• No results found

The role of the ubiquitin system in human cytomegalovirus-mediated degradation of MHC class I heavy chains

N/A
N/A
Protected

Academic year: 2021

Share "The role of the ubiquitin system in human cytomegalovirus-mediated degradation of MHC class I heavy chains"

Copied!
25
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

The role of the ubiquitin system in human cytomegalovirus-mediated

degradation of MHC class I heavy chains

Hassink, G.C.

Citation

Hassink, G. C. (2006, May 22). The role of the ubiquitin system in human

cytomegalovirus-mediated degradation of MHC class I heavy chains. Retrieved from

https://hdl.handle.net/1887/4414

Version:

Corrected Publisher’s Version

(2)

CHAPTER

1

I

NTRODUCTI

ON

(3)
(4)

VIRAL IMMUNE EVASION AND QUALITY CONTROL

OF THE ENDOPLASMIC RETICULUM

Protein degradation and MHC class I-restricted antigen

presentation

Duri

ng

t

hei

r

l

i

fecycl

e,

prot

ei

ns

are

prone

t

o

agi

ng

as

a

resul

t

of

di

fferent

cel

l

ul

ar

st

resses

and

prot

ei

n

oxi

dat

i

on.

Cel

l

s

renew

t

hei

r

prot

ei

n

popul

at

i

on

by

means

of

const

ant

prot

ei

n

synt

hesi

s

and

degradat

i

on

1

.

Al

t

hough

many

cel

l

ul

ar

organel

l

es

have

t

he

means

of

prot

ei

n

degradat

i

on,

t

he

prot

easome

pl

ays

a

domi

nant

rol

e

i

n

t

hi

s

process

2

.

The

prot

easome

i

s

formed

by

a

l

arge

barrel

-shaped

20S

prot

ei

n

compl

ex

consi

st

i

ng

of

four

ri

ngs

of

t

wo

t

i

mes

seven

al

pha

and

t

wo

t

i

mes

seven

bet

a

subuni

t

s

resi

di

ng

i

n

t

he

cyt

osol

and

nucl

eus

3

.

The

prot

easome

can

ei

t

her

be

i

n

compl

ex

wi

t

h

t

wo

11S

caps,

whi

ch

faci

l

i

t

at

es

t

he

accel

erat

ed

degradat

i

on

of

l

arge

pept

i

de

fragment

s

4,5

,

or

i

n

compl

ex

wi

t

h

t

wo

19S

caps

on

bot

h

si

des,

whi

ch

faci

l

i

t

at

es

t

he

degradat

i

on

of

whol

e

prot

ei

ns

6-8

.

Bot

h

t

he

19S

and

t

he

11S

caps

have

ATPase

act

i

vi

t

y.

The

19S

cap

t

oget

her

wi

t

h

t

he

20S

prot

easome

core

forms

a

26S

compl

ex.

Thi

s

compl

ex

recogni

zes

pol

y-ubi

qui

t

i

n

t

rees

t

hat

are

at

t

ached

t

o

prot

ei

n

subst

rat

es

(

see

l

at

er

on

i

n

t

hi

s

i

nt

roduct

i

on)

and

removes

t

hem

9-11

.

At

t

he

same

t

i

me,

t

he

19S

cap

unfol

ds

t

he

subst

rat

e

and

aqui

res

access

t

o

t

he

port

s

of

t

he

out

er

ri

ng

of

t

he

prot

easome

12-14

.

In

t

he

core

of

t

he

prot

easome,

a

set

of

t

rypsi

n-,

chemot

rypsi

n-

and

post

-gl

ut

amyl

pept

i

dyl

hydrol

yt

i

c-l

i

ke

prot

eases

cl

eave

t

he

unfol

ded

prot

ei

ns

i

nt

o

ol

i

gopept

i

des

wi

t

h

an

average

l

engt

h

of

4

t

o

20

ami

no

aci

ds

1,15

.

These

short

pept

i

de

fragment

s

l

eave

t

he

prot

easome

and

are

furt

her

processed

by

endopept

i

dase

t

o

ret

ri

eve

t

he

ori

gi

nal

ami

no

aci

ds

16

.

A

smal

l

number

of

pept

i

de

fragment

s,

however,

are

used

for

an

i

mport

ant

defense

syst

em

of

t

he

organi

sm.

(5)

solubilize the unfolded proteins, attract and coordinate the binding of

additional chaperones, and form disulfide bridges, respectively (figure 1).

Shortly after this process has been completed, the specialized

chaperone tapasin mediates the binding of the newly synthesized MHC class I

complexes to the transporter associated with antigen presentation (TAP). In

addition, the luminal lectin calreticulin

18

and Ƣ2m bind the heavy chain,

together forming the MHC class I peptide-loading complex

19

. The

proteasome-derived peptides are transported into the ER via the TAP

complex

20

. In the ER, the N-terminal ends of the peptides are trimmed to

8-10 residues

21-23

, in order to fit into the newly synthesized MHC class I

molecules

24,25

. Recently the ER localized exopeptidase ERAAP was found to

be required for this trimming

26

. Only high affinity binding peptides allow

further maturation and transport of MHC class I molecules to the cell surface

27-30

.

At the cell surface, the protein complex is screened by another

specialized polymorphic molecule, the T-cell receptor, which is present on

cytotoxic T-lymphocytes, among others. The polymorphic TCRs are required

to recognize all the possible peptides presented by the MHC class I molecules.

A process of clonal selection ensures that only T-cells that fail to recognize

peptides derived from “self-proteins” are able to survive their maturation into

T-cells. W hen in addition to the MHC class I-peptide-TCR interaction, the

appropriate co-stimulatory signals are present, a signaling cascade is induced

in the T-cell resulting in the release of cytotoxic granules, which will ultimately

result in the death of the target cell in a process called apoptosis. Viruses have,

however, developed counter measures against this immune strategy.

Figure 1. M odel for translocation and formation of the M H C class I complex.

(6)

Life-long CMV infection and its risks

Human cytomegalovirus belongs to the herpes viridae, which are

common among all vertebrates. Cytomegaloviruses have a very large double

stranded DNA genome of 125-235 kbp. An important feature of these viruses

is that they are able to produce a life-long infection. The infection is usually

without symptoms as a long co-evolution of host and virus has resulted in a

delicate balance;

the host’

s immune system constantly represses upsurges of

viral expansion, while the virus continuously avoids detection by the host.

Yet CMV is a well-known pathogen. It is associated with acute and

chronic rejection of transplanted organs

31-33

and has also been implicated as a

causative agent of atherosclerosis

34,35

. Furthermore, it is known to cause

congenital defects in fetuses of pregnant women who have had a primary

CMV infection

36-38

. Finally, CMV frequently causes serious disease in

immuno-compromised individuals. When the immune system of the host is

suppressed in periods of stress or fatigue, the virus may activate and use the

metabolism of the cells to proliferate. During sustained immuno-suppression,

e.g. in HIV-infected individuals or individuals using immuno-suppressive

agents to prevent rejection of transplanted tissue, CMV can cause hepatitis,

encephalitis, or pneumonia, all of which may be fatal

39,40

.

To study the mechanisms of the HCMV-induced pathology, rat and

mouse CMV are often used as models. Mouse CMV (MCMV) is used to study

retinitis

41

, vasculitis

42

, pneumonitis

43

, and the role of the immune system in

the pathogenesis of human CMV in general

44

. Rat CMV (RCMV), on the

other hand, is used particularly in models of hart transplant-induced vascular

scleroses

45-47

, lung transplant-induced obliterative bronchiolitis

48

, and in the

study of diabetes

49

.

(7)

Viral immune escape by targeting of MHC class I molecules is not

restricted to human CMV; cytomegaloviruses infecting other mammalians

also interfere with MHC class I expression. Mouse CMV encodes three

proteins affecting MHC class I expression at the cell surface. The m152 gene

product prevents export of class I complexes from the post-ER/early Golgi

66,67

. A second protein, the m06 gene product, induces lysosomal degradation

of the MHC complex after a transient interaction in the ER

68

. A third gene

product, the m04-encoded gp34, associates with properly folded MHC class I

molecules in the ER. The resulting complex is transported to the cell

membrane where it may modulate the function of NK cells

69,70

. Together,

these MCMV-encoded immune evasion genes appear to counteract MHC

class I restricted T cell activation

71,72

. Although RCMV also establishes life

long infection, the genes facilitating this persistence are yet to be identified.

HCMV, MCMV and RCMV all encode homologues of MHC class I heavy

chains

73-75

, which may deviate NK cell

74,76

, and B-Cell- and

macrophage-responses

77

. Although RCMV has been found to downregulate MHC class I

expression

78

, the viral genes underlying this phenotype remain to be

identified.

ER protein quality control and degradation of ER proteins

HCMV US2 and US11-mediated dislocation of MHC class I molecules to the cytosol

The discovery that US2 and US11-dependent degradation of MHC

class I heavy chains takes place at the level of the ER, suggested that US2 and

US11 used a pre-existing ER degradation pathway to downregulate MHC

class I molecules at the cell’s surface

54,55

. This pathway was shown to dispose

of misfolded ER proteins like CFTR

79,80

, CPY*

81

, and ste6

82

. Since the

discovery that ER proteins can in fact be retro-translocated into the cytosol

where they are degraded by the proteasome, substantial effort has been made

to elucidate the precise mechanism behind this process.

(8)

in the nucleus

83

. Therefore, protein expression is tightly controlled in

eukaryotic cells. Prolonged production of misfolded proteins in the ER will

induce the translation of chaperones

84

, reduce translation of mRNA’s to

relieve protein folding machineries

85

, and may eventually even induce

apoptosis

86

. This process has been called the unfolded protein response

(UPR).

Amino acid chains produced by the ribosome are recognized as

nascent ER proteins when they contain either one or more transmembrane

domains, a signal sequence, or a signal anchor. As soon as the ribosome

recognizes one of those domains in the newly synthesized amino acid chain,

further translation is halted until the ribosome has bound to the signal

recognition particle adjacent to a conduit for translocation in the ER which is

called the translocon (figure 2) (reviewed in

87,88

). The core of the translocon

consists of a complex of the subunits Sec61ơ, Ƣ, and ƣ, which together form a

50 Ångström pore. The translocon can be equipped with a translocation-

associated membrane protein (TRAM), which is thought to facilitate a lateral

opening in the wall of this pore through which newly-translated

transmembrane-regions (TM) can move laterally into the lipid bi-layer

89-91

.

Domain structure and polartity of TM-adjacent residues ultimately determine

the orientation of TM regions in the membrane

92-96

. Adjacent to the cytosolic

side of the complex is the oligosaccharyltransferase (OST), which provides the

core glycan to be attached to asparagine residues present in NXT/S motifs

(where X is any amino acid except proline). Besides these complexes, a fourth

comlex, the translocon-associated complex (TRAP) can be found in complex

with the translocon and enhances translocation of proteins that have

prolonged acces to the cytoplasm

97

. Recently, a closer analysis using electron

micrograph analysis (EMAN), shows that the ribosome actually binds four

Sec61 complexes and two TRAP complexes at the same time

98

. Only one of

the Sec61 complexes, however, seems to be active in translocation

98

.

(9)

In the ER, the N-linked glycans play an important role in protein

folding and quality control, as has been shown for yeast alpha1-antitrypsin

and mammalian IgM heavy chains, and many others

113,114

. Glucosidases I and

II remove glucose moieties from GlcNAc2Man9Glc3-containing

glycoproteins to form GlcNAc2Man9 proteins. The membrane-bound

calnexin and soluble calreticulin recognize GlcNAc2Man9Glc1 sugar moieties

on the unfolded proteins and allow protein disulfide isomerase (PDI) ERP57

to bind these complexes. The PDI can now establish disulfide bridges within

the newly synthesized molecule and between subunits of protein complexes

115,116

. When all the glucose residues have been removed and the protein has

folded correctly, it is promoted further into the secretion pathway, the Golgi

network, where the N-linked glycans are processed to larger structures. If the

complex has not acquired the correct conformation, however, it is recognized

by UDP-glucose:

glycoprotein glucosyltransferase (UGTR) which reattaches a

glucose residue to the N-linked glycan, thus enhancing the affinity for

calnexin and calreticulin to allow another round of folding

113,117

. At the same

time, there is a chance that ER manosidase I will remove a mannose residue,

forming GlcNAc2Man8Glc1

113

. The GlcNAc2Man8Glc1 glycan is a poor

substrate for glucosidase II. Therefore, it does not remove the last glucose

residue, and the protein is not released to the Golgi

118,119

. Prolonged ER

retention eventually leads to the degradation of ER proteins.

This mannose clock is just one of the many checkpoints in the

maturation of proteins. A variety of specialized ER-associated compartments

are implied in quality control

120,121

. Furthermore, en route to the cell surface,

misfolded or mal-complexed proteins are degraded in a pathway involving

lysomes

122,123

, the membrane-enveloped organelles responsible for

Figure 2. Model of the ribosome channel complex in a sheet of ER membrane.

(10)

degradation of the bulk of endocytosed material.

Retro-translocation of proteins from the ER to the cytosol (‘Dislocation’)

ER-resident proteins that do not pass the ER quality control are

destined for degradation by the proteasome which resides in the cytosol. How

exactly proteins are transported from the ER to the cytosol, a mechanism

termed retro-translocation or dislocation, is an intriguing question which is

not well understood (figure 3). The first step for ER proteins destined for

degradation, is the transport of these substrates through a conduit which may

be formed by components of the Sec61 complex

54,124-126

. Once in the cytosol,

any N-linked glycans are removed by a recently identified N-glycanase

127,128

.

Ultimately, the dislocated substrates are provided with polyubiquitin chains

and degraded by the proteasome. The energy that drives dislocation may

come from several sides. The translocon itself is a passive conduit.

Dislocation, however, involves ER luminal ATP

129

and several cytolic factors,

namely the proteasome, the p97-ufd-npl4 AAA ATPase complex, and the

ubiquitin system.

Figure 3. Translocation of a protein from the ER into the cytosol.

(11)

The role of the proteasome in dislocation of proteins from the ER

A proportion of the 20S and 26S proteasomes have been found to be

associated with the ER membrane

130-132

. Involvement of the proteasome in

the extraction of proteins from the ER membrane is suggested because the

ER-to-cytosol dislocation of many ER substrates can be prevented by the

removal or inhibition of proteasomes

126,133-136

. On the other hand, US2 and

US11-dependent dislocation of MHC class I heavy chains

54,55

, and the

dislocation of TCR-ơ

133

, have been clearly shown to occur in spite of

proteasomal inhibition.

The Role of the p97 complex

The AAA ATPases form an abundant family, of which the members

are involved in various cellular functions, such as membrane trafficking,

proteasome function, organelle biogenesis, and microtubule regulation

(reviewed in

137,138

). Their modes of action are not well understood

139

, but one

of their functions is the ability to unfold molecules. For example, a ring of six

AAA ATPases comprise the regulatory lid in of the 20S proteasome in

eukaryotes

140

where they might drive unfolding of substrates before entrance

of the proteolitic core

12-14

. Two other AAA ATPase rings, N-ethylmaleimide

sensitive factor (NSF) and the p97-p47 complex, drive vesicle fusion and

transport by dissociating soluble NSF attachment protein receptor (SNARE)

complexes which are otherwise stable up to temperatures of 90 ºC

141

. Yet

another function may be that they act as microtubule-based motors

142

.

A role for AAA ATPase p97 in the degradadation of ER proteins was

established when Ye et al. discovered that MHC class I heavy chain

dislocation in yeast and US11-dependent dislocation of MHC class I heavy

chains was dependent on the vasolin-containing protein(VCP)/p97 in

complex with Ufd1 and Npl4

143

. The P97 protein constitutes 1% of the

cytosol and it is, therefore, one of the most abundant proteins in the cell

144

.

Every p97 molecule harbors two ATPase domains which form two hexameric

rings stacked on top of one another

145,146

. Exactly how p97-Ufd1-Npl4

extracts the polypeptides from the ER membrane during dislocation is unclear

147

.

(12)

has remained unclear, however, whether the actual dislocation from the ER is

also dependent on ubiquitination

148

.

Attaching ubiquitin to a protein involves three steps, each providing

the versatility required to regulate the cellular process in question (Figure 4).

The first step is to activate the ubiquitin itself in order to prepare it for

attachment. This is accomplished by the E1 ubiquitin activating enzyme. The

amount of activated ubiquitin that is present in the cell ultimately limits the

many processes that depend on ubiquitination. E2 conjugating enzymes

subsequently bind the activated ubiquitin on a conserved cysteine residue in

the second step. It is thought that an E2 enzyme, probably in association with

an E3 enzyme, determines how the ubiquitin molecules will be attached to

one another and the substrate. A single, mono-ubiquitin attachment, or a

poly-ubiquitin chain linked at a certain lysine in the ubiquitin molecule is thus

formed on the target protein. Ubiquitin molecules can form isopeptide

linkages between the C-terminal glycine residue and one of the lysines at

positions 6, 11, 29, 48 or 63 on the next ubiquitin molecule. K63-based

ubiquitin linkages have been associated with DNA repair, INB kinase

activation and endocytosis, whereas K48 is mostly involved in proteasomal

degradation. K11 and K29 have also been associated with protein degradation

by the proteasome (reviewed in

149,150

). The E3 ubiquitin protein ligase

enzymes accomplish the third step in the process. They bind the substrate and

facilitate the actual transfer of one or more ubiquitin molecules from the E2

enzyme to a lysine or the N-terminus of the selected protein.

Ubiquitin does not only have a role in proteasomal degradation but

also in a large number of other cellular functions. DNA repair in the nucleus,

regulation of translation, activation of transcription factors and kinases, and

proteasomal degradation all require ubiquitination. Protein sorting and

trafficking within the secretory system of the cell is also regulated by

(13)

Moreover, it has become clear in the past few years that ubiquitin also

plays a role in the degradation of ER proteins (See also chapter 3). Hrd1p of

S. cerevisiae was the first ER-resident E3 ubiquitin ligase to be identified as

having a direct role in the degradation of ER proteins

158,159

. Hrd1p/Der3p of

S. cerevisiae was identified independently by Hampton and co-workers

160

and

by Wolf and co-workers

161

. Hrd1p/Der3p functions as a ubiquitin ligase in

the degradation of S. cerevisiae Hmg2p, one of the yeast isozymes of

3-hydroxy-3-methylglutaryl-coenzyme A reductase (HMGR)

158,160

.

Hrd1p/Der3p is also involved in degradation of other ER proteins,

including CPY* and sec61-2p

162,163

. Hrd1p/Der3p is a multispanning

membrane protein with six transmembrane domains, and its C-terminal

RING-H2 domain is located in the cytosol

159

. In yeast, Hrd1p is found in a

stoichiometric complex with Hrd3p, a lumen-oriented ER membrane protein

that stabilizes Hrd1p and modulates its ligase activity

164

. Hrd3p is, therefore,

essential for a functional Hrd1p E3 ligase in yeast. Hrd1p acts with the E2

enzymes Ubc1p and Ubc7p, of which Ubc7p is more important in the

degradation of ER proteins

158

. Initially, it was thought that Hrd1p might be

the central E3 ligase that served the degradation of yeast ER proteins in

general, together with Hrd1 associates Hrd3p and Der1p, and the E2’s Ubc1p

and Ubc7p

158

. However, while a number of substrates depend on Hrd1p for

their degradation from the ER, a number of others do not

165,166

.

Besides Hrd1p, two other E3 ligases dedicated to degradation of ER

Figure 4. Ubiquitin cascade.

(14)

proteins have been identified in yeast to date, Rsp5p and Doa10. Human

CFTR degradation has recently been shown to depend on Hrd1p in yeast, as

well as on Doa10, illustrating that the different E3 ligases active in

degradation of ER proteins, may cooperate to create a versatile degradation

machinery

167

. In mammalians, two homologues of Hrd1p have been

identified: AMFR/gp78

168

and HRD1

169

. AMFR facilitates the degradation

of apoliprotein B100, whereas HRD1 facilitates the degradation of TCR-ơ

168,169

. Additionally, HRD1 and AMFR both facilitate the degradation of

CD3-Ƥ

168,169

, stressing the versatility of the system.

Outline of this thesis

It is clear from the above that the degradation of ER (glyco-)proteins is

a tightly regulated and complicated process involving a diversity of players not

necessarily dedicated to ER-associated degradation alone. The discovery that

viruses may co-opt this process to evade detection by the immune system has

significantly boosted research in this direction and forms the basis for this

thesis.

In this thesis we explored the mechanism by which the

HCMV-encoded proteins US2 and US11 alter the expression of MHC class I

molecules in order to evade MHC class I-dependent recognition by T-cells. In

the second chapter it was investigated whether, like human CMV and mouse

CMV, rat CMV also evades the host’s immune system using cell surface

downregulation of the MHC class I molecules.

In the third chapter we studied the role of the ubiquitin system in the

CMV US11-dependent dislocation of the MHC class I heavy chains. As

US11-mediated dislocation was indeed found to depend on a functional

ubiquitin system, a number of E3 ligases potentially involved in the

degradation of ER proteins were characterized. One of these potential E3

ligases is HRD1, a yet uncharacterized human homologue of the yeast Hrd1p.

In chapter four we characterized this homologue as an ER-resident ubiquitin

ligase involved in the degradation of TCR-ơ and CD3-Ƥ. Another potential E3

ligase is TEB4, a yet uncharacterized human homologue of the yeast Doa10.

This E3 was characterized in chapter five as an ER-resident ubiquitin ligase.

The crucial role of the ubiquitin system in protein dislocation suggested

that ubiquitination of lysines within the target proteins would be required to

facilitate dislocation. In chapter six we investigated whether lysines in the

MHC class I molecules themselves were indeed required for the dislocation of

these proteins in the presence of US2 and US11. Interestingly, this was not

the case. Nevertheless, a functional ubiquitin system was required for

(15)

on the role of the ubiquitin system in the dislocation process.

In chapter seven, we investigated whether Ƣ2m association and folding

of MHC class I heavy chains influenced US2 and US11-dependent

dislocation. Finally, in chapter eight we studied the class I dislocation process

in more detail. In particular, we investigated the order in which the amino acid

chains of MHC class I molecules were transported through the dislocon,

another important but hitherto unexplored aspect of the dislocation pathway.

References

1 Yewdell, J., Reits, E. and Neefjes, J. (2003) Making sense of mass destruction: quantitating MHC class I antigen presentation. Nat.Rev.Immunol. 3, 952-961

2 Fenteany, G., Standaert, R., Lane, W., Choi, S., Corey, E. and Schreiber, S. (1995) Inhibition of proteasome activities and subunit-specific amino-terminal threonine modification by lactacystin. Science 268, 726-731

3 Groll, M., Ditzel, L., Lowe, J., Stock, D., Bochtler, M., Bartunik, H. and Huber, R. (1997) Structure of 20S proteasome from yeast at 2.4 A resolution. Nature 386, 463-471

4 Dubiel, W., Pratt, G., Ferrell, K. and Rechsteiner, M. (1992) Purification of an 11 S regulator of the multicatalytic protease. J.Biol.Chem. 267, 22369-22377

5 Ma, C., Willy, P., Slaughter, C. and DeMartino, G. (1993) PA28, an activator of the 20 S proteasome, is inactivated by proteolytic modification at its carboxyl terminus. J.Biol.Chem. 268, 22514-22519

6 Orlowski, M. (1990) The multicatalytic proteinase complex, a major extralysosomal proteolytic system. Biochemistry 29, 10289-10297

7 Rechsteiner, M., Hoffman, L. and Dubiel, W. (1993) The multicatalytic and 26 S proteases. J.Biol.Chem. 268, 6065-6068

8 Hilt, W. and Wolf, D. (1996) Proteasomes: destruction as a programme. Trends Biochem.Sci. 21, 96-102

9 Eytan, E., Armon, T., Heller, H., Beck, S. and Hershko, A. (1993) Ubiquitin C-terminal hydrolase activity associated with the 26 S protease complex. J.Biol.Chem. 268, 4668-4674

10 Glickman, M., Rubin, D., Coux, O., Wefes, I., Pfeifer, G., Cjeka, Z., Baumeister, W., Fried, V. and Finley, D. (1998) A subcomplex of the proteasome regulatory particle required for ubiquitin-conjugate degradation and related to the COP9-signalosome and eIF3. Cell 94, 615-623 11 Guterman, A. and Glickman, M. (2004) Complementary roles for Rpn11 and Ubp6 in

deubiquitination and proteolysis by the proteasome. J.Biol.Chem. 279, 1729-1738

12 Benaroudj, N., Zwickl, P., Seemuller, E., Baumeister, W. and Goldberg, A. (2003) ATP hydrolysis by the proteasome regulatory complex PAN serves multiple functions in protein degradation. Mol.Cell 11, 69-78

13 Navon, A. and Goldberg, A. (2001) Proteins are unfolded on the surface of the ATPase ring before transport into the proteasome. Mol.Cell 8, 1339-1349

(16)

15 Ciechanover, A. (1998) The ubiquitin-proteasome pathway: on protein death and cell life. EMBO J. 17, 7151-7160

16 Reits, E., Griekspoor, A., Neijssen, J., Groothuis, T., Jalink, K., van Veelen, P., Janssen, H., Calafat, J., Drijfhout, J. and Neefjes, J. (2003) Peptide diffusion, protection, and degradation in nuclear and cytoplasmic compartments before antigen presentation by MHC class I. Immunity. 18, 97-108 17 Zinkernagel, R. (1974) Restriction by H-2 gene complex of transfer of cell-mediated immunity to

Listeria monocytogenes. Nature 251, 230-233

18 Gao, B., Adhikari, R., Howarth, M., Nakamura, K., Gold, M., Hill, A., Knee, R., Michalak, M. and Elliott, T. (2002) Assembly and antigen-presenting function of MHC class I molecules in cells lacking the ER chaperone calreticulin. Immunity. 16, 99-109

19 Cresswell, P. (2000) Intracellular surveillance: controlling the assembly of MHC class I-peptide complexes. Traffic. 1, 301-305

20 Rudd, P., Elliott, T., Cresswell, P., Wilson, I. and Dwek, R. (2001) Glycosylation and the immune system. Science 291, 2370-2376

21 Serwold, T., Gaw, S. and Shastri, N. (2001) ER aminopeptidases generate a unique pool of peptides for MHC class I molecules. Nat.Immunol. 2, 644-651

22 Brouwenstijn, N., Serwold, T. and Shastri, N. (2001) MHC class I molecules can direct proteolytic cleavage of antigenic precursors in the endoplasmic reticulum. Immunity. 15, 95-104

23 Fruci, D., Niedermann, G., Butler, R. and van Endert, P. (2001) Efficient MHC class I-independent amino-terminal trimming of epitope precursor peptides in the endoplasmic reticulum. Immunity. 15, 467-476

24 Van Bleek, G. and Nathenson, S. (1990) Isolation of an endogenously processed immunodominant viral peptide from the class I H-2Kb molecule. Nature 348, 213-216

25 Yewdell, J. and Bennink, J. (1992) Cell biology of antigen processing and presentation to major histocompatibility complex class I molecule-restricted T lymphocytes. Adv.Immunol. 52, 1-123 26 Serwold, T., Gonzalez, F., Kim, J., Jacob, R. and Shastri, N. (2002) ERAAP customizes peptides

for MHC class I molecules in the endoplasmic reticulum. Nature 419, 480-483

27 Townsend, A., Ohlen, C., Bastin, J., Ljunggren, H., Foster, L. and Karre, K. (1989) Association of class I major histocompatibility heavy and light chains induced by viral peptides. Nature 340 , 443-448

28 Townsend, A., Elliott, T., Cerundolo, V., Foster, L., Barber, B. and Tse, A. (1990) Assembly of MHC class I molecules analyzed in vitro. Cell 62, 285-295

29 Paulsson, K., Wang, P., Anderson, P., Chen, S., Pettersson, R. and Li, S. (2001) Distinct differences in association of MHC class I with endoplasmic reticulum proteins in wild-type, and beta 2-microglobulin- and TAP-deficient cell lines. Int.Immunol. 13, 1063-1073

30 Williams, A., Peh, C., Purcell, A., McCluskey, J. and Elliott, T. (2002) Optimization of the MHC class I peptide cargo is dependent on tapasin. Immunity. 16, 509-520

31 Almond, P., Matas, A., Gillingham, K., Dunn, D., Payne, W., Gores, P., Gruessner, R. and Najarian, J. (1993) Predictors of chronic rejection in renal transplant recipients. Transplant.Proc. 25 , 936-

(17)

33 McLaughlin, K., Wu, C., Fick, G., Muirhead, N., Hollomby, D. and Jevnikar, A. (2002) Cytomegalovirus seromismatching increases the risk of acute renal allograft rejection. Transplantation 74, 813-816

34 Epstein, S., Speir, E., Zhou, Y., Guetta, E., Leon, M. and Finkel, T. (1996) The role of infection in restenosis and atherosclerosis: focus on cytomegalovirus. Lancet 348 Suppl 1, s13-s17

35 Grahame-Clarke, C., Chan, N., Andrew, D., Ridgway, G., Betteridge, D., Emery, V., Colhoun, H. and Vallance, P. (2003) Human cytomegalovirus seropositivity is associated with impaired vascular function. Circulation 108, 678-683

36 Altshuler, G. (1974) Pathogenesis of congenital herpesvirus infection. Case report including a description of the placenta. Am.J.Dis.Child 127, 427-429

37 Pass, R., Stagno, S., Myers, G. and Alford, C. (1980) Outcome of symptomatic congenital cytomegalovirus infection: results of long-term longitudinal follow-up. Pediatrics 66, 758-762 38 Gaytant, M., Steegers, E., Semmekrot, B., Merkus, H. and Galama, J. (2002) Congenital

cytomegalovirus infection: review of the epidemiology and outcome. Obstet.Gynecol.Surv. 57, 245-256

39 Ljungman, P. (1996) Cytomegalovirus infections in transplant patients. Scand.J.Infect.Dis.Suppl 100, 59-63

40 Ramsay, M., Miller, E. and Peckham, C. (1991) Outcome of confirmed symptomatic congenital cytomegalovirus infection. Arch.Dis.Child 66, 1068-1069

41 Atherton, S., Newell, C., Kanter, M. and Cousins, S. (1991) Retinitis in euthymic mice following inoculation of murine cytomegalovirus (MCMV) via the supraciliary route. Curr.Eye Res. 10, 667-677

42 Hamamdzic, D., Harley, R., Hazen-Martin, D. and LeRoy, E. (2001) MCMV induces neointima in IFN-gammaR-/- mice: intimal cell apoptosis and persistent proliferation of myofibroblasts. BMC.Musculoskelet.Disord. 2, 3

43 Tanaka, K., Noda, S., Sawamura, S., Kabir, A. and Koga, Y. (2001) Nitric oxide targets bronchiolar epithelial cells in murine cytomegalovirus-associated disease in lungs that are free of the virus. Arch.Virol. 146, 1499-1515

44 Hengel, H., Reusch, U., Gutermann, A., Ziegler, H., Jonjic, S., Lucin, P. and Koszinowski, U. (1999) Cytomegaloviral control of MHC class I function in the mouse. Immunol.Rev. 168, 167-176 45 Streblow, D., Kreklywich, C., Smith, P., Soule, J., Meyer, C., Yin, M., Beisser, P., Vink, C., Nelson,

J. and Orloff, S. (2005) Rat cytomegalovirus-accelerated transplant vascular sclerosis is reduced with mutation of the chemokine-receptor R33. Am.J.Transplant. 5, 436-442

46 Zhou, Y., Leon, M., Waclawiw, M., Popma, J., Yu, Z., Finkel, T. and Epstein, S. (1996) Association between prior cytomegalovirus infection and the risk of restenosis after coronary atherectomy. N.Engl.J.Med. 335, 624-630

47 Zhou, Y., Shou, M., Harrell, R., Yu, Z., Unger, E. and Epstein, S. (2000) Chronic non-vascular cytomegalovirus infection: effects on the neointimal response to experimental vascular injury. Cardiovasc.Res. 45, 1019-1025

48 Tikkanen, J., Krebs, R., Bruggeman, C., Lemstrom, K. and Koskinen, P. (2004) Platelet-derived growth factor regulates cytomegalovirus infection-enhanced obliterative bronchiolitis in rat tracheal allografts. Transplantation 77, 655-658

49 van der, Werf, Hillebrands, J., Klatter, F., Bos, I., Bruggeman, C. and Rozing, J. (2003)

(18)

50 Beersma, M., Bijlmakers, M. and Ploegh, H. (1993) Human cytomegalovirus down-regulates HLA class I expression by reducing the stability of class I H chains. J.Immunol. 151, 4455-4464 51 Warren, A., Ducroq, D., Lehner, P. and Borysiewicz, L. (1994) Human cytomegalovirus-infected

cells have unstable assembly of major histocompatibility complex class I complexes and are resistant to lysis by cytotoxic T lymphocytes. J.Virol. 68, 2822-2829

52 Jones, T., Hanson, L., Sun, L., Slater, J., Stenberg, R. and Campbell, A. (1995) Multiple independent loci within the human cytomegalovirus unique short region down-regulate expression of major histocompatibility complex class I heavy chains. J.Virol. 69, 4830-4841

53 Jones, T. and Sun, L. (1997) Human cytomegalovirus US2 destabilizes major histocompatibility complex class I heavy chains. J.Virol. 71, 2970-2979

54 Wiertz, E., Tortorella, D., Bogyo, M., Yu, J., Mothes, W., Jones, T., Rapoport, T. and Ploegh, H. (1996) Sec61-mediated transfer of a membrane protein from the endoplasmic reticulum to the proteasome for destruction. Nature 384, 432-438

55 Wiertz, E., Jones, T., Sun, L., Bogyo, M., Geuze, H. and Ploegh, H. (1996) The human cytomegalovirus US11 gene product dislocates MHC class I heavy chains from the endoplasmic reticulum to the cytosol. Cell 84, 769-779

56 Jun, Y., Kim, E., Jin, M., Sung, H., Han, H., Geraghty, D. and Ahn, K. (2000) Human cytomegalovirus gene products US3 and US6 down-regulate trophoblast class I MHC molecules. J.Immunol. 164, 805-811

57 Ahn, K., Angulo, A., Ghazal, P., Peterson, P., Yang, Y. and Fruh, K. (1996) Human cytomegalovirus inhibits antigen presentation by a sequential multistep process. Proc.Natl.Acad.Sci.U.S.A 93, 10990-10995

58 Park, B., Kim, Y., Shin, J., Lee, S., Cho, K., Fruh, K., Lee, S. and Ahn, K. (2004) Human cytomegalovirus inhibits tapasin-dependent peptide loading and optimization of the MHC class I peptide cargo for immune evasion. Immunity. 20, 71-85

59 Jones, T., Wiertz, E., Sun, L., Fish, K., Nelson, J. and Ploegh, H. (1996) Human cytomegalovirus US3 impairs transport and maturation of major histocompatibility complex class I heavy chains. Proc.Natl.Acad.Sci.U.S.A 93, 11327-11333

60 Lee, S., Yoon, J., Park, B., Jun, Y., Jin, M., Sung, H., Kim, I., Kang, S., Choi, E., Ahn, B. and Ahn, K. (2000) Structural and functional dissection of human cytomegalovirus US3 in binding major histocompatibility complex class I molecules. J.Virol. 74, 11262-11269

61 Ahn, K., Gruhler, A., Galocha, B., Jones, T., Wiertz, E., Ploegh, H., Peterson, P., Yang, Y. and Fruh, K. (1997) The ER-luminal domain of the HCMV glycoprotein US6 inhibits peptide translocation by TAP. Immunity. 6, 613-621

62 Hengel, H., Koopmann, J., Flohr, T., Muranyi, W., Goulmy, E., Hammerling, G., Koszinowski, U. and Momburg, F. (1997) A viral ER-resident glycoprotein inactivates the MHC-encoded peptide transporter. Immunity. 6, 623-632

63 Hewitt, E., Gupta, S. and Lehner, P. (2001) The human cytomegalovirus gene product US6 inhibits ATP binding by TAP. EMBO J. 20, 387-396

64 Kyritsis, C., Gorbulev, S., Hutschenreiter, S., Pawlitschko, K., Abele, R. and Tampe, R. (2001) Molecular mechanism and structural aspects of transporter associated with antigen processing inhibition by the cytomegalovirus protein US6. J.Biol.Chem. 276, 48031-48039

65 Lehner, P., Karttunen, J., Wilkinson, G. and Cresswell, P. (1997) The human cytomegalovirus US6 glycoprotein inhibits transporter associated with antigen processing-dependent peptide

(19)

66 Ziegler, H., Muranyi, W., Burgert, H., Kremmer, E. and Koszinowski, U. (2000) The luminal part of the murine cytomegalovirus glycoprotein gp40 catalyzes the retention of MHC class I molecules. EMBO J. 19, 870-881

67 Ziegler, H., Thale, R., Lucin, P., Muranyi, W., Flohr, T., Hengel, H., Farrell, H., Rawlinson, W. and Koszinowski, U. (1997) A mouse cytomegalovirus glycoprotein retains MHC class I complexes in the ERGIC/cis-Golgi compartments. Immunity. 6, 57-66

68 Reusch, U., Muranyi, W., Lucin, P., Burgert, H., Hengel, H. and Koszinowski, U. (1999) A cytomegalovirus glycoprotein re-routes MHC class I complexes to lysosomes for degradation. EMBO J. 18, 1081-1091

69 Kavanagh, D., Koszinowski, U. and Hill, A. (2001) The murine cytomegalovirus immune evasion protein m4/gp34 forms biochemically distinct complexes with class I MHC at the cell surface and in a pre-Golgi compartment. J.Immunol. 167, 3894-3902

70 Kleijnen, M., Huppa, J., Lucin, P., Mukherjee, S., Farrell, H., Campbell, A., Koszinowski, U., Hill, A. and Ploegh, H. (1997) A mouse cytomegalovirus glycoprotein, gp34, forms a complex with folded class I MHC molecules in the ER which is not retained but is transported to the cell surface. EMBO J. 16, 685-694

71 Kavanagh, D., Gold, M., Wagner, M., Koszinowski, U. and Hill, A. (2001) The multiple immune-evasion genes of murine cytomegalovirus are not redundant: m4 and m152 inhibit antigen presentation in a complementary and cooperative fashion. J.Exp.Med. 194, 967-978

72 LoPiccolo, D., Gold, M., Kavanagh, D., Wagner, M., Koszinowski, U. and Hill, A. (2003) Effective inhibition of K(b)- and D(b)-restricted antigen presentation in primary macrophages by murine cytomegalovirus. J.Virol. 77, 301-308

73 Beck, S. and Barrell, B. (1988) Human cytomegalovirus encodes a glycoprotein homologous to MHC class-I antigens. Nature 331, 269-272

74 Farrell, H., Vally, H., Lynch, D., Fleming, P., Shellam, G., Scalzo, A. and Davis-Poynter, N. (1997) Inhibition of natural killer cells by a cytomegalovirus MHC class I homologue in vivo. Nature 386, 510-514

75 Beisser, P., Kloover, J., Grauls, G., Blok, M., Bruggeman, C. and Vink, C. (2000) The r144 major histocompatibility complex class I-like gene of rat cytomegalovirus is dispensable for both acute and long-term infection in the immunocompromised host. J.Virol. 74, 1045-1050

76 Cretney, E., Degli-Esposti, M., Densley, E., Farrell, H., Davis-Poynter, N. and Smyth, M. (1999) m144, a murine cytomegalovirus (MCMV)-encoded major histocompatibility complex class I homologue, confers tumor resistance to natural killer cell-mediated rejection. J.Exp.Med. 190, 435-444

77 Cosman, D., Fanger, N., Borges, L., Kubin, M., Chin, W., Peterson, L. and Hsu, M. (1997) A novel immunoglobulin superfamily receptor for cellular and viral MHC class I molecules. Immunity. 7, 273-282

78 Hassink, G., Duijvestijn-van Dam, J., Koppers-Lalic, D., Gaans-van den Brink, J., van Leeuwen, D., Vink, C., Bruggeman, C. and Wiertz, E. (2005) Rat cytomegalovirus induces a temporal downregulation of major histocompatibility complex class I cell surface expression. Viral Immunol. 18, 607-615

79 Jensen, T., Loo, M., Pind, S., Williams, D., Goldberg, A. and Riordan, J. (1995) Multiple proteolytic systems, including the proteasome, contribute to CFTR processing. Cell 83, 129-135

(20)

81 Biederer, T., Volkwein, C. and Sommer, T. (1997) Role of Cue1p in ubiquitination and degradation at the ER surface. Science 278, 1806-1809

82 Kolling, R. and Hollenberg, C. (1994) The ABC-transporter Ste6 accumulates in the plasma membrane in a ubiquitinated form in endocytosis mutants. EMBO J. 13, 3261-3271

83 Menendez-Benito, V., Verhoef, L., Masucci, M. and Dantuma, N. (2005) Endoplasmic reticulum stress compromises the ubiquitin-proteasome system. Hum.Mol.Genet. 14, 2787-2799 84 Schroder, M. and Kaufman, R. (2005) The mammalian unfolded protein response.

Annu.Rev.Biochem. 74, 739-789

85 Brewer, J. and Hendershot, L. (2005) Building an antibody factory: a job for the unfolded protein response. Nat.Immunol. 6, 23-29

86 Zhang, C., Cai, Y., Adachi, M., Oshiro, S., Aso, T., Kaufman, R. and Kitajima, S. (2001) Homocysteine induces programmed cell death in human vascular endothelial cells through activation of the unfolded protein response. J.Biol.Chem. 276, 35867-35874

87 Johnson, A. and van Waes, M. (1999) The translocon: a dynamic gateway at the ER membrane. Annu.Rev.Cell Dev.Biol. 15, 799-842

88 Higy, M., Junne, T. and Spiess, M. (2004) Topogenesis of membrane proteins at the endoplasmic reticulum. Biochemistry 43, 12716-12722

89 Martoglio, B., Hofmann, M., Brunner, J. and Dobberstein, B. (1995) The protein-conducting channel in the membrane of the endoplasmic reticulum is open laterally toward the lipid bilayer. Cell 81, 207-214

90 Hegde, R., Voigt, S., Rapoport, T. and Lingappa, V. (1998) TRAM regulates the exposure of nascent secretory proteins to the cytosol during translocation into the endoplasmic reticulum. Cell 92 , 621-631

91 Meacock, S., Lecomte, F., Crawshaw, S. and High, S. (2002) Different transmembrane domains associate with distinct endoplasmic reticulum components during membrane integration of a polytopic protein. Mol.Biol.Cell 13, 4114-4129

92 von Heijne, G. (1994) Signals for protein targeting into and across membranes. Subcell.Biochem. 22, 1-19

93 Denzer, A., Nabholz, C. and Spiess, M. (1995) Transmembrane orientation of signal-anchor proteins is affected by the folding state but not the size of the N-terminal domain. EMBO J. 14, 6311-6317

94 Wahlberg, J. and Spiess, M. (1997) Multiple determinants direct the orientation of signal-anchor proteins: the topogenic role of the hydrophobic signal domain. J.Cell Biol. 137, 555-562

95 Moss, K., Helm, A., Lu, Y., Bragin, A. and Skach, W. (1998) Coupled translocation events generate topological heterogeneity at the endoplasmic reticulum membrane. Mol.Biol.Cell 9, 2681-2697 96 Harley, C., Holt, J., Turner, R. and Tipper, D. (1998) Transmembrane protein insertion orientation

in yeast depends on the charge difference across transmembrane segments, their total hydrophobicity, and its distribution. J.Biol.Chem. 273, 24963-24971

97 Fons, R. and Bogert, B. (2003) Substrate-specific function of the translocon-associated protein complex during translocation across the ER membrane. 160, 529-539

(21)

99 Farmery, M., Allen, S., Allen, A. and Bulleid, N. (2000) The role of ERp57 in disulfide bond formation during the assembly of major histocompatibility complex class I in a synchronized semipermeabilized cell translation system. J.Biol.Chem. 275, 14933-14938

100 High, S., Lecomte, F., Russell, S., Abell, B. and Oliver, J. (2000) Glycoprotein folding in the endoplasmic reticulum: a tale of three chaperones? FEBS Lett. 476, 38-41

101 Haas, I. and Wabl, M. (1983) Immunoglobulin heavy chain binding protein. Nature 306, 387-389 102 Bole, D., Hendershot, L. and Kearney, J. (1986) Posttranslational association of immunoglobulin heavy chain binding protein with nascent heavy chains in nonsecreting and secreting hybridomas. J.Cell Biol. 102, 1558-1566

103 Liebermeister, W., Rapoport, T. and Heinrich, R. (2001) Ratcheting in post-translational protein translocation: a mathematical model. J.Mol.Biol. 305, 643-656

104 Rapoport, T., Matlack, K., Plath, K., Misselwitz, B. and Staeck, O. (1999) Posttranslational protein translocation across the membrane of the endoplasmic reticulum. Biol.Chem. 380, 1143-1150 105 Matlack, K., Misselwitz, B., Plath, K. and Rapoport, T. (1999) BiP acts as a molecular ratchet during

posttranslational transport of prepro-alpha factor across the ER membrane. Cell 97, 553-564 106 Lee, A., Wells, S., Kim, K. and Scheffler, I. (1986) Enhanced synthesis of the

glucose/calcium-regulated proteins in a hamster cell mutant deficient in transfer of oligosaccharide core to polypeptides. J.Cell Physiol 129, 277-282

107 Hamman, B., Hendershot, L. and Johnson, A. (1998) BiP maintains the permeability barrier of the ER membrane by sealing the lumenal end of the translocon pore before and early in translocation. Cell 92, 747-758

108 Haigh, N. and Johnson, A. (2002) A new role for BiP: closing the aqueous translocon pore during protein integration into the ER membrane. J.Cell Biol. 156, 261-270

109 van Laar, T., van der Eb, A. and Terleth, C. (2001) Mif1: a missing link between the unfolded protein response pathway and ER-associated protein degradation? Curr.Protein Pept.Sci. 2, 169-190 110 Sommer, T. and Wolf, D. (1997) Endoplasmic reticulum degradation: reverse protein flow of no

return. FASEB J. 11, 1227-1233

111 Sommer, T. and Jarosch, E. (2002) BiP binding keeps ATF6 at bay. Dev.Cell 3, 1-2

112 Shen, X., Ellis, R., Lee, K., Liu, C., Yang, K., Solomon, A., Yoshida, H., Morimoto, R., Kurnit, D., Mori, K. and Kaufman, R. (2001) Complementary signaling pathways regulate the unfolded protein response and are required for C. elegans development. Cell 107, 893-903

113 Liu, Y., Choudhury, P., Cabral, C. and Sifers, R. (1999) Oligosaccharide modification in the early secretory pathway directs the selection of a misfolded glycoprotein for degradation by the proteasome. J.Biol.Chem. 274, 5861-5867

114 Fagioli, C., Mezghrani, A. and Sitia, R. (2001) Reduction of interchain disulfide bonds precedes the dislocation of Ig-mu chains from the endoplasmic reticulum to the cytosol for proteasomal degradation. J.Biol.Chem. 276, 40962-40967

115 Oliver, J., van der Wal, F., Bulleid, N. and High, S. (1997) Interaction of the thiol-dependent reductase ERp57 with nascent glycoproteins. Science 275, 86-88

116 Oliver, J., Roderick, H., Llewellyn, D. and High, S. (1999) ERp57 functions as a subunit of specific complexes formed with the ER lectins calreticulin and calnexin. Mol.Biol.Cell 10, 2573-2582 117 Parodi, A. (1999) Reglucosylation of glycoproteins and quality control of glycoprotein folding in the

(22)

118 Wilson, C., Farmery, M. and Bulleid, N. (2000) Pivotal role of calnexin and mannose trimming in regulating the endoplasmic reticulum-associated degradation of major histocompatibility complex class I heavy chain. J.Biol.Chem. 275, 21224-21232

119 Fagioli, C. and Sitia, R. (2001) Glycoprotein quality control in the endoplasmic reticulum. Mannose trimming by endoplasmic reticulum mannosidase I times the proteasomal degradation of unassembled immunoglobulin subunits. J.Biol.Chem. 276, 12885-12892

120 Vashist, S., Kim, W., Belden, W., Spear, E., Barlowe, C. and Ng, D. (2001) Distinct retrieval and retention mechanisms are required for the quality control of endoplasmic reticulum protein folding. J.Cell Biol. 155, 355-368

121 Huyer, G., Longsworth, G., Mason, D., Mallampalli, M., McCaffery, J., Wright, R. and Michaelis, S. (2003) A striking quality control subcompartment in Saccharomyces cerevisiae: The endoplasmic reticulum-associated compartment (ERAC). Mol.Biol.Cell

122 Fayadat, L. and Kopito, R. (2003) Recognition of a single transmembrane degron by sequential quality control checkpoints. Mol.Biol.Cell 14, 1268-1278

123 Reggiori, F., Black, M. and Pelham, H. (2000) Polar transmembrane domains target proteins to the interior of the yeast vacuole. Mol.Biol.Cell 11, 3737-3749

124 Bebok, Z., Mazzochi, C., King, S., Hong, J. and Sorscher, E. (1998) The mechanism underlying cystic fibrosis transmembrane conductance regulator transport from the endoplasmic reticulum to the proteasome includes Sec61beta and a cytosolic, deglycosylated intermediary. J.Biol.Chem. 273, 29873-29878

125 Plemper, R., Bohmler, S., Bordallo, J., Sommer, T. and Wolf, D. (1997) Mutant analysis links the translocon and BiP to retrograde protein transport for ER degradation. Nature 388, 891-895 126 Plemper, R., Egner, R., Kuchler, K. and Wolf, D. (1998) Endoplasmic reticulum degradation of a

mutated ATP-binding cassette transporter Pdr5 proceeds in a concerted action of Sec61 and the proteasome. J.Biol.Chem. 273, 32848-32856

127 Hirsch, C., Blom, D. and Ploegh, H. (2003) A role for N-glycanase in the cytosolic turnover of glycoproteins. EMBO J. 22, 1036-1046

128 Suzuki, T., Kwofie, M. and Lennarz, W. (2003) Ngly1, a mouse gene encoding a deglycosylating enzyme implicated in proteasomal degradation: expression, genomic organization, and chromosomal mapping. Biochem.Biophys.Res.Commun. 304, 326-332

129 Albring, J., Koopmann, J., Hammerling, G. and Momburg, F. (2004) Retrotranslocation of MHC class I heavy chain from the endoplasmic reticulum to the cytosol is dependent on ATP supply to the ER lumen. Mol.Immunol. 40, 733-741

130 Rivett, A., Palmer, A. and Knecht, E. (1992) Electron microscopic localization of the multicatalytic proteinase complex in rat liver and in cultured cells. J.Histochem.Cytochem. 40, 1165-1172 131 Palmer, A., Rivett, A., Thomson, S., Hendil, K., Butcher, G., Fuertes, G. and Knecht, E. (1996)

Subpopulations of proteasomes in rat liver nuclei, microsomes and cytosol. Biochem.J. 316 ( Pt 2), 401-407

132 Brooks, P., Fuertes, G., Murray, R., Bose, S., Knecht, E., Rechsteiner, M., Hendil, K., Tanaka, K., Dyson, J. and Rivett, J. (2000) Subcellular localization of proteasomes and their regulatory complexes in mammalian cells. Biochem.J. 346 Pt 1, 155-161

133 Yang, M., Omura, S., Bonifacino, J. and Weissman, A. (1998) Novel aspects of degradation of T cell receptor subunits from the endoplasmic reticulum (ER) in T cells: importance of

(23)

134 Xiong, X., Chong, E. and Skach, W. (1999) Evidence that endoplasmic reticulum (ER)-associated degradation of cystic fibrosis transmembrane conductance regulator is linked to retrograde translocation from the ER membrane. J.Biol.Chem. 274, 2616-2624

135 Mayer, T., Braun, T. and Jentsch, S. (1998) Role of the proteasome in membrane extraction of a short-lived ER-transmembrane protein. EMBO J. 17, 3251-3257

136 Elkabetz, Y., Shapira, I., Rabinovich, E. and Bar-Nun, S. (2004) Distinct steps in dislocation of luminal endoplasmic associated degradation substrates: roles of endoplamic reticulum-bound p97/Cdc48p and proteasome. J.Biol.Chem. 279, 3980-3989

137 Confalonieri, F. and Duguet, M. (1995) A 200-amino acid ATPase module in search of a basic function. Bioessays 17, 639-650

138 Patel, S. and Latterich, M. (1998) The AAA team: related ATPases with diverse functions. Trends Cell Biol. 8, 65-71

139 Vale, R. (2000) AAA proteins. Lords of the ring. J.Cell Biol. 150, F13-F19

140 Baumeister, W., Walz, J., Zuhl, F. and Seemuller, E. (1998) The proteasome: paradigm of a self-compartmentalizing protease. Cell 92, 367-380

141 Fasshauer, D., Otto, H., Eliason, W., Jahn, R. and Brunger, A. (1997) Structural changes are associated with soluble N-ethylmaleimide-sensitive fusion protein attachment protein receptor complex formation. J.Biol.Chem. 272, 28036-28041

142 Neuwald, A., Aravind, L., Spouge, J. and Koonin, E. (1999) AAA+: A class of chaperone-like ATPases associated with the assembly, operation, and disassembly of protein complexes. Genome Res. 9 , 27-43

143 Ye, Y., Meyer, H. and Rapoport, T. (2001) The AAA ATPase Cdc48/p97 and its partners transport proteins from the ER into the cytosol. Nature 414, 652-656

144 Rabouille, C., Levine, T., Peters, J. and Warren, G. (1995) An NSF-like ATPase, p97, and NSF mediate cisternal regrowth from mitotic Golgi fragments. Cell 82, 905-914

145 Rouiller, I., Butel, V., Latterich, M., Milligan, R. and Wilson-Kubalek, E. (2000) A major conformational change in p97 AAA ATPase upon ATP binding. Mol.Cell 6, 1485-1490 146 Zhang, X., Shaw, A., Bates, P., Newman, R., Gowen, B., Orlova, E., Gorman, M., Kondo, H.,

Dokurno, P., Lally, J., Leonard, G., Meyer, H., van Heel, M. and Freemont, P. (2000) Structure of the AAA ATPase p97. Mol.Cell 6, 1473-1484

147 Ye, Y., Meyer, H. and Rapoport, T. (2003) Function of the p97-Ufd1-Npl4 complex in retrotranslocation from the ER to the cytosol: dual recognition of nonubiquitinated polypeptide segments and polyubiquitin chains. J.Cell Biol. 162, 71-84

148 Shamu, C., Story, C., Rapoport, T. and Ploegh, H. (1999) The pathway of US11-dependent degradation of MHC class I heavy chains involves a ubiquitin-conjugated intermediate. J.Cell Biol. 147, 45-58

149 Fang, S. and Weissmann, A. (2004) Ubiquitin-proteasome systemA field guide to ubiquitylation. Cell Mol.Life Sci. 61, 1546-1561

150 Varadan, R., Assfalg, M., Haririnia, A., Raasi, S., Pickart, C. and Fushman, D. (2004) Solution conformation of Lys63-linked di-ubiquitin chain provides clues to functional diversity of polyubiquitin signaling. J.Biol.Chem. 279, 7055-7063

(24)

152 Hicke, L. and Dunn, R. (2003) Regulation of membrane protein transport by ubiquitin and ubiquitin-binding proteins. Annu.Rev.Cell Dev.Biol. 19, 141-172

153 Marmor, M. and Yarden, Y. (2004) Role of protein ubiquitylation in regulating endocytosis of receptor tyrosine kinases. Oncogene 23, 2057-2070

154 Stahl, P. and Barbieri, M. (2002) Multivesicular bodies and multivesicular endosomes: the "ins and outs" of endosomal traffic. Sci.STKE. 2002, E32

155 Strous, G. and van Kerkhof, P. (2002) The ubiquitin-proteasome pathway and the regulation of growth hormone receptor availability. Mol.Cell Endocrinol. 197, 143-151

156 Sigismund, S., Polo, S. and Di Fiore, P. (2004) Signaling through monoubiquitination. Curr.Top.Microbiol.Immunol. 286, 149-185

157 Shcherbik, N. and Haines, D. (2004) Ub on the move. J.Cell Biochem. 93, 11-19 158 Bays, N., Gardner, R., Seelig, L., Joazeiro, C. and Hampton, R. (2001) Hrd1p/Der3p is a

membrane-anchored ubiquitin ligase required for ER-associated degradation. Nat.Cell Biol. 3, 24-29

159 Deak, P. and Wolf, D. (2001) Membrane topology and function of Der3/Hrd1p as a ubiquitin-protein ligase (E3) involved in endoplasmic reticulum degradation. J.Biol.Chem. 276, 10663-10669 160 Hampton, R., Gardner, R. and Rine, J. (1996) Role of 26S proteasome and HRD genes in the

degradation of 3-hydroxy-3-methylglutaryl-CoA reductase, an integral endoplasmic reticulum membrane protein. Mol.Biol.Cell 7, 2029-2044

161 Knop, M., Finger, A., Braun, T., Hellmuth, K. and Wolf, D. (1996) Der1, a novel protein specifically required for endoplasmic reticulum degradation in yeast. EMBO J. 15, 753-763 162 Bordallo, J., Plemper, R., Finger, A. and Wolf, D. (1998) Der3p/Hrd1p is required for endoplasmic

reticulum-associated degradation of misfolded lumenal and integral membrane proteins. Mol.Biol.Cell 9, 209-222

163 Bordallo, J. and Wolf, D. (1999) A RING-H2 finger motif is essential for the function of Der3/Hrd1 in endoplasmic reticulum associated protein degradation in the yeast Saccharomyces cerevisiae. FEBS Lett. 448, 244-248

164 Gardner, R., Swarbrick, G., Bays, N., Cronin, S., Wilhovsky, S., Seelig, L., Kim, C. and Hampton, R. (2000) Endoplasmic reticulum degradation requires lumen to cytosol signaling. Transmembrane control of Hrd1p by Hrd3p. J.Cell Biol. 151, 69-82

165 Hill, K. and Cooper, A. (2000) Degradation of unassembled Vph1p reveals novel aspects of the yeast ER quality control system. EMBO J. 19, 550-561

166 Wilhovsky, S., Gardner, R. and Hampton, R. (2000) HRD gene dependence of endoplasmic reticulum-associated degradation. Mol.Biol.Cell 11, 1697-1708

167 Gnann, A., Riordan, J. and Wolf, D. (2004) CFTR Degradation Depends on the Lectins Htm1p/EDEM and the Cdc48 Protein Complex in Yeast. Mol.Biol.Cell 15, 4125-4135

168 Fang, S., Ferrone, M., Yang, C., Jensen, J., Tiwari, S. and Weissman, A. (2001) The tumor autocrine motility factor receptor, gp78, is a ubiquitin protein ligase implicated in degradation from the endoplasmic reticulum. Proc.Natl.Acad.Sci.U.S.A 98, 14422-14427

(25)

Referenties

GERELATEERDE DOCUMENTEN

License: Licence agreement concerning inclusion of doctoral thesis in the Institutional Repository of the University of Leiden Downloaded from: https://hdl.handle.net/1887/4294.

License: Licence agreement concerning inclusion of doctoral thesis in the Institutional Repository of the University of Leiden Downloaded from: https://hdl.handle.net/1887/4294.

Ubiquitination is essential for human cytomegalovirus US11-mediated dislocation of MHC class I molecules from the endoplasmic reticulum to the cytosol. Human HRD1 is an E3

License: Licence agreement concerning inclusion of doctoral thesis in the Institutional Repository of the University of Leiden Downloaded.

The research presented in this thesis was performed in the Laboratory of Vaccine Research in the National Institute of Public Health and the Environment, Bilthoven, and in

In addition to gene products that affect the expression of MHC class I molecules on the surface of infected cells, HCMV, MCMV and also RCMV encode homologs of MHC class I heavy

The MHC class I breakdown intermediate was observed in E36 cells expressing HLA-A2 and US11, but not in ts20 cells expressing HLA-A2 and US11, incubated at 40°C in the presence

In the presence of the ubiquitin- conjugating enzyme UBC7, the RING-H2 finger has in vitro ubiquitination activity for Lys48-specific polyubiquitin linkage, suggesting that human HRD1