for a consensus genome (Chee, 1990; Davison, 2003; Murphy, 2003), while various interpretations to the coding potential of AD169 (Chee, 1990) has led to several revisions of its genomic content. Sequence comparisons with the closely related chimpanzee cytomegalovirus (CCMV) discounted 51 previously identified ORFs, re-interpreted 24 ORFs, as well as predicting the existence of 10 novel genes, including five in AD169 (UL15A, UL21A, UL128, UL131A, and US34A) (Davison, 2003). Another study, using the Bio-Dictionary-based Gene Finder (BDGF) analysis tool, found that 37 ORFs from the original AD169 sequence can be discarded as non-polypeptide encoding as well as the addition of up to

13 12 ORFs that had not been previously identified (Murphy, 2003). In addition, sequence differences have been pointed out in ORF UL102 (Smith, 1995a), US28 (Neote, 1993), and UL15 (Davison, 2003). Certain stocks of AD169 were also found to contain an additional 929 base pairs that were not in the original sequence of AD169 and which resulted in alterations to the UL42 and UL43 genes (Dargan, 1997; Mocarski, 1997) as well as a reinterpretation of the UL41 gene (Dargan, 1997). Several ORFs have also been modified to account for splicing events, including UL111A, UL118, UL119, UL22, US3 (Rawlinson, 1993), and UL33 (Davis-Poynter, 1997). Furthermore, various regions of the genome are known to be hypervariable among clinical isolates (Prichard, 2001). These include: three major genotype variations in UL11 (Hitomi, 1997), UL144 (Lurain, 1999), four major genotype variations in US9, US28 (Rasmussen, 2003), UL4 (Bar, 2001), UL73 (Dal Monte, 2004; Pignatelli, 2001), UL55 (Chou, 1991; Meyer-König, 1998), and five major genotype variations in UL37 (Hayajneh, 2001), as well as several associated with UL146 (Hassan-Walker, 2004). Nevertheless, laboratory strain AD169 is still widely regarded as the consensus strain for interpretation of coding.

1.4 CMV growth cycle

HCMV generally replicates slowly in cell culture, taking days for cytopathic effects to appear (Britt, 1996). However, in vivo, CMV has been observed to have a doubling time of approximately 1-2 days, depending on the methods used (Emery, 1999). The HCMV growth cycle consists of cell attachment and viral entry, transport to the nucleus, transcription and translation of viral proteins, DNA replication, capsid assembly, and maturation and egress from the cell (Figure 2). The CMV replication cycle has not been completely elucidated, although it shares numerous homologies to HSV-1 replication (Boehmer, 2003).

14

Figure 2 CMV growth cycle

1. Cell attachment and viral entry; 2. Transport to the nucleus; 3. Transcription/translation of viral proteins; 4. DNA replication; 5. Capsid assembly; and 6. Maturation and egress

1.4.1 Cell attachment and viral entry HCMV cell attachment and penetration of the host cell occurs by fusion with the plasma membrane (Compton, 2004) and is mainly mediated by envelope glycoproteins (Arvin, 2007; Mach, 2005). Entry of HCMV involves the interaction of these envelope glycoproteins with a number of distinct receptors, initiated by the use of the cell surface proteoglycan heparan sulfate by glycoprotein B (gB) (Compton, 1993). HCMV gB (UL55) and the gH:gL (UL75:UL115) complex have been the most extensively studied (Baldwin, 2000; Feire, 2004; Lopper, 2004; Navarro, 1993) while recent evidence suggests core envelope glycoproteins gM (UL100) and gN (UL73) may also be important for viral entry (Mach, 2005). HCMV and other herpesviruses exploit normal cytoplasmic transport systems of the host (Dohner, 2005; Ogawa-Goto, 2003) to control nucleocapsid transit through the cytoplasm. It relies on microtubules to gain access to the nucleus and nuclear pores where un-coating is completed and the viral genome is released into the nucleoplasm (Dohner, 2005). The importance of essential protein-

15 protein interactions in receptor recognition and signaling for viral attachment and entry underscores the potential for developing drugs which target other aspects of the CMV replication cycle.

1.4.2 HCMV replication genes HCMV replication is associated with eleven loci that have been identified as necessary for transient complementation of oriLyt-dependent DNA replication (Pari, 1993). Six of these were found to encode homologs of herpes simplex virus (HSV-1) DNA replication proteins (Anders, 1996; Pari, 1993), including: a two-subunit DNA polymerase (UL54 and UL44), a single-stranded DNA binding protein (UL57), a primase (UL70), a helicase (UL105), and a primase-associated factor (UL102). In addition to these six core replication genes, five other loci supplemented in regulatory and transactivation roles. The UL36-38 locus is involved in the up-regulation of viral transcription, inhibition of apoptosis, and growth in human cells (Colberg-Poley, 1996; Goldmacher, 1999; Pari, 1993a; Smith, 1995; Terhune, 2007), while the four spliced variants of UL112-113 cooperate with each other to relocate UL44 and possibly other core replication proteins to the pre-replication foci (Iskenderian, 1996; Park, 2006). The TRS1/IRS1 locus is involved in the disruption of host cell antiviral response pathways, the production of viable virus, as well as having a regulatory function for the other immediate early genes (Adamo, 2004; Blankenship, 2002; Child, 2004; Hakki, 2005; Romanowski, 1997). The major immediate-early region (IE1/IE2) plays a role in regulating the cell cycle, metabolism and apoptosis (Chiou, 2006; Lukac, 1999; Yu, 2002; Zhu, 1995). Although not considered to be part of the core set of replication machinery proteins, UL84 has been shown to have an essential role in oriLyt-dependent DNA replication (Sarisky, 1996; Xu, 2004). It is also involved in promoting the formation of replication compartments (Sarisky, 1996; Xu, 2002) as well as interacting with IE2 to regulate transactivation activity in the UL112-113 locus (Gebert, 1997; Spector, 1994).

1.4.2.1 DNA polymerase (UL54) HCMV encodes a two-subunit DNA polymerase which comprises a catalytic subunit, UL54 (Heilbronn, 1987) and an accessory protein, UL44, which acts as a processivity factor (Ertl, 1992). Initial studies revealed that UL54 possesses DNA-dependent DNA

16 polymerase activity (Mar, 1981) as well as 3’-5’ exonuclease activity (Nishiyama, 1983) and was highly homologous to the UL30 protein of HSV-1 (Gottlieb, 1990; Hernandez, 1990). UL54 shares a series of conserved regions and structural domains with the α- family of DNA polymerases. These conserved regions are located within amino acid residues 379–1100 and include conserved domains I–VII, which are involved in the DNA polymerisation reaction (Hwang, 1992; Wong, 1988), the δ-region C (Zhang, 1991) and three motifs designated Exo I–III which are located within conserved domains IV and δ-region C (Bernad, 1989). Mutations conferring resistance to GCV, CDV, and FOS map to several of these conserved motifs (Lurain, 1992; Sullivan, 1993) while mutations induced in these regions leading to a loss of function (Ye, 1993) demonstrate the importance of these conserved domains. In particular, domains I, II and III have been shown to directly participate in binding to deoxynucleoside triphosphates, in chelating Mg2+ ions and in interacting with primer and template (Ye, 1993) while it has been suggested that domains III, V, and δ -region C can also form part of the dNTP- binding site (Ye, 1993). The region involved with interacting with the UL44 subunit has been mapped to C terminal residues 1221-1242 of UL54 and appears to be dependent on hydrophobic interactions (Loregian, 2003).

1.4.2.2 DNA polymerase accessory protein (UL44) Initial studies on the HCMV DNA polymerase (UL54) and its accessory subunit (UL44) were based on the HSV-1 equivalent, UL30 and UL42. UL54 is a functional homolog of the HSV-1 UL30 DNA polymerase, both sharing polymerase activity (Mar, 1981) as well as 3’-5’ exonuclease activity (Nishiyama, 1983) that are dependent on salt concentration (Cihlar, 1997; Weiland, 1994). UL44 is analogous to the HSV-1 processivity factor UL42 (Gottlieb, 1990), functioning to bind double-stranded DNA and stimulating DNA synthesis by the main catalytic subunit UL54 (Appleton, 2004; Ertl, 1992; Loregian, 2004a; Weiland, 1994). The observation that both UL54 and UL44 are required for DNA synthesis and that disruption to either subunit severely inhibits viral DNA replication (Digard, 1993; Pari, 1993a; Ripalti, 1995) provides a strong basis to study the UL54/UL44 interaction. The UL44 binding site on UL54 had previously been elucidated to the C-terminal 22 amino-acids of UL54 with two cysteine residues suggested to play a key role in binding or stabilizing the UL54/UL44 interaction (Loregian, 2004; Loregian, 2004a; Loregian, 2003). Alanine mutants of the C-terminal

17 region in UL54 further identified Leu1227 and Phe1231 as key residues for subunit interaction (Loregian, 2004). Although the HSV-1 homolog interaction between UL30/UL42 had been mapped to an analogous region earlier (Monahan, 1993; Zuccola, 2000), neither UL30 nor UL54 shared any sequence similarity with each other (Loregian, 2005).

Analysis of residues 1-290 from the UL44 crystal structure revealed an overall fold that was similar to other processivity factors including UL42 of HSV-1 and the proliferating cell nuclear antigen (PCNA) of eukaryotic DNA polymerases, despite sharing no obvious sequence similarity (Appleton, 2004). Crystallization of the N-terminal two- thirds of UL44 revealed a C-clamp-shaped head-to-head homodimer (Appleton, 2004) in contrast to UL42, which carries out its functions as a monomer (Gottlieb, 1990; Randell, 2004; Zuccola, 2000). Analysis of the dimer binding region revealed specific residue interactions involving six main-chain-to-main-chain hydrogen bonds and extensive packaging of hydrophobic side chains at the interface (Appleton, 2004). At the binding interface, F121 of each monomer is buried against a hydrophobic loop composed of P85, L86 and L87 of the other monomer. L86 and L87 also pack against M123 and L93 of the opposite monomer (Appleton, 2004). Furthermore, L86 and L87 are conserved as hydrophobic residues in UL44 homologs of other β-herpesvirus (Appleton, 2004), suggesting an importance in specific binding mechanisms. Alanine substitutions of the two lysine residues in the hydrophobic loop disrupted DNA binding by up to 100 fold (Appleton, 2004; Loregian, 2005).

In addition, UL44 shares a ‘connector loop’ structure similar to UL42 which was shown to connect two topologically similar domains of UL42 (Zuccola, 2000) as well as playing a crucial role in interacting with UL30 (Bridges, 2001). Indeed, recent studies have shown that the connector loop in UL44 has functional similarities to its homolog in HSV-1 (Appleton, 2004; Loregian, 2004). Constructed alanine mutants of the connector loop region showed that substitutions that affected residues 133-136 severely reduced physical and functional interaction with UL54 while a substitution at residue I135 disrupted UL54/UL44 binding and long-chain DNA synthesis completely (Appleton, 2004; Loregian, 2004). Similar studies done with the UL42 connector loop identified the glutamine residue at position 171 as crucial for subunit interaction (Bridges, 2001). However, subtle differences between the two systems have also been 18 described, suggesting that despite the similarities in function and analogous positioning, the mechanism of binding is probably different. Firstly, the residues important for binding the two HCMV subunits are hydrophobic while their counterparts in the HSV-1 interaction are hydrophilic (Bridges, 2001; Loregian, 2004; Zuccola, 2000). This is further supported by findings that the accessory subunit of HSV-1 cannot stimulate the catalytic subunit of HCMV, and vice-versa (Loregian, 2005). Secondly, despite the substitution of residue Q171 in UL42, weak binding could still be detected in maltose binding protein (MBP)-pulldown assays (Bridges, 2001), whereas substitution of the I135 residue in UL44 abolished binding completely (Loregian, 2004). The observation that even a single substitution can disrupt key interactions to such an extent has provided some insight into possible anti-viral treatments such as the use of small inhibitory molecules (Loregian, 2004; Pilger, 2004).

Although the crucial functions of UL44 had been mapped to the N-terminal two-thirds of the gene, the region beyond residue 290 has been less well defined functionally. Earlier studies had determined that the C-terminus of the gene was dispensable for UL44 activity (Loregian, 2004a; Weiland, 1994), despite the presence of salient glycine-rich strings (Ertl, 1992). However, more recent studies suggest the C-terminus to be involved in nuclear transport with the key nuclear localization site (NLS) located at residues 425-431 of the C-terminus (Alvisi, 2005). This is consistent with similar studies on nuclear localization sites flanked by protein kinase phosphorylation sites of other homologous herpesviruses (Loh, 1999; Takeda, 2000). Apart from its function as the polymerase processivity factor, UL44 has also been associated with mediating cell adhesion via a putative integrin binding RGD motif that has been studied in murine CMV (Loh, 2000).

1.4.2.3 Single-stranded DNA binding protein (UL57) HCMV UL57 encodes an early single-stranded DNA-binding protein that is homologous to the HSV-1 major DNA-binding protein, ICP8 (Anders, 1996; Pari, 1993). ICP8 has been studied more extensively than UL57 and as such, much of our understanding of UL57 is based on observations derived from studies on ICP8. Studies in the HSV-1 major DNA-binding protein ICP8 have been linked to multiple functions within the HSV replication cycle. Its ability to modulate activity of the HSV-1 DNA

19 polymerase (Hernandez, 1990), affect genome recombination and processing (Bortner, 1993; Dutch, 1993; Nimonkar, 2003), regulate late viral gene expression (Chen, 1996; Gao, 1989; Gao, 1991), organize DNA replication enzymes into nuclear replication compartments (Bush, 1991; de Bruyn Kops, 1988) as well as interactions with the helicase-primase complex (Boehmer, 1998; Boehmer, 1993; Crute, 1991; Falkenberg, 1998; Hamatake, 1997; He, 2001; Lee, 1997; Makhov, 1996) provides an insight into potential functional parallels that may be found with its HCMV counterpart. However, it has already been established that like ICP8, UL57 serves as an essential component for transient complementation in HCMV oriLyt-mediated DNA replication (Pari, 1993; Sarisky, 1996). Physical aspects of the ICP8 protein have also been characterized to a certain extent. Cooperative DNA binding has been mapped to the C-terminus of ICP8 (Dudas, 1998; Mapelli, 2000) with the possible involvement of two cysteine residues at 245 and 455 playing a role in modulation (Dudas, 1998). The DNA binding region has been harder to elucidate with studies mapping regions which range from residues 564- 1160 (Gao, 1989), 300-849 (Wang, 1990), and 368-902 (White, 1999) as potential active sites. Several other functional regions have been mapped including: a nuclear localization signal at the C-terminal 28 amino acid residues (Gao, 1992), an intranuclear localization site located in the region between residues 1080-1135 (Taylor, 2003), and a putative zinc binding motif between residues 499-512 (Gao, 1988; Gupte, 1991).

UL57 is located about 1kb downstream of oriLyt, raising the possibility that UL57 promoter elements contribute to oriLyt function (Kiehl, 2003). Indeed, mapping the 5’ ends of the UL57 transcripts elucidated a 42 base-pair sequence spanning the oriLyt- proximal start site for UL57 that was found to play a role for both UL57 transcription and oriLyt function, albeit not an essential one for the latter (Kiehl, 2003). Furthermore, the 1kb non-coding region between oriLyt and UL57 consists of numerous potential transcriptional control signals, including transcription factor binding sites and polyadenylation signals, all of which have been shown to participate in replicator functions in other herpesvirus systems (Nguyen-Huynh, 1998; Schepers, 1993).

1.4.2.4 Primase-helicase complex (UL70, UL102, and UL105) Despite several studies detailing the functionality of the helicase-primase complex, most of them were done in the context of the herpes simplex virus model (Barnard, 1997;

20 Biswas, 1999; Dracheva, 1995; Graves-Woodward, 1996; Graves-Woodward, 1997; Klinedinst, 1994). Apart from the initial characterization of the HCMV helicase-primase proteins (Pari, 1993; Smith, 1995a; Smith, 1996), there have been relatively few publications on the specific interactions of these proteins. Although studies have shown that the UL70, UL102, and UL105 proteins interact with each other to form a heterotrimeric complex (McMahon, 2002), the specific regions that bind the complex together have not been studied. Based on sequence and positional similarities with the herpes simplex virus counterparts, the helicase protein (UL105) is predicted to be involved in unwinding the DNA helix, while the HCMV primase protein (UL70) is predicted to be involved in initiating DNA synthesis. However, the primase-helicase associated protein (UL102) is not specifically involved in the activity of either, yet it interacts with both UL70 and UL105 to enhance their activities. By elucidating the binding relationship of these three proteins, we hope to understand more about the role of UL102.

The HCMV primase ORF (UL70) encodes a 947 amino acid protein that shares 27% homology with the HSV-1 primase, UL52 (Chee, 1990). Similar to the HSV-1 primase, pUL70 contains several conserved regions including a putative DXD motif that is associated with primase catalysis (Dracheva, 1995; Klinedinst, 1994). Furthermore, substitutions of the aspartate residues in the motif have been shown to abolish primase but not helicase or ATPase activity in vitro (Dracheva, 1995; Klinedinst, 1994). The HSV-1 UL52 protein contains a putative zinc-finger motif at its C terminus that is highly conserved among herpesviruses as well as other prokaryotic, and eukaryotic primases (Ilyina, 1992; Mendelman, 1994). Although the role of the zinc finger is not well defined, alanine substitutions of the third and fourth conserved cysteines resulted in the loss of primase, helicase, ATPase, and DNA-binding activities of the HSV helicase- primase complex (Biswas, 1999). UL70 is predicted to engage in primer synthesis during DNA replication on the basis of its relative homology to the HSV-1 primase, which has been shown to produce short oligoribonucleotide primers up to 10-13 nucleotides long on ssDNA (Crute, 1991; Ramirez-Aguilar, 1995), allowing the DNA polymerase to begin replicating DNA via dNTP polymerization (Gottlieb, 1994; Nimonkar, 2003). In addition, the HSV-1 primase was found to initiate primer synthesis with a purine, at the second nucleotide of a 3'-deoxyguanylate-pyrimidine-pyrimidine

21 (3'-G-pyr-pyr) template sequence (Ramirez-Aguilar, 1995), which may be similar in the HCMV primase.

HCMV also encodes a helicase-primase associated protein (UL102) that shares little sequence homology to its HSV counterpart, UL8 (Chee, 1990), but instead is recognized as a positional homolog (McGeoch, 1988; Pari, 1993). Initial identification of the UL102 ORF revealed a 798 amino acid protein that was essential for origin- dependent DNA replication (Chee, 1990; Pari, 1993). In addition, the UL101 ORF located upstream of UL102 was also shown to be an essential component in replication (Pari, 1993). However, comparison against cDNA and subsequent re-sequencing of UL102 in Towne and AD169 (Smith, 1995a) revealed that the genomic stop codon for UL101 was an error in the original published sequence (Chee, 1990). As a result, UL102 has been redefined as a 2.7kb un-spliced transcript encoding an 873 amino acid protein that has the capacity to encode several smaller proteins, all within frame of the UL102 stop codon (Smith, 1995a). Although further elucidations of the biochemical activities of UL102 have yet to be published, its positional and functional homology to HSV UL8 may provide an insight into potential functional parallels derived from studies on UL8. Firstly, studies looking at primase activity have demonstrated that UL8 may serve a key role in stimulating the synthesis of RNA primers (Tenney, 1994) as well as increasing the efficiency of primer utilization by DNA polymerase (Sherman, 1992). A central segment of UL8 has been attributed to interact with both the HSV helicase (UL5) and primase (UL52) subunits (Barnard, 1997). However, in the absence of UL8, the HSV primase and helicase protein form a sub-complex that is enzymatically indistinguishable from a complex formed by all three proteins (Calder, 1992; Crute, 1991), suggesting that UL8 may play a role in promoting nuclear localization of the complex (Calder, 1992). In addition, UL8 also interacts directly with the origin binding protein (UL9) and at least functionally with the single-stranded DNA binding protein (ICP8) (Hamatake, 1997; McLean, 1994). UL8 has also been hypothesized to help direct the polymerase to the initiation complex, where it coordinates polymerase and helicase-primase activities (Marsden, 1997).

The HCMV helicase ORF (UL105) encodes a 982 amino acid protein that shares 34% sequence homology with the HSV-1 helicase, UL5 (Chee, 1990), and is present in infected cells as early as 24 hrs post-infection (Smith, 1996). Like the HSV-1 helicase, 22 the HCMV helicase protein (pUL105) contains six conserved helicase motifs common to superfamily-1 DNA helicases (Gorbalenya, 1989; Zhu, 1992). Motif I consists of a GxxGxGKT/S Walker A motif associated with binding the di- or triphosphate moiety of the nucleotide cofactor (Walker, 1982) while Motif II consists of a Walker B motif containing a group of hydrophobic residues terminated by an Asp and a Glu residue and is associated with the stabilization of the coordinated Mg2+ ion (Walker, 1982). In HSV, motif I has been shown to be directly involved in ATP binding and/or hydrolysis while motif II appears to be required for coupling of DNA binding to ATP hydrolysis (Graves-Woodward, 1997). The functional significance of motifs III, IV, V, and VI are yet to be fully elucidated but the strong conservation of these six motifs suggests that they may be important for helicase activity. However, mutations generated in motifs III, IV, V, and VI did not eliminate ATP hydrolysis nor affect DNA binding and therefore is postulated to be involved in the coupling of these two activities to the process of DNA unwinding (Graves-Woodward, 1997). Other studies on the HSV helicase have shown that single amino acid substitutions in the most conserved residues of these motifs abolish the ability of the HSV helicase to support DNA replication in vivo suggesting that these conserved residues are essential to DNA replication (Zhu, 1992).

1.4.2.5 Putative Initiator Protein (UL84) HCMV UL84 encodes a 586 amino acid protein that is detected in infected cells as early as 2.5 hrs post-infection (He, 1992). Currently, unlike the other replication genes, UL84 shares no functional or positional homology to any HSV replication gene. Analysis of the UL84 amino acid sequence has revealed the presence of two potential leucine zipper domain in the N-terminal half of the protein, at amino acids 114–135 and at 325-373 (He, 1992). The leucine zipper domain between residues 114-135 has been implicated in the interaction of UL84 with the IE2 protein (Colletti, 2004), an immediate-early protein, that has been identified as one of the auxiliary components required for origin- dependent DNA replication in human fibroblasts (Pari, 1993). Substitutions of the leucine residues in this domain rendered UL84 incapable of complementing oriLyt- dependent DNA replication (Colletti, 2004) while further studies on the functional role of the UL84–IE2 interaction revealed that an over-expression of UL84 interfered with the IE2-mediated transactivation of the UL112/113 promoter, decreasing IE2-mediated transient transactivation (Gebert, 1997). UL84 has also been shown to undergo

23 oligomerization with key interaction domain mapped to residues 151-201 (Colletti, 2004). Initial studies identified UL84 as utilizing a nuclear localization signal similar to that of the simian virus 40 large T antigen and IE2 nuclear localization signals (Xu, 2002), although subsequent studies report that UL84 interacts with at least four members of the α importin protein family in vitro and in vivo to utilize a classic importin-mediated pathway for nuclear import (Lischka, 2003). The region of UL84 responsible for the interaction with α importin has been identified as a complex domain spanning residues 226-508 and containing a cluster of basic amino acids similar to that of the classical nuclear localization signals (NLS), albeit lacking classical NLS activity (Lischka, 2003). This domain also contains two motifs homologous to leucine-rich nuclear export signals (NES) which have been shown to enable UL84 to shuttle between the nucleus and the cytoplasm (Lischka, 2006). There are also evidence to suggest that UL84 may have enzymatic properties that is similar to that of a helicase (Colletti, 2005). Studies have demonstrated UTPase activity in UL84 that has been suggested to be part of an energy-generating system for helicase activity (Colletti, 2005; Davison, 2005). In addition, the amino acid sequence of UL84 displays some structural homology to the DExD/H box family of RNA helicases (Colletti, 2005). However, most importantly are recent implications that UL84 provides the role as an initiator of DNA replication, with several studies demonstrating that UL84 interacts with DNA replication genes UL57 and UL44 (Gao, 2007).

1.4.3 DNA replication Herpesvirus DNA replication is thought to begin with genome circularization and theta form replication which then proceeds to a rolling circle form which has been well documented in HSV (Boehmer, 2003). Initiation of CMV DNA synthesis has yet to be elucidated. However, in general, initiation of DNA synthesis involves targeted unwinding of particular DNA sequences to enable assembly of a replisome. It has been proposed that DNA synthesis in herpesvirus initially proceeds by a theta type mechanism in which DNA replication proceeds bi-directionally from the lytic origin of replication, although this has yet to be seen in vitro (Boehmer, 2003). HCMV has a highly complex oriLyt region that spans over a kilo-base of DNA upstream of the UL57 ORF (Kiehl, 2003). However, unlike HSV, it is the only lytic-phase replicator identified in HCMV (Anders, 1992). Herpesviruses encode a virion-associated transcript that

24 associates with oriLyt to form a three-stranded structure whose precise role in DNA synthesis is yet to be elucidated (Prichard, 1998). In general, binding of the Origin Binding Protein (OBP) with specific sites in OriLyt is followed by an interaction with single stranded DNA binding proteins which lead to localized unwinding and access of replication fork proteins. Here, the replication fork machinery consisting of the DNA polymerase catalytic and accessory subunits, and the hetero-trimeric helicase-primase complex direct continuous, leading strand viral DNA replication in a rolling circle mechanism (Boehmer, 2003).

1.4.4 Capsid assembly, maturation and egress Rolling circle replication generates long head-to-tail concatemers consisting of multiple unit-length genomes that accumulate in the nuclei. These concatemers are cleaved at regular intervals and packaged into individual, pre-formed capsids by a mechanism that is conserved throughout the entire herpesviridae family (Homa, 1997). Structural studies of the HSV capsid have indicated that structural and functional protein homologues of HSV capsid proteins are present in the capsid of HCMV (Gibson, 1996). In addition, cryo-electron microscopic analysis of the capsids has revealed that HCMV and HSV have near identical structures (Butcher, 1998; Chen, 1999) suggesting that the assembly of the HCMV capsid follows a very similar assembly pathway as that of HSV (Grunewald, 2003). The HCMV capsid is composed of 162 capsomeres consisting of 150 hexons and 12 pentons (Butcher, 1998; Chen, 1999) of which the most abundant protein components consist of the major capsid protein (MCP, UL86) and the smallest capsid protein (SCP, UL48–49). Two copies of the minor capsid protein (MnCP, UL85) combined with a single copy of the minor capsid binding protein (MnCP-BP, UL46) form the triplexes that are located between adjacent pentons and hexons (Butcher, 1998; Chen, 1999). The process of capsid assembly involves UL86 and UL48-49 interacting in the cytoplasm in the presence of assembly protein precursor gene UL80a before translocating to the nucleus. UL85 and UL46 are also translocated to the nucleus, albeit separately from the UL86/UL48-49 complex. Once in the nucleus, self-interaction domains in the products of the UL80a lead to formation of pentons and hexons and the generation of the capsid scaffold. The UL86/UL48-49 complex then interacts with UL85 and UL46 to form the immature shell of the capsid (Gibson, 1996). In HCMV, the packaging of viral DNA involves a portal protein, proposed to be encoded by

25 UL104 (Komazin, 2004), and is mediated through virus encoded protein recognition of two conserved sequence motifs, the pac-1 and pac-2 sequences, that are located at each end of the viral genome (Mocarski, 2001). In HSV, this process is also coupled with capsid maturation (Heymann, 2003), which involves the proteolytic cleavage of the carboxyl terminal of the MCP binding domain (Gibson, 1996). The process of HCMV egress is not well known, but is thought to involve an initial envelope obtained at the inner nuclear membrane that is lost by a fusion event at the outer nuclear membrane, releasing free nucleocapsids into the cytoplasm, where egress continues through a second and final envelopment step as the virion leaves the host cell (Mettenleiter, 2004).

1.4.5 Latency and reactivation A unique biological property common to all herpesviruses, including HCMV, is the ability of the virus to establish lifelong persistence within the host following initial infection. In general, HCMV reactivation from latency is observed in 13% of healthy adults (Ling, 2003), and though episodes of sporadic reactivation may occur, they are generally well-controlled by cell-mediated immuno-surveillance. However, as with primary infection, reactivation within an immunocompromised individual can lead to high levels of morbidity and mortality. Recent studies have implicated endothelial cells and specific cell types of the myeloid lineage as sites of HCMV persistence and latency (Jarvis, 2002). The factors leading to reactivation from latency are not completely understood, although studies have shown that myeloid differentiation can stimulate the production of viral gene expression and in some cases, the production of infectious virus (Reeves, 2005; Soderberg-Naucler, 2001; Taylor-Wiedeman, 1994). In addition, the production of the stress hormones cortisol, adrenocorticotrophic hormone, epinephrine and norepinephrine has been implicated in the reactivation and shedding of CMV in urine (Mehta, 2000).

1.5 Aims and Objectives

There is a fundamental need for understanding the role of other HCMV genes essential for DNA replication as a foundation for developing new antiviral treatments. In this study we looked at six HCMV replication genes encoding for the DNA polymerase accessory protein (UL44), single stranded DNA binding protein (UL57), primase

26 (UL70), helicase (UL105), primase-helicase associated protein (UL102), and the putative initiator protein (UL84) in order to increase our understanding of their role in DNA replication (Figure 3). In particular, UL70 is an essential component of the primase-helicase complex and elucidation of its functional domains is essential for understanding its role in DNA replication and subsequently, ways to inhibit its activity as a potential drug target. Hence, the aims of this project are to:

1. Identify strain variation within the UL44, UL57, UL70, UL102, UL105, and UL84 genes. 2. Predict putative functional domains and motifs for these genes using the ScanProsite software, and 3. Express and study the UL70 (Primase) gene through the use of recombinant mutants.

Figure 3 Relative positions of genes investigated in this study

27 2 Reagents and equipment

All reagents and equipment used in the study are detailed in this section. Individual experimental methods are outlined in their corresponding chapters. In general, most chemicals used were molecular biology grade and solutions sterilized by autoclaving or filtration. Processing of biological specimens was carried out in Class II cabinets using sterile techniques. In addition all purchased reagents were tested sterile and pyrogen- free by the manufacturers. Biological waste was sterilized by autoclaving prior to disposal.

2.1 Viral stocks

HCMV low- Clinical isolates 44A, 70A, 77A, 90A, 91A, 16B, 21B, 80B, 4E, passage strains and 30E were obtained from the Prince of Wales Hospital.

2.2 DNA extraction

2.2.1 Buffers

T10E50 T10E50 buffer consisted of 10mM Tris-HCl (pH 8.0), 50mM EDTA in buffer MilliQ water and autoclaved before use. TE buffer TE buffer consisted of 10mM Tris-HCl (pH 8.0), 1mM EDTA in MilliQ water and autoclaved before use.

2.2.2 Chenicals Chloroform Chloroform was obtained from Sigma-Aldrich (USA). Ethanol 99.5% ethanol was obtained from Sigma-Aldrich (USA). NaCl >98% (titration) sodium chloride for molecular biology was obtained from Sigma-Aldrich (USA). Phenol Phenol was obtained from Sigma-Aldrich (USA).

28 2.2.3 Kits QIAamp DNA QIAamp DNA mini kits (Qiagen, USA) were used to extract DNA Mini Kit from frozen cell-associated viral stock using a QIAamp spin-column procedure.

2.3 Polymerase chain reaction (PCR)

2.3.1 Buffers ExpandLong 10X Expand Long Template Buffer 1 was obtained from Roche template reaction (Germany) and is supplemented with 17.5mM MgCl2. buffer 1 GoTaq green reaction 5X GoTaq green reaction buffer was obtained from buffer Promega (USA) and is supplemented with 7.5mM MgCl2. Platinum Taq High 10X Platinum Taq high fidelity PCR buffer was obtained from

Fidelity PCR buffer Invitrogen (USA) and consisted of 600mM Tris-SO4 buffer (pH 8.9), 180mM ammonium sulfate. Red Hot reaction 10X Red Hot reaction buffer was obtained from AB gene (UK) buffer and consisted of 750mM Tris-HCl (pH 8.8), 200mM

(NH4)2SO4, 0.1% (v/v) Tween 20.

2.3.2 Enzymes ExpandLong template Expand Long Template enzyme mix was obtained from enzyme Roche (Germany) and stored at -20°C before use. GoTaq Flexi DNA GoTaq Flexi DNA polymerase was obtained from polymerase Promega (USA) and stored at -20°C before use. Platinum Taq High Fidelity Platinum Taq DNA polymerase was obtained from DNA Polymerase Invitrogen (USA) and stored at -20°C before use. Red Hot DNA polymerase Red Hot DNA polymerase was obtained from ABgene (UK) and stored at -20°C before use.

29 2.3.3 Equipment Thermal PTC 200 Peltier Thermal cyclers (GMI, USA) were used for cycler amplification of DNA fragments.

2.3.4 Reagents dNTP Deoxynucleoside triphosphates, dATP, dCTP, dGTP and dTTP, each at stock concentrations of 100mM (Promega, USA) were diluted 1:10 and 1:100 with sterile water (Baxter, Australia) to a final concentration of 10mM and 1mM respectively. Diluted dNTP stocks were stored at -20°C.

MgCl2 25mM magnesium chloride was obtained from AB gene (UK) or Promega (USA) and stored at -20°C before use.

MgSO4 50mM magnesium sulfate was obtained from Invitrogen (USA) and stored at - 20°C before use. Primers Primers were synthesized by Sigma Genosys (Australia) as 400µM stock solutions. These were further diluted to 10µM with sterile water (Baxter, Australia) and stored at -20°C before use.

2.4 Electrophoresis

2.4.1 Buffers TBE 10X TBE consisted of Trizma base (Sigma-Aldrich, USA), boric acid (Sigma- buffer Aldrich, USA) and EDTA (Sigma-Aldrich, USA) in MilliQ water and autoclaved. 0.5X working stock made by diluting 1:20 in MilliQ water.

30 2.4.2 DNA Markers 1kb step-ladder 0.3µg/µl 1kb step-ladder markers were obtained from Promega marker (USA), diluted 1:10 with sterile water (Baxter, Australia) and stored at -20°C before use. BenchTop pGEM BenchTop pGEM markers were obtained from Promega (USA) DNA markers and stored at room temperature. Lambda/HindIII Lambda/HindIII markers were obtained from Promega (USA), marker diluted 1:10 with sterile water (Baxter, Australia) and stored at - 20°C before use.

2.4.3 Equipment

Electrophoresis The Mini-Sub Cell GT System (Biorad, USA) was used to Equipment run 2D agarose gels.

2.4.4 Reagents Agarose Agarose (analytical grade) was obtained from Promega (USA) and used for gels at concentrations of 0.75% to 1.5% depending on PCR product size. Ethidium 10mg/ml ethidium bromide was obtained from Sigma-Aldrich (USA) bromide and diluted in sterile water (Baxter, Australia) to a final concentration of 5mg/ml. This was further diluted to 0.5mg/ml in agarose gels and TBE running buffer. SYBR safe 10,000X SYBR Safe gel stain was obtained from Invitrogen (USA) and DNA gel stored at room temperature. stain

31 2.5 DNA sequencing

2.5.1 Equipment DNA A 3730 DNA Analyzer (Applied Biosystems, USA) was used for sequencer analysis of sequencing products at the Ramaciotti Centre, University of New South Wales.

2.5.2 Reagents PEG PCR Polyethylene glycol PCR mix consisted of 26.7% (w/v) polyethylene mix glycol (PEG) 8000 (Sigma-Aldrich, USA), 0.6M NaOAc (pH 5.2),

6.5mM MgCl2. Primers Primers were synthesized by Sigma Genosys (Australia) as 400µM stock solutions. These were further diluted to 10µM with sterile water (Baxter, Australia) and stored at -20°C before use. Sequencing BigDye Terminator v3.1 was obtained from Applied Biosystems (USA) mix and stored at -20°C before use. Sequencing 5X sequencing buffer was obtained from Applied Biosystems (USA) buffer and stored at -20°C before use.

2.6 Restriction enzymes

2.6.1 Buffers CIAP stop CIAP stop buffer consisted of 10mM Tris-HCl (pH 7.5), 1mM EDTA buffer (pH 7.5), 200mM NaCl, and 0.5% (w/v) SDS. Multi-core 10X multi-core buffer was obtained from Promega (USA) and consisted buffer of 10mM Tris-HCl (pH 7.5), 100mM NaCl, 50% glycerol.

32 2.6.2 Enzymes BamHI BamHI was obtained from New England BioLabs (USA) with NEBuffer 3 and stored at -20°C before use. HindIII HindIII was obtained from Promega (USA) with buffer B and stored at -20°C before use. SalI SalI was obtained from New England BioLabs (USA) with NEBuffer 3 and stored at -20°C before use. SpeI SpeI was obtained from Promega (USA) with buffer B and stored at -20°C before use.

2.6.3 Reagents BSA 1µg/µl Bovine serum albumin (BSA) was obtained from Promega (USA) and stored at -20°C before use. CIAP Calf intestinal alkaline phosphatase (CIAP) was obtained from Promega (USA) and stored at -20°C before use. TSAP Thermosensitive alkaline phosphatase (TSAP) was obtained from Promega (USA) and stored at -20°C before use.

2.7 Cloning reagents

2.7.1 Antibiotics Ampicillin 91.0-100.5% Ampicillin was obtained from Sigma-Aldrich (USA), diluted with sterile water (Baxter, USA) to a working concentration of 10mg/ml, filtered and stored at -20°C before use. Tetracycline 100mg/ml Tetracycline was obtained from Invitrogen (USA) and stored at -20°C before use. Zeocin 100mg/ml Zeocin was obtained from Invitrogen (USA) and stored at - 20°C before use.

33 2.7.2 Cells

Library efficiency DH5α DH5α competent cells were obtained from Invitrogen competent cells (USA) and stored at -80°C before use. XL10 Gold Ultracompetent XL10 Gold ultracompetent cells were obtained from cells Stratagene (USA) and stored at -80°C before use.

2.7.3 Equipment Low-salt LB Low salt LB consisted of 1% (w/v) tryptone (Oxoid, Australia), 0.5% zeocin plates NaCl (Sigma-Aldrich, USA), 0.5% yeast (Oxoid, Australia), 1.5% (w/v) agar (Promega, USA), adjusted to pH 7.5 and autoclaved. Allowed to cool to 55°C and zeocin (Invitrogen, USA) added to final concentration of 50µg/ml before pouring. LB 2% (w/v) agar (Promega, USA) added to LB broth and autoclaved. Ampicillin Allowed to cool to 55°C and Ampicillin (Sigma-Aldrich, USA) added plates to final concentration of 100µg/ml before pouring. LB 2% (w/v) agar (Promega, USA) added to LB broth and autoclaved. tetracycline Allowed to cool to 55°C and tetracycline (Sigma-Aldrich, USA) added plates to final concentration of 125µg/ml before pouring.

34 2.7.4 Kits DNA ligation kit DNA ligation kit was obtained from Stratagene (USA) and stored at -20°C before use. Includes pUC18 (BamH I digested) and cI857 wild-type lambda (Hind III digested) control DNA, 10mM rATP (pH 7.5), T4 DNA ligase, and 10X ligase buffer consisting of

500mM Tris-HCl (pH 7.5), 70mM MgCl2, 10mM dithiothreitol (DTT). Gateway BP BP Clonase II kit was obtained from Invitrogen (USA) and stored Clonase II at -20°C before use. enzyme mix Wizard Plus Wizard Plus midiprep kits (Promega, USA) were used to extract Midipreps DNA plasmid DNA using a column and vacuum procedure. Reagents purification were stored at room temperature. system Wizard Plus Wizard Plus miniprep kits (Promega, USA) were used to extract Minipreps DNA plasmid DNA using a column and vacuum procedure. Reagents purification were stored at room temperature. system Wizard SV Wizard PCR clean-up kits (Promega, USA) were used to purify Gel/PCR Clean- restriction digested DNA. Reagents were stored at room up System temperature.

35 2.7.5 Media Luria Luria broth (LB) consisted of 1% (w/v) NaCl (Sigma-Aldrich, USA), 1% broth (w/v) tryptone (Oxoid, Australia), 0.5% (w/v) yeast (Oxoid, Australia), adjusted to pH 7.0, and autoclaved prior to use. NZY NZY broth consisted of 1% (w/v) casein hydrolysate (Merck, USA), 0.5% broth (w/v) yeast (Oxoid, Australia), and 0.5% (w/v) NaCl (Sigma-Aldrich, USA),

adjusted to pH 7.5 and autoclaved. Add 12.5ml of 1M MgCl2 (Sigma, USA),

12.5ml of 1M MgSO4 (BDH, UK), and 10ml of 2M glucose (Gibco BRL, USA) per liter of broth prior to use. S.O.C S.O.C media consisted of 2% (w/v) tryptone (Oxoid, Australia), 0.5% (w/v) media yeast (Oxoid, Australia), 10mM NaCl (Sigma-Aldrich, USA), 2.5mM KCl

(BDH, UK), 10mM MgCl2 (Sigma, USA), 10mM MgSO4 (BDH, UK), 20mM glucose (Gibco BRL, USA), and autoclaved before use.

2.7.6 Reagents IPTG 100mM Isopropyl B-D-Thiogalactopyranoside was obtained from Promega (USA). Proteinase 2µg/µl Proteinase K was obtained from Invitrogen (USA) and stored at - K 20°C before use. X-gal 50mg/ml 5-bromo-4-chloro-3-indolyl-B-D-galactosidase was obtained from Promega (USA).

2.7.7 Vectors pBluescript II pBluescript II SK(+) vectors were obtained from Stratagene (USA) SK(+) and stored at -20°C before use. pCITE-4A(+) pCITE-4A(+) vectors were obtained from Novagen (Germany) and stored at -20°C before use. pDONR/Zeo pDONR/Zeo vectors were obtained from Invitrogen (USA) and stored at -20°C before use.

36 2.8 Protein expression and detection

2.8.1 Antibodies Anti-His G Antibody Anti-His G antibodies (Invitrogen, USA) were used for protein detection and stored at -20°C before use. S-Protein alkaline S-Protein alkaline phosphatase conjugate (S-Tag) was phosphatase conjugate obtained from Novagen (Germany) and stored at -20°C before use.

2.8.2 Buffers Loading 4X loading buffer consisted of 0.25M Tris-Cl (Sigma-Aldrich, USA), 40% buffer (w/v) glycerol, 8% (w/v) SDS, 0.4% (w/v) Bromophenol Blue (BDH, UK), 8% (v/v) β-mercaptoethanol (Sigma-Aldrich, USA) PBS 10X phosphate buffered solution (PBS) were obtained from Gibco (USA). Running 1X running buffer consisted of 25mM Trizma base (Sigma-Aldrich, USA), buffer 250mM (pH 8.3) glycine (Sigma-Aldrich, USA), and 0.1% (w/v) SDS in MilliQ water. Transfer 1X transfer buffer consisted of 25mM Trisma base (Sigma-Aldrich, USA), buffer 250mM (pH 8.3) glycine (Sigma-Aldrich, USA), and 20% methanol (Mallinckrodt Chemical, UK) in MilliQ water.

37 2.8.3 Equipment 2-D gel The Mini-PROTEAN 3 Cell electrophoresis system (Biorad, electrophoresis USA) was used to run 2D poly-acrylamide gels. equipment Nitrocellulose Hybond-ECL nitrocellulose was obtained from Amersham Biosciences. Resolving gel 10% sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) resolving gels consisted of 34% (v/v) 30% Bis/Acrylamide, 0.39M Tris (pH 8.8), 0.1% SDS, 0.1% APS, and 0.04% TEMED. Stacking gel 5% sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) stacking gels consisted of 16.5% (v/v) 30% Bis/Acrylamide, 0.125M Tris (pH 6.8), 0.1% SDS, 0.1% APS, and 0.1% TEMED.

2.8.4 Kits TNT Quick Coupled TNT Quick Coupled Transcription/Translation Transcription/Translation System (Promega, USA) was used to express protein System utilizing the pCITE expression vector.

2.8.5 Protein Markers Kaleidoscope marker Precision Plus Protein Kaleidoscope Standards were obtained from Biorad (USA) and stored at -20°C before use. Perfect Protein Perfect protein western markers were obtained from Novagen Western markers (Germany) and stored at -20°C before use.

38 2.8.6 Reagents APS >98% ammonium persulfate (APS) was obtained from Sigma-Aldrich (USA) and stored at room temperature. Bis/Acrylamide 30% Acrylamide/Bis Solution (37.5:1) was obtained from Biorad (USA) and stored at 4°C before use. Gelcode blue stain Gelcode blue stain (Pierce, USA) was used to stain PAGE gels for total protein. Glycerol >99% glycerol was obtained from Sigma-Aldrich (USA) Ponceau S 0.1% (w/v) Ponceau S in 5% acetic acid (Sigma-Aldrich, USA) was used to stain nitrocellulose membranes for total protein. SDS >98.5% Sodium dodecyl sulfate (for molecular biology, .98.5%) was obtained from Sigma-Adrich (USA). SuperSignal West Pico SuperSignal West Pico Chemiluminescent Substrate kit chemiluminescent substrate was obtained from Pierce (USA) and stored at 4°C before use. TEMED >99% N, N, N’, N’-Tetramethylethylenediamine was obtained from Sigma-Aldrich (USA) and stored at room temperature.

2.8.7 Solutions Blocking Blocking solution consisted of 5% skim milk powder (Diploma, solution Australia) in PBST. PBST Phosphate buffered solution with Tween (PBST) consisted of 1X PBS, solution and 0.1% Tween (Sigma-Aldrich, USA)

39 3 Identification of putative functional motifs in viral proteins essential for human cytomegalovirus DNA replication

3.1 Introduction

HCMV DNA replication is associated with eleven loci necessary for transient complementation of oriLyt-dependent DNA replication (Pari, 1993). These include six core genes homologous to six of the seven essential replication genes in Herpes Simplex Virus (HSV) (Anders, 1996) (Table 6). In HCMV, they consist of a two-subunit DNA polymerase (UL54 and UL44), single-stranded DNA binding protein (UL57), primase (UL70), helicase (UL105), and the primase-helicase associated factor (UL102). In addition, five other loci (TRS1/IRS1, IE1/IE2, UL112-113, UL84, and UL36-38) have supporting regulatory and trans-activation roles in HCMV replication (Iskenderian, 1996; Pari, 1993). DNA replication in HCMV is considered to be analogous to that in HSV, with the function of several HCMV proteins being predicted on the basis of sequence and functional homology with their HSV counterparts (Anders, 1996; Chee, 1990).

Table 6 Essential DNA replication proteins in HCMV and HSV-1 Function HCMV HSV1 DNA Polymerase UL54 UL30 DNA Polymerase accessory protein UL44 UL42 Single-Stranded DNA Binding Protein UL57 UL29 Primase UL70 UL52 Helicase UL105 UL5 Primase-Helicase Associated Protein UL102 UL8 Initiator Protein UL84 UL9

The role of the HCMV DNA polymerase accessory protein (UL44) has been well documented over the years with the elucidation of key functional motifs and residues (Appleton, 2004; Loregian, 2004) as well as a partial crystal structure of the functional N terminal half of the protein (Appleton, 2006; Appleton, 2004). In contrast, the HCMV single stranded DNA binding protein (UL57) has been less well elucidated although early studies established biochemical properties that were comparable to that of the

40 HSV major DNA-binding protein, ICP8 (Anders, 1987). The HCMV primase (UL70), helicase (UL105) and primase-helicase associated factor (UL102) form a trimeric complex (McMahon, 2002) that shares biologically important motifs with the HSV primase-helicase complex (Dracheva, 1995; Klinedinst, 1994; Zhu, 1992). The UL84 ORF encodes the only non-core replication protein that is required for origin-dependent DNA replication in human fibroblasts (Sarisky, 1996). Although initially classed as a regulatory protein for HCMV replication (Pari, 1993), recent functional studies (Colletti, 2004; Xu, 2004) have suggested that the UL84 ORF encodes the initiator protein for HCMV DNA synthesis (Xu, 2004a).

The residues comprising a protein are not all equally important. Some serve as indispensable mediators of protein function while others can be readily replaced without impacting on functionality. As the sequences of biologically significant residues are evolutionarily conserved, conservation analysis has become one of the most widely used techniques for predicting protein function. In this study we report our findings on the variation of the HCMV replication proteins amongst isolate strains.

3.2 Methods

3.2.1 Viral stocks and DNA extraction Viral DNA from 10 low-passage isolates (44A, 70A, 77A, 90A, 91A, 16B, 21B, 80B, 4E, and 30E) were extracted from frozen cell-associated viral stock using the QIAamp DNA Mini Kit (Qiagen, USA) as per manufacturer’s instructions, with DNA eluted in 50µl of buffer AE after 5 min incubation instead of the recommended 200µl.

3.2.2 Polymerase chain reaction (PCR) amplification Full-length ORFs encoding the accessory protein (UL44), helicase-primase complex, (UL70, UL102, and UL105), single-stranded DNA-binding protein (UL57), and putative initiator protein (UL84) were amplified in overlapping segments from viral

DNA by single rounds of PCR. PCR reaction mixes consisted of 1x buffer, 2mM MgCl2, 0.25mM deoxynucleoside triphosphates (dNTP), 0.2µM forward and reverse primers (Table 7), 1 U/µl Red Hot DNA polymerase (AB Gene), and 5µl of extracted DNA template (approximately 100ng). Single round cycling conditions were 94°C for 5 min 41 followed by 30 cycles of 94°C for 1 min, 55°C for 30 sec, and 72°C for 2 min. PCR products were visualized following electrophoresis on 1.5% agarose gels.

42 Table 7 Primers used for PCR amplification of HCMV DNA fragments Gene Primer** Region Amplified* Sequence (5' to 3') UL44 UL44.1 [F] 172843 - 173547 ATGGATCGCAAGACGCGCCT UL44.2 [R] GACGGCGCAATTGAGCAGCG UL44.3 [F] 173438 - 174199 GCGAGCTGGAATTCACGGCC UL44.4 [R] TTTCTCCATCACGGGACCGCG UL57 UL57.1 [F] 138815 - 140062 CATCACGCTATTTTCGCGGGC UL57.2 [R] CGGTGATCGGTTGCGTTGGTC UL57.3 [F] 139960 - 141135 TGGCCTTCAGTTTTGCCTCGG UL57.4 [R] AATAGGCGTAGGCTAGCAGCG UL57.5 [F] 141020 - 142123 TTCGGAGGTGCGCCTCAAGAG UL57.6 [R] CCGTTGCCCGAAAAGTAGCCG UL57.7 [F] 141978 - 142993 ACTGTGGCGCGTGAATCGTTG UL57.8 [R] TTCGGGGCTGTTGCTGCTGTT UL70 UL70.1 [F] 125951 - 126606 GCTGGGCCAGCTGCATCGTGCCGGCGCGACG UL70.2 [R] GCGAGCCCAGTAGCAGACGCGCGAA UL70.3 [F] 126497 - 128309 CATCTCCACCGTGGAGGAGTACGTG UL70.4 [R] GGGGCAGATTTTGTCGCGTACGCTG UL70.5 [F] 128221 - 128853 CCCAACTGGATCTGCGTAACCTGCT UL70.6 [R] GGCGGGTCACCGGCGCCGTGGAAAGTGAGGC UL102 UL102.1 [F] 147750 - 148466 GCTCACGACGCGGTTTGAGCA UL102.2 [R] ATACTTGTGAGTGCCGACGCG UL102.3 [F] 148341 - 148966 CGTGTGGTCACGCATGCCGCGGAAC UL102.4 [R] GCCGTGGCATTGAGACGCACGGCGC UL102.5 [F] 148896 - 149960 GTGCCCGAGGATGAGTGGCAGGTCT UL102.6 [R] GAGGATCTCGCGATACACGGCTTCC UL102.7 [F] 149870 - 150480 CGCGGCCGGTGGCGACTGGCTCTCG UL102.8 [R] TACCCACGTAGTTCCCCTACGTGACTCG UL105 UL105.1 [F] 153111 - 153909 AGAGCAGCTGCTTGCGCAGCA UL105.2 [R] CATTTGTCCACGATGTCGGCG UL105.3 [F] 153748 - 155593 CTTCGTCAGCAAGCACGTGCCGCTG UL105.4 [R] CTGCGAGCTGATCTGACAGTTGGCC UL105.5 [F] 155478 - 156130 TCAAGCGCTACCAGCTCATGCAGCG UL105.6 [R] GTCACGTTTTCCTTACACGGTGTTGTG UL84 UL84.1 [F] 108076 - 109175 AGAAGGGCGACGCGCTATGCG UL84.2 [R] TCTTCTTGCGACGTCGCGGGG UL84.3 [F] 109101 - 110188 TCTGCTCTCTACGCCGCTGCA UL84.4 [R] ACGTGGCGCCATTCTCGTCGC UL84.5 [F] 110115 - 111151 GCCCGTTACGGTCTGGCTACC UL84.6 [R] CCGACGGGTAGTGGTGCACGT *Given in relation to AD169 for UL44, UL57, and UL70, Merlin for UL102 and UL105, and Towne for UL84. **[F] denotes forward primer and [R] denotes reverse primer

43 3.2.3 DNA sequencing and analysis All PCR products were precipitated from solution with polyethylene glycol as previously described (Craxton, 1991). Sequencing was carried out using a reaction mix consisting of Big Dye Terminator (Applied Biosystems, USA), premix buffer, 0.25µM forward or reverse primer (Table 7), and approximately 50ng/300bp of DNA. Cycling conditions consisted of 25 cycles of 96°C for 10 sec, 50°C for 5 sec, and 60°C for 4 min. Sequencing products were purified and analyzed at the University of New South Wales DNA analysis facility. Forward and reverse sequences were compared for sequencing accuracy using GAP (GCG) (Accelrys).

3.2.4 Phylogenetic analysis Full length nucleotide sequences for UL44, UL70, UL102, UL105, UL57, and UL84 were compiled for each individual strain, translated to protein using Translate (GCG) (Accelrys) and aligned with three laboratory strains (AD169, Towne, and Merlin) using ClustalW (Thompson, 1994). Genetic distances between pairs of nucleotide and amino acid sequences were calculated with DNAdist and Protdist programs (Felsenstein, 1989) respectively. Homologous proteins from other virus or cellular systems were also identified using BLASTP (Altschul, 1997) and the results of the searches aligned to their respective CMV counterpart using ClustalW (fast) (Thompson, 1994). All of these programs were accessed through the Australian National Genomic Information Service (ANGIS). The PROSITE database of biologically meaningful motif descriptors derived from multiple alignments and ProRule (Hulo, 2006; Sigrist, 2005) were used to scan the sequences of each HCMV replication protein. Potential motifs were identified using the ScanProsite program (de Castro, 2006) from the Expert Protein Analysis System (ExPASy) proteomics server of the Institute of Bioinformatics.

3.2.5 Genbank Accession Numbers Sequences derived from each of the HCMV clinical isolates were submitted to Genbank and have the accession numbers: UL44 (EU294431 - EU294440), UL57 (EU294441 - EU294450), UL70 (EU294451 - EU294460), UL84 (EU294461 - EU294469), UL102 (EU294470 - EU294479), and UL105 (EU294480 - EU294489).

44 3.3 Results

3.3.1 Strain variation amongst HCMV isolates

The DNA polymerase accessory protein (UL44), single-stranded DNA binding protein (UL57), primase (UL70), helicase (UL105), primase-helicase associated factor (UL102), and putative initiator protein (UL84) are highly conserved among the ten HCMV clinical isolates and three reference strains (AD169, Towne, and Merlin). All six genes studied were found to be highly conserved with nucleotide (nt) identity ranging from 97.9% to 100% and amino acid (aa) identity ranging from 97.46% to 100% (Table 8). Major alterations of the coding sequence, such as frame-shifts or premature stop codons, were not observed. Amino acid polymorphisms in UL44, UL57, UL70, UL102, UL105, and UL84 are summarized in Table 8. Polymorphisms existed outside of known functional domains with the exceptions of a K430R substitution in the nuclear localization signal (NLS) of UL44 as well as various polymorphisms in the NLS of UL84. Results of phylogenetic analysis showed no preferential clustering of isolates although certain amino acid changes were linked in UL84 and UL44. In particular, a S296G substitution was consistently associated with an I319L substitution in UL44 (AD169, Towne, 44A, 70A, 77A, and 4E) while three isolates (16B, 21B, and 30E) had a S27T substitution in UL84 which correlated with A390T substitutions further towards the carboxy terminal end of the protein.

Table 8 Sequence variation within HCMV replication genes Gene Protein Nucleotide Amino Acid Amino Acid Length Identity (%) Identity (%) Variations UL44 433 98.1 - 100 98.6 - 100 18 UL57 1235 99.1 - 100 99.7 - 100 16 UL70 946 98.6 - 100 99.1 - 100 36 UL105 956 99.0 - 100 99.0 - 100 31 UL102 875 98.6 - 100 99.2 - 100 39 UL84 587 97.9 - 100 97.5 - 100 73

3.3.2 Prediction of potential motifs No motifs were detected in the HCMV replication genes when ScanProsite was designated to exclude motifs with a high probability of occurrence. However, this also tended to omit documented motifs which had previously been shown to be biologically significant (Alvisi, 2005; Colletti, 2004; Graves-Woodward, 1997; Lischka, 2006). By 45 contrast, the inclusion of motifs with a high probability of occurrence generated a range of motifs associated with post-translational modifications as well as several domains unique to specific genes (Table 9). In addition, several compositionally biased regions (regions consisting of a high proportion of one amino acid type) that are often abundant in certain motifs were also observed. For example, glycine-rich domains were evident for UL44, UL57 and UL102. A number of conserved functionally significant motifs were identified by comparison of the HCMV proteins with other Herpesvirus homologues, and related viral and cellular proteins discovered by the BLASTP search. Excluding HCMV, we were able to identify 12 homologues for UL44, 34 for UL57, 29 for UL70, 12 for UL102, 32 for UL105, and 4 for UL84.

Table 9 Summary of the putative motifs identified in the HCMV replication genes Gene Motif UL44 UL57 UL70 UL102 UL105 UL84 N-linked glycosylation site 4 3 3 1 9 4 cAMP-dependent protein kinase phosphorylation site 1 0 1 0 2 1 Casein kinase II phosphorylation site 5 14 20 8 15 12 Protein kinase C phosphorylation site 10 15 10 7 10 12 Amidation site 1 1 1 0 0 0 N-myristoylation site 27 42 5 22 11 9 Tyrosine kinase phosphorylation site 0 1 2 2 2 0 Cell attachment sequence 1 0 0 1 0 0 Leucine zipper 0 0 0 0 0 1 Microbodies targeting signal 0 1 0 0 0 0 ATP/GTP-binding site motif A (P-loop) 0 0 0 0 1 0 Total 49 77 42 41 50 39

3.3.3 Strain variation and identification of protein motifs in UL44 Sequence alignments of the ten HCMV strains show a highly conserved protein with variations outside of essential UL44 domains such as the DNA polymerase connector loop (Loregian, 2004) and dimer binding domains (Appleton, 2004). A lysine to arginine substitution was observed at residue 430 of 4E in a region that was previously identified as a nuclear localization signal (NLS) (Alvisi, 2005). However, the significance of the substitution is unknown although both amino acids share similar basic properties. Interestingly, a serine to glycine substitution at residue 296 of AD169, Towne, 44A, 70A, 77A, and 4E correlated with an isoleucine to leucine substitution in the same isolates (Figure 4). An unusual integrin binding motif was identified at residue

46 274 for HCMV, but was not present in other homologues except pongine Herpesvirus 4. More than half of the potential motifs identified for UL44 by ScanProsite (Table 9) were found in the carboxy-terminal third of the gene (codons 290-433), including a putative casein kinase-2 (CKII) signal thought to enhance the carboxy-terminal NLS of UL44 (Alvisi, 2005). In addition, half of the motifs identified by ScanProsite also occurred in the glycine rich region of UL44 (codons 284-397).

47 α δ α δ α δ γ UL44 1 11 21 31 41 51 61 71 81 91 Consensus MDRKTRLSEPPTLALRLKPYKTAIQQLRSVIRALKENTTVTFLPTPSLILQTVRSHCVSKITFNSSCLYITDKSFQPKTINNSTPLLGNFMYLTSSKDLT AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... V...... 77A ...... 90A ...... H.. 91A ...... 16B ...... 21B ...... 80B ...... 4E ...Q...... 30E ......

δ α 101 111 121 131 141 151 161 171 181 191 Consensus KFYVQDISDLSAKISMCAPDFNMEFSSACVHGQDIVRESENSAVHVDLDFGVVADLLKWIGPHTRVKRNVKKAPCPTGTVQILVHAGPPAIKFILTNGSE AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

48 γ δ γ ε δ η 201 211 221 231 241 251 261 271 281 291 Consensus LEFTANNRVSFHGVKNMRINVQLKNFYQTLLNCAVTKLPCTLRIVTEHDTLLYVASRNGLFAVENFLTEEPFQRGDPFDKNYVGNSGKSRGGGGGSGSLS AD169 ...... G.... Merlin ...... Towne ...... G.... 44A ...... I...... G.... 70A ...... G.... 77A ...... G.... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... G.... 30E ......

η η η ζ η δ β η δ γ δ η η η 301 311 321 331 341 351 361 371 381 391 Consensus SLANAGGLHDDGPGLDNDIMNEPMGLGGLGGGGGGGGKKHDRGGGGGSGTRKMSSGGGGGDHDHGLSSKEKYEQHKITSYLTSKGGSGGGGGGGGGGLDR AD169 ...... L...... Merlin ...... Towne ...... L...... 44A ...... L...... 70A ...... L...... 77A ...... L...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... L...... S...... 30E ......

49 η γ δ 401 411 421 431 Consensus NSGNYFNDAKEESDSEDSVTFEFVPNTKKQKCG AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... R.. 30E ...... Figure 4 HCMV strain alignments of UL44

Residues highlighted in yellow indicate functionally significant amino acids (Alvisi, 2005; Appleton, 2004; Loregian, 2004) while underlined sequences represent putative motifs identified with ScanProsite. α = N-linked glycosylation site (N-x-T/S-x), β = cAMP-dependent protein kinase phosphorylation site (RR/KK/RK-x-S/T), γ = Casein kinase II phosphorylation site (S/T-xx-D/E), δ = Protein kinase C phosphorylation site (S/T-x-R/K), ε = Cell attachment sequence (R-G-D), ζ = Amidation site (x-G- RR/KK/RK), η = N-myristoylation site (G-A/N/C/Q/G/I/L/M/S/T/V-xx-all except P).

50 3.3.4 Strain variation and identification of protein motifs in UL57 Although no well defined functional domains exist in UL57, sequence alignments indicate a highly conserved protein with very little variation between strains (Figure 5). An unusual micro-bodies targeting signal was identified at the extreme C-terminus of UL57 (residues 1233-1235) while two short glycine-rich domains spanning residues 536-591 (score = 18.78) and 1167–1226 (score =13.29) were also observed. An alignment of the UL57 protein and its 34 homologues identified several highly conserved cysteine residues at amino acid positions Cys107, Cys483, Cys472, and Cys475, with the latter two forming a putative canonical disulfide bond motif that was conserved in 97% of the homologues. Three of the conserved cysteine residues (the disulfide bond residues Cys472 and Cys475, as well as Cys483) are homologous to a putative zinc binding motif identified in the HSV1 homolog, ICP8 (Wang, 1990). In this study, we observed that this domain was retained in 16 of the 35 (46%) UL57 homologues (Figure 6).

51 γ γ δ η UL57 1 11 21 31 41 51 61 71 81 91 Consensus MSHEELTALAPVGPAAFLYFSRLNAETQEILATLSLCDRSSSVVIAPLLAGLTVEADFGVSVRTPVLCYDGGVLTKVTSFCPFALYFHHTQGIVAFTEDH AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... T...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

η δ 101 111 121 131 141 151 161 171 181 191 Consensus GDVHRLCEDARQKYALEAYTPEADRVPTDLAALCAAVGCQASETTVHVVVGNGLKEFLFAGQLIPCVEEATTVRLHGGEAVRVPLYPPTLFNSLQLDAEA AD169 ...... M...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... V...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

52 δ δ δ γ 201 211 221 231 241 251 261 271 281 291 Consensus DEVSLDARSAFVEARGLYVPAVSETLFYYVYTSWCQSLRFSEPRVLIEAALRQFVHDSQQSVKLAPHKRYLGYMSQRLSSLEKDHLMLSDAVVCELAFSF AD169 ...... Merlin ...... Towne ...... 44A ...... F...... 70A ...... F...... 77A ...... 90A ...... F...... 91A ...... 16B ...... 21B ...... 80B ...... F...... 4E ...... F...... 30E ......

α δ η η 301 311 321 331 341 351 361 371 381 391 Consensus ASVFFDSAYQPAESMLFSEWPLVTNATDHRDLIRALTELKLHLSTHVAALVFSANSVLYQHRLVYLQSSARHPSAGGTASQETLLKAIQFTNGLSAACED AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

53 η η η 401 411 421 431 441 451 461 471 481 491 Consensus VYNDARKVLKFQGAPLKDERYGPQHLALVCGTCPQLVSGFVWYLNRVSVYNTGLSGSSTLTNHLVGCAAGLCEACGGTCCHTCYQTAFVRVRTRLPVVPK AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

ζ η η η η η η δ 501 511 521 531 541 551 561 571 581 591 Consensus QPKKEPCVITVQSRFLNDVDILGSFGRRYNVDAKDGGLDGKGDDGVPGGGAGGGGGRDVSGGPSDGLGGGRGGGGGGDSGGMMGRGGRMLGASVDRTYRL AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

54 γ δ γ η 601 611 621 631 641 651 661 671 681 691 Consensus NRILDYCRKMRLIDPVTGEDTFSAHGKSDFVAVFSALNKFVDDEALGFVSEVRLKSSRDEVAGATQAFNLDLNPYAVAFQPLLAYAYFRSVFYVIQNVAL AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

η δ δ γ β 701 711 721 731 741 751 761 771 781 791 Consensus ITATSYIVDNPLTTNLVSKWMTQHFQSIHGAFSTTSSRKGFLFTKQIKSSKNSDHDRLLDFRLYAQGTYAVVPMEIKLSRLSVPTLIMVRVKNRPIYRAG AD169 ...... Merlin ...... Towne ...... 44A ....A...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

55 η γ 801 811 821 831 841 851 861 871 881 891 Consensus KGNAGSVFFRRDHVPRRNPAKGCLGFLLYRHHERLFPECGLPCLQFWQKVCSNALPKNVPIGDMGEFNAFVKFLVAVTADYQEHDLLDVAPDCVLSYVES AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

η δ δ γ η γ η α 901 911 921 931 941 951 961 971 981 991 Consensus RFHNKFLCYYGFKDYIGSLHGLTTRLTTQNHAQFPHVLGASPRFSSPAEFALHVKGLKTAGVPAPMAATVARESLVRSVFEHRSLVTVPVSVEKYAGINN AD169 ...... Merlin ...... Towne ...... 44A ...... S...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

56 η α γ γ δ δ γ 1001 1011 1021 1031 1041 1051 1061 1071 1081 1091 Consensus SKEIYQFGQIGYFSGNGVERSLNVSSMSGQDYRFMRQRYLLATRLADVLIKRSRRENVLFDADLIKNRVMLALDAENLDCDPEVMAVYEILSVREEIPAS AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ...... γ η η η 1101 1111 1121 1131 1141 1151 1161 1171 1181 1191 Consensus DDVLFFVDGCEALAASLMDKFAALQEQGVEDFSLENLRRVLDADAQRLTDAAGGEVHDLSALFAPSGVGAASGVGGGGLLLGESVAGNSICFGVPGETGG AD169 ...... Merlin ...... Towne ...... T...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... D...... 4E ...... 30E ......

57 η η ε 1201 1211 1221 1231 Consensus GCFLVNAGEDEAGGVGGSSGGGGGSGLLPAKRSRL AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... S...... 77A ...... S...... 90A ...... S...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... S...... 30E ...... Figure 5 HCMV strain alignments of UL57

Underlined sequences represent putative motifs identified with ScanProsite. α = N-linked glycosylation site (N-x-T/S-x), β = Tyrosine kinase phosphorylation site (R/K-xx- D/E-xxx-Y/R/K-xxx-D/E-xxx-Y), γ = Casein kinase II phosphorylation site (S/T-xx-D/E), δ = Protein kinase C phosphorylation site (S/T-x-R/K), ε = Microbodies C-terminal targeting signal (S/T/A/G/C/N-R/K/H-L/I/V/M/A/F/Y) ζ = Amidation site (x-G-RR/KK/RK), η = N-myristoylation site (G-A/N/C/Q/G/I/L/M/S/T/V-xx-all except P).

58 AHV1 (O36360) YNITEYVGSAANSP-VCSLCSGQCPCVCINTLF AtHV3 YSVAQYVGTVAVSE-LCELCQGKCPAACIHTLF BHV1 (Q89549) DALKYVASTLEGDV-PCGLCSRDDRHACAHTTL BHV4 YNISNYVGTAASSD-MCNLCQGKKPAVCINTLF BHV5 (Q6X241) DALKYVASTLEGDV-PCGLCSRDDRHACAHTTL CHV3 (Q993K9) HEIGSYVAGAATSS-VCSLCEGSTPAVCLNTLF EBV (U75698) PETGSYVAGAAASP-MCSLCEGRAPAVCLNTLF EHV1 DALRYLANTLESDV-PCGLCNQATRPACAHTTL EHV2 (Q66611) YNVPTYVGTAANTP-MCELCRGSCPASCVNTLF GHV1 (Q9QH63) SISNSGKAPCTGAVPECRWCNDESRNHCIRYTM GTH (Q6VQA1) TLLQYLNAKTEEGA--CDLCDVETRHVCPATTF HGTH (Q5ZR67) TLLQYLNAKTEEGA--CDLCDVETRHVCPATTF HSV1 (X14112) DVFRYVADSNQTDV-PCNLCTFDTRHACVHTTL HSV2 DVFRYVADSGQTDV-PCNLCTFETRHACAHTTL HHV6 TEIYNHLVNCSAN--LCEFCDGKCCQSCIGTAM HHV7 SEIYNHIVNCSSN--LCEFCEGKCCHSCIGTAL HHV8 (P88904) YNVVDYVGTAAPSQ-MCDLCQGQCPAVCINTLF HCMV STLTNHLVGCAAG--LCEACGGTCCHTCYQTAF LHV1 (Q82172) YNIAQYISTA------MCMV DTVFSHIVNAGSK--LCGACGGRCCHTCYATSF PorcHV4 (Q8QS31) STLSNHIVGCASS--LCEACGGTCCHTCYNTAF PorcHV1 (Q8JYD5) YNISQYVGSAAVSN-VCQQCKGSYPCVCINTLF PorcHV2 (Q8B424) YNICQYVGSAAVSN-VCQQCKGNYPCVCINTLF PorcHV3 (Q8B408) YSISQYVGSVAGSN-VCQLCHGTCPCTCLNTLF PHV1 (Q6UDK2) NTLKFVAAETTMAA-ECRWCTETTRQYCVRHTL RCMV HAVYQHVVHSVGN--LCEACGGRCCHTCYATPF (Q9WRU1) FNMVHYVGTAANSE-MCTLCHGNTPATCLNTLF RhCMV STLSNHLIGCSSS--LCGACGGTCCHTCYNTAF SaimHV2 YSVPQYVGTAAASD-LCELCQGTCPASCIHTLF SCMV STLSNHLIGCSSS--LCGACGGTCCHTCYNTAF SuidHV1 (Q5PPC5) DAVRYVAGSLDAEV-PCSLCDRASRPACAHTTL THV1 SALYEHLVHCAVN--LCPACRGRCCQSCYQTAF THV2 SALYEHLVHCAVN--LCPACRGRCCQSCYQTAF Turtle HV (Q5Y969) TLLQYLNAKTEEGA--CDLCDVETRHVCPATTF VZV DALKYVTGTFDSEI-PCSLCEKHTRPVCAHTTV Figure 6 Putative zinc finger domain of UL57.

The DNA binding protein from 35 homologues are aligned to show conservation of the cysteines residues (grey) in a region that is otherwise poorly conserved. Accession numbers (where available) are denoted in brackets. Species abbreviations are as follows: AHV1: Alcelaphine herpesvirus 1 (wildebeest herpesvirus); AtHV3: Ateline herpesvirus 3; BHV1/4/5: Bovine herpesvirus 1/4/5; CHV3: Callitrichine herpesvirus 3 (marmoset ); EBV: Epstein-Barr virus; EHV1/2: Equine herpesvirus 1/2; FATP: Fibropapilloma-associated turtle herpesvirus; GHV1: Gallid herpesvirus 1; GTH: Green turtle herpesvirus; HGTH: Hawaiian green turtle herpesvirus; HCMV: Human cytomegalovirus; HHV6/7/8: Human herpesvirus 6/7/8; HSV1/2: Herpes simplex virus 1/2; LHV1: Leporid herpesvirus 1; MCMV: Murine cytomegalovirus; MMR: Macca Mulatta rhadinovirus; PongHV4: Pongine herpesvirus 4; PorcHV1/2/3: Porcine herpesvirus 1/2/3; PHV1: Psittacid herpesvirus 1; RCMV: Rat cytomegalovirus; RhCMV: Rhesus cytomegalovirus; SaimHV2: Saimiriine herpesvirus 2; SCMV: Simian cytomegalovirus; SuidHV1: virus; THV1/2: Tupaiid herpesvirus 1/2; VZV: Varicella-Zoster virus

59 3.3.5 Strain variation and identification of protein motifs in UL70 Sequence alignments of the ten HCMV strains show that most variations occur in the C- terminal third of the protein (Figure 7) and that key functional motifs predicted in HSV are also highly conserved in HCMV (Biswas, 1999; Dracheva, 1995) (Figure 8A). In addition, alignment of the UL70 homologues reaffirms the importance of the highly conserved DxD catalytic motif (Dracheva, 1995) in a region that is otherwise poorly conserved. The HCMV primase also retains a putative zinc finger motif originally identified in HSV (Biswas, 1999). The zinc finger motif, spanning residues 881-920, remained highly conserved amongst the HCMV isolates studied and was present in 90% of the aligned homologues. In addition, a key phenylalanine residue involved in zinc finger structural stability (Michael, 1992) was present in all HCMV strains at position 896 and conserved in 22 of the 30 (73%) homologues (Figure 8D). Further analysis also revealed a pair of cysteines conserved across all 30 homologues (Figure 8C) while a previously uncharacterized domain spanning amino acids 566-572 contained several highly conserved aromatic residues (Figure 8B).

60 γ γ δ δ δ α γ UL70 1 11 21 31 41 51 61 71 81 91 Consensus MTLVLFATEYDSAHIVANVLSQTPTDHCVFPLLVKHQVSRRVYFCLQTQKCSDSRRVAPVFAVNNETLQLSRYLAARQPIPLSALIASLDEAETQPLYRH AD169 ...... Merlin ...... R..... Towne ...... 44A ...... R..... 70A ...... R..... 77A ...... 90A ...... 91A ...... R..... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

η γ γ 101 111 121 131 141 151 161 171 181 191 Consensus LFRTPVLSPEHGGEVREFKHLVYFHHAAVLRHLNQVFLCPTSPSWFISVFGHTEGQVLLTMAYYLFEGQYSTISTVEEYVRSFCTRDLGTIIPTHASMGE AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

61 γ ζ 201 211 221 231 241 251 261 271 281 291 Consensus FARLLLGSPFRQRVSAFVAYAVARNRRDYTELEQVDTQINAFRERARLPDTVCVHYVYLAYRTALARARLLEYRRVVAYDADAAPEAQCTREPGFLGRRL AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

β η η γ γ δ 301 311 321 331 341 351 361 371 381 391 Consensus STELLDVMQKYFSLDNFLHDYVETHLLRLDESPHSATSPHGLGLAGYGGRIDGTHLAGFFGTSTQLARQLERINTLSESVFSPLERSLSGLLRLCASLRT AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

62 δ γ 401 411 421 431 441 451 461 471 481 491 Consensus AQTYTTGTLTRYSQRRYLLPEPALAPLLERPLPVYRVHLPNDQHVFCAVASETWHRSLFPRDLLRHVPDSRFSDEALTETVWLHDDDVASTSPETQFYYT AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

γ γ γ γ 501 511 521 531 541 551 561 571 581 591 Consensus RHEVFNERLPVFNFVADFDLRLRDGVSGLARHTVFELCRGLRRVWMTVWASLFGYTHPDRHPVYFFKSACPPNSVPVDAAGAPFDDDDYLDYRDERDTEE AD169 ...... Merlin ...... Towne ...... 44A ...... I...... S...... 70A ...... 77A ...... 90A ...... S...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... S...... 30E ......

63 γ γ γ γ δ 601 611 621 631 641 651 661 671 681 691 Consensus DEDGKEDKNNVPDNGVFQKTTSSVDTSPPYCRCKGKLGLRIITPFPACTVAVHPSVLRAVAQVLNHAVCLDAELHTLLDPISHPESSLDTGIYHHGRSVR AD169 ...... Merlin ...... G...... I...... Towne ...... N...... 44A ...... 70A ...... G...... I...... 77A ...... I...... 90A ...... P..T...... I...... 91A ...... G...... I...... 16B ...... L...... 21B ...... L...... 80B ...... G...... 4E ...... P..T...... 30E ...... L......

ε γ δ α ε 701 711 721 731 741 751 761 771 781 791 Consensus LPYMYKMDQDDGYFMHRRLLPLFIVPDAYREHPLGFVRAQLDLRNLLHHHPPHDLPALPLSPPPRVILSVRDKICPSTEANFIETRSLNVTRYRRRGLTE AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... H...... 21B ...... H...... 80B ...... S...... 4E ...... Q...... 30E ...... H......

64 η γ α δ 801 811 821 831 841 851 861 871 881 891 Consensus VLAYHLYGGDGATAAAISDTDLQRLVVTRVWPPLLEHLTQHYEPHVSEQFTAPHVLLFQPHGACCVAVKRRDGARTRDFRCLNYTHRNPQETVQVFIDLR AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ...... δ δ 901 911 921 931 941 Consensus TEHSYALWASLWSRCFTKKCHSNAKNVHISIKIRPPDAPMPPATAV AD169 ...... V...... Merlin ...... Towne ...... 44A ...... 70A ...... V...... 77A ...... V...... 90A ...... 91A ...... 16B ...... V...... 21B ...... V...... 80B ...... 4E ...... 30E ...... V...... Figure 7 HCMV strain alignments of UL70

Residues highlighted in yellow indicate functionally significant amino acids (Biswas, 1999; Dracheva, 1995) while underlined sequences represent putative motifs identified with ScanProsite. α = N-linked glycosylation site (N-x-T/S-x), β = cAMP-dependent protein kinase phosphorylation site (RR/KK/RK-x-S/T), γ = Casein kinase II phosphorylation site (S/T-xx-D/E), δ = Protein kinase C phosphorylation site (S/T-x-R/K), ε = Tyrosine kinase phosphorylation site (R/K-xx-D/E-xxx-Y/R/K-xxx-D/E-xxx- Y), ζ = Amidation site (x-G-RR/KK/RK), η = N-myristoylation site (G-A/N/C/Q/G/I/L/M/S/T/V-xx-all except P).

65 518 A 526 562 B 566 633 C 644 883 D 924 HCMV FVADFDLRL PVYFFKSACP PYCRCKGKLGLRI FRCLNYTHRN.24.SRCFTKKCHS AHV1 WVLDLDLPI EVFFFKSACI TFCTCTEKLGMRV FKCLRYNHRG.24.SQCFANKCQS AtHV3 LVLDLDLHI PVYFFKSSCE KFCYCTKKLGFRI FSCLNFKHKL.24.SQCFASKCNN BHV1 LVMDFDLKV PVYFFKSAC- RFCNCNSKIGLRI FRCLTFNHRG.24.SQCFANKCNS BHV4 IVLDVDFRL PCFFYKSACP AACGCEDKMGFRV FGCLRAAHGR.25.QQCFATKCGS BHV5 IVLDVDFRL PCFFYKSACP TACGCDDKMGFRV FGCLRAAHGR.25.QQCFATKCGS CercoHV1 IILDLDVAL PCYFFKSACS PRCTCAKKIGLRV LSCLRFKHGR.25.QQCFASKCDS CercoHV9 IILDVDFHV PCYFYKSSC- NTCSCNKNLGFRV FTCLRYPH-R.25.QQCFATKCDS CHV3 LVLDMDVKI PVYFFKSACP PFCICREKLGLRV FKCLRYQHRN.24.SQCFSGRCGS EBV LVLDLDLKI PVYFFKSACP PFCICTGKLGFRV FACLRHTHRA.24.SQCFAGRCGA EHV1 IILDVDFGI PCYFYKTSCP DACECTEKMGFRV FSCLRAKHLR.25.QQCFATKCGN EHV2 LVLDFDLPL PVYFFKSACP PFCTCDAKLGLRI FVCISSNHRN.24.SQCFAHKCNS EHV4 IILDVDFDI PCYFYKTQCP NACQCTEKMGFRI FSCLRAKHLR.25.QQCFATKCGN GHV1 LILDIDIPL PVYFYKSHCS TLCKCKEKLGFRV FTCLKHSHRS.25.D------HHV6 YIGDLDLPL PIFFFKTQCD QFCVCRKKIGLRI FSCLNRQHRG.24.STCFATKCQS HHV7 YIGDLDLPL PIFFFKTTCS AFCVCKKKIGLRI FSCLTRNHKG.24.SKCFTTKCKS HHV8 WVLDFDLPV PVYFFKSACP SFCRCHDKLGMRI FLCINHNHKN.24.SQCFASKCNN HSV1 IILDLDIAL PCYFFKSACR PVCSCTDKIGLRV LSCLRFKHGR.25.QQCFAAKCDS HSV2 IILDLDIAL PCYFFKSACR VVCSCADKIGLRV LSCLRFKHGR.25.QQCFATKCDN MCMV FVGDVDLKL PVFFFKSAC- PFCVCRRKLGLRV FSCLARETYT.24.SRCFTRRCNS MDHV1 IVLDVDIHL PCYFYKSSCK KPCGCHDKIGLRV FTCVRFKHAR.25.QQCFAAKCGN MMR WVLDFDLKV PVYFFKSACP AFCHCDAKIGMRI FFCINHKHRN.24.SQCFAAKCNH PhoHV2 LVLDFDLPL QVYFFKSSCP MFCNCSEKLGLRV ------PHV1 LPLDLDITL PVYFYKTQCD RFCRCERKIGFRI FTCLKYQHRG.25.TRCFATKCGS PongHV4 FVADMDLRL PVYFFKSACR LYCRCTEKLGLRI FRCLNYVHRN.24.SRCFTKKCHS RCMV FVGDLDLKL PVYFFKSACD -FCTCRRKIGMRI FRCLTTEHHF.25.SRCFANKCQS RhCMV FIADFDLRL PVYFFKSACK DYCKCSEKLGLRI FRCLNYTHRN.24.SRCFTKKCHS SaimHV2 LVLDLDLHI HVYFFKSACE KFCYCTKKLGFRI FSCLNFKHKL.24.SQCFASKCNS THV1 LALDVDLAL PVYFFKSACP DFCACRAKLGFRV FRCLSRAHRI.24.SRCFATKCRS VZV LILDVDFHI PCYFYKTACP LPCNCKEKIGFRV FTCLRYPH-R.25.QQCFATKCDS * * * * * * * * * * * * ** *

Figure 8 Conserved domains between UL70 homologues.

(A) The DxD catalytic motif. (B) Domain consisting of several conserved aromatic residues. (C) Conserved cysteines. (D) Putative zinc finger domain. Species abbreviations are as follows: AHV1: Alcelaphine herpesvirus 1 (wildebeest herpesvirus); AtHV3: Ateline herpesvirus 3; BHV1/4/5: Bovine herpesvirus 1/4/5; CercoHV1: Cercopothecine herpesvirus 1 (Simian herpes ); CercoHV9: Cercopothecine herpesvirus 9 (Simian varicella virus); CHV3: Callitrichine herpesvirus 3 (marmoset lymphocryptovirus); EBV: Epstein-Barr virus; EHV1/2/4: Equine herpesvirus 1/2/4; GHV1: Gallid herpesvirus 1; HCMV: Human cytomegalovirus; HHV6/7/8: Human herpesvirus 6/7/8; HSV1/2: Herpes simplex virus 1/2; MCMV: Murine cytomegalovirus; MDHV1: Meleagrid herpesvirus 1; MMR: Macca Mulatta rhadinovirus; PongHV4: Pongine herpesvirus 4; PorcHV1/2/3: Porince herpesvirus 1/2/3; PhoHV2: Phocid herpesvirus 2; PHV1: Psittacid herpesvirus 1; RCMV: Rat cytomegalovirus; RhCMV: Rhesus cytomegalovirus; SaimHV2: Saimiriine herpesvirus 2; THV1: Tupaiid herpesvirus 1; VZV: Varicella-Zoster virus. * denotes highly conserved residues

66 3.3.6 Strain variation and identification of protein motifs in UL102 Sequence alignments of the UL102 isolates show a highly conserved protein, although alignments with homologues yielded a low degree of conservation for any region. As with UL44, a putative integrin binding motif was also detected in UL102 at amino acid residue 58. An alanine to glycine substitution at residue 76 in AD169 allowed the formation of another N-myristoylation site while the insertion of additional glycines in isolates 44A and 80B also increased the number of N-myristoylation sites formed (Figure 9). ScanProsite detected a glycine-rich domain (score = 12.88) between residues 553-687 and a short serine-rich domain spanning residues 795-836 (score = 12.8; normative score = 7.22). Similar to UL44, the glycine-rich domain in UL102 accounted for a significant number of the motifs identified. Isolate variation within the glycine-rich region spanning residues 629-633 was observed, with some isolates containing additional and deleted glycine residues. Similar variations were not observed in the glycine rich regions of UL44 or UL57.

67 η η ζ η γ UL102 1 11 21 31 41 51 61 71 81 91 Consensus MTAQPPLHHRHHPYTLFGTSCHLSWYGLLEASVPIVQCLFLDLGGGRAEPRLHTFVVRGDRLPPAEVRAVHRASYAALASAVTTDADERRRGLEQRSAVL AD169 ...... -G...... Merlin ...... Towne ...... A...... 44A .....S...... 70A ...... C...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... V...... 30E ......

γ η γ 101 111 121 131 141 151 161 171 181 191 Consensus ARVLLEGSALIRVLARTFTPVQIQTDASGVEILEAAPALGVETAALSNALSLFHVAKLVVIGSYPEVHEPRVVTHAAERVSEEYGTHAHKKLRRGYYAYD AD169 ...... T...... Merlin ...... T...... Towne ...... S...... 44A ...... T...... 70A ...... 77A ...... 90A ...... 91A ...... T...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

68 δ δ η 201 211 221 231 241 251 261 271 281 291 Consensus LAMSFRVGTHKYVLERDDEAVLARLFEVREVCFLRTCLRLVTPVGFVAVAVTDEQCCLLLQSAWTHLYDVLFRGFAGQPPLRDYLGPDLFETGAARSFFF AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

γ η α 301 311 321 331 341 351 361 371 381 391 Consensus PGFPPVPVYAVHGLHTLMRETALDAAAEVLSWCGLPDIVGSAGKLEVEPCALSLGVPEDEWQVFGTEAGGGAVRLNATAFRERPAGGDRRWLLPPLPRDD AD169 ...... Merlin ...... Towne ...... S...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

69 β δ δ 401 411 421 431 441 451 461 471 481 491 Consensus GDGENNVVEVSSSTGGAHPPSDDATFTVHVRDATLHRVLIVDLVERVLAKCVRARDFNPYVRYSHRLHTYAVCEKFIENLRFRSRRAFWQIQSLLGYISE AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... L...... 21B ...... L...... 80B ...... 4E ...... 30E ...... L......

β α δ η 501 511 521 531 541 551 561 571 581 591 Consensus HVTSACASAGLLWVLSRGHREFYVYDGYSGHGPVSAEVCVRTVVDCYWRKLFGGDDPGPTCRVQESAPGVLLVWGDERLVGPFNFFYGNGGAGGSPLHGV AD169 ...... Merlin ...... Towne ...... 44A ...... L...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... F... 30E ......

70 η η η δ η η η δ η 601 611 621 631 641 651 661 671 681 691 Consensus VGGFAAGHCGGACCAGCVVTHRHSSGGGG--SGVGDADHASGGGLDAAAGSGHNGGSDRVSPSTPPAALGGCCCAAGGDWLSAVGHVLGRLPALLRERVS AD169 ...... Merlin ...... Towne ...... 44A ...... GG...... D...... G...... 70A ...... 77A ...... G...... 90A ...... 91A ...... 16B ...... 21B ...... -..--...... G...... 80B ...... GG...... 4E ...... -...... 30E ...... -..--...... G......

γ η 701 711 721 731 741 751 761 771 781 791 Consensus VSELEAVYREILFRFVARRNDVDFWLLRFQPGENEVRPHAGVIDCAPFHGVWAEQGQIIVQSRDTALAADIGYGVYVDKAFAMLTACVEVWARELLSSST AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

71 γ 801 811 821 831 841 851 861 871 Consensus ASTTACSSSSVLSSALPSVTSSSSGTATVSPPSCSSSSATWLEERDEWVRSLAVDAQHAAKRVASEGLRFFRLNA AD169 ...... Merlin ....T...... Towne ...... 44A ...... 70A ...... -...... 77A ...... 90A ...... 91A ....T...... 16B ...... 21B ...... 80B ....T...... 4E ....T....-...... 30E ...... Figure 9 HCMV strain alignments for UL102

Underlined sequences represent putative motifs identified with ScanProsite. α = N-linked glycosylation site (N-x-T/S-x), β = Tyrosine kinase phosphorylation site (R/K-xx- D/E-xxx-Y/R/K-xxx-D/E-xxx-Y), γ = Casein kinase II phosphorylation site (S/T-xx-D/E), δ = Protein kinase C phosphorylation site (S/T-x-R/K), ζ = Cell attachment site (R- G-D), η = N-myristoylation site (G-A/N/C/Q/G/I/L/M/S/T/V-xx-all except P).

72 3.3.7 Strain variation and identification of protein motifs in UL105 Sequence alignments of the UL105 isolates show a highly conserved protein with sparse variation occurring outside of predicted functional domains (Figure 10). UL105 is postulated to contain six motifs typical of superfamily-1 class of helicase proteins (Zhu, 1992) based on studies of the HSV helicase, UL5 (Graves-Woodward, 1997). ScanProsite detected an ATP/GTP-binding site in motif I as well as a putative cAMP/cGMP-dependent protein kinase phosphorylation site in Motif IV. Protein alignments of the UL105 homologues predicted 2 potential N-linked glycosylation sites as well as a potential CKII phosphorylation site based on conservation across 33 homologues (Figure 11).

73

δ δ α γ UL105 1 11 21 31 41 51 61 71 81 91 Consensus MSMTASSSTPRPTPKYDDALILNLSSAAKIERIVDKVKSLSRERFAPEDFSFQWFRSISRVERTTDNNPSAATTAAATTTVHSSASSSAAAAASSEAGGT AD169 ...... Merlin ...... A.....V...... Towne ...... 44A ...... 70A ..I...... S...... 77A ...... 90A ...... 91A ...... A.....V...... 16B ...... 21B ...... 80B ...... 4E ...... A...... 30E ......

ζ α α 101 111 121 131 141 151 161 171 181 191 Consensus RVPCVDRWPFFPFRALLVTGTAGAGKTSSIQVLAANLDCVITGTTVIAAQNLSAILNRTRSAQVKTIYRVFGFVSKHVPLADSAVSHETLERYRVCEPHE AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

74

201 211 221 231 241 251 261 271 281 291 Consensus ETTIQRLQINDLLAYWPVIADIVDKCLNMWERKAASASAAAAAAACEDLSELCESNIIVIDECGLMLRYMLQVVVFFYYFYNALGDTRLYRERRVPCIIC AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... G...... VVG..L...... 21B ...... 80B ...... 4E ...... 30E ......

γ γ α δ β δ 301 311 321 331 341 351 361 371 381 391 Consensus VGSPTQTEALESRYDHYTQNKSVRKGVDVLSALIQNEVLINYCDIADNWVMFIHNKRCTDLDFGDLLKYMEFGIPLKEEHVAYVDRFVRPPSSIRNPSYA AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

75 γ δ ε α α 401 411 421 431 441 451 461 471 481 491 Consensus AEMTRLFLSHVEVQAYFKRLHEQIRLSERHRLFDLPVYCVVNNRAYQELCELADPLGDSPQPVELWFRQNLARIINYSQFVDHNLSSEITKEALRPAADV AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B D...... C...... 80B ...... F...... 4E ...... 30E ......

α η η δ δ η γ γ ε 501 511 521 531 541 551 561 571 581 591 Consensus VATNNSSVQAHGGGGSVIGSTGGNDETAFFQDDDTTTAPDSRETLLTLRITYIKGSSVGVNSKVRACVIGYQGTVERFVDILQKDTFIERTPCEQAAYAY AD169 ...... Merlin .....P...... Towne ...... 44A ...... T...... 70A ...... 77A .....P...... 90A ...... I...... 91A .....P...... 16B ...... 21B ...... 80B ..N.-P..K...... 4E .....P...... 30E ......

76 η γ γ δ η 601 611 621 631 641 651 661 671 681 691 Consensus SLVSGLLFSAMYYFYVSPYTTEEMLRELARVELPDVSSLCAAAAATAAAPAWSGGENPINNHVDADSSQGGQSVPVSQRMEHGQEETHDIPCLSNHHDDS AD169 ...... Merlin ...... Towne ...... 44A ...... I...... 70A ...... L 77A ...... 90A ...... K...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... L...... 30E ......

γ γ δ γ η η 701 711 721 731 741 751 761 771 781 791 Consensus DAITDAELMDHTSLYADPFFLKYVKPPSLALLSFEETVHMYTTFRDIFLKRYQLMQRLTGGRFATLPLVTYNRRNVVFKANCQISSQTGSFVGMLSHVSP AD169 ...... Merlin ...... Towne ...... Y...... 44A ...... 70A .V...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ......

77 δ α γ δ 801 811 821 831 841 851 861 871 881 891 Consensus AQTYTLEGYTSDNVLSLPSDRHRIHPEVVQRGLSRLVLRDALGFLFVLDVNVSRFVESAQGKSLHVCTTVDYGLTSRTAMTIAKSQGLSLEKVAVDFGDH AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ...... γ β 901 911 921 931 941 951 Consensus PKNLKMSHIYVAMSRVTDPEHLMMNVNPLRLPYEKNTAITPYICRALKDKRTTLIF AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 80B ...... 4E ...... 30E ...... Figure 10 HCMV strain alignments for UL105

Residues highlighted in yellow indicate functionally significant motifs (Graves-Woodward, 1996; Graves-Woodward, 1997) while underlined sequences represent putative motifs identified with ScanProsite. α = N-linked glycosylation site (N-x-T/S-x), β = cAMP-dependent protein kinase phosphorylation site (RR/KK/RK-x-S/T), γ = Casein kinase II phosphorylation site (S/T-xx-D/E), δ = Protein kinase C phosphorylation site (S/T-x-R/K), ε = Tyrosine kinase phosphorylation site (R/K-xx-D/E-xxx-Y/R/K-xxx- D/E-xxx-Y), ζ = ATP/GTP-binding site motif A (A/G-xxxx-G-K-S/T), η = N-myristoylation site (G-A/N/C/Q/G/I/L/M/S/T/V-xx-all except P).

78 20 A 27 409 B 415 473 C 481 HCMV LILNLSSAA MTRLFLSHVEVQA RIINYSQFV AHV1 FILNMTSEA WTRLFLSHSEVKQ RLSNYSQFI AtHV3 FILNMTSDA WTRLFLSHAEVKS RLSNYSQFI BHV1 FILNMTSDA WTRLFLSHNEVKA RLSNYSQFI BHV4 VFLNFTSMH WTRLYSSHREVSA RITNYSQSR BHV5 VFLNFTSMH WTRLYSSHREVSA RITNYSQSR CHV3 FMLNMTSDA WTRLFLSHAEVKQ RLGNYSQFA CercoHV1 VYLNFTSMH WTRLYSSHKEVSA RLNNYSQSR CercoHV9 AFLNFTSMH WTRLFSSHKEVSA RITNYSQSQ EBV FMLNMTSDA WTRLFLSHAEVKT RLGNYSQFA EHV1 VYLNFTSMH WTRLYSSHKEVSA RVSNYSQSR EHV2 FILNMTSDS WTRLFVSHREVKA RLSNYSQFI EHV4 VYLNFTSMH WTRLYSSHKEVSA RVSNYSQSR GHV1 TYLNFTAMH WTRLYSSHKEVSA RLSNWSQSR GHV3 VYLNFSAMQ WTRLFSSHEEVKE RLGNYSQSR HSV1 AFLNFTSMH WTRLFSSHKEVSA RITNYSQSQ HSV2 AFLNFTSMH WTRLFSSHKEVSA RITNYSQSQ HHV6 FLLNMSSAP TTRLFLSHNEVKN RLNTYSQFA HHV7 FLLNMSSAA MTRLFLSHYEVKS RLNTYSQFA HHV8 FILNMTSDA WTRLFISHQEVKS RLSNYSQFA MCMV FVLNMSSAS VTRLFISHAEVKR RISNYSQFT MDHV1 TYLNFTAMH WTRLYSSHKEVST RLSNWSQSR MD TYLNFTAMH WTRLYSSHKEVSA RLSNWSQSR MMR FILNMTSDA WTRLFLSHSEVKA KLSNYSQFV PorcHV1 FFLNMTSDS WTRLFLSHIEVKN RLSNYSQFV PorcHV2 FFLNMTSDA WTRLFLSHAEVKN RLSNYSQFV PongHV4 LILNLSSAA MTRLFLSHAEVQL RIINYSQFV Pseudorabies TYLNFTSMH WTRLYSSHREVSA RITNYSQSR RCMV FVLNTSSAL ATRLFVSHREVKD RIGNYSQFT RhCMV LILNLSSAA MTRLFLSHVEVQA RIINYSQFV SuidHV1 TYLNFTSMH WTRLYSSHREVSA RITNYSQSR THV1 FILNMSSAV MTRLFISHAEVKR RIINYSQFV VZV VYLNFTSMH WTRLYSSHKEVSA RLHNYSQSR *** * * *

Figure 11 Putative motifs conserved across UL105 homologues (A) and (C) Conserved putative N-linked glycosylation sites. (B) Conserved putative CK2 phosphorylation site. Species abbreviations are as follows: AHV1: Alcelaphine herpesvirus 1 (wildebeest herpesvirus); AtHV3: Ateline herpesvirus 3; BHV1/4/5: Bovine herpesvirus 1/4/5; CercoHV1: Cercopothecine herpesvirus 1 (Simian herpes B virus); CercoHV9: Cercopothecine herpesvirus 9 (Simian varicella virus); CHV3: Callitrichine herpesvirus 3 (marmoset lymphocryptovirus); EBV: Epstein-Barr virus; EHV1/2/4: Equine herpesvirus 1/2/4; GHV1/3: Gallid herpesvirus 1/3; HCMV: Human cytomegalovirus; HHV6/7/8: Human herpesvirus 6/7/8; HSV1/2: Herpes simplex virus 1/2; MCMV: Murine cytomegalovirus; MDHV1: Meleagrid herpesvirus 1; MD: Marek’s disease; MMR: Macca Mulatta rhadinovirus; PongHV4: Pongine herpesvirus 4; PorcHV1/2: Porince herpesvirus 1/2; RCMV: Rat cytomegalovirus; RhCMV: Rhesus cytomegalovirus; SuidHV1: Pseudorabies virus; THV1: Tupaiid herpesvirus 1; VZV: Varicella-Zoster virus.

79 3.3.8 Strain variation and identification of protein motifs in UL84 Sequence alignments revealed a high degree of variation amongst the isolates (Figure 12), especially in a region that has been associated with oligomerization of the UL84 protein (Colletti, 2004). In particular, deletions of glutamine residues in three clinical isolates (16B, 21B and 30E) were observed in a region that was previously associated with nuclear import (Xu, 2002) and is part of a previously described hydrophobic region (He, 1992). Two overlapping CKII phosphorylation sites (consensus site = S/TxxD/E) (Pinna, 1990) TLQE and SQEE, were detected flanking this region (Figure 13) and interestingly, three strains contained only a single CKII site; two strains (AD169 and 70A) had substitutions which abolished the second motif while another isolate (77A) abolished the first motif. ScanProsite also detected a previously described leucine zipper domain (He, 1992) at the amino-terminus of UL84.

80 η η η η δ α δ η γ γ γ η γ δ UL84 1 11 21 31 41 51 61 71 81 91 consensus MPRVDPNLRNRARRPRARRGGGGGVGSNSSRHSGKCRRQRRALSAPPLTFLATTTTTTMMGVASTDDDSLLLKTPDELDKHSGSPQTILTLTDKHDIRQP AD169 ...... Y...... Merlin ...... T...... Towne ...A...... 44A ...... 70A ...... Y...... 77A ...... T...... 90A ...... T...... 91A ...... T...... 16B ...... T...... 21B ...... T...... 4E ...... R...... 30E ...... T......

ε γ γ γ 101 111 121 131 141 151 161 171 181 191 consensus RVHRGTYHLIQLHLDLRPEELRDPFQILLSTPLQLGEANGESQTAPATSQEEETAASHEPEKKKEKEE-KKEEDEDDRNDDRERGILCVVSNEDSDVRPA AD169 ...... D...... L...... Q...... Merlin ...... K.E...... Towne ...... L...... K.E...... 44A ...... DL...... K...... 70A ...... L...... Q...... 77A ...... TM...... K.E...... 90A ...... T...... K.E...... 91A ...... K.E...... 16B ...... -...... 21B ...... -...... 4E ...... 30E ...... -......

81 γ α γ γ 201 211 221 231 241 251 261 271 281 291 consensus FSLFPARPGCHILRSVIDQQLTRMAIVRLSLNLFALRIITPLLKRLPLRRKAAHHTALHDCLALHLPELTFEPTLDINNVTENAASVAD AESTDADLTP AD169 ...... T...... Merlin ...... T...... Towne ...... P...V...... M...... S...... A...... 44A ...... 70A ...... T...... 77A ...... V.----.....T.... 90A ...... T...... 91A ...... T...... 16B ...... A..A...... 21B ...... A..A...... 4E ...... TA..A...... 30E ...... A..A......

δ δ δ δ γ α β 301 311 321 331 341 351 361 371 381 391 consensus TLTVRVRHALCWHRVEGGISGPRGLTSRISARLSETTAKTLGPSVFGRLELDPNESPPDLTLSSLTLYQDGILRFNVTCDRTEAPADPVAFRLRLRRETV AD169 ...... Merlin ...... Towne ...... V...... M...... 44A ...... M...... 70A ...... 77A ...... M...... 90A ...... 91A ...... 16B ...... T...... 21B ...... T...... 4E ...... 30E ...... T......

82 δ δ δ γ 401 411 421 431 441 451 461 471 481 491 consensus RRPFFSDAPLPYFVPPRSGAADEGLEVRVPYELTLKNSHTLRIYRRFYGPYLGVFVPHNRQGLKMPVTVWLPRSWLELTVLVSDENGATFPRDALLGRLY AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... 77A ...... H...... 90A ...... 91A ...... 16B ...... 21B ...... 4E ...... C...... 30E ......

δ η η α δ 501 511 521 531 541 551 561 571 581 consensus FISSKHTLNRGCLSAMTHQVKSTLHSRSTSHSPSQQQLSVLGASIALEDLLPMRLASPETEPQDCKLTENTTEKTSPVTLAMVCGDL AD169 ...... Merlin ...... Towne ...... 44A ...... 70A ...... Y...... 77A ...... 90A ...... 91A ...... 16B ...... 21B ...... 4E ...... M....G...... 30E ...... Figure 12 HCMV strain alignments for UL84

Residues highlighted in yellow indicate functionally significant motifs (Colletti, 2004; He, 1992; Lischka, 2006) while underlined sequences represent putative motifs identified with ScanProsite. α = N-linked glycosylation site (N-x-T/S-x), β = cAMP-dependent protein kinase phosphorylation site (RR/KK/RK-x-S/T), γ = Casein kinase II phosphorylation site (S/T-xx-D/E), δ = Protein kinase C phosphorylation site (S/T-x-R/K), ε = Leucine zipper (L-xxxxxx-L-xxxxxx-L-xxxxxx-L), η = N-myristoylation site (G-A/N/C/Q/G/I/L/M/S/T/V-xx-all except P).

83 CKII Consensus PATSQEEETAASHEPEKKKEKEE-KKEEDEDDRNDDRERGILCVVSNEDSDVRPA AD169 PATLQEEETAASHEPEKKKEKQEKKEE-DEDDRNDDRERGILCVVSNEDSDVRPA Merlin PATSQEEETAASHEPEKKKEKEEKKEEEDEDDRNDDRERGILCVVSNEDSDVRPA Towne PATSQEEETAASHELEKKKEKEEKKEEEDEDDRNDDRERGILCVVSNEDSDVRPA 44A LATSQEEETAASHEPEKKKEKEKKKEE-DEDDRNDDRERGILCVVSNEDSDVRPA 70A PATLQEEETAASHEPEKKKEKQEKKEE-DEDDRNDDRERGILCVVSNEDSDVRPA 77A PTMSQEEETAASHEPEKKKEKEEKKEEEDEDDRNDDRERGILCVVSNEDSDVRPA 90A PTTSQEEETAASHEPEKKKEKEEKKEEEDEDDRNDDRERGILCVVSNEDSDVRPA 91A PATSQEEETAASHEPEKKKEKEEKKEEEDEDDRNDDRERGILCVVSNEDSDVRPA 16B PATSQEEETAASHEPEKKKEKE-KKEE-DEDDRNDDRERGILCVVSNEDSDVRPA 21B PATSQEEETAASHEPEKKKEKE-KKEE-DEDDRNDDRERGILCVVSNEDSDVRPA 4E PATSQEEETAASHEPEKKKEKEEKKEE-DEDDRNDDRERGILCVVSNEDSDVRPA 30E PATSQEEETAASHEPEKKKEKE-KKEE-DEDDRNDDRERGILCVVSNEDSDVRPA **** ************* *** **** ********************** Figure 13 Overlapping putative casein kinase-2 (CKII) phosphorylation sites of UL84

84 3.4 Discussion

Several putative motifs were identified using protein scanning software and although the core HCMV replication genes have already been characterized (Anders, 1988; Ertl, 1992; Pari, 1993; Smith, 1995a; Smith, 1996; Weiland, 1994), their significance in replication was mainly predicted from similarities with the HSV counterparts (Anders, 1988; Pari, 1993). By combining the novel approach of motif scanning and phylogenetic analysis between isolates and across species homologues, we have identified several putative functional motifs in HCMV that may contribute to a complete understanding of the CMV replication cycle.

One particular class of motifs has gained much attention in recent times. CKII is a cellular serine-threonine kinase that has multiple target substrates involved in nucleic acid replication, transcription, and protein synthesis among other biological processes (Meggio, 2003). Active CKII is incorporated into the HCMV virion resulting in immediate post-infection phosphorylation of cellular proteins that enhance HCMV replication (Nogalski, 2007). There is also increasing evidence that CKII enhances HCMV replication through phosphorylation of several HCMV proteins involved in DNA replication and other processes (Alvisi, 2005; Barrasa, 2005; Jarvis, 2004). This includes the UL44 DNA polymerase accessory protein that contains a (S/T)xx(D/E) consensus CKII phosphorylation motif (also confirmed here) and is phosphorylated at serine and/or threonine residues (Alvisi, 2005). In this study, we have identified interesting and potential CKII phosphorylation motifs in the HCMV helicase (UL105) and putative initiator protein (UL84), the latter recently shown to be a substrate of CKII in pull-down assays (Gao, 2007). Although the previously described putative NLS spanning residues 160-171 (Xu, 2002) has been disproved of its role in nuclear import (Lischka, 2003), the position of the CKII phosphorylation site relative to this putative NLS in UL84 (Xu, 2002) resemble a pattern shared by the CKII enhancing the NLS of UL44 (Alvisi, 2005), which also binds with high affinity to the importin α/β heterodimer (Alvisi, 2005). In addition, both these motifs fit the ‘CcN’ profile described for phosphorylation regulated nuclear localization (Jans, 1996).

UL84 is an essential regulatory protein that is required for oriLyt-dependent DNA replication (Lischka, 2003; Xu, 2004; Yu, 2003). Of the six replication genes studied, it

85 also had the lowest amino acid identity when compared against clinical isolates with a score ranging from 97.46% to 100% identity. Initial characterization of UL84 identified two potential leucine zippers between amino acids 114 to 135 and 325 to 373 (He, 1992), as well as a highly charged region starting at amino acid 161 (He, 1992). Coincidentally, this region was initially associated with nuclear import (Xu, 2002), although subsequent studies later mapped nuclear localization to a complex domain downstream of this region (Lischka, 2003). In addition, two nuclear export signals (NES) starting at amino acids 229 and 360 were also found to act as a non-conventional NLS which allowed pUL84 to shuttle to the cytoplasm (Lischka, 2006). Recent studies have shown that several regions of pUL84 share similar homology to the DExD/H box family helicases (Colletti, 2005) while another bioinformatic study has shown homology to dUTPases (Davison, 2005), both of which support the hypothesis that pUL84 acts as an initiator protein for viral-DNA synthesis of HCMV (Xu, 2004a). PROSITE scanning for potential motifs yielded three potential N-linked glycosylation sites while in earlier studies, five were detected (He, 1992). However, on closer inspection one of the original five N-linked glycosylation sites, beginning at amino acid 354, would have been discounted as having functional significance on the basis of proline being present within the consensus sequence (Bause, 1983; Gavel, 1990). BLAST searching revealed only four sequence homologues, though when aligned showed a high degree of conservation for both of the NES as well as a putative CKII phosphorylation site at UL84 amino acid residue 474 identified in our studies. In addition, the putative CKII phosphorylation site lies upstream of a motif suggested to be associated with helicase function (Colletti, 2005) and may have a regulatory role. The putative leucine zipper starting at amino acid 114 in HCMV was also detected in our scanning programs, while the putative zipper starting at residue 325 was not, albeit neither were conserved structures when aligned against the BLAST result homologues.

Among the ten HCMV isolates studied for UL44, variations in peptide sequence occurred outside of regions currently known to have a functional role with the exception of a K431R substitution in the previously identified NLS (Alvisi, 2005). Both the HCMV UL44 and the HSV-1 homolog (UL42) share a “connector loop” structure that plays a crucial role in binding to the DNA polymerase (Appleton, 2004; Bridges, 2001; Loregian, 2004a). However, the sequence structure for the HCMV connector loop is better conserved across the species, with more than 50% of the sequence alignments 86 having a similar structure of at least 57% sequence homology in that region. Although the crucial functions of UL44 has previously been mapped to the N-terminal two-thirds of the gene (Weiland, 1994), more than half of the potential motifs identified for UL44 in this study were found in the C terminal third of the gene, which has been determined to be functionally dispensable for replication (Ertl, 1992). However, a recent study elucidating the presence of a nuclear localization signal located at residues 425-431 of the C-terminus (Alvisi, 2005) suggests that further study into this region needs to be done, considering that the putative casein kinase-2 (CK2) site enhancing the NLS (Alvisi, 2005) was also identified in our motif scans. Further analysis revealed an unusual integrin binding motif in UL44 that was later shown to be conserved in pongine Herpesvirus 4 as well. Integrins are known to mediate cell-cell association as well as interactions between cells and extra-cellular proteins (Haynes, 2002). In addition, a previous study on the murine CMV M44 homologue protein identified a putative integrin binding RGD motif that was capable of mediating cell adhesion (Loh, 2000), suggesting that UL44 could play a novel role in CMV infection.

Structure-function prediction for the UL57 single stranded DNA binding protein in HCMV were largely based on studies of the HSV-1 homolog ICP8, which has so far mapped the cooperative DNA binding region as well as the nuclear localization signal to the C-terminus (Dudas, 1998; Gao, 1992; Mapelli, 2000), while an intranuclear localization site was found to be located between residues 1080-1135 (Taylor, 2003). A putative zinc binding motif between amino acids 499-512 in ICP8 (Gao, 1988; Gupte, 1991) was also found to be conserved in 16 of the 35 (46%) homolog sequences for the single stranded DNA binding protein, albeit it was found to be non-essential for DNA binding in HSV-1 (Wang, 1990). Although formal elucidation of its function in other species of herpesvirus is yet to be validated, the structural nature of the motif is missing key elements including a conserved phenylalanine and leucine involved in stabilizing the motif (Jasanoff, 1993; Michael, 1992). However, the N terminal fragment of the putative zinc finger beginning at amino acid 472 of UL57 showed a potential disulfide bond that was conserved in 34 of the 35 (97%) homolog alignments. The conservation of this putative disulfide bond motif across the species indicates an evolutionarily conserved functional domain. Currently, the exact DNA binding region has yet to be elucidated with numerous studies detailing a range of potential active sites (Gao, 1989; Wang, 1990; White, 1999). The presence of a micro-bodies targeting signal observed in 87 the C-terminus of UL57, while yet to be functionally elucidated, suggests this protein may be involved with the non-viral, non-lysosomal microbodies which have been observed in CMV infected cells for decades (Craighead, 1972; Ruebner, 1966).

The primase-helicase complex of HCMV consisting of UL70, UL102, and UL105 is considered to be homologous to the HSV-1 primase-helicase complex consisting of UL52, UL8, and UL5 respectively (McMahon, 2002; Pari, 1993; Smith, 1996). All three genes were highly conserved amongst the clinical isolates ranging from 98% to 100% sequence identity and 99 to 100% amino acid identity. Although the HCMV primase (UL70) has not been formally characterized, it shares two functionally distinct motifs that have been documented in its HSV-1 counterpart, UL52. The DxD catalytic motif is essential for initiating DNA synthesis via the primer synthesis (Dracheva, 1995; Klinedinst, 1994) while substitutions of the third and fourth conserved cysteines in the zinc-finger at the C terminal end resulted in the loss of primase, helicase, ATPase, and DNA-binding activities of the HSV helicase-primase complex (Biswas, 1999). The zinc binding residues CHCC were found to be conserved in 26 of the 30 (87%) herpesvirus homologues, while a key phenylalanine involved in structural stability (Michael, 1992) was conserved in 22 of the 30 (73%) sequences. Interestingly, putative zinc fingers lacking the conserved phenylalanine or an ‘aromatic swap’ (Kochoyan, 1991) equivalent have been shown to be non-essential for key biological functions such as DNA binding and replication (Hui, 2003; Wang, 1990). Analysis of the UL70 homologues show a highly conserved cluster of aromatic residues preceding a conserved cysteine which has been associated with conferring resistance to a compound which currently inhibits the HCMV primase (Chen, 2007). Aromatic interactions have been attributed to involve up to 60% of the aromatic residues in a protein of which 80% of these contribute to tertiary structure (Burley, 1985) and may provide some insight into the structure-function relationship of that domain. Further analysis also revealed a pair of cysteines conserved across all 30 homologues while a previously uncharacterized domain spanning amino acids 631-646 contained several highly conserved residues amongst the isolates and homologues in an otherwise poorly conserved region (Figure 14B). Protein disulfide isomerase (PDI), with the canonical CxxC motif, is the most efficient known catalyst of oxidative protein folding (Weissman, 1993). Our phylogenetic analyses uncovered two highly conserved cysteine residues forming a CxC motif in a previously uncharacterized domain of UL70 primase 88 and 30 primase homologues. The conservation of this CxC motif suggest that the CMV primase may utilize a rare thiol-disulfide exchange mechanism (Woycechowsky, 2003) that is present in only a handful of organisms including the Mengo encephalomyelitis virus coat protein (Krishnaswamy, 1990), Bacillus Ak.1 protease (Smith, 1999) and the E.Coli chaperone, Hsp33 (Jakob, 1999). The CxC motif is known to mediate disulfide bond interactions between other conserved cysteines to form the well defined tertiary structure for the chemokine class of proteins (Fernandez, 2002), and hence may have a similar role in stabilizing the UL70 tertiary structure of HCMV here. In addition, an earlier study of the HSV homologue demonstrated that substitutions of the arginine residue seven amino acids downstream of the putative domain decreased the replication efficiency of HSV by more than 50% (Klinedinst, 1994), suggesting that functionally significant residues lie in close proximity to the CxC motif.

UL102 is a positional homolog of the HSV-1 UL8 primase associated factor (PAF) and is shown to be involved in modulating primase-helicase activity (Barnard, 1997). Apart from it’s initial characterization (Smith, 1995a), there have been few studies on UL102, which, along with UL44 is the least well conserved of the core replication proteins in HCMV, with BLAST searches yielding only 12 and 14 homologues respectively. However, the functional domains of HSV-1 UL8 has been mapped to the regions spanning amino acids 6–23 at the N-terminus, 718–750 at the C-terminus, and 78–339 internally where deletion mutants inhibit the ability of UL8 to support origin-dependent DNA synthesis (Barnard, 1997). Several variations involving the distribution of glycine residues were observed between residues 625-635 (Figure 14E) amongst the isolates while the short serine-rich domain spanning residues 795-836 (Figure 14F) may have implications for phosphorylation (Barrasa, 2005). Although UL102 was highly conserved amongst the clinical isolates, alignments with the BLAST results showed no distinguishable motifs conserved across the homologues. However, similar to UL44, an unusual integrin binding motif was identified in the N-terminal of UL102 (Figure 14D), although whether it plays a role in infection has yet to be seen.

UL105 is predicted to encode the HCMV helicase protein, based on sequence homology to its HSV-1 UL5 counterpart which also contains six motifs typical of superfamily-1 of helicase proteins (Gorbalenya, 1989; Graves-Woodward, 1997). However, of the six motifs found to be essential for helicase activity in HSV-1 (Zhu, 1992), only two have 89 been functionally characterized. Motif I (also called the Walker A motif) follows the consensus sequence GxxGxGKT/S and is believed to be involved in binding ATP (Walker, 1982) while motif II (also called the Walker B motif) contains a series of hydrophobic residues terminated by an aspartic and glutamic acid residue and is believed to be involved in stabilizing Mg2+ (Walker, 1982). The functional significance of Motifs III, IV, V, and VI is yet to be elucidated although a G815A substitution in motif V of HSV-1 UL5 altered its ability to bind ATP, but otherwise left primase and helicase activities intact (Graves-Woodward, 1996). In this study, a putative cAMP/cGMP-dependent protein kinase phosphorylation site was detected in Motif IV suggesting that this domain may interact with the cAMP signaling pathway as part of its function. These six motifs were found to be highly conserved amongst the HCMV clinical isolates. Comparisons with homologues revealed a high degree of conservation for the six motifs amongst all 33 sequence homologues, emphasizing the strict structure- function relationship of this class of proteins. In addition to the six helicase motifs, two potential N-linked glycosylation sites as well as a potential casein-kinase 2 phosphorylation site was found to be highly conserved in the helicase homologues. All three motifs were previously identified from the results of PROSITE scanning (Figure 14M/N/O). N-linked glycosylation sites follow a distinct Nx(T/S)x motif and have been associated with aiding correct protein folding as well as cell-cell adhesion while casein kinase 2 phosphorylation sites follow a (S/T)xx(D/E) consensus motif (Pinna, 1990) and has already been shown to be involved with several biological processes of HCMV replication (Alvisi, 2005; Barrasa, 2005; Jarvis, 2004).

90

Figure 14 Isolate variation in the HCMV primase-helicase complex.

Green lines represent amino acid substitutions amongst isolates while pink segments represent known functional domains and tan segments represent interesting regions identified by phylogenetic analysis: A) Catalytic domain (Dracheva, 1995), B) Novel conserved domain, C) Putative zinc finger motif (Biswas, 1999) D) Putative integrin binding motif, E) Glycine-rich domain, F) Serine-rich domain, G-L) Helicase motifs I-VI (Zhu, 1992), M-N) Putative N-linked glycosylation sites, O) Putative CKII phosphorylation sit

The functional domains of a protein are often found to be highly conserved amongst homologues (Dracheva, 1995; Gorbalenya, 1988), with the implication that these domains remained unchanged over time and during species differentiation due to the crucial role they play in propagating the organism. In this study we have identified nearly 300 potential motifs spanning five core and one non-core replication gene of HCMV, many of which are ubiquitous commonly-occurring sequences. Motif numbers can be overestimated by ScanProsite in glycine rich regions, where a string of glycines can give rise to the same motif repeated several times. The 27 N-myristoylation sites identified for UL44, for example, may be an overestimation, and caution should be applied when using ScanProsite to determine a definitive number of particular motifs within a given gene. However, as we have demonstrated, motif scanning programs 91 combined with phylogenetic analysis are useful tools for the discovery of previously unidentified motifs likely to be important for overall protein activity. This provides focus areas for future studies of these important HCMV replication proteins including investigations of potential antiviral inhibitors.

92 4 HCMV primase expression

4.1 Introduction

Despite the efficacy of HCMV anti-viral therapy in treating HCMV-related disease (Goldberg, 2003; Salzberger, 2005), long-term administration of ganciclovir (GCV) or foscarnet (FOS) can lead to severe toxicity or the emergence of drug-resistant virus strains (Baldanti, 1995). Therefore, there is a need to develop new therapeutic drugs that combine efficacy and safety with a novel mechanism of action, thus excluding the possibility of cross-resistance with existing therapeutics (Biron, 2006; Wathen, 2002). Among the new anti-viral therapies currently being developed, Maribivir targets the CMV protein kinase gene (Evers, 2004) and has been shown to be effective against strains resistant to GCV, FOS, and Cidofovir (CDV) (Drew 2006). However, resistant strains have already begun to emerge, with mutations conferring resistance identified in the UL27 and UL97 genes (Chou, 2007; Komazin, 2003).

The HCMV primase, encoded by the UL70 ORF, shares 27% homology with the HSV- 1 primase, UL52 (Chee, 1990) and is an essential component of the viral DNA replication apparatus (Pari, 1993). All known DNA polymerases are unable to initiate synthesis of a new DNA strand de novo, while almost all require the activity of a DNA primase to prime the template (Arezi, 2000). However, unlike eukaryotic primases, which initiate primer synthesis at preferred but nonspecific template sequences, prokaryotic and viral primases usually have specific template sequence requirements for primer initiation (Frick, 2001). The HCMV primase is part of a trimeric complex which also includes the helicase (UL105) and the primase-associated protein (UL102) (McMahon, 2002). The complex is thought to unwind and prime the DNA template for replication, which has been shown to be the case for homologous proteins of the HSV-1 (Graves-Woodward, 1997; Ramirez-Aguilar, 2004). Currently, high-throughput screening has identified a series imidazolyl-pyrimidine compounds that can inhibit the primase (Cushing, 2006) Although the mechanism of action of these compounds are largely unknown, resistance conferred by a P570S point mutation in UL70 suggests that the primase is inactivated via irreversible binding of the primase with the compound (Chen, 2007). The importance of the primase-helicase complex in CMV DNA replication presents as a logical target for future anti-viral therapeutics. Compounds

93 which can target either the primase activity or its ability to associate with complement members of the essential complex can provide a novel mechanism of action that is unique from existing treatments. In addition, they may be used in conjunction with existing anti-viral therapies to provide effective treatment and reduce the incidence of specific resistance mutants that often emerge with long term single course treatments. By elucidating the functional aspects of the primase, we can increase our understanding of the complex CMV replication cycle and provide the groundwork for developing new anti-viral strategies in the future.

Due to the large genome size and prolonged replication cycle of CMV, generating recombinant HCMV has been difficult and time-consuming (Wang, 2004), while previous efforts to study the HCMV helicase–primase proteins in particular have been hampered by their low abundance in HCMV-infected cells (McMahon, 2002). However, the advent of Bacterial Artificial Chromosome (BAC) technology solves these problems as well as providing a robust genetic system for constructing viral mutants (Jarvis 2007). BAC technology has already been used in several studies ranging from systematic profiling of the CMV genome (Dunn, 2003; Yu, 2003) to elucidating mutations conferring anti-viral resistance (Chou, 2007a; Martin, 2006). We proposed using the BAC system for the purpose of creating recombinant mutants to study the importance of the HCMV primase for viral replication. Furthermore, we pursued the expression of wild-type and mutant UL70 from baculovirus constructs in insect cells, as well as in vitro expression using a streamlined TNT Quick Coupled Transcription/Translation Kit (Promega, USA), with a view towards functional and protein-protein interaction analyses.

94 4.2 Methods

4.2.1 Cloning of UL70 into high-copy pBluescript II primase vector

4.2.1.1 Viral stocks and DNA extraction Viral DNA from high-passage HCMV isolates 29A and 5B were extracted from frozen cell-associated viral stock using the QIAamp DNA Mini Kit (Qiagen, USA) as per manufacturer’s instructions. Briefly, specimens were incubated in lysis buffer and Proteinase K to release cellular DNA. Ethanol was used to precipitate the DNA which was subsequently captured in microfuge spin columns (Qiagen, USA). The DNA was then washed and eluted in 50µl of buffer AE after 5 min incubation instead of the recommended 200µl to increase DNA concentration.

4.2.1.2 Polymerase chain reaction and purification of UL70 for pBluescript The full length ORF encoding the HCMV primase (UL70) was amplified from the DNA of high-passage isolates using primers containing the SpeI and HindIII restriction sites in the 5’ and 3’ ends respectively (Table 10). Two sets of primers were created such that UL70.SpeI and UL70.HindIII produced a ~3kb fragment containing only the UL70 ORF (UL70a) while UL70.3SpeI and UL70.4HindIII produced a ~5kb fragment that contained the UL70 ORF as well as approximately 1kb genomic material flanking the 5’ and 3’ ends (UL70b) (Table 10). A single round PCR was carried out in a 50µl reaction consisting of 1x buffer (supplemented with 1.75mM MgCl2 ), 400µM dNTPs, 300nM primers, 100µg/ml bovine serum albumin (BSA) (Promega, US), 3.75U ExpandLong template enzyme (Roche) and 5µl DNA template. The single round cycling conditions consisted of denaturation at 94°C for 3 min; 10 cycles of 94°C for 10 sec, 65°C for 30 sec, 68°C for 4 min; 20 cycles of 94°C for 15 sec, 65°C for 30 sec, 68°C for 4 min (increasing by 20 sec for each subsequent cycle); and a final elongation at 68°C for 7 min. PCR products were visualized following electrophoresis on 0.75% agarose gels and purified using polyethylene glycol (PEG). Briefly, this involved mixing the PCR product with 2x PEG (26.7% PEG 8000, 0.6 M NaOAc (pH 5.2), 6.5 mM MgCl2) in a 1:1 ratio. The mixture is incubated at room temperature for 10 min and centrifuged at

95 8000xg, with the supernatant removed afterwards. The DNA pellet was then washed with 96% ethanol, air-dried and re-suspended in 20µl distilled water (Baxter, Australia).

Table 10 Nucleotide primers used for UL70/pBluescript cloning

Primer Name Primer sequence* UL70.SpeI 5' - GCACTAGTTGACGTCGGTCCGAAACCTCC UL70.HindIII 5' - CGCAAGCTTATCGTGCGGGTTGACGGGTAG UL70.3SpeI 5' - TAGAGGACTAGTCAGCACGTAATCGGCCACAGG UL70.4HindIII 5' - CGATGAAAGCTTTATAGATGAGACCGCTGCCGG UL70.4 5' – CCCAACTGGATCTGCGTAACCTGCT T3 5' - AATTAACCCTCACTAAAGGG T7 5' - GTAATACGACTCACTATAGGGC *Underlined sequences denote restriction sites

4.2.1.3 Molecular cloning of UL70 into pBluescript II Both the 3kb and 5kb UL70 PCR products as well as the pBluescript II vector were double digested with SpeI and HindIII restriction enzymes in separate 20µl reaction consisting of 1x buffer B (Promega, US), 2 µg/µl acetylated BSA (Promega, US), 5U SpeI (Promega, US), 5U HindIII (Promega, US), and ~1µg DNA, or a 5U/µg over- digest. Restriction digest reactions were incubated at 37°C for 1.5hrs followed by heat inactivation at 74°C for 15 min. Ligation was performed immediately after in a 10µl reaction consisting of 1x ligase buffer (Stratagene, USA), 1mM rATP (Stratagene, USA), 2U T4 DNA ligase (Stratagene, USA), and DNA in an insert to vector ratio of 3:1 as calculated by the following formula:

Figure 15 Insert to vector ratio formula

A reaction consisting of pUC18 plasmid in place of the sample was used as a ligation control. The ligation reactions were incubated at room temperature for 1 hr followed by transformation to XL10-Gold Ultracompetent Cells (Stratagene, USA) as per

96 manufacturer’s protocol. Briefly, this involved thawing a 100µl aliquot of cells on ice before adding 4µl β-Mercaptoethanol and 2µl ligation mixture. The reaction was heat pulsed at 42°C for 30 sec followed by the addition of 900µl NZY broth. The cells were grown in a shaking incubator at 37°C and 225rpm for 1 hr prior to plating on LB- Ampicillin agar plates supplemented with 100µl 1M IPTG and 20µl X-Gal and incubated at 37°C for 16hrs. Plates were then placed in a 4°C fridge for 1 hr to enhance the blue/white screening process.

Single white colonies were re-streaked and screened via colony PCR using an internal UL70 primer and a pBluescript T7 primer (Table 10). A single round colony PCR was carried out in a 50µl reaction consisting of 1x GoTaq Flexi buffer (Promega), 4mM

MgCl2, 200µM dNTPs, 0.2µM primers, 1.25U GoTaq DNA polymerase (Promega), and 5µl DNA template. The cycling conditions consisted of an initial denaturation at 95°C for 2 min followed by 30 cycles at 95°C for 1 min, 65°C for 1 min, 72°C for 2 min and a final elongation at 72°C for 5 min. PCR products were visualized following electrophoresis on 1.5% agarose gels. PCR positive colonies were grown overnight in 2ml NZY broth supplemented with 100µg/ml Ampicillin at 37°C with 225rpm shaking. Plasmid DNA was extracted using a Wizard Plus Minipreps DNA Purification System (Promega, USA) as per manufacturer’s instruction and sequenced to check for genomic integrity of the vector/insert construct. Sequencing reactions consisted of Big Dye Terminator (Applied Biosystems, USA), premix buffer, 0.25µM primer (Table 7 and Table 10), and approximately 75ng/300bp of DNA. Cycling conditions consisted of 25 cycles of 96°C for 10 sec, 50°C for 5 sec, and 60°C for 4 min. Sequencing products were purified and analyzed at the University of New South Wales DNA analysis facility (Figure 16).

97

Figure 16 Overview of cloning UL70 into pBluescript 98

4.2.2 Construction of baculovirus entry clone for HCMV primase expression

4.2.2.1 Viral stocks and DNA extraction Plasmids containing the UL70 ORF were extracted from the UL70a pBluescript II stock using the Wizard Plus Midipreps DNA Purification System (Promega, USA) as per manufacturer’s instruction. Briefly, plasmids were inoculated and grown overnight in 5ml NZY broth supplemented with 10mg/ml Ampicillin at 37°C with 225rpm shaking. The cells are centrifuged at 10,000xg for 10 min and the supernatant is discarded. Cell re-suspension, lysis, and neutralization solutions are added to lyse the cells and precipitate the protein from solution. The lysate was centrifuged at 14,000xg for 15 min and the supernatant is decanted to a new microfuge tube for purification. The DNA is transferred to a Wizard midi-column (Promega, USA) and purified under vacuum with diatomaceous resin and washed with ethanol before a final elution with 100µl hot (65°C -70°C) distilled water (Baxter, Australia).

4.2.2.2 Polymerase chain reaction and purification of UL70 for pDONR The full length UL70 ORF was amplified from midipreps of the UL70a pBluescript stock with primers which incorporated attachment sites into the product (Table 11). A single round PCR to amplify the ~3kb product was carried out in a 50µl reaction consisting of 1x High Fidelity buffer (Invitrogen, USA), 0.2mM dNTPs, 2mM MgSO4, 400nM primers, 1U Platinum Taq High Fidelity enzyme (Invitrogen, USA) and 5µl DNA template. The single round cycling conditions consisted of denaturation at 94°C for 3 min followed by 35 cycles of 94°C for 30 sec, 55°C for 30 sec, and 68°C for 3 min. PCR products were visualized following electrophoresis on 1.0% agarose gels and purified using a 30% PEG 8000/30mM MgCl2 Solution (Invitrogen, USA). Briefly, this involved diluting the PCR product 4-fold with TE buffer, followed by adding 0.5 times volume of 30% PEG 8000/30mM MgCl2 solution. The mixture is centrifuged for 15 min at 8000xg, with the supernatant removed afterwards and the clear pellet suspended in 20µl TE.

99 Table 11 Nucleotides primers used for UL70/pDONR cloning

Primer Name Primer Sequence* UL70attb1 5' - GGGGACAAGTTTGTACAAAAAAGCAGGCTTCCTGCATCGTGCCGGCGCGACG UL70attb2 5' - CCCCTGGTGAAACATGTTCTTTCGACCCAGCGCCGTGGAAAGTGAGGCTAG UL70.3 5' - GCGAGCCCAGTAGCAGACGCGCGAA UL70.4 5' - CCCAACTGGATCTGCGTAACCTGCT M13 Forward 5' - GTAAAACGACGGCCAG M13 Reverse 5' - CAGGAAACAGCTATGAC * Underlined sequences denote attachment sites

4.2.2.3 pDONR recombination and molecular cloning of UL70 The UL70 PCR product was recombined with the pDONR donor vector (Invitrogen, USA) (Figure 17) using a Gateway BP Clonase II enzyme (Invitrogen, USA).

Figure 17 pDONR/Zeocin Vector pDONR/Zeocin vector with key features including: pUC Ori: pUC Origin; Zeocin: Zeocin resistance gene; T1/T2: Transcription termination sequence; attp1/2: Recombination sites; ccdB: ccdB gene (negative

100 selection of plasmid); CmR: Chloramphenicol resistance gene; M13 Forward/Reverse: Sequencing priming sites.

A 10µl BP reaction consisting of 20-50fmol PCR product (Figure 18), 150ng pDONR vector, TE Buffer, and 2µl BP Clonase II enzyme mix (Invitrogen, USA) was incubated overnight at 25°C, followed by the addition of 1µl Proteinase K (Invitrogen, USA) and a further incubation of 37°C for 10 min. A pEXP7-tet positive control was used in conjunction with each experiment. A 1-2µl sample of the BP reaction was then immediately transformed to Library Efficient DH5α competent cells (Invitrogen, USA). Briefly, this involved thawing a 50µl aliquot of cells on ice before adding the sample and heat shocking the reaction at 42°C for 30 sec followed by the addition of 250µl SOC media pre-warmed to 37°C. The cells were grown in a shaking incubator at 37°C and 225rpm for 1 hr prior to plating on LB-Zeocin agar plates while positive controls were plated on LB-tetracycline plates. All plates were incubated overnight at 37°C.

Figure 18 converting femtomoles to nanograms

Single colonies were re-streaked and screened via colony PCR using an internal UL70 primer and a pDONR M13 primer (Table 11). A single round colony PCR as outlined in section 4.2.1.3 was carried out to check for the insert. PCR positive colonies were grown overnight in 5ml SOC media supplemented with 50µg/ml Zeocin at 37°C with 225rpm shaking. Plasmid DNA was extracted using a Wizard Plus Midipreps DNA Purification System (Promega, USA) as per manufacturer’s instruction and sequenced to check for genomic integrity of the construct. Sequencing reactions and cycling conditions were previously outlined in section 4.2.1.3. A brief overview of the process is outlined in Figure 19.

101

Figure 19 Overview of UL70/pDONR recombination

4.2.3 In vitro UL70 primase expression using TNT Quick Coupled Transcription/Translation

4.2.3.1 Viral stocks and DNA extraction Plasmids containing the UL70 ORF were extracted from the UL70a pBluescript II stock using the Wizard Plus Midipreps DNA Purification System (Promega, USA) as per manufacturer’s instruction and as outlined in previous sections.

102 4.2.3.2 Polymerase chain reaction and purification of UL70 The full length UL70 ORF was amplified from midipreps of the UL70a pBluescript stock using primers which incorporated the BamHI and SalI restriction sites at the 5’ and 3’ ends of UL70 respectively (Table 12). A single round PCR to amplify the ~2.9kb product was carried out in a 50µl reaction consisting of 1x High Fidelity buffer

(Invitrogen, USA), 0.2mM dNTPs, 2mM MgSO4, 400nM primers, 1U Platinum Taq High Fidelity enzyme (Invitrogen, USA) and 5µl DNA template. The single round cycling conditions consisted of denaturation at 94°C for 3 min followed by 35 cycles of 94°C for 30 sec, 60°C for 30 sec, and 68°C for 3 min. PCR products were visualized following electrophoresis on 1.0% agarose gels and purified using a Wizard SV Gel and PCR Clean-Up Kit (Promega, USA) as per manufacturer’s protocols. Briefly, the PCR products were mixed in a 1:1 ratio with Membrane Binding Solution (Promega, USA), incubated at room temperature and transferred to a SV mini-column (Promega, USA) where it is centrifuged at 16,000xg to collect the DNA in the membrane. The membrane is washed several times with Membrane Wash Solution (Promega, USA) before a final elution of 30µl with nuclease-free water (Promega, USA).

Table 12 Nucleotide primers used for UL70/pCITE cloning

Primer Name Primer Sequence* UL70pCITE.3 5' - ATTGGATCCCTGCATCGTGCCGGCGCGACG UL70pCITE.4 5' - ATTGTCGACTGACGGCGGTCGCCGGCGGCAT UL70.3 5' – GCGAGCCCAGTAGCAGACGCGCGAA T7 5' - GTAATACGACTCACTATAGGGC * Underlined sequences denote restriction sites

4.2.3.3 Molecular cloning of UL70 into pCITE 4A Both the UL70 PCR product and pCITE vector were sequentially digested, initially with SalI followed by BamHI. Restriction digests with SalI were carried out in a 20µl reaction consisting of 1x NEBuffer 3 (New England BioLabs (NEB), USA), 100µg/ml BSA (NEB, USA), 30U SalI (NEB, USA), and ~3µg DNA, or a 10U/µg DNA over- digest. Digest reactions were incubated at 37°C for 3 hrs followed by heat inactivation at 65°C for 20 min. Digest products were then purified using the Wizard SV Gel and

103 PCR Clean-Up Kit (Promega, USA) as per manufacturer’s protocols outlined in the previous section prior to the second digest. Restriction digests with BamHI were also carried out in 20µl reaction consisting of 1x NEBuffer 3 (NEB, USA), 100µg/ml BSA (NEB, USA), 20U BamHI (NEB, USA), and ~2µg DNA, or a 10U/µg DNA over-digest. Digest reactions were incubated at 37°C for 2 hrs and immediately followed by processing with the Wizard SV Gel and PCR Clean-Up Kit (Promega, USA) as a means to stop the digest reaction and purify the restriction products. Products were ligated in a 20µl reaction consisting of 1x ligase duffer (Stratagene, USA), 1mM rATP (Stratagene, USA), 4U T4 DNA ligase (Stratagene, USA), and DNA in an insert to vector ratio of 3:1 as calculated by the formula in section 4.2.1.3. A reaction consisting of pUC18 plasmid in place of the sample was used as a ligation control. The ligation reactions were incubated at room temperature for 16 hrs followed by transformation to Library Efficient DH5α Competent Cells (Invitrogen, USA) as per manufacturer’s protocol, albeit omitting the dilution steps. Briefly, this involved thawing a 100µl aliquot of cells on wet ice before adding 1-10ng DNA and incubated on ice for 30 min. The mixture was heat shocked at 42°C for 45 sec followed by the addition of 900µl SOC media. The cells were grown in a shaking incubator at 37°C and 225rpm for 2 hrs prior to plating on LB-Ampicillin agar plates.

Single colonies were re-streaked and screened via colony PCR using an internal UL70 primer and a pCITE T7 primer (Table 12). A single round colony PCR as outlined in section 4.2.1.3 was carried out to check for the insert. PCR positive colonies were grown overnight in 5ml SOC media supplemented with 100µg/ml Ampicillin at 37°C with 225rpm shaking. Plasmid DNA was extracted using a Wizard Plus Midipreps DNA Purification System (Promega, USA) as per manufacturer’s instruction and sequenced to check for genomic integrity of the construct. The sequencing reaction and cycling conditions used were previously outlined in section 4.2.1.3. An overview of the process is outlined in Figure 20.

104

Figure 20 Overview of cloning UL70 into pCITE expression vector

105 4.2.3.4 Protein expression and detection UL70 expression was carried out using a TNT Quick Coupled Transcription/Translation Kit (Promega, USA) designed to express protein in an in vitro environment. Protein expression was carried out in 50µl reactions consisting of 40µl TNT Quick Master-Mix (Promega, USA), 20µM methionine (Promega, USA), and ~1ng DNA. The reaction was incubated at 30°C for 90 min, followed by a 1:4 dilution with 4X loading dye and finally denatured at 100°C for 5 min. A T7 luciferase which produced a 61kDa monomeric protein was used as a translation control. Translation products were visualized with a 10% SDS-PAGE resolving gel using a Mini-PROTEAN 3 Cell electrophoresis system (Biorad, USA). The protein samples were loaded alongside 10µl Kaleidoscope markers (Biorad, USA) and Magic Mark protein markers (Novagen, Germany). All samples were run through the 5% stacking gel at a constant voltage of 80V for 30 min prior to running the samples through the resolving gel at a constant voltage of 110V until the dye front left the gel. Protein bands were transferred from the gel to nitrocellulose membrane using Western blotting techniques. The transfer was run overnight at a constant current of 90mA (30V) using a magnetic stirrer and ice-pack to keep the system from overheating. The gel was stained with 20ml Gelcode Blue Stain (Pierce, USA) post-transfer to check for total protein. The S-Tag and His Tag antibodies were used to detect for the UL70 protein product. Briefly, the nitrocellulose membrane was incubated with gentle shaking in blocking solution for 1 hr, followed by thorough washing with PBST solution. For membranes being detected with the S-Tag, the S- Protein Alkaline Phosphatase Conjugate (Novagen, Germany) was diluted 1:5000 with PBST and incubated with the membrane at room temperature for 30 min with gentle shaking. Detection with the S-Tag did not require a secondary antibody. For membranes being detected with the His-Tag, c-terminal Anti-His antibodies (Invitrogen, USA) were diluted 1:5000 with PBST and incubated with the membrane at room temperature for 1- 2 hrs with gentle shaking. The membrane was washed thoroughly with PBST prior to incubation with the secondary antibody. Immun-Star goat anti-mouse IgG-HRP (Australian Laboratory Services, Australia) was diluted 1:4000 with blocking solution and incubated with the membrane at room temperature with gentle shaking for 1 hr. The membranes are thoroughly washed again with PBST after antibody incubation and prepared for chemiluminescence visualization with SuperSignal West Pico Chemiluminescent Substrate (Pierce, USA), as per manufacturer’s instructions. Briefly,

106 a working solution is prepared by mixing the Luminol/Enhancer Solution and the Stable Peroxide Solution in a ratio of 1:1. The working solution is then carefully applied over the membrane using a pipette and incubated at room temperature for 5 min. The excess reagent is drained and the membrane is exposed to a camera set to detect chemiluminescence (Figure 21). The membrane was also stained with Ponceau S (Sigma-Aldrich, USA) on several occasions to check for total protein.

Figure 21 Overview of UL70 in vitro protein expression

107 4.3 Results

4.3.1 UL70 was cloned into pBluescript II high-copy vector Both the UL70 ORF spanning 2,985bp from high-passage isolate 5B and the 5,250bp fragment from high-passage isolate 29A were successfully cloned into pBluescript and designated as UL70a and UL70b respectively. UL70a consisting of the 3kb pBluescript vector and the 3kb UL70 insert (Figure 22A) produces a 6kb band when extracted from a plasmid preparation. In addition, we also detected an extra band several kilo-bases heavier than our anticipated product indicating that nicked DNA may also be present (Figure 22B). However, sequencing of the plasmid confirmed the insert ligated correctly into the vector without mutations.

A B

Figure 22 pBluescript UL70a

(A) UL70a plasmid map indicating position of UL70 (red) relative to key vector domains (orange). (B) UL70a produces two products bands, the anticipated plasmid (II) and nicked DNA (I). Lambda DNA HIndIII markers (Promega) on the left are indicated in base-pairs.

UL70b consists of the 3kb pBluescript vector and the 3kb UL70 ORF as well as approximately 1kb of flanking genomic region either end of the UL70 ORF (Figure 23A). Plasmid preparations produces an anticipated band of approximately 8kb as well

108 as a high molecular weight product predicted to be nicked DNA (Figure 23B). Sequencing of the plasmid confirmed that the 5kb insert ligated into pBluescript and without mutations.

A B

Figure 23 pBluescript UL70b

(A) UL70b plasmid map indicating position of UL70 (red) relative to key vector domains (orange). (B) UL70b produces two products bands, the anticipated plasmid (II) and nicked DNA (I). Lambda DNA HIndIII markers (Promega) Markers on the right are indicated in base-pairs.

109 4.3.2 Generating Baculovirus entry clones via BP recombination Experiments to create a UL70/pDONR recombinant have been unsuccessful with several instances of the PCR incorporated attachment site failing to bind to their correct complement on the pDONR vector, resulting in recombinants which truncated the UL70 ORF. As a result, subsequent experiments to recombine the pDONR/UL70 entry clone with the baculovirus expression DNA have not been pursued. The length of the excised regions varied, with designate samples BPR4, BPR5, and BPR8 missing 1,919bp, 36bp, and 505bp respectively, all from the 3’ end of UL70. However, the 5’ end of UL70 recombined with pDONR without problems. Other attempts to generate an entry clone failed to produce colonies at the screening stage despite the transformation controls producing visible colonies. Experiment BPR1 and BPR2 both used BP Clonase I (Invitrogen, USA) in the BP recombination reaction (c.f. BP Clonase II in later experiments), with the former transformed into One Shot TOP10 competent cells (Invitrogen, USA) while latter experiments used the DH5α competent cells. BPR3 was carried out as a time-course experiment to determine the optimal BP reaction time, and though colonies were generated at the longest incubation, all samples were found negative for UL70. Samples BPR4 and BPR5 were generated using increasing amounts of total DNA during the BP reaction, with BPR5 grown for an additional hour during the transformation step to increase the number of cells prior to plating. BPR4 was selected from several colonies of which only one produced a band at the expected size of 780bp when screened using M13 forward and UL70.3 reverse primers (Figure 24A). However, plasmid preparations of the construct revealed a prominent band at 2.3kb, the expected band of 4.9kb, as well as a band exceeding 9kbp predicted to be nicked DNA (Figure 24B).

110 A B

Figure 24 Screening of entry clone BPR4

(A) PCR screening of BPR4 generated the expected 780bp fragment. (B) Plasmid preparation of BPR4 revealed 3 products: a band >9kbp predicted to be nicked DNA (I), band at ~6kbp predicted to be the expected entry clone, and an unknown prominent band at 2.3kbp (III). Lambda DNA HIndIII Markers (Promega) (right) and BenchTop pGEM DNA markers (Promega) (left) sizes are indicated in base-pairs.

Sequence analysis of the BPR4 plasmid preparation revealed a truncated 3kb product with UL70 missing the C-terminal 1,919bp, albeit retaining frame with the rest of pDONR. The initial PCR screening was a false positive as the region amplified did not involve the truncation. In addition, the truncation removes the catalytic motif and the putative zinc finger binding domain (Figure 25).

Figure 25 The BPR4 truncation removes key motifs from UL70 Key UL70 motifs highlighted in purple while red box denotes section of gene truncated during recombination. Orientation and length of te UL70 protein is denoted by numbers.

111 Despite the unfavorable recombination in BPR4, another attempt was carried out on the assumption that increasing the overall number of cells prior to plating would increase all populations universally, including the constructs which contained our desired product. In addition, PCR screening of the colonies involved the use of both the M13 forward/UL70.3 reverse primers as well as a set of internal UL70 primers such that colonies containing the truncation in BPR4 would be selected out. BPR5 was selected on the basis of colony PCR (Figure 26).

Figure 26 Colony PCR screening for entry clone BPR5

Internal UL70 primers were used to amplify a 1.2kbp region as well as intrinsically select out recombinants carrying the BPR4 truncation. BenchTop pGEM DNA marker (Promega) sizes on the right are indicated in base-pairs.

However, after sequencing several BPR5 samples, a design flaw was identified at the 3’ end of the UL70 insert product, which caused a short truncation and amplification of non-specific DNA (Figure 27). An inverted attachment sequence was identified for the 3’ primer, resulting in complications during the amplification as well as the subsequent BP recombination.

112

UL70 junk Primer attb1 pDONR ------|------|------|------|------

Figure 27 The BPR5 truncation of UL70

A primer design flaw caused the 36bp truncation of UL70 as well as the amplification of non-specific sequences (green). Key UL70 motifs highlighted in purple while red box denotes section of gene truncated during recombination. Orientation and length of the UL70 protein is denoted by numbers.

Subsequent experiments, including BPR6, BPR7 and BPR8 were carried out using a modified primer with the correct attachment sequence. BPR6 and BPR7 failed to generate colonies despite using the previously optimized protocol while BPR8, transformed in a new batch of DH5α cells produced several colonies. PCR positive colonies were sequenced, revealing another novel truncation at the 3’ end of UL70. A 505bp fragment at the c-terminal end of UL70 was shown to be excised, removing a putative zinc finger domain in the process (Figure 28).

Figure 28 The BPR8 truncation removes the putative zinc finger from UL70

Key UL70 motifs highlighted in purple while red box denotes section of gene truncated during recombination. Orientation and length of the UL70 protein is denoted by numbers.

113 4.3.3 UL70 was cloned into pCITE expression vector The full length UL70 ORF was successfully cloned into expression vector pCITE 4A with a functional S-Tag and c-terminal His-tag (Figure 29A). The original UL70 stop codon, TGA, was replaced with TCA to ensure that subsequent translation incorporated the down-stream His-tag (Figure 29A). Plasmid preparations of the DNA gave a distinct band at the expected size of 6.7kbp (Figure 29B) while sequencing confirmed that the 3kbp insert ligated into pCITE without mutations.

A B

Figure 29 UL70/pCITE expression vector (A) UL70/pCITE plasmid map indicating position of UL70 (red) relative to key vector domains (orange). (B) UL70/pCITE produces one distinct band at 6.7kbp. Lambda DNA HIndIII markers (Promega) on the right are indicated in base-pairs.

114 4.3.4 UL70 in vitro expression using the TNT system The expression vector created in 4.3.3 was used with the TNT Quick Coupled Transcription/Translation Kit (Promega, USA) to express the HCMV primase. However, attempts at protein expression failed to produce a protein of the predicted 107kDa size. In addition, an unknown product of ~55kDa was detected in all lanes (Figure 30). The method of detection was by using the S Tag protein, which is present in the pCITE expression vector, but not in the luciferase control DNA, hence the control reaction was not carried out.

Figure 30 Expression of UL70 in TNT system

Typical protein expression produces an unknown product of ~55kDa (I) in both the UL70 sample (A) and the negative control (B). Precision Plus Kaleidoscope markers (Biorad) on the left are given in Kilo- Daltons (kDa).

Since previous attempts were based on a single incubation time of 90 min as recommended per manufacturer’s instructions, we carried out a time course experiment to determine the incubation time that would yield the optimal amount of protein.

115 However, samples taken at 30 min intervals up to duration of 3 hrs yielded similar results to incubations at 90 min (Figure 31).

Figure 31 Time course expression of UL70 with TNT system

Samples were taken at 30 min intervals starting at 1hr (A) to 3 hrs (E). The negative control was taken at 3 hrs (F). Markers on the left are given in kDa.

116 4.4 Discussion

The ability to replicate the UL70 ORF at high yields using a high copy vector as opposed to traditional cell culturing is an invaluable tool for elucidating the functions of the primase in future expression studies. UL70b, which contains flanking regions either side of the 5’ and 3’ ends of the UL70 ORF, was designed as a precursor to express the UL70 gene in a Bacterial Artificial Chromosome (BAC) system. The BAC system has been shown to be an effective tool for studying the importance of various genes in HCMV (AuCoin, 2006; Hahn, 2002; Hahn, 2004; Jiang, 2008; Komazin, 2004; Lorz, 2006; Martin, 2006; Michel, 2005; Stropes, 2008; Xu, 2004) and the creation of a UL70 precursor may serve as a contribution to future studies on the HCMV helicase-primase complex, which has yet to be studied in detail.

Protein expression involving baculovirus requires the creation of an expression virus through recombination between an entry clone and the baculovirus linear DNA. The entry clone is created through another recombination reaction involving an entry vector and a PCR generated DNA fragment utilizing attachment sites. In this study, attempts to generate entry clones for this process resulted in recombinants that truncated UL70 and although initial results reflected the consequences of a faulty primer, later experiments persisted with the truncation of the 3’ end. The original 3’ primer sequence was found to have an inverted attachment sequence, which resulted in the PCR fragment failing to recognize the complement sequence on pDONR. Initially this was thought to have caused the truncations as demonstrated in BPR4 and BPR5. However, closer inspection of the plasmid sequences suggests that the truncations may have been caused by the recognition sequences binding to domains within UL70 that were similar to the attachment sites (Figure 32). During sequence analysis we noticed that the recombination site remained continuous between UL70 and pDONR, indicating that the plasmid was circularized. In addition, BPR4 remained in frame with pDONR, suggesting that expression of the partial UL70 protein should still be possible, and indeed could be used to determine if the N-terminal third of the HCMV primase played a role in primase function or in interacting with other proteins, in particular, other members of the primase-helicase complex. Unfortunately, the design of the baculovirus system meant that BPR4 could not be carried forward to express protein since the recombination between the entry clone and baculovirus linear DNA requires another

117 recombination sequence that is created upon successful recombination between the PCR fragment and pDONR.

A GGACGTCATGCAAAAATACTTCTCGCTCGACAACTTTCTACACGATTACGTGGAGACGCATCTACTACGT | | || |||||| | | CCGTCCTAGCCTCACTTTCCACGGCGGACCCAGCTTTCTTGTACAAAGTTGGCATTATAAGAAAGCATTG

B AAGCCAACTTTATCGAGACGCGCTCGCTTAACGTGACGCGTTATCGACGCCGCGGTCTCACCGAGGTGCT | ||||| | | | |||| || | | | AGCCTCACTTTCCACGGCGGACCCAGCTTTCTTGTACAAAGTTGGCATTATAAGAAAGCATTGCTTATCA Figure 32 Sequence analysis of BPR4 and BPR8

Comparisons of the recombination site for BPR4 (A) and BPR8 (B) reveals a degree of similarity between the expected recombination sites (grey) and the random sequence within UL70 (as underlined) that marked the recombination site for BPR4 and BPR8 with pDONR (purple).

Experiments to express the HCMV primase failed to produce detectable levels of the 107kDa protein. The absence of detectable protein could have been the result of low translation efficiency caused by contaminants. In particular, the presence of calcium may reactivate nucleases in the TNT lysate mix that were used to remove endogenous mRNA and result in degradation of the DNA template (TNT Technical manual). The unexpected 55kDa band which appeared during expression may indicate that another peptide was translated from the template due to translation initiating at internal methionines or that an artifact was created by the methods used for expression. Increasing the fidelity for the correct initiation methionines can done by optimizing the Mg2+ or K+ concentration (Hurst, 1996). Antibody detection was done using the S-Tag system exclusively, which was a unique feature of the pCITE vector. The S-Tag system is based on the interaction between a 15 amino acid S-Tag peptide and a ribonuclease S- protein, which differs from current conventional His-Tag antibody detection in that it does not require a secondary anti-body and hence thought to reduce the amount of background on the Western Blot.

118 5 Concluding discussion and future directions

In this study we have shown that essential HCMV replication genes UL44, UL57, UL70, UL102, UL105, and UL84 are highly conserved amongst isolates with variations occurring outside of known or otherwise predicted functional domains. We have also shown that the certain components of HCMV primase-helicase complex, UL70 and UL105, as well as the single-stranded DNA binding protein (UL57) were highly conserved amongst homologues, affirming the premise that functional domains remain conserved as species diverged. Then putative functional motifs predicted from protein motif scanning software were used as a basis for identifying potential functional domains amongst the homologues under this premise.

We identified two N-linked glycosylation sites and a casein-kinase II phosphorylation site that were highly conserved amongst 33 homologues of the HCMV helicase (UL105). In addition, a cAMP-dependent protein kinase phosphorylation site was identified in motif IV of the six putative helicase motifs previously predicted to exist in UL105 (Graves-Woodward, 1996), providing some insight into the mechanistic properties of that domain. In UL70, we identified a highly conserved domain consisting of aromatic residues preceding a conserved C570 residue that has been implicated in the structural inhibition of a compound currently being studied for its inhibitory effects on the HCMV primase (Chen, 2007). In addition, a pair of highly conserved cysteines was also identified although guesses as to their function, if any, are highly speculative at this stage of the study. On one side, the domain is reminiscent of a rare PDI motif found in only a handful of organisms in nature (Woycechowsky, 2003) while on the other, it may be an important motif for protein structure, interacting with other conserved cysteines, possibly with C570, or cysteines from the putative zinc finger domain (Chen, 2005), to stabilize the tertiary structure in a manner that is similar to that of the chemokine class of proteins (Fernandez, 2002). Similar conserved cysteine residues were previously predicted to reside in UL57 (Gao, 1988) and were shown here to be highly conserved across 35 homologues. In particular, a classic disulfide bond motif with the canonical CxxC sequence was shown to be present in 34 of the 35 homologues studied and may serve as the starting point for understanding the structural properties for this class of proteins.

119 While conservation across homologous proteins may act as an indicator of evolutionarily conserved functional domains, the lack of homology may also demonstrate the ability of genes to adapt over time to specific roles unique to a particular species. The DNA polymerase accessory proteins of HSV (UL42) and HCMV (UL44) are both well characterized and document this distinction well. While homologous in function and position (Ertl, 1992), UL42 operates as a monomer (Gottlieb, 1990) and relies on basic interactions (Bridges, 2001; Zuccola, 2000), while UL44 was shown to form a dimer and rely on hydrophobic interactions (Appleton, 2004). Similarly, UL102 shares no homology with its HSV equivalent, UL8, despite the primase and helicase components sharing homologies of 27% and 34% respectively to the HSV equivalents (Chee, 1990) while UL84 has no homolog equivalent in other herpesviruses, despite being an essential component of HCMV DNA replication (Pari, 1993). Although we cannot use conservation as a means to identify putative domains for these genes, the putative motifs predicted by the scanning software in this study can be used as a reference for future studies elucidating protein function.

In this study, three different protein expression techniques were explored and though we failed to express the HCMV primase, insights can be gained from the methods used in constructing the vectors and attempts at expression in vitro. In addition, the creation of a high copy vector for UL70 featuring recombinant arms can be the basis of future experiments to express the HCMV primase in a BAC system.

120 6 References

Abdel-Haq, N., Chearskul, P., Al-Tatari, H., Asmar, B. (2006) New antiviral agents. Indian Journal of Pediatrics 73(4), 313-21. Adamo, J.E., Schroer, J., Shenk, T. (2004) Human cytomegalovirus TRS1 protein is required for efficient assembly of DNA-containing capsids. Journal of Virology 78(19), 10221-9. Adjei, A., Armah, H., Narter-Olaga, E. (2006) Seroprevalence of cytomegalovirus among some voluntary blood donors at the 37 military hospital, Accra, Ghana. Ghana Medical Journal 40(3), 99-104. Adler, S.P. (1988) Cytomegalovirus transmission among children in day care, their mothers and caretakers. The Pediatric Infectious Disease Journal 7(4), 279-85. Adler, S.P. (1989) Cytomegalovirus and child day care. Evidence for an increased infection rate among day-care workers. The New England Journal of Medicine 321(19), 1290-6. Ahlfors, K., Ivarsson, S.A., Harris, S. (1999) Report on a long-term study of maternal and congenital cytomegalovirus infection in Sweden. Review of prospective studies available in the literature. Scandinavian Journal of Infectious Diseases 31(5), 443-57. Alanen, A., Kahala, K., Vahlberg, T., Koskela, P., Vainionpää, R. (2005) Seroprevalence, incidence of prenatal infections and reliability of maternal history of varicella zoster virus, cytomegalovirus, herpes simplex virus and parvovirus B19 infection in South-Western Finland. BJOG: An International Journal of Obstetrics and Gynaecology 112(1), 50-6. Altschul, S.F., Madden, T.L., Schäffer, A.A., Zhang, J., Zhang, Z., Miller, W. & Lipman, D.J. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acid Research 25, 3389-3402. Alvisi, G., Jans, D.A., Guo, J., Pinna, L.A., and Ripalti, A. (2005) A protein kinase CK2 site flanking the nuclear targeting signal enhances nuclear transport of human cytomegalovirus ppUL44. Traffic 6(11), 1002-1013. Anders, D.G., Gibson, W. (1988) Location, transcript analysis, and partial nucleotide sequence of the cytomegalovirus gene encoding an early DNA-binding protein with similarities to ICP8 of herpes simplex virus type. Journal of Virology 62(4), 1364-1372. Anders, D.G., Kacica, M.A., Pari, G., Punturieri, S.M. (1992) Boundaries and structure of human cytomegalovirus oriLyt, a complex origin for lytic-phase DNA replication. Journal of Virology 66(6), 3373-84. Anders, D.G., Kidd, J.R., Gibson, W. (1987) Immunological characterization of an early cytomegalovirus single-strand DNA-binding protein with similarities to the HSV major DNA-binding protein. Virology 161(2), 579-88. Anders, D.G., McCue, L.A. (1996) The human cytomegalovirus genes and proteins required for DNA synthesis. Intervirology 39(5-6), 378-88. Andersson-Ellström, A., Svennerholm, B., Forssman, L. (1995) Prevalence of antibodies to herpes simplex virus types 1 and 2, Epstein-Barr virus and cytomegalovirus in teenage girls. Scandinavian Journal of Infectious Diseases 27(4), 315-8. Appleton, B.A., Brooks, J., Loregian, A., Filman, D.J., Coen, D.M., and Hogle, J.M. (2006) Crystal structure of the cytomegalovirus DNA polymerase subunit UL44

121 in complex with the C terminus from the catalytic subunit. The Journal of Biological Chemistry 281(8), 5224-5232. Appleton, B.A., Loregian, A., Filman, D.J., Coen, D.M., and Hogle J.M. (2004) The cytomegalovirus DNA polymerase subunit UL44 forms a C clamp-shaped dimer. Molecular Cell 15, 233-244. Arezi, B., Kuchta, R.D. (2000) Eukaryotic DNA primase. Trends in Biochemical Sciences 25(11), 572-6. Arvin, A. (2007) Human Herpesviruses: Biology, Therapy, and Immunoprophylaxis, 1408 pp. 1st ed, edited by A. Arvin, Mocarski, E., Moore, P.S., Roizman, B., Whitley, R., Yamanishi, K., 1. 1 vols. Cambridge University Press, New York. AuCoin, D.P., Smith, G.B., Meiering, C.D., Mocarski, E.S. (2006) Betaherpesvirus- conserved cytomegalovirus tegument protein ppUL32 (pp150) controls cytoplasmic events during virion maturation. Journal of Virology 80(16), 8199- 210. Baldanti, F., Silini, E., Sarasini, A., Talarico, C.L., Stanat, S.C., Biron, K.K., Furione, M., Bono, F., Palù, G., Gerna, G. (1995) A three-nucleotide deletion in the UL97 open reading frame is responsible for the ganciclovir resistance of a human cytomegalovirus clinical isolate. Journal of Virology 69(2), 796-800. Baldanti, F., Underwood, M.R., Talarico, C.L., Simoncini, L., Sarasini, A., Biron, K.K., Gerna, G. (1998) The Cys607-->Tyr change in the UL97 phosphotransferase confers ganciclovir resistance to two human cytomegalovirus strains recovered from two immunocompromised patients. Antimicrobial Agents and Chemotherapy 42(2), 444-6. Baldwin, B.R., Zhang, C.O., Keay, S. (2000) Cloning and epitope mapping of a functional partial fusion receptor for human cytomegalovirus gH. The Journal of General Virology 81(1), 27-35. Bale, J.F.J., Zimmerman, B., Dawson, J.D., Souza, I.E., Petheram, S.J., Murph, J.R. (1999) Cytomegalovirus transmission in child care homes. Archives of Pediatrics and Adolescent Medicine 153(1), 75-9. Bar, M., Shannon-Lowe, C., Geballe, A.P. (2001) Differentiation of human cytomegalovirus genotypes in immunocompromised patients on the basis of UL4 gene polymorphisms. Journal of Infectious Diseases 183(2), 218-225. Barnard, E.C., Brown, G., and Stow, N.D. (1997) Deletion mutants of the herpes simplex virus type 1 UL8 protein effect on DNA synthesis and ability to interact with and influence the intracellular localization of the UL5 and UL52 proteins. Virology 237, 97–106. Barrasa, M.I., Harel, N.Y., Alwine, J.C. (2005) The phosphorylation status of the serine-rich region of the human cytomegalovirus 86-kilodalton major immediate-early protein IE2/IEP86 affects temporal viral gene expression. Journal of Virology 79(3), 1428-37. Bause, E. (1983) Structural requirements of N-glycosylation of proteins. Studies with proline peptides as conformational probes. The Biochemical Journal 209(2), 331-6. Beresford, C.H., Storey, G., Faed, J.M., Milligan, L.M. (1988) Cytomegalovirus infection and blood transfusion. The New Zealand Medical Journal 101(838), 2- 4. Bernad, A., Blanco, L., Lázaro, J.M., Martín, G., Salas, M. (1989) A conserved 3'----5' exonuclease active site in prokaryotic and eukaryotic DNA polymerases. Cell 59(1), 219-28.

122 Biggar, R.J., Andersen, H.K., Ebbesen, P., Melbye, M., Goedert, J.J., Mann, D.L., Strong, D.M. (1983) Seminal fluid excretion of cytomegalovirus related to immunosuppression in homosexual men. British Medical Journal 286(6383), 2010-2. Biron, K.K. (2006) Antiviral drugs for cytomegalovirus diseases. Antiviral Research 71(2-3), 154-63. Biron, K.K., Harvey, R.J., Chamberlain, S.C., Good, S.S., Smith, A.A. 3rd, Davis, M.G., Talarico, C.L., Miller, W.H., Ferris, R., Dornsife, R.E., Stanat, S.C., Drach, J.C., Townsend, L.B., Koszalka, G.W. (2002) Potent and selective inhibition of human cytomegalovirus replication by 1263W94, a benzimidazole L-riboside with a unique mode of action. Antimicrobial Agents and Chemotherapy 46(8), 2365-72. Biswas, N., and Weller, S.K. (1999) A mutation in the C-terminal putative Zn2+ finger motif of UL52 severely affects the biochemical activities of the HSV-1 helicase- primase subcomplex. The Journal of Biological Chemistry 274(12), 8068–8076. Blankenship, C.A., Shenk, T. (2002) Mutant human cytomegalovirus lacking the immediate-early TRS1 coding region exhibits a late defect. Journal of Virology 76(23), 12290-9. Boehmer, P.E. (1998) The herpes simplex virus type-1 single-stranded DNA-binding protein, ICP8, increases the processivity of the UL9 protein DNA helicase. The Journal of Biological Chemistry 273, 2676–2683. Boehmer, P.E., Lehman, I.R. (1993) Herpes simplex virus type 1 ICP8: helix- destabilizing properties. Journal of Virology 67, 711–715. Boehmer, P.E., Nimonkar, A.V. (2003) Herpes Virus Replication. Life 55, 13-22. Bogner, E. (2001) Human cytomegalovirus terminase as a target for antiviral chemotherapy. Reviews in Medical Virology 12(2), 115-27. Boivin, G., Gilbert, C., Gaudreau, A., Greenfield, I., Sudlow, R., Roberts, N.A. (2001) Rate of emergence of cytomegalovirus (CMV) mutations in leukocytes of patients with acquired immunodeficiency syndrome who are receiving valganciclovir as induction and maintenance therapy for CMV retinitis. The Journal of Infectious Diseases 184(12), 1598-602. Bortner, C., Hernandez, T.R., Lehman, I.R., Griffith, J. (1993) Herpes simplex virus 1 single-strand DNA-binding protein (ICP8) will promote homologous pairing and strand transfer. Journal of Molecular Biology 231, 241-250. Bridges, K.G., Chow, C.S., and Coen, D.M. (2001) Identification of crucial hydrogen- bonding residues for the interaction of herpes simplex virus DNA polymerase subunits via peptide display, mutational, and calorimetric approaches. Journal of Virology 75, 4990–4998. Britt, W.J., and Alford, C.A. (1996) Cytomegalovirus. Fields Virology, edited by B.N. Fields, Knipe, D.M., Howley, P.M. Lippincott-Raven Publishers, Philadelphia. Burley, S.K., Petsko, G.A. (1985) Aromatic-aromatic interaction: a mechanism of protein structure stabilization. Science 229(4708), 23-8. Bush, M., Yager, D.R., Gao, M., Weisshart, K., Marcy, A.I., Coen, D.M., Knipe, D.M. (1991) Correct intranuclear localization of herpes simplex virus DNA polymerase requires the viral ICP8 DNA-binding protein. Journal of Virology 65, 1082-1089. Butcher, S.J., Aitken, J., Mitchell, J., Gowen, B., Dargan, D.J. (1998) Structure of the human cytomegalovirus B capsid by electron cryomicroscopy and image reconstruction. Journal of Structural Biology 124(1), 70-6.

123 Calder, J.M., Stow, E.C., Stow, N.D. (1992) On the cellular localization of the components of the herpes simplex virus type 1 helicase-primase complex and the viral origin-binding protein. The Journal of General Virology 73(3), 531-8. Cervia, J.S., Wenz, B., Ortolano, G.A. (2007) Leukocyte reduction's role in the attenuation of infection risks among transfusion recipients. Clinical Infectious Diseases 45(8), 1008-13. Chee, M.S., Bankier, A.T., Beck, S., Bohni, R., Brown, C.M., Cerny, R., Horsnell, T., Hutchison, C.A., Kouzarides, T., Mertigneti, J.A., Preddie, E., Satchwell, S.C., Tomlinson, P., Weston, K.M., Barrell, B.G. (1990) Analysis of the protein- coding content of the sequence of human cytomegalovirus strain AD169. Current Topics in Microbiology and Immunology 154, 125-69. Chen, D.H., Jiang, H., Lee, M., Liu, F., Zhou, Z.H. (1999) Three-dimensional visualization of tegument/capsid interactions in the intact human cytomegalovirus. Virology 260(1), 10-6. Chen, X., Adrian, J., Cushing, T., DiMaio, D., Liang, L., Mayorga, V., Miao, S., Peterson, M.G., Powers, J.P., Spector, F., Stein, C., Wright, M., Xu, D., Ye Q., Jaen, J. (2007) SAR studies on a novel series of human cytomegalovirus primase inhibitors. Bioorganic & Medicinal Chemistry Letters 17(8), 2188-2192. Chen, Y., Carrington-Lawrence, S.D., Bai, P., Weller, S.K. (2005) Mutations in the putative zinc-binding motif of UL52 demonstrate a complex interdependence between the UL5 and UL52 subunits of the human herpes simplex virus type 1 helicase/primase complex. Journal of Virology 79(14), 9088-96. Chen, Y.M., Knipe, D.M. (1996) A dominant mutant form of the herpes simplex virus ICP8 protein decreases viral late gene transcription. Virology 221, 281-290. Child, S.J., Hakki, M., De Niro, K.L., Geballe, A.P. (2004) Evasion of Cellular Antiviral Responses by Human Cytomegalovirus TRS1 and IRS1. Journal of Virology 78(1), 197-205. Chiou, S.H., Yang, Y.P., Lin, J.C., Hsu, C.H., Jhang, H.C., Yang, Y.T., Lee, C.H., Ho, L.L., Hsu, W.M., Ku, H.H., Chen, S.J., Chen, S.S., Chang, M.D., Wu, C.W., Juan, L.J. (2006) The immediate early 2 protein of human cytomegalovirus (HCMV) mediates the apoptotic control in HCMV retinitis through up- regulation of the cellular FLICE-inhibitory protein expression. Journal of Immunology 177(9), 6199-206. Chou, S. (1999) Antiviral drug resistance in human cytomegalovirus. Transplant Infectious Disease 1(2), 105-14. Chou, S., Dennison, K. (1991) Analysis of interstrain variation in cytomegalovirus glycoprotein B sequences encoding neutralization-related epitopes. Journal of Infectious Diseases 163, 1229-1234. Chou, S., Erice, A., Jordan, M.C., Vercellotti, G.M., Michels, K.R., Talarico, C.L., Stanat, S.C., Biron, K.K. (1995) Analysis of the UL97 phosphotransferase coding sequence in clinical cytomegalovirus isolates and identification of mutations conferring ganciclovir resistance. The Journal of Infectious Diseases 171(3), 576-83. Chou, S., Lurain, N.S., Thompson, K.D., Miner, R.C., Drew, W.L. (2003) Viral DNA polymerase mutations associated with drug resistance in human cytomegalovirus. The Journal of Infectious Diseases 188(1), 32-9. Chou, S., Marousek, G., Guentzel, S., Follansbee, S.E., Poscher, M.E., Lalezari, J.P., Miner, R.C., Drew, W.L. (1997) Evolution of mutations conferring multidrug resistance during prophylaxis and therapy for cytomegalovirus disease. Journal of Infectious Diseases 176(3), 786-9.

124 Chou, S., Marousek, G., Parenti, D.M., Gordon, S.M., LaVoy, A.G., Ross, J.G., Miner, R.C., Drew, W.L. (1998) Mutation in region III of the DNA polymerase gene conferring foscarnet resistance in cytomegalovirus isolates from 3 subjects receiving prolonged antiviral therapy. The Journal of Infectious Diseases 178(2), 526-30. Chou, S., Marousek, G.I., Senters, A.E., Davis, M.G., Biron, K.K. (2004) Mutations in the human cytomegalovirus UL27 gene that confer resistance to maribavir. Journal of Virology 78(13), 7124-30. Chou, S., Marousek, G.I., Van Wechel, L.C., Li, S., Weinberg, A. (2007a) Growth and drug resistance phenotypes resulting from cytomegalovirus DNA polymerase region III mutations observed in clinical specimens. Antimicrobial Agents and Chemotherapy 51(11), 4160-2. Chou, S., Meichsner, C.L. (2000a) A nine-codon deletion mutation in the cytomegalovirus UL97 phosphotransferase gene confers resistance to ganciclovir. Antimicrobial Agents and Chemotherapy 44(1), 183-5. Chou, S., Miner, R.C., Drew, W.L. (2000) A deletion mutation in region V of the cytomegalovirus DNA polymerase sequence confers multidrug resistance. Journal of Infectious Diseases 182(6), 1765-8. Chou, S., Van Wechel, L.C., Lichy, H.M., Marousek, G.I. (2005) Phenotyping of cytomegalovirus drug resistance mutations by using recombinant viruses incorporating a reporter gene. Antimicrobial Agents and Chemotherapy 49(7), 2710-5. Chou, S., Waldemer, R.H., Senters, A.E., Michels, K.S., Kemble, G.W., Miner, R.C., Drew, W.L. (2002) Cytomegalovirus UL97 phosphotransferase mutations that affect susceptibility to ganciclovir. The Journal of Infectious Diseases 185(2), 162-9. Chou, S., Wechel, L.C., Marousek, G.I. (2007) Cytomegalovirus UL97 kinase mutations that confer maribavir resistance. Journal of Infectious Diseases 196(1), 91-4. Chou, S.W. (2001) Cytomegalovirus drug resistance and clinical implications. Transplant Infectious Disease 3(2), 20-4. Cihlar, T., Fuller, M.D., Cherrington, J.M. (1997) Expression of the catalytic subunit (UL54) and the accessory protein (UL44) of human cytomegalovirus DNA polymerase in a coupled in vitro transcription/translation system. Protein Expression and Purification 11(2), 209-18. Cihlar, T., Fuller, M.D., Cherrington, J.M. (1998) Characterization of drug resistance- associated mutations in the human cytomegalovirus DNA polymerase gene by using recombinant mutant viruses generated from overlapping DNA fragments. Journal of Virology 72(7), 5927-36. Cihlar, T., Fuller, M.D., Mulato, A.S., Cherrington, J.M. (1998a) A point mutation in the human cytomegalovirus DNA polymerase gene selected in vitro by cidofovir confers a slow replication phenotype in cell culture. Virology 248(2), 382-93. Cocohoba, J.M., McNicholl, I.R. (2002) Valganciclovir: an advance in cytomegalovirus therapeutics. The Annals of Pharmacotherapy 36(6), 1075-9. Colberg-Poley, A.M. (1996) Functional roles of immediate early proteins encoded by the human cytomegalovirus UL36-38, UL115-119, TRS1/IRS1 and US3 loci. Intervirology 39(5-6), 350-60. Colletti, K.S., Xu, Y., Cei, S.A., Tarrant, M., Pari, G.S. (2004) Human cytomegalovirus UL84 oligomerization and heterodimerization domains act as transdominant inhibitors of oriLyt-dependent DNA replication: evidence that IE2-UL84 and

125 UL84-UL84 interactions are required for lytic DNA replication. Journal of Virology 78(17), 9203-14. Colletti, K.S., Xu, Y., Yamboliev, I., Pari, G.S. (2005) Human cytomegalovirus UL84 is a phosphoprotein that exhibits UTPase activity and is a putative member of the DExD/H box family of proteins. The Journal of Biological Chemistry 280(12), 11955-60. Compton, T. (2004) Receptors and immune sensors: the complex entry path of human cytomegalovirus. Trends in Cell Biology 14(1), 5-8. Compton, T., Nowlin, D.M., Cooper, N.R. (1993) Initiation of human cytomegalovirus infection requires initial interaction with cell surface heparan sulfate. Virology 193(2), 834-41. Craig, J.M., Macauley, J.C., Weller, T.H., Wirth, P. (1957) Isolation of intranuclear inclusion producing agents from infants with illnesses resembling cytomegalic inclusion disease. Proceedings of the Society for Experimental Biology and Medicine 94(1), 4-12. Craighead, J.E., Kanich, R.E., Almeida, J.D. (1972) Nonviral microbodies with viral antigenicity produced in cytomegalovirus-infected cells. Journal of Virology 10(4), 766-75. Craxton, M. (1991) Linear amplification sequencing, a powerful method for sequencing DNA. Methods 3, 20-24. Crute, J.J., Lehman, I.R. (1991) Herpes simplex virus-1 helicase-primase: physical and catalytic properties. The Journal of Biological Chemistry 266, 4484-4488. Cushing, T.D., Adrian, J., Chen, X., Dimaio, H., Doughan, B., Flygare, J., Liang, L., Mayorga, V., Miao, S., Mellon, H., Peterson, M.G., Powers, J.P., Spector, F., Stein, C., Wright, M., Xu, D., Ye, Q., Jaen, J. (2006) Discovery of a novel series of inhibitors of human cytomegalovirus primase. Bioorganic and medicinal Chemistry Letters 16(18), 4879-83. Dahle, A.J., Fowler, K.B., Wright, J.D., Boppana, S.B., Britt, W.J., Pass, R.F. (2000) Longitudinal investigation of hearing disorders in children with congenital cytomegalovirus. Journal of the American Academy of Audiology 11(5), 283-90. Dal Monte, P., Pignatelli, S., Rossini, G., Landini, M.P. (2004) Genomic variants among human cytomegalovirus (HCMV) clinical isolates: the glycoprotein n (gN) paradigm. Human Immunology 65(5), 387-394. Dargan, D.J., Jamieson, F.E., MacLean, J., Dolan, A., Addison, C., McGeoch, D.J. (1997) The published DNA sequence of human cytomegalovirus strain AD169 lacks 929 base pairs affecting genes UL42 and UL43. Journal of Virology 71(12), 9833-6. Davis-Poynter, N.J., Lynch, D.M., Vally, H., Shellam, G.R., Rawlinson, W.D., Barrell, B.G., Farrell, H.E. (1997) Identification and characterization of a G protein- coupled receptor homolog encoded by murine cytomegalovirus. Journal of Virology 71(2), 1521-1529. Davison, A.J. (2002) Evolution of the herpesviruses. Veterinary Microbiology 86(1-2), 69-88. Davison, A.J., Dolan, A., Akter, P., Addison, C., Dargan, D.J., Alcendor, D.J., McGeoch, D.J., Hayward, G.S. (2003) The human cytomegalovirus genome revisited: comparison with the chimpanzee cytomegalovirus genome. The Journal of General Virology 84(1), 17-28. Davison, A.J., Stow, N.D. (2005) New genes from old: redeployment of dUTPase by herpesviruses. Journal of Virology 79(20), 12880-92.

126 de Bruyn Kops, A., Knipe, D. M. (1988) Formation of DNA replication structures in herpes virus-infected cells requires a viral DNA-binding protein. Cell 55, 857- 868. de Castro, E., Sigrist, C.J., Gattiker, A., Bulliard, V., Langendijk-Genevaux, P.S., Gasteiger, E., Bairoch, A., Hulo, N. (2006) ScanProsite: detection of PROSITE signature matches and ProRule-associated functional and structural residues in proteins. Nucleic Acid Research 34(Web server issue), W362-5. De Clercq, E. (2001) Antiviral drugs: current state of the art. Journal of Clinical Virology 22(1), 73-89. de Mello, A.L., Ferreira, E.C., Vilas Boas, L.S., Pannuti, C.S. (1996) Cytomegalovirus infection in a day-care center in the municipality of São Paulo. Revista do Instituto de Medicina Tropical de Sao Paulo 38(3), 165-9. de Ory, F., Ramírez, R., García Comas, L., León, P., Sagües, M.J., Sanz, J.C. (2004) Is there a change in cytomegalovirus seroepidemiology in Spain? European Journal of Epidemiology 19(1), 85-9. Digard, P., Bebrin, W., Weisshart, K., Coen, D.M. (1993) The extreme C-terminus of herpes simplex virus DNA polymerase is crucial for functional interaction with processivity factor UL42 and for viral replication. Journal of Virology 67, 398– 406. Dohner, K., Sodeik, B. (2005) The role of the cytoskeleton during viral infection. Current Topics in Microbiology and Immunology 285, 67-108. Dolan, A., Cunningham, C., Hector, R.D., Hassan-Walker, A.F., Lee, L., Addison, C., Dargan, D.J., McGeoch, D.J., Gatherer, D., Emery, V.C., Griffiths, P.D., Sinzger, C., McSharry, B.P., Wilkinson, G.W., Davison, A.J. (2004) Genetic content of wild-type human cytomegalovirus. Journal of General Virology 85(5), 1301-12. Dollard, S.C., Grosse, S.D., Ross, D.S. (2007) New estimates of the prevalence of neurological and sensory sequelae and mortality associated with congenital cytomegalovirus infection. Reviews in Medical Virology 17(5), 355-63. Dracheva, S., Koonin, E.V., and Crute, J.J. (1995) Identification of the primase active site of the herpes simplex virus type 1 helicase-primase. The Journal of Biological Chemistry 270(23), 14148-14153. Drew, W.L., Miner, R.C., Marousek, G.I., Chou, S. (2006) Maribavir sensitivity of cytomegalovirus isolates resistant to ganciclovir, cidofovir or foscarnet. Journal of Clinical Virology 37(2), 124-7. Ducancelle, A., Champier, G., Alain, S., Petit, F., Le Pors, M.J., Mazeron, M.C. (2006) A novel mutation in the UL54 gene of human cytomegalovirus isolates that confers resistance to foscarnet. Antiviral Therapy 11(4), 537-40. Dudas, K.C., And Ruyechan, W.T. (1998) Identification of a region of the Herpes Simplex Virus single-stranded DNA-binding protein involved in cooperative binding. Journal of Virology 72(1), 257–265. Dummer, J.S., Hardy, A., Poorsattar, A., Ho, M. (1983) Early infections in kidney, heart, and liver transplant recipients on cyclosporine. Transplantation 36(3), 259-67. Dunn, W., Chou, C., Li, H., Hai, R., Patterson, D., Stolc, V., Zhu, H., Liu, F. (2003) Functional profiling of a human cytomegalovirus genome. Proceedings of the National Academy of Sciences of the United States of America 100(24), 14223- 8. Dutch, R.E., Lehman, I.R. (1993) Renaturation of complementary DNA strands by herpes simplex virus type 1 ICP8. Journal of Virology 67(12), 6945–6949. Eckle, T., Lang, P., Prix, L., Jahn, G., Klingebiel, T., Handgretinger, R., Selle, B., Niethammer, D., Hamprecht, K. (2002) Rapid development of ganciclovir-

127 resistant cytomegalovirus infection in children after allogeneic stem cell transplantation in the early phase of immune cell recovery. Bone Marrow Transplantation 30(7), 433-9. Emery, V.C., Cope, A.V., Bowen, E.F., Gor, D., Griffiths, P.D. (1999) The dynamics of human cytomegalovirus replication in vivo. The journal of Experimental Medicine 190(2), 177-82. Emery, V.C., Griffiths, P.D. (2000) Prediction of cytomegalovirus load and resistance patterns after antiviral chemotherapy. Proceedings of the National Academy of Sciences of the United States of America 97(14), 8039-44. Ertl, P.F., and Powell, K.L. (1992) Physical and functional interaction of human cytomegalovirus DNA polymerase and its accessory protein (ICP36) expressed in insect cells. Journal of Virology 66(7), 4126-4133. Evers, D.L., Komazin, G., Ptak, R.G., Shin, D., Emmer, B.T., Townsend, L.B., Drach, J.C. (2004) Inhibition of human cytomegalovirus replication by benzimidazole nucleosides involves three distinct mechanisms. Antimicrobial Agents and Chemotherapy 48(10), 3918-27. Faizi-Khan, R., Mori, S., Eizuru, Y., Kumura-Ishii. K., Minamishima, Y. (1998) Genetic analysis of a ganciclovir-resistant human cytomegalovirus mutant. Antiviral Research 40(1-2), 95-103. Falkenberg, M., Elias, P., Lehman, I.R. (1998) The herpes simplex virus type 1 helicase-primase. The Journal of Biological Chemistry 273, 32154–32157. Feire, A.L., Koss, H., and Compton, T. (2004) Cellular integrins function as entry receptors for human cytomegalovirus via a highly conserved disintegrin-like domain. Proceedings of the National Academy of Sciences of the United States of America 101, 15470–15475. Felsenstein, J. (1989) PHYLIP -- Phylogeny Inference Package (Version 3.2). Cladistics 5, 164-166. Fernandez, E.J., Lolis, E. (2002) Structure, function, and inhibition of chemokines. Annual Review of Phamacology and Toxicology 42, 469-99. Fishman, J.A., Rubin, R.H. (1998) Infection in organ-transplant recipients. The New England Journal of Medicine 338(24), 1741-51. Frick, D.N., Richardson, C.C. (2001) DNA primases. Annual Review of Biochemistry 70, 39-80. Gandhi, M.K., Khanna, R. (2004) Human cytomegalovirus: clinical aspects, immune regulation, and emerging treatments. The Lancet Infectious Diseases 4(12), 725- 38. Gao, M., And Knipe, D.M. (1989) Genetic evidence for multiple nuclear functions of the Herpes Simplex Virus ICP8 DNA-binding protein. Journal of Virology 63(12), 5258-5267. Gao, M., Bouchey, J., Curtin, K., Knipe, D.M. (1988) Genetic identification of a portion of the herpes simplex virus ICP8 protein required for DNA-binding. Virology 163, 319–329. Gao, M., Knipe, D.M. (1991) Potential role for herpes simplex virus ICP8 DNA replication protein in stimulation of late gene expression. Journal of Virology 65, 2666–2675. Gao, M., Knipe, D.M. (1992) Distal protein sequences can affect the function of a nuclear localization signal. Molecular Cell Biology 12, 1330-9. Gao, Y., Colletti, K., Pari, G.S. (2007) Identification of Human Cytomegalovirus UL84 viral and cellular encoded binding partners using Proteomics Analysis. Journal of Virology To be published.

128 Gavel, Y., von Heijne, G. (1990) Sequence differences between glycosylated and non- glycosylated Asn-X-Thr/Ser acceptor sites: implications for protein engineering. Protein Engineering 3(5), 433-42. Gaytant, M.A., Steegers, E.A., Semmekrot, B.A., Merkus, H.M., Galama, J.M. (2002) Congenital cytomegalovirus infection: review of the epidemiology and outcome. Obstetrical and Gynecological Survey 57(4), 245-56. Geary, R.S., Henry, S.P., Grillone, L.R. (2002) Fomivirsen: clinical pharmacology and potential drug interactions. Clinical Pharmacokinetics 41(4), 255-60. Gebert, S., Schmolke, S., Sorg, G., Floss, S., Plachter, B., Stamminger, T. (1997) The UL84 protein of human cytomegalovirus acts as a transdominant inhibitor of immediate-early-mediated transactivation that is able to prevent viral replication. Journal of Virology 71(9), 7048-60. Gibson, W. (1996) Structure and assembly of the virion. Intervirology 39(5-6), 389-400. Gibson, W. (1999) Cytomegaloviruses (Herpesviridae) | Molecular Biology (Human). In: A. Granoff, Webster, R.G. (Ed), Encyclopedia of Virology, 2nd Edition ed., pp. 351-357. 3 vols. Academic Press. Gilbert, C., Boivin, G. (2005a) New reporter cell line to evaluate the sequential emergence of multiple human cytomegalovirus mutations during in vitro drug exposure. Antimicrobial Agents and Chemotherapy 49(12), 4860-6. Goldberg, D.E., Wang, H., Azen, S.P., Freeman, W.R. (2003) Long term visual outcome of patients with cytomegalovirus retinitis treated with highly active antiretroviral therapy. The British Journal of Ophthalmology 87(7), 853-5. Goldmacher, V.S., Bartle, L.M., Skaletskaya, A., Dionne, C.A., Kedersha, N.L., Vater, C.A., Han, J.W., Lutz, R.J., Watanabe, S., Cahir McFarland, E.D., Kieff, E.D., Mocarski, E.S., Chittenden, T. (1999) A cytomegalovirus-encoded mitochondria-localized inhibitor of apoptosis structurally unrelated to Bcl-2. Proceedings of the National Academy of Sciences of the United States of America 96(22), 12536-41. Gorbalenya, A.E., Koonin, E.V., Donchenko, A.P., Blinov, V.M. (1988) A conserved NTP-motif in putative helicases. Nature 333(6168), 22. Gorbalenya, A.E., Koonin, E.V., Donchenko, A.P., Blinov, V.M. (1989) Two related superfamilies of putative helicases involved in replication, recombination, repair and expression of DNA and RNA genomes. Nucleic Acid Research 17, 4713- 4730. Gottlieb, J., and Challberg, M.D. (1994) Interaction of herpes simplex virus type 1 DNA polymerase and the UL42 accessory protein with a model primer template. Journal of Virology 68(8), 4937-4945. Gottlieb, J., Marcy, A.I., Coen, D.M., Challberg, M.D. (1990) The herpes simplex virus type 1 UL42 gene product: a subunit of DNA polymerase that functions to increase processivity. Journal of Virology 64(12), 5976-87. Grattan, M.T., Moreno-Cabral, C.E., Starnes, V.A., Oyer, P.E., Stinson, E.B., Shumway, N.E. (1989) Cytomegalovirus infection is associated with cardiac allograft rejection and atherosclerosis. JAMA: The Journal of the American Medical Association 261(24), 3561-6. Graves-Woodward, K.L., and Weller, S.K. (1996) Replacement of gly815 in helicase motif V alters the single-stranded DNA-dependent ATPase activity of the herpes simplex virus type 1 helicase-primase. The Journal of Biological Chemistry 271(23), 13629–13635. Graves-Woodward, K.L., Gottlieb, J., Challberg, M.D., and Weller, S.K. (1997) Biochemical analyses of mutations in the HSV-1 helicase-primase that alter ATP

129 hydrolysis, DNA unwinding, and coupling between hydrolysis and unwinding. The Journal of Biological Chemistry 272(7), 4623–4630. Group., V.S. (2002) Safety of intravitreous fomivirsen for treatment of cytomegalovirus retinitis in patients with AIDS. American Journal of Ophthalmology 133(4), 484–98. Grunewald, K., Desai, P., Winkler, D.C., Heymann, J.B., Belnap, D.M., Baumeister, W., Steven, A.C. (2003) Three-dimensional structure of herpes simplex virus from cryo-electron tomography. Science 302(5649), 1396-8. Gupte, S.S., Olson, J.W., and Ruyechan, W.T. (1991) The major herpes simplex virus type-1 DNA-binding protein is a zinc metalloprotein. The Journal of Biological Chemistry 266(18), 11413-11416. Hahn, G., Khan, H., Baldanti, F., Koszinowski, U.H., Revello, M.G., Gerna, G. (2002) The human cytomegalovirus ribonucleotide reductase homolog UL45 is dispensable for growth in endothelial cells, as determined by a BAC-cloned clinical isolate of human cytomegalovirus with preserved wild-type characteristics. Journal of Virology 76(18), 9551-5. Hahn, G., Revello, M.G., Patrone, M., Percivalle, E., Campanini, G., Sarasini, A., Wagner, M., Gallina, A., Milanesi, G., Koszinowski, U., Baldanti, F., Gerna, G. (2004) Human cytomegalovirus UL131-128 genes are indispensable for virus growth in endothelial cells and virus transfer to leukocytes. Journal of Virology 78(18), 10023-33. Hakki, M., Geballe, A.P. (2005) Double-Stranded RNA Binding by Human Cytomegalovirus pTRS1. Journal of Virology 79(12), 7311-8. Hamatake, R.K., Bifano, M., Hurlburt, W.W., Tenney, D.J. (1997) A functional interaction of ICP8, the herpes simplex virus single-stranded DNA-binding protein, and the helicase-primase complex that is dependent on the presence of the UL8 subunit. Journal of General Virology 78, 857–865. Handsfield, H.H., Chandler, S.H., Caine, V.A., Meyers, J.D., Corey, L., Medeiros, E., McDougall, J.K. (1985) Cytomegalovirus infection in sex partners: evidence for sexual transmission. The Journal of Infectious Diseases 151(2), 344-8. Hansen, S.G., Strelow, L.I., Franchi, D.C., Anders, D.G., Wong, S.W. (2003) Complete sequence and genomic analysis of rhesus cytomegalovirus. Journal of Virology 77(12), 6620-36. Hanson, M.N., Preheim, L.C., Chou, S., Talarico, C.L., Biron, K.K., Erice, A. (1995) Novel mutation in the UL97 gene of a clinical cytomegalovirus strain conferring resistance to ganciclovir. Antimicrobial Agents and Chemotherapy 39(5), 1204- 5. Hantz, S., Michel, D., Fillet, A.M., Guigonis, V., Champier, G., Mazeron, M.C., Bensman, A., Denis, F., Mertens, T., Dehee, A., Alain, S. (2005) Early selection of a new UL97 mutant with a severe defect of ganciclovir phosphorylation after valaciclovir prophylaxis and short-term ganciclovir therapy in a renal transplant recipient. Antimicrobial Agents and Chemotherapy 49(4), 1580-3. Hassan-Walker, A.F., Okwuadi, S., Lee, L., Griffiths, P.D., Emery, V.C. (2004) Sequence variability of the alpha-chemokine UL146 from clinical strains of human cytomegalovirus. Journal of Medical Virology 74(4), 573-579. Hassan, J., Connell, J. (2007) Translational mini-review series on infectious disease: congenital cytomegalovirus infection: 50 years on. Clinical and Experimental Immunology 149(2), 205-10. Hayajneh, W.A., Colberg-Poley, A.M., Skaletskaya, A., Bartle, L.M., Lesperance, M.M., Contopoulos-Ioannidis, D.G., Kedersha, N.L., Goldmacher, V.S. (2001)

130 The sequence and antiapoptotic functional domains of the human cytomegalovirus UL37 exon 1 immediate early protein are conserved in multiple primary strains. Virology 279(1), 233-40. Haynes, R.O. (2002) Integrins: bidirectional, allosteric signaling machines. Cell 110(6), 673-87. He, X., Lehman, I.R. (2001) An initial ATP-independent step in the unwinding of a herpes simplex virus type 1 origin of replication by a complex of the viral origin-binding protein and single-strand DNAbinding protein. Proceedings of the National Academy of Sciences of the United States of America 98, 3024– 3028. He, Y.S., Xu, L., Huang, E.S. (1992) Characterization of human cytomegalovirus UL84 early gene and identification of its putative protein product. Journal of Virology 66(2), 1098-108. Hecker, M., Qiu, D., Marquardt, K., Bein, G., Hackstein, H. (2004) Continuous cytomegalovirus seroconversion in a large group of healthy blood donors. Vox Sanguinis 86(1), 41-4. Heilbronn, R., Jahn, G., Bürkle, A., Freese, U.K., Fleckenstein, B., zur Hausen, H. (1987) Genomic localization, sequence analysis, and transcription of the putative human cytomegalovirus DNA polymerase gene. Journal of Virology 61(1), 119- 24. Hernandez, T.R., Lehman, I.R. (1990) Functional interaction between the herpes simplex-1 DNA polymerase and UL42 protein. The Journal of Biological Chemistry 265, 11227–11232. Heymann, J.B., Cheng, N., Newcomb, W.W., Trus, B.L., Brown, J.C., Steven, A.C. (2003) Dynamics of herpes simplex virus capsid maturation visualized by time- lapse cryo-electron microscopy. Nature Structural Biology 10(5), 334-41. Hitomi, S., Kozuka-Hata, H., Chen, Z., Sugano, S., Yamaguchi, N., Watanabe, S. (1997) Human cytomegalovirus open reading frame UL11 encodes a highly polymorphic protein expressed on the infected cell surface. Archives of Virology 142(7), 1407-27. Hizel, S., Parker, S., Onde, U. (1999) Seroprevalence of cytomegalovirus infection among children and females in Ankara, Turkey, 1995. Pediatrics International 41(5), 506-9. Hjelmesaeth, J., Sagedal, S., Hartmann, A., Rollag, H., Egeland, T., Hagen, M., Nordal, K.P., Jenssen, T. (2004) Asymptomatic cytomegalovirus infection is associated with increased risk of new-onset diabetes mellitus and impaired insulin release after renal transplantation. Diabetologia 47(9), 1550-6. Ho, E.S., Lin, D.C., Mendel, D.B., Cihlar, T. (2000) Cytotoxicity of antiviral nucleotides adefovir and cidofovir is induced by the expression of human renal organic anion transporter 1. Journal of the American Society of Nephrology 11(3), 383-93. Homa, F.L., Brown, J.C. (1997) Capsid assembly and DNA packaging in herpes simplex virus. Reviews in Medical Virology 7(2), 107-122. Hui, E.K., Ralston, K., Judd, A.K., Nayak, D.P. (2003) Conserved cysteine and histidine residues in the putative zinc finger motif of the influenza A virus M1 protein are not critical for influenza virus replication. Journal of General Virology 84(11), 3105-13. Hulo, N., Bairoch, A., Bulliard, V., Cerutti, L., De Castro, E., Langendijk-Genevaux, P.S., Pagni, M., Sigrist, C.J. (2006) The PROSITE database. Nucleic Acid Research 34, D227-30.

131 Hurst, R., Maffitt, M., Murray, E., Kappelman, J., Xu, Q., Mendez, J., Butler, B., Ryan A., and Beckler, G.S. (1996) The TNT® T7 Quick Coupled Transcription/Translation System. Promega Notes 58, 8-11. Hwang, C.B., Ruffner, K.L., Coen, D.M. (1992) A point mutation within a distinct conserved region of the herpes simplex virus DNA polymerase gene confers drug resistance. Journal of Virology 66(3), 1774-6. Ijichi, O., Michel, D., Mertens, T., Miyata, K., Eizuru, Y. (2002) GCV resistance due to the mutation A594P in the cytomegalovirus protein UL97 is partially reconstituted by a second mutation at D605E. Antiviral Research 53(2), 135-42. Ilyina, T.V., Gorbalenya, A.E., Koonin, E.V. (1992) Organization and evolution of bacterial and bacteriophage primase-helicase systems. Journal of Molecular Evolution 34(4), 351–357. Iskenderian, A.C., Huang, L., Reilly, A., Stenberg, R.M., Anders, D.G. (1996) Four of eleven loci required for transient complementation of human cytomegalovirus DNA replication cooperate to activate expression of replication genes. Journal of Virology 70(1), 383-92. Jakob, U., Muse, W., Eser, M., Bardwell, J.C. (1999) Chaperone activity with a redox switch. Cell 96(3), 341-52. Jans, D.A., Hübner, S. (1996) Regulation of protein transport to the nucleus: central role of phosphorylation. Physiological Reviews 76(3), 651-85. Jarvis, M.A., Jones, T.R., Drummond, D.D., Smith, P.P., Britt, W.J., Nelson, J.A., Baldick, C.J. (2004) Phosphorylation of human cytomegalovirus glycoprotein B (gB) at the acidic cluster casein kinase 2 site (Ser900) is required for localization of gB to the trans-Golgi network and efficient virus replication. Journal of Virology. Jarvis, M.A., Nelson, J.A. (2002) Human cytomegalovirus persistence and latency in endothelial cells and macrophages. Current Opinion in Microbiology 5(4), 403-7. Jasanoff, A., Weiss, M.A. (1993) Aromatic-histidine interactions in the zinc finger motif: structural inequivalence of phenylalanine and tyrosine in the hydrophobic core. Biochemistry 32(6), 1423-32. Jiang, X.J., Adler, B., Sampaio, K.L., Digel, M., Jahn, G., Ettischer, N., Stierhof, Y.D., Scrivano, L., Koszinowski, U., Mach, M., Sinzger, C. (2008) UL74 of human cytomegalovirus contributes to virus release by promoting secondary envelopment of virions. Journal of Virology 82(6), 2802-12. Kääriäinen, L., Klemola, E., Paloheimo, J. (1966) Rise of cytomegalovirus antibodies in an infectious-mononucleosis-like syndrome after transfusion. British Medical Journal 1(5498), 1270-2. Kangro, H.O., Osman, H.K., Lau, Y.L., Heath, R.B., Yeung, C.Y., Ng, M.H. (1994) Seroprevalence of antibodies to human herpesviruses in England and Hong Kong. Journal of Medical Virology 43(1), 91-6. Kiehl, A., Huang, L., Franchi, D., and Anders, D.G. (2003) Multiple 5' ends of human cytomegalovirus UL57 transcripts identify a complex, cycloheximide-resistant promoter region that activates oriLyt. Virology 314, 410–422. Kiss, P., De Bacquer, D., Sergooris, L., De Meester, M., VanHoorne, M. (2002) Cytomegalovirus infection: an occupational hazard to kindergarten teachers working with children aged 2.5-6 years. International Journal of Occupational and Environmental Health 8(2), 79-86. Klinedinst, D.K., and Challberg, M.D. (1994) Helicase-primase complex of herpes simplex virus type 1 a mutation in the UL52 subunit abolishes primase activity. Journal of Virology 68(6), 3693-3701.

132 Knowles, S.J., Grundy, K., Cahill, I., Cafferkey, M.T., Geary, M. (2005) Low cytomegalovirus sero-prevalence in Irish pregnant women. Irish Medical Journal 98(7), 210-2. Kochoyan, M., Keutmann, H.T., Weiss, M.A. (1991) Architectural rules of the zinc- finger motif: comparative two-dimensional NMR studies of native and "aromatic-swap" domains define a "weakly polar switch". Proceedings of the National Academy of Sciences of the United States of America 88(19), 8455-9. Komazin, G., Ptak, R.G., Emmer, B.T., Townsend, L.B., Drach, J.C. (2003) Resistance of human cytomegalovirus to the benzimidazole L-ribonucleoside maribavir maps to UL27. Journal of Virology 77(21), 11499-506. Komazin, G., Townsend, L.B., Drach, J.C. (2004) Role of a mutation in human cytomegalovirus gene UL104 in resistance to benzimidazole ribonucleosides. Journal of Virology 78(2), 710-5. Kothari, A., Ramachandran, V.G., Gupta, P., Singh, B., Talwar, V. (2002) Seroprevalence of cytomegalovirus among voluntary blood donors in Delhi, India. Journal of Health, Population, and Nutrition 20(4), 348-51. Krishnaswamy, S., Rossmann, M.G. (1990) Structural refinement and analysis of Mengo virus. Journal of Molecular Biology 211(4), 803-44. Krosky, P.M., Baek, M-C., Jahng, W.J., Barrera, I., Harvey, R.J., Biron, K.K., Coen, D.M., and Sethna P.B. (2003) The human cytomegalovirus UL44 protein Is a substrate for the UL97 protein kinase. Journal of Virology 77(14), 7720–7727. Krosky, P.M., Baek, M.C., Coen, D.M. (2003a) The human cytomegalovirus UL97 protein kinase, an antiviral drug target, is required at the stage of nuclear egress. Journal of Virology 77(2), 905-14. Krosky, P.M., Underwood, M.R., Turk, S.R., Feng, K.W., Jain, R.K., Ptak, R.G., Westerman, A.C., Biron, K.K., Townsend, L.B., Drach, J.C. (1998) Resistance of human cytomegalovirus to benzimidazole ribonucleosides maps to two open reading frames: UL89 and UL56. Journal of Virology 72(6), 4721-8. Kubo, T., Rai, S.K., Nakanishi, M., Yamano, T. (1991) Seroepidemiological study of herpes viruses in Nepal. The Southeast Asian Journal of Tropical Medicine and Public Health 22(3), 323-5. Lang, D.J., Kummer, J.F. (1975) Cytomegalovirus in semen: observations in selected populations. The Journal of Infectious Diseases 132(4), 472-3. Lee, S.S.-K., Lehman, I.R. (1997) Unwinding of the box I element of a herpes simplex virus type 1 origin by a complex of the viral origin binding protein, single-strand DNA binding protein, and singlestranded DNA. Proceedings of the National Academy of Sciences of the United States of America 94, 2838–2842. Lemström, K., Koskinen, P., Krogerus, L., Daemen, M., Bruggeman, C., Häyry, P. (1995) Cytomegalovirus antigen expression, endothelial cell proliferation, and intimal thickening in rat cardiac allografts after cytomegalovirus infection. Circulation 92(9), 2594-604. Limaye, A.P., Corey, L., Koelle, D.M., Davis, C.L., Boeckh, M. (2000) Emergence of ganciclovir-resistant cytomegalovirus disease among recipients of solid-organ transplants. Lancet 356(9230), 645-9. Ling, P.D., Lednicky, J.A., Keitel, W.A., Poston, D.G., White, Z.S., Peng, R., Liu, Z., Mehta, S.K., Pierson, D.L., Rooney, C.M., Vilchez, R.A., Smith, E.O., Butel, J.S. (2003) The dynamics of herpesvirus and polyomavirus reactivation and shedding in healthy adults: a 14-month longitudinal study. The Journal of Infectious Diseases 187(10), 1571-80.

133 Lischka, P., Rauh, C., Mueller, R., Stamminger, T. (2006) Human cytomegalovirus UL84 protein contains two nuclear export signals and shuttles between the nucleus and the cytoplasm. Journal of Virology 80(20), 10274-80. Lischka, P., Sorg, G., Kann, M., Winkler, M., Stamminger, T. (2003) A nonconventional nuclear localization signal within the UL84 protein of human cytomegalovirus mediates nuclear import via the importin alpha/beta pathway. Journal of Virology 77(6), 3734-48. Liu, Z., Wang, E., Taylor, W., Yu, H., Wu, T., Wan, Z., Huang, Y., Ni, Z., Sackett, D. (1990) Prevalence survey of cytomegalovirus infection in children in Chengdu. American Journal of Epidemiology 131(1), 143-50. Loh, L.C., Keeler, V.D., Shanley, J.D. (1999) Sequence requirements for the nuclear localization of the murine cytomegalovirus M44 gene product pp50. Virology 259, 43–59. Loh, L.C., Locke, D., Melnychuk, R., and Laferte, S. (2000) The RGD sequence in the cytomegalovirus DNA polymerase accessory protein can mediate cell adhesion. Virology 272, 302-314. Lopper, M., Compton, T. (2004) Coiled-coil domains in glycoproteins B and H are involved in human cytomegalovirus membrane fusion. Journal of Virology 78(15), 8333-41. Loregian, A., Appleton, B.A., Hogle, J.M., and Coen, D.M. (2004) Specific residues in the connector loop of the human cytomegalovirus DNA polymerase accessory protein UL44 are crucial for interaction with the UL54 catalytic subunit. Journal of Virology 78(17), 9084–9092. Loregian, A., Appleton, B.A., Hogle, J.M., and Coen, D.M. (2004a) Residues of human cytomegalovirus DNA polymerase catalytic subunit UL54 that are necessary and sufficient for interaction with the accessory protein UL44. Journal of Virology 78(1), 158-167. Loregian, A., Palu, G. (2005) Disruption of the interactions between the subunits of herpesvirus DNA polymerases as a novel antiviral strategy. Clinical Microbiology and Infection 11(6), 437-46. Loregian, A., Rigatti, R., Murphy, M., Schievano, E., Palu, G., Marsden, H.S. (2003) Inhibition of human cytomegalovirus DNA polymerase by C-terminal peptides from the UL54 subunit. Journal of Virology 77(15), 8336-44. Lorz, K., Hofmann, H., Berndt, A., Tavalai, N., Mueller, R., Schlotzer-Schrehardt, U., Stamminger, T. (2006) Deletion of open reading frame UL26 from the human cytomegalovirus genome results in reduced viral growth, which involves impaired stability of viral particles. Journal of Virology 80(11), 5423-34. Lukac, D.M., Alwine, J.C. (1999) Effects of human cytomegalovirus major immediate- early proteins in controlling the cell cycle and inhibiting apoptosis: studies with ts13 cells. Journal of Virology 73(4), 2825-31. Lurain, N.S., Kapell, K.S., Huang, D.D., Short, J.A., Paintsil, J., Winkfield, E., Benedict, C.A., Ware., C.F., Bremer, J.W. (1999) Human cytomegalovirus UL144 open reading frame: sequence hypervariability in low-passage clinical isolates. Journal of Virology 73(12), 10040-10050. Lurain, N.S., Spafford, L.E., Thompson, K.D. (1994) Mutation in the UL97 open reading frame of human cytomegalovirus strains resistant to ganciclovir. Journal of Virology 68(7), 4427-31. Lurain, N.S., Thompson, K.D., Holmes, E.W., Read, G.S. (1992) Point mutations in the DNA polymerase gene of human cytomegalovirus that result in resistance to antiviral agents. Journal of Virology 66(12), 7146-52.

134 Mach, M., Kropff, B., Kryzaniak, M., and Britt, W. (2005) Complex formation by glycoproteins M and N of human cytomegalovirus: structural and functional aspects. Journal of Virology 79, 2160–2170. Makhov, A.M., Boehmer, P.E., Lehman, I.R., Griffith, J.D. (1996) Visualization of the unwinding of long DNA chains by the herpes simplex virus type 1 UL9 protein and ICP8. Journal of Molecular Biology 258, 789–799. Malm, G., Engman, M.L. (2007) Congenital cytomegalovirus infections. Seminars in Fetal and Neonatal Medicine 12(3), 154-9. Mansat, A., Mengelle, C., Chalet, M., Boumzebra, A., Mieusset, R., Puel, J., Prouheze, C., Segondy, M. (1997) Cytomegalovirus detection in cryopreserved semen samples collected for therapeutic donor insemination. Human Reproduction 12(8), 1663-6. Mapelli, M., Muhleisen, M., Persico, G., Van Der Zandt, H., and Tucker, P.A. (2000) The 60-residue C-terminal region of the single-stranded DNA binding protein of Herpes Simplex Virus type 1 Is required for cooperative DNA binding. Journal of Virology 74(19), 8812–8822. Mar, E.C., Patel, P.C., Huang, E.S. (1981) Human cytomegalovirusassociated DNA polymerase and protein kinase activities. Journal of General Virology 57, 149– 156. Marfori, J.E., Exner, M.M., Marousek, G.I., Chou, S., Drew, W.L. (2007) Development of new cytomegalovirus UL97 and DNA polymerase mutations conferring drug resistance after valganciclovir therapy in allogeneic stem cell recipients. Journal of Clinical Virology 38(2), 120-5. Marschall, M., Freitag, M., Suchy, P., Romaker, D., Kupfer, R., Hanke, M., Stamminger, T. (2003) The protein kinase pUL97 of human cytomegalovirus interacts with and phosphorylates the DNA polymerase processivity factor pUL44. Virology 311(1), 60-71. Marsden, H.S., McLean, G.W., Barnard, E.C., Francis, G.J., MacEachran, K., Murphy, M., McVey, G., Cross, A., Abbotts, A.P., Stow, N.D. (1997) The catalytic subunit of the DNA polymerase of herpes simplex virus type 1 interacts specifically with the C terminus of the UL8 component of the viral helicase- primase complex. Journal of Virology 71(9), 6390-7. Martin, M., Gilbert, C., Covington, E., Boivin, G. (2006) Characterization of human cytomegalovirus (HCMV) UL97 mutations found in a valganciclovir/oral ganciclovir prophylactic trial by use of a bacterial artificial chromosome containing the HCMV genome. Journal of Infectious Diseases 194(5), 579-83. McGeoch, D.J., Cook, S., Dolan, A., Jamieson, F.E., Telford, E.A.R. (1995) Molecular Phylogeny and Evolutionary Timescale for the Family of Mammalian Herpesviruses. Journal of Molecular Biology 247, 443–458. McGeoch, D.J., Dalrymple, M.A., Dolan, A., McNab, D., Perry, L.J., Taylor, P., Challberg, M.D. (1988) Structures of herpes simplex virus type 1 genes required for replication of virus DNA. Journal of Virology 62(2), 444-53. McGeoch, D.J., Gatherer, D. (2005) Integrating reptilian herpesviruses into the family herpesviridae. Journal of Virology 79(2), 725-31. McGowan, M.P., Hayes, K., Kovacs, G.T., Leydon, J.A. (1983) Prevalence of cytomegalovirus and herpes simplex virus in human semen. International Journal of Andrology 6(4), 331-6. McLean, G.W., Abbotts, A.P., Parry, M.E., Marsden, H.S., Stow, N.D. (1994) The herpes simplex virus type 1 origin-binding protein interacts specifically with the viral UL8 protein. The Journal of General Virology 75(10), 2699-706.

135 McMahon, T.P., and Anders, D.G. (2002) Interactions between Human Cytomegalovirus Helicase-Primase Proteins. Virus Research 86, 39–52. McSharry, J.J., McDonough, A., Olson, B., Hallenberger, S., Reefschlaeger, J., Bender, W., Drusano, G.L. (2001) Susceptibilities of human cytomegalovirus clinical isolates to BAY38-4766, BAY43-9695, and ganciclovir. Antimicrobial Agents and Chemotherapy 45(10), 2925-7. Meggio, F., Pinna, L.A. (2003) One-thousand-and-one substrates of protein kinase CK2? FASEB 17(3), 349-68. Mehta, S.K., Stowe, R.P., Feiveson, A.H., Tyring, S.K., Pierson, D.L. (2000) Reactivation and shedding of cytomegalovirus in astronauts during spaceflight. The Journal of Infectious Diseases 182(6), 1761-4. Mendelman, L.V., Beauchamp, B.B., Richardson, C.C. (1994) Requirement for a zinc motif for template recognition by the bacteriophage T7 primase. The EMBO Journal 13(16), 3909-16. Mercorelli, B., Sinigalia, E., Loregian, A., Palù, G. (2007) Human cytomegalovirus DNA replication: antiviral targets and drugs. Reviews in Medical Virology In Print. Mettenleiter, T.C. (2004) Budding events in herpesvirus morphogenesis. Virus Research 106(2), 167-80. Meyer-König, U., Haberland, M., von Laer, D., Haller, O., Hufert, F.T. (1998) Intragenic variability of human cytomegalovirus glycoprotein B in clinical strains. The Journal of Infectious Diseases 177(5), 1162-9. Michael, S.F., Kilfoil, V.J., Schmidt, M.H., Amann, B.T., Berg, J.M. (1992) Metal binding and folding properties of a minimalist Cys2His2 zinc finger peptide. Proceedings of the National Academy of Sciences of the United States of America 89(11), 4796-800. Michel, D., Milotić, I., Wagner, M., Vaida, B., Holl, J., Ansorge, R., Mertens, T. (2005) The human cytomegalovirus UL78 gene is highly conserved among clinical isolates, but is dispensable for replication in fibroblasts and a renal artery organ- culture system. Journal of General Virology 86(2), 297-306. Miles, D.J., van der Sande, M., Jeffries, D., Kaye, S., Ismaili, J., Ojuola, O., Sanneh, M., Touray, E.S., Waight, P., Rowland-Jones, S., Whittle, H., Marchant, A. (2007) Cytomegalovirus infection in Gambian infants leads to profound CD8 T-cell differentiation. Journal of Virology 81(11), 5766-76. Miller, W.J., McCullough, J., Balfour, H.H. Jr, Haake, R.J., Ramsay, N.K., Goldman, A., Bowman, R., Kersey, J. (1991) Prevention of cytomegalovirus infection following bone marrow transplantation: a randomized trial of blood product screening. Bone Marrow Transplantation 7(3), 227-34. Mocarski, E.S., Prichard, M.N., Tan, C.S., Brown, J.M. (1997) Reassessing the organisation of the UL42-UL43 region of the human cytomegalovirus strain AD169 genome. Virology 239(1), 169-175. Mocarski, E.S., Tan Courcelle, C. (2001) Cytomegalovirus and Their Replication. In: D.M. Knipe, Howley, P.M., (Ed), Field's Virology, 4th ed., pp. 2629-2673. Vol. 2. Lippencott Williams and Wilkins, Philadelphia. Monahan, S.J., Barlam, T.F., Crumpacker, C.S., Parris, D.S. (1993) Two regions of the herpes simplex virus type 1 UL42 protein are required for its functional interaction with viral DNA polymerase. Journal of Virology 67, 5922–5931. Mousavi-Jazi, M., Schloss, L., Drew, W.L., Linde, A., Miner, R.C., Harmenberg, J., Wahren, B., Brytting, M. (2001) Variations in the cytomegalovirus DNA

136 polymerase and phosphotransferase genes in relation to foscarnet and ganciclovir sensitivity. Journal of Clinical Virology 23(1-2), 1-15. Mousavi-Jazi, M., Schloss, L., Wahren, B., Brytting, M. (2003) Point mutations induced by foscarnet (PFA) in the human cytomegalovirus DNA polymerase. Journal of Clinical Virology 26(3), 301-6. Munro, S.C., Hall, B., Whybin, L.R., Leader, L., Robertson, P., Maine, G.T., Rawlinson, W.D. (2005) Diagnosis of and screening for cytomegalovirus infection in pregnant women. Journal of Clinical Microbiology 43(9), 4713-8. Murphy, E., Yu, D., Grimwood, J., Schmutz, J., Dickson, M., Jarvis, M.A., Hahn, G., Nelson, J.A., Myers, R.M., Shenk, T.E. (2003) Coding potential of laboratory and clinical strains of human cytomegalovirus. Proceedings of the National Academy of Sciences of the United States of America 100(25), 14976-81. Natali, A., Valcavi, P., Medici, M.C., Dieci, E., Montali, S., Chezzi, C. (1997) Cytomegalovirus infection in an Italian population: antibody prevalence, virus excretion and maternal transmission. The New Microbiologica 20(2), 123-33. Navarro, D., Paz, P., Tugizov, S., Topp, K., La, Vail. J, Pereira, L. (1993) Glycoprotein B of human cytomegalovirus promotes virion penetration into cells, transmission of infection from cell to cell, and fusion of infected cells. Virology 197(1), 143-58. Neote, K., DiGregorio, D., Mak, J.Y., Horuk, R., Schall, T.J. (1993) Molecular cloning, functional expression, and signaling characteristics of a C-C chemokine receptor. Cell 72(3), 415-25. Nguyen-Huynh, A.T., Schaffer, P.A. (1998) Cellular transcription factors enhance herpes simplex virus type 1 oriS-dependent DNA replication. Journal of Virology 72(5), 3635–3645. Nimonkar, A.V., Boehmer, P.E. (2003) Reconstitution of recombination-dependent DNA synthesis in herpes simplex virus 1. Proceedings of the National Academy of Sciences of the United States of America 100(18), 10201-6. Nishiyama, Y., Maeno, K., Yoshida, S. (1983) Characterization of human cytomegalovirus-induced DNA polymerase and the associated 3'-to-5' exonuclease. Virology 124, 221–231. Nogalski, M.T., Podduturi, J.P., DeMeritt, I.B., Milford, L.E., Yurochko, A.D. (2007) The human cytomegalovirus virion possesses an activated casein kinase II that allows for the rapid phosphorylation of the inhibitor of NF-kappaB, IkappaBalpha. Journal of Virology 81(10), 5305-14. Ogawa-Goto, K., Tanaka, K., Gibson, W., Moriishi, E., Miura, Y., Kurata, T., Irie, S., Sata, T. (2003) Microtubule network facilitates nuclear targeting of human cytomegalovirus capsid. Journal of Virology 77(15), 8541-7. Pari, G.S., and Anders, D.G. (1993) Eleven loci encoding trans-acting factors are required for transient complementation of human cytomegalovirus oriLyt- dependent DNA replication. Journal of Virology 67(12), 6979-6988. Pari, G.S., Kacica, M.A., and Anders, D.A. (1993a) Open reading frames UL44, IRS1/TRS1, and UL36-38 are required for transient complementation of human cytomegalovirus oriLyt-dependent DNA synthesis. Journal of Virology 67, 2575–2582. Park, M.Y., Kim, Y.E., Seo, M.R., Lee, J.R., Lee, C.H., Ahn, J.H. (2006) Interactions among four proteins encoded by the human cytomegalovirus UL112-113 region regulate their intranuclear targeting and the recruitment of UL44 to prereplication foci. Journal of Virology 80(6), 2718-27.

137 Pass, R.F. (2005) Congenital cytomegalovirus infection and hearing loss. Herpes: The Journal of the IHMF 12(2), 50-5. Pass, R.F., Hutto, C., Lyon, M.D., Cloud, G. (1990) Increased rate of cytomegalovirus infection among day care center workers. The Pediatric Infectious Disease Journal 9(7), 465-70. Pass, R.F., Hutto, S.C., Reynolds, D.W., Polhill, R.B. (1984) Increased frequency of cytomegalovirus infection in children in group day care. Pediatrics 74(1), 121-6. Pescovitz, M.D., Rabkin, J., Merion, R.M., Paya, C.V., Pirsch, J., Freeman, R.B., O'Grady, J., Robinson, C., To, Z., Wren, K., Banken, L., Buhles, W., Brown, F. (2000) Valganciclovir results in improved oral absorption of ganciclovir in liver transplant recipients. Antimicrobial Agents and Chemotherapy 44(10), 2811-5. Pignatelli, S., Dal Monte, P., Landini, M.P. (2001) gpUL73 (gN) genomic variants of human cytomegalovirus isolates are clustered into four distinct genotypes. Journal of General Virology 82(11), 2777-84. Pilger, B.D., Cui, C., and Coen, D.M. (2004) Identification of a small molecule that inhibits herpes simplex virus DNA polymerase subunit interactions and viral replication. Chemical Biology 11, 647–654. Pinna, L.A. (1990) Casein kinase 2: an 'eminence grise' in cellular regulation? Biochemica et Biophysica atca 1054(3), 267-84. Pouteil-Noble, C., Ecochard, R., Landrivon, G., Donia-Maged, A., Tardy, J.C., Bosshard, S., Colon, S., Betuel, H., Aymard, M., Touraine, J.L. (1993) Cytomegalovirus infection--an etiological factor for rejection? A prospective study in 242 renal transplant patients. Transplantation 55(4), 851-7. Prabhakar, P., Bailey, A., Smikle, M.F., Ashley, D. (1992) Seroprevalence of cytomegalovirus infection in a selected population in Jamaica. The West Indian Medical Journal 41(4), 133-5. Prichard, M.N., Jairath, S., Penfold, M.E., St Jeor, S., Bohlman, M.C., Pari, G.S. (1998) Identification of persistent RNA-DNA hybrid structures within the origin of replication of human cytomegalovirus. Journal of Virology 72(9), 6997-7004. Prichard, M.N., Penfold, M.E., Duke, G.M., Spaete, R.R., Kemble, G.W. (2001) A review of genetic differences between limited and extensively passaged human cytomegalovirus strains. Reviews in Medical Virology 11(3), 191-200. Prior, J.R., Morroll, D.R., Birks, A.G., Matson P.L. and Lieberman, B.A. (1994) The screening for cytomegalovirus antibody in semen donors and recipients within a donor insemination program. Human Reproduction 9(11), 2076-2078. Pultoo, A., Meetoo, G., Pyndiah, M.N., Khittoo, G. (2001) Seroprevalence of cytomegalovirus infection in Mauritian volunteer blood donors. Indian Journal of Medical Sciences 55(2), 73-8. Ramirez-Aguilar, K.A., Kuchta, R.D. (2004) Mechanism of primer synthesis by the herpes simplex virus 1 helicase-primase. Biochemistry 43(6), 1754-62. Ramirez-Aguilar, K.A., Low-Nam, N.A., Kuchta, R.D. (1995) Key Role of Template Sequence for Primer Synthesis by the Herpes Simplex Virus 1 Helicase-Primase. Biochemistry 41(49), 14569-79. Randell, J.C., Coen, D.M. (2004) The herpes simplex virus processivity factor, UL42, binds DNA as a monomer. Journal of Molecular Biology 335(2), 409-13. Rasmussen, L., Geissler, A., Winters, M. (2003) Inter- and intragenic variations complicate the molecular epidemiology of human cytomegalovirus. The Journal of Infectious Diseases 187(5), 809-19. Rawlinson, W.D., Barrell, B.G. (1993) Spliced transcripts of human cytomegalovirus. Journal of Virology 67(9), 5502-13.

138 Rawlinson, W.D., Farrell, H.E., Barrell, B.G. (1996) Analysis of the complete DNA sequence of murine cytomegalovirus. Journal of Virology 70(12), 8833-49. Razonable, R.R., Emery, V.C; 11th Annual Meeting of the IHMF (International Herpes Management Forum). (2004) Management of CMV infection and disease in transplant patients. 27-29 February 2004. Herpes: The Journal of the IHMF 11(3), 77-86. Reefschlaeger, J., Bender, W., Hallenberger, S., Weber, O., Eckenberg, P., Goldmann, S., Haerter, M., Buerger, I., Trappe, J., Herrington, J.A., Haebich, D., Ruebsamen-Waigmann, H. (2001) Novel non-nucleoside inhibitors of cytomegaloviruses (BAY 38-4766): in vitro and in vivo antiviral activity and mechanism of action. The Journal of Antimicrobial Chemotherapy 48(6), 757-67. Reeves, M.B., MacAry, P.A., Lehner, P.J., Sissons, J.G., Sinclair, J.H. (2005) Latency, chromatin remodeling, and reactivation of human cytomegalovirus in the dendritic cells of healthy carriers. Proceedings of the National Academy of Sciences of the United States of America 102(11), 4140-5. Ricart, M.J., Malaise, J., Moreno, A., Crespo, M., Fernández-Cruz, L; Euro-SPK Study Group. (2005) Cytomegalovirus: occurrence, severity, and effect on graft survival in simultaneous pancreas-kidney transplantation. Nephrology, Dialysis, Transplantation 20(Suppl 2:ii25-ii32), ii62. Ripalti, A., Boccuni, M.C., Campanini, F., Landini, M.P. (1995) Cytomegalovirus- mediated induction of antisense mRNA expression to UL44 inhibits virus replication in an astrocytoma cell line: identification of an essential gene. Journal of Virology 69(4), 2047–2057. Romanowski, M.J., Shenk, T. (1997) Characterization of the human cytomegalovirus irs1 and trs1 genes: a second immediate-early transcription unit within irs1 whose product antagonizes transcriptional activation. Journal of Virology 71(2), 1485-96. Ross, S.A., Boppana, S.B. (2005) Congenital cytomegalovirus infection: outcome and diagnosis. Seminars in Pediatirc Infectious Diseases 16(1), 44-9. Ross, S.A., Fowler, K.B., Ashrith, G., Stagno, S., Britt, W.J., Pass, R.F., Boppana, S.B. (2006) Hearing loss in children with congenital cytomegalovirus infection born to mothers with preexisting immunity. The Journal of Pediatrics 148(3), 332-6. Rowe, W.P., Hartley, J.W., Waterman, S., Turner, H.C., Huebner, R.J. (1956) Cytopathogenic agent resembling human salivary gland virus recovered from tissue cultures of human adenoids. Proceedings of the Society for Experimental Biology and Medicine 92(2), 418-24. Ruebner, B.H., Hirano, T., Slusser, R., Osborn, J., Medearis, D.N. Jr. (1966) Cytomegalovirus infection. Viral ultrastructure with particular reference to the relationship of lysosomes to cytoplasmic inclusions. The American Journal of Pathology 48(6), 971-89. Salzberger, B., Hartmann, P., Hanses, F., Uyanik, B., Cornely, O.A., Wöhrmann, A., Fätkenheuer, G. (2005) Incidence and prognosis of CMV disease in HIV- infected patients before and after introduction of combination antiretroviral therapy. Infection 33(5-6), 345-9. Sarisky, R.T., Hayward, G.S. (1996) Evidence that the UL84 gene product of human cytomegalovirus is essential for promoting oriLyt-dependent DNA replication and formation of replication compartments in cotransfection assays. Journal of Virology 70(11), 7398-413.

139 Schepers, A., Pich, D., Mankertz, J., Hammerschmidt, W. (1993) Cis-acting elements in the lytic origin of DNA replication of Epstein-Barr virus. Journal of Virology 67(7), 4237–4245. Scott, G.M., Weinberg, A., Rawlinson, W.D., Chou, S. (2007) Multidrug resistance conferred by novel DNA polymerase mutations in human cytomegalovirus isolates. Antimicrobial Agents and Chemotherapy 51(1), 89-94. Seale, H., MacIntyre, C.R., Gidding, H.F., Backhouse, J.L., Dwyer, D.E., Gilbert, L. (2006) National Serosurvey of Cytomegalovirus in Australia. Clinical and Vaccine Immunology 13(11), 1181–1184. Shen, C.Y., Chang, S.F., Lin, H.J., Ho, H.N., Yeh, T.S., Yang, S.L., Huang, E.S., Wu, C.W. (1994) Cervical cytomegalovirus infection in prostitutes and in women attending a sexually transmitted disease clinic. Journal of Medical Virology 43(4), 362-6. Shen, C.Y., Chang, S.F., Yang, S.L., Wu, C.W., Yang, Y.S., Hong Nerng Ho, Huang, E.S., Bantel-Schaal, U., Haefelin, D.N., Schieferstein, G. (1994a) Cytomegalovirus is present in semen from a population of men seeking fertility evaluation. Journal of Infectious Diseases 169(1), 222-223. Shen, C.Y., Chang, W.W., Chang, S.F., Chao, M.F., Huang, E.S., Wu, C.W. (1992) Seroepidemiology of cytomegalovirus infection among children between the ages of 4 and 12 years in Taiwan. Journal of Medical Virology 37(1), 72-5. Sherman, G., Gottlieb, J., Challberg, M.D. (1992) The UL8 subunit of the herpes simplex virus helicase-primase complex is required for efficient primer utilization. The Journal of Virology 66(8), 4884-92. Sigrist, C.J., De Castro, E., Langendijk-Genevaux, P.S., Le Saux, V., Bairoch, A., Hulo, N. (2005) ProRule: a new database containing functional and structural information on PROSITE profiles. Bioinformatics 21(21), 4060-6. Smith, C.A., Toogood, H.S., Baker, H.M., Daniel, R.M., Baker, E.N. (1999) Calcium- mediated thermostability in the subtilisin superfamily: the crystal structure of Bacillus Ak.1 protease at 1.8 A resolution. Journal of Molecular Biology 294(4), 1027-40. Smith, J.A., and Pari, G.S. (1995a) Human cytomegalovirus UL102 gene. Journal of Virology 69(3), 1734–1740. Smith, J.A., Jairath, S., Cryte, J.J., and Pari, G.S. (1996) Characterization of the human cytomegalovirus UL105 gene and identification of the putative helicase protein. Virology 220, 251–255. Smith, J.A., Pari, G.S. (1995) Expression of human cytomegalovirus UL36 and UL37 genes is required for viral DNA replication. Journal of Virology 69(3), 1925-31. Smith, M.G. (1956) Propagation in tissue cultures of a cytopathogenic virus from human salivary gland virus (SGV) disease. Proceedings of the Society for Experimental Biology and Medicine 92(2), 424-30. Soderberg-Naucler, C., Streblow, D.N., Fish, K.N., Allan-Yorke, J., Smith, P.P., Nelson, J.A. (2001) Reactivation of latent human cytomegalovirus in CD14(+) monocytes is differentiation dependent. Journal of Virology 75(16), 7543-54. Spaete, R.R., Mocarski, E.S. (1987) Insertion and deletion mutagenesis of the human cytomegalovirus genome. Proceedings of the National Academy of Sciences of the United States of America 84(20), 7213-7. Spector, D.J., Tevethia, M.J. (1994) Protein-protein interactions between human cytomegalovirus IE2-580aa and pUL84 in lytically infected cells. Journal of Virology 68(11), 7549-53.

140 Springer, K.L., Chou, S., Li, S., Giller, R.H., Quinones, R., Shira, J.E., Weinberg, A. (2005) How evolution of mutations conferring drug resistance affects viral dynamics and clinical outcomes of cytomegalovirus-infected hematopoietic cell transplant recipients. Journal of Clinical Microbiology 43(1), 208-13. Stagno, S., Pass, R.F., Cloud, G., Britt, W.J., Henderson, R.E., Walton, P.D., Veren, D.A., Page, F., Alford, C.A. (1986) Primary cytomegalovirus infection in pregnancy. Incidence, transmission to fetus, and clinical outcome. JAMA: The Journal of the American Medical Association 256(14), 1904-8. Stagno, S., Pass, R.F., Dworsky, M.E., Henderson, R.E., Moore, E.G., Walton, P.D., Alford, C.A. (1982) Congenital cytomegalovirus infection: The relative importance of primary and recurrent maternal infection. The New England Journal of Medicine 306(16), 945-9. Staras, S.A., Dollard, S.C., Radford, K.W., Flanders, W.D., Pass, R.F., Cannon, M.J. (2007) Seroprevalence of cytomegalovirus infection in the United States, 1988- 1994. Clinical Infectious Diseases 43(9), 1143-51. Steininger, C. (2007) Clinical relevance of cytomegalovirus infection in patients with disorders of the immune system. Clinical Microbiology and Infection 13(10), 953-63. Stropes, M.P., Miller, W.E. (2008) Functional analysis of human cytomegalovirus pUS28 mutants in infected cells. Journal of General Virology 89(1), 97-105. Sullivan, V., Biron, K.K., Talarico, C., Stanat, S.C., Davis, M., Pozzi, L.M., Coen, D.M. (1993) A point mutation in the human cytomegalovirus DNA polymerase gene confers resistance to ganciclovir and phosphonylmethoxyalkyl derivatives. Antimicrobial Agents and Chemotherapy 37(1), 19-25. Sullivan, V., Talarico, C.L., Stanat, S.C., Davis, M., Coen, D.M., Biron, K.K. (1992) A protein kinase homologue controls phosphorylation of ganciclovir in human cytomegalovirus-infected cells. Nature 358(6382), 162-4. Takeda, K., Haque, M., Nagoshi, E., Takemoto, M., Shimamoto, T., Yoneda, Y., Yamanishi, K. (2000) Characterization of human herpesvirus 7 U27 gene product and identification of its nuclear localization signal. Virology 272, 394– 401. Taylor-Wiedeman, J., Sissons, P., Sinclair, J. (1994) Induction of endogenous human cytomegalovirus gene expression after differentiation of monocytes from healthy carriers. Journal of Virology 68(3), 1597-604. Taylor, T.J., Knipe, D.M. (2003) C-terminal region of herpes simplex virus ICP8 protein needed for intranuclear localization. Virology 309(2), 219-31. Tenney, D.J., Hurlburt, W.W., Micheletti, P.A., Bifano, M., Hamatake, R.K. (1994) The UL8 component of the herpes simplex virus helicase-primase complex stimulates primer synthesis by a subassembly of the UL5 and UL52 components. The Journal of Biological Chemistry 269(7), 5030-5. Terhune, S., Torigoi, E., Moorman, N., Silva, M., Qian, Z., Shenk, T., Yu, D. (2007) Human Cytomegalovirus UL38 Protein Blocks Apoptosis. Journal of Virology In Print. Thompson, J.D., H.D.G.G.T.J. (1994) CLUSTAL W: improving the sensitivity of progressive multiple sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acid Research 22, 4673-4680. Tjiam, K.H., van Heijst, B.Y., Polak-Vogelzang, A.A., Rothbarth, P.H., van Joost, T., Stolz, E., Michel, M.F. (1987) Sexually communicable micro-organisms in human semen samples to be used for artificial insemination by donor. Genitourinary Medicine 63(2), 116-8.

141 Urwijitaroon, Y., Teawpatanataworn, S., Kitjareontarm, A. (1993) Prevalence of cytomegalovirus antibody in Thai-northeastern blood donors. The Southeast Asian Journal of Tropical Medicine and Public Health 24(1), 180-2. Vink, C., Beuken, E., Bruggeman, C.A. (2000) Complete DNA sequence of the rat cytomegalovirus genome. Journal of Virology 74(16), 7656-65. Wagstaff, A.J., Bryson, H.M. (1994) Foscarnet. A reappraisal of its antiviral activity, pharmacokinetic properties and therapeutic use in immunocompromised patients with viral infections. Drugs 48(2), 199-226. Walker, J.E., Saraste, M., Runswick, M.J., and Gay, N.J. (1982) Distantly related sequences in the alpha- and beta-subunits of ATP synthase, myosin, kinases and other ATP-requiring enzymes and a common nucleotide binding fold. The EMBO Journal 1(8), 945-51. Wang, W., Patterson, C.E., Yang, S., Zhu, H. (2004) Coupling generation of cytomegalovirus deletion mutants and amplification of viral BAC clones. Journal of Virological Methods 121(2), 137-43. Wang, Y.S., Hall, J.D. (1990) Characterization of a major DNA-binding domain in the herpes simplex virus type 1 DNA-binding protein (ICP8). Journal of Virology 64(5), 2082-2089. Wathen, M.W. (2002) Non-nucleoside inhibitors of herpesviruses. Reviews in Medical Virology 12(3), 167-78. Weber, O., Bender, W., Eckenberg, P., Goldmann, S., Haerter, M., Hallenberger, S., Henninger, K., Reefschläger, J., Trappe, J., Witt-Laido, A., Ruebsamen- Waigmann, H. (2001) Inhibition of murine cytomegalovirus and human cytomegalovirus by a novel non-nucleosidic compound in vivo. Antiviral Research 49(3), 179-89. Weiland, K.L., Oien, N.L., Homa, F., Wathen, M.W. (1994) Functional analysis of human cytomegalovirus polymerase accessory protein. Virus Research 34(3), 191-206. Weinberg, A., Jabs, D.A., Chou, S., Martin, B.K., Lurain, N.S., Forman, M.S., Crumpacker, C; Cytomegalovirus Retinitis and Viral Resistance Study Group; Adult AIDS Clinical Trials Group Cytomegalovirus Laboratories. (2003) Mutations conferring foscarnet resistance in a cohort of patients with acquired immunodeficiency syndrome and cytomegalovirus retinitis. The Journal of Infectious Diseases 187(5), 777-84. Weissman, J.S., Kim, P.S. (1993) Efficient catalysis of disulphide bond rearrangements by protein disulphide isomerase. Nature 365(6442), 185-8. White, E.J., Boehmer, P.E. (1999) Photoaffinity labeling of the herpes simplex virus type-1 single-strand DNA-binding protein (ICP8) with oligodeoxyribonucleotides. Biochemical and Biophysical Research Communications 264(2), 493-7. Wolf, D.G., Smith, I.L., Lee, D.J., Freeman, W.R., Flores-Aguilar, M., Spector, S.A. (1995) Mutations in human cytomegalovirus UL97 gene confer clinical resistance to ganciclovir and can be detected directly in patient plasma. The Journal of Clinical Investigation 95(1), 257-63. Wong, A., Tan, K.H., Tee, C.S., Yeo, G.S. (2000) Seroprevalence of cytomegalovirus, toxoplasma and parvovirus in pregnancy. Singapore Medical Journal 41(4), 151- 5. Wong, S.W., Wahl, A.F., Yuan, P.M., Arai, N., Pearson, B.E., Arai, K., Korn, D., Hunkapiller, M.W., Wang, T.S. (1988) Human DNA polymerase alpha gene expression is cell proliferation dependent and its primary structure is similar to

142 both prokaryotic and eukaryotic replicative DNA polymerases. The EMBO Journal 7(1), 37-47. Woycechowsky, K.J., and Raines, R.T. (2003) The CXC Motif: A Functional Mimic of Protein Disulfide Isomerase. Biochemistry 42(18), 5387-94. Xu, Y., Cei, S.A., Huete, A.R., Pari, G.S. (2004) Human cytomegalovirus UL84 insertion mutant defective for viral DNA synthesis and growth. Journal of Virology 78(19), 10360-9. Xu, Y., Cei, S.A., Rodriguez, Huete, A., Colletti, K.S., Pari, G.S. (2004a) Human cytomegalovirus DNA replication requires transcriptional activation via an IE2- and UL84-responsive bidirectional promoter element within oriLyt. Journal of Virology 78(21), 11664-77. Xu, Y., Colletti, K.S., Pari, G.S. (2002) Human cytomegalovirus UL84 localizes to the cell nucleus via a nuclear localization signal and is a component of viral replication compartments. Journal of Virology 76(17), 8931-8. Yang, Y.S., Ho, H.N., Chen, H.F., Chen, S.U., Shen, C.Y., Chang, S.F., Huang, E.S., Wu, C.W. (1995) Cytomegalovirus infection and viral shedding in the genital tract of infertile couples. Journal of Medical Virology 45(2), 179-82. Ye, L.B., Huang, E.S. (1993) In vitro expression of the human cytomegalovirus DNA polymerase gene: effects of sequence alterations on enzyme activity. Journal of Virology 67(11), 6339-47. Yu, D., Silva, M.C., Shenk, T. (2003) Functional map of human cytomegalovirus AD169 defined by global mutational analysis. Proceedings of the National Academy of Sciences of the United States of America 100(21), 12396–12401. Yu, Y., Alwine, J.C. (2002) Human cytomegalovirus major immediate-early proteins and simian virus 40 large T antigen can inhibit apoptosis through activation of the phosphatidylinositide 3'-OH kinase pathway and the cellular kinase Akt. Journal of Virology 76(8), 3731-8. Yust, I., Fox, Z., Burke, M., Johnson, A., Turner, D., Mocroft, A., Katlama, C., Ledergerber, B., Reiss, P., Kirk, O; EuroSIDA. (2004) Retinal and extraocular cytomegalovirus end-organ disease in HIV-infected patients in Europe: a EuroSIDA study, 1994-2001. European Journal of Clinical Microbiology and Infectious Diseases 23(7), 550-9. Zacny, V.L., Gershburg, E., Davis, M.G., Biron, K.K., Pagano, J.S. (1999) Inhibition of Epstein-Barr virus replication by a benzimidazole L-riboside: novel antiviral mechanism of 5, 6-dichloro-2-(isopropylamino)-1-beta-L-ribofuranosyl-1H- benzimidazole. Journal of Virological Methods 73(9), 7271-7. Zhang, J., Chung, D.W., Tan, C.K., Downey, K.M., Davie, E.W., So, A.G. (1991) Primary structure of the catalytic subunit of calf thymus DNA polymerase delta: sequence similarities with other DNA polymerases. Biochemistry 30(51), 11742-50. Zhu, H., Shen, Y., Shenk, T. (1995) Human cytomegalovirus IE1 and IE2 proteins block apoptosis. Journal of Virology 69(12), 7960-70. Zhu, L., Weller, S.K. (1992) The six conserved helicase motifs of the UL5 gene product, a component of the herpes simplex virus type 1 helicase-primase, are essential for its function. Journal of Virology 66(1), 469-79. Zuccola, H.J., Filman, D.J., Coen, D.M., and Hogle J.M. (2000) The crystal structure of an unusual processivity factor, herpes simplex virus UL42, bound to the C terminus of its cognate polymerase. Molecular Cell 5, 267-278.

143