TWO NOVEL MECHANISMS OF MHC CLASS I DOWN- REGULATION IN HUMAN CANCER: ACCELERATED DEGRADATION OF TAP-1 mRNA AND DISRUPTION OF TAP-1 FUNCTION

DISSERTATION

Presented in Partial Fulfillment of the Requirement for the Degree Doctor of

Philosophy in the Graduate School of The Ohio State University

by

Tianyu Yang

* * * * *

The Ohio State University 2004

Dissertation Committee: Approved by Professor Pan Zheng, Adviser

Professor Yang Liu

Professor Joan Durbin

Professor Xuefeng Bai Adviser Department of Pathology Professor William Lafuse

ABSTRACT

Both viruses and tumors evade cytotoxic T lymphocyte (CTL)-mediated host

immunity by down-regulation of major histocompatibility complex (MHC) class I

antigen presentation machinery. The transporter associated with antigen processing

(TAP), a heterodimer of TAP-1 and TAP-2, plays essential role in the MHC class I- restricted antigen presentation pathway by translocating antigenic peptides from cytosol into the ER lumen, where the assembly of MHC class I complex takes place. TAP deficiency is a frequent observation in human cancers, which is one of the causes of

MHC class I down-regulation. However, the underlying molecular mechanism has been limited.

In search for novel mechanisms of TAP deficiency and MHC class I down- regulation in human cancer cells, we characterized an MHC class I-deficient melanoma cell line SK-MEL-19. The expression of TAP-1 mRNA was found deficient in this cell line even after interferon-gamma (IFN-γ) stimulation, despite its active transcription.

This abnormality is caused by a single nucleotide deletion at position +1489 of the TAP-

1 , which results in rapid degradation of the TAP-1 mRNA. Subsequently, using this

TAP-1 deficient cell line, we studied the function of TAP-1 variants containing amino acid changes within or close to the signature motif that exist in the normal population or were generated by site-directed mutagenesis. All the TAP-1 variants showed decreased ii

function compared to the wild type, which supports an essential role of the signature motif in TAP transporter function. One of these variants, TAP-1 R648Q, with the alteration immediately C-terminal to the signature motif, was found to occur 17.5 times more frequently in HLA- colon cancers than those with normal HLA levels. Functional

analysis revealed that the Q648 variant retained only about 20% peptide translocation

activity compared to TAP-1 (R648).

To our knowledge, the two mechanisms we identified in this work, which lead to abrogation of TAP-1 expression and decreased TAP transporter function, respectively, have not been addressed before. The determination of transporter activity of a natural

TAP-1 variant, TAP-1 R648Q, and its increased presence in HLA- colon cancer samples

may help to develop diagnostic and therapeutic methods for colon cancer immunotherapy.

iii

DEDICATION

Dedicated to my parents.

iv

ACKNOWLEDGMENTS

I wish to thank my adviser, Dr. Pan Zheng for her support and encouragement, and for her patience in correcting my stylistic and scientific errors. I thank Dr. Yang Liu for his valuable advice. Without them, this thesis would not be possible.

I thank Philip E. Lapinski and Haotian Zhao for their contribution to this work.

I’m in debted to Dr. Soledad Fernandez for the statistical analysis and Dr. Qunmin Zhou for sharing normal population blood DNA samples.

I thank Dr. Jian-Xin Gao, Dr. Huiming Zhang, Dr. Kurtis Yearsley and Paul

Rangel for technical assistance.

I am grateful to Dr. Soldano Ferrone for providing us the SK-MEL-19 cell line,

Dr. Malini Raghavan for providing the peptide and technical assistance for the peptide translocation assay and Dr. Peter Cresswell for rabbit antiserum against TAP1 and .

This work was supported by NIH grants CA82355, CA58033 and CA69091,

Department of Defense grant DAMD17-00-1-0041 and the Ohio State University

Comprehensive Cancer Center.

v

VITA

November 8, 1974 ………………………Born, Hebei, People’s Republic of China

1998 ……………………………………..Bachelor in Medicine and Master in Medicine,

China medical University

1998 - present ……………………………Graduate Research Assistant, The Ohio State

University

PUBLICATIONS

Research Publication

1. T. Yang, B.A. McNally, S. Ferrone, Y. Liu, and P. Zheng, “A single-nucleotide deletion leads to rapid degradation of TAP-1 mRNA in a melanoma cell line”. J Biol Chem, 278(17), 15291-6, (2003).

2. Y. Guo, T. Yang, X. Liu, S. Lu, J. Wen, J.E. Durbin, Y. Liu, and P. Zheng, “Cis elements for transporter associated with antigen-processing-2 transcription: two new promoters and an essential role of the IFN response factor binding element in IFN- gamma-mediated activation of the transcription initiator”. Int Immunol, 14(2), 189-200, (2002).

FIELDS OF STUDY

Major Field: Immunology.

vi

TABLE OF CONTENTS

Page

ABSTRACT...... ii DEDICATION ...... iv ACKNOWLEDGMENTS...... v VITA ...... vi LIST OF TABLES ...... ix LIST OF FIGURES...... x 1. INTRODUCTION ...... 1 1.1 Major histocompatibility complex (MHC) class I-restricted antigen processing and presentation...... 1 1.2 Presentation of exogenous antigens by the class I MHC molecules...... 7 1.3 TAP-1 and protein function,...... 11 1.4 Impaired MHC Class I expression in viral infection...... 16 1.5 MHC Class I down-regulation in cancer...... 18 2. MATERIALS AND METHODS...... 30 2.1 Cell lines ...... 30 2.2 ...... 30 2.3 Flow cytometry ...... 31 2.4 Northern blot...... 31 2.5 Southern blot...... 32 2.6 Generation of TAP-1 cDNA constructs and stable transfection...... 32 2.7 Generation of TAP-1 constructs and Tet-Off SK-MEL-19 cell lines for RNA stability analyses ...... 34 2.8 Generation of luciferase reporter constructs and assay for promoter activity ...... 34 2.9 Nuclear run-on assay...... 35 2.10 Restriction fragment length polymorphism (RFLP)...... 36 2.11 RNase protection assay ...... 36 2.12 Immunofluorescence microscopy ...... 37 2.13 Immunoprecipitation...... 38 2.14 Microsome-based peptide translocation assay ...... 39 2.15 Colon cancer tissue sections, immunohistochemistry and statistical analysis...... 40 2.16 Isolation of genomic DNA and genotyping ...... 41 3. A SINGLE-NUCLEOTIDE DELETION LEADS TO RAPID DEGRADATION OF TAP-1 MRNA IN A MELANOMA CELL LINE ...... 42 3.1 Introduction...... 42 3.2 Results and discussion...... 44 3.2.1 Down-regulation of TAP-1 mRNA by a posttranscriptional mechanism in melanoma cell line SK-MEL-19 ...... 44 3.2.2 A single-nucleotide deletion leads to accelerated decay of TAP-1 mRNA ...... 47 vii

4. FUNCTIONAL CONSERVATION BETWEEN THE SIGNATURE MOTIF OF TAP-1 AND CYSTIC FIBROSIS TRANSMEMBRANE REGULATOR PROTEIN: A ROLE OF ITS POLYMORPHISM IN HLA DOWN-REGULATION IN CANCER CELLS...... 65 4.1 Introduction...... 65 4.2 Results...... 68 4.2.1 CF-like mutations in TAP-1 signature motif abrogate TAP-1 activity...... 68 4.2.2 TAP-1 G646D is expressed in the ER and remains associated with the peptide- loading complex...... 69 4.2.3 TAP-1 G646D failed to transport peptide across the ER membrane...... 71 4.2.4 Increase of the G1943A allele frequency in HLAlow colon cancer tissues...... 72 4.2.5 Function of TAP-1 R648Q...... 73 4.3 Discussion...... 75 5. CONCLUSION...... 96

6. BIBILIOGRAPHY...... 99

viii

LIST OF TABLES

Table Page

Table 4.1 Nucleotide changes and amino acid substitutions generated by the site directed mutagenesis. ....79

Table 4.2 TAP-1G1943A genotyping in different types of human cancer...... 80

ix

LIST OF FIGURES

Figure Page

Figure 1.1 Conventional MHC class I antigen processing and presentation pathway...... 26

Figure 1.2 Genomic localization and promoter structure of human TAP-1 gene...... 28

Figure 1.3 Alignment of the nucleotide-binding domains (NBDs) of ABC transporters...... 29

Figure 3.1 Deficiency of surface HLA class I expression in melanoma cell line SK-MEL-19 was due to the TAP-1 down-regulation...... 52

Figure 3.2 Posttranscriptional mechanisms are responsible for poor accumulation of TAP-1 mRNA...... 55

Figure 3.3 TAP-1 mRNA level in SK-MEL-19 cells was increased by CHX...... 58

Figure 3.4 A homozygous single-nucleotide deletion was identified in the TAP-1 gene at position +1489 that resulted in premature termination codons...... 60

Figure 3.5 The single-nucleotide deletion D1489 in the TAP-1 gene results in accelerated decay of the TAP-1 mRNA by a non-NMD-related mechanism...... 62

Figure 4.1 Sequence alignment of the NBDs of ABC transporters...... 81

Figure 4.2 Expression of TAP-1 mutant and cell surface MHC class I in cells stably transfected with wild type or mutant TAP-1...... 82

Figure 4.3 TAP-1 G646D was expressed in the ER and associated with TAP-2, tapasin, and MHC class I heavy chain...... 84

Figure 4.4 TAP-1 G646D failed to transport peptide across ER membrane...... 88

Figure 4.5 Immunohistochemical staining of colon cancer with HC-10 mAb...... 90

Figure 4.6 Defective activity of TAP-1 R648Q...... 92

x

CHAPTER 1

INTRODUCTION

1.1 Major histocompatibility complex (MHC) class I-restricted antigen processing

and presentation.

CD8+ cytotoxic T lymphocytes (CTLs) are the major players of immune system to

fight against tumor cells and viral infections. CTLs recognize antigenic peptides

presented by class I MHC molecules on the target cell surface. The interaction between T

cell receptors (TCRs) and peptide-MHC complexes is a fundamental event for the

adaptive immune system. It mediates the positive and negative selections during the

thymus development of CD8+ T cells. It is indispensable for the survival and activation of

naïve T cells, effector function of activated T cells and their transition to memory cells

[1, 2]. Blocking surface expression of MHC class I in host cells is one of the strategies

certain viruses use to escape immune surveillance by CTLs [3-5]. Down-regulation of

MHC class I on tumor cells has also been frequently observed in a variety of cancers,

which is thought to facilitate tumor cells to evade immune surveillance by CTLs and

poses a major obstacle to the T-cell based cancer immunotherapy [6, 7].

1

MHC class I molecules are expressed by virtually all the nucleated cells except those of the immune-privileged organs, such as brain, testis and eyes. The MHC class I complex expressed on the cell surface consists of two non-covalently associated polypeptide chains, an α or heavy chain (HC) and a β2- (β2m) or light chain, as well as the antigenic peptide. The gene encoding for the class I HC is located in the MHC region on the short arm of human 6 and the corresponding position of murine chromosome 17. The HC of class I MHC is a trans-membrane glycoprotein, whose extra-cellular part folds into three domains: α1, α2 and α3. The α3 domain interacts with the non-membrane spanning β2m and the folded structure they form closely resembles that of the immunoglobulin constant domain. The α1 and α2 domains, which are distal to the membrane, form the peptide-binding platform and make direct contact with TCR [8]. Like other membrane proteins, MHC class I heavy chain is synthesized on the ER membrane with its extra-cellular domain exposed to the ER lumen.

It is inside the ER where HC associates with β2m and is subsequently loaded with antigenic peptides.

The generation of antigenic peptides from intact proteins, their assembly with class I HC and β2m, as well as their final display on the cell surface have been studied extensively in the past two decades [9-11]. Figure 1.1 shows a simplified representation of the conventional class I antigen processing and presentation pathway. To briefly describe, antigenic peptides are generated from endogenous proteins by the proteasome, and then transported into the ER lumen by the transporters associated with antigen

2

processing (TAP). Inside the ER, some of the antigenic peptides will be loaded onto the empty class I HC-β2m heterodimer with the help of a so-called peptide-loading complex, which is comprised of tapasin, calreticulin, ERp57 and TAP.

Class I restricted antigenic peptides are mostly generated in the cytosol from endogenous cellular proteins or intracellular pathogen derived proteins, mostly from virus. Proteasome, the major proteolytic machinery in eukaryotic cells that degrades ubiquitin-conjugated proteins or polypeptides, provides the vast majority of antigenic peptides for the class I restricted presentation [12-14]. The 20S proteasome is composed of two heptameric outer rings of α subunits and two hexameric inner rings of β subunits arranged in a hollow-cylinder shape. Upon cytokine stimulation, such as gamma- interferon (IFN-γ), three active β subunits of the 20S proteasome, X, Y and Z, are replaced by the multicatalytic complex-like 1 (MECL1), low molecular weight proteasome subunit (LMP)-2 and LMP-7 [15-18]. The LMP-2/7-containing proteasome exhibits enhanced trypsin- and chymotrypsin- like activity, which generates peptides with a basic or hydrophobic C-terminal residue that favors their subsequent binding to the TAP transporter and MHC class I molecules [19-22]. Although LMP-2 and

LMP-7 are dispensable for the overall class I antigen presentation in some studies [23-

25], decreased cell surface class I expression and defects in class I-restricted T cell response have been observed with the LMP-2 or LMP-7 deficient mice [26, 27].

The 20S proteasome can function independently or with regulatory subunits associated at both ends, which greatly enhances the proteolytic activity of the 20S proteasome. Such regulatory subunits include the 19S (PA700) subunit [28], which is

3

comprised of multiple subunits including six ATPases, and the 11S (PA28; REG) regulatory complex [29], which consists of two IFN-γ inducible subunits PA28α and

PA28β. Together with the 20S proteasome core, they form the 26S proteasome (19S-20S-

19S), the PA28-proteasome complex (11S-20S-11S) or the 19S-20S-PA28 complex, which are all involved in the generation of antigenic peptides for MHC class I-restricted presentation. However, proteasome is not the only source of antigenic peptides presented by class I MHC. Recently, cytosolic peptidase that can generate antigenic peptides independent of proteasome system has been identified, such as II

(TPPII) [30]. In human, some of the signal peptides produced in the ER during protein synthesis can bind to the MHC class I molecules directly [31].

The carboxyl termini of antigenic peptides generated by the proteasome are hydrophobic or charged, which favors their bindings to MHC [32]. The amino termini, however, are mostly extended and vary greatly. Some cytosolic have been identified that are involved in trimming the amino termini of these peptides, which include tripeptidyl peptidase II, puromycin-sensitive (PSA), bleomycin (BH) and the IFN-γ inducible (LAP) [33, 34]. In the cytosol, antigenic peptides are subjected to the attack by other peptidases, such as thimet oligopeptidase (TOP), and some may be destroyed completely [35]. In fact, most peptides are lost at this stage before being presented by class I MHC molecules [36].

4

It is not clear yet whether the peptides that survive are chaperoned in the cytosol

while being modified or protected on their way to the ER. Cytosolic heat shock proteins

(HSPs), such as Hsp70 and Hsp90, are the possible candidates. The important role of cytosolic HSPs in class I antigen presentation pathway is supported by studies by Binder

et al [37], in which they treated cells with deoxyspergualin, a drug that binds hsp70 and

hsp90, and observed abrogation of MHC I antigen presentation, which was restored by

introduction of additional hsp70 into the cytosol. In another study, overexpression of

hsp72 in the B16 melanoma cell line leads to enhanced antigen presentation by class I

MHC on the cell surface [38]. Although physical association of HSPs with proteasomes

and TAP transporters has been detected, respectively [39, 40], the association of

antigenic peptides or their precursors with HSPs or other chaperone proteins in the cytosol has not been directly detected.

The peptides that manage to reach the ER membrane may bind to the TAP transporters on the ER membrane. TAP transporters are capable of translocating the peptides from cytosol into the ER lumen utilizing energy released from ATP hydrolysis.

Inside the ER, the newly synthesized MHC class I HC binds to ER chaperone protein calnexin or immunoglobulin binding protein (BiP), which facilitates its proper folding and assembly with β2m [41]. The empty HC-β2m heterodimer will subsequently associate with multiple ER-resident proteins to form a peptide-loading complex, which consists of: calreticulin, another ER chaperone protein [42, 43]; ERp57, a thio-reductase involved in disulfide bond formation of MHC class I molecules [44-46]; tapasin, a transmembrane glycoprotein that mediates the association of HC-β2m heterodimer with

5

TAP transporter [42, 47]; and the TAP transporter, which forms the “core” of the loading complex and provides antigenic peptides generated in the cytosol. Tapasin is the key component in this complex. Unlike other chaperone proteins, such as calnexin and calreticulin, which are dispensable for MHC class I expression, tapasin is specific for this process and its deficiency results in decreased cell surface MHC class I level [42, 47].

Tapasin is capable of interacting with both TAP-1 and TAP-2 through its C-terminal region and with MHC class I HC via the 50 residues on its N-terminus [48]. The interaction between tapasin and TAP is sufficient to stabilize TAP proteins and increase their expression levels without significantly affecting the intrinsic translocation activity

[48, 49]. Tapasin is also linked to ERp57 via disulfide-bond formation, which has been shown required for the complete oxidation of class I HC in the loading complex [50].

However, the interaction between tapasin-MHC complex and TAP is not necessary for the peptide loading as a soluble form of tapasin, which is unable to bind TAP, still supports the loading of MHC class I molecules with antigenic peptides [51, 52].

It is thought that formation of this multimolecular peptide-loading complex may allow the assembly of MHC class I molecules with their peptide cargo to optimize over time. The antigenic peptides transported by the TAP transporters are usually 8-15 amino acid long, which is longer than those presented by the class I MHC molecules (8-10 amino-acid long). As mentioned above, although the carboxyl termini of these peptides are hydrophobic or charged, which favor their bindings to MHC, their amino termini can be varied and extended, which are subjected to further trimming by the ER aminopeptidases until the best fitters are selected by the empty HC-β2m dimers. Such an

6

ER resident aminopeptidases has been identified recently, named ER aminopeptidase I

(ERAPI). The expression of ERAPI can be induced by IFN-γ and its deficiency results in

decreased cell surface expression of class I MHC complexes in IFN-treated cells [53-55].

ERAPI was shown capable of cleaving peptide precursors of 10-residue or longer,

generated in cytosol or ER, to 8-9 residues that are fit for presentation by class I MHC

molecules [53-55]. Peptides of 8-residue or shorter are spared by ERAPI. It destroys

about half of the 9-residue peptides in vitro, which may limit antigen presentation and

explain the phenomenon that without IFN-γ treatment, lack of ERAPI increases class I

surface expression [54, 55].

After the MHC class I complexes are loaded with proper antigenic peptides, they

will dissociate from the peptide-loading complex, be released from ER and go to the cell

surface via the secretory expression pathway. Those that fail to bind any peptides will be

retained in the ER and eventually translocated out of the ER via the translocon and

subsequently degraded by the proteasome in the cytoplasm [56, 57]. Only under

temperature below physiological level, the unstable empty MHC class I molecules can be

expressed on the cell surface [58].

1.2 Presentation of exogenous antigens by the class I MHC molecules.

The conventional class I antigen presentation pathway described above is the

machinery by which endogenous antigens are processed and displayed on the cell

surface, which then can be scrutinized by CD8+ T lymphocytes, the major players of

immune system to fight against viral infection and cancer cells. However, the naïve CD8+

7

T cells cannot be activated unless two conditions are fulfilled: 1) two stimulatory signals

are required [59], one delivered through the engagement of MHC class I/antigen and T

cell receptor/CD8, the other through the interaction between the co-stimulatory molecules

on T cells (such as CD28) and their ligands on the antigen presenting cells (APCs) (such

as B7.1 and B7.2). Although MHC class I complexes are universally expressed, ligands

for the costimulatory molecules are only expressed by the professional APCs, including

dendritic cells (DCs), macrophages and B cells. 2) The antigens have to be seen by the

naïve CD8+ T cells, which, however, circulate in the blood and lymphatics after their

maturation in thymus and do not go to the peripheral tissues [60]. Lymphoid organs,

especially lymph nodes (LNs), are the most likely sites of T cell activation [61].

Therefore, when tissue cells other then the professional APCs are infected with virus or

undergoing malignant transformation, they are unable to alert the CD8+ T cells by

themselves. The immune system has evolved a way to activate the naïve CD8+ T cells in

such a situation that the antigens are picked up by the professional APCs that are billeted

in the peripheral tissues, especially the bone marrow-derived DCs, which if turning

activated and mature, will migrate to the draining LN and present the anithgens to the

naïve CTLs [62-64]. The process that exogenous antigens are processed and presented by

the class I MHC molecules on the professional APCs is termed “cross-presentation”.

Traditionally, exogenous antigens are presented by the class II MHC complexes, which, unlike the class I MHC, are expressed only by the professional APCs on their surface. Since the emphasis of this thesis is on the class I antigen presentation machinery, the class II-restricted antigen presentation pathway is summarized here in an over-

8

simplified way. The exogenous antigens are taken up by the professional APCs via

endocytosis or phagocytosis, and then degraded by in the acidic endocytic compartments. The class II molecules are synthesized in the ER, where they associate with the invariant chain (Ii) molecule, which not only prevents the binding of antigenic peptides in the ER, but also targets the complexes to a specialized endosomal compartment, MHC class II-containing compartment (MIIC), where they will be loaded with the antigenic peptides derived from the exogenous antigens and then expressed on the cell surface.

The presentation of exogenous antigens by class I MHC molecules is distinct from both the conventional class I antigen presentation and the class II antigen presentation machinery. Up to date, although not fully understood, two cross-presentation pathways have been proposed, the phagosome-to-cytosol pathway and the post-Golgi loading pathway or vacuolar pathway.

In both pathways, the exogenous antigens are internalized by phagocytosis or endocytosis. The internalization process is much more efficient when the exogenous antigens are associated with antibodies, which can bind to the Fc receptors on the surface of APCs, or when the exogenous antigens are chaperoned by the HSPs, which APCs also express surface receptors for, such as CD91 [65, 66]. In the phagosome-to-cytosol pathway, upon entry into the cells, phagosomes containing exogenous antigens may fuse with the ER and thereby acquire the empty class I MHC HC-β2m dimers as well as other

ER resident proteins, including those involved in the conventional class I restricted antigen presentation pathway [67, 68]. The exogenous antigens are then translocated out

9

of the phagosomes via a translocon Sec61 (or possibly by other mechanisms) and ubiquitinated by the ubiquitin-conjugating (UBCs) nearby [67, 68]. The antigens are subsequently cleaved by the proteasomes that are associated with the phagosome membrane. Or they may diffuse through the cytosol, which subjects them to degradation by the free proteasomes and cytosolic proteases. The antigenic peptides with proper length and sequence will be transported back into the ER-fused phagosomes or ER by the TAP transporters and then follow the conventional class I antigen presentation pathway as described above.

In the vacuolar pathway, the exogenous antigens remain in the phagosomes and are degraded in situ by the proteases present in the phagosomes. It is not clear yet where the MHC class I molecules come from. One possibility is that the phagosomes fuse with the ER, as in the phagosome-to-cytosol pathway, and thus acquire ER proteins including the empty MHC class I HC-β2m dimers and tapasin. The other possibility is that the cell surface MHC class I molecules have been taken into the phagosomes together with the exogenous antigens. In the latter case, the MHC class I molecules are very likely to be peptide-loaded and a peptide exchange process might occur in the endosomal compartments, probably with the help of tapasin [69]. After loaded with antigenic peptides, these MHC class I complexes could recycle to the cell surface. In support of the existence of such a vacuolar pathway, peptide exchange on class I molecules at low pH condition has been observed in vitro [70]. Mice that express transgenic GPI-linked MHC class I molecules, which can not be recycled, fail to generate CTL response against

10

particular antigenic epitope, which indicates that MHC class I recycling is required for presentation of certain antigenic epitopes [71]. However, the vacuolar pathway is rather speculative and needs more experimental evidence.

1.3 TAP-1 gene expression and protein function,

As described above, TAP transporter plays pivotal role in both the conventional class I MHC restricted antigen presentation and the cross-presentation pathway. It is the only known transporter on the ER membrane providing access for the antigenic peptides generated in the cytosol to the ER lumen.

The importance of TAP transporter to the class I restricted-antigen presentation was first noticed in studies on several mutant cell lines that were mutated and then selected for the loss of class I expression on the cell surface. These cell lines include the derivatives of a human B-lymphoblastoid cell line B-LCL 721, 721.134 and 721.174, as well the mutant RMA-S cell line generated from murine lymphoma RMA cells [72, 73].

In these mutant cell lines, MHC class I HC and β2m are expressed but fail to bind antigenic peptides, and thus unable to be presented on the cell surface under physiological condition. The surface expression of MHC class I molecules can be restored when peptides are given exogenously [74]. The genetic defects were later mapped to the MHC class II gene locus, from where two were cloned, now known as TAP-1 and TAP-2 [75-78]. The class I expression defects in the mutant cell lines were corrected when TAP-1 and/or TAP-2 were expressed [79-81]. Subsequently biochemical experiments show that TAP-1 and TAP-2 polypeptides form a heterodimeric transporter

11

on the ER membrane and transport antigenic peptides across the ER membrane in an

ATP-dependent manner [82-84]. The necessary role of TAP in the class I antigen presentation pathway was further supported by the TAP-1 knockout mice, which showed

severely decreased surface MHC class I expression and impaired CD8+ T cell development [85].

Genes that encode for TAP-1 and TAP-2 are clustered with those for the proteasome subunits, LMP-2 and LMP-7 (Figure 1.2). TAP-1 and LMP-2 genes share a bi-directional promoter of 593 base-pair long [86]. Although there is no TATA box homology at either end, a prevalence of GC boxes is prominent [86]. The region proximal to the TAP-1 gene has been shown to control the maximal basal level expressions of both

TAP-1 and LMP-2 and several cis elements have been identified in this region, including an Sp1 element, an NF-kappa B , the partially overlapping IFN consensus sequence (ICS)-1/2 and a gamma activating sequence (GAS) [86]. The Sp1 site is required for the basal expression of both genes and the NF-kappa B binding site is essential for the induction of both genes by tumor necrosis factor-alpha (TNFα) [86]. The rapid induction of the promoter by IFN-γ is mediated by STAT-1 that binds to the GAS element [87, 88]. The binding of IRF-1 to the ICS/GAS region supports constitutive expression of both genes and augments the induction by IFN-γ [89]. In the IRF-1- deficient mice, expressions of TAP-1 and LMP-2, as well as the surface class I MHC are greatly reduced, which leads to the defective CD8+ T cell development [89, 90]. In

human melanoma cells, TAP-1 is constitutively expressed, the expression of LMP-2

12

gene, however, requires IFN-γ stimulation [91]. Further analysis show that whereas

binding of either IRF-1 or STAT-1 is sufficient for the transcription of TAP-1 gene,

binding of both factors is required for LMP-2 gene transcription [91].

TAP-1 and TAP-2 polypeptides belong to the ATP-binding-cassette (ABC)

superfamily, which is comprised of a diverse class of proteins including not only the transmembrane transporters that carry various substances, ranging from ions to proteins, across membrane, but also the non-transporter enzymes that are involved in DNA repair or structural maintenance of [92, 93]. A prototype ABC transporter is composed of two paired membrane integral domains and two paired nucleotide-binding domains (NBDs). For some ABC transporters, such as CFTR, these domains are fused as a single polypeptide chain, while for others, like TAP, two polypeptides are separately encoded and each consists of one transmembrane domain and one NBD.

In ABC transporters, the paired hydrophobic membrane-spanning regions are thought to form the transmembrane pore through which substrates get across the membrane. The precise membrane topology of TAP-1 and TAP-2 is still controversial. It is predicted that the pore-forming domains of TAP-1 and TAP-2 assume a head-head/tail- tail orientation to each other [94]. Both N- and C- termini of TAP-1 reside in the cytoplasm, whereas for TAP-2, the N-terminus is facing the ER lumen and the C- terminus remains in the cytosol [94-97]. At least 8 and 7 transmembrane fragments have been proposed for TAP-1 and TAP-2, respectively [94-97]. Despite the lack of precise

13

structure, the membrane-spanning regions of TAP-1 and TAP-2 have been shown important for their ER retention, dimerization, interaction with tapasin, peptide binding, as well as peptide translocation [94-97].

The substrate binding is not conserved among ABC family members. The antigenic peptide binding sites of TAP have been mapped to the regions of the cytosol-

ER membrane boundaries of the two transmembrane segments close to the ATP binding site of NBDs [98-100]. TAP preferentially transports peptides 8-15 residue long [101].

Peptides longer or shorter can still be translocated, but with much reduced efficiency. The amino acid composition of peptide, especially on the carboxyl terminus, is also critical for TAP-mediated translocation [102-104]. Human TAP has been shown to efficiently transport peptides with hydrophobic or basic C-termini, which are characteristic of those produced by the IFN-induced proteasome [102-104]. In addition, the N-termini of peptides may also limit their access to the ER lumen [105].

The NBDs of ABC transporters, including TAP, are paired and serve as an

ATPase that binds and hydrolyzes ATP molecules to provide energy for the substrate translocation process. Like other members of the ABC superfamily, TAP-1 and TAP-2 contain several conserved sequence motifs in their NBDs, including Walker A and

Walker B, signature motif or C loop, Q loop, D loop and switch region (Figure 1.3). The

Walker A and Walker B domains are required for the ATP binding and ATP hydrolysis, whereas roles of the other motifs are largely unknown. In the past ten years, the peptide translocation process of TAP has begun to be elucidated. By disrupting the Walker domains of TAP-1 and TAP-2 individually, it has been shown that nucleotide binding by

14

TAP precedes peptide binding [106]. TAP-1 and TAP-2 play distinct roles in this process that nucleotide binding and hydrolysis by TAP-2 is required for the binding of peptide, whereas hydrolysis by TAP-1 is not required [107-109]. Successful peptide binding then induces both the ATPase activity of TAP and the conformational changes associated with peptide translocation [110, 111], for which the integrity of the Walker domains of both

TAP subunits is necessary [107, 108].

It has been shown that the NBDs of ABC transporters, including TAP, interact with each other, which is thought to mediate the domain communication and conformational transmission in the substrate translocation process [112]. The signature motif with consensus LSGGQ/E sequence has been shown important for the NBD interaction [113-115]. Signature motif is conserved among all the ABC transporters and is specific for the ABC superfamily. Recent studies on the crystal structure of several

ABC ATPase have shown that in the biologically relevant dimer of NBDs, the signature motif of one NBD forms a composite ATP-binding site at the dimer interface with the

Walker motifs of the other NBD [113, 116]. The two most conserved residues of signature motif, the serine and second glycine, form specific hydrogen bonds to the phosphate oxygens of the ATP molecule [113, 116]. In this way, the signature motif may act as a γ-phosphate sensor and mediate the ATP-driven conformational changes during the substrate translocation process.

The important role of signature motif in the ABC transporter function has been reflected by the fact that its mutations led to loss or impairment of the transporter function of several ABC transporters, including the human cystic fibrosis transmembrane

15

conductance regulator (CFTR) [117-120], yeast cadmium factor (Ycf1p)[121], yeast a- factor transporter Ste6p[122, 123], the bacterial maltose-transport system ATPase subunit

MalK [124], and TAP [125]. However, it is not known whether the function of signature motif is conserved among these ABC family members.

In humans, mutations in the signature motif of CFTR are among the frequent causes of cystic fibrosis (CF) [126]. Whether the signature motifs of other ABC transporters are also prone to mutations has not been addressed. Although natural variants of both human TAP-1 and TAP-2 have been identified in the normal population, the amino acid changes in these variants are not in the signature motif. Since congenital TAP deficiency in humans results in lack of cell surface MHC class I expression, the type 1 bare lymphocyte syndrome [127], and TAP deficiency has been frequently abserved in a variety of human cancers, which causes surface MHC class I down-regulation of cancer cells, it will be interesting to examine whether the signature motif of TAP are frequently mutated in these disease conditions.

1.4 Impaired MHC Class I expression in viral infection.

CTLs are the soldiers that fight against viral infection in our body. They need to see the targets, the antigenic peptides presented by the cell surface MHC class I molecules, to fulfill their job. To disguise themselves, the intruding viruses have evolved complicated strategies to block the expression of MHC class I on their host cells. Every step in the class I antigen processing and presentation pathway can be interfered by these sophisticated mechanisms. For example, the Epstein-Barr virus encoded nuclear antigen

16

(EBNA)-1 contains internal glycine-alanine repeats that inhibit its proteasomal degradation, therefore the antigenic peptides cannot be produced [128]. The Herpes simplex virus (HSV)-1 and HSV-2 express an infected cell protein (ICP)-47. ICP-47 inhibits translocation of antigenic peptides from cytosol into the ER by binding to the

TAP transporters and thus preventing other peptides from binding to TAP [129-131]. The human cytomegalovirus (HCMV) genomic unique short (US) region encodes a family of homologous genes essential for the inhibition of MHC class I-mediated antigen presentation during viral infection. Among them, US6, a transmembrane protein on the

ER membrane, interacts with TAP transporter via its ER luminal domain [132]. Unlike

ICP-47, US6 blocks the peptide translocation activity of TAP without affecting peptide binding [133].

The MHC class I molecules themselves can also become the targets of viruses.

US2 and US11, glycoproteins expressed by HCMV, bind to and dislocate nascent MHC class I HC from the ER into the cytosol through translocon Sec61 [56, 134, 135]. The class I HC subsequently will be deglycosylated and degraded by the proteasome [56, 134,

135]. The mK3 protein of γ2-Herpesvirus 68 also targets the degradation of nascent class

I molecules via the ubiquitin/proteasome pathway [136-139]. However, the mK3 protein has to be associated with members of the peptide-loading complex (TAP/tapasin) [139].

The association is not only critical for stable mK3 expression, but also required by mK3 to specifically recognize and ubiquitinate class I MHC molecules [139]. Adenovirus type

2 and 5 express a 19KD glycoprotein, E3/19K, that has two ways to interfere with class I surface expression: 1) it binds to TAP and acts as a tapasin inhibitor, thus preventing

17

class I/TAP association, which delays the maturation of class I MHC complexes [140]; 2) it retains the peptide-loaded MHC class I molecules in the ER/cis-Golgi compartment

[141, 142]. Another viral product that can cause retention of peptide-associated MHC class I molecules in the ER is US3, an ER resident glycoprotein expressed by HCMV

[143, 144]. After being expressed on the cell surface, MHC class I complexes can still be

down-regulated. For example, the Nef protein of human immunodeficiency virus (HIV)

is able to target surface class I MHC to accelerated endocytosis [145].

Above all, all the key steps in the class I restricted antigen presentation pathway

can be targeted by viruses to decrease the surface presentation of viral antigens by the

host cells. One virus may utilize different escape strategies at different times during the

infectious cycle, such as the US proteins of HCMV.

1.5 MHC Class I down-regulation in cancer.

MHC class I down-regulation has been frequently encountered in a variety of

cancers, which is thought to facilitate tumor cell escape from immune surveillance by

CTLs [7, 146, 147]. Loss of HLA class I in human cancers can be at different levels, i.e.

total loss of class I, loss of expression of one locus or one haplotype, or even one specific

allele [148]. Loss of class I expression is observed more frequently in metastatic lesions

than in primary lesions from studies of various of cancers, such as breast cancer,

squamous cell carcinoma of the head and neck, cervical carcinoma, prostate cancer, colon

cancer, etc [149-156]. It has been correlated with disease progression and may serve as a

marker of poor prognosis for certain human cancer [153, 157, 158]. In a mouse model,

18

mice that develop breast carcinoma spontaneously due to the expression of HER-2/Neu

oncogene showed down-regulation of MHC class I expression as the tumor progressed

[159]. In transgenic mice that develop melanoma as a result of the expression of SV40 large T antigen in melanocytes, MHC class I expression was also found to decrease with

tumor growth [160]. It has been shown that tumor cells that have lost surface MHC class

I grow in the presence of pre-existing strong anti-tumor immune response, which

provides direct evidence that loss of MHC class I expression may result in immune

evasion of tumor cells [161].

In the past two decades, research on tumor immunology and human tumor

immunotherapy has made tremendous progress [162]. Since treatments are not

effective except those using antibodies against growth factor receptors, tumor

immunotherapy in human has been focused on generating tumor specific cellular immune

responses, which can be achieved by active immunization or passive transfer and further

enhanced by cytokine treatment, such as interleukin (IL)-2 [162]. As more and more class

I-restricted tumor associated antigens (TAAs) are identified, both the active and the

passive approaches have been improved [163]. It is now feasible to generate and expand

enough tumor specific CD8+ T cells in vitro or in vivo to kill tumor cells in patients [162,

164]. Though complete and partial remission has been reported in cancer patients

receiving T cell-based immunotherapy in clinical trials, relapse is not uncommon [165,

166]. Factors that limit the effectiveness of anti-tumor immune response include, among

others, lack of help from CD4+ T cells, active tolerization of T cells due to absence of co-

19

stimulatory molecules on the tumor cells, apoptosis of T cells induced by interacting with tumor targets, loss of surface MHC class I expression or loss of tumor antigen expression in tumor cells [7].

In patients, relapse due to loss of class I expression after effective immunotherapy has been observed [167]. Apparently, lack of class I HLA on the surface makes tumor cells invisible to the tumor specific CD8+ T lymphocytes. The number of HLA molecules present on the surface of tumor cells has been shown to affect their lysis by CTL quantitatively in situations with borderline amounts of peptide and/or MHC [168]. As class I down-regulation is frequently encountered in a variety of cancers, it has become one of the major obstacles to the T cell based tumor immunotherapy [169]. In some cases, the decreased surface expression of MHC class I complex can be restored by treatment with cytokines, such as IFN-γ [170-173]. For those that do not respond to the cytokine treatment, elucidation of the underlying mechanisms for loss of class I expression in cancer cells may provide information for developing effective diagnostic or therapeutic methods to improve cancer immunotherapy. However, unlike viral infection, the causes of MHC class I down-regulation in cancer can vary greatly due to the genetic instability and rapid growth of cancer cells. Reports on causes of MHC class I down- regulation in tumor cells have been diverse, which are summarized here.

As described above, antigen presentation by MHC class I molecules requires coordinated expression of class I HC, β2m, proteasome and its inducible subunits

(MECL, LMP-2/7), the ER aminopeptidases, the TAP transporter, tapasin and other ER chaperone proteins. The deficiency of any of these molecules may affect the class I-

20

restricted antigen processing and presentation quantitatively or qualitatively. In some cases, these factors are coordinated down-regulated [171, 172, 174, 175]. However, selective down-regulation of individual components of the class I antigen presentation machinery is more often reported. Lack of functional β2m causes total loss of class I expression as the class I HC is unstable without its association, which, though rare, has been found in metastatic melanomas and lung cancers [167, 176]. Defects in the expression of tapasin, LMP-2/7 or TAP-1/2 were found to cause down-regulation of total level of class I HLA on the surface of human cancer cells [173, 177-181]. Decreased

PA28 expression in tumor cells leads to impaired presentation of a human TRP2 tumor antigen [29].

Some of the underlying molecular mechanisms have been identified, though limited, for the class I antigen presentation machinery defects. Loss of heterozygosity

(LOH) is an important mechanism that leads to loss of gene expression in cancer cells. It contributes to the defective MHC class I antigen presentation as well. LOH occurs at frequencies of 25%, 29% and 28%, respectively, in the MHC region on chromosome 6p,

β2m gene on chromosome 15q and the putative HLA class I modifier of methylation gene (MEMO)-1 on chromosome 1p in breast cancer patients [183]. LOH causes β2m inactivation and thus total loss of MHC class I in microsatelite instability (MSI)-positive colorectal cancer cells [184]. HLA deficiency has been attributed to LOH in MHC class I region in melanoma and laryngeal carcinomas [185-188]. The allelic imbalances on

21

chromosome 6p has been frequently observed in primary mediastinal B-cell lymphoma

(PMBL) [189]. In one case, loss of HLA-A2 melanoma cell surface expression is

associated with a complex rearrangement of the short arm of [190].

Transcriptional dysregulation of MHC class I gene is another frequent cause of

MHC class I down-regulation in cancer. The human CCAAT displacement protein

CDP/cut has been shown as a locus-specific repressor of MHC class I gene transcription in human tumor cells [191]. Lack of transcription factor binding, such as NFκB, to the class I enhancer sequence has been observed in tumor cells [192] and in tumorigenic adenovirus transformed cells [193]. In murine insulinoma model, the decreased expression of MHC class I antigen was found controlled by the R1 element of the H-2 enhancer [194]. Reduced co-expression of the c-fos and c-jun proto-oncogenes has been correlated with the down regulation of class I MHC in high-metastatic cell lines derived

from the Lewis lung carcinoma, B16 melanoma and the K1735 melanoma [195].

Overexpression of certain oncogenes, such as n-myc and c-myc, has been shown to down- regulate MHC class I expression on the transcription level [196, 197]. Class I MHC expression can also be suppressed by cytokine IL-10, which act through NKκB inhibition

in the short term and may repress TAP-1/2 transcription in the long term [198-200].

Hypermethylation is another cause of transcription inactivation, which has been found in

the class I MHC gene region in both human [201, 202] and murine tumor cells [203,

204].

22

Post-transcriptional defects also contribute to loss of class I expression in tumor cells, reports about which, however, have been limited. Aberrant pre-mRNA splicing causes selective HLA-A2 loss and β2m deficiency in melanoma cells [205, 206].

Premature stop codons resulting from nonsense mutation or frame-shifting insertion or deletion mutation have been shown to cause allelic loss of MHC class I expression in cervical cancer [207].

Due to its essential role in the MHC class I antigen processing and presentation,

deficiency of TAP may cause loss of MHC class I surface expression in tumor cells

[208]. In fact, loss of TAP expression has been frequently observed in a variety of

cancers [208]. In melanoma and cervical cancer, TAP-1 expression in primary lesions

may represent an independent prognostic marker [209, 210]. Presentation of specific

tumor associated antigenic epitope has been found limited by the level of TAP expression

in tumor cells [211]. In mouse model, TAP-deficient tumor cells show higher

tumorigenicity compared to the wild type cells in the immune competent host but not in

the athymic nude mice [212]. However, the underlying molecular mechanisms of TAP

deficiency in tumor cells have been limited. In a renal cell carcinoma cell line, defect in

the earliest steps of the IFN-γ signaling pathway causes loss of IFN-γ inducibility of

TAP-1 and LMP-2 genes [213]. Expression of a defective TAP-1 protein variant

(R659Q) results in loss of MHC class I surface expression in a human small cell lung

cancer (SCLC) cell line [214].

23

In search for new mechanisms that contribute to the down-regulation of MHC class

I expression in human cancers, we started this work by analyzing class I-deficient tumor cell lines. A melanoma cell line, SK-MEL-19, which had been shown lack surface expression of MHC class I, was found deficient in TAP-1 mRNA expression even after

IFN-γ treatment. A single nucleotide deletion (D1489) was then identified in the exon 7, which causes accelerated degradation of TAP-1 mRNA. To our knowledge, this is the first example of down-regulation of MHC class I antigen presentation machinery by decreasing mRNA stability.

Considering that although deficiency in TAP expression is a common observation in human cancers, TAP transporter activity can be disrupted without being downregulated on the expression level., we then looked for mutations that abrogate TAP protein function and may contribute to MHC class I down-regultion in human cancers. Since the signature motif has been known as a mutation “hot spot” in the CFTR protein of CF patients, we introduced the analogous CF-mutations into TAP-1 protein and characterized the function of these mutants in the SK-MEL-19 cells. Though all the CF-mutations abrogate TAP function, they are not present in the 228 human cancer samples we screened, 66 of which have down-regulated MHC class I expression. However, a rare allele of TAP-1, R648Q

(G1943A in DNA sequence), which harbors a replacement just C-terminal to the signature motif, occurs at a significantly higher frequency in the class I MHC- colon

cancer samples than those with normal class I expression level. The functional analysis

showed that the Q648 TAP-1 has only about 20% transporter activity of the R648 TAP-1.

24

Above all, we identified two mechanisms in this work that have not been addressed before, which down-regulate the TAP-1 mRNA level and disrupt TAP-1 protein function, respectively, and as a result cause the loss of MHC class I expression in human cancer cells.

25

Figure 1.1 Conventional MHC class I antigen processing and presentation pathway.

The antigenic peptides presented by class I MHC molecules are mostly derived from the

endogenous cellular proteins or viral proteins, which are degraded by the cytosolic

proteasome. TAP transporters on the ER membrane are responsible for transporting the

antigenic peptides from cytosol to the ER lumen, where they associate with the empty

MHC class I HC-β2m heterodimers with the help of the “peptide-loading complex” that consists of tapasin, calreticulin and ERp57.

26

~

~

Golgi

~ ER Proteasome ~ ~ ~ ~ ~ Cellular or peptides~ viral protein

Nucleus

MHC class I HC calnexin

β2m ERp57

tapasin ER aminopeptidase

calreticulin TAP1/2

27

MHC Class II MHC Class I

DP DN DO DQ2 DQ1 DR B C

LMP2 TAP1 LMP7 TAP2 DO DQ2

-594 +1 -427 ICS-1/2 GAS kB SP1-93-90-86

Figure 1.2 Genomic localization and promoter structure of human TAP-1 gene.

The genes that encode for TAP-1/2 and LMP-2/7 are clustered in the MHC class II locus

on the short arm of human chromosome 6. TAP-1 and LMP-2 share a bi-directional promoter of 593 nucleotides, which harbors cis elements including the IFN consensus sequence (ICS-1/2), a gamma activated sequence (GAS), an NFκB binding site and an

Sp1 site. The four transcription start sites for TAP-1 gene are shown as arrowheads with

position labeled accordingly (the first ATG codon for TAP-1 as +1).

28

Walker A

(529) 529 540 550 560 570 580 590 607 TAP1(493) LLTPLHLEGLVQFQDVSFAYPNRPDVLVLQGLTFTLRPGEVTALVGPNGSGKSTVAALLQNLYQPTGGQLLLDGKPLPQ TAP2(458) TLAPTTLQGVVKFQDVSFAYPNRPDRPVLKGLTFTLRPGEVTALVGPNGSGKSTVAALLQNLYQPTGGQVLLDEKPISQ CFTR-N(413) KQNNNNRKTSNGDDSLFFSNFSLLGTPVLKDINFKIERGQLLAVAGSTGAGKTSLLMMIMGELEPSEGKIKHSGR---- CFTR-C(500) KDDIWPSGGQMTVKDLTAKYTEG-GNAILENISFSISPGQRVGLLGRTGSGKSTLLSAFLRLLN-TEGEIQIDGVSWDS MDR1-N(382) GHKPDNIKGNLEFRNVHFSYPSRKEVKILKGLNLKVQSGQTVALVGNSGCGKSTTVQLMQRLYDPTEGMVSVDGQDIRT MDR1-C(365) GLMPNTLEGNVTFGEVVFNYPTRPDIPVLQGLSLEVKKGQTLALVGSSGCGKSTVVQLLERFYDPLAGKVLLDGKEIKR Consensus(529) L P L GNV F DVSFAYPSRPDVPVLKGLSFTIRPGQVLALVG SGSGKSTVLALLQRLYQPTEG VLLDGK I

Q loop C loop/signature motif (608) 608 620 630 640 650 660 670 686 TAP1(572) YEHRYLHRQVAAVGQEPQVFGRSLQENIAYGLTQK-PTMEEITAAAVKSGAHSFISGLPQGYDTEVDEAGSQLSGGQRQ TAP2(537) YEHCYLHSQVVSVGQEPVLFSGSVRNNIAYGLQS--CEDDKVMAAAQAAHADDFIQEMEHGIYTDVGEKGSQLAAGQKQ CFTR-N(488) ------ISFCS-QFSWIMPGTIKENIIFGVSYD---EYRYRSVIKACQLEEDISKFAEKDNIVLGEGGITLSGGQRA CFTR-C(577) ITLQQWRKAFGVIPQKVFIFSGTFRKNLDPYEQWS---DQEIWKVADEVGLRSVIEQFPGKLDFVLVDGGCVLSHGHKQ MDR1-N(461) INVRFLREIIGVVSQEPVLFATTIAENIRYGREN--VTMDEIEKAVKEANAYDFIMKLPHKFDTLVGERGAQLSGGQKQ MDR1-C(444) LNVQWLRAHLGIVSQEPILFDCSIAENIAYGDNSRVVSQEEIVRAAKEANIHAFIESLPNKYSTKVGDKGTQLSGGQKQ Consensus(608) I V YLR VGVV QEPVLFSGSIRENIAYGLQ TDDEI KAAKEANAHSFI LP KYDTVVGEKGSQLSGGQKQ

Walker B D loop Switch

(687) 687 700 710 720 730 740 750 765 TAP1(650) AVALARALIRKPCVLILDDATSALDANSQLQVEQLLYESPERYSRSVLLITQHLSLVEQADHILFLEGGAIREGGTHQQ TAP2(614) RLAIARALVRDPRVLILDEATSALDVQCEQAKTLWKFMIF------CFTR-N(555) RISLARAVYKDADLYLLDSPFGYLDVLTEKEIFESCVCKLMA-NKTRILVTSKMEHLKKADKILILHEGSSYFYGTFSE CFTR-C(653) LMCLARSVLSKAKILLLDEPSAHLDPVTYQIIRRTLKQAFA--DCTVILCEHRIEAMLECQQFLVIEENKVRQYDSIQK MDR1-N(538) RIAIARALVRNPKILLLDEATSALDTESEAVVQVALDKARK--GRTTIVIAHRLSTVRNADVIAGFDDGVIVEKGNHDE MDR1-C(523) RIAIARALVRQPHILLLDEATSALDTESEKVVQEALDKARE--GRTCIVIAHRLSTIQNADLIVVFQNGRVKEHGTHQQ Consensus(687) RIAIARALVR PKILLLDEATSALD SE VV AL A RT ILI HRLS V NAD ILVLEEG IREYGTHQ

Figure 1.3 Alignment of the nucleotide-binding domains (NBDs) of ABC transporters.

The NBDs of selective ABC transporters, including human CFTR, human MDR1,

TAP-1/2 and bacterial HisP, are compared. The conserved sequence motifs, Walker A,

Walker B, C loop/signature motif, D loop and switch region, are marked with rectangles.

29

CHAPTER 2

MATERIALS AND METHODS

2.1 Cell lines

Human melanoma cell lines 1195, 1102, and SK-MEL-19 were cultured as

described previously [215]. The breast cancer cell line SK-BR-3 was obtained from

ATCC (HTB-30; ATCC, Manassas, VA). All cells were grown in Dulbecco's modified

Eagle's medium (DMEM) supplemented with 10% heat-inactivated fetal calf serum, 2

mM L-glutamine, 100IU/ml penicillin, and 100µg/ml streptomycin (Invitrogen Inc.

Carlsbad, CA). For induction of HLA class I expression, cells were cultured in medium supplemented with recombinant human IFN-γ (R&D Systems, Minneapolis, MN) at 1000 units/ml.

2.2 Antibodies.

The following antibodies were used: R.RING4C, a rabbit anti-peptide antibody to

the C-terminus of TAP-1 (a gift from Dr. Peter Cresswell); HC10, a mAb recognizing

free class I heavy chain; W6/32, a β2m-dependent anti-class I heavy chain mAb;

R.gp48N, a rabbit anti-peptide Ab to the N-terminal regions of tapasin (a gift from Dr.

30

Peter Cresswell); a rabbit anti-human TAP-1 peptide (735-748) Ab (Calbiochem-

Novabiochem corporation); an anti-human calnexin mAb (Affinity Bioreagents, Golden,

CO); an anti-human TAP-2 mAb (BD PharMingen, San Diego, CA).

2.3 Flow cytometry

Cell surface HLA class I expression was examined by flow cytometry as described

previously [174]. Briefly, viable cells were incubated with PE-conjugated mouse IgG1

(BD Pharmingen), as isotype control, and PE-conjugated anti-HLA-A, -B, and -C

antibody (BD Pharmingen), respectively, at 4 °C for 2 h. After three washes with

phosphate-buffered saline (PBS) containing 1% fetal calf serum, cells were fixed with 1%

paraformaldehyde and examined by flow cytometry.

2.4 Northern blot

Cells were either treated with IFN-γ (R&D Systems) at 1000 units/ml for 48 h or left untreated. For cycloheximide (CHX; Sigma-Aldrich Corp. St. Louis, MO) treatment,

SK-MEL-19 cells were cultured with IFN-γ at 1000 units/ml for 48 h, and then CHX was

added to the cells for a final concentration of either 5 or 10 µg/ml, respectively, for up to

16 h. Total RNA was isolated using TRIzol reagent (Invitrogen). Hybridization conditions

followed the instructions of the Northern hybridization kit (Eppendorf Scientific,

Westbury, NY). The cDNA probes for TAP-1, TAP-2, LMP-2, LMP-7, HLA class I heavy

chain, and β2M were made from PCR products using primers listed previously [161]. A human splenocyte cDNA library from Invitrogen was used as the template for PCR

31

reactions. All PCR products were subcloned into pBluescript vector (Stratagene, La Jolla,

CA) and sequenced and confirmed to be identical to published sequences. The probes were labeled with [α-32P]dCTP (PerkinElmer Life Sciences, Boston MA) using the

DECAprimeTM II kit (Ambion, Austin, TX).

2.5 Southern blot

Genomic DNA was isolated from SK-MEL-19 cells, SK-BR-3 cells, and HeLa cells. Genomic DNA (20 µg) was digested with Afl III (Invitrogen) and separated in 0.8%

agarose gel. The TAP-1 promoter probe was made by PCR from normal human

lymphocyte genomic DNA with sense primer 5'-

TCCCGCCTCGAGCATCCCTGCAAGGCA-3' and antisense primer 5'-

TGCAGTAGCCTGGTGCTATCCG-3'. Probes were labeled with [α-32P]dCTP

(PerkinElmer Life Sciences) using the DECAprimeTM II kit (Ambion).

2.6 Generation of TAP-1 cDNA constructs and stable transfection

The human small cell carcinoma H146 cell line (provided by Dr. N. P. Restifo;

National Cancer Institute, Bethesda, MD) was incubated with IFN-γ at 1000 units/ml for

48 h. Total RNA was isolated as described above. Reverse transcription was done using

the SUPERSCRIPT First-Strand cDNA Synthesis System (Invitrogen). TAP-1 cDNA was

amplified by PCR in three fragments. Primers were as follows: hTAP.f1, 5'-

GCGGCCGCTTTCGATTTCGCTTTC-3'; hTAP.r1, 5'-

TGCAGTAGCCTGGTGCTATCCG-3'; hTAP.f2, 5'-

32

CTTGCCTTGTTCCGAGAGCTGA-3'; hTAP.r2, 5'-CTCGTTGGCAAAGCTTCGAAC-

3'; hTAP.f3, 5'-CGGCCATGCCTACAGTTCGAAG-3'; and hTAP.r3, 5'-

ATAAATATCAAGAACCTACAGGG-3'. The three fragments were cloned into pBS-KS vector (Stratagene) at NotI/SmaI, SmaI/HindIII, and HindIII/XhoI sites, respectively, and

sequenced to confirm that the cDNA has a wild type sequence. When SK-MEL-19 cells grew to 70% confluence, 0.2 µg of pcDNA3.1/Hyg(+) vector (Invitrogen) and pcDNA3.1/Hyg(+) vector with TAP-1 cDNA insert were respectively transfected into

each well of a 24-well plate, using 6 µl of FuGENE 6 transfection reagent (Roche

Diagnostics Corporation, Indianapolis, IN) according to the manual. 48 h later, the

transfected cells were replated into 96-well plates and cultured in the presence of

0.5mg/ml hygromycin B (Invitrogen). Single cell clones were selected for further culture

and analyzed for HLA class I antigen expression.

Constructs for all the TAP-1 signature motif mutants and TAP-1 R648Q were

made by overlapping PCR and expressed in the pcDNA3.1/Hyg(+) vector (Invitrogen).

Nucleotide changes and amino acid changes are summarized in Table 2.1. To ensure that

any functional defects of the TAP-1 mutants were due to the mutations we generated, all

the constructs were sequenced before further analyses. Stable cell clones were established

as described above. Expression of TAP-1 protein was examined by Western blotting

using a rabbit anti-human TAP-1 peptide (735-748) Ab (Calbiochem, EMD Biosciences

Inc., San Diego, CA). Cell clones expressing TAP-1 protein were selected for flow

cytometric analysis of cell surface MHC class I.

33

2.7 Generation of TAP-1 constructs and Tet-Off SK-MEL-19 cell lines for RNA

stability analyses

TAP-1 D1489 was generated by PCR using the total cDNA from the SK-MEL-19

cells as the template. Site-directed mutagenesis by overlapping PCR was performed to

make TAP-1 Del3 cDNA. After confirming their sequences, all three TAP-1 cDNAs were

inserted into the multiple cloning site of the pBI-EGFP vector (BD Bioscience, Clontech,

Palo Alto, CA) by blunt-end ligation. The three constructs, pBI-EGFP-TAP-1, pTet-Off

(Clontech), and pcDNA3.1/Hyg(+) (Invitrogen), were co-transfected into SK-MEL-19

cells using FuGENE 6 transfection reagent (Roche). Stably transfected cell clones were

selected in 96-well plates in the ordinary DMEM culture medium supplemented with

0.5mg/ml hygromycin (Invitrogen). To confirm the efficiency of the Tet-Off construct in the tumor cell line, the green fluorescence protein-positive cell clones were treated with 1

µg/ml tetracycline (Roche) for 24 h. The total RNA was extracted from the cell clones

with or without tetracycline treatment. The cell clones in which TAP-1 mRNA expression was inhibited by at least 95% by tetracycline were used for the study of mRNA stability.

2.8 Generation of luciferase reporter constructs and assay for promoter activity

The TAP-1 promoter was amplified from SK-MEL-19 cell genomic DNA by PCR using the following primers: hTAP-1.Pr1, 5'-

GCTCTAGATGGCACTCGGACGCCGTC-3'; and hLMP2.Pf1, 5'-

GCTCTAGACCCTGCAAGGCACCGCTC-3'. The PCR products were subcloned using

the Zero Blunt TOPO PCR cloning kit (Invitrogen) and then cloned into pGL2-basic

34

vector (Promega, Madison, WI) at XhoI and HindIII sites. All constructs were confirmed

by DNA sequencing. Expression level of the firefly luciferase from the pGL2 constructs

(basic, SV40, pTAP-1/T, and pTAP-1/G) was normalized to the internal control pRL-

SV40 Renilla luciferase level. Results were shown as the fold increase compared with the

pGL2-basic. The dual luciferase assay was carried out according to the manufacturer's

instructions (Promega).

2.9 Nuclear run-on assay

The assay was performed as described elsewhere [216]. Briefly, nuclei were

extracted from 107 to 108 SK-MEL-19 cells treated with or without 1000 units/ml IFN-γ.

The transcripts were labeled in vitro with 40 nM biotin-16-UTP (Roche) in the presence

of 3.75 mM ATP, GTP, and CTP; 25 mM Tris-HCl; 12.5 mM MgCl2; and 750 mM KCl. cDNA fragments of LMP-2 and GAPDH were amplified by PCR from cloned cDNA constructs [161]. TAP-1 cDNA fragment was amplified by PCR from cloned cDNA constructs using primers hTAP-1.f1 and hTAP-1.r1 described above. The

pcDNA3.1/Hyg(+) vector was linearized with HindIII. All the DNA was immobilized on

nitrocellulose membrane using S&S Minifold II slot blot apparatus according to the

manual (Schleicher & Schüll Bioscience Inc., Keene NH). Hybridization conditions were

as described previously [216], and the biotin-labeled transcripts were detected using

streptavidin-alkaline phosphatase conjugate (Roche) and CDP-star Ready-To-Use with

Nitro-Block-II reagent (Tropix, Bedford, MA).

35

2.10 Restriction fragment length polymorphism (RFLP)

To detect the D1489 mutation in TAP-1 exon 7, primers hTAP-1CE7.f (5'-

GCACCCCTCGCTGCCTACCCAGTGGTCT-3') and hTAP-1CE7.r (5'-

TACAGGGAGTGGTAGGTTGTACCTG-3') were used to amplify from genomic DNA

the region of TAP-1 exon 7 where the single-nucleotide deletion resides. The region was also amplified from cDNA using primers hTAP-1CE7.f and hTAP-1CE7.r PCR products

were separated by gel electrophoresis and purified using Qiagen gel extraction kit (Qiagen

Inc., Valencia, CA). The purified PCR products were incubated with BslI (New England

Biolabs Inc., Beverly, MA) at 55 °C overnight and then separated in 5% agarose gel.

To screen the G1943A polymorphism in normal population, a forward primer, E10

Ahd.f 5’-ACCGTTCTCATCTTGGCCCTTTGCTCTG-3’, was designed to generate an

Ahd I site in the G allele but not in the A allele. Together, with the reverse primer, E10

Ahd.r 5’- ATCAATGCTCGGGCCAACGCGACTGCCT-3’, TAP-1 exon 10 was amplified from the genomic DNA. PCR products were then separated by gel electrophoresis and purified using QIAGEN gel extraction kit (Qiagen). The purified

PCR products were incubated with Ahd I (New England Biolabs) at 37oC overnight and

then separated in 5% agarose gel.

2.11 RNase protection assay

The TAP-1 cDNA construct used to make the TAP-1 antisense transcript was the

same as the one used for the Northern blot. The GAPDH cDNA fragment was generated

by PCR using the primers hGAPDH.f (5'-TGAGAACGGGAAGCTTGTCATCAA-3')

36

and hGAPDH.r (5'-CAGCCTTCTAGATGGTGGTGAAGA-3'). The EGFP cDNA

fragment was also generated by PCR using the primers EGFP.f (5'-

TCCAGCAGGATCCTGTGATCGCGCT-3') and EGFP.r (5'-

ACCTACGGCCTCGAGTGCTTCAGCC-3'). The antisense probes were made using the

Riboprobe in vitro transcription system (Promega). The RNase protection assay was conducted with the RPA III ribonuclease protection assay kit (Ambion) according to the instruction manual. After separation of protected fragments on a 6% sequencing gel,

signals were quantified by phosphorimaging (Amersham Biosciences Inc., Piscataway,

NJ). The TAP-1 signal intensity was normalized by the signal intensity of GAPDH, which

served as a loading control. The percentages of the amount of remaining TAP-1 mRNA at

different time points after tetracycline was added compared with time 0 were calculated.

2.12 Immunofluorescence microscopy

Cells were plated on coverslips in a 24-well plate and were allowed to grow

overnight. Cells were fixed for 15 min in a freshly prepared solution of 4% formaldehyde

in PBS, pH 7.4, and then permeabilized in washing buffer (PBS with 0.2% Triton X-100)

for 5 minutes at room temperature. After three washes, the primary antibodies were

added, rabbit anti-human TAP-1 (Calbiochem) and mouse anti-human calnexin (Affinity

Bioreagents, Golden, CO) or anti-human TAP-2 mAb (BD Pharmingen), which were

diluted in 200 µl of dilution buffer (PBS containing 3% bovine serum and 0.2% Triton X-

100). After incubation for 1 hour at room temperature, coverslips were washed three times in washing buffer and then incubated with the secondary antibodies, fluorescein-

37

conjugated donkey anti-rabbit Ig (Amersham) for TAP-1 and Texas red dye-linked goat anti-mouse Ig (Jackson ImmunoResearch Laboratories Inc., West Grove, PA) for calnexin or TAP-2, for 1 hour at room temperature. After three washes, the coverslips were dipped in water and then mounted on glass slides in the mounting media with DAPI

(Vector Laboratories, Burlingame, CA).

2.13 Immunoprecipitation

Immunoprecipitation procedures described [42] were followed. Cells were extracted for 30 min on ice at 2 × 106 cells per ml in 1% digitonin (Sigma) in TBS (10 mM Tris, 150 mM NaCl [pH 7.4]) containing 0.5 mM phenylmethyl sulforyl fluoride

(Sigma), 0.1 mM N-tosyl-L-lysine chloromethylketone (Sigma), and 5 mM iodoacetamide (Sigma). Postnuclear supernatant was precleared overnight at 4°C with 5

µl normal rabbit serum (Invitrogen) and 50 µl Zysorbin (Zymed Laboratories, Inc., San

Francisco, CA) per ml of extract. Aliquots were then incubated with the appropriate antibodies, R.RING4C, a rabbit anti-peptide antibody to the C-terminus of TAP-1 (a gift from Dr. Peter Cresswell), W6/32, a mAb detecting the β2m-associated class I heavy chain and an anti-human TAP-2 mAb (BD PharMingen), and protein A–Sepharose for 1 hr each at 4°C. Protein G–Sepharose (Amersham) was used with anti-TAP-2 mAb and

W6/32. The immunoprecipitates were washed twice with TBS, 0.1% digitonin and once with distilled water before being separated by SDS–PAGE. TAP-1, TAP-2 and tapasin were detected by Western blotting using the rabbit anti-human TAP-1 peptide (735-748)

38

Ab (Calbiochem), the anti-human TAP-2 mAb (BD PharMingen), and R.gp48N, a polyclonal rabbit Ab against the N-terminal regions of tapasin (a gift from Dr. Peter

Cresswell). The same antibodies were used for the following Western blotting analyses.

2.14 Microsome-based peptide translocation assay

Microsomes were prepared as described previously [108]. Briefly, 1 × 108 SK-

MEL-19 cell clones that were stably transfected with wild-type TAP-1, TAP-1 G646D,

TAP-1 R648Q and pcDNA3.1/Hyg vector, respectively, were harvested. Cells were washed once in ice-cold PBS, and then resuspended in 800 µl of ice-cold cavitation buffer (250 mM sucrose, 25 mM potassium acetate, 5 mM magnesium acetate, 0.5 mM calcium acetate, 50 mM Tris, pH 7.4) supplemented with a inhibitor mixture

(Sigma). Cells were lysed by repeatedly drawing the suspension through a 26-gauge needle at least 10 times. After centrifugation at ~500 × g for 5 min, the supernatant was thoroughly mixed with 5 ml of 2.5 M sucrose in gradient buffer (2.5 M sucrose, 150 mM potassium acetate, 5 mM magnesium acetate, 50 mM Tris, pH 7.4), which was then overlaid with 3 ml each of 2.0 M sucrose in gradient buffer, then 1.3 M sucrose in gradient buffer, and finally with 800 µl of cavitation buffer. This was centrifuged overnight at 80,000 × g, and the microsomal fraction at the interface of the 2.0 and 1.3 M sucrose layers was collected. This fraction was diluted into 5 ml of PBS, 1 mM dithiothreitol, and centrifuged for 1 h at 100,000 × g. The pellet was resuspended in

39

100µl of PBS, 1 mM dithiothreitol and frozen in aliquots at -70 °C. The total protein

content was determined by a BCA assay (Pierce Biotechnology Inc., Rockford, IL). The

expressions of TAP-1 and TAP-2 proteins were examined by Western blotting.

Translocation assays were carried out as previously described [108]. An iodinated

model peptide, RRYNASTEL, with a specific activity of 100-150 µCi/µg was used for

the study. Briefly, 30 µg of microsomes containing wild-type TAP-1, TAP-1 G646D,

TAP-1 R648Q or pcDNA3.1/Hyg vector were added to 150 µl of assay buffer (PBS,

0.1% bovine serum , 1 mM dithiothreitol, pH 7.3) containing 10 mM MgCl2 and

5 mM ATP (+ATP samples) and then incubated with the radioiodinated RRYNASTEL

peptide at 37 °C for 15 min. The samples were then centrifuged at 4 °C at 8,800 × g and washed once with 250 µl of the assay buffer. The pellets were subsequently resuspended

in 250 µl of lysis buffer (50 mM Tris, 150 mM NaCl, 1% Nonidet P-40, pH 7.4) and

incubated on ice for 1 h. After centrifugation at 8,800 × g for 5 min at 4 °C, the

supernatants were transferred to ConA-Sepharose (Amersham) beads and incubated for 2 h at 4 °C with agitation. Beads were washed twice with the assay buffer before the

radioactivity was determined using a Beckman 5500 gamma-counter.

2.15 Colon cancer tissue sections, immunohistochemistry and statistical analysis

Colon cancer tissue microarray slides were acquired from National Cancer

Institute (NCI) and Zymed Laboratories Incorporation. All specimens contained in the

array were formalin-fixed and paraffin-embedded. Mouse monoclonal antibody HC-10

that detects HLA-A,B,C heavy chain was used in immunohistochemical staining. Tissue

40

samples were graded as MHC class I negative when at least 75% of the tumor cells were

not stained or weakly stained. The comparisons of the allelic frequencies were tested

using two-tailed z-tests and an alpha level of 0.05. The p values are reported.

2.16 Isolation of genomic DNA and genotyping

The genomic DNA from blood of normal population was extracted as described

elsewhere [217] and then subjected to the restriction fragment length polymorphism

(RFLP) analysis as described above. Tumor tissues were collected either by laser capture

microdissection (LCM) or manually using sterile tweezers. DNA was then isolated using

a lysis buffer containing 1×high fidelity PCR buffer (Roche Diagnostics Corporation,

Indianapolis, IN), 1% tween-20 and 4mg/ml proteinase K and incubated at 55 oC for 72 hours. Nested PCR was designed using the following primers to amplify the exon 10 region of TAP-1 gene: ouE10.F 5’-GTTCTCATCTTGGCCCTTTGCTCTG-3’ and inE10.R1: 5’- AGAAGATGACTGCCTCACCTGTAAC-3’ for the first-round amplification; E10.F 5’-CCTTTGCTCTGCAGAGGTAGACGAG-3’ and inE10.R2 5’-

TGCCTCACCTGTAACTGGCTGTTTG -3’ for the second round. The final PCR products were purified with QIAGEN PCR purification kit (Qiagen) before sequencing.

41

CHAPTER 3

A SINGLE-NUCLEOTIDE DELETION LEADS TO RAPID DEGRADATION OF TAP-1 MRNA IN A MELANOMA CELL LINE

3.1 Introduction.

Recent studies demonstrate that patients with malignant melanoma often have a

high number of cytotoxic T lymphocytes specific for melanoma-associated antigens [218,

219]. The co-existence of T cells and tumor cells even in the draining lymph nodes

suggests that the tumors were able to evade destruction by host cytolytic T lymphocytes.

Accumulating evidence supports the notion that both malfunction of T cells and down- regulation of antigen presentation machinery in tumors can be responsible for tumor

evasion of host immunity [148, 161, 174, 218-220]. In fact, a high proportion of

malignant tumors, including melanoma, have severely depressed cell surface expression

of class I HLA antigens [221], the target molecules that present tumor antigenic peptide to cytolytic T lymphocytes. Understanding the mechanisms underlying the T-cell

malfunction or antigen presentation defects may thus provide insight for immunotherapy

of melanoma and other cancers.

42

Optimal cell surface expression of HLA molecules requires the coordinated expression of several genes, such as transporters associated with antigen processing

(TAP)-1/2, low molecular weight peptide (LMP)-2/7, and tapasin, as well as HLA class I

heavy chain and β2-microglobulin (β2M). In cases of both tumorigenesis and viral infection, expression of these genes and the function of the encoded proteins are often impaired [222-224]. The mechanisms for such down-regulation have been studied extensively. Theoretically, gene expression can be modulated by transcriptional, posttranscriptional, translational, and posttranslational mechanisms. The mechanisms that have been shown to underlie the antigen presentation abnormalities are transcriptional suppression of antigen presentation genes and/or functional inactivation of their gene products, either by missense mutation or by protein-protein interactions [200, 213, 214,

225]. Here we show that actively transcribed TAP-1 mRNA in the melanoma cell line

SK-MEL-19 is rapidly degraded even after stimulation with IFN-γ. Cloning and sequencing analysis have revealed a single-nucleotide deletion at position +1489. This mutation results in substantial reduction of the stability of TAP-1 mRNA by mechanisms unrelated to nonsense-mediated mRNA decay (NMD). These results reveal a new potential mechanism for tumor evasion of host T-cell recognition.

43

3.2 Results and discussion.

3.2.1 Down-regulation of TAP-1 mRNA by a posttranscriptional mechanism in

melanoma cell line SK-MEL-19

Three human melanoma cell lines (1102, 1195, and SK-MEL-19) were examined by

flow cytometry for their cell surface HLA class I expression with or without IFN-γ stimulation. A PE-conjugated anti-human HLA-A, -B, and -C antibody was used to detect all HLA class I alleles, and a PE-conjugated mouse IgG1 was used as isotype control. As

shown in Figure 3.1a, 1102 and 1195 cells had significant HLA class I that was further

up-regulated by incubation with 1000 units/ml IFN-γ for 3 days. Confirming previous

studies[215], we found that SK-MEL-19 cells had no cell surface HLA. Surprisingly,

whereas other melanoma cell lines up-regulated their cell surface HLA in response to

IFN-γ, very little HLA class I antigen could be found on the SK-MEL-19 cells even after

IFN-γ-treatment.

Because optimal cell surface HLA class I expression requires the coordinated

expression of multiple genes, including TAP-1/2, LMP-2/7, and β2M as well as HLA

class I heavy chain, a Northern blot analysis was performed to detect the expression of

these genes (Figure 3.1b). In 1102 and 1195 cells, all six genes were expressed at low but

detectable levels. IFN-γ treatment drastically induced expression of all six genes.

Interestingly, in the SK-MEL-19 cells, whereas β2M, HLA heavy chain, LMP-2, LMP-7,

44

and TAP-2 were present at low levels without induction, no TAP-1 mRNA was detected.

After IFN-γ treatment, β2M, HLA heavy chain, TAP-2, LMP-2, and LMP-7 were

expressed at high levels, yet TAP-1 was still expressed at low levels.

It has been known that TAP-deficient cells can express HLA class I after

transfection with the TAP-1 or TAP-2 gene [79, 226, 227]. To test whether the lack of

TAP-1 expression was responsible for the barely detectable expression of HLA class I

antigen on the surface of SK-MEL-19 cells, we transfected the cells with TAP-1 cDNA.

As shown in Figure 3.1c, the TAP-1 cDNA-transfected SK-MEL-19 cells expressed

significant levels of HLA class I antigen even before IFN-γ-treatment. Moreover, the

TAP-1 transfectants were as responsive to IFN-γ as the other melanoma cell lines. Based

on these results, it is likely that the primary defect of antigen presentation in SK-MEL-19 cells is attributable to defects in TAP-1 expression.

Because the TAP-1 expression was low at the mRNA level, we hypothesized that

the TAP-1 down-regulation was caused by defective transcription or malfunction in RNA

metabolism. The TAP-1 expression is under the control of a bidirectional promoter, as

characterized by Wright et al. [86]. We cloned and sequenced the 593-bp TAP-1 promoter

from SK-MEL-19 cells. In comparison with the published sequence [86], a single-

nucleotide G ->T replacement was identified at position -446 (the first ATG of the TAP-1

gene is designated as +1), which is close to the first transcription start site at -427 [86]

(Figure 3.2a). Because the T allele results in a loss of restriction site AflIII, we did a

Southern blot hybridization using AflIII to confirm the mismatch. As shown in Figure

3.2a, whereas the HeLa cell line contained homozygous G alleles as described previously

45

[86], both SK-MEL-19 and the breast cancer cell line SK-BR-3 were homozygous for T

alleles that lack the restriction site for AflIII. To test whether this single-nucleotide replacement results in reduced promoter activity, both alleles of the TAP-1 promoter were

cloned into the pGL2-basic vector that contains the luciferase reporter gene. As shown in

Figure 3.2b, the T allele TAP-1 promoter retained 50% of the promoter activity compared

with the G allele. However, given the significant variation in transient transfection and

luciferase assays, it is unclear whether the G ->T change has a significant effect on TAP-1

transcription. However, both promoters were equally efficiently induced by IFN-γ

treatment, whereas the TAP-1 mRNA in the original SK-MEL-19 cell line was not

induced by IFN-γ treatment (Figure 1.1b). Moreover, our analysis of normal human

peripheral blood lymphocyte samples revealed that both alleles were present at a high frequency, and individuals that carry either G or T alleles have equivalent cell surface

HLA class I antigen expression (data not shown). We therefore performed a nuclear run-

on assay to directly evaluate the transcription of the TAP-1 gene. LMP-2 transcription,

which is under the control of the same bidirectional promoter, was also evaluated. As shown in Figure 3.2c, TAP-1 was transcribed at high levels in SK-MEL-19 cells under

basal conditions, although IFN-γ appeared to up-regulate TAP-1 transcription somewhat.

In contrast, LMP-2 was transcribed at an undetectable level but was induced to high levels by IFN-γ (Figure 3.2c). The lack of LMP-2 transcription at basal condition may reflect the

IFN-γ-inducible expression pattern of this gene. These results demonstrate that lack of

TAP-1 mRNA in SK-MEL-19 cells was not due to defective transcription. Taken together, the results demonstrate that a posttranscriptional defect is responsible for poor 46

TAP-1 expression in SK-MEL-19 cells even after IFN-γ stimulation. Numerous studies

have revealed defective TAP-1 expression among tumor cells [214, 221]. To our

knowledge, however, this is the first example of a posttranscriptional defect of TAP-1

expression.

3.2.2 A single-nucleotide deletion leads to accelerated decay of TAP-1 mRNA

A major mechanism responsible for posttranscriptional regulation of mRNA is

RNA degradation, which can be prevented by CHX, a protein synthesis inhibitor of

mammalian cells. It is well established that the turnover of mRNA is closely linked to the

translation process and that blocking translation can stabilize mRNA, especially those mRNA with short half-lives [228-230]. To test whether accelerated RNA degradation is

responsible for the lack of TAP-1 in SK-MEL-19 cells, we treated the SK-MEL-19 and

control HeLa cells with CHX after incubation with or without IFN-γ (1000 units/ml) for

48 h at 37 °C. At different time points after the CHX was added to the cell culture, cells

were harvested, and the total cellular RNA was analyzed for TAP-1 mRNA. The intensity

of each band was quantified using ImageQuant 5.0 software (Amersham Biosciences)

after exposure to a phosphorimaging screen (Figure 3.3). For a better comparison, TAP-1

mRNA levels were normalized to the endogenous housekeeping gene GAPDH level, and

the fold increase compared with non-CHX-treated cells was calculated. Under basal

conditions, the TAP-1 mRNA was up-regulated 4.8-fold in SK-MEL-19 cells. After IFN-γ induction, the TAP-1 mRNA was up-regulated 25.7-fold. In comparison, CHX caused a

less significant increase of TAP-1 mRNA in both IFN-γ-treated and untreated HeLa cells. 47

Taken together, the lack of TAP-1 mRNA, the normal transcription of TAP-1, and the

rescue of TAP-1 mRNA by CHX treatment suggest that the TAP-1 mRNA was rapidly

degraded in the SK-MEL-19 cells. It is noteworthy that in SK-MEL-19 cells, the effect of

CHX was significantly stronger when used in combination with IFN-γ. This finding

cannot be fully explained by the fact the IFN-γ is a transcriptional activator for antigen

presentation genes, because its effect on TAP-1 transcription is not so obvious in SK-

MEL-19 cells as shown in Figure 3.2c. It is likely that IFN-γ stabilized mRNA in SK-

MEL-19 cells, although this possibility remains to be tested formally.

The rapid degradation of TAP-1 mRNA can be due to a genetic lesion in the TAP-

1 gene. Alternatively, it is possible that the tumor cell line expresses factors that can cause TAP-1 mRNA degradation. The successful rescue of cell surface HLA class I antigen expression by wild type TAP-1 in SK-MEL-19 cells favors the first hypothesis

because a wild type cDNA can be expressed in the tumor cell line. As the first step to test

this hypothesis, we cloned TAP-1 cDNA from SK-MEL-19 cells that were treated with

both IFN-γ and CHX as described above. All of the three clones sequenced showed a

single-nucleotide deletion at position +1489 (Figure 3.4a), which resides in exon 7 in the

TAP-1 gene. Further analysis showed that multiple downstream premature termination

codons (the closest one is at position +1555) were present due to this nucleotide deletion

(two of them are shown in Figure 3.4c). To confirm that the mutation was in the TAP-1

gene, we amplified exon 7 of the TAP-1 gene from the SK-MEL-19 cells by PCR. The

PCR products were digested with BslI because this restriction recognizes the

deletion mutant but does not recognize the wild type exon 7. Because complete digestion 48

was obtained, it appears that the SK-MEL-19 cells are homozygous for the frameshift

mutation (Figure 3.4b), even though the cytogenetic analysis revealed that there are four

copies of chromosome 6 present in the SK-MEL-19 cells (data not shown). We subsequently amplified exon 7 of the TAP-1 gene from 50 normal human peripheral

lymphocyte genomic DNA samples by PCR and subjected the PCR products to BstI digestion. Because none of the PCR products from the 50 samples could be digested by

BslI, it is most likely that the single-nucleotide deletion in SK-MEL-19 cells resulted from

a somatic mutation (data not shown) and that the apparent homozygosity of the TAP-1

locus caused aneuploidy.

Premature termination codons resulting from frameshift mutation or nonsense

mutation have been shown to interfere with the metabolism of many different mRNAs in

mammalian cells, leading to nonsense-mediated altered RNA splicing, such as exon skipping and intron retention and/or NMD [228-230]. Alternatively, a mutation may disrupt a cis-element that is necessary for mRNA stability and thereby cause RNA decay.

To differentiate the two possible mechanisms, we designed another TAP-1 mutant (TAP-

1 Del3) that has an in-frame 3-nucleotide deletion at the same position as TAP-1 D1489

(Figure 3.5a) and compared the mRNA half-lives of TAP-1 WT, TAP-1 D1489, and

TAP-1 Del3. If the degradation of TAP-1 D1489 is via the NMD pathway, then the half-

life of the TAP-1 Del3 message should be comparable with that of TAP-1 WT.

Otherwise, if the message of TAP-1 Del3 is comparable with or even more unstable than

that of the TAP-1 D1489, then it is the deletion itself, but not the resulting premature

termination codons, that leads to the accelerated decay of TAP-1 mRNA in SK-MEL-19

49

cells. The Tet-Off gene expression system was adapted (Figure 3.5a) to compare the stabilities of various TAP-1 mRNAs. We first selected the SK-MEL-19 transfectants with an induction ratio (Tet 0/Tet 24 h) of 20 or greater to test mRNA half-lives. This ensured that more than 95% of transcription was blocked by the tetracycline after 24 h. The

selected cell clones were then treated with 1 µg/ml tetracycline for different lengths of

time before the total RNA was harvested. The amounts of the mRNA were quantified by

phosphorimaging. In addition, the intensity of the TAP-1 signal was normalized to that of

GAPDH before the percentage of remaining mRNA was calculated. Four experiments

were conducted, and an average percentage value was used to derive the mRNA half-

lives (t1/2) via regression analysis. As shown in Figure 3.5, b and c, with a t1/2 of 7.2 h, the

TAP-1 WT mRNA is considerably more stable than that of the TAP-1 D1489 (t1/2 = 3.5

h). However, the mRNA derived from TAP-1 Del3 (t1/2 = 2.7 h), which has an in-frame 3- nucleotide deletion, was degraded at least as fast as that of TAP-1 D1489. Because the

TAP-1 Del3 mRNA has no premature termination codon downstream to the deletion, the accelerated decay in mutant TAP-1 mRNA is most likely through mechanisms other than

NMD. It is more likely that the mutation disrupts a cis-element critical for the stability of

TAP-1 mRNA. Whereas few cis-elements that help to stabilize mRNA have been

reported, at least two have been reported by others [231, 232] and our group [233].

Preliminary analysis showed no similarity between the region surrounding the mutation

and the previously reported cis-element.

50

Nevertheless, the NMD is a well-conserved cellular surveillance mechanism.

Whereas our work revealed a non-NMD mechanism for TAP-1 mRNA degradation, it is

still possible that mRNA derived from the endogenous mutant TAP-1 gene can also be

degraded by NMD. It has been shown recently that several criteria have to be met for the

pathway to degrade a premature termination codon-containing message. First, at least one

downstream spliceable intron is required for optimal NMD [234-236]. The intron is

thought to help recruit NMD factors, such as hUpf3, to the mRNA via the spliceosome

[237-239]. Second, the premature termination codon should be at least 45-55 nucleotides away from the next spliceable intron [234, 235]. The lack of introns in our constructs may

have prevented us from revealing NMD in TAP-1 mRNA degradation. However,

intronless premature termination codon-containing HEXA mRNA was shown to be

subject to NMD, although at a lower efficiency than that seen when multiple downstream

introns are present [240]. In preliminary studies, when we made pBI-EGFP/TAP-1

constructs with intron 7 or 8, the results also failed to support a role for NMD in degradation of the mutated TAP-1 mRNA (data not shown).

Posttranscriptional regulations of other genes involved in antigen presentation have

been reported previously [241]. The increased turnover of HLA-C heavy chain mRNA

has been suggested to contribute to the low level of HLA-C surface expression [241]. Our

work shows that mutations in the TAP-1 gene in a tumor cell line can modulate its mRNA

stability. This mechanism may be exploited by tumors to evade host immunity.

51

Figure 3.1 Deficiency of surface HLA class I expression in melanoma cell line SK-

MEL-19 was due to the TAP-1 down-regulation.

a, HLA class I expression in three melanoma cell lines, SK-MEL-19, 1102, and

1195. Bold black lines depict the staining by PE-conjugated anti-human HLA-A, -B, and

-C antibody in untreated cells; dotted lines represent the staining by PE-conjugated

mouse IgG1 as isotype control; and red lines represent anti-HLA-A, -B, and -C antibody

staining after stimulation with 1000 units/ml IFN-γ for 72 h. b, expression of HLA class I

heavy chain (MHC I), β2M, TAP-1, TAP-2, LMP-2, and LMP-7 in each cell line with or without IFN-γ induction (1000 units/ml for 72 h). Total RNA loading to each well was shown as 28 S rRNA (28S) and 18 S rRNA (18S). c, transfection with wild type TAP-1, but not vector alone, restored HLA class I expression in the SK-MEL-19 cells. SK-MEL-

19 cells were transfected with either vector alone (top panels) or vector with TAP-1 cDNA insert (bottom panels). These stable clones from each group were stimulated with or without IFN-γ and analyzed for cell surface HLA-A, -B, and -C, as detailed in a.

52

Figure 3.1

a

SK-MEL-19 1102 1195 Ctrl -IFNγ +IFNγ Cell numbers Cell numbers log Fluorescence (PE-anti-human HLA-A, B, C)

b SK-MEL-19 1102 1195 - + - + - + IFNγ

β2M

MHC I LMP-2

LMP-7

TAP-1

TAP-2

28S 18S

(continued)

53

Figure 3.1 (continued)

c.

F2 D9 E6 Ctrl -IFNγ +IFNγ vector

5G5 5C6 3A2 Cell numbers

TAP-1

log Fluorescence (PE-anti-human HLA-A, B, C)

54

Figure 3.2 Posttranscriptional mechanisms are responsible for poor accumulation

of TAP-1 mRNA.

a, a single-nucleotide polymorphism, adjacent to the first transcription start site (-427), was identified at -446 in the bidirectional promoter shared by the TAP-1 and LMP-2

genes. The G-> T change results in the loss of the AflIII restriction site. Southern blot

hybridization was performed using AflIII and detected by a DNA probe that encompasses

the downstream region of the polymorphism site. SK-MEL-19 cells showed one 5.6-kb

band that represents homozygous T allele, as did the breast cancer cell line SK-BR-3,

which has significant cell surface HLA class I surface expression (data not shown). HeLa

cells, in contrast, are homozygous for the G allele. b, activities of T and G alleles of TAP-

1 promoter (pTAP1/T and pTAP1/G, respectively) in SK-MEL-19 cells. The two allelic

forms of TAP-1 promoter were cloned into pGL2-basic vector (basic) that did not contain any promoter or enhancer but encoded firefly luciferase. The pGL2-SV40 construct

(SV40) that had both SV40 promoter and SV40 enhancer as well as the firefly luciferase reporter gene was used as positive control. After transfection, IFN-γ was added to the cell culture at 1000 units/ml. Cells were lysed 48 h after transfection, and luciferase expression was tested using a luminometer. Data shown are representative of at least five independent experiments. c, the TAP-1 gene was actively transcribed in SK-MEL-19 cells in the presence and absence of IFN-γ in nuclear run-on assay. Endogenous GAPDH expression was used as a positive control, and the pcDNA3.1/Hyg(+) vector was used as a negative control. The run-on experiments were repeated three times with similar results.

55

Figure 3.2 a.

-93 -593 -446 -427 -90 +1 -86 LMP-2 G/T TAP-1

AflIII AflIII AflIII 2.9kb 2.7kb

probe SK-MEL-19 HeLa SK-BR-3

5.6kb

2.7kb

(continued)

56

Figure 3.2 (continued)

b. -593 -446 +1 pTAP1/G G LUC

pTAP1/T T LUC

SV40 SV40 promoter LUC SV40 enhancer

basic LUC

40

35 IFN-r(-)

30 IFN-r(+)

25

20

15

10

5

0 Basic SV40 pTAP/T pTAP/G

c. SK-MEL-19 -IFNγ +IFN γ LMP-2

TAP-1

pcDNA3.1/Hyg(+)

GAPDH

57

Figure 3.3 TAP-1 mRNA level in SK-MEL-19 cells was increased by CHX.

The protein synthesis inhibitor CHX was added to the SK-MEL-19 cells and HeLa cells that had normal TAP-1 and HLA class I expression. Total RNA was isolated from both cells at different time points and subjected to Northern blot hybridization to detect TAP-1 expression. The blot was exposed to a PhosphorImager, and the signal intensity was quantified using ImageQuant 5.0 software (Amersham). After normalization of TAP-1 signal to endogenous GAPDH signal in each sample, the signals in the CHX-treated group were compared with those that received no CHX treatment. These signals were quantitated as fold of those in untreated cells.

58

Figure 3.3

SK-MEL-19

+ IFNγ -IFNγ CHX 16hr 8hr 4hr 2hr - 16hr 8hr 4hr 2hr - Folds 10.6 25.7 15.2 16.8 1 1.4 4.8 1.5 1.9 1

TAP1

GAPDH

Hela

+ IFNγ -IFNγ CHX 16hr 8hr 4hr 2hr - 16hr 8hr 4hr 2hr - Folds 2.7 2.4 1.8 1.0 1.0 3.9 3.4 2.2 1.4 1.0

TAP1

GAPDH

59

Figure 3.4 A homozygous single-nucleotide deletion was identified in the TAP-1 gene at position +1489 that resulted in premature termination codons. a, sequencing chromogram. The arrow points to the deletion site. b, primers hTAP1E7.f and hTAP1E7.r (arrows) were used to amplify the deletion region in TAP-1 exon 7 (E7).

The arrowhead points to the position of the deletion resulting in a new BslI site. The PCR

products of genomic DNA were purified and digested with BslI. Gel electrophoresis data

showed that SK-MEL-19 cells were homozygous for the +1489 deletion. U, uncut; B,

BslI-digested; M, molecular weight. c, the sequence of the deletion region is shown, and

the downstream premature termination codons (X) are underlined.

60

Figure 3.4 a. b.

E7

1102 SK-MEL-19 1480 1496 M U B U B

500bp 200bp 100bp

c.

T P Y T W R A L S S S K M S P Mu: 1483 ACTCCC-TACACTTGGAGGGCCTTGTCCAGTTCCAAGATGTCTCCT 1528 WT: 1483 ACTCCCTTACACTTGGAGGGCCTTGTCCAGTTCCAAGATGTCTCCT 1528 T P L H L E G L V Q F Q D V S

L P T Q T A Q M S X C Y R G X Mu: 1529 TTGCCTACCCAAACCGCCCAGATGTCTTAGTGCTACAGGGGCTGA 1572 WT: 1529 TTGCCTACCCAAACCGCCCAGATGTCTTAGTGCTACAGGGGCTGA 1573 F A Y P N R P D V L V L Q G L

61

Figure 3.5 The single-nucleotide deletion D1489 in the TAP-1 gene results in

accelerated decay of the TAP-1 mRNA by a non-NMD-related mechanism.

RNase protection assay of three Tet-Off SK-MEL-19 cell clones stably transfected with

TAP-1 WT, TAP-1 D1489, and TAP-1 Del3 cDNA constructs was conducted using the antisense transcripts as probe. a, depiction of the Tet-Off system and the pBI-EGFP/TAP-

1 constructs containing TAP-1 WT, TAP-1 D1489, and TAP-1 Del3 used to measure the mRNA half-lives. b, The nucleotide sequences around the mutation site in the cDNA constructs are denoted. c, RNase protection assay to demonstrate the accelerated decay of

TAP-1 D1489 and TAP-1 Del3 message compared with that of the wild type TAP-1. The antisense probe protects a 345-nucleotide TAP-1 mRNA (labled as B). The protected message for EGFP and GAPDH are labeled as A and C, respectively. One representative assay for each TAP-1 construct is shown. d, decay kinetics of the TAP-1 WT, TAP-1

D1489, and TAP-1 Del3 mRNA. Means ± S.D. of four experiments are presented. The half-lives of the mRNA were determined by manual best-fit regression analysis.

62

Figure 3.5 a.

b.

(continued)

63

Figure 3.5 (continued) c.

Tet. (Hrs) Tet. (Hrs) Tet. (Hrs) Ctrl RNase (-) (-) Ctrl RNase Ctrl RNase (-) (-) Ctrl RNase

Ctrl RNase (-) (-) Ctrl RNase 0 0.5 1 2 4 6 0 1 2 4 8 0 0.5 1 1.5 3 6 (+) Ctrl RNase Ctrl RNase (+) (+) Ctrl RNase Ctrl RNase (+) (+) Ctrl RNase A A A

B B B

C C C

WT D1489 Del3 d. 10

WT( t1/2=7.2 hours)

RNA remaining % D1489 (t1/2=3.5 hour) Del3 (t1/2=2.7 hours) 1 0 1 2 3 4 5 6 7 8 9 1 Hours after tetracycline

64

CHAPTER 4

FUNCTIONAL CONSERVATION BETWEEN THE SIGNATURE MOTIF OF TAP-1 AND CYSTIC FIBROSIS TRANSMEMBRANE REGULATOR PROTEIN: A ROLE OF ITS POLYMORPHISM IN HLA DOWN-REGULATION IN CANCER CELLS.

4.1 Introduction.

Major histocompatibility complex (MHC) class I molecules are ligands for cytotoxic T lymphocytes on the surface of most tissue cells and are important for the immune recognition of intracellular pathogens and cancer cells. The transporter associated with antigen processing (TAP), a heterodimer of TAP-1 and TAP-2, plays an essential role in the MHC class I-restricted antigen processing and presentation pathway.

TAP translocates antigenic peptides ranging from 8-15 residues across the ER membrane.

These peptides are typically generated in the cytosol by the 20S proteasome. Upon being translocated into the ER lumen, some of the peptides bind to MHC class I heavy chain/β2-microglobulin (β2m) heterodimers with the help of a peptide-loading complex consisting of tapasin, ERp57 and calreticulin (reviewed in ref. [9, 11]). Recent studies by

several groups also provide convincing evidence that through fusion of phagosome and

ER, exogenous antigens can be presented on MHC class I molecules by professional

antigen presenting cells in a TAP-dependent manner [67, 68, 242].

65

TAP-1 and TAP-2 belong to the ATP-binding-cassette (ABC) superfamily, which is comprised of a diverse class of proteins that carry various substances, ranging from ions to proteins, across cellular membranes [92, 93]. The NBDs of the ABC transporters are structurally related to non-transporter enzymes that are involved in DNA repair or structural maintenance of chromosomes [92, 93, 113]. Defects in several ABC transporter members result in human diseases. Mutations in the cystic fibrosis transmembrane conductance regulator (CFTR), a chloride channel located in the apical membrane of epithelial cells, causes cystic fibrosis (CF) in humans [92]. Over- expression of P-glycoprotein or multiple-drug resistance protein (MDR), a transporter that actively extrudes cytotoxic drugs to maintain low intracellular levels, leads to multiple-drug resistance of certain cancer cells [92]. A third medically important ABC transporter is the TAP transporter, disruption of which results in a lack of cell surface

MHC class I expression (human type 1 bare lymphocyte syndrome) [127]. Disruption of

TAP activity in cancer cells also leads to class I down-regulation on the surface, which facilitates immune evasion [208].

A prototype ABC transporter is composed of two paired membrane integral domains that are thought to form the translocation pore, and two paired nucleotide- binding domains (NBD), that serve as an ATPase to fuel the translocation process. For some ABC transporters, such as CFTR, these domains are fused as a single polypeptide chain, while for others, like TAP, two polypeptides are separately encoded, each containing one trans-membrane domain and one NBD. Members of the ABC transporter superfamily contain several conserved sequence motifs in their NBDs, including the

66

Walker A and Walker B sequences, the signature motif or C loop, Q loop, and D loop.

With the exception of the Walker A and B motifs, which are required for the ATP

binding and ATP hydrolysis, roles of the other conserved motifs in the ABC transporter

function are largely unknown. The well-conserved signature motif with the consensus

LSGGQ/E sequence is the most intriguing, since it is conserved in all ABC transporter

family members, but not found outside this superfamily. The importance of the signature

motif in human disease is demonstrated by the fact that mutations in this motif lead to loss or impairment of the transporter function of human CFTR [117, 119, 120, 243]. It was therefore of great interest as to whether this structurally conserved signature motif is

also functionally conserved in TAP, and if so, whether variations of this region contribute

to frequently observed down-regulation of HLA in human cancer. If this is the case, the

naturally occurring mutations that abrogate the CFTR activity should also inactivate

TAP-1. We introduced into TAP-1 mutations analogous to those frequently identified in

CF patients [117-120] and tested the effect of these mutations on TAP-1 function in a

TAP-1-deficient melanoma cell line, which lacked cell surface HLA. Our results showed

that all the CF-like mutations severely impaired the ability of TAP-1 to restore surface

HLA expression. One of the mutants, G646D, lost transporter function completely, even

though its localization in the ER and association with the peptide-loading complex were

unaffected. These results are consistent with the hypothesis that the signature motif is

also functionally conserved between TAP and CFTR. These data prompt us to search for

natural TAP-1 variants at or near the signature motif. Our sequencing of the DNA

fragments encoding the signature motif from 103 colon cancer samples revealed a 17.5-

67

fold increase in the frequency of a rare TAP-1 allele among the HLAlow samples in

comparison to those with normal cell surface HLA. This allele, G1943A, results in R->Q

replacement of the amino acid immediately adjacent to the signature motif (R648Q). We

analyzed the transporter activity of TAP-1 R648Q, and found that it was reduced by more

than 80%. These results raise the intriguing possibility that variations in the vicinity of

the TAP-1 signature motif play a role in down-regulation of cell surface HLA class I in

human colon cancer.

4.2 Results.

4.2.1 CF-like mutations in TAP-1 signature motif abrogate TAP-1 activity.

The signature motif is one of the most conserved motifs in the NBDs of all the ABC

family members (examples shown in Figure 4.1a). By site-directed mutagenesis, we

introduced the following mutations into the signature motif of TAP-1: S644R, G645R,

G646S and G646D. Nucleotide changes and amino acid changes are summarized in

Table 4.1. Each of these mutations on CFTR has been shown to cause CF [126].

We have identified a homozygous single-nucleotide deletion (D1489) of the TAP-1

gene in a melanoma cell line, SK-MEL-19, which lacks surface MHC class I expression.

The D1489 mutation caused rapid degradation of TAP-1 mRNA. By introducing wild-

type TAP-1, surface MHC class I expression could be restored in this cell line [244]. In

order to characterize the function of the TAP-1 mutants, we reconstituted the SK-MEL-

19 cells with either empty vector or wild type or mutant TAP-1 constructs. After the

stable cell clones were established, the expression of TAP-1 proteins was examined by

68

Western blotting using a rabbit anti-human TAP-1 polyclonal antibody. Cell clones that

expressed mutant protein at levels greater than or comparable to wild type TAP-1

proteins were selected (Figure 4.2c), and their cell-surface MHC class I levels were evaluated by flow cytometry. As reported previously, vector-transfected cells expressed little cell surface HLA, and wild-type TAP-1 restored HLA expression (Figure 4.2a).

Cells transfected with mutant TAP-1 constructs had reduced restoration of cell surface

HLA, to between 1-20% compared to cells that were transfected with wild type TAP-1

(Figure 4.2b). These results demonstrate that mutations analogous to those that inactivate

CFTR function also have a detrimental effect on TAP-1 function.

4.2.2 TAP-1 G646D is expressed in the ER and remains associated with the peptide- loading complex.

To define the intracellular localization of TAP-1 mutants, we stained stably transfected cells with an antiserum against TAP-1 and an ER-resident protein, calnexin, and analyzed the stained cells by immunofluorescence microscopy. As shown in Figure

4.3a, a perinuclear reticular staining pattern characteristic of the ER was observed in cells expressing both the wild type and the G646D mutant TAP-1 protein. In addition, both the wild-type and G646D TAP-1 were co-localized with calnexin (red), as demonstrated by the yellow fluorescence in the merged images. The expression of TAP-2 protein was also examined by immunofluorescence microscopy (Figure 4.3b). Although TAP-2 protein was observed (red) expressed in a similar pattern in both the wild type and G646D TAP-1 transfected clones, it was barely detectable when TAP-1 was absent.

69

It is known that antigenic peptides transported by the TAP transporters bind to the

empty heterodimer of MHC class I heavy chain and β2m in the ER lumen with the help

of several chaperone proteins, such as ERp57 and calreticulin. Bridged by tapasin, MHC

class I molecules associate with the TAP transporter to form the peptide-loading complex

[9, 11]. We tested whether the G646D mutation in the TAP-1 protein disrupted the

formation of interactions within the peptide-loading complex. Immunoprecipitation

assays were performed using the rabbit anti-TAP-1 antiserum, a monoclonal anti-TAP-2

antibody, and W6/32, a mAb recognizing β2m-associated MHC class I heavy chains. The

precipitated lysates were then analyzed by Western blotting to detect TAP-1, TAP-2 and

tapasin. As shown in Figure 4.3c, the G646D mutation did not interfere with the

formation of the tapasin/TAP complex (left and middle panels, Figure 4.3c). Both wild-

type TAP-1 and G646D TAP-1 were detectable in immunoprecipitations with W6/32,

indicating that the G646D mutation also did not interfere with the class I/TAP interaction

(Figure 4.3c, right panel). The thiol-reductase ERp57 was also detected by sequencing

after we separated the cell lysates precipitated by the antiserum against TAP-1 in a two-

dimensional SDS-PAGE (data not shown).

It was noticed that without TAP-1 proteins in the cell, TAP-2 proteins could not

be detected in the lysates of vector-transfected clone that were precipitated by the

antibodies against TAP-1 or TAP-2 (left and middle panels, Figure 4.3c). Only a weak band was observed when the lysates pulled down by W6/32 mAb were blotted with anti-

TAP-2 mAb after long-time exposure (data not shown). Previously we have shown that the TAP-2 mRNA level in the SK-MEL-19 cells was greatly enhanced by IFN-γ [244].

70

However, after IFN-γ treatment at 1000u/ml for 48 hours, the TAP-2 protein level was

only slightly increased. Since the TAP-2 protein levels are greatly increased (Figure 4.3b)

and support the surface MHC class I expression when TAP-1 is expressed (Figure 4.2a), it appears that the TAP-2 protein is unstable when TAP-1 is absent.

4.2.3 TAP-1 G646D failed to transport peptide across the ER membrane.

We then evaluated the transporter function of G646D TAP-1 by a microsome-based peptide translocation experiment described previously [99]. Microsomes were prepared

from stable cell clones expressing the wild-type TAP-1, G646D TAP-1 and vector

control, respectively. Each microsome preparation was quantified for total protein and

examined for TAP-1/2 protein expression by immunoblotting (Figure 4.4b). A

radioiodinated model peptide RR125-IYNASTEL was incubated with the microsome in the

presence or absence of 5 mM ATP for 15 min at 37°C. Membranes were separated from

the un-translocated peptides by centrifugation. Since the peptide contains an N-linked glycosylation consensus sequence, N-glycosylation occurs upon translocation into the

microsome lumen. Translocated radiolabeled peptides were recovered and then

quantified using the lectin concanavalin A coupled to Sepharose beads. As shown in

Figure 4.4a, microsomes containing the wild-type TAP-1 had a ratio of average

cpm+ATP/cpm-ATP about 10, which indicated a successful specific peptide-translocation.

The microsomes expressing the G646D TAP-1 or vector only did not show a significant

71

increase in signal when ATP was present in the assays. We concluded that the mutant

G646D TAP-1 had lost transporter function completely. Similar results were obtained in three independent experiments.

4.2.4 Increase of the G1943A allele frequency in HLAlow colon cancer tissues.

Down-regulation of cell surface MHC class I expression is one of the strategies tumor cells used to evade immune surveillance by CD8+ T lymphocytes [161, 208], which also poses a major obstacle for T-cell based cancer immunotherapy [7].

Disruption of the TAP transporter is one of the reasons for the decreased surface MHC class I level, that has been frequently encountered in a variety of tumors [208, 214, 244,

245]. The CF-like mutations that we introduced into the TAP-1 protein are frequently identified in CF patients. It was therefore interesting to investigate whether such functionally defective TAP-1 variants contribute to MHC class I down-regulation in cancer.

Tissue microarrays were acquired from the National Cancer Institute and Zymed

Laboratories Incorporation, which consist of various types of tumors, including colon cancer, breast cancer, lung cancer, melanoma, lymphoma, prostatic adenocarcinoma, ovarian adenocarcinoma and brain tumor. Individual cancer tissues were collected by

LCM or manually for DNA extraction. The exon 10 region of TAP-1 gene encompassing the signature motif was amplified by nested PCR, for which forward primers were designed to reside in the intron 9 to avoid contamination from the TAP-1 cDNA constructs.

72

All the cancer samples were also analyzed for their MHC class I expression by

immunohistochemistry using monoclonal antibody HC-10, which recognizes the heavy

chain of HLA-A,B,C. The standard by which we graded a sample as MHC class I

negative was that at least 75% of tumor cells showed severely decreased level of MHC

class I compared to the normal tissues. Typical patterns of HC-10 staining for MHC class

I positive and negative samples of colon cancer are shown in Figure 4.5 a-c, respectively.

Our screening of 228 human cancer tissues, 66 of which were negative for class I MHC

expression, failed to identify any mutation within the signature motif (data not shown).

However, an increase in the frequency of a rare allele, hereby called G1943A, was noted

in the HLA- cancer samples (Table 4.2). Since 103 of the 228 samples were colon cancer, we compared HLA+ and HLA- colon cancer for the allelic frequency of G1943A,

primarily because the sample size was large enough for us to do statistical analyses.

Among the colon cancer samples studied, 88 were MHC class I positive, and 15 were

negative, which was similar to the frequency reported previously [184, 246]. The

frequency of the G1943A allele occurring in the MHC class I negative samples (10.00%,

3 out of 30) was much higher than that in the positive group (0.57%, 1 out of 176), and

the difference was statistically very significant (p=0.0005).

4.2.5 Function of TAP-1 R648Q.

We tested the function of the protein product of the G1943A TAP-1 allele, R648Q.

By comparing the NBD sequences of ABC transporters (Figure 4.1a) and that of TAP-1

from different species (Figure 4.1b), we found that either lysine or arginine, both being

73

basic amino acids, are present at position 648 in most cases. Since the residue is just next

to the essential LSGGQ/E sequence, we hypothesized that the replacement of the arginine

residue by glutamine at position 648 may affect the TAP function. We cloned the TAP-1

R648Q by site-directed mutagenesis and transfected it into the SK-MEL-19 cells. Three

stable cell clones that expressed comparable levels of TAP-1 protein to the wild-type

clone were examined for their surface MHC class I by flow cytometry. As shown in

Figure 4.6 a and b, SK-MEL-19 cells reconstituted with R648Q TAP-1 expressed MHC

class I on the surface to about 28.62% of that with the wild type. Significant defects remained even after treatment with interferon (Figure 4.6a right panels).

To test the transporter function of R648Q TAP-1, we prepared microsomes from two of the three stable cell clones expressing R648Q TAP-1. Microsomes containing the wild-type TAP-1 and the cloning vector only or the nonfunctional mutant G646D TAP-1 were isolated to serve as positive control and negative control, respectively. Wild-type

TAP-1 had a ratio of average cpm+ATP/cpm-ATP about 10 (Figure 4.6e). The microsomes expressing the G646D TAP-1 or vector alone did not show significant increase of signal when ATP was included in the assays. As expected, the R648Q TAP-1-expressing microsomes did transport peptide, but with much reduced efficiency. The two microsome-preparations we used exhibited an average of 19.81% of transporter activity relative to microsomes containing the wild-type TAP-1. The reduction of TAP activity is similar for two different clones, and these results were reproducible in two independent

74

translocation analyses. We concluded that peptide translocation by R648Q TAP-1 was impaired, which resulted in the reduced cell surface expression of MHC class I molecules.

4.3 Discussion

It has been shown that the ABC signature motif is well conserved among all the

ABC transporter family members and its disruption results in the loss of protein function

[117-120]. As a first step to determining whether the signature motifs of TAP-1 and

CFTR serve the same function, we introduced the mutations that inactivate CFTR in CF patients into TAP-1 protein. Our results demonstrate that these mutations also inactivated

TAP-1, as analyzed by its ability to restore cell surface HLA in a TAP-1-deficient cancer cell line and by ATP-driven peptide translocation assay. These data suggest that the signature motif likely serves a similar function in TAP-1 and CFTR. Recently, one study showed that simultaneous replacement of the conserved serine and second glycine in the signature motif of TAP-1 with alanines resulted in inactivation of TAP-1 transporter function [125], which is also consistent with the essential function of TAP-1 signature motif. The crystal structures of the ABC transporter MJ0796 and the DNA repair enzyme

Rad50 (that contains an NBD that is structurally related to ABC transporter NBD) show that the signature motif of one NBD interacts with the Walker A region of a second NBD in a dimeric arrangement [113, 116]. Specific hydrogen bonds are formed between the serine and the second glycine residues of the signature motif and the phosphate oxygens of ATP [113, 116]. In this way, the signature motif may not only mediate the interaction

75

between the two NBDs but also act as a sensor for the hydrolysis of ATP γ-phosphate and

mediate conformational changes driven by ATP binding and hydrolysis during the

substrate translocation process [113-115]. Additional studies are required to elucidate the

precise roles of the signature motif residues during peptide translocation.

The fact that the replacement of a single amino acid in the signature motif can

inactivate TAP-1 raises the possibility that such mutations may be observed among

cancer samples that have low cell surface HLA, although our survey of 228 samples has

not identified such a mutation. Nevertheless, we did observe a 17.5-fold increase in the

frequency of a rare allele among the colon cancer samples that showed reduced cell

surface HLA. Functional analysis suggests that TAP-1 encoded by this allele has only

19.81% of transporter activity compared to the wild type. The surface MHC class I

expression restored by this TAP-1 variant was about 30% of the wild type.

Although the significant increase of the frequency of the defective A allele among

the HLAlow colon cancer samples strongly suggests a role of genetic polymorphism of

TAP in HLA expression, several caveats deserve consideration. First, when over- expressed, this allele can restore cell surface HLA to about 30% of what was induced by wild-type TAP-1 (Figure 4.6 a and b). One may thus wonder whether cells expressing this allele alone may show severe reduction in cell surface HLA. However, it should be pointed out that in our assay, the expression of the variant TAP-1 protein was driven by a

CMV promoter and consequently over-expressed. In fact, in a previous study [214], no

76

cell surface HLA was detectable in a small cell lung cancer cell line due to selective

expression of another TAP-1 variant, R659Q, which was later shown to have as much as

50% of wild-type transporter activity [109].

Second, based on the allele-A frequency derived from the normal population, which is 3.01% in our screening of 166 normal blood donors (data not shown) and 2.6% reported by another study [247], the chance that a homozygous A being identified is less

than 0.1% according to the Hardy-Weinberg equilibrium, which explains the absence of

homozygous A allele in our samples. It is not clear how the defective A allele, in a

heterozygous individual, could contribute to MHC class I down-regulation in cancer

cells. One potential mechanism would be the loss of heterozygosity (LOH). LOH has

been found to contribute to β2m deficiency in melanoma [206] and HLA haplotype loss

in colorectal and laryngeal carcinomas [248]. Unfortunately, this possibility could not be experimentally assessed, as the HLAlow colon cancer samples from which the A allele

was identified were prepared in a fashion that was not suitable for micro-dissection, and

the source materials are no longer available from the supplier. It is also possible that in

some tumor cells, only one allele of TAP-1 gene is transcribed. It has been reported

previously that a defective TAP-1 allele, but not the wild-type one, was expressed in a

human small cell lung cancer cell line, which led to loss of cell surface MHC class I

[214]. It is also possible that other components of the antigen presentation pathway, in

addition to the defective TAP-1 alleles, contributed to the class I down-modulations we

observed in the colons cancer samples.

77

Taken together, the data presented in this manuscript showed functional conservation of the signature motif between CFTR and TAP-1. Moreover, the fact that a defective allele is present with increased frequency among colon cancer with HLA down- regulation suggests that genetic polymorphism in the TAP-1 gene can be a contributing factor for tumor evasion of host immunity.

78

Signature motif sequence Nucleotide change Amino acid change

LSGGQ TCA>CGA S644R

LSGGQ GGG>GCG G645R

LSGGQ GGT>GAT G646D

LSGGQ GGT>AGT G646S

Table 4.1 Nucleotide changes and amino acid substitutions generated by the site- directed mutagenesis.

79

Cancer tissue MHC (+) MHC (-) number G/G G/A A/A G/G G/A A/A Breast adenocarcinoma 22 12 1 0 9 0 0

Brain tumor 6 3 0 0 3 0 0

Melanoma 7 5 0 0 2 0 0

Colonic adenocarcinoma 103 87 1 0 12 3 0

Lymphoma 25 23 2 0 0 0 0

Lung cancer 21 11 0 0 9 1 0

Ovarian adenocarcinoma 18 10 0 0 8 0 0

Prostatic adenocarcinoma 26 7 0 0 18 1 0

Total 228 158 4 0 61 5 0

Table 4.2 TAP-1G1943A genotyping in different types of human cancer.

80

a.

(670) 670 680 690 700 710 720 730 7 HisP(130) ERAVKYLAKVGIDERAQGKYPVHLSGGQQQRVSIARALAMEPEVLLFDEPTSA hCFTR-C(512) PGKLDFVLVDG------GCVLSHGHKQLMCLARSVLSKAKILLLDEPSAH hCFTR-N(534) AEKDNIVLGE------GGITLSGGQRARISLARAVYKDADLYLLDSPFGY hMDR1-N(517) PHKFDTLVGE------RGAQLSGGQKQRIAIARALVRNPKILLLDEATSA hMDR1-C(462) PNKYSTKVGD------KGTQLSGGQKQRIAIARALVRQPHILLLDEATSA hTAP1(629) PQGYDTEVDE------AGSQLSGGQRQAVALARALIRKPCVLILDDATSA hTAP2(593) EHGIYTDVGE------KGSQLAAGQKQRLAIARALVRDPRVLILDEATSA

b.

(620) 620 630 640 650 660 670 680 human(620) GAHSFISGLPQGYDTEVDEAGSQLSGGQRQAVALARALIRKPCVLILDDATSAL gorilla(620) GAHSFISGLPQGYDTEVGEAGSQLSGGQRQAVALARALIRKPCVLILDDATSAL mouse(596) GAHDFISGFPQGYDTEVGETGNQLSGGQRQAVALARALIRKPLLLILDDATSAL rat(597) GAHDFISGFPQGYDTEVGETGNQLSGGQRQAVALARALIRKPRLLILDDATSAL horn shark(419) NAHRFITELKDGYNTDAGEKGGQLSGGQKQRVAIARALIRDPRVLILDDATSCL rainbow trout(590) NAHKFISDLPNGYDTDAGEKGGQVSGGQKQRIAIARALIRKPRILVLDDATSNL

Figure 4.1 Sequence alignment of the NBDs of ABC transporters.

The signature motifs of several human ABC transporters, including TAP-1/2, CFTR,

MDR1, and the bacterial histidine permease HisP were aligned in a. The signature motifs of TAP-1 from different species were compared in b. The conserved LSGGQ/E sequence of signature motif is marked by a rectangle.

81

Figure 4.2 Expression of TAP-1 mutant proteins and cell surface MHC class I in

cells stably transfected with wild type or mutant TAP-1.

a. SK-MEL-19 cell clones stably transfected with the TAP-1 mutants, G644R, G645R

G646S and G646D as indicated were examined for their surface MHC class I by flow

cytometry after being stained with a PE-conjugated anti-human HLA-A,B,C antibody

(bold line) or isotype control (PE-conjugated mouse IgG1, dotted lines). The numbers in the panels indicate the MFI of each clone after the MFI of isotype control staining was subtracted. The results shown are representatives of at least three clones for each mutant. b. The level of restored MHC class I was calculated by subtracting the MFI of vector transfectant from that of the TAP-1-transfectants. The % HLA restored by each mutant clone in comparison to the wild-type TAP-1 (defined as 100%) was plotted. c. The expression of TAP-1 protein was detected by immunoblotting of proteins in cell lysates with a rabbit polyclonal antibody, which demonstrated that the TAP-1 protein levels were comparable between the clones containing the mutants and the wild type. The expression of beta-actin was examined as a loading control.

82

Figure 4.2

a Vector WT S644R 100 100 100 0.40 177.35 4.07 80 80 80

60 60 60

40 40 40

20 20 20

0 0 0 1 10 100 1000 100 1 10 100 1000 100 1 10 100 1000 100 Counts G645R G646S G646D 100 100 100 41.15 16.94 1.61 80 80 80

60 60 60 % of Max % of Max 40 40 40

20 20 20

0 0 0 1 10 100 1000 100 1 10 100 1000 100 1 10 100 1000 100 Log Fluorescence Intensity

b ) 120 100% 100

80

60

40 23.03% 20 2.07% 9.35% 0.68% relative to WT (% 0 WT S644R G645R G646S G646D

c

R WT G645R G646S G646D S644 Vector TAP1

β-actin

83

Figure 4.3 TAP-1 G646D was expressed in the ER and associated with TAP-2, tapasin, and MHC class I heavy chain. a. Immunofluorescence staining with antibodies against calnexin (red), an ER-resident protein, and TAP-1 are shown, respectively. Nuclei were stained blue by DAPI. A yellow fluorescence was observed in the merged images, which indicates the co-localization of the TAP-1 proteins, both wild type and the G646D mutant, with the ER marker calnexin. b. Similarly, immunofluorescence staining with antibodies against TAP-2 (red) and TAP-

1 (green) are shown, respectively. Nuclei were stained blue by DAPI. A yellow fluorescence was observed in the merged images, which indicates the co-localization of the TAP-1 proteins, both wild type and the G646D mutant, with TAP-2. Without TAP-1 expression, no TAP-2 protein was detected. c. Immunoprecipitations were performed using antibodies against TAP-1 (rabbit antiserum), TAP-2 (monoclonal) and β2m-

associated MHC class I heavy chain (monoclonal, W6/32), respectively. The precipitated

proteins were separated by SDS-PAGE and then blotted with anti-TAP-1 antiserum, mAb

against TAP-2 and anti-tapasin antiserum. V: vector-tranfected cells were used.

84

Figure 4.3

a

(continued)

85

Figure 4.3 (continued)

b

(continued)

86

Figure 4.3 (continued)

c

87

Figure 4.4 TAP-1 G646D failed to transport peptide across ER membrane. a. Peptide translocation assay. For each of the wild type, G646D or the vector-containing microsomes, the average cpm from three tests with or without ATP was plotted. The result is representative of three independent analyses. b. Preparations of microsomes used in the translocation assay that contain the wild-type TAP-1, G646D TAP-1 and vector control (V), respectively, were evaluated for the expression level of TAP-1 and TAP-2 by immunoblotting.

88

Figure 4.4

a

14000 -ATP +ATP 12000

10000

8000

cpm 6000

4000

2000

0 Vector WT G646D

b

V WT G646D

TAP1

TAP2

89

Figure 4.5 Immunohistochemical staining of colon cancer with HC-10 mAb.

a. and b. show HLAHi and HLALow colon cancer samples, respectively. Class I heavy

chain expression was detected as brown color in the cytoplasm as well as on the cell

surface. c. An HLA- colon cancer sample harboring heterozygous G1943A allele, in

which the surface MHC class I staining was completely lost. H&E staining of the same samples are shown in the bottom panel as (d-f). Magnification: 400×.

90

Figure 4.5

91

Figure 4.6 Defective activity of TAP-1 R648Q.

a. Restoration of cell-surface HLA expression in the SK-MEL-19 cells. Three clones that expressed TAP-1 R648Q and one clone containing the wild-type TAP-1 or the pcDNA3.1/Hyg(+) vector were examined for the surface MHC class I by flow cytometry

using a PE-conjugated anti-HLA-A,B,C antibody (bold line). A PE-conjugated mouse

IgG1 antibody was used as isotype control (dotted line). The MHC class I level of these

cell clones was also examined after being treated with IFN-γ for 48 hours at 1000

units/ml (right panels). After substraction of MFI from the isotype control, the MFI was

indicated in each panel. b. TAP-1 and TAP-2 protein expressions in each clone were

detected by immunoblotting of cell lysates using the anti-TAP-1 antiserum and anti-TAP-

2 mAb, respectively. The beta-actin level is shown as a loading control. c. Cell surface

HLA restored by mutant R648Q as a percentage of that restored by wild-type TAP-1. d.

Transporter activity of R648Q. Peptide translocation assay was performed using

microsomes prepared from cell clones that expressed vector, wild-type TAP-1, G646D

TAP-1 and R648Q TAP-1, respectively. For each microsome, the average cpm from

three tests with or without ATP was plotted. One of two independent assays is shown

and two different clones were tested in each assay.

92

Figure 4.6

a -IFN- +IFN-γ 100 100 3.87 30.5 80 80

60 60

40 40 Vector

20 20

0 0 1 10 100 1000 100 1 10 100 1000 100 100 100 139.00 416.14

80 80 Counts

60 60

40 40 WT

20 20

0 0 1 10 100 1000 100 1 10 100 1000 100 100 100 36.61 196.07 80 80

60 60 R648Q-1 40 40

20 20

0 0 1 10 100 1000 100 1 10 100 1000 100 100 100 51.10 202.50 80 80

60 60 R648Q-2

40 40

20 20

0 0 1 10 100 1000 100 1 10 100 1000 100 100 100 39.94 176.66 80 80

60 60 R648Q-3

40 40

20 20

0 0 1 10 100 1000 100 1 10 100 1000 100 Log Fluorescence Intensity (continued) 93

Figure 4.6 (continued)

b

120 - IFN-γ 100% 100 + IFN-γ

80

60 42.93% 44.60% 37.90% 40 34.95% 24.23% 26.69% Relative toRelative WT (%) 20

0 WT R648Q-1 R648Q-2 R648Q-3

c R648Q-2 R648Q-3 Vector WT R648Q-1

TAP1

TAP2 -IFN-γ

β-actin

(continued)

94

Figure 4.6 (continued)

d

7000 -ATP 6000 +AT P 5000 4000

cpm 3000 2000 1000 0 Vector WT G646D R648Q-1 R648Q-2

95

CHAPTER 5

CONCLUSION

In summary, we have identified two novel mechanisms that cause loss of TAP-1

expression post-transcriptionally and defective function of TAP transporter, respectively,

which lead to the down-regulation of surface MHC class I in human cancer cells.

One mechanism we identified is through rapid turnover of TAP-1 mRNA that contains a frame-shifting mutation D1489. By comparing the stability of TAP-1 mRNA that harbors the D1489 mutation or three-nucleotide deletion at the same position without interrupting the reading frame, we found that mechanisms other than NMD may mediate the accelerated degradation of TAP-1 (D1489) mRNA. Our finding suggests that the sequence where this mutation resides may be important for maintaining the stability of

TAP-1 mRNA. It will be of great interest to examine whether in human cancer cells the

D1489 mutation occurs frequently.

The other mechanism we found is through abrogation of TAP transporter function.

First we provide evidence that the signature motif of TAP-1 protein is essential for normal TAP transporter activity and the function of signature motif is conserved between

96

TAP and CFTR. Though the signature motif of CFTR has been known as a mutation “hot

spot”, we were unable to identify any mutations in the signature motif of TAP-1 in 228 human cancer samples, among them 66 samples have down-regulated MHC class I expression. Interestingly a polymorphic form of TAP-1 with an R->Q replacement at position just C-terminal to the consensus signature sequence LSGGQ/E was found to occur at significantly higher frequency in the class I HLA- colon cancer samples than

those with normal HLA expression. Further analysis showed that the activity of TAP-1

Q648 was severely decreased compared to TAP-1 R648, which indicates that this

defective allele may be a contributing factor for the class I down-regulation in human

colon cancer. It will be interesting to examine whether it is also involved in the loss of

class I expression in other kinds of human cancers.

Since this allele (Q648) occurs at high frequency in the HLA down-regulated

colon cancer samples, colon cancer patients who have received T cell based cancer

immunotherapy but remain nonresponsive or undergo relapse may be screened for this

allele. Cytokine treatment, such as with IFN-γ, or other methods to increase cell surface

MHC class I expression may be beneficial to the individuals that harbor this allele.

Along with the above discoveries, we also made interesting observations about

TAP-1 gene expression. 1) In the nuclear run-on assay, we found that TAP-1 gene is constitutively transcribed, whereas the transcription of LMP-2 gene needs cytokine (IFN-

γ) induction. However, the mRNA levels of both TAP-1 and LMP-2 in melanoma cell lines are not detectable by Northern blot unless being stimulated with IFN-γ. Therefore, it

is likely that the TAP-1 mRNA is unstable under physiological condition. Cytokines that

97

are enriched when immune responses take place, such as IFN-γ, not only enhance the

transcription of TAP-1 but may also stabilize the message. In this way, the expression of

TAP-1 serves as a checkpoint for the surface level of class I MHC. 2) Without TAP-1 protein in the SK-MEL-19 cells, TAP-2 protein is undetectable. Upon overexpression of

TAP-1 protein into the cell, TAP-2 protein level is upregulated and the level of TAP-2

seems to be correlated with that of TAP-1. Therefore, it is likely that without TAP-1,

TAP-2 polypeptide is unstable. However, these interesting findings and their significance

need more work to prove.

98

BIBILIOGRAPHY

1. Tanchot, C., F.A. Lemonnier, B. Perarnau, A.A. Freitas, and B. Rocha, Differential requirements for survival and proliferation of CD8 naive or memory T cells. Science, 1997. 276(5321): p. 2057- 62.

2. Sprent, J. and C.D. Surh, T cell memory. Annu Rev Immunol, 2002. 20: p. 551-79.

3. Brodsky, F.M., L. Lem, A. Solache, and E.M. Bennett, Human pathogen subversion of antigen presentation. Immunol Rev, 1999. 168: p. 199-215.

4. Fruh, K., A. Gruhler, R.M. Krishna, and G.J. Schoenhals, A comparison of viral immune escape strategies targeting the MHC class I assembly pathway. Immunol Rev, 1999. 168: p. 157-66.

5. Hewitt, E.W., The MHC class I antigen presentation pathway: strategies for viral immune evasion. Immunology, 2003. 110(2): p. 163-9.

6. Seliger, B., M.J. Maeurer, and S. Ferrone, Antigen-processing machinery breakdown and tumor growth. Immunol Today, 2000. 21(9): p. 455-64.

7. Khong, H.T. and N.P. Restifo, Natural selection of tumor variants in the generation of "tumor escape" phenotypes. Nat Immunol, 2002. 3(11): p. 999-1005.

8. van der Merwe, P.A. and S.J. Davis, Molecular interactions mediating T cell antigen recognition. Annu Rev Immunol, 2003. 21: p. 659-84.

9. Cresswell, P., N. Bangia, T. Dick, and G. Diedrich, The nature of the MHC class I peptide loading complex. Immunol Rev, 1999. 172: p. 21-8.

10. Gromme, M. and J. Neefjes, Antigen degradation or presentation by MHC class I molecules via classical and non-classical pathways. Mol Immunol, 2002. 39(3-4): p. 181-202.

11. Antoniou, A.N., S.J. Powis, and T. Elliott, Assembly and export of MHC class I peptide ligands. Curr Opin Immunol, 2003. 15(1): p. 75-81.

12. Rock, K.L., I.A. York, T. Saric, and A.L. Goldberg, Protein degradation and the generation of MHC class I-presented peptides. Adv Immunol, 2002. 80: p. 1-70.

13. Kloetzel, P.M., Antigen processing by the proteasome. Nat Rev Mol Cell Biol, 2001. 2(3): p. 179- 87.

14. York, I.A., A.L. Goldberg, X.Y. Mo, and K.L. Rock, Proteolysis and class I major histocompatibility complex antigen presentation. Immunol Rev, 1999. 172: p. 49-66.

99

15. Akiyama, K., K. Yokota, S. Kagawa, N. Shimbara, T. Tamura, H. Akioka, H.G. Nothwang, C. Noda, K. Tanaka, and A. Ichihara, cDNA cloning and interferon gamma down-regulation of proteasomal subunits X and Y. Science, 1994. 265(5176): p. 1231-4.

16. Akiyama, K., S. Kagawa, T. Tamura, N. Shimbara, M. Takashina, P. Kristensen, K.B. Hendil, K. Tanaka, and A. Ichihara, Replacement of proteasome subunits X and Y by LMP7 and LMP2 induced by interferon-gamma for acquirement of the functional diversity responsible for antigen processing. FEBS Lett, 1994. 343(1): p. 85-8.

17. Fruh, K., M. Gossen, K. Wang, H. Bujard, P.A. Peterson, and Y. Yang, Displacement of housekeeping proteasome subunits by MHC-encoded LMPs: a newly discovered mechanism for modulating the multicatalytic proteinase complex. Embo J, 1994. 13(14): p. 3236-44.

18. Groettrup, M., R. Kraft, S. Kostka, S. Standera, R. Stohwasser, and P.M. Kloetzel, A third interferon-gamma-induced subunit exchange in the 20S proteasome. Eur J Immunol, 1996. 26(4): p. 863-9.

19. Driscoll, J., M.G. Brown, D. Finley, and J.J. Monaco, MHC-linked LMP gene products specifically alter peptidase activities of the proteasome. Nature, 1993. 365(6443): p. 262-4.

20. Kuckelkorn, U., S. Frentzel, R. Kraft, S. Kostka, M. Groettrup, and P.M. Kloetzel, Incorporation of major histocompatibility complex--encoded subunits LMP2 and LMP7 changes the quality of the 20S proteasome polypeptide processing products independent of interferon-gamma. Eur J Immunol, 1995. 25(9): p. 2605-11.

21. Ehring, B., T.H. Meyer, C. Eckerskorn, F. Lottspeich, and R. Tampe, Effects of major- histocompatibility-complex-encoded subunits on the peptidase and proteolytic activities of human 20S proteasomes. Cleavage of proteins and antigenic peptides. Eur J Biochem, 1996. 235(1-2): p. 404-15.

22. Gaczynska, M., K.L. Rock, and A.L. Goldberg, Gamma-interferon and expression of MHC genes regulate peptide hydrolysis by proteasomes. Nature, 1993. 365(6443): p. 264-7.

23. Arnold, D., J. Driscoll, M. Androlewicz, E. Hughes, P. Cresswell, and T. Spies, Proteasome subunits encoded in the MHC are not generally required for the processing of peptides bound by MHC class I molecules. Nature, 1992. 360(6400): p. 171-4.

24. Momburg, F., V. Ortiz-Navarrete, J. Neefjes, E. Goulmy, Y. van de Wal, H. Spits, S.J. Powis, G.W. Butcher, J.C. Howard, P. Walden, and et al., Proteasome subunits encoded by the major histocompatibility complex are not essential for antigen presentation. Nature, 1992. 360(6400): p. 174-7.

25. Yewdell, J., C. Lapham, I. Bacik, T. Spies, and J. Bennink, MHC-encoded proteasome subunits LMP2 and LMP7 are not required for efficient antigen presentation. J Immunol, 1994. 152(3): p. 1163-70.

26. Van Kaer, L., P.G. Ashton-Rickardt, M. Eichelberger, M. Gaczynska, K. Nagashima, K.L. Rock, A.L. Goldberg, P.C. Doherty, and S. Tonegawa, Altered peptidase and viral-specific T cell response in LMP2 mutant mice. Immunity, 1994. 1(7): p. 533-41.

100

27. Fehling, H.J., W. Swat, C. Laplace, R. Kuhn, K. Rajewsky, U. Muller, and H. von Boehmer, MHC class I expression in mice lacking the proteasome subunit LMP-7. Science, 1994. 265(5176): p. 1234-7.

28. Voges, D., P. Zwickl, and W. Baumeister, The 26S proteasome: a molecular machine designed for controlled proteolysis. Annu Rev Biochem, 1999. 68: p. 1015-68.

29. Sijts, A., Y. Sun, K. Janek, S. Kral, A. Paschen, D. Schadendorf, and P.M. Kloetzel, The role of the proteasome activator PA28 in MHC class I antigen processing. Mol Immunol, 2002. 39(3-4): p. 165-9.

30. Seifert, U., C. Maranon, A. Shmueli, J.F. Desoutter, L. Wesoloski, K. Janek, P. Henklein, S. Diescher, M. Andrieu, H. de la Salle, T. Weinschenk, H. Schild, D. Laderach, A. Galy, G. Haas, P.M. Kloetzel, Y. Reiss, and A. Hosmalin, An essential role for tripeptidyl peptidase in the generation of an MHC class I epitope. Nat Immunol, 2003. 4(4): p. 375-9.

31. Wei, M.L. and P. Cresswell, HLA-A2 molecules in an antigen-processing mutant cell contain signal sequence-derived peptides. Nature, 1992. 356(6368): p. 443-6.

32. Bouvier, M. and D.C. Wiley, Importance of peptide amino and carboxyl termini to the stability of MHC class I molecules. Science, 1994. 265(5170): p. 398-402.

33. Levy, F., L. Burri, S. Morel, A.L. Peitrequin, N. Levy, A. Bachi, U. Hellman, B.J. Van den Eynde, and C. Servis, The final N-terminal trimming of a subaminoterminal proline-containing HLA class I-restricted antigenic peptide in the cytosol is mediated by two peptidases. J Immunol, 2002. 169(8): p. 4161-71.

34. Beninga, J., K.L. Rock, and A.L. Goldberg, Interferon-gamma can stimulate post-proteasomal trimming of the N terminus of an antigenic peptide by inducing leucine aminopeptidase. J Biol Chem, 1998. 273(30): p. 18734-42.

35. York, I.A., A.X. Mo, K. Lemerise, W. Zeng, Y. Shen, C.R. Abraham, T. Saric, A.L. Goldberg, and K.L. Rock, The cytosolic endopeptidase, thimet oligopeptidase, destroys antigenic peptides and limits the extent of MHC class I antigen presentation. Immunity, 2003. 18(3): p. 429-40.

36. Reits, E., A. Griekspoor, J. Neijssen, T. Groothuis, K. Jalink, P. van Veelen, H. Janssen, J. Calafat, J.W. Drijfhout, and J. Neefjes, Peptide diffusion, protection, and degradation in nuclear and cytoplasmic compartments before antigen presentation by MHC class I. Immunity, 2003. 18(1): p. 97-108.

37. Binder, R.J., N.E. Blachere, and P.K. Srivastava, Heat shock protein-chaperoned peptides but not free peptides introduced into the cytosol are presented efficiently by major histocompatibility complex I molecules. J Biol Chem, 2001. 276(20): p. 17163-71.

38. Wells, A.D., S.K. Rai, M.S. Salvato, H. Band, and M. Malkovsky, Hsp72-mediated augmentation of MHC class I surface expression and endogenous antigen presentation. Int Immunol, 1998. 10(5): p. 609-17.

39. Verma, R., S. Chen, R. Feldman, D. Schieltz, J. Yates, J. Dohmen, and R.J. Deshaies, Proteasomal proteomics: identification of nucleotide-sensitive proteasome-interacting proteins by mass spectrometric analysis of affinity-purified proteasomes. Mol Biol Cell, 2000. 11(10): p. 3425-39.

101

40. Chen, D. and M.J. Androlewicz, Heat shock protein 70 moderately enhances peptide binding and transport by the transporter associated with antigen processing. Immunol Lett, 2001. 75(2): p. 143-8.

41. Bouvier, M., Accessory proteins and the assembly of human class I MHC molecules: a molecular and structural perspective. Mol Immunol, 2003. 39(12): p. 697-706.

42. Sadasivan, B., P.J. Lehner, B. Ortmann, T. Spies, and P. Cresswell, Roles for calreticulin and a novel glycoprotein, tapasin, in the interaction of MHC class I molecules with TAP. Immunity, 1996. 5(2): p. 103-14.

43. Harris, M.R., Y.Y. Yu, C.S. Kindle, T.H. Hansen, and J.C. Solheim, Calreticulin and calnexin interact with different protein and glycan determinants during the assembly of MHC class I. J Immunol, 1998. 160(11): p. 5404-9.

44. Morrice, N.A. and S.J. Powis, A role for the thiol-dependent reductase ERp57 in the assembly of MHC class I molecules. Curr Biol, 1998. 8(12): p. 713-6.

45. Lindquist, J.A., O.N. Jensen, M. Mann, and G.J. Hammerling, ER-60, a chaperone with thiol- dependent reductase activity involved in MHC class I assembly. Embo J, 1998. 17(8): p. 2186-95.

46. Hughes, E.A. and P. Cresswell, The thiol ERp57 is a component of the MHC class I peptide-loading complex. Curr Biol, 1998. 8(12): p. 709-12.

47. Ortmann, B., J. Copeman, P.J. Lehner, B. Sadasivan, J.A. Herberg, A.G. Grandea, S.R. Riddell, R. Tampe, T. Spies, J. Trowsdale, and P. Cresswell, A critical role for tapasin in the assembly and function of multimeric MHC class I-TAP complexes. Science, 1997. 277(5330): p. 1306-9.

48. Bangia, N., P.J. Lehner, E.A. Hughes, M. Surman, and P. Cresswell, The N-terminal region of tapasin is required to stabilize the MHC class I loading complex. Eur J Immunol, 1999. 29(6): p. 1858-70.

49. Garbi, N., N. Tiwari, F. Momburg, and G.J. Hammerling, A major role for tapasin as a stabilizer of the TAP peptide transporter and consequences for MHC class I expression. Eur J Immunol, 2003. 33(1): p. 264-73.

50. Dick, T.P., N. Bangia, D.R. Peaper, and P. Cresswell, Disulfide bond isomerization and the assembly of MHC class I-peptide complexes. Immunity, 2002. 16(1): p. 87-98.

51. Lehner, P.J., M.J. Surman, and P. Cresswell, Soluble tapasin restores MHC class I expression and function in the tapasin-negative cell line .220. Immunity, 1998. 8(2): p. 221-31.

52. Lauvau, G., B. Gubler, H. Cohen, S. Daniel, S. Caillat-Zucman, and P.M. van Endert, Tapasin enhances assembly of transporters associated with antigen processing-dependent and - independent peptides with HLA-A2 and HLA-B27 expressed in insect cells. J Biol Chem, 1999. 274(44): p. 31349-58.

53. Serwold, T., F. Gonzalez, J. Kim, R. Jacob, and N. Shastri, ERAAP customizes peptides for MHC class I molecules in the endoplasmic reticulum. Nature, 2002. 419(6906): p. 480-3.

102

54. Saric, T., S.C. Chang, A. Hattori, I.A. York, S. Markant, K.L. Rock, M. Tsujimoto, and A.L. Goldberg, An IFN-gamma-induced aminopeptidase in the ER, ERAP1, trims precursors to MHC class I-presented peptides. Nat Immunol, 2002. 3(12): p. 1169-76.

55. York, I.A., S.C. Chang, T. Saric, J.A. Keys, J.M. Favreau, A.L. Goldberg, and K.L. Rock, The ER aminopeptidase ERAP1 enhances or limits antigen presentation by trimming epitopes to 8-9 residues. Nat Immunol, 2002. 3(12): p. 1177-84.

56. Wiertz, E.J., D. Tortorella, M. Bogyo, J. Yu, W. Mothes, T.R. Jones, T.A. Rapoport, and H.L. Ploegh, Sec61-mediated transfer of a from the endoplasmic reticulum to the proteasome for destruction. Nature, 1996. 384(6608): p. 432-8.

57. Hughes, E.A., C. Hammond, and P. Cresswell, Misfolded major histocompatibility complex class I heavy chains are translocated into the cytoplasm and degraded by the proteasome. Proc Natl Acad Sci U S A, 1997. 94(5): p. 1896-901.

58. Ljunggren, H.G., N.J. Stam, C. Ohlen, J.J. Neefjes, P. Hoglund, M.T. Heemels, J. Bastin, T.N. Schumacher, A. Townsend, K. Karre, and et al., Empty MHC class I molecules come out in the cold. Nature, 1990. 346(6283): p. 476-80.

59. Janeway, C.A., Jr. and K. Bottomly, Signals and signs for lymphocyte responses. Cell, 1994. 76(2): p. 275-85.

60. Mackay, C.R., Homing of naive, memory and effector lymphocytes. Curr Opin Immunol, 1993. 5(3): p. 423-7.

61. Bai, X.F., J.X. Gao, J. Liu, J. Wen, P. Zheng, and Y. Liu, On the site and mode of antigen presentation for the initiation of clonal expansion of CD8 T cells specific for a natural tumor antigen. Cancer Res, 2001. 61(18): p. 6860-7.

62. Heath, W.R. and F.R. Carbone, Cytotoxic T lymphocyte activation by cross-priming. Curr Opin Immunol, 1999. 11(3): p. 314-8.

63. Heath, W.R. and F.R. Carbone, Cross-presentation, dendritic cells, tolerance and immunity. Annu Rev Immunol, 2001. 19: p. 47-64.

64. Guermonprez, P., J. Valladeau, L. Zitvogel, C. Thery, and S. Amigorena, Antigen presentation and T cell stimulation by dendritic cells. Annu Rev Immunol, 2002. 20: p. 621-67.

65. Basu, S., R.J. Binder, T. Ramalingam, and P.K. Srivastava, CD91 is a common receptor for heat shock proteins gp96, hsp90, hsp70, and calreticulin. Immunity, 2001. 14(3): p. 303-13.

66. Li, Z., A. Menoret, and P. Srivastava, Roles of heat-shock proteins in antigen presentation and cross-presentation. Curr Opin Immunol, 2002. 14(1): p. 45-51.

67. Guermonprez, P., L. Saveanu, M. Kleijmeer, J. Davoust, P. Van Endert, and S. Amigorena, ER- phagosome fusion defines an MHC class I cross-presentation compartment in dendritic cells. Nature, 2003. 425(6956): p. 397-402.

68. Houde, M., S. Bertholet, E. Gagnon, S. Brunet, G. Goyette, A. Laplante, M.F. Princiotta, P. Thibault, D. Sacks, and M. Desjardins, Phagosomes are competent organelles for antigen cross- presentation. Nature, 2003. 425(6956): p. 402-6.

103

69. Abdel Motal, U.M., X. Zhou, A. Joki, A.R. Siddiqi, B.R. Srinivasa, K. Stenvall, J. Dahmen, and M. Jondal, Major histocompatibility complex class I-binding peptides are recycled to the cell surface after internalization. Eur J Immunol, 1993. 23(12): p. 3224-9.

70. Gromme, M., F.G. Uytdehaag, H. Janssen, J. Calafat, R.S. van Binnendijk, M.J. Kenter, A. Tulp, D. Verwoerd, and J. Neefjes, Recycling MHC class I molecules and endosomal peptide loading. Proc Natl Acad Sci U S A, 1999. 96(18): p. 10326-31.

71. Abdel Motal, U.M., C.L. Sentman, X. Zhou, P.J. Robinson, J. Dahmen, and M. Jondal, Glycosylphosphatidylinositol-linked Db does not induce an influenza-specific cytotoxic T lymphocyte response or recycle membrane-bound peptides. Eur J Immunol, 1995. 25(4): p. 1121- 4.

72. DeMars, R., R. Rudersdorf, C. Chang, J. Petersen, J. Strandtmann, N. Korn, B. Sidwell, and H.T. Orr, Mutations that impair a posttranscriptional step in expression of HLA-A and -B antigens. Proc Natl Acad Sci U S A, 1985. 82(23): p. 8183-7.

73. Salter, R.D. and P. Cresswell, Impaired assembly and transport of HLA-A and -B antigens in a mutant TxB cell hybrid. Embo J, 1986. 5(5): p. 943-9.

74. Townsend, A., C. Ohlen, J. Bastin, H.G. Ljunggren, L. Foster, and K. Karre, Association of class I major histocompatibility heavy and light chains induced by viral peptides. Nature, 1989. 340(6233): p. 443-8.

75. Deverson, E.V., I.R. Gow, W.J. Coadwell, J.J. Monaco, G.W. Butcher, and J.C. Howard, MHC class II region encoding proteins related to the multidrug resistance family of transmembrane transporters. Nature, 1990. 348(6303): p. 738-41.

76. Trowsdale, J., I. Hanson, I. Mockridge, S. Beck, A. Townsend, and A. Kelly, Sequences encoded in the class II region of the MHC related to the 'ABC' superfamily of transporters. Nature, 1990. 348(6303): p. 741-4.

77. Spies, T., M. Bresnahan, S. Bahram, D. Arnold, G. Blanck, E. Mellins, D. Pious, and R. DeMars, A gene in the human major histocompatibility complex class II region controlling the class I antigen presentation pathway. Nature, 1990. 348(6303): p. 744-7.

78. Monaco, J.J., S. Cho, and M. Attaya, Transport protein genes in the murine MHC: possible implications for antigen processing. Science, 1990. 250(4988): p. 1723-6.

79. Spies, T. and R. DeMars, Restored expression of major histocompatibility class I molecules by gene transfer of a putative peptide transporter. Nature, 1991. 351(6324): p. 323-4.

80. Powis, S.J., A.R. Townsend, E.V. Deverson, J. Bastin, G.W. Butcher, and J.C. Howard, Restoration of antigen presentation to the mutant cell line RMA-S by an MHC-linked transporter. Nature, 1991. 354(6354): p. 528-31.

81. Attaya, M., S. Jameson, C.K. Martinez, E. Hermel, C. Aldrich, J. Forman, K.F. Lindahl, M.J. Bevan, and J.J. Monaco, Ham-2 corrects the class I antigen-processing defect in RMA-S cells. Nature, 1992. 355(6361): p. 647-9.

82. Neefjes, J.J., F. Momburg, and G.J. Hammerling, Selective and ATP-dependent translocation of peptides by the MHC-encoded transporter. Science, 1993. 261(5122): p. 769-71.

104

83. Androlewicz, M.J., K.S. Anderson, and P. Cresswell, Evidence that transporters associated with antigen processing translocate a major histocompatibility complex class I-binding peptide into the endoplasmic reticulum in an ATP-dependent manner. Proc Natl Acad Sci U S A, 1993. 90(19): p. 9130-4.

84. Shepherd, J.C., T.N. Schumacher, P.G. Ashton-Rickardt, S. Imaeda, H.L. Ploegh, C.A. Janeway, Jr., and S. Tonegawa, TAP1-dependent peptide translocation in vitro is ATP dependent and peptide selective. Cell, 1993. 74(3): p. 577-84.

85. Van Kaer, L., P.G. Ashton-Rickardt, H.L. Ploegh, and S. Tonegawa, TAP1 mutant mice are deficient in antigen presentation, surface class I molecules, and CD4-8+ T cells. Cell, 1992. 71(7): p. 1205-14.

86. Wright, K.L., L.C. White, A. Kelly, S. Beck, J. Trowsdale, and J.P. Ting, Coordinate regulation of the human TAP1 and LMP2 genes from a shared bidirectional promoter. J Exp Med, 1995. 181(4): p. 1459-71.

87. Cramer, L.A., S.L. Nelson, and M.J. Klemsz, Synergistic induction of the Tap-1 gene by IFN- gamma and lipopolysaccharide in macrophages is regulated by STAT1. J Immunol, 2000. 165(6): p. 3190-7.

88. Min, W., J.S. Pober, and D.R. Johnson, Kinetically coordinated induction of TAP1 and HLA class I by IFN-gamma: the rapid induction of TAP1 by IFN-gamma is mediated by Stat1 alpha. J Immunol, 1996. 156(9): p. 3174-83.

89. White, L.C., K.L. Wright, N.J. Felix, H. Ruffner, L.F. Reis, R. Pine, and J.P. Ting, Regulation of LMP2 and TAP1 genes by IRF-1 explains the paucity of CD8+ T cells in IRF-1-/- mice. Immunity, 1996. 5(4): p. 365-76.

90. Matsuyama, T., T. Kimura, M. Kitagawa, K. Pfeffer, T. Kawakami, N. Watanabe, T.M. Kundig, R. Amakawa, K. Kishihara, A. Wakeham, and et al., Targeted disruption of IRF-1 or IRF-2 results in abnormal type I IFN gene induction and aberrant lymphocyte development. Cell, 1993. 75(1): p. 83-97.

91. Chatterjee-Kishore, M., R. Kishore, D.J. Hicklin, F.M. Marincola, and S. Ferrone, Different requirements for signal transducer and activator of transcription 1alpha and interferon regulatory factor 1 in the regulation of low molecular mass polypeptide 2 and transporter associated with antigen processing 1 gene expression. J Biol Chem, 1998. 273(26): p. 16177-83.

92. Borst, P. and R.O. Elferink, Mammalian ABC transporters in health and disease. Annu Rev Biochem, 2002. 71: p. 537-92.

93. Holland, I.B. and M.A. Blight, ABC-ATPases, adaptable energy generators fuelling transmembrane movement of a variety of molecules in organisms from bacteria to humans. J Mol Biol, 1999. 293(2): p. 381-99.

94. Vos, J.C., E.A. Reits, E. Wojcik-Jacobs, and J. Neefjes, Head-head/tail-tail relative orientation of the pore-forming domains of the heterodimeric ABC transporter TAP. Curr Biol, 2000. 10(1): p. 1-7.

105

95. Gileadi, U. and C.F. Higgins, Membrane topology of the ATP-binding cassette transporter associated with antigen presentation (Tap1) expressed in Escherichia coli. J Biol Chem, 1997. 272(17): p. 11103-8.

96. Vos, J.C., P. Spee, F. Momburg, and J. Neefjes, Membrane topology and dimerization of the two subunits of the transporter associated with antigen processing reveal a three-domain structure. J Immunol, 1999. 163(12): p. 6679-85.

97. Koch, J., R. Guntrum, S. Heintke, C. Kyritsis, and R. Tampe, Functional dissection of the transmembrane domains of the transporter associated with antigen processing (TAP). J Biol Chem, 2003.

98. Nijenhuis, M. and G.J. Hammerling, Multiple regions of the transporter associated with antigen processing (TAP) contribute to its peptide binding site. J Immunol, 1996. 157(12): p. 5467-77.

99. Nijenhuis, M., S. Schmitt, E.A. Armandola, R. Obst, J. Brunner, and G.J. Hammerling, Identification of a contact region for peptide on the TAP1 chain of the transporter associated with antigen processing. J Immunol, 1996. 156(6): p. 2186-95.

100. Neefjes, J., E. Gottfried, J. Roelse, M. Gromme, R. Obst, G.J. Hammerling, and F. Momburg, Analysis of the fine specificity of rat, mouse and human TAP peptide transporters. Eur J Immunol, 1995. 25(4): p. 1133-6.

101. Momburg, F., J. Roelse, G.J. Hammerling, and J.J. Neefjes, Peptide size selection by the major histocompatibility complex-encoded peptide transporter. J Exp Med, 1994. 179(5): p. 1613-23.

102. Heemels, M.T. and H.L. Ploegh, Substrate specificity of allelic variants of the TAP peptide transporter. Immunity, 1994. 1(9): p. 775-84.

103. Momburg, F., J. Roelse, J.C. Howard, G.W. Butcher, G.J. Hammerling, and J.J. Neefjes, Selectivity of MHC-encoded peptide transporters from human, mouse and rat. Nature, 1994. 367(6464): p. 648-51.

104. Schumacher, T.N., D.V. Kantesaria, M.T. Heemels, P.G. Ashton-Rickardt, J.C. Shepherd, K. Fruh, Y. Yang, P.A. Peterson, S. Tonegawa, and H.L. Ploegh, Peptide length and sequence specificity of the mouse TAP1/TAP2 translocator. J Exp Med, 1994. 179(2): p. 533-40.

105. Lauvau, G., K. Kakimi, G. Niedermann, M. Ostankovitch, P. Yotnda, H. Firat, F.V. Chisari, and P.M. van Endert, Human transporters associated with antigen processing (TAPs) select epitope precursor peptides for processing in the endoplasmic reticulum and presentation to T cells. J Exp Med, 1999. 190(9): p. 1227-40.

106. Knittler, M.R., P. Alberts, E.V. Deverson, and J.C. Howard, Nucleotide binding by TAP mediates association with peptide and release of assembled MHC class I molecules. Curr Biol, 1999. 9(18): p. 999-1008.

107. Karttunen, J.T., P.J. Lehner, S.S. Gupta, E.W. Hewitt, and P. Cresswell, Distinct functions and cooperative interaction of the subunits of the transporter associated with antigen processing (TAP). Proc Natl Acad Sci U S A, 2001. 98(13): p. 7431-6.

106

108. Lapinski, P.E., R.R. Neubig, and M. Raghavan, Walker A lysine mutations of TAP1 and TAP2 interfere with peptide translocation but not peptide binding. J Biol Chem, 2001. 276(10): p. 7526- 33.

109. Saveanu, L., S. Daniel, and P.M. van Endert, Distinct functions of the ATP binding cassettes of transporters associated with antigen processing: a mutational analysis of Walker A and B sequences. J Biol Chem, 2001. 276(25): p. 22107-13.

110. Neumann, L. and R. Tampe, Kinetic analysis of peptide binding to the TAP transport complex: evidence for structural rearrangements induced by substrate binding. J Mol Biol, 1999. 294(5): p. 1203-13.

111. Gorbulev, S., R. Abele, and R. Tampe, Allosteric crosstalk between peptide-binding, transport, and ATP hydrolysis of the ABC transporter TAP. Proc Natl Acad Sci U S A, 2001. 98(7): p. 3732- 7.

112. Lapinski, P.E., G.G. Miller, R. Tampe, and M. Raghavan, Pairing of the nucleotide binding domains of the transporter associated with antigen processing. J Biol Chem, 2000. 275(10): p. 6831-40.

113. Hopfner, K.P., A. Karcher, D.S. Shin, L. Craig, L.M. Arthur, J.P. Carney, and J.A. Tainer, Structural biology of Rad50 ATPase: ATP-driven conformational control in DNA double-strand break repair and the ABC-ATPase superfamily. Cell, 2000. 101(7): p. 789-800.

114. Loo, T.W., M.C. Bartlett, and D.M. Clarke, The "LSGGQ" motif in each nucleotide-binding domain of human P-glycoprotein is adjacent to the opposing walker A sequence. J Biol Chem, 2002. 277(44): p. 41303-6.

115. Loo, T.W., M.C. Bartlett, and D.M. Clarke, Drug binding in human P-glycoprotein causes conformational changes in both nucleotide-binding domains. J Biol Chem, 2003. 278(3): p. 1575- 8.

116. Smith, P.C., N. Karpowich, L. Millen, J.E. Moody, J. Rosen, P.J. Thomas, and J.F. Hunt, ATP binding to the motor domain from an ABC transporter drives formation of a nucleotide sandwich dimer. Mol Cell, 2002. 10(1): p. 139-49.

117. Logan, J., D. Hiestand, P. Daram, Z. Huang, D.D. Muccio, J. Hartman, B. Haley, W.J. Cook, and E.J. Sorscher, Cystic fibrosis transmembrane conductance regulator mutations that disrupt nucleotide binding. J Clin Invest, 1994. 94(1): p. 228-36.

118. Gregory, R.J., D.P. Rich, S.H. Cheng, D.W. Souza, S. Paul, P. Manavalan, M.P. Anderson, M.J. Welsh, and A.E. Smith, Maturation and function of cystic fibrosis transmembrane conductance regulator variants bearing mutations in putative nucleotide-binding domains 1 and 2. Mol Cell Biol, 1991. 11(8): p. 3886-93.

119. Li, C., M. Ramjeesingh, W. Wang, E. Garami, M. Hewryk, D. Lee, J.M. Rommens, K. Galley, and C.E. Bear, ATPase activity of the cystic fibrosis transmembrane conductance regulator. J Biol Chem, 1996. 271(45): p. 28463-8.

120. Derand, R., L. Bulteau-Pignoux, and F. Becq, The cystic fibrosis mutation G551D alters the non- Michaelis-Menten behavior of the cystic fibrosis transmembrane conductance regulator (CFTR)

107

channel and abolishes the inhibitory Genistein binding site. J Biol Chem, 2002. 277(39): p. 35999-6004.

121. Falcon-Perez, J.M., M.J. Mazon, J. Molano, and P. Eraso, Functional domain analysis of the yeast ABC transporter Ycf1p by site-directed mutagenesis. J Biol Chem, 1999. 274(33): p. 23584-90.

122. Ketchum, C.J., W.K. Schmidt, G.V. Rajendrakumar, S. Michaelis, and P.C. Maloney, The yeast a- factor transporter Ste6p, a member of the ABC superfamily, couples ATP hydrolysis to pheromone export. J Biol Chem, 2001. 276(31): p. 29007-11.

123. Browne, B.L., V. McClendon, and D.M. Bedwell, Mutations within the first LSGGQ motif of Ste6p cause defects in a-factor transport and mating in Saccharomyces cerevisiae. J Bacteriol, 1996. 178(6): p. 1712-9.

124. Schmees, G., A. Stein, S. Hunke, H. Landmesser, and E. Schneider, Functional consequences of mutations in the conserved 'signature sequence' of the ATP-binding-cassette protein MalK. Eur J Biochem, 1999. 266(2): p. 420-30.

125. Hewitt, E.W. and P.J. Lehner, The ABC-transporter signature motif is required for peptide translocation but not peptide binding by TAP. Eur J Immunol, 2003. 33(2): p. 422-7.

126. Bobadilla, J.L., M. Macek, Jr., J.P. Fine, and P.M. Farrell, Cystic fibrosis: a worldwide analysis of CFTR mutations--correlation with incidence data and application to screening. Hum Mutat, 2002. 19(6): p. 575-606.

127. Gadola, S.D., H.T. Moins-Teisserenc, J. Trowsdale, W.L. Gross, and V. Cerundolo, TAP deficiency syndrome. Clin Exp Immunol, 2000. 121(2): p. 173-8.

128. Levitskaya, J., M. Coram, V. Levitsky, S. Imreh, P.M. Steigerwald-Mullen, G. Klein, M.G. Kurilla, and M.G. Masucci, Inhibition of antigen processing by the internal repeat region of the Epstein-Barr virus nuclear antigen-1. Nature, 1995. 375(6533): p. 685-8.

129. Hill, A., P. Jugovic, I. York, G. Russ, J. Bennink, J. Yewdell, H. Ploegh, and D. Johnson, Herpes simplex virus turns off the TAP to evade host immunity. Nature, 1995. 375(6530): p. 411-5.

130. Galocha, B., A. Hill, B.C. Barnett, A. Dolan, A. Raimondi, R.F. Cook, J. Brunner, D.J. McGeoch, and H.L. Ploegh, The of ICP47, a herpes simplex virus-encoded inhibitor of the major histocompatibility complex (MHC)-encoded peptide transporter associated with antigen processing (TAP), maps to the NH2-terminal 35 residues. J Exp Med, 1997. 185(9): p. 1565-72.

131. Tomazin, R., A.B. Hill, P. Jugovic, I. York, P. van Endert, H.L. Ploegh, D.W. Andrews, and D.C. Johnson, Stable binding of the herpes simplex virus ICP47 protein to the peptide binding site of TAP. Embo J, 1996. 15(13): p. 3256-66.

132. Ahn, K., A. Gruhler, B. Galocha, T.R. Jones, E.J. Wiertz, H.L. Ploegh, P.A. Peterson, Y. Yang, and K. Fruh, The ER-luminal domain of the HCMV glycoprotein US6 inhibits peptide translocation by TAP. Immunity, 1997. 6(5): p. 613-21.

133. Hengel, H., J.O. Koopmann, T. Flohr, W. Muranyi, E. Goulmy, G.J. Hammerling, U.H. Koszinowski, and F. Momburg, A viral ER-resident glycoprotein inactivates the MHC-encoded peptide transporter. Immunity, 1997. 6(5): p. 623-32.

108

134. van der Wal, F.J., M. Kikkert, and E. Wiertz, The HCMV gene products US2 and US11 target MHC class I molecules for degradation in the cytosol. Curr Top Microbiol Immunol, 2002. 269: p. 37-55.

135. Wiertz, E.J., T.R. Jones, L. Sun, M. Bogyo, H.J. Geuze, and H.L. Ploegh, The human cytomegalovirus US11 gene product dislocates MHC class I heavy chains from the endoplasmic reticulum to the cytosol. Cell, 1996. 84(5): p. 769-79.

136. Boname, J.M. and P.G. Stevenson, MHC class I ubiquitination by a viral PHD/LAP finger protein. Immunity, 2001. 15(4): p. 627-36.

137. Stevenson, P.G., S. Efstathiou, P.C. Doherty, and P.J. Lehner, Inhibition of MHC class I-restricted antigen presentation by gamma 2-herpesviruses. Proc Natl Acad Sci U S A, 2000. 97(15): p. 8455-60.

138. Yu, Y.Y., M.R. Harris, L. Lybarger, L.A. Kimpler, N.B. Myers, H.W.t. Virgin, and T.H. Hansen, Physical association of the K3 protein of gamma-2 herpesvirus 68 with major histocompatibility complex class I molecules with impaired peptide and beta(2)-microglobulin assembly. J Virol, 2002. 76(6): p. 2796-803.

139. Lybarger, L., X. Wang, M.R. Harris, H.W.t. Virgin, and T.H. Hansen, Virus subversion of the MHC class I peptide-loading complex. Immunity, 2003. 18(1): p. 121-30.

140. Bennett, E.M., J.R. Bennink, J.W. Yewdell, and F.M. Brodsky, Cutting edge: adenovirus E19 has two mechanisms for affecting class I MHC expression. J Immunol, 1999. 162(9): p. 5049-52.

141. Andersson, M., S. Paabo, T. Nilsson, and P.A. Peterson, Impaired intracellular transport of class I MHC antigens as a possible means for adenoviruses to evade immune surveillance. Cell, 1985. 43(1): p. 215-22.

142. Burgert, H.G. and S. Kvist, An adenovirus type 2 glycoprotein blocks cell surface expression of human histocompatibility class I antigens. Cell, 1985. 41(3): p. 987-97.

143. Jones, T.R., E.J. Wiertz, L. Sun, K.N. Fish, J.A. Nelson, and H.L. Ploegh, Human cytomegalovirus US3 impairs transport and maturation of major histocompatibility complex class I heavy chains. Proc Natl Acad Sci U S A, 1996. 93(21): p. 11327-33.

144. Ahn, K., A. Angulo, P. Ghazal, P.A. Peterson, Y. Yang, and K. Fruh, Human cytomegalovirus inhibits antigen presentation by a sequential multistep process. Proc Natl Acad Sci U S A, 1996. 93(20): p. 10990-5.

145. Schwartz, O., V. Marechal, S. Le Gall, F. Lemonnier, and J.M. Heard, Endocytosis of major histocompatibility complex class I molecules is induced by the HIV-1 Nef protein. Nat Med, 1996. 2(3): p. 338-42.

146. Seliger, B., T. Cabrera, F. Garrido, and S. Ferrone, HLA class I antigen abnormalities and immune escape by malignant cells. Semin Cancer Biol, 2002. 12(1): p. 3-13.

147. Hicklin, D.J., F.M. Marincola, and S. Ferrone, HLA class I antigen downregulation in human cancers: T-cell immunotherapy revives an old story. Mol Med Today, 1999. 5(4): p. 178-86.

109

148. Garrido, F., F. Ruiz-Cabello, T. Cabrera, J.J. Perez-Villar, M. Lopez-Botet, M. Duggan-Keen, and P.L. Stern, Implications for immunosurveillance of altered HLA class I phenotypes in human tumours. Immunol Today, 1997. 18(2): p. 89-95.

149. Redondo, M., J. Garca, E. Villar, I. Rodrigo, E. Perea-Milla, A. Serrano, and M. Morell, Major histocompatibility complex status in breast carcinogenesis and relationship to apoptosis. Hum Pathol, 2003. 34(12): p. 1283-9.

150. Andratschke, M., C. Pauli, M. Stein, S. Chaubal, and B. Wollenberg, MHC-class I antigen expression on micrometastases in bone marrow of patients with head and neck squamous cell cancer. Anticancer Res, 2003. 23(2B): p. 1467-71.

151. Ryu, K.S., Y.S. Lee, B.K. Kim, Y.G. Park, Y.W. Kim, S.Y. Hur, T.E. Kim, I.K. Kim, and J.W. Kim, Alterations of HLA class I and II antigen expression in preinvasive, invasive and metastatic cervical cancers. Exp Mol Med, 2001. 33(3): p. 136-44.

152. Grandis, J.R., D.M. Falkner, M.F. Melhem, W.E. Gooding, S.D. Drenning, and P.A. Morel, Human leukocyte antigen class I allelic and haplotype loss in squamous cell carcinoma of the head and neck: clinical and immunogenetic consequences. Clin Cancer Res, 2000. 6(7): p. 2794- 802.

153. Kageshita, T., S. Hirai, T. Ono, D.J. Hicklin, and S. Ferrone, Down-regulation of HLA class I antigen-processing molecules in malignant melanoma: association with disease progression. Am J Pathol, 1999. 154(3): p. 745-54.

154. Blades, R.A., P.J. Keating, L.J. McWilliam, N.J. George, and P.L. Stern, Loss of HLA class I expression in prostate cancer: implications for immunotherapy. Urology, 1995. 46(5): p. 681-6; discussion 686-7.

155. Cromme, F.V., P.F. van Bommel, J.M. Walboomers, M.P. Gallee, P.L. Stern, P. Kenemans, T.J. Helmerhorst, M.J. Stukart, and C.J. Meijer, Differences in MHC and TAP-1 expression in cervical cancer lymph node metastases as compared with the primary tumours. Br J Cancer, 1994. 69(6): p. 1176-81.

156. Goepel, J.R., R.C. Rees, K. Rogers, C.J. Stoddard, W.E. Thomas, and L. Shepherd, Loss of monomorphic and polymorphic HLA antigens in metastatic breast and colon carcinoma. Br J Cancer, 1991. 64(5): p. 880-3.

157. Zia, A., F.W. Schildberg, and I. Funke, MHC class I negative phenotype of disseminated tumor cells in bone marrow is associated with poor survival in R0M0 breast cancer patients. Int J Cancer, 2001. 93(4): p. 566-70.

158. Concha, A., F. Esteban, T. Cabrera, F. Ruiz-Cabello, and F. Garrido, Tumor aggressiveness and MHC class I and II antigens in laryngeal and breast cancer. Semin Cancer Biol, 1991. 2(1): p. 47- 54.

159. Lollini, P.L., G. Nicoletti, L. Landuzzi, C. De Giovanni, I. Rossi, E. Di Carlo, P. Musiani, W.J. Muller, and P. Nanni, Down regulation of major histocompatibility complex class I expression in mammary carcinoma of HER-2/neu transgenic mice. Int J Cancer, 1998. 77(6): p. 937-41.

110

160. Milling, S.W., W.K. Silvers, T. Sai, and B. Mintz, Decline in MHC class I expression with increasing thickness of cutaneous melanomas in standard-strain transgenic mouse models. Melanoma Res, 2002. 12(3): p. 221-30.

161. Zheng, P., S. Sarma, Y. Guo, and Y. Liu, Two mechanisms for tumor evasion of preexisting cytotoxic T-cell responses: lessons from recurrent tumors. Cancer Res, 1999. 59(14): p. 3461-7.

162. Rosenberg, S.A., Progress in human tumour immunology and immunotherapy. Nature, 2001. 411(6835): p. 380-4.

163. Wang, R.F. and S.A. Rosenberg, Human tumor antigens for cancer vaccine development. Immunol Rev, 1999. 170: p. 85-100.

164. Dudley, M.E. and S.A. Rosenberg, Adoptive-cell-transfer therapy for the treatment of patients with cancer. Nat Rev Cancer, 2003. 3(9): p. 666-75.

165. Ridolfi, R., E. Flamini, A. Riccobon, F. De Paola, R. Maltoni, A. Gardini, L. Ridolfi, L. Medri, G. Poletti, and D. Amadori, Adjuvant adoptive immunotherapy with tumour-infiltrating lymphocytes and modulated doses of interleukin-2 in 22 patients with melanoma, colorectal and renal cancer, after radical metastasectomy, and in 12 advanced patients. Cancer Immunol Immunother, 1998. 46(4): p. 185-93.

166. Lee, D.S., D.E. White, R. Hurst, S.A. Rosenberg, and J.C. Yang, Patterns of relapse and response to retreatment in patients with metastatic melanoma or renal cell carcinoma who responded to interleukin-2-based immunotherapy. Cancer J Sci Am, 1998. 4(2): p. 86-93.

167. Restifo, N.P., F.M. Marincola, Y. Kawakami, J. Taubenberger, J.R. Yannelli, and S.A. Rosenberg, Loss of functional beta 2-microglobulin in metastatic melanomas from five patients receiving immunotherapy. J Natl Cancer Inst, 1996. 88(2): p. 100-8.

168. Rivoltini, L., K.C. Barracchini, V. Viggiano, Y. Kawakami, A. Smith, A. Mixon, N.P. Restifo, S.L. Topalian, T.B. Simonis, S.A. Rosenberg, and et al., Quantitative correlation between HLA class I allele expression and recognition of melanoma cells by antigen-specific cytotoxic T lymphocytes. Cancer Res, 1995. 55(14): p. 3149-57.

169. Ruiz-Cabello, F., T. Cabrera, M.A. Lopez-Nevot, and F. Garrido, Impaired surface antigen presentation in tumors: implications for T cell-based immunotherapy. Semin Cancer Biol, 2002. 12(1): p. 15-24.

170. Propper, D.J., D. Chao, J.P. Braybrooke, P. Bahl, P. Thavasu, F. Balkwill, H. Turley, N. Dobbs, K. Gatter, D.C. Talbot, A.L. Harris, and T.S. Ganesan, Low-dose IFN-gamma induces tumor MHC expression in metastatic malignant melanoma. Clin Cancer Res, 2003. 9(1): p. 84-92.

171. Restifo, N.P., F. Esquivel, Y. Kawakami, J.W. Yewdell, J.J. Mule, S.A. Rosenberg, and J.R. Bennink, Identification of human cancers deficient in antigen processing. J Exp Med, 1993. 177(2): p. 265-72.

172. Johnsen, A., J. France, M.S. Sy, and C.V. Harding, Down-regulation of the transporter for antigen presentation, proteasome subunits, and class I major histocompatibility complex in tumor cell lines. Cancer Res, 1998. 58(16): p. 3660-7.

111

173. Ritz, U., F. Momburg, H. Pilch, C. Huber, M.J. Maeurer, and B. Seliger, Deficient expression of components of the MHC class I antigen processing machinery in human cervical carcinoma. Int J Oncol, 2001. 19(6): p. 1211-20.

174. Zheng, P., Y. Guo, Q. Niu, D.E. Levy, J.A. Dyck, S. Lu, L.A. Sheiman, and Y. Liu, Proto- oncogene PML controls genes devoted to MHC class I antigen presentation. Nature, 1998. 396(6709): p. 373-6.

175. Garcia-Lora, A., M. Martinez, I. Algarra, J.J. Gaforio, and F. Garrido, MHC class I-deficient metastatic tumor variants immunoselected by T lymphocytes originate from the coordinated downregulation of APM components. Int J Cancer, 2003. 106(4): p. 521-7.

176. Chen, H.L., D. Gabrilovich, A. Virmani, I. Ratnani, K.R. Girgis, S. Nadaf-Rahrov, M. Fernandez- Vina, and D.P. Carbone, Structural and functional analysis of beta2 microglobulin abnormalities in human lung and breast cancer. Int J Cancer, 1996. 67(6): p. 756-63.

177. Yoon, S.J., J.O. Kang, J.S. Park, N.K. Kim, and D.S. Heo, Reduced expression of MHC class I antigen in human cancer cell lines with defective LMP-7. Anticancer Res, 2000. 20(2A): p. 949- 53.

178. Singal, D.P., M. Ye, and X. Qiu, Molecular basis for lack of expression of HLA class I antigens in human small-cell lung carcinoma cell lines. Int J Cancer, 1996. 68(5): p. 629-36.

179. Singal, D.P., M. Ye, J. Ni, and D.P. Snider, Markedly decreased expression of TAP1 and LMP2 genes in HLA class I-deficient human tumor cell lines. Immunol Lett, 1996. 50(3): p. 149-54.

180. Seliger, B., A. Hohne, A. Knuth, H. Bernhard, B. Ehring, R. Tampe, and C. Huber, Reduced membrane major histocompatibility complex class I density and stability in a subset of human renal cell carcinomas with low TAP and LMP expression. Clin Cancer Res, 1996. 2(8): p. 1427- 33.

181. Seliger, B., K. Schreiber, K. Delp, M. Meissner, S. Hammers, T. Reichert, K. Pawlischko, R. Tampe, and C. Huber, Downregulation of the constitutive tapasin expression in human tumor cells of distinct origin and its transcriptional upregulation by cytokines. Tissue Antigens, 2001. 57(1): p. 39-45.

182. McEvoy, C.R., R. Seshadri, A.A. Morley, and F.A. Firgaira, Frequency and genetic basis of MHC, beta-2-microglobulin and MEMO-1 loss of heterozygosity in sporadic breast cancer. Tissue Antigens, 2002. 60(3): p. 235-43.

183. Cabrera, C.M., P. Jimenez, T. Cabrera, C. Esparza, F. Ruiz-Cabello, and F. Garrido, Total loss of MHC class I in colorectal tumors can be explained by two molecular pathways: beta2- microglobulin inactivation in MSI-positive tumors and LMP7/TAP2 downregulation in MSI- negative tumors. Tissue Antigens, 2003. 61(3): p. 211-9.

184. Geertsen, R., R. Boni, R. Blasczyk, P. Romero, D. Betts, D. Rimoldi, X. Hong, E. Laine, J. Willers, and R. Dummer, Loss of single HLA class I allospecificities in melanoma cells due to selective genomic abbreviations. Int J Cancer, 2002. 99(1): p. 82-7.

185. Mendez, R., A. Serrano, E. Jager, I. Maleno, F. Ruiz-Cabello, A. Knuth, and F. Garrido, Analysis of HLA class I expression in different metastases from two melanoma patients undergoing peptide immunotherapy. Tissue Antigens, 2001. 57(6): p. 508-19.

112

186. Wang, Z., B. Seliger, N. Mike, F. Momburg, A. Knuth, and S. Ferrone, Molecular analysis of the HLA-A2 antigen loss by melanoma cells SK-MEL-29.1.22 and SK-MEL-29.1.29. Cancer Res, 1998. 58(10): p. 2149-57.

187. Maleno, I., M.A. Lopez-Nevot, T. Cabrera, J. Salinero, and F. Garrido, Multiple mechanisms generate HLA class I altered phenotypes in laryngeal carcinomas: high frequency of HLA haplotype loss associated with loss of heterozygosity in chromosome region 6p21. Cancer Immunol Immunother, 2002. 51(7): p. 389-96.

188. Rigaud, G., P.S. Moore, D. Taruscio, M. Scardoni, M. Montresor, F. Menestrina, and A. Scarpa, Alteration of chromosome arm 6p is characteristic of primary mediastinal B-cell lymphoma, as identified by genome-wide allelotyping. Genes Chromosomes Cancer, 2001. 31(2): p. 191-5.

189. Maeurer, M.J., S.M. Gollin, W.J. Storkus, W. Swaney, J. Karbach, D. Martin, C. Castelli, R. Salter, A. Knuth, and M.T. Lotze, Tumor escape from immune recognition: loss of HLA-A2 melanoma cell surface expression is associated with a complex rearrangement of the short arm of chromosome 6. Clin Cancer Res, 1996. 2(4): p. 641-52.

190. Snyder, S.R., J. Wang, J.F. Waring, and G.D. Ginder, Identification of CCAAT displacement protein (CDP/cut) as a locus-specific repressor of major histocompatibility complex gene expression in human tumor cells. J Biol Chem, 2001. 276(7): p. 5323-30.

191. Blanchet, O., J.F. Bourge, H. Zinszner, A. Israel, P. Kourilsky, J. Dausset, L. Degos, and P. Paul, Altered binding of regulatory factors to HLA class I enhancer sequence in human tumor cell lines lacking class I antigen expression. Proc Natl Acad Sci U S A, 1992. 89(8): p. 3488-92.

192. Smirnov, D.A., S. Hou, X. Liu, E. Claudio, U.K. Siebenlist, and R.P. Ricciardi, Coup-TFII is up- regulated in adenovirus type 12 tumorigenic cells and is a repressor of MHC class I transcription. Virology, 2001. 284(1): p. 13-9.

193. Ye, X., A. Kralli, R. Ge, R.P. Ricciardi, and B.B. Knowles, Down-regulation of MHC class I antigen in insulinoma cells controlled by the R1 element of the H-2 enhancer. Oncogene, 1994. 9(4): p. 1195-204.

194. Yamit-Hezi, A., D. Plaksin, and L. Eisenbach, c-fos and c-jun overexpression in malignant cells reduces their tumorigenic and metastatic potential, and affects their MHC class I gene expression. Oncogene, 1994. 9(4): p. 1065-79.

195. van 't Veer, L.J., R.L. Beijersbergen, and R. Bernards, N-myc suppresses major histocompatibility complex class I gene expression through down-regulation of the p50 subunit of NF-kappa B. Embo J, 1993. 12(1): p. 195-200.

196. Peltenburg, L.T., R. Dee, and P.I. Schrier, Downregulation of HLA class I expression by c-myc in human melanoma is independent of enhancer A. Nucleic Acids Res, 1993. 21(5): p. 1179-85.

197. Terrazzano, G., M.F. Romano, M.C. Turco, S. Salzano, A. Ottaiano, S. Venuta, S. Fontana, C. Manzo, S. Zappacosta, and E. Carbone, HLA class I antigen downregulation by interleukin (IL)-10 is predominantly governed by NK-kappaB in the short term and by TAP1+2 in the long term. Tissue Antigens, 2000. 55(4): p. 326-32.

113

198. Yue, F.Y., R. Dummer, R. Geertsen, G. Hofbauer, E. Laine, S. Manolio, and G. Burg, Interleukin- 10 is a growth factor for human melanoma cells and down-regulates HLA class-I, HLA class-II and ICAM-1 molecules. Int J Cancer, 1997. 71(4): p. 630-7.

199. Salazar-Onfray, F., J. Charo, M. Petersson, S. Freland, G. Noffz, Z. Qin, T. Blankenstein, H.G. Ljunggren, and R. Kiessling, Down-regulation of the expression and function of the transporter associated with antigen processing in murine tumor cell lines expressing IL-10. J Immunol, 1997. 159(7): p. 3195-202.

200. Serrano, A., S. Tanzarella, I. Lionello, R. Mendez, C. Traversari, F. Ruiz-Cabello, and F. Garrido, Rexpression of HLA class I antigens and restoration of antigen-specific CTL response in melanoma cells following 5-aza-2'-deoxycytidine treatment. Int J Cancer, 2001. 94(2): p. 243-51.

201. Nie, Y., G. Yang, Y. Song, X. Zhao, C. So, J. Liao, L.D. Wang, and C.S. Yang, DNA hypermethylation is a mechanism for loss of expression of the HLA class I genes in human esophageal squamous cell carcinomas. Carcinogenesis, 2001. 22(10): p. 1615-23.

202. Nava, G., R. Ocadiz, V. Ortega, and G. Alfaro, Damage in B2m genes and DNA methylation of H- 2 genes are involved in loss of expression of class I MHC products on the membrane of LR.4, a cell line derivative of the T-cell lymphoma L5178Y. Eur J Immunogenet, 1992. 19(3): p. 141-58.

203. Rubocki, R.J., B.E. Berrigan, S.L. Speaks, and J.L. Wisecarver, Locus-specific de novo methylation down-regulates MHC class I in S49 lymphomas. Immunogenetics, 1996. 43(1-2): p. 63-7.

204. Wang, Z., F.M. Marincola, L. Rivoltini, G. Parmiani, and S. Ferrone, Selective histocompatibility leukocyte antigen (HLA)-A2 loss caused by aberrant pre-mRNA splicing in 624MEL28 melanoma cells. J Exp Med, 1999. 190(2): p. 205-15.

205. Hicklin, D.J., Z. Wang, F. Arienti, L. Rivoltini, G. Parmiani, and S. Ferrone, beta2-Microglobulin mutations, HLA class I antigen loss, and tumor progression in melanoma. J Clin Invest, 1998. 101(12): p. 2720-9.

206. Koopman, L.A., A.R. van Der Slik, M.J. Giphart, and G.J. Fleuren, Human leukocyte antigen class I gene mutations in cervical cancer. J Natl Cancer Inst, 1999. 91(19): p. 1669-77.

207. Seliger, B., M.J. Maeurer, and S. Ferrone, TAP off--tumors on. Immunol Today, 1997. 18(6): p. 292-9.

208. Kamarashev, J., S. Ferrone, B. Seifert, R. Boni, F.O. Nestle, G. Burg, and R. Dummer, TAP1 down-regulation in primary melanoma lesions: an independent marker of poor prognosis. Int J Cancer, 2001. 95(1): p. 23-8.

209. Gostout, B.S., G.A. Poland, E.S. Calhoun, Y.R. Sohni, R.L. Giuntoli, 2nd, R.M. McGovern, J.A. Sloan, S.S. Cha, and D.H. Persing, TAP1, TAP2, and HLA-DR2 alleles are predictors of cervical cancer risk. Gynecol Oncol, 2003. 88(3): p. 326-32. 210. Evans, M., L.K. Borysiewicz, A.S. Evans, M. Rowe, M. Jones, U. Gileadi, V. Cerundolo, and S. Man, Antigen processing defects in cervical carcinomas limit the presentation of a CTL epitope from human papillomavirus 16 E6. J Immunol, 2001. 167(9): p. 5420-8.

114

211. Johnsen, A.K., D.J. Templeton, M. Sy, and C.V. Harding, Deficiency of transporter for antigen presentation (TAP) in tumor cells allows evasion of immune surveillance and increases tumorigenesis. J Immunol, 1999. 163(8): p. 4224-31.

212. Dovhey, S.E., N.S. Ghosh, and K.L. Wright, Loss of interferon-gamma inducibility of TAP1 and LMP2 in a renal cell carcinoma cell line. Cancer Res, 2000. 60(20): p. 5789-96.

213. Chen, H.L., D. Gabrilovich, R. Tampe, K.R. Girgis, S. Nadaf, and D.P. Carbone, A functionally defective allele of TAP1 results in loss of MHC class I antigen presentation in a human lung cancer. Nat Genet, 1996. 13(2): p. 210-3.

214. Hicklin, D.J., T. Kageshita, and S. Ferrone, Development and characterization of rabbit antisera to human MHC-linked transporters associated with antigen processing. Tissue Antigens, 1996. 48(1): p. 38-46.

215. Greenberg, M.E. and E.B. Ziff, Stimulation of 3T3 cells induces transcription of the c-fos proto- oncogene. Nature, 1984. 311(5985): p. 433-8.

216. Zhou, Q., K. Rammohan, S. Lin, N. Robinson, O. Li, X. Liu, X.F. Bai, L. Yin, B. Scarberry, P. Du, M. You, K. Guan, P. Zheng, and Y. Liu, CD24 is a genetic modifier for risk and progression of multiple sclerosis. Proc Natl Acad Sci U S A, 2003.

217. Romero, P., P.R. Dunbar, D. Valmori, M. Pittet, G.S. Ogg, D. Rimoldi, J.L. Chen, D. Lienard, J.C. Cerottini, and V. Cerundolo, Ex vivo staining of metastatic lymph nodes by class I major histocompatibility complex tetramers reveals high numbers of antigen-experienced tumor-specific cytolytic T lymphocytes. J Exp Med, 1998. 188(9): p. 1641-50.

218. Lee, P.P., C. Yee, P.A. Savage, L. Fong, D. Brockstedt, J.S. Weber, D. Johnson, S. Swetter, J. Thompson, P.D. Greenberg, M. Roederer, and M.M. Davis, Characterization of circulating T cells specific for tumor-associated antigens in melanoma patients. Nat Med, 1999. 5(6): p. 677-85.

219. Seliger, B., C. Harders, S. Lohmann, F. Momburg, S. Urlinger, R. Tampe, and C. Huber, Down- regulation of the MHC class I antigen-processing machinery after oncogenic transformation of murine fibroblasts. Eur J Immunol, 1998. 28(1): p. 122-33.

220. Seliger, B., U. Ritz, R. Abele, M. Bock, R. Tampe, G. Sutter, I. Drexler, C. Huber, and S. Ferrone, Immune escape of melanoma: first evidence of structural alterations in two distinct components of the MHC class I antigen processing pathway. Cancer Res, 2001. 61(24): p. 8647-50.

221. Rotem-Yehudar, R., M. Groettrup, A. Soza, P.M. Kloetzel, and R. Ehrlich, LMP-associated proteolytic activities and TAP-dependent peptide transport for class 1 MHC molecules are suppressed in cell lines transformed by the highly oncogenic adenovirus 12. J Exp Med, 1996. 183(2): p. 499-514.

222. Rotem-Yehudar, R., H. Shechter, and R. Ehrlich, Transcriptional regulation of class-I major histocompatibility complex genes transformed in murine cells is mediated by positive and negative regulatory elements. Gene, 1994. 144(2): p. 265-70. 223. Rotem-Yehudar, R., S. Winograd, S. Sela, J.E. Coligan, and R. Ehrlich, Downregulation of peptide transporter genes in cell lines transformed with the highly oncogenic adenovirus 12. J Exp Med, 1994. 180(2): p. 477-88.

115

224. Ahn, K., T.H. Meyer, S. Uebel, P. Sempe, H. Djaballah, Y. Yang, P.A. Peterson, K. Fruh, and R. Tampe, Molecular mechanism and species specificity of TAP inhibition by herpes simplex virus ICP47. Embo J, 1996. 15(13): p. 3247-55.

225. Kallfelz, M., D. Jung, C. Hilmes, A. Knuth, E. Jaeger, C. Huber, and B. Seliger, Induction of immunogenicity of a human renal-cell carcinoma cell line by TAP1-gene transfer. Int J Cancer, 1999. 81(1): p. 125-33.

226. Singal, D.P., M. Ye, and D. Bienzle, Transfection of TAP 1 gene restores HLA class I expression in human small-cell lung carcinoma. Int J Cancer, 1998. 75(1): p. 112-6.

227. Liu, H.X., S.L. Chew, L. Cartegni, M.Q. Zhang, and A.R. Krainer, Exonic splicing enhancer motif recognized by human SC35 under splicing conditions. Mol Cell Biol, 2000. 20(3): p. 1063-71.

228. Liu, H.X., L. Cartegni, M.Q. Zhang, and A.R. Krainer, A mechanism for exon skipping caused by nonsense or missense mutations in BRCA1 and other genes. Nat Genet, 2001. 27(1): p. 55-8.

229. Lozano, F., B. Maertzdorf, R. Pannell, and C. Milstein, Low cytoplasmic mRNA levels of immunoglobulin kappa light chain genes containing nonsense codons correlate with inefficient splicing. Embo J, 1994. 13(19): p. 4617-22.

230. Mullner, E.W. and L.C. Kuhn, A stem-loop in the 3' untranslated region mediates iron-dependent regulation of receptor mRNA stability in the cytoplasm. Cell, 1988. 53(5): p. 815-25.

231. Mullner, E.W., B. Neupert, and L.C. Kuhn, A specific mRNA binding factor regulates the iron- dependent stability of cytoplasmic transferrin receptor mRNA. Cell, 1989. 58(2): p. 373-82.

232. Zhou, Q., Y. Guo, and Y. Liu, Regulation of the stability of heat-stable antigen mRNA by interplay between two novel cis elements in the 3' untranslated region. Mol Cell Biol, 1998. 18(2): p. 815- 26.

233. Thermann, R., G. Neu-Yilik, A. Deters, U. Frede, K. Wehr, C. Hagemeier, M.W. Hentze, and A.E. Kulozik, Binary specification of nonsense codons by splicing and cytoplasmic translation. Embo J, 1998. 17(12): p. 3484-94.

234. Zhang, J., X. Sun, Y. Qian, J.P. LaDuca, and L.E. Maquat, At least one intron is required for the nonsense-mediated decay of triosephosphate mRNA: a possible link between nuclear splicing and cytoplasmic translation. Mol Cell Biol, 1998. 18(9): p. 5272-83.

235. Brocke, K.S., G. Neu-Yilik, N.H. Gehring, M.W. Hentze, and A.E. Kulozik, The human intronless melanocortin 4-receptor gene is NMD insensitive. Hum Mol Genet, 2002. 11(3): p. 331-5.

236. Kim, V.N., N. Kataoka, and G. Dreyfuss, Role of the nonsense-mediated decay factor hUpf3 in the splicing-dependent exon-exon junction complex. Science, 2001. 293(5536): p. 1832-6.

237. Lykke-Andersen, J., M.D. Shu, and J.A. Steitz, Human Upf proteins target an mRNA for nonsense-mediated decay when bound downstream of a termination codon. Cell, 2000. 103(7): p. 1121-31. 238. Lykke-Andersen, J., M.D. Shu, and J.A. Steitz, Communication of the position of exon-exon junctions to the mRNA surveillance machinery by the protein RNPS1. Science, 2001. 293(5536): p. 1836-9.

116

239. Rajavel, K.S. and E.F. Neufeld, Nonsense-mediated decay of human HEXA mRNA. Mol Cell Biol, 2001. 21(16): p. 5512-9.

240. McCutcheon, J.A., J. Gumperz, K.D. Smith, C.T. Lutz, and P. Parham, Low HLA-C expression at cell surfaces correlates with increased turnover of heavy chain mRNA. J Exp Med, 1995. 181(6): p. 2085-95.

241. Ackerman, A.L., C. Kyritsis, R. Tampe, and P. Cresswell, Early phagosomes in dendritic cells form a cellular compartment sufficient for cross presentation of exogenous antigens. Proc Natl Acad Sci U S A, 2003. 100(22): p. 12889-94.

242. Rich, D.P., R.J. Gregory, M.P. Anderson, P. Manavalan, A.E. Smith, and M.J. Welsh, Effect of deleting the R domain on CFTR-generated chloride channels. Science, 1991. 253(5016): p. 205-7.

243. Yang, T., B.A. McNally, S. Ferrone, Y. Liu, and P. Zheng, A single-nucleotide deletion leads to rapid degradation of TAP-1 mRNA in a melanoma cell line. J Biol Chem, 2003. 278(17): p. 15291-6.

244. Maeurer, M.J., S.M. Gollin, D. Martin, W. Swaney, J. Bryant, C. Castelli, P. Robbins, G. Parmiani, W.J. Storkus, and M.T. Lotze, Tumor escape from immune recognition: lethal recurrent melanoma in a patient associated with downregulation of the peptide transporter protein TAP-1 and loss of expression of the immunodominant MART-1/Melan-A antigen. J Clin Invest, 1996. 98(7): p. 1633-41.

245. Cabrera, T., A. Collado, M.A. Fernandez, A. Ferron, J. Sancho, F. Ruiz-Cabello, and F. Garrido, High frequency of altered HLA class I phenotypes in invasive colorectal carcinomas. Tissue Antigens, 1998. 52(2): p. 114-23.

246. Tang, J., D.O. Freedman, S. Allen, E. Karita, R. Musonda, C. Braga, J. Margolick, and R.A. Kaslow, TAPI polymorphisms in several human ethnic groups: characteristics, evolution, and genotyping strategies. Hum Immunol, 2001. 62(3): p. 256-68.

247. Ramal, L.M., I. Maleno, T. Cabrera, A. Collado, A. Ferron, M.A. Lopez-Nevot, and F. Garrido, Molecular strategies to define HLA haplotype loss in microdissected tumor cells. Hum Immunol, 2000. 61(10): p. 1001-12.

117