Download View Full Page PDF

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Organ-on-a-chip wikipedia , lookup

Cytosol wikipedia , lookup

Proteasome wikipedia , lookup

Interferon wikipedia , lookup

Signal transduction wikipedia , lookup

JADE1 wikipedia , lookup

List of types of proteins wikipedia , lookup

Transcript
Physiol Rev
82: 187–204, 2002; 10.1152/physrev.00025.2001.
The Transporter Associated With Antigen Processing:
Function and Implications in Human Diseases
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
Institut für Physiologische Chemie, Philipps-Universität Marburg, Marburg, Germany
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
I. Introduction
A. Overview of antigen processing
II. Discovery, Genomic Organization, and Homology of the Transporter Associated With Antigen
Processing Genes
A. Cell lines with defects in MHC class I presentation
B. Characterization
C. Homologies
D. Genomic organization
E. Polymorphism
III. Regulation of the Transporter Associated With Antigen Processing Expression
A. Induction of TAP by IFN-␥ and other factors
B. Promoter of the TAP1 gene
IV. Structural Organization of the Transporter Associated With Antigen Processing Complex
A. A model for ABC transporters
B. The TAP complex
V. Substrate Specificity of the Transporter Associated With Antigen Processing
A. Peptides optimal for transport
B. Polymorphism and substrate specificity of TAP
VI. Transport Mechanism of the Transporter Associated With Antigen Processing
A. Binding of peptides
B. Translocation of peptides
C. Protein machinery for peptide recognition, transport, and loading of MHC class I molecules
VII. Human Diseases Associated With the Transporter Associated With Antigen Processing
A. Viral infections and virus persistence
B. Genetic diseases
C. Autoimmune disease and transplantation
D. TAP deficiency and tumor development
VIII. Concluding Remarks
188
188
190
190
190
191
191
191
191
191
192
192
192
192
193
193
194
194
194
195
195
196
196
197
198
198
199
Lankat-Buttgereit, Brigittte, and Robert Tampé. The Transporter Associated With Antigen Processing:
Function and Implications in Human Diseases. Physiol Rev 82: 187–204, 2002; 10.1152/physrev.00025.2001.—The
adaptive immune systems have evolved to protect the organism against pathogens encountering the host.
Extracellular occurring viruses or bacteria are mainly bound by antibodies from the humoral branch of the
immune response, whereas infected or malignant cells are identified and eliminated by the cellular immune
system. To enable the recognition, proteins are cleaved into peptides in the cytosol and are presented on the
cell surface by class I molecules of the major histocompatibility complex (MHC). The transport of the antigenic
peptides into the lumen of the endoplasmic reticulum (ER) and loading onto the MHC class I molecules is an
essential process for the presentation to cytotoxic T lymphocytes. The delivery of these peptides is performed
by the transporter associated with antigen processing (TAP). TAP is a heterodimer of TAP1 and TAP2, each
subunit containing transmembrane domains and an ATP-binding motif. Sequence homology analysis revealed
that TAP belongs to the superfamily of ATP-binding cassette transporters. Loss of TAP function leads to a loss
of cell surface expression of MHC class I molecules. This may be a strategy for tumors and virus-infected cells
to escape immune surveillance. Structure and function of the TAP complex as well as the implications of loss
or downregulation of TAP is the topic of this review.
www.prv.org
0031-9333/02 $15.00 Copyright © 2002 the American Physiological Society
187
188
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
I. INTRODUCTION
A. Overview of Antigen Processing
FIG. 1. Overview of the antigen processing and presentation pathway via
major histocompatibility complex (MHC)
class I molecules. Endogenous proteins
are degraded by the proteasome. Cleavage products are transported into the lumen of the endoplasmic reticulum (ER)
by the transporter associated with antigen
processing (TAP) complex. Several molecules are involved in assembly and loading
of the MHC class I molecules, including
calnexin, calreticulin, tapasin, and ERp57
(for details, see Fig. 6). Stable MHC-peptide complexes leave the ER via the Golgi
compartment to the cell surface for recognition by cytotoxic T lymphocytes.
Physiol Rev • VOL
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
The adaptive immune system has evolved to protect
the organism against pathogens. The recognition and
elimination of mutated or infected cells are performed by
the cellular immune system, which can be subdivided into
the class I major histocompatibility complex (MHC) and
the class II MHC pathways. MHC class I molecules are
present on the surface of nearly all nucleated cells,
whereas class II molecules are restricted to the cells of
the immune system. Here, we focus on antigen processing
via MHC class I molecules. The peptides needed for antigen presentation are generated in the cytosol by cleavage
of endogenous proteins and are displayed on the cell
surface by MHC class I molecules (Fig. 1). The chronic
presentation of self-protein-derived peptides does not
lead to a stimulation of T cells. During the development of
the thymus, the T cells with the ability to recognize the
self-peptides are deleted. In case of malignant transformation or viral infection of the cells, an additional set of
peptides is presented to the cytotoxic T lymphocytes
(CTL). The intracellular pathways leading to the generation of peptides, their binding to MHC molecules, and
presentation on the cell surface are called antigen pro-
cessing and presentation. The recognition of MHC class I
molecules as “self component” loaded with peptides from
“non-self” proteins by CTL leads to apoptosis or lysis of
malignant or infected cells. This is an essential process of
the cellular immune response (for review, see Refs. 116,
165). The pathway of antigen presentation by the MHC
class I complex is constitutively active in all nucleated
cells and is upregulated by inflammatory cytokines. Interference with this pathway has evolved as an effective
strategy for pathogens to evade immune response, leading
to chronic or latent infections. However, human patients
deficient in MHC class I antigen presentation do not appear to have a higher incidence of viral infections (33).
In the absence of infection, normal endogenous proteins are degraded in the cytosol, but MHC class I-associated peptides can be also derived from viral antigens in
the cytosol (159) as well as from proteins that are artificially introduced into the cytosol (104, 181). Inhibition of
protein breakdown in lysosomes did not affect MHC class
I presentation (106) and implied that the presented peptides were generated by proteases in an extralysosomal
pathway. The major mechanism for degrading proteins in
the cytosol is ATP dependent and highly conserved from
yeast to mammals (52). The resulting peptides are generated by a proteolytic particle, the 20S/26S proteasome (15,
28, 42, 182). The contribution of the 20S proteasome for
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
Physiol Rev • VOL
encoded subunit ␤2-microglobulin and peptide (for review, see Refs. 91, 116). ␤2-Microglobulin makes extensive contacts with all three domains of the heavy chain
(21). For this reason, the conformation of the heavy chain
is dependent on the presence of ␤2-microglobulin. Allelic
polymorphism in the heavy chain primarily occurs at
residues involved in peptide binding and, therefore, alters
substrate specificity of the class I molecules (95, 96, 117).
In humans, there are three to six MHC class I alleles. This
small set of molecules has to bind the large pool of
peptides. Crystal structures of MHC class I molecules
provided further insight into this problem. How does a
single protein bind many different peptides? The antigenic peptide binds in a groove, which is formed by two
␣-helices on the rim and eight ␤-strands on the bottom of
the ␣1- and ␣2-domain of the heavy chain. The peptide is
fixed at the free COOH and NH2 terminus. Anchor residues at position two or three and at the COOH-terminal
residue pointing into the groove are also important for
binding. The groove accommodates peptides with eight or
nine residues in an extended conformation. The side
chains between most of these anchor residues point outside the groove. This binding motif explains the large pool
of peptides that is presented by one MHC class I allele.
Due to this arrangement, the variable region of the bound
peptide is monitored by the T-cell receptor (47, 48).
Efficient binding of peptides to the MHC class I molecules requires a macromolecular loading complex; the
assembly of this complex takes place in the ER and is
dependent on different chaperones that ensure correct
folding. After cotranslational translocation, the MHC
␣-chain associates with calnexin, a transmembrane calcium-dependent lectin with chaperone activity, followed
by formation of a noncovalent dimer with ␤2-microglobulin (for review, see Ref. 116). Calnexin is then replaced by
calreticulin, another calcium-dependent lectin, and
ERp57, a thiol reductase, is added to the complex, but the
exact time point of the interaction is unclear. Class I
heterodimers, associated with calreticulin and ERp57,
form the macromolecular loading complex with tapasin
and TAP. The glycoprotein tapasin is required for efficient
interaction of TAP and class I molecules (135). Meanwhile, there is evidence that calnexin and ERp57 bind to
the TAP-tapasin complex, generating an intermediate
which is capable of binding to the MHC class I-␤2-microglobulin dimers (37). The formation of this macromolecular complex is accompanied by the loss of calnexin and
binding of calreticulin. Furthermore, it could be demonstrated that the association of ERp57 to the complex is
tapasin dependent and in concordance with calreticulin,
not with calnexin (55). Despite the physical association of
TAP and the MHC class I molecules, it was shown by
inhibition of binding after antipeptide antibody application that possibly most TAP-transported peptides diffuse
through the lumen of the ER before being loaded onto
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
the generation of antigenic peptides is further strengthened by the observation that proteasomal subunits, LMP2
and LMP7 (low-molecular-mass polypeptides), are encoded within the MHC locus (22, 51, 97). The catalytic
core of the proteasome is a 20S (700 kDa) cylindrical
particle composed of 28 subunits arranged in four heptameric rings. The outer rings are made up of seven
␣-subunits; the inner rings are composed of seven ␤-subunits. Whereas the ␣-subunits are thought to be primarily
responsible for structural and regulatory functions, the
␤-subunits harbor the catalytic centers (40, 93, 142). Another form of the proteasome is the 26S (1,500 kDa)
particle. It contains the 20S complex and additional subunits associated with regulation of its activity (118, 126).
The proteasome produces peptides with a size distribution of 3–30 residues with an optimum of 6 –11 residues
(39, 77). This size corresponds in part to the size of
antigenic peptides bound to MHC class I molecules. During interferon (IFN)-␥ stimulation, the three active proteasomal ␤-subunits are replaced by ␤i-subunits (131) to
build up a so-called “immunoproteasome.” Immunoproteasomes show a different cleavage pattern compared
with the proteasomes, thereby generating more peptides
with hydrophobic and basic COOH termini (131); both of
them are favored for uptake by transporters associated
with antigen processing (TAP) into the endoplasmic reticulum (ER) and for optimal binding to MHC class I
molecules. About one-third of newly synthesized proteins
are rapidly degraded by proteasomes under physiological
conditions (139, 162). However, peptides are also generated, at least in part, by other proteases, such as the
IFN-␥-inducible leucine aminopeptidase (19) or a giant
cytosolic protease system, the tripeptidyl peptidase II (49,
50). It has been shown that blocking proteasomal activity
does not inhibit the presentation of antigenic peptides
derived from influenza viral proteins (173). Meanwhile,
there are results that breakdown of ovalbumin by the
proteasome primarily produced peptide variants with
NH2-terminal extensions compared with the immunodominant peptide, suggesting that MHC class I-bound peptides
are NH2-terminally trimmed by cytosolic aminopeptidases
(23). In addition, proteolytic trimming in the ER may also
produce epitopes after transport (132).
The TAP translocates peptides generated by the proteasome complex from the cytosol into the lumen of the
ER. In the ER, peptides are loaded onto the newly synthesized MHC class I molecules. Deletion or mutation of
TAP severely affects the translocation of peptides into the
ER. Because binding of peptides stabilizes the MHC complex and induces subsequent export to the cell surface for
presentation to T-cell receptors (Fig. 1), any defect, which
impairs the transport of peptides, results in reduced surface expression of MHC class I molecules. MHC class I
heterotrimers consist of a MHC encoded polymorphic
␣-chain (heavy chain) as well as the invariant non-MHC
189
190
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
cell lines were unable to present intracellular antigens on
the cell surface. It soon became clear that this defect was
due to deletions in the region of the MHC locus and most
likely involved a gene or genes that were responsible for
delivering peptides to the lumen of the ER and/or loading
of these peptides onto newly synthesized class I molecules. These genes were tentatively mapped to the MHC
locus by analysis of gene deletions in one of these mutant
cells (24). In the following, four groups independently
described candidate genes from the MHC region for a
factor that would transport peptides across the ER membrane (36, 103, 148, 160). In 1991, a World Health Organization (WHO) nomenclature committee for factors in the
HLA system renamed these genes to TAP (155). RING4,
PSF1, mtp1, and HAM1 are now TAP1, and RING11, PSF2,
mtp2, and HAM2 are renamed TAP2.
II. DISCOVERY, GENOMIC ORGANIZATION,
AND HOMOLOGY OF THE TRANSPORTER
ASSOCIATED WITH ANTIGEN
PROCESSING GENES
B. Characterization
A. Cell Lines With Defects in MHC Class I
Presentation
Peptides produced in the cytosol must pass into the
lumen of the ER before they can bind to MHC class I
molecules. The first indication that this translocation into
the ER requires a transporter was based on studies dealing with various cell lines with a strongly reduced level of
MHC class I molecules on the cell surface (91, 158). These
mutant cell lines exhibit a low cell-surface expression of
MHC class I and are deficient in antigen presentation.
Although the expression levels of MHC class I heavy chain
and ␤2-microglobulin are normal and peptides exogenously added or introduced into the ER by signal sequences were efficiently presented (8), these defective
The genes for human TAP1 and TAP2 are located in
the MHC II locus of chromosome 6 and comprise 8 –12 kb
each (Fig. 2) (161). Deletions of either or both of the TAP
genes result in greatly reduced surface expression of
MHC class I molecules (24, 65, 158). These defects can be
partly corrected by addition of exogenously supplied
class I-binding peptides (157). The functional impact of
TAP1 and TAP2 was proven by transfection of defective
cell lines with TAP2 and/or TAP1 genes, restoring antigenpresenting activity (12, 124, 149, 150). These observations
implicated that MHC class I molecules are unstable under
physiological conditions in the absence of appropriately
bound peptides and that the majority of binding peptides
are provided by a TAP-mediated transport. Moreover,
TAP1-deficient mice created by gene targeting exhibit a
phenotype consistent with that observed in TAP-defective
cell lines. These mice show a severely reduced class I cell
surface expression and fail to present cytosolic antigens
to cytotoxic T cells (171).
FIG. 2. The location of the TAP genes in the
human MHC class II complex (HLA) in a schematic representation. The scale is intended as a
rough guide only. Boxes indicate the approximate position of the genes. The figure underneath the schematic location of the genes illustrates the participation of the gene products in
peptide generation and transport.
Physiol Rev • VOL
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
MHC class I molecules (63). The binding of peptides to
class I heterodimers releases the loaded MHC class I
molecules from the assembly complex for transport to the
cell surface via the Golgi apparatus and the trans-Golgi
network. The export from the ER by ER-budding vesicles
is the rate-limiting step, and the multitude of ER-to-Golgi
transport is independent of peptide loading or dissociation from TAP (151).
During the last years, the comprehension of structure
and functions of TAP increased significantly, stressing the
importance of this transporter for the adaptive immune
response. A growing number of studies deal with the
implications of disturbed TAP function in various human
diseases from viral infections to tumor development. In
this review, the current knowledge about TAP and its
implications for diseases are summarized.
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
191
E. Polymorphism
The sequences of the TAP1 and TAP2 coding regions
were determined, and alignment of TAP1 and TAP2
showed the highest homology in a stretch of 200 amino
acids near the COOH terminus. This region contains three
characteristic motifs. The Walker A (P loop) and Walker B
motifs form a highly conserved ATP-binding cassette
(ABC). ATP and other nucleotides bind to these sites and
are hydrolyzed in a Mg2⫹-dependent manner (107, 134,
177). A so-called C loop, which consists of six to eight
conserved amino acids, is located in between the Walker
A and B motif. There is genetic evidence that generally a
so-called “EAA” sequence found in the last cytosolic loop
of the membrane-spanning domain of bacterial ABC
transporters interacts with the C loop, thereby coupling
ATP hydrolysis to the transmembrane spanning domain
(27, 32). Due to the sequence homology, TAP1 and TAP2
belong to the superfamily of ABC transporters and, in
more detail, to subfamily B. A conserved nucleotide-binding domain (NBD) and a transmembrane domain (TMD)
of about six membrane-spanning segments characterize
this protein family. A functional unit comprises two NBDs
and two TMDs. ABC transporters are found in all domains
of life throughout the animal and plant kingdoms, bacteria, and archea (61). They form the largest and most
diverse family of membrane-spanning transport proteins.
TAP1 and TAP2 are polymorphic in all species examined, and there is clear evidence that this polymorphism can
affect the substrate specificity of the transporter (56, 101,
121, 122, 180). On the other hand, a functional polymorphism was only demonstrated for rat TAP (112) and for the
human TAP2iso splice variant (180). The rat TAP from the
RT1u strain as well as human TAPiso most efficiently translocates peptides with hydrophobic COOH-terminal residues,
while TAP from the rat RT1a strain is permissive for peptides
containing hydrophobic and basic COOH termini (56, 101).
Human TAP apparently resembles more closely rat TAP
from the RT1a strain with a broader specificity, whereas
mouse TAP has a higher similarity to rat TAP from the RT1u
strain. The level of structural polymorphism in human and
mouse TAP (an average of 2–5 amino acid substitutions
between alleles) (26, 123) is reduced compared with that
seen in rat (25 positions of allelic variations in the rat TAP2
gene) (121, 125).
Nonetheless, the TAP genes are polymorphic both in
mice and humans. Several alleles of human TAP1 and TAP2
have been identified in different populations (133). This
suggests that functional differences between alleles exist,
even though initial attempts to demonstrate such functional
polymorphism have been unsuccessful (31, 114, 141, 168).
On the other hand, data concerning polymorphism of the
TAP genes in different human populations point to little, if
any, functional significance (69, 75, 99, 133).
D. Genomic Organization
The genomic structures of the human and mouse
TAP genes have been established (17, 160). As mentioned
in section IIB, the human TAP1 and TAP2 genes are located in the MHC II locus of chromosome 6 and are 8 –12
kb in size. They consist of 11 exons (54); 8 of them are of
the same length, although the intron sizes vary significantly. The remaining three exons 1, 9, and 11 differ in
length by 100, 3, and 78 nucleotides, respectively. All 11
intron/exon boundaries are identical in their classes and
follow the GT/AG rule. Sequence comparisons from human to salmon TAP1 showed the expected phylogenetic
differences (for example, 98.8% homology of human with
gorilla TAP1, 69.2% with hamster, and 40% with salmon).
The analysis indicates that the degree of relatedness between human TAPs is similar to that between each gene
and its homologs in rodents (66). The homology between
TAP1 and TAP2 in all species examined up to now is
⬃35%, although they share a similar predicted membrane
topology. Therefore, it is speculated that the genes have
evolved from a common ancestral gene by duplication
before the development of the adaptive immune system in
vertebrates (16).
Physiol Rev • VOL
III. REGULATION OF THE TRANSPORTER
ASSOCIATED WITH ANTIGEN
PROCESSING EXPRESSION
A. Induction of TAP by IFN-␥ and Other Factors
MHC class I molecules are expressed at low levels in
most cells and are strongly induced by cytokines such as
IFN-␥ (for a review, see Ref. 42). Increased expression of
MHC class I molecules correlates with increased CTL function. Therefore, it was of special interest to study the regulation of TAP gene expression. Both TAP1 and TAP2 mRNAs
are upregulated ⬃10- to 20-fold within 12 h by IFN-␥,
whereas HLA class I heavy chain mRNA is induced more
slowly, increasing only 10-fold in 24 h (13, 94). Although
TAP1 and TAP2 proteins are also rapidly induced, HLA class
I heavy chains and surface expression increase more slowly,
accompanied by an increased peptide transport capacity.
Comparable to these data, gene transfer of a cDNA encoding
IFN-␥ into a human renal carcinoma cell line results in a high
constitutive expression of TAP, but allospecific CTL reaction was not better stimulated with IFN-␥ transductants than
with unmodified tumor cells (138). In some cases, TNF-␣
can lead to an increase in TAP1 levels by in vitro treatment
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
C. Homologies
192
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
belong to the superfamily of ABC transporters, comprising a large number of polytopic membrane proteins. Members of this family transport a diverse set of molecules
ranging from ions to large proteins across cell membranes
in an ATP-dependent manner (for review, see Refs. 38,
61). ABC transporters play an important role in different
(patho)physiological processes, such as cystic fibrosis
(cystic fibrosis transmembrane conductance regulator),
multidrug resistance (e.g., MDR1 or MRP1), high-density
lipoprotein deficiency (Tangier disease, ABCA1), retinitis
pigmentosa 19 and age-related macular degeneration
(ABCR), hyperinsulinemic hypoglycemia of infancy
(SUR1), and adrenoleukodystrophy (ALD) (for review,
see Ref. 79). As mentioned in section II, all ABC transporters share two conserved cytoplasmic ATP-binding domains and two hydrophobic domains, consisting of 5–10
transmembrane stretches. These stretches are composed
of ␣-helices, lining up the putative translocation pore. One
to four genes can encode all four domains.
B. Promoter of the TAP1 Gene
B. The TAP Complex
The human TAP genes contain no TATA box motifs
in the 5⬘-flanking sequences but putative GC-rich elements (Sp1 binding sites, 128 nucleotides upstream of the
translation start codon for TAP1 and 79 nucleotides upstream in the case of TAP2) (17). Site-directed mutagenesis of the Sp1 binding site leads to a threefold reduction
of basal promoter activity of the TAP1 gene (179). The
promoter is induced by several IFN-␥-responsive elements and a p53-responsive element, which act independently from each other (185). However, the activity of the
p53-responsive element was proven in vivo only in context with a heterologous minimal promoter, and no data
are available concerning the significance of this element
in the homologous promoter region. It should be noted
that the TAP1 gene is coordinately regulated by a bidirectional promoter of 593 bp, also directing the divergently
transcribed gene of LMP2, the ␤-type proteasomal subunit
␤1i. Both genes are induced by tumor necrosis factor
(TNF)-␣ via a NF-␬B-like enhancer (68). The cytokinestimulated expression of TAP1 and LMP2 concordantly
with class I genes suggests a mechanism linking transporter levels with class I production.
Immunoelectron and immunofluorescence microscopy analyses have localized TAP1 and TAP2 in the ER
and the cis-Golgi (78, 98), retarded by a so far uncharacterized ER retention signal. This subcellular distribution
is consistent with the location where MHC molecules bind
peptides. The homology with other ABC transporters
points to a functional TAP protein either as a homodimer
of TAP1 or TAP2 or a heterodimer. Indeed, immunoprecipitations with antisera directed against TAP1 coprecipitated TAP1 and TAP2 (76). Comparisons of the phenotypes of different mutant cell lines suggest, and
heterologous coexpression of TAP1 and TAP2 in insect
cells and yeast demonstrate, that neither TAP1 nor TAP2
forms a functional homodimer (98, 166). Furthermore, no
additional factors of the immune system are required for
TAP function. TAP1 and TAP2 are found to be nonglycosylated, although TAP1 has three putative glycosylation
sites; however, two are facing the cytosol, and one is
placed in an ER loop too short for effective glycosylation
(98).
Hydrophobicity analysis and sequence alignments of
TAP proteins with related ABC transporters point to a
two ⫻ six transmembrane helix model of TAP, extended
by additional four and three transmembrane helices predicted for the highly diverse NH2-terminal domain (Fig. 3)
(152). Up to now, the function of these NH2-terminal
hydrophobic extensions is unclear. As also found for
other ABC transporters located in the ER, both TAP molecules lack a known NH2-terminal signal sequence for the
import into the ER, indicating that an internal signal
sequence may exist. Experiments with truncated proteins
suggest that TAP1 and TAP2 contain multiple ER reten-
IV. STRUCTURAL ORGANIZATION OF THE
TRANSPORTER ASSOCIATED WITH
ANTIGEN PROCESSING COMPLEX
A. A Model for ABC Transporters
As mentioned before, human TAP1 (748 amino acids)
and human TAP2 (686 amino acids, 653 for TAP2iso)
Physiol Rev • VOL
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
of tumor samples (109). Another cytokine, interleukin-10,
expressed in a murine tumor cell line, has an opposite effect;
a reduction of TAP level was observed (136).
Tumor cells may evade tumor surveillance by acquiring mutations that inhibit antigen processing via MHC
class I. Because more than 50% of human tumors have a
dysfunctional p53, this mutation may diminish or abrogate the host tumor surveillance. Therefore, the effect of
p53 on TAP1 expression was investigated (185). The authors observed that p53 as well as DNA-damaging agents
could increase TAP1 levels and subsequent MHC class I
cell surface expression, indicating that a dysfunctional
p53 in human tumor cells would not induce TAP1 after
genotoxic stress. These results suggest that tumor surveillance may be a mechanism by which p53 acts as a tumor
suppressor. Furthermore, several breast cancer cell lines
showed lower expression levels of TAP mRNA compared
with normal breast epithelial cells and a cell cycle-dependent regulation (6).
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
193
FIG. 3. Membrane topology of the human TAP complex. The transmembrane helices are predicted from hydrophobicity plots and sequence alignments with other
ABC transporters. The unique NH2-terminal domains (N)
of TAP1 and TAP2 are very hydrophobic, comprising four
and three predicted transmembrane helices, respectively
(orange cylinders). Numbers in the cylinders indicate the
amino acid residues beginning and ending the transmembrane domains. The nucleotide binding domain consists of
the Walker A and B motifs (A, B) and the so-called C loop
(C), possibly interacting with membrane-spanning domains, in particular with the EAA-like motif (E). The orange lines illustrate binding regions for peptides as identified by photo-cross-linking experiments (113).
Physiol Rev • VOL
transporter P-glycoprotein (92). Because of the discrepancies in the models, further studies are needed to clarify
the structural organization and membrane topology of
TAP.
V. SUBSTRATE SPECIFICITY OF THE
TRANSPORTER ASSOCIATED WITH
ANTIGEN PROCESSING
A. Peptides Optimal for Transport
The characterization of antigenic epitopes is very
important for the knowledge about the immunologic response and the development of vaccination strategies.
Therefore, the substrate specificity of TAP is well studied.
The first results were obtained by trapping transported
peptides in the ER via glycosylation (110). The comparison of glycosylated peptides differing in length or amino
acid composition provided information about length and
sequence preferences of the transported peptides (for
review, see Refs. 10, 165). Peptides with a length of 8 –16
amino acids bind to TAP with equal affinity (170); translocation into the ER was most efficient for peptides of
8 –12 amino acids (83). Transport of peptides with even 40
amino acids was observed, but with lower efficiency.
Therefore, TAP transports preferentially peptides with a
similar or slightly larger size suitable for MHC class I
binding. Another approach to study TAP selectivity was
realized determining the equilibrium binding affinity constant (164, 170). Equal binding affinities were found for
peptides ranging from 8 to 16 amino acids (170). The only
discrepancies exist in the absolute values between binding affinity and transport efficiency (10, 164, 169). The
glycosylation method includes several side reactions such
as peptide transport, glycosylation, degradation, and export and, therefore, does not necessarily reflect transport
rates. Thus the binding assay seems to be more accurate
for the analysis of substrate specificity.
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
tion signals in the transmembrane regions, including the
first transmembrane segment (176). The hydrophobic
transmembrane regions are linked to the nucleotide binding domains, which contain the conserved Walker A and B
motifs for ATP-binding and hydrolysis and, based on this
model, are located in the cytosol. According to this model,
the complex is highly asymmetric with respect to the
cytosolic and ER-luminal regions; large cytosolic loops
and both NBDs are located in the cytosol, but only a small
portion (⬍10%) reaches toward the ER lumen (152). It
was further proposed that the pore-forming domains of
the TAP1 and TAP2 heterodimer are aligned in a headhead/tail-tail orientation (175), which however contradicts the TMD organization of P-glycoprotein (92).
Peptides are photo-cross-linked to TAP1 and TAP2,
suggesting that both subunits are involved in peptide
binding (10, 11). More recently, the peptide-binding region
was mapped after peptide photo-cross-linking, digestion
of TAP by trypsin and/or bromocyan, and subsequent
immunoprecipitation with antibodies directed against different epitopes of TAP (113). This analysis revealed a
similar binding region for TAP1 and TAP2, comprising the
cytosolic loops between TM4 and TM5 and a COOHterminal stretch of ⬃15 amino acids after TM6 (Fig. 3).
Deletion of some of these potential peptide-binding sites
between residues 366 and 405 of TAP1 resulted in defective peptide transport (130). Due to the topological model
of TAP (1, 152), all these regions are expected to be
exposed to the cytosol. Recently, a slightly different
model of the TAP complex was presented (176). With a
set of COOH-terminal deletions for the TAP1 and TAP2
subunits, the topology of TAP was investigated. Eight and
seven transmembrane segments were identified for TAP1
and TAP2, respectively. Due to this model, no evidence
for membrane integration was obtained for the two hydrophobic regions adjacent to the nucleotide binding domains for both subunits. However, this model is in contrast to the established membrane topology of the ABC
194
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
The contribution of each peptide residue to the affinity for TAP was determined systemically by screening
combinatorial peptide libraries (163) (Fig. 4). With this
method, the average affinity of a randomized peptide
mixture with one defined residue is compared with a
totally randomized peptide library. Therefore, the influence of each peptide residue was analyzed independent of
a given sequence context on the affinity of TAP. The effect
of each amino acid residue on the stabilization of peptide
binding to TAP was determined. The strongest differences
were observed for the COOH-terminal amino acid and the
first three NH2-terminal residues. The amino acids in be-
tween do not significantly contribute to substrate specificity of TAP but are necessary for recognition of a MHC
class I-bound peptide by the T-cell receptor. This binding
motif combines two advantages: 1) maximal diversity of
residues 5– 8 in the antigenic peptide where T-cell receptor recognition and no selection by TAP occurs, and 2)
maximal binding affinity at the termini for TAP. The
matched preferences of human TAP and MHC class I
suggest a coevolution of the genes (165) as it was also
proposed for rat TAP and MHC class I molecules (72).
B. Polymorphism and Substrate Specificity of TAP
4. Peptide specificity for the TAP transporter as determined by
combinatorial peptide libraries (164). Top panel: substrate specificity for
TAP. The first three NH2-terminal amino acids and the last COOHterminal amino acid contribute significantly to the stabilization of peptide binding to TAP. Middle panel: favored amino acids at the individual
positions with negative ⌬⌬G values (favored residues) are shown in
blue, and positive ⌬⌬G values (disfavored residues) are in red. For
example, for the first position the amino acids K, N, and R are favored,
and D, E, and F are disfavored. Bottom panel: a model of the substratebinding pocket of TAP.
FIG.
Physiol Rev • VOL
VI. TRANSPORT MECHANISM OF THE
TRANSPORTER ASSOCIATED WITH
ANTIGEN PROCESSING
A. Binding of Peptides
Peptide transport by TAP is a multistep process
(170). The peptide-binding pocket is formed by TAP1 and
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
The strongest differences in peptide binding affinity
to TAP were detected at the COOH terminus of peptides
(101, 103, 141, 163). As mentioned in section II, a limited
degree of sequence polymorphism has been found for
human and mouse TAP genes by restriction length polymorphism and sequence analysis, but no measurable alteration in function of these gene products was observed
(133, 141). However, it has been shown that in rats, TAP
polymorphism has an influence on peptide selectivity.
TAP from the rat strain RT1a as well as human TAP
translocates peptides with a broad specificity (hydrophobic or basic amino acids at the COOH terminus), whereas
TAP from the rat strain RT1u and mouse TAP prefers
peptides with hydrophobic COOH termini. This pattern
correlates with the predominant peptide binding profiles
of human and mouse MHC class I molecules.
Although human TAP genes are less allelically diverse than rat genes, another possibility to generate different substrate specificity in the human transporter was
recently reported (180). An alternative splicing variant of
the human TAP2 gene was described termed TAP2iso. It
lacks exon 11, a part of the coding region and the original
3⬘-untranslated region, but contains the newly identified
exon 12. TAP2iso mRNA was normally coexpressed with
TAP2 mRNA in all human lymphocyte cell lines examined.
TAP1-TAP2iso transporters exhibited distinct and opposing influences on peptide selectivity, more resembling the
TAP complex from rat strain RTu and mouse TAP. The
common expression of an alternative splice product of
the TAP2 gene may contribute to a broadened immune
diversity.
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
195
B. Translocation of Peptides
The translocation strictly requires hydrolysis of ATP,
because nonhydrolyzable ATP analogs do not promote
peptide transport. In addition to ATP, UTP, CTP, and GTP
(9) can energize peptide translocation. Direct binding of
nucleotides was demonstrated by 8-azido-ATP photocross-linking experiments (107, 134, 176). ATP and ADP
have similar affinities for TAP; therefore, ADP and other
nucleoside diphosphates competing for ATP binding inhibit peptide translocation. The ATPase activity of TAP is
substrate specific and tightly coupled to peptide binding,
indicating that peptide binding is a prerequisite for ATP
hydrolysis, thereby possibly preventing waste of ATP
without transport of peptides (53). ATP hydrolysis follows Michaelis-Menten kinetics with a turnover number
kcat of ⬃5 ATP/s. This turnover rate is sufficient to account for the essential role of TAP in the overall process
of antigen processing. Interestingly, sterically restricted
peptides that bind to but that are not transported by TAP
do not stimulate the ATPase activity of TAP (53). Both
nucleotide-binding domains of TAP interact with ATP,
even if expressed separately (107, 176). A structural rearrangement of the NBDs by peptide binding seems to
function as a molecular switch to activate the ATPase of
TAP, since the NBDs alone are unable to hydrolyze ATP
(107, 177). It is known that both NBDs are essential for
TAP function, since mutation of one NBD leads to a loss
of transport action (25). It may be that hydrolysis at one
NBD starts the peptide transport, whereas hydrolysis at
the second NBD is necessary for completion of the translocation cycle and perhaps for the reconversion of the
initial substrate-binding site. A model of the peptide translocation by the TAP complex is presented in Figure 5.
Lysine mutation of the Walker A motif of TAP1 and TAP2
points to distinct functional properties of both NBDs (84,
Physiol Rev • VOL
FIG. 5. Current model of peptide binding and transport by the TAP
complex. ATP and/or ADP as well as peptide bind independently to TAP
in the cytosol. Peptides are shown as blue triangles. It is unknown if
both nucleotide binding domains are loaded with ATP at the beginning
of the cycle. Peptide binding induces a conformational change of the
TAP complex. This structural reorganization may trigger the ATP hydrolysis and translocation of the peptide.
137). Even if the data concerning nucleotide and peptide
binding after mutation of human TAP1 or TAP2 are somehow contradictory, it seems that ATP binding to both subunits is tightly regulated and transport cycle dependent.
Results with glycine mutations in the Walker A motif of rat
TAP1 or TAP2 indicate that binding of ATP to TAP1 is the
first step in energizing the transport cycle and induces ATP
binding to TAP2 (4). Future investigations are needed to
clarify these conflicting observations.
TAP only transports peptides longer than 7 residues;
the efficiency drops off with peptides longer than 12
amino acids, but there is no clear limit (9, 22, 101, 145). A
complicating factor for analyzing TAP-dependent transport is the presence of an efficient ATP-dependent peptide
export system out of the ER and into the cytoplasm (82,
132, 140). Peptides with low affinities may not be rapidly
bound to the MHC class I molecules and are therefore
available for export; this may be another mechanism to
select for high-affinity peptides binding to the MHC class
I molecules.
C. Protein Machinery for Peptide Recognition,
Transport, and Loading of MHC Class I
Molecules
The view that TAP plays an important role in class I
assembly derives from the observation that it not only
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
TAP2. Peptides associate with TAP in an ATP-independent manner at low temperatures (164). So far, no indication for a second binding site was found by direct
peptide binding or competition assays as well as by photocross-linking experiments. With the use of fluorescencelabeled peptides, the association pathway to TAP could
be kinetically dissected in real time (112). In a fast bimolecular association step, peptide binds to TAP, followed
by a slow isomerization of the TAP complex. It is suggested that this structural reorganization of the molecule
triggers ATP hydrolysis and peptide translocation across
the membrane. The binding step determines the selectivity of TAP (112). The conformational change of TAP also
influences its lateral mobility, shown by tagging the TAP1
subunit with green fluorescent protein (127). The mobility
of the TAP complex increases when it is inactive and
decreases during peptide translocation.
196
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
Peptide binding to MHC class I molecules is a prerequisite for dissociation of TAP-MHC complexes. This
peptide-mediated detachment depends on conformational
signals from the TAP complex induced by ATP binding
(80). Recent data suggest that release of peptide-loaded
MHC class I molecules is predominantly dependent on the
conformation of TAP1 (4). Therefore, the dynamic activity of the TAP-MHC class I complex is synchronized with
the peptide binding and translocation cycle of TAP.
VII. HUMAN DISEASES ASSOCIATED WITH
THE TRANSPORTER ASSOCIATED WITH
ANTIGEN PROCESSING
A. Viral Infections and Virus Persistence
CTL recognize and eliminate virus-infected cells. This
requires the presentation of viral peptide fragments by
MHC class I molecules at the cell surface. Persistent
viruses, such as the large DNA-containing herpes viruses,
have been forced to evolve different strategies to evade
the immune defense, which means especially by interfering with antigen processing and presentation (for review,
see Refs. 59, 120).
The immediate early protein ICP47 encoded by herpes simplex virus type 1 (HSV-1) was identified to be
responsible for the downregulation of the MHC class I
surface expression in human fibroblasts (183). Therefore,
HSV-infected cells are masked for immune recognition by
CTL. ICP47 inhibits the TAP-mediated peptide transport
into the ER by binding to the TAP complex (41, 62). It was
demonstrated that ICP47 binds with high affinity at least
in part to the peptide-binding site of TAP, thereby blocking the first and essential step in the translocation pathway (3, 156). Recently, a similar downregulation of peptide transport activity was observed in bovine epithelial
cells infected with bovine herpesvirus-1 (64). The interaction of human ICP47 and the TAP complex is species
specific, since ICP47 has a 100-fold higher affinity for
human than for murine TAP (3, 156). Meanwhile, the
critical amino acids of ICP47, which are necessary to
inhibit TAP function, have been identified (111, 159).
These amino acids comprise the NH2-terminal region
FIG. 6. The macromolecular transport and chaperone
complex. The de novo-synthesized MHC-encoded heavy
chains (left) assemble with calnexin and ␤2-microglobulin.
After binding to ERp57 and calreticulin, the MHC class I
molecules can associate via tapasin to the TAP complex.
Peptides are translocated into the lumen of the ER by TAP
and loaded onto the MHC class I complex, which leaves
the ER to the cell surface.
Physiol Rev • VOL
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
delivers peptides to the ER, but also is involved in the
formation of a large loading complex critical for class I
maturation (Fig. 6). A number of other proteins are required for the assembly and loading of MHC class I molecules, and several chaperone-like proteins have been
identified (for review, see Refs. 29, 87). At least four
proteins, calnexin, calreticulin, ERp57, and tapasin, are
involved in the assembly of heavy chain and ␤2-microglobulin. The dominant chaperones associated with newly
synthesized class I heavy chains are calnexin and calreticulin (67, 135). Calreticulin is a soluble homolog to calnexin and shares its specificity for monoglycosylated Nlinked glycans. The mechanism how calnexin/calreticulin
facilitates folding is not understood, but it is suggested to
be involved in a binding and release cycle. If the association of these chaperones is disturbed by glucosidase
inhibitors, the folding and subsequent surface expression
of MHC class I molecules are dramatically reduced (153,
172). An additional housekeeping molecule, called ERp57,
is part of the large multisubunit TAP-tapasin-MHC complex (67, 90, 105). ERp57 is a member of the thioredoxin
family of enzymes and also binds to the complex in a very
early step of the maturation, probably supporting the
correct formation of disulfide bridges.
The MHC class I molecules interact with the TAP
complex via an ER-resident type I glycoprotein, named
tapasin for TAP-associated glycoprotein (88, 115). Tapasin is a transmembrane protein that, like MHC molecules,
is a member of the immunoglobulin superfamily. It has an
apparent molecular mass of 48 kDa. Tapasin mediates
complex formation and the cross-talk of structural information of MHC and TAP and is therefore important for
class I assembly and editing (135). In its absence, the
assembly in the ER is impaired, and MHC class I surface
expression is reduced. Tapasin has two independent functions: 1) it increases the level of TAP, thereby increasing
the efficiency of peptide transport, and 2) it associates
with MHC class I molecules, thereby facilitating directly
loading and assembly of class I molecules (14, 86). Approximately four MHC class I molecules seem to be linked
via four tapasin to one TAP complex (115). Truncated
tapasin missing the transmembrane region and the cytosolic tail appears to be not associated with TAP, but still
rescues MHC class I expression (86).
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
Physiol Rev • VOL
disease recurrence. These findings suggest that HPV may
evade immune recognition by downregulating MHC class
I cell surface expression via decreased TAP1 levels. The
expression of TAP1 could be used for prognostic evaluation of disease severity. Moreover, a similar mechanism
with decreased TAP1 and MHC class I protein was described in other malignancies induced by HPV, for example, HPV-16- and HPV-18-infected carcinomas of the cervix (30). Similar to HPV, the Epstein-Bar virus (EBV) has
evolved a strategy to avoid immune surveillance by downregulation of TAP1 (184). EBV expresses a protein, vIL-10,
that is similar in sequence to human interleukin-10. vIL-10
downregulates the expression of the TAP1 gene, but not
of the TAP2 gene, thereby affecting the transport of peptides into the ER. This mechanism might contribute to the
persistence of EBV-infected cells in the human host.
B. Genetic Diseases
There is little knowledge about human genetic TAP
defects. Bare lymphocyte syndrome (BLS) is known to be
characterized by a severe downregulation or deficiency of
MHC class I and/or MHC class II molecules. In type 1 BLS,
the defects are assigned to the MHC class I, whereas in
type 2, the MHC class II molecules are affected. The
characterization of 22 patients with type 1 BLS revealed
the existence of 3 clinically and immunologically distinct
subsets (for review, see Ref. 45). Two groups include
patients with a defective transcription of MHC class I
molecules and ␤2-microglobulin. The third group is the
best-characterized subset of type 1 BLS and comprises 15
patients from different families. The reduced MHC class I
cell surface expression manifests with necrotizing granulomatous skin lesions and recurrent bacterial infections.
Tissue typing of 10 patients indicated that a defect TAP
complex is responsible for the downregulation of MHC
class I molecules. Studies of two families revealed that
the disease is caused by deficiency of TAP2, which is
mainly due to a premature stop codon resulting in a
nonfunctional TAP complex (33, 154). The homozygous
TAP2 ⫺/⫺ siblings of one family have reduced expression
of MHC class I molecules on the cell surface and reduced
amounts of CTL. On the other hand, these people do not
show an increased susceptibility to viral infections. A
fraction of antigenic peptides seems to be transported
into the ER lumen in a TAP-independent manner, although TAP appears to be the dominant pathway under
normal conditions (34).
Five adults with a syndrome resembling Wegener’s
granulomatosis were also found to exhibit low surface
expression of MHC class I molecules and defective expression of the TAP complex (100). In two patients, a
mutation in the TAP2 gene was identified, responsible for
this defective expression of the TAP complex.
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
from residue 3 to 34 (111). This active domain of ICP47
appears to be mainly unstructured in aqueous solution
(18). In the presence of lipidlike environment, an ␣-helical
structure is induced, which is composed of two helical
regions extending from residues 4 –15 and 22–32 connected by a flexible loop (18, 119). On the basis of these
results, therapeutic drugs could be designed that are potent immune suppressors or that are applicable in novel
vaccination strategies against HSV, restoring the ability of
the immune system to recognize the infected cells.
Human cytomegalovirus (HCMV) encodes at least
four proteins that decrease the cell surface expression of
MHC class I molecules (for review, see Ref. 59). An ERresident glycoprotein, called gpUS6, inhibits the TAPmediated peptide transport. The mechanism is probably
due to binding of gpUS6 to the ER-luminal part of the TAP
complex (2, 58, 85). Peptide binding is not affected by
gpUS6; also, the association of tapasin, calreticulin, and
MHC class I molecules to TAP appears normal. Recent
findings indicate that gpUS6 binds to TAP and stabilizes a
TAP1 conformation, which is unable to bind ATP, and
subsequently, peptide translocation is inhibited (60).
Overexpression of TAP with IFN-␥ induction overrides
the inhibition of peptide transport by gpUS6, but there are
other proteins of HCMV, expressed in the early phase of
infection, which interfere in a different way with the MHC
class I pathway of antigen presentation (57; for a review,
see Ref. 178).
Another herpes virus that hampers the MHC class I
immune response is the porcine pseudorabies virus (PrV).
It causes a highly infectious disease of pigs and can exist
in a latent form in several tissues, but mainly in the
trigeminal ganglia. PrV infects epithelial cells of mucosal
surfaces and spreads by cell-to-cell contact. It could be
shown that inhibition of peptide transport activity of the
TAP complex is one of the mechanisms by which an early
PrV protein downregulates MHC class I expression (7).
Two viruses that do not belong to the herpes virus
family but interfere with the MHC class I pathway and
TAP function are the adenovirus and the human papilloma virus (HPV) (20, 167). Adenoviruses encode a protein, E3/19K, that directly binds MHC class I molecules,
trapping them in the ER (20). Moreover, E3/19K seems to
bind to TAP and inhibits the tapasin action, thereby preventing MHC class I/TAP association (20). HPV causes
recurrent papillomatosis that is characterized by a variable clinical course and that can include significant morbidity and occasional mortality. Complete eradication of
HPV is rare with repeated surgical excision. The knowledge about the mechanism responsible for the persistence of latent HPV would be valuable for treatment. In
laryngeal papilloma tissue biopsies and in cell culture of
primary explants, a downregulation of TAP1 was observed (167). More significantly, the amount of TAP1 protein expression correlated inversely with the frequency of
197
198
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
C. Autoimmune Disease and Transplantation
The participation of TAP in autoimmune diseases is
discussed controversially. For several patients with diverse MHC-linked autoimmune diseases, such as type I
diabetes, Sjogren’s syndrome, Graves’ disease, and Hashimoto’ disease, a downregulation of mRNA levels for TAP1
and TAP2 was found in varying degrees (43). These data
suggest that defective transcription of TAP genes can
contribute to reduced MHC class I cell surface expression
in autoimmune diseases. Cells lacking MHC class I molecules might then become targets for natural killer cells.
Recently, the regulation of TAP by phosphorylation was
examined (89). Phosphorylated TAP complexes can bind
ATP and peptides but are not capable of transporting
peptides. Several TAP polymorphic residues are potential
phosphorylation sites. Therefore, it was suggested that
abnormal TAP phosphorylation leads to cytotoxic immune responses and subsequent development of certain
autoimmune diseases, which are known to have associations with TAP polymorphism.
MHC class I molecules are not only critical mediators
in host defense against intracellular pathogens, but also
can stimulate strong alloreactions against transplanted
tissue. The incidence and severity of acute rejection after
renal transplantations seem to be influenced by TAP2
gene polymorphism (81). It appears that donor’s antigen
presenting cells with the TAP2*0103 allele have attenuated efficacy in the presentation of allospecific antigens to
Physiol Rev • VOL
the recipient’s T cells, but further investigations are necessary to evaluate the significance of the TAP complex in
rejection after transplantations.
D. TAP Deficiency and Tumor Development
Many tumors escape recognition by CTLs. In some
cancers, including melanomas, this has been associated
with ineffective antigen processing and presentation of
tumor-specific peptides because of low levels of MHC
class I molecules (144, 146). In tumor cell lines from
breast cancer (5) or human small cell lung cancer (144),
low to nondetectable levels of TAP1 and/or TAP2 mRNA
were found, which can be restored by IFN-␥. In murine as
well as human cancers, a downregulation of TAP1 expression by an unknown mechanism or a mutation of TAP
resulting in a loss or decrease in class I surface expression has been demonstrated (25, 128, 143). In some malignant tumor cells, the reduced levels of TAP mRNA
could be restored by IFN-␥ stimulation (143). One type of
human tumor characterized by more than one group of
investigators is breast carcinomas. TAP1 downregulation
was found in 13 and 44% of the lesions (73, 174). The
variations in the results may be determined by characteristics of the patients, the tumor type, or the proliferative
status. The frequency of TAP2 downregulation correlates
with that of TAP1 but was rarely examined. The mRNA
levels of TAP1 and TAP2 can be restored by treatment
with IFN-␥. Recently, it was observed in metastatic melanoma cells that, despite an increase in TAP1 and TAP2
mRNA levels by IFN-␥, a similar increase in cytotoxicity
did not occur (108).
The downregulation of TAP in several tumor cell
lines can be corrected by transfection with the wild-type
form or the treatment with several cytokines, such as
IFN-␥ and TNF-␣. The phenotypic changes are accompanied by an increase in MHC class I expression. Recently,
in a small cell lung carcinoma cell line the potential of
providing exogenous TAP in cancer therapies was tested
(5). With vaccinia virus containing the TAP1 gene, it could
be shown that this virus could reduce tumors in mice and,
therefore, TAP should be considered for inclusion in cancer therapies. In contrast to the above-mentioned cytokines, interleukin-10 leads to TAP downregulation and
subsequently to a reduced cell surface expression of class
I antigens (184). This fact may be of clinical relevance,
since a large number of human tumors secrete interleukin-10.
Downregulation of TAP expression and, thereby, loss
of surface HLA class I molecules, can be a tumor escape
mechanism, but also mutations in the TAP genes, as
shown for example for TAP1 in a small cell lung cancer
(25). This mutation introduces an amino acid exchange at
position 659 in TAP1, resulting in a nonfunctional TAP1
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
In genetic diseases, not only TAP2, but also TAP1
may be involved. In one patient suffering type 1 BLS, a
splice site mutation in the TAP1 gene was described (44).
Interestingly, TAP2 protein was undetectable. Two other
patients, displaying not only the typical lung syndrome of
MHC class I deficiency but also skin lesions, were found
to be deficient of TAP1 (35). In both cases, mutations
introducing a frame shift into the first third of the coding
sequence were observed. The syndromes associated with
TAP1 and TAP2 deficiency are similar and are characterized by the development of chronic inflammation of the
respiratory tract in late childhood and vasculitis at variable stages, but deficiency of TAP is compatible with life.
In later stages, recurrent bacterial infections may lead to
progressive lung damage, respiratory failure, and death.
Early recognition, treatment of the respiratory infections,
and cleansing of skin ulcers of TAP-deficient patients are
the major aims of therapy. Surgical intervention for
chronic sinusitis seems to promote progression of nasal
disease leading to bronchial infections; immunosuppressive and immunomodulatory treatments are associated
with progression of skin lesions (45). Due to the ubiquitous tissue distribution of MHC class I molecules, a gene
therapy of TAP-deficiency syndrome may not be practicable.
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
VIII. CONCLUDING REMARKS
The TAP peptide transporters are a key link between
antigen processing and presentation. The evaluation of its
molecular structure and function will contribute to the
understanding of (patho)physiological processes in which
the TAP complex may be involved. However, many questions remain open to fully understand the whole process
of T-cell immunity, especially regarding the mechanisms
governing peptide binding to MHC class I molecules in the
ER. On one level, the process seems to represent a complex folding event; on another level it is a highly regulated
system ensuring the surface expression of high-affinity
MHC class I-peptide complexes. These complexes are
essential for CD8⫹ T cell-mediated immunity. The elucidation of all steps in the antigen presentation pathway is
Physiol Rev • VOL
essential for many new therapeutic approaches in different diseases. Knowledge about the molecular mechanisms of how viruses evade the immune defense will
identify novel pathogenic factors and will probably allow
the development of new vaccination strategies. Moreover,
studies have revealed that loss of TAP function occurs
more often in metastatic than in primary tumors. The
ability of tumor cells to evade T-cell responses has important implications for the understanding of tumorigenesis. The potential efficacy of immunotherapy for cancer
is dependent on our knowledge of the underlying mechanisms. If mutations in TAP function allow tumor cells to
escape immune surveillance, this may have implications
in the therapeutic strategies. Further analysis will clarify
the interaction of malignant cells with the immune system
of the host and the significance of the loss of TAP function.
The Deutsche Forschungsgemeinschaft funded this research.
Address for reprint requests and other correspondence: R.
Tampé, Institut für Biochemie, Biozentrum Frankfurt, GoetheUniversität Frankfurt, Marie-Curie-Str. 9, D-60439 Frankfurt
a.M., Germany (E-mail: [email protected]).
REFERENCES
1. ABELE R AND TAMPÉ R. Function of the transport complex TAP in
cellular immune recognition. Biochim Biophys Acta 1461: 405– 419,
1999.
2. AHN K, GRUHLER A, GALOCHA B, JONES TR, WIERTZ EJHJ, PLOEGH HL,
PETERSON PA, YANG Y, AND FRÜH K. The ER-luminal domain of the
HCMV glycoprotein US6 inhibits peptide translocation by TAP.
Immunity 6: 613– 621, 1997.
3. AHN K, MEYER TH, UEBEL S, SEMPÉ P, DJABALLAH H, YANG Y, PETERSON
PA, FRÜH K, AND TAMPÉ R. Molecular mechanism and species specificity of TAP inhibition by herpes simplex virus protein ICP47.
EMBO J 15: 3247–3255, 1996.
4. ALBERTS P, DAUMKE O, DEVERSON EV, HOWARD JC, AND KNITTLER MR.
Distinct functional properties of the TAP subunits coordinate the
nucleotide-dependent transport cycle. Curr Biol 11: 242–251, 2001.
5. ALIMONTI J, ZHANG QJ, GABATHULER R, REID G, CHEN S, AND JEFFERIES
WA. Tap expression provides a general method for improving the
recognition of malignant cells in vivo. Nature Biotechnol 18: 515–
520, 2000.
6. ALPAN RS, ZHANG M, AND PARDEE AB. Cell cycle-dependent expression of TAP1, TAP2, and HLA-B27 messenger RNAs in a human
breast cancer cell line. Cancer Res 56: 4358 – 4361, 1996.
7. AMBAGALA APN, HINKLEY S, AND SRIKUMARAN S. An early pseudorabies virus protein down-regulates porcine MHC class I expression
by inhibition of transporter associated with antigen processing.
J Immunol 164: 93–99, 2000.
8. ANDERSON K, CRESSWELL P, GAMMON M, HERMES J, WILLIAMSON A, AND
ZWEERINK H. Endogenously synthesized peptide with an endoplasmic reticulum signal sequence sensitizes antigen processing mutant cells to class I-restricted cell-mediated lysis. J Exp Med 174:
489 – 492, 1991.
9. ANDROLEWICZ MJ, ANDERSON KS, AND CRESSWELL P. 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 USA 90: 9130 –9134, 1993.
10. ANDROLEWICZ MJ AND CRESSWELL P. Human transporters associated
with antigen processing possess a promiscuous peptide-binding
site. Immunity 1: 7–14, 1994.
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
protein. MHC class I molecule expression on the cell
surface could be restored by expression of wild-type
TAP1 or addition of exogenous peptides. The mutation of
TAP1 is located between the C loop and the Walker B
motif, indicating that this mutation might affect ATP binding or hydrolysis. Interestingly, mutations that affect
TAP1 occur frequently in a variety of human tumors, for
example, in primary breast cancer (73, 174). The expression of TAP1 in carcinoma lesions was significantly associated with tumor grading and reduced survival. Like
normal mammary tissue, low-grade breast carcinoma lesions showed no reduced TAP1 protein (174). A similar
situation was found for colorectal carcinomas (74),
whereas adenomas and normal mucose tissue always exhibit no loss of TAP1.
Even if some studies correlate the presence of deficiencies in the MHC class I antigen processing pathway
with tumor progression (73, 144, 174), it is still unclear
whether these deficits have any consequence for tumor
development. To clarify this point, matched panels of
TAP1-positive and TAP1-negative tumor cell lines were
generated (71). Inoculation of mice with TAP1-negative
cells produced large and persistent tumors, whereas in
contrast, TAP1-positive cells did not generate lasting tumors. In athymic mice, both cell lines produced tumors,
confirming that TAP-dependent differences in tumorigenicity were due to T cell-dependent immune responses. It
can be concluded that loss of TAP function allows some
tumor cells to escape immune recognition.
There is evidence that downregulation of MHC class
I cell surface expression is associated with malignant
transformation, and each step in the antigen-processing
pathway can be affected. The target and frequency of the
mutations vary markedly among different types of tumors
and, thereby, the clinical impact of such deficiencies will
vary from one type of tumor to another and needs further
investigations.
199
200
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
Physiol Rev • VOL
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
ated with antigen processing polymorphism for peptide selection.
J Immunol 159: 2350 –2357, 1997.
DASSA E AND HOFNUNG M. Sequence of gene malG in E. coli K12:
homologies between integral membrane components from binding
protein-dependent transport systems. EMBO J 4: 2287–2293, 1985.
DE LA SALLE H, HANAU D, FRICKER D, URLACHER A, KELLY A, SALAMERO
J, POWIS SH, DONATO L, BAUSINGER H, LAFORET M, JERAS M, SPEHNER
D, BIEBER T, FALKENRODT A, CAZENAVE JP, TROWSDALE J, AND TONGIO
MM. Homozygous human TAP peptide transporter mutation in HLA
class I deficiency. Science 265: 237–241, 1994.
DE LA SALLE H, HOUSSAINT E, PEYRAT MA, ARNOLD D, SALAMERO J,
PINCZON D, STEVANOVIC S, BAUSINGER H, FRICKER D, AND GOMARD E.
Human peptide transporter deficiencies: importance of HLA-B in
the presentation of TAP-independent EBV antigens. J Immunol
158: 4555– 4563, 1997.
DE LA SALLE H, ZIMMER J, FRICKER D, ANGENIEUX C, CAZENAVE JP,
OKUBO M, MAEDA H, PLEBANI A, TONGIO MM, DORMOY A, AND HANAU
D. HLA class I deficiencies due to mutations in subunit 1 of the
peptide transporter TAP1. J Clin Invest 103: R9 –R13, 1999.
DEVERSON EV, GOW IR, COADWELL WJ, MONACO JJ, BUTCHER GW, AND
HOWARD JC. MHC class II region encoding proteins related to the
multidrug resistance family of transmembrane transporters. Nature 348: 738 –741, 1990.
DIEDRICH G, BANGIA N, PAN M, AND CRESSWELL P. A role for calnexin
in the assembly of the MHC class I loading complex in the endoplasmic reticulum. J Immunol 166: 1703–1709, 2001.
DOIGE CA AND AMES GFL. ATP-dependent transport systems in
bacteria and humans: relevance to cystic fibrosis and multidrug
resistance. Annu Rev Microbiol 47: 291–319, 1993.
EHRING B, MEYER TH, ECKERSKORN C, LOTTSPEICH F, AND TAMPÉ R.
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 235:
404 – 415, 1996.
FENTEANY G, STANDAERT RF, LANE WS, CHOI S, COREY EJ, AND SCHREIBER SL. Inhibition of proteasome activities and subunit-specific
amino-terminal threonine modification by lactacystin. Science 268:
726 –731, 1995.
FRÜH K, AHN K, DJABALLAH H, SEMPÉ P, VAN ENDERT PM, TAMPÉ R,
PETERSON PA, AND YANG Y. A viral inhibitor of peptide transporters
for antigen presentation. Nature 375: 415– 418, 1995.
FRÜH K AND YANG Y. Antigen presentation by MHC class I and its
regulation by interferon gamma. Curr Opin Immunol 11: 76 – 81,
1999.
FU Y, YAN G, SHI L, AND FAUSTMAN D. Antigen processing and
autoimmunity. Evaluation of mRNA abundance and function of
HLA-linked genes. Ann NY Acad Sci 842: 138 –155, 1998.
FURUKAWA H, MURATA S, YABE T, SHIMBARA N, KEICHO N, KASHIWASE K,
WATANABE K, ISHIKAWA Y, AKAZA T, TADOKORO K, TOHMA S, INOUE T,
TOKUNAGA K, YAMAMOTO K, TANAKA K, AND JUJI T. Splice acceptor site
mutation of the transporter associated with antigen processing-1
gene in human bare lymphocyte syndrome. J Clin Invest 103:
755–758, 1999.
GADOLA D, MOINS-TESSERENC HT, TROWSDALE J, GROSS WL, AND
CERUNDOLO V. Tap deficiency syndrome. Clin Exp Immunol 121:
173–178, 2000.
GALOCHA B, HILL A, BARNETT BC, DOLAN A, RAIMONDI A, COOK RF,
BRUNNER J, MCGEOCH DJ, AND PLOEGH HL. The active site 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 185: 1565–1572, 1997.
GARBOCZI DN, GHOSH P, UTZ U, FAN QR, BIDDISON WE, AND WILEY DC.
Structure of the complex between human T-cell receptor, viral
peptide and HLA-A2. Nature 384: 134 –141, 1996.
GARCIA KC, DEGANO M, STANFIELD RL, BRUNMARK A, JACKSON MR,
PETERSON PA, TEYTON L, AND WILSON IA. An alpha/beta T-cell receptor structure at 2.5 Å and its orientation in the TCR-MHC complex.
Science 274: 209 –219, 1996.
GEIER E, PFEIFER G, WILM M, LUCCHIARI-HARTZ M, BAUMEISTER W,
EICHMANN K, AND NIEDERMANN G. A giant protease with potential to
substitute for some functions of the proteasome. Science 283:
978 –981, 1999.
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
11. ANDROLEWICZ MJ, ORTMANN B, VAN ENDERT PM, SPIES T, AND CRESSWELL P. Characteristics of peptide and major histocompatibility
complex class I/beta 2-microglobulin binding to the transporter
associated with antigen processing (TAP1 and TAP2). Proc Natl
Acad Sci USA 91: 12716 –12720, 1994.
12. ATTAYA M, JAMESON S, MARTINEZ CK, HERMEL E, ALDRICH C, FORMAN J,
LINDAHL KF, BEVAN MJ, AND MONACO JJ. Ham-2 corrects the class I
antigen-processing defect in RMA-S cells. Nature 355: 647– 649,
1992.
13. AYALON O, HUGHES EA, CRESSWELL P, LEE J, O’DONNELL L, PARDI R,
AND BENDER JR. Induction of transporter associated with antigen
processing by interferon gamma confers endothelial cell cytoprotection against natural killer-mediated lysis. Proc Natl Acad Sci
USA 95: 2435–2440, 1998.
14. BANGIA N, LEHNER PJ, HUGHES EA, SURMAN M, AND CRESSWELL P. The
N-terminal region of tapasin is required to stabilize the MHC class
I loading complex. Eur J Immunol 29: 1858 –1870, 1999.
15. BAUMEISTER W, WALZ J, ZÜHL F, AND SEEMÜLLER E. The proteasome:
paradigm of a self-compartmentalizing protease. Cell 92: 367–380,
1998.
16. BECK S, ABDULLA S, ALDERTON R, GLYNNE PR, GUT II, HOSKING GL,
JACKSON KA, KELLY A, NEWELL W, SANSEAU RP, RADLEY E, THORPE K,
AND TROWSDALE J. Evolutionary dynamics of non-coding sequences
within the class II region of the human MHC. J Mol Biol 255: 1–13,
1996.
17. BECK S, KELLY A, RADLEY E, KHURSHID F, ALDERTON RP, AND TROWSDALE J. DNA sequence analysis of 66 kb of the human MHC class II
region encoding a cluster of genes for antigen processing. J Mol
Biol 228: 433– 441, 1992.
18. BEINERT D, NEUMANN L, UEBEL S, AND TAMPÉ R. Structure of the viral
TAP-inhibitor ICP47 induced by membrane association. Biochemistry 36: 4694 – 4700, 1997.
19. BENINGA J, ROCK KL, AND GOLDBERG AL. Interferon-gamma can stimulate post-proteasomal trimming of the N terminus of an antigenic
peptide by inducing leucine aminopeptidase. J Biol Chem 273:
18734 –18742, 1998.
20. BENNETT EM, BENNINK JR, YEWDELL JW, AND BRODSKY FM. Cutting
edge: adenovirus E19 has two mechanisms for affecting class I
MHC expression. J Immunol 162: 5049 –5052, 1999.
21. BJORKMAN PJ, SAPER MA, AND WILEY DC. Structure of human class I
histocompatibility antigen, HLA-A2. Nature 329: 506 –512, 1987.
22. BROWN MG, DRISCOLL J, AND MONACO JJ. Structural and serological
similarity of MHC-linked LMP and proteasome (multicatalytic proteinase) complexes. Nature 353: 355–357, 1991.
23. CASCIO P, HILTON C, KISSELEV AF, ROCK KL, AND GOLDBERG AL. 26S
proteasomes and immunoproteasomes produce mainly N-extended
versions of an antigenic peptide. EMBO J 20: 2357–2366, 2001.
24. CERUNDOLO V, ALEXANDER J, ANDERSON K, LAMB C, CRESSWELL P,
MCMICHEAL A, GOTCH F, AND TOWNSEND A. Presentation of viral
antigen controlled by a gene in the major histocompatibility complex. Nature 345: 449 – 452, 1990.
25. CHEN HL, GABRILOVICH D, TAMPÉ R, GIRGIS KR, NADAF S, AND CARBONE
DP. A functionally defective allele of TAP1 results in loss of MHC
class I antigen presentation in a human lung cancer. Nat Genet 13:
210 –213, 1996.
26. COLONNA M, BRESNAHAN M, BAHRAM S, STROMINGER JL, AND SPIES T.
Allelic variants of the human putative peptide transporter involved
in antigen processing. Proc Natl Acad Sci USA 89: 3932–3936, 1992.
27. COTTEN JF, OSTEDGAARD LS, CARSON MR, AND WELSH MJ. Effect of
cystic fibrosis-associated mutations in the fourth intracellular loop
of cystic fibrosis transmembrane conductance regulator. J Biol
Chem 271: 21270 –21284, 1996.
28. COUX O, TANAKA K, AND GOLDBERG AL. Structure and functions of the
20S and 26S proteasomes. Annu Rev Biochem 65: 801– 847, 1996.
29. CRESSWELL P, BANGIA N, DICK T, AND DIEDRICH G. The nature of the
MHC class I peptide loading complex. Immunol Rev 172: 21–28,
1999.
30. CROMME FV, AIREY J, HEEMELS MT, PLOEGH HL, KEATING PJ, STERN PL,
MEIJER CJ, AND WALBOOMERS JM. Loss of transporter protein, encoded by the TAP-1 gene, is highly correlated with loss of HLA
expression in cervical carcinomas. J Exp Med 179: 335–340, 1994.
31. DANIEL S, CAILLAT-ZUCMAN S, HAMMER J, BACH JF, AND VAN ENDERT
PM. Absence of functional relevance of human transporter associ-
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
Physiol Rev • VOL
73. KAKLAMANIS L, LEEK R, KOUKOURAKIS M, GATTER KC, AND HARRIS AL.
Loss of transporter antigen processing 1 transport protein and
major histocompatibility complex class I molecules in metastatic
versus primary breast cancer. Cancer Res 55: 5191–5194, 1995.
74. KAKLAMANIS L, TOWNSEND A, DOUSSIS-ANAGNOSTOPOULOU IA,
MORTENSEN N, HARRIS AL, AND GATTER KC. Loss of major histocompatibility complex-encoded transporter associated with antigen
presentation (TAP) in colorectal cancer. Am J Pathol 145: 505–509,
1994.
75. KELLAR-WOOD HF, POWIS SH, GRAY J, AND COMPSTON DAS. MHCencoded TAP1 and TAP2 dimorphisms in multiple sclerosis. Tissue
Antigens 43: 129 –132, 1994.
76. KELLY A, POWIS SH, KERR LA, MOCKRIDGE I, ELLIOTT T, BASTIN J,
UCHANSKA-ZIEGLER B, ZIEGLER A, TROWSDALE J, AND TOWNSEND A.
Assembly and function of the two ABC transporter proteins encoded in the human major histocompatibility complex. Nature 355:
641– 644, 1992.
77. KISSELEV AF, AKOPIAN TN, WOO KM, AND GOLDBERG AL. The sizes of
peptides generated from protein by mammalian 26 and 20 S proteasomes. Implications for understanding the degradative mechanism and antigen presentation. J Biol Chem 274: 3363–3371, 1999.
78. KLEIJMEER MJ, KELLY A, GEUZE HJ, SLOT JW, TOWNSEND A, AND
TROWSDALE J. Location of MHC-encoded transporters in the endoplasmic reticulum and cis-Golgi. Nature 357: 342–344, 1992.
79. KLEIN I, SARKADI B, AND VÁRADI A. An inventory of the human ABC
proteins. Biochim Biophys Acta 1461: 237–262, 1999.
80. KNITTLER MR, ALBERTS P, DEVERSON EV, AND HOWARD JC. Nucleotide
binding by TAP mediates association with peptide and release of
assembled MHC class I molecules. Curr Biol 9: 999 –1008, 1999.
81. KOBAYASHI T, YOKOYAMA I, INOKO H, NARUSE T, HAYASHI S, MOROZUMI
K, UCHIDA K, AND NAKAO A. Significance of transporter associated
with antigen processing gene polymorphism in living related renal
transplantation. Hum Immunol 61: 670 – 674, 2000.
82. KOOPMANN JO, ALBRING J, HUTER E, BULBUC N, SPEE P, NEEFJES J,
HÄMMERLING GJ, AND MOMBURG F. Export of antigenic peptides from
the endoplasmic reticulum intersects with retrograde protein translation through the Sec61p channel. Immunity 13: 117–127, 2000.
83. KOOPMANN JO, POST M, NEEFJES JJ, HÄMMERLING GJ, AND MOMBURG F.
Translocation of long peptides by transporter associated with antigen processing (TAP). Eur J Immunol 26: 1729 –1728, 1996.
84. LAPINSKI PE, NEUBIG RR, AND RAGHAVAN M. Walker A lysine mutations of TAP1 and TAP2 interfere with peptide translocation but
not peptide binding. J Biol Chem 276: 7526 –7533, 2001.
85. LEHNER PJ, KARTTUNEN JT, WILKINSON GWG, AND CRESSWELL P. The
human cytomegalovirus US6 glycoprotein inhibits transporter associated with antigen processing-dependent peptide translocation.
Proc Natl Acad Sci USA 94: 6904 – 6909, 1997.
86. LEHNER PJ, SURMAN MJ, AND CRESSWELL P. Soluble tapasin restores
MHC class I expression and function in the tapasin-negative cell
line 220. Immunity 8: 221–231, 1998.
87. LEHNER PJ AND TROWSDALE J. Antigen presentation: coming out
gracefully. Curr Biol 8: R605–R608, 1998.
88. LI SL, SJOGREN HO, HELLMAN U, PETTERSSON RF, AND WANG P. Cloning and functional characterization of a subunit of the transporter
associated with antigen processing. Proc Natl Acad Sci USA 94:
8708 – 8713, 1997.
89. LI Y, SALTER-CID L, VITIELLO A, PRECKEL T, LEE JD, ANGULO A, CAI Z,
PETERSON PA, AND YANG Y. Regulation of transporter associated
with antigen processing by phosphorylation. J Biol Chem 275:
24130 –24135, 2000.
90. LINDQUIST JA, JENSEN ON, MANN M, AND HÄMMERLING GJ. Er-60, a
chaperone with thiol-dependent reductase activity involved in
MHC class I assembly. EMBO J 17: 2186 –2195, 1998.
91. LJUNGGREN HG, STAM NJ, OHLEN C, NEEFJES JJ, HOGLUND P, HEEMELS
MT, BASTIN J, SCHUMACHER TN, TOWNSEND A, KARRE K, AND PLOEGH
HL. Empty MHC class I molecules come out in the cold. Nature
346: 476 – 480, 1990.
92. LOO TW AND CLARKE DM. Membrane topology of a cystein-less
mutant of human P-glycoprotein. J Biol Chem 270: 843– 838, 1995.
93. LÖWE J, STOCK D, JAP B, ZWICKL P, BAUMEISTER W, AND HUBER R.
Crystal structure of the 20S proteasome from the archaeon T.
acidophilum at 34 Å resolution. Science 268: 533–539, 1995.
94. MA W, LEHNER PJ, CRESSWELL P, POBER JS, AND JOHNSON DR. Inter-
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
50. GLAS R, BOGYO M, MCASTER JS, GACZYNSKA M, AND PLOEGH HL. A
proteolytic system that compensates for loss of proteasome function. Nature 392: 618 – 622, 1998.
51. GLYNNE R, POWIS SH, BECK S, KELLY A, KERR LA, AND TROWSDALE J. A
proteasome-related gene between the two ABC transporter loci in
the class II region of the human MHC. Nature 353: 357–360, 1991.
52. GOLDBERG AL AND ST. JOHN AC. Intracellular protein degradation in
mammalian and bacterial cells: part 2. Annu Rev Biochem 45:
747– 803, 1976.
53. GORBULEV S, ABELE R, AND TAMPÉ R. Allosteric crosstalk between
peptide-binding, transport, and ATP hydrolysis of the ABC transporter TAP. Proc Natl Acad Sci USA 98: 3732–3737, 2001.
54. HANSON IM AND TROWSDALE J. Colinearity of novel genes in the class
II regions of the MHC in mouse and human. Immunogenetics 34:
5–11, 1991.
55. HARRIS MR, LYBARGER L, YU YYL, MYERS NB, AND HANSEN TH. Association of ERp57 with mouse MHC class I molecules is tapasin
dependent and mimics that of calreticulin and not calnexin. J Immunol 166: 6686 – 6692, 2001.
56. HEEMELS MT, SCHUMACHER TN, WONIGEIT K, AND PLOEGH HL. Peptide
translocation by variants of the transporter associated with antigen
processing. Science 262: 2059 –2063, 1993.
57. HENGEL H, FLOHR T, HÄMMERLING GJ, KOSZINOWSKI UH, AND MOMBURG
F. Human cytomegalovirus inhibits peptide translocation into the
endoplasmic reticulum for MHC class I assembly. J Gen Virol 77:
2287–2296, 1996.
58. HENGEL H, KOOPMANN JO, FLOHR T, MURANYI W, GOULMY E, HÄMMERLING GJ, KOSZINOWSKI UH, AND MOMBURG F. A viral ER-resident
glycoprotein inactivates the MHC-encoded peptide transporter. Immunity 6: 623– 632, 1997.
59. HENGEL H AND KOSZINOWSKI UH. Interference with antigen processing by viruses. Curr Opin Immunol 9: 470 – 476, 1997.
60. HEWITT EW, GUPTA SS, AND LEHNER PJ. The human cytomegalovirus
gene product US6 inhibits ATP binding by TAP. EMBO J 20: 387–
396, 2001.
61. HIGGINS CF. ABC transporters: from microorganisms to man. Annu
Rev Cell Biol 8: 67–113, 1992.
62. HILL A, JUGOVIC R, YORK I, RUSS G, BENNINK J, YEWDELL J, PLOEGH H,
AND JOHNSON D. Herpes simplex virus turns off the TAP to evade
host immunity. Nature 375: 411– 415, 1995.
63. HILTON CJ, DAHL AM, AND ROCK KL. Anti-peptide antibody blocks
peptide binding to MHC class I molecules in the endoplasmic
reticulum. J Immunol 166: 3952–3956, 2001.
64. HINKLEY S, HILL AB, AND SRIKUMARAN S. Bovine herpesvirus-1 infection affects the peptide transport activity in bovine cells. Virus Res
53: 91–96, 1998.
65. HOSKEN NA AND BEVAN MJ. Defective presentation of endogenous
antigen by a cell line expressing class I molecules. Science 248:
367–370, 1990.
66. HUGHES AL. Evolution of the ATP-binding-cassette transmembrane
transporters of vertebrates. Mol Biol Evol 11: 899 –910, 1994.
67. HUGHES EA AND CRESSWELL P. The thiol oxidoreductase ERp57 is a
component of the MHC class I peptide-loading complex. Curr Biol
8: 709 –712, 1998.
68. ISRAEL A, LE BAIL O, HATAT D, PIETTE J, KIERAN M, LOGEAT F, FELLOUS
M, AND KOURILSKY P. TNF stimulates expression of mouse MHC
class I genes by inducing an NF kappa B-like enhancer binding
activity which displaces constitutive factors. EMBO J 8: 3793–3800,
1989.
69. JACKSON DG AND CAPRA JD. Tap1 alleles in insulin-dependent diabetes mellitus: a newly defined centromeric boundary of disease
susceptibility. Proc Natl Acad Sci USA 90: 11079 –11083, 1993.
70. JACKSON MR, COHEN-DOYLE MF, PETERSON PA, AND WILLIAMS DB.
Regulation of MHC class I transport by the molecular chaperone
calnexin. Science 263: 384 –387, 1994.
71. JOHNSEN AK, TEMPLETON DJ, SY MS, AND HARDING CV. Deficiency of
transporter for antigen presentation (TAP) in tumor cells allows
evasion of immune surveillance and increases tumorigenesis. J Immunol 163: 4224 – 4231, 1999.
72. JOLY AL, LE ROLLE AL, GONZALEZ AL, MEHLING WJ, STEVENS WJ,
COADWELL WJ, HUNIG JC, HOWARD JC, AND BUTCHER GW. Co-evolution of rat TAP transporters and MHC class I RT1-A molecules.
Curr Biol 8: 169 –172, 1998.
201
202
95.
96.
97.
98.
99.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
feron-gamma rapidly increases peptide transporter (TAP) subunit
expression and peptide transport capacity in endothelial cells.
J Biol Chem 272: 16585–16590, 1997.
MADDEN DR, GARBOCZI DN, AND WILEY DC. The antigenic identity of
peptide-MHC complexes: a comparison of the conformations of
five viral peptides presented by HLA-A2. Cell 75: 693–708, 1993.
MADDEN DR, GORGA JC, STROMINGER JL, AND WILEY DC. The structure
of HLA-B27 reveals nonamer self-peptides bound in an extended
conformation. Nature 353: 321–325, 1991.
MARTINEZ CK AND MONACO JJ. Homology of proteasome subunits to
a major histocompatibility complex-linked LMP gene. Nature 353:
664 – 667, 1991.
MEYER TH, VAN ENDERT PM, UEBEL S, EHRING B, AND TAMPÉ R.
Functional expression and purification of the ABC transporter
complex associated with antigen processing (TAP) in insect cells.
FEBS Lett 35: 443– 447, 1994.
MOINS-TEISSERENC H, SEMANA G, ALIZADEH M, LOISEAU P, BOBRYNINA V,
DESCHAMPS I, EDAN G, BIREBENT B, GENETET B, SABOURAUD O, AND
CHARRON D. Tap2 gene polymorphism contributes to genetic susceptibility to multiple sclerosis. Hum Immunol 42: 195–202, 1995.
MOINS-TEISSERENC HT, GADOLA SD, CELLA M, DUNBAR PR, EXLEY A,
BLAKE N, BAYCAL C, LAMBERT J, BIGLIARDI P, WILLEMSEN M, JONES M,
BUECHNER S, COLONNA M, GROSS WL, AND CERUNDOLO V. Association
of a syndrome resembling Wegener’s granulomatosis with low
surface expression of HLA class I molecules. Lancet 354: 1598 –
1603, 1999.
MOMBURG F, ROELSE J, HÄMMERLING GJ, AND NEEFJES JJ. Peptide size
selection by the major histocompatibility complex-encoded peptide transporter. J Exp Med 179: 1613–1623, 1994.
MOMBURG F, ROELSE J, HOWARD JC, BUTCHER GW, HÄMMERLING GJ,
AND NEEFJES JJ. Selectivity of MHC-encoded peptide transporters
from human, mouse and rat. Nature 367: 648 – 651, 1994.
MONACO JJ, CHO S, AND ATTAYA M. Transport protein genes in the
murine MHC: possible implications for antigen processing. Science
250: 1723–1726, 1990.
MOORE MW, CARBONE FR, AND BEVAN MJ. Introduction of soluble
protein into class I pathway of antigen processing and presentation. Cell 54: 777–785, 1988.
MORRICE NA AND POWIS SJ. A role for the thiol-dependent reductase
ERp57 in the assembly of MHC class I molecules. Curr Biol 8:
713–716, 1998.
MORRISON LA, LUKACHER AE, BRACIALE VL, FAN DP, AND BRACIALE TJ.
Difference in antigen presentation to MHC class I- and class IIrestricted influenza virus-specific cytolytic T-lymphocyte clones. J
Exp Med 163: 903–921, 1986.
MÜLLER KM, EBENSPERGER C, AND TAMPÉ R. Nucleotide binding to the
hydrophilic C-terminal domain of the transporter associated with
antigen processing (TAP). J Biol Chem 269: 14032–14037, 1994.
MURRAY JL, HUDSON M, ROSS MI, ZHANG HZ, AND IOANNIDES CG.
Reduced recognition of metastatic melanoma cells by autologous
MART-1 specific CTL: relationship to TAP expression. J Immunotherapy 23: 28 –35, 2000.
NAGY N, VANKY F, AND KLEIN E. Tumor surveillance: expression of
the transporter associated with antigen processing (TAP-1) in ex
vivo human tumor samples and its elevation by in vitro treatment
with IFN-␥ and TNF-␣. Immunol Lett 64: 153–160, 1998.
NEEFJES JJ, MOMBURG F, AND HÄMMERLING GJ. Selective and ATPdependent translocation of peptides by the MHC-encoded transporter. Science 261: 769 –771, 1993.
NEUMANN L, KRAAS W, UEBEL S, JUNG G, AND TAMPÉ R. The activation
domain of the herpes simplex virus protein ICP47: a potent inhibitor of the transporter associated with antigen processing (TAP). J
Mol Biol 272: 484 – 492, 1997.
NEUMANN L AND TAMPÉ R. Kinetic analysis of peptide binding to the
TAP transport complex: evidence for structural rearrangements
induced by substrate binding. J Mol Biol 294: 1203–1213, 1999.
NIJENHUIS M AND HÄMMERLING GJ. Multiple regions of the transporter
associated with the antigen processing (TAP) contribute to its
peptide binding site. J Immunol 157: 5467–5477, 1996.
OBST R, ARMANDOLA EA, NIJENHUIS M, MOMBURG F, AND HÄMMERLING
GJ. Tap polymorphism does not influence transport of peptide
variants in mice and humans. Eur J Immunol 25: 2170 –2176, 1995.
Physiol Rev • VOL
115. ORTMANN B, COPEMAN J, LEHNER PJ, SADASIVAN B, HERBERG JA, GRANDEA AG, RIDDELL SR, TAMPÉ R, SPIES T, AND TROWSDALE J. A critical
role for tapasin in the assembly and function of multimeric MHC
class I-TAP complexes. Science 277: 1306 –1309, 1997.
116. PAMER E AND CRESSWELL P. Mechanisms of MHC class I-restricted
antigen processing. Annu Rev Immunol 16: 323–358, 1998.
117. PARHAM P, ADAMS EJ, AND ARNETT KL. The origins of HLA-A, B, C
polymorphism. Immunol Rev 143: 141–180, 1995.
118. PETERS JM. Proteasomes: protein degradation machines of the cell.
Trends Biochem Sci 19: 377–382, 1994.
119. PFÄNDER R, NEUMANN L, ZWECHSTETTER M, SEGER C, HOLAK TA, AND
TAMPÉ R. Structure of the active domain of the herpes simplex virus
protein ICP47 in water/sodium dodecyl sulfate solution determined
by nuclear magnetic resonance spectroscopy. Biochemistry 38:
13692–13698, 1999.
120. PLOEGH HL. Viral strategies of immune evasion. Science 280: 248 –
253, 1998.
121. POWIS SJ, DEVERSON EV, COADWELL WJ, CIRUELA A, HUSKISSON NS,
SMITH H, BUTCHER GW, AND HOWARD JC. Effect of polymorphism of
an MHC-linked transporter on the peptides assembled in a class I
molecule. Nature 357: 211–215, 1992.
122. POWIS SH, MOCKRIDGE I, KELLY A, KERR LA, GLYNNE R, GILEADI U,
BECK S, AND TROWSDALE J. Polymorphism in a second ABC transporter gene located within the class II region of the human major
histocompatibility complex. Proc Natl Acad Sci USA 89: 1463–1467,
1992.
123. POWIS SJ, TONKS S, MOCKRIDGE I, KELLY AP, BODMER JG, AND TROWSDALE J. Alleles and haplotypes of the MHC-encoded ABC transporters TAP1 and TAP2. Immunogenetics 37: 373–380, 1993.
124. POWIS SJ, TOWNSEND AR, DEVERSON EV, BASTIN J, BUTCHER GW, AND
HOWARD JC. Restoration of antigen presentation to the mutant cell
line RMA-S by an MHC-linked transporter. Nature 354: 528 –531,
1991.
125. POWIS SJ, YOUNG LL, JOLY E, BARKER PJ, RICHARDSON L, BRANDT RP,
MELIEF CJ, HOWARD JC, AND BUTCHER GW. The rat cim effect: TAP
allele-dependent changes in a class I MHC anchor motif and evidence against C-terminal trimming of peptides in the ER. Immunity 4: 159 –162, 1996.
126. RECHSTEINER M, HOFFMAN L, AND DUBIEL W. The multicatalytic and
26S proteases. J Biol Chem 268: 6065– 6068, 1993.
127. REITS EAJ, VOS JC, GROMME M, AND NEEFJES J. The major substrates
for TAP in vivo are derived from newly synthesized proteins.
Nature 404: 774 –778, 2000.
128. RESTIFO NP, ESQUIVEL F, ASTHER AL, STOTTER H, BARTH RJ, BENNINK
JR, MULE JJ, YEWDELL JW, AND ROSENBERG SA. Defective presentation of endogenous antigens by a murine sarcoma. Implications for
the failure of an anti-tumor immune response. J Immunol 147:
1453–1459, 1991.
129. RESTIFO NP, ESQUIVEL F, KAWAKAMI Y, YEWDELL JW, MULE JJ, ROSENBERG SA, AND BENNINK JR. Identification of human cancers deficient
in antigen processing. J Exp Med 177: 265–272, 1993.
130. RITZ U, MOMBURG F, PIRCHNER HP, STRAND D, HUBER C, AND SELIGER
B. Identification of sequences in the human peptide transporter
subunit TAP1 required for transporter associated with antigen
processing (TAP) function. Int Immunol 13: 31– 41, 2001.
131. ROCK KL AND GOLDBERG AL. Degradation of cell proteins and the
generation of MHC class I-presented peptides. Annu Rev Immunol
17: 739 –779, 1999.
132. ROELSE J, GROMME M, MOMBURG F, HÄMMERLING G, AND NEEFJES J.
Trimming of TAP-translocated peptides in the endoplasmic reticulum and in the cytosol during recycling. J Exp Med 180: 1591–1597,
1994.
133. RUEDA FAUCZ FR, MACAGNAN PROBST C, AND PETZL-ERLER ML. Polymorphism of LMP2, TAP1, LMP7 and TAP2 in Brazilian amerindians and caucasoids: implications for the evolution of allelic and
haplotypic diversity. Eur J Immunogenet 27: 5–16, 2000.
134. RUSS G, ESQUIVEL F, YEWDELL JW, CRESSWELL P, SPIES T, AND BENNINK
JR. Assembly, intracellular localization, and nucleotide binding
properties of the human peptide transporters TAP1 and TAP2
expressed by recombinant vaccinia viruses. J Biol Chem 270:
21312–21318, 1995.
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
100.
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ
INTRACELLULAR PEPTIDE TRANSPORT IN ANTIGEN PROCESSING
Physiol Rev • VOL
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
I processing and presentation pathway. Immunol Lett 57: 183–187,
1997.
THE WHO NOMENCLATURE COMMITTEE. Factors of the HLA system.
Immunogenetics 36: 135–148, 1991.
TOMAZIN R, HILL AB, JUGOVIC P, YORK I, VAN ENDERT P, PLOEGH HL,
ANDREWS DW, AND JOHNSON DC. Stable binding of the herpes simplex virus ICP47 protein to the peptide binding site of TAP. EMBO
J 15: 3256 –3266, 1996.
TOWNSEND A, ÖHLEN C, BASTIN J, LJUNGGREN HG, FOSTER L, AND KÄRRE
K. Association of class I major histocompatibility heavy and light
chains induced by viral peptides. Nature 340: 443– 448, 1989.
TOWNSEND A, ÖHLEN C, FOSTER L, BASTIN J, LJUNGGREN HG, AND KÄRRE
K. A mutant cell in which association of class I heavy and light
chains is induced by viral peptides. Cold Spring Harb Symp Quant
Biol 54: 299 –308, 1989.
TOWNSEND AR, BASTIN J, GOULD K, AND BROWNLEE GG. Cytotoxic
T-lymphocytes recognize influenza haemagglutinin that lacks a signal sequence. Nature 324: 575–577, 1986.
TROWSDALE J, HANSON I, MOCKRIDGE I, BECK S, TOWNSEND A, AND KELLY
A. Sequences encoded in the class II region of the MHC related to
the ABC superfamily of transporters. Nature 348: 741–744, 1990.
TROWSDALE J, RAGOUSSIS J, AND CAMPBELL DR. Map of the human
MHC. Immunol Today 12: 443– 446, 1991.
TURNER GC AND VARSHAVSKI A. Detecting and measuring cotranslational protein degradation in vivo. Science 289: 2117–2120, 2000.
UEBEL S, KRAAS W, KIENLE S, WIESMÜLLER KH, JUNG G, AND TAMPÉ R.
Recognition principle of the TAP transporter disclosed by combinatorial peptide libraries. Proc Natl Acad Sci USA 94: 8976 – 8981,
1997.
UEBEL S, MEYER TH, KRASS W, KIENLE S, JUNG G, WIESMÜLLER KH, AND
TAMPÉ R. Requirements for peptide binding to the human transporter associated with antigen processing revealed by peptide
scans and complex peptide libraries. J Biol Chem 270: 18512–
18516, 1995.
UEBEL S AND TAMPÉ R. Specificity of the proteasome and the TAP
transporter. Curr Opin Immunol 11: 203–208, 1999.
URLINGER S, KUCHLER K, MEYER TH, UEBEL S, AND TAMPÉ R. Intracellular location, complex formation, and function of the transporter
associated with antigen processing in yeast. Eur J Biochem 245:
266 –272, 1997.
VAMBUTAS A, BONAGURA VR, AND STEINBERG BM. Altered expression
of TAP1 and major histocompatibility complex class I in laryngeal
papillomatosis: correlation of TAP1 with disease. Clin Diagn Lab
Immunol 7: 79 – 85, 2000.
VAN ENDERT PM, LIBLAU RS, PATEL SD, FUGGER L, LOPEZ T, POCIOT F,
NERUP J, AND MCDEVITT HO. Major histocompatibility complexencoded antigen processing gene polymorphism in IDDM. Diabetes
43: 110 –117, 1994.
VAN ENDERT PM, RIGANELLI D, GRECO G, FLEISCHHAUER K, SIDNEY J,
SETTE A, AND BACH JF. The peptide-binding motif for the human
transporter associated with antigen processing. J Exp Med 182:
1883–1895, 1995.
VAN ENDERT PM, TAMPÉ R, MEYER TH, TISCH R, BACH JF, AND MCDEVITT HO. A sequential model for peptide binding and transport by
the transporters associated with antigen processing. Immunity 1:
491–500, 1994.
VAN KAER L, ASHTON-RICKARDT PG, PLOEGH HL, AND TONEGAWA S.
Tap1 mutant mice are deficient in antigen presentation, surface
class I molecules, and CD4 – 8⫹ T-cells. Cell 71: 1205–1214, 1992.
VASSILAKOS A, COHEN-DOYLE MF, PETERSON PA, JACKSON MR, AND
WILLIAMS DB. The molecular chaperone calnexin facilitates folding
and assembly of class I histocompatibility molecules. EMBO J 15:
1495–1506, 1996.
VINITSKY A, ANTON LC, SNYDER HL, ORLOWSKI M, BENNINK JR, AND
YEWDELL JW. The generation of MHC class I-associated peptides is
only partially inhibited by proteasome inhibitors: involvement of
nonproteasomal cytosolic proteases in antigen processing? J Immunol 159: 554 –564, 1997.
VITALE M, REZZANI R, RODELLA L, ZAULI G, GRIGOLATO P, CADEI M,
HICKLIN DJ, AND FERRONE S. HLA class I antigen and transporter
associated with antigen processing (TAP1 and TAP2) down-regu-
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
135. SADASIVAN B, LEHNER PJ, ORTMANN B, SPIES T, AND CRESSWELL P.
Roles for calreticulin and a novel glycoprotein, tapasin, in the
interaction of MHC class I molecules with TAP. Immunity 5:
103–114, 1996.
136. SALAZAR-ONFRAY F, CHARO J, PETERSSON M, FRELAND S, NOFFZ G, QIN
Z, BLANKENSTEIN T, LJUNGGREN HG, AND KIESSLING R. Down-regulation of the expression and function of the transporter associated
with antigen processing in murine tumor cell lines expressing IL-10.
J Immunol 159: 3159 –3202, 1997.
137. SAVEANU L, DANIEL S, AND VAN ENDERT PM. Distinct functions of the
ATP-binding cassettes of the transporters associated with antigen
processing: a mutational analysis of Walker A and B sequences.
J Biol Chem 276: 22107–22113, 2001.
138. SCHENDEL DJ, FALK CS, NOSSNER E, MAGET B, KRESSENSTEIN S, URLINGER S, TAMPÉ R, AND GANSBACHER B. Gene transfer of human
interferon gamma complementary DNA into a renal cell carcinoma
line enhances MHC-restricted cytotoxic T-lymphocyte recognition
but does not suppress non-MHC-restricted effector cell activity.
Gene Ther 7: 950 –959, 2000.
139. SCHUBERT U, ANTÓN LC, GIBBS J, NORBURY CC, YEWDELL JW, AND
BENNINK JR. Rapid degradation of a large fraction of newly synthesized proteins by proteasomes. Nature 404: 770 –774, 2000.
140. SCHUMACHER TN, KANTESARIA DY, HEEMELS MT, ASHTON-RICKARDT PG,
SHEPHERD JC, FRÜH K, YANG Y, PETERSON PA, TONEGAWA S, AND
PLOEGH HL. Peptide length and sequence specificity of the mouse
TAP1/TAP2 translocator. J Exp Med 179: 533–540, 1994.
141. SCHUMACHER TNM, KANTESARIA DV, SEREZE DV, ROOPENIAN DC, AND
PLOEGH HL. Transporters from H-2b, H-2d, H-2s, H-2k and H-2g7
(NOD/Lt) haplotype translocate similar sets of peptides. Proc Natl
Acad Sci USA 91: 13004 –13008, 1994.
142. SEEMÜLLER E, LUPAS A, STOCK D, LÖWE J, HUBER R, AND BAUMEISTER
W. Proteasome from Thermoplasma acidophilum: a threonine
protease. Science 268: 579 –582, 1995.
143. SELIGER B, HOHNE A, JUNG D, KALLFELZ M, KNUTH A, JAEGER E,
BERNHARD H, MOMBURG F, TAMPÉ R, AND HUBER C. Expression and
function of the peptide transporters in escape variants of human
renal cell carcinomas. Exp Hematol 25: 608 – 614, 1997.
144. SELIGER B, MAEURER MJ, AND FERRONE S. Tap off—tumors on. Immunol Today 18: 292–299, 1997.
145. SHEPHERD JC, SCHUMACHER TNM, ASHTON-RICKARDT PG, IMAEDA S,
PLOEGH HL, JANEWAY CA, AND TONEGAWA S. Tap1-dependent peptide
translocation in vitro is ATP-dependent and peptide selective. Cell
74: 577–584, 1993.
146. SHERMAN LA, THEOBALD M, MORGAN D, HERNANDEZ J, BACIK I, YEWDELL
J, BENNINK J, AND BIGGS J. Strategies for tumor elimination by
cytotoxic T-lymphocytes. Crit Rev Immunol 18: 47–54, 1998.
147. SINGAL DP, YE M, AND QIU X. Molecular basis for lack of expression
of HLA class I antigens in human small-cell lung carcinoma cell
lines. Int J Cancer 68: 629 – 636, 1996.
148. SPIES T, BRESNAHAN M, BAHRAM S, ARNOLD D, BLANCK G, MELLINS E,
PIOUS D, AND DEMARS R. A gene in the human major histocompatibility complex class II region controlling the class I antigen presentation pathway. Nature 348: 744 –747, 1990.
149. SPIES T, CERUNDOLO V, COLONNA M, CRESSWELL P, TOWNSEND A, AND
DEMARS R. Presentation of viral antigen by MHC class I molecules
is dependent on a putative peptide transporter heterodimer. Nature
355: 644 – 646, 1992.
150. SPIES T AND DEMARS R. Restored expression of major histocompatibility class I molecules by gene transfer of a putative peptide
transporter. Nature 351: 323–324, 1991.
151. SPILIOTIS ET, OSORIO M, ZÚNIGA MC, AND EDIDIN M. Selective export
of MHC class I molecules from the ER after their dissociation from
TAP. Immunity 13: 841– 851, 2000.
152. TAMPÉ R, URLINGER S, PAWLITSCHKO K, AND UEBEL S. Unusual Secretory Pathways: From Bacteria to Man, edited by Kuchler K, Rubartelli A, and Holland B. New York: Springer, 1997, p. 115–136.
153. TECTOR M AND SALTER RD. Calnexin influences folding of human
class I histocompatibility proteins but not their assembly with beta
2-microglobulin. J Biol Chem 270: 19638 –19642, 1995.
154. TEISSERENC H, SCHMITT W, BLAKE N, DUNBAR R, GADOLA S, GROSS WL,
EXLEY A, AND CERUNDOLO V. A case of primary immunodeficiency
due to a defect of the major histocompatibbility gene complex class
203
204
175.
176.
177.
178.
179.
lation in high-grade primary breast carcinoma lesions. Cancer Res
58: 737–742, 1998.
VOS JC, REITS EAJ, WOJCIK-JACOBS E, AND NEEFJES J. Head-head/tailtail relative orientation of the pore-forming domains of the heterodimeric ABC transporter TAP. Curr Biol 10: 1–7, 2000.
VOS JC, SPEE P, MOMBURG F, AND NEEFJES J. Membrane topology and
dimerization of the two subunits of the transporter associated with
antigen processing reveal a three domain structure. J Immunol
163: 6679 – 6685, 1999.
WANG K, FRÜH K, PETERSON PA, AND YANG Y. Nucleotide binding of
the C-terminal domains of the major histocompatibility complexencoded transporter expressed in Drosophila melanogaster cells.
FEBS Lett 350: 337–341, 1994.
WIERTZ E, HILL A, TORTORELLA D, AND PLOEGH H. Cytomegaloviruses
use multiple mechanisms to elude the host immune response.
Immunol Lett 57: 213–216, 1997.
WRIGHT KL, WHITE LC, KELLY A, BECK S, TROWSDALE J, AND TING JPY.
Coordinate regulation of the human TAP1 and LMP2 genes from a
shared bidirectional promoter. J Exp Med 181: 1459 –1471, 1995.
YAN G, SHI L, AND FAUSTMAN D. Novel splicing of the human MHC-
Physiol Rev • VOL
181.
182.
183.
184.
185.
encoded peptide transporter confers unique properties. J Immunol
162: 852– 859, 1999.
YEWDELL JW, BENNINCK JR, AND OSAKA YH. Cells process exogenous
proteins for recognition by cytotoxic T-lymphocytes. Science 239:
637– 640, 1988.
YORK IA, GOLDBERG AL, MO XY, AND ROCK KL. Proteolysis and class
I major histocompatibility complex antigen presentation. Immunol
Rev 172: 49 – 66, 1999.
YORK IA, ROOP C, ANDREWS DW, RIDDELL SR, GRAHAM FL, AND
JOHNSON DC. A cytosolic herpes simplex virus protein inhibits
antigen presentation to CD8⫹ T-lymphocytes. Cell 77: 525–535,
1994.
ZEIDLER R, EISSNER G, MEISSNER P, UEBEL S, TAMPÉ R, LAZIS S, AND
HAMMERSCHMIDT W. Down-regulation of TAP1 in B lymphocytes by
cellular and Epstein-Barr virus-encoded interleukin-10. Blood 90:
2390 –2397, 1997.
ZHU K, WANG J, ZHU J, SHOU J, AND CHEN X. p53 induces TAP1 and
enhances the transport of MHC class I peptides. Oncogene 18:
7740 –7747, 1999.
82 • JANUARY 2002 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.1 on July 4, 2017
180.
BRIGITTE LANKAT-BUTTGEREIT AND ROBERT TAMPÉ