Molecular mechanisms in the TCR (TCR
ßCD3
,
) interaction with
2 homodimers: clues from a `phenotypic revertant' clone
Eric P. G. Martin,
Jacques Arnaud,
Laeticia Alibaud,
Cécile Gouaillard,
Régine Llobera,
Anne Huchenq-Champagne and
Bent Rubin
Unité de Physiopathologie Cellulaire et Moléculaire, CNRS, ERS 1590, IFR 30 d'Immunologie Cellulaire et Moléculaire, CHU de Purpan, 31059 cedex 03 Toulouse, France
Correspondence to:
B. Rubin
 |
Abstract
|
---|
The association between the TCR
ßCD3


hexamers and
2
homodimers in the endoplasmic reticulum (ER) constitutes a key step in TCR assembly and export to the T cell surface. Incompletely assembled TCRCD3 complexes are degraded in the ER or the lysosomes. A previously described Jurkat variant (J79) has a mutation at position 195 on the TCR C
domain causing a phenylalanine to valine exchange. This results in a lack of association between TCR
ßCD3


hexamers and
2 homodimers. Two main hypotheses could explain this phenomenon in J79 cells: TCRCD3 hexamers may be incapable of interacting with
2 due to a structural change in the TCR C
region; alternatively, TCRCD3 hexamers may be incapable of interacting with
2 due to factors unrelated to either molecular complex. In order to assess these two possibilities, the TCRCD3 membrane-negative J79 cells were treated with ethylmethylsulfonate and clones positive for TCR membrane expression were isolated. The characterization of the J79r58 phenotypic revertant cell line is the subject of this study. The main question was to assess the reason for the TCR re-expression. The TCR on J79r58 cells appears qualitatively and functionally equivalent to wild-type TCR complexes. Nucleotide sequence analysis confirmed the presence of the original mutation in the TCR C
region but failed to detect compensatory mutations in
, ß,
,
,
or
chains. Thus, mutated J79-TCRCD3 complexes can interact with
2 homodimers. Possible mechanisms for the unsuccessful TCRCD3 interaction with
2 homodimers are presented and discussed.
Keywords: endoplasmic reticulum, Jurkat cell, molecular chaperones, point mutation, TCR
 |
Introduction
|
---|
The TCR
ß is a structurally variable disulfide-linked heterodimer associated with invariant components of the CD3 molecule (
,
,
,
) on the T cell surface (1). While
ß dimers recognize peptide fragment antigens embedded in MHC molecules, the CD3 subunit transduces this recognition event into an intracellular signal, which activates the T cells (2,3). Membrane expression of TCRCD3 complexes is tightly regulated, with only completely assembled complexes being transported through the secretory pathway to the cell surface (4). Lack of, or mutations in, any one of the chains is sufficient to stop surface expression by retaining partially assembled or unassembled TCRCD3 components in the endoplasmic reticulum (ER) before targeting them for proteasomal or lysosomal degradation (510).
In thymocytes, normal T cells, T leukemia cells or hybridoma T cells, partial TCRCD3 complexes are formed following a precisely regulated scheme (1113). CD3
dimers associate with TCRß chains to form relatively stable molecules, whereas TCR
chains, which are rather labile, form stable complexes with CD3
dimers. Associations between TCR
and CD3
or TCRß and CD3
appear to take place via the transmembrane (TM) or the extracellular regions respectively (1416). The interaction between TCRßCD3
and TCR
CD3
complexes (TCRCD3 trimers) is followed by the creation of disulfide linkages between TCR
and ß chains.
The limiting step in TCR assembly seems to be the interaction between correctly assembled TCR
ßCD3
,
complexes (TCRCD3 hexamers) and
2 homodimers (17,18):
2 homodimers do not associate with single TCR or CD3 chains or with partial complexes such as TCR
CD3
or TCRßCD3
(11,18,
19). Thus, it would seem reasonable to assume that the interaction site of TCR
ßCD3 hexamers for
2 homodimers requires the interaction of several molecular components (20).
Several TCRCD3 membrane-negative variants of the human T cell leukemia line, Jurkat, have shown the importance of conserved amino acids in the Ig superfamily domain C1 (IgSF-C1) (16,18,19,21,22), in the connecting peptide (15) or in the TM region (13) of TCR C
or TCR Cß regions for the assembly of TCR
ßCD3
,
/
complexes. The 3P11 cells have the second of the TCR Cß intrachain disulfide-forming cysteines replaced by a tyrosine (19). In the J79 cells, a phenylalanine is substituted by a valine in position 195 of the TCR C
region (TCR
FV) (18). A phenylalanine to valine exchange in the equivalent position of the TCR Cß region has the same effect. In all cases, the precise phenotype is: (i) disulfide-linked TCR
ßCD3 hexamers are formed apparently normally, but they do not interact with
2 homodimers, (ii) the glycosylation of TCR
ßCD3 hexamers is incomplete and the hexamers are not transported to the cis-Golgi, and (iii) the TCR ß chain is unusually stable (16,1820). Thus, one interpretation of the data would be that the mutations in the TCR C
or TCR Cß regions decrease the avidity of the hexamer
2 homodimer interaction, i.e. the mutated amino acids are directly or indirectly (through steric interactions) implicated in the hexamer
2 interaction (20, 22).
The precise environment for the TCRCD3 +
2 interaction is not known (23). The formation of TCRCD3 trimers or hexamers is catalyzed or controlled by chaperones like BIP, calnexin, CD3
, glucose-regulated protein 94 or protein disulfide isomerase in the ER (2431). The specificity of interactions between chaperones and polypeptide chains is not completely defined: chaperones react with heterogeneous protein structures or certain glycoside structures on glycoproteins (29,32). If the TCRCD3 hexamer interaction with
2 homodimers is controlled by ER chaperones, catalytic enzymes or transporter molecules, mutations in such molecules could explain the defective phenotype of 3P11 or J79 cells: the amino acid exchanges might not influence the physical interaction between TCRCD3 complexes with mutated TCR chains (TCRMCD3) and
2 homodimers, but could play an important role in the interaction between TCRCD3 hexamers and a molecule(s) permitting the subsequent association with
2 homodimers. This role could involve control of the hexamer structure, the transport of hexamers from one intracellular compartment to another or facilitating hexamer interaction with
2 homodimers.
In order to distinguish between these two possibilities, we generated phenotypic revertant clones from J79 cells by treatment with ethylmethylsulfonate (EMS). The data in the present paper characterize one such TCRCD3 membrane-positive J79 phenotypic revertant cell line. The main question asked is whether the surface expression of TCRCD3 complexes is due to a compensatory mutation in one of the TCRCD3 chains or in a molecule implicated in the TCRCD3 hexamer interaction with the
2 homodimers.
 |
Methods
|
---|
Cells, culture and activation.
Jurkat T cells (E6.1) and the membrane TCRCD3-negative variant J79 (18,33) were cultured in 8% FCS/RPMI 1640glutamax medium supplemented with PSN (penicillin/streptomycin/neomycin) antibiotics (Life Technologies, Grand Island, NY) and 5x105 M 2-mercaptoethanol. Since the isolation of J79 cells in 1988, they have been maintained in culture for several years and neither cell cloning at limited dilution (9 times) nor cell-sorting has revealed TCRCD3 membrane-positive revertant cells. Treatment with EMS was performed as described previously (33
). J79 cells were transfected with a TCR V
4 C
cDNA (34) in the pMH-neo vector (35) as described previously (19). LYON cells express TCR
CD3 complexes at the surface membrane and they contain functionally competent, intracellular TCR ß chains (22).
E6.1, J79 and revertant cells were cultured in Albumax medium for 18 h in the presence of 1 ng/ml phorbol myristate acetate (PMA), or F101.01, OKT3, JOVI or anti-Vß8 mAb (100 ng/ml). After this incubation, the cells were washed and incubated with phycoerythrin (PE)-labeled anti-CD69 mAb. CD69 expression is given as percent CD69+ cells and mean fluorescence intensity (MFI). Internalization was measured by incubating the cells with F101.01, OKT3, JOVI or anti-Vß8 mAb for 18 h. After thorough washing, cells were incubated with PEUCHT1 anti-CD3
mAb and analyzed for percent CD3+ cells and MFI (13,16). For mAb, see below.
mAb, flow cytometry and cell sorting
PE-labeled UCHT1 anti-CD3
mAb were obtained from Dakopatts (Glostrup, Denmark). The
F1 (anti-TCR C
) and ßF1 (anti-TCR Cß) mAb were purchased from T Cell Science (Cambridge, MA); mAb against CD3
(HMT-3.2), CD3
(APA-1/2), CD3
(SP34),
(H146.968) and TCR Cß1 (JOVI) were obtained as culture supernatants from hybridomas provided by Drs Alarcon, Terhorst, Kubo and Owen respectively. OKT3 (anti-CD3
) hybridoma cells were obtained from ATCC (Rockville, MD). F101.01 (anti-TCRCD3 mAb reacting against a conformational epitope on CD3
and CD3
dimers in association with TCR chains) producing hybridoma cells (36) were kindly provided by Dr T. Plesner (Copenhagen, Denmark). PE-labeled anti-CD69 mAb and anti-Vß8 mAb were purchased from Immunotech (Marseille, France).
Cells were washed twice in ice-cold Dulbecco's phosphate buffered saline/0.3% BSA, and then incubated with PE-labeled mAb or first mAb for 25 min at 4°C followed by washing and an eventual second incubation with FITC-labeled F(ab')2 fragments of anti-mouse Ig (Dakopatts). For cell sorting, the mAb was dialyzed and sterile-filtered before incubation with the cells. Between 106 and 5x106 cells were sorted whenever necessary, and membrane CD3+ cells were grown in bulk cultures and cloned at limited dilution. Analyses and cell sorting were carried out on a Coulter Elite cytofluorometer equipped for cell sorting.
As described previously for membrane TCRCD3-negative Jurkat variants (37), the phenotypic variants studied in the present paper were analyzed for surface expression of CD2, CD5, CD7, CD11A, CD28, CD45, CD54 or CD58 molecules using reagents from Becton Dickinson (Mountain View, CA): all 16 revertant clones had normal expression of these membrane markers.
In certain experiments, the cells were treated with saponin buffer in order to render them permeable for mAb and FITC-labeled secondary antibodies. This was particularly important in the quantitative measurements of
chain content of the different Jurkat clones.
Biosynthetic and biotin labeling, immunoprecipitation, and SDSPAGE
Metabolic labeling with [35S]methionine/cysteine of Jurkat T cells, solubilization (1% digitonin or 2% NP-40), immunoprecipitation with anti-TCRCD3 mAb precoated onto Protein ASepharose, one-dimensional SDSPAGE as well as deglycosylation with endoglycosidase H were performed as described previously (13,15,16). All immunoprecipitation experiments are performed under non-reducing conditions. Pulsechase experiments were carried out as described before and the following mAb were used specifically for immunoprecipitation experiments: anti-calnexin (ref. 804-014-R100) from Alexis (San Diego, CA) or (AF8) (38) kindly provided by Dr M. Brenner (Boston, MA). Biotin labeling was performed according to the protocol described by Drs F. Lenfant (39) and T. Saito (40).
cDNA sequence analysis
Total RNA was prepared from 107 Jurkat cells by means of the guanidium isothiocyanatecesium chloride method (41). The first-strand cDNA was synthesized with reverse transcriptase and oligo-dT primer. PCR amplification was done as described previously (13,15,16) with the primers described in Table 1
. Amplified products from at least two different PCR were analyzed. A total of eight cDNA sequences were performed for each of apparently allelically excluded genes [TCR
, TCRß or CD3
(47)]; 12 cDNA sequence determinations were carried out with each of the non-allelically excluded genes (CD3
, CD3
and
) (see Results).
 |
Results
|
---|
General design of the experiments
The treatment of Jurkat cells with EMS was adjusted to induce single nucleotide exchanges preferentially. Thus, in J79 cells the first nucleotide, thymidine (T) in the Phe195 codon, i.e. TTC, was changed to a guanosine (G), GTC = valine (18). This nucleotide exchange seems to be the only TCRCD3-related mutation, since TCRCD3 structure and function is reconstituted upon transfection of J79 cells with wild-type TCR
cDNA (20). Furthermore, transfection of TCR
FV cDNA into a TCR
mRNA-negative Jurkat variant or into LYON cells generated intracellular TCR
FV chains, which form TCR
FVßCD3 complexes but lead to no surface expression of TCR
ßCD3 complexes (18,22).
The basic idea underlying the experiments in the present article was that induction in J79 cells of a novel mutation in any genes encoding a molecule controlling TCRCD3
2 association, in terms of either inter- or intramolecular interactions, could possibly compensate for the original mutation. Therefore, the J79 Jurkat variant cells were treated with EMS and analyzed by cytofluometry for surface expression of TCRCD3 complexes. Eventual phenotypic revertant cells were cloned and analyzed with respect to TCRCD3 assembly and TCRCD3-mediated cell function. Finally, the compensatory mutation was chased by cDNA sequencing of TCR
, TCRß,
, CD3
, CD3
or CD3
chains.
J79 phenotypic revertants (J79r58)
J79 cells were treated with EMS and analyzed for emergence of TCRCD3 surface-expressing cells. As no TCRCD3high surface-positive cells were detected, we tried to sort EMS-treated J79 cells that had the same mean fluorescence as control E6.1 cells when analyzed with PEUCHT1 mAb (Fig. 1
). A few thousand cells were sorted, expanded and cloned. Sixteen clones were obtained that expressed TCRCD3 surface levels comparable to E6.1 cells (Figs 1 and 2
) after a few weeks in culture. Such cells could in principle be E6.1 contaminants, back mutations of J79 TCR
chains (V
F195) or compensatory mutations. In order to distinguish among these possibilities, genomic DNA was produced from the phenotypic revertant clones and digested with HincII [J79 cell TCR
gene DNA contains an additional HincII site (18)] or BamHI (control). Southern blot analysis with 32P-labeled TCR C
probe demonstrated that all 16 revertant clones possessed the original FV195 mutation, a finding that was confirmed by nucleotide sequence analysis (see below). Seven cloning experiments with non-mutagenized J79 cells carried out during the last 10 years have not produced phenotypic revertant cells. Thus, phenotypic J79 revertant clones with surface membrane TCRCD3 complexes were obtained and we subsequently attempted to identify putative compensatory mutation(s). It is unknown whether the 16 phenotypic revertant clones are all different or derive from a single precursor clone (see Fig. 2
legend). The phenotypic revertant clone 58 (J79r58) was chosen for further analysis.

View larger version (18K):
[in this window]
[in a new window]
|
Fig. 1. Selection of J79 phenotypic revertants by cell sorting and cloning at limiting dilution. The data show cytofluometrical analysis of TCR membrane-positive (mTCR+) Jurkat T cells (E6.1), of TCR membrane-negative (mTCR) variant J79 cells, of EMS-treated J79 cells, of sorted, EMS-treated J79 cells and subclones of sorted cells. In the cloning experiment, 48 clones were analyzed: 29 clones were mTCR+, three clones were weakly mTCR+ and 16 clones were mTCR. mTCR+ clones were tested three consecutive times during 2 months and 16 clones maintained the high TCRCD3 membrane expression. Cell sorting of the variants presented here was performed 1 week after the limiting dilution step.
|
|
TCRCD3 surface structure and signaling of J79r58 cells
J79r58 cells have been grown in culture for several months without losing their high TCRCD3 surface expression. The level of TCRCD3 surface expression was determined with the PEUCHT1 mAb directed against an epitope on the CD3
molecule. In order to determine whether there was any difference among anti-TCRCD3 mAb-defined epitopes on TCRCD3 complexes of E6.1 and J79r58 cells, cytofluorometric analyses were performed with all available mAb: anti-Vß8 (recognizing the TCR Vß domain on Jurkat cells), anti-Cß [JOVI mAb recognizing an epitope on Cß1 domains and not on Cß2 domains; this epitope is destroyed when the intrachain disulfide bond is absent (19)], F101.01 [defining epitopes on CD3
and CD3
heterodimers, only when these are associated with TCR
or ß chains (36)] and OKT3 (anti-CD3
) mAb. Both E6.1 and J79r58 cell TCRCD3 complexes expressed the different mAb-defined epitopes at similar levels (Fig. 3A
). None of the mAb reacted with J79 cells (not shown).

View larger version (40K):
[in this window]
[in a new window]
|
Fig. 3. The TCRCD3 complex is structurally and functionally identical on both E6.1 and J79r58 cells. (A) E6.1 (grey bars) or J79r58 (open bars) cells were incubated with 10 ng/ml, 100 ng/ml, 1 µg/ml or 10 µg/ml of the individual mAb (OKT3 and W6.32 mAb only with 1 µg), washed and then incubated with saturating amounts of FITC-labeled anti-Ig Fab2 fragments. It can be seen that E6.1 and J79r58 cells express very similar amounts of the tested epitopes (W6.32 mAb is directed against a conserved epitope of MHC class I molecules). (B) E6.1 (grey bars) or J79r58 (open bars) cells were incubated with 10 ng/ml, 100 ng/ml, 1 µg/ml or 10 µg/ml of the indicated mAb at 4°C (OKT3 and W6.32 mAb only at 1 µg/ml), then washed and incubated with PEUCHT1 mAb (600 ng/ml). The data represent the MFI obtained with the PEUCHT1 mAb without competitor (0) or with competitive mAb. MFI without competitor was given the value 100% and the other experimental MFI values are expressed as percentage of this 100% value. It can be seen that only OKT3 and F101.01 mAb inhibit the interaction between TCRCD3 and PEUCHT1 mAb on E6.1 cells. The results obtained were very similar for E6.1 and J79r58 cells, indicating that the geometry of the tested epitopes is similar on the two cells. (C) E6.1 (grey bars), J79 (black bars) or J79r58 (open bars) cells were stimulated with PMA, PHA or different amounts of mAb for 24 h at 37°C as described in Methods. The cells were subsequently washed and analyzed with PE-labeled anti-CD69 mAb. The data represent the MFI obtained with the PE-labeled anti-CD69 mAb. The value obtained after PMA activation was defined as 100% and the other MFI values are expressed as 100% of this 100% value. The results indicate that TCRCD3-mediated T cell activation is absent in J79 cells, in contrast to both E6.1 and J79r58 cells. PMA-induced T cell activation is similar in all three Jurkat T cell lines. It should be noted that TCRCD3-mediated T cell activation induced by PHA is dependent on membrane expression of the TCRCD3 complex. (D) E6.1 (grey bars) and J79r58 (open bars) cells were incubated with PMA or different concentrations of PHA or F101.01 mAb for 24 h at 37°C. The cells were then washed and incubated with PEUCHT1 mAb. The data represent the MFI obtained with the PEUCHT1 mAb on Jurkat cells incubated in culture medium alone (defined as 100%). The experimental MFI values are expressed as percentage of this 100% value. An aliquot of the cells was incubated with FITC-labeled anti-Ig Fab2 fragments; no labeling was observed, i.e. the lack of reaction of the UCHT1 mAb was not due to competitive blockage by the mAb used for inducing the internalization. It can be seen that both the recycling internalization pathway (induced by PMA) and the anti-TCRCD3 mAb-induced internalization followed by the degradation pathway (47) is very similar in the two cell types.
|
|
Next, we asked whether the geometry between these epitopes was the same on TCRCD3 complexes from E6.1 or J79r58 cells. The cells were preincubated with `cold' anti-Vß8, anti-Cß, F101.01 or OKT3 mAb, washed and then PEUCHT1 mAb was added. The capacity of the four non-labeled mAb to inhibit TCRCD3 labeling by PEUCHT1 mAb was determined. The data in Fig. 3
(B) demonstrate that for both cell clones only OKT3 and F101.01 mAb inhibited the interaction between TCRCD3 complexes and PEUCHT1 mAb. We next compared the ability of E6.1, J79 or J79r58 cells to transmit signals to the interior of the cell. In the first analysis, the three cell lines were induced to express the CD69 early activation molecule on the plasma membrane. PMA stimulates T cells independently of TCRCD3 surface expression and PHA stimulates T cells with TCRCD3 surface complexes, as do anti-Vß8, JOVI, F101.01 or OKT3 mAb. The data in Fig. 3
(C) demonstrate that PMA induced CD69 expression in all three cell lines. In contrast, PHA and the mAb induced CD69 expression in a similar way in E6.1 and J79r58 cells, but not in J79 cells (Fig. 3C
).
Finally, TCRCD3 complex internalization induced by anti-TCR reagents was investigated. As can be seen in Fig. 3
(C), F101.01 mAb, PHA and PMA induce TCRCD3 internalization to a similar extent in both E6.1 and J79r58 cells. Within the limits of the presented experiments, our results indicate that TCRCD3 complexes on the surface membrane of E6.1 and J79r58 cells are similar regarding both tertiary structure and function.
Assembly of TCRCD3 complexes in J79r58 cells
J79r58 cells express a TCRCD3 complex that contains the J79-mutated TCR
chain. The main question is whether the TCRCD3/
assembly and stoichiometry are the same in J79r58 cells as in E6.1 cells or whether partial TCRCD3/
complexes are formed in J79r58 cells (45
). Metabolic labeling of TCRCD3 chains in J79 cells has shown that TCRCD3 complexes do not associate with
chains. Moreover, TCR
ß heterodimers do not mature, indicating ER retention of incompletely processed TCR chains (they do not proceed beyond the high mannose form sensitive to endo-H glycosidase) and TCR ß chains are much more stable in J79 cells compared to E6.1 cells (20). Thus, E6.1, J79 and J79r58 cells were cultured for 90 min in the presence of 35S-labeled methionine/cysteine and soluble TCRCD3 proteins from lysed cells were immunoprecipitated with mAb against the six polypeptide chains. The results in Fig. 4
demonstrate that TCR
, ß and CD3
,
,
or
chains reacted normally with their respective mAb. However,
chains were co-precipitated with anti-TCRCD3 mAb only in E6.1 or J79r58 cells, but not in J79 cells (note that the autoradiogram of J79r58 cells was obtained after 3 days of exposure, whereas the autoradiograms of E6.1 or J79 cells were obtained after 21 days). In addition, it can be seen that the high mol. wt J79-TCR
ß heterodimers are not present in J79r58 cells. Furthermore, a pulsechase experiment shows normal maturation of TCR
ß heterodimers in J79r58 cells (Fig. 5
; note that the exposure time for J79r58 cell immunoprecipitates was 4 days, and for E6.1 or J79 cell immunoprecipitates, the exposure time was 15 days). The results in Figs 4 and 5
suggest that the TCRCD3/
associations occur normally in J79r58 cells. In order to determine whether the TCRCD3/
interactions in J79r58 cells take place quantitatively the same way as in E6.1 cells, [35S]methionine/cysteine-labeled TCRCD3 proteins from E6.1, J79 and J79r58 cells were immunoprecipitated with different quantities of mAb directed against CD3
, CD3
, CD3
chains or CD3
CD3
complexes. There appears to be no quantitative differences in TCRCD3 assembly between E6.1 and J79r58 cells (not shown). Thus, the immunoprecipitation experiments did not elucidate whether the compensatory mutation has taken place in TCR or CD3 chains.

View larger version (77K):
[in this window]
[in a new window]
|
Fig. 5. Pulsechase immunoprecipitation experiments with E6.1, J79 and J79r58 cells. Jurkat T cells were labeled for 30 min with [35S]methionine/cysteine, washed, and aliquots of cells were lysed in 1% digitonin after 0, 30 min, 1 h, 2 h and 4 h of culture in normal medium (tracks 1, 2, 3, 4 and 5 respectively). Immunoprecipitations were performed with 5 µl Protein ASepharose beads saturated with OKT3 mAb. Raji B lymphoma cells labeled metabolically for 30 min and subsequently lysed in 1% digitonin served as negative control (non-reducing conditions). Whereas both E6.1 and J79r58TCR ß heterodimers mature during the chase period of 4 h, the J79TCR ß heterodimers remain incompletely glycosylated and not associated with 2 homodimers.
|
|
Nucleotide sequence analysis of TCRCD3/
chains in J79r58 cells
Karyotype analysis (Kuhlmann, unpublished) and analysis of cell cycling by cytometry have shown that Jurkat E6.1, J79 and J79r58 cells possess 46 chromosomes (2n) as compared to peripheral blood lymphocytes (also 2n) or to LYON TCR
(4n) cells (data not shown). At least two different RT-PCR were performed for each of the individual TCRCD3 chains from total RNA of E6.1, J79 and J79r58 cells. Each resulting cDNA was cloned in the pCRII (Invitrogen) vector and prepared for nucleotide sequence analysis. The TCR
and ß chains undergo allelic exclusion: previous studies have shown that Jurkat cell RNA only hybridizes with V
1 (out of 24 TCR families) and Vß8 (out of 21 TCR families) cDNA probes (46). Moreover, Jurkat cells have been described to only express one allele of the CD3
chain (47). Therefore, eight cDNA sequence determinations (four sequences for each PCR product) seemed reasonable to ascertain the nucleotide sequence of TCR
, TCR ß and CD3
chains. When two alleles are expressed (in the case of CD3
,
and
chains), 12 sequences were analyzed (six for each PCR product) giving a probability of 12(0.5)12 > 99% to have analyzed two equally expressed alleles.
Nucleotide sequence analysis of V
, C
, Vß and Cß demonstrated that there were no compensatory mutations in these genes; however, the original phenylalanine
valine mutation was found in all C
sequences. The same result was obtained from sequence analysis of CD3
, CD3
, CD3
and
cDNA in either J79 or J79r58 cells (not shown). These data suggest that the compensatory mutation has not occurred in the individual TCRCD3 chains but possibly in an as yet unidentified molecule involved in the TCRCD3
2 homodimer interaction. Thus, TCRCD3 complexes with a phenylalanine
valine mutated TCR
chain can interact with
2 homodimers and form a functional TCR
Chaperone function in J79r58 cells
ER chaperones control proper folding and egress of newly synthesized membrane proteins (29,32,43). Three chaperones have been reported to transiently interact with TCRCD3 subunits in T cells: BIP (25), CD3
(24,27) and calnexin (26,28,30,31). As the compensatory mutation in J79r58 cells may have occured in a chaperone-like molecule, we attempted to compare chaperone behavior in J79r58 cells with that of J79 or E6.1 cells. BIP has been implicated in the assembly of TCR
chains with CD3
heterodimers (25). Since this interaction appears to take place normally in J79 cells, we did not pursue BIP as a possible target for the compensatory mutation.
It has been shown that the CD3
interaction with CD3 complexes appears similar in E6.1 and J79 cells as judged by co-precipitation experiments achieved with anti-CD3
mAb (33). Several studies indicate that CD3
interacts primarily with CD3
heterodimers (12,24,27). Therefore, J79r58 cells (with E6.1 and J79 cells as controls) were labeled with [35S]methionine/cysteine and chased for 30 or 90 min. Individual lysates were immunoprecipitated with anti-CD3
(not shown) or anti-CD3
mAb (Fig. 6
). OKT3 anti-CD3
mAb (and less strongly APA-1/2 anti-CD3
mAb) co-precipitate CD3
molecules labeled with [35S]methionine/cysteine for 20 min; the association of CD3
with CD3
molecules or CD3
-containing complexes was stable for ~30 min (Fig. 6
). Most importantly, these results were very reproducible. One study has indicated that CD3
can physically associate with TCRCD3 hexamers and actually be exchanged with
2 homodimers (27). Our results from experiments with normal or PMA-stimulated Jurkat cells solubilized in digitonin, NP-40 or Triton-X100 were unable to demonstrate any association between TCRCD3 hexamers and CD3
. However, CD3
is associated to a similar extent to CD3
heterodimers or CD3
homodimers in E6.1, J79 or J79r58 cells (in preparation).
Calnexin is essential for TCRCD3 trimer formation (26,28,30,31), i.e. early in the TCRCD3 assembly pathway. Immunoprecipitation of biosynthetically labeled lysates from E6.1, J79 and J79r58 cells with two different anti-calnexin mAb, with
F1 anti-TCR C
mAb and with OKT3 anti-CD3
mAb, showed little if any co-precipitation of TCR components with anti-calnexin mAb; and when present, similar results were obtained with lysates from all three cells (not shown). The conclusion from these data is that in J79 or J79r58 cells nothing seems abnormal in the initial assembly interactions leading to formation of TCRCD3 trimers or hexamers. Actually, in all three cells, disulfide-linked TCRCD3 hexamers are normally formed (Figs 4 and 5
); the major difference being that in J79 cells the hexamers remain in the ER in an incompletely glycosylated form, whereas J79r58 hexamers associate with
2 homodimers and are transported to the plasma membrane.
It has been shown previously that
2 homodimers can be expressed on the T cell surface membrane in an autonomous way; in addition, it has been suggested that the
2 molecules may serve as transporters of complete TCRCD3 complexes (in addition to their function in signal transduction) (40). Quantitative cytofluometric analyses using anti-
mAb demonstrated that the concentration of intracellular
chains was very similar in the three cells (Fig. 2C
). Furthermore, the data in Fig. 7
show clearly that all three Jurkat T cell lines express similar amounts of TCRß and
2 molecules. In order to ascertain whether TCRCD3 hexamer
2 homodimer interaction takes place in a similar way in E6.1 or J79r58 cells, these cells were surface biotinylated and a pulsechase experiment was carried out (Fig. 8
). It can be seen that in all three cell lines,
2 homodimers are well labeled at time zero and the biotinylated
2 molecules disappear from the cell surface following 2 h of in vitro culture. In addition, relabeling with biotin after 2 h of culture shows the same quantity of
2 homodimers in E6.1, J79 and J79r58 cells. Thus, the transport function of
2 molecules seems to work normally in the three Jurkat cell lines.
 |
Discussion
|
---|
Newly synthesized membrane proteins fold rapidly upon entry into the ER, where they assemble with other polypeptides in order to acquire a correct quaternary structure. The ER provides an oxidizing environment that favors formation of disulfide bonds. In addition, a number of molecules that facilitate the folding and subunit assembly are present in the ER. Such molecules include soluble chaperones like BIP, protein disulfide isomerase and glucose-regulated protein 94, and chaperones that are integral components of the ER membrane (e.g. calnexin) (32). Chaperones may have at least three functions: (i) quality control over protein folding, e.g. retention of proteins that are incorrectly folded or delayed in acquiring their native structure, which might prolong their retention in the ER, (ii) a provision of increased energy for proteinprotein interactions, i.e. a kind of enzyme activity, and (iii) a transporter function, i.e. increasing the concentration of interacting components in a particular subcompartment of the ER (4,25,27,2931). Calnexin and calreticulin possess a lectin-like domain, that has specificity for oligosaccharide moieties characteristic of incompletely folded glycoproteins (26,30,31,32,43). Other chaperones, while interacting with non-assembled proteins, dissociate rapidly as new polypeptides assemble to the complex (BIP and CD3
) (24,25). In TCRCD3 assembly, BIP, calnexin, calreticulin, CD3
and many others appear to play important roles in TCRCD3 trimer and hexamer formation (30,31,32,38,43,4850).
The data presented in this paper bear on the mechanism of assembly and membrane expression in phenotypic revertant cells derived from a TCRCD3 membrane-negative variant J79 (20). This TCRCD3 surface membrane-negative Jurkat T cell variant has a phenylalanine
valine exchange in position 195 of the external TCR C
region; this defect causes production of partially glycosylated TCR
ßCD3 hexamers that do not interact with
2 homodimers (20). Many experiments have indicated that the Phe195 of TCR C
and the equivalent phenylalanine in TCR Cß regions are involved in a molecular interaction site for
2 homodimers (18,20,21). However, the crystal structure of the TCR
ß heterodimers (51,52,53) show that the phenylalanine in question is situated in a hydrophobic pocket facing the TCR V region rather than the plasma membrane. Since
chains have only nine external amino acids, these observations made the hypothesis of a mutated interaction site for
2 on the C
/Cß region unlikely. It appears now clear that TCR
FVßCD3 hexamers can interact with
2 homodimers. Moreover, those mutant TCR
FVß CD3/
2 complexes on the J79r58 cell surface have apparently the same structure, epitope expression or geometry and signaling function capacities as TCRCD3 complexes on wild-type Jurkat T cells. Most importantly, the compensatory mutation permitting TCRMCD3 membrane expression on J79r58 cells did not occur in any of the TCRCD3 chains:
, ß,
,
,
or
. Thus, with our present state of knowledge, it is likely that the compensatory mutation has occurred in a chaperone-like molecule(s), which normally controls the TCRCD3 hexamer interaction with
2 homodimers. The J79 phenylalanine
valine mutation causes lack of or inefficient interaction between TCRMCD3 hexamers and the hypothesized chaperone, rather than between TCRMCD3 hexamers and
2 homodimers.
Both calnexin, calreticulin and CD3
have been implicated in prolonged interactions with TCRCD3 complexes. Thus, calnexin seems to interact transiently with all TCRCD3 components except
chains (12,26,28,30,31), whereas calreticulin reacts transiently only with TCR
or TCRß chains (30,31,32). Our analysis of calnexin association with TCRCD3 components in E6.1, J79 or J79r58 cells showed no significant differences; in contrast, calnexin is indeed strongly associated with TCR ß chains in the TCR
chain-negative JR3.11 variant cells (16). These data indicate that calnexin is not involved in the defective TCRMCD3
2 interaction in J79 cells.
Most data available suggest that CD3
plays a role early in the TCRCD3 assembly pathway, in particular in CD3 heterodimer and TCRCD3 trimer formation (12,24,27,49,50). However, one study suggests that CD3
may stay associated with TCRCD3 complexes until the interaction with
2 homodimers; in addition, it has been suggested that the TCRCD3
2 interaction may dissociate CD3
from the TCRCD3 complexes (27). Therefore, we analyzed possible differences in the interaction between CD3
and TCRCD3 components among E6.1, J79 or J79r58 cells. The results of these experiments carried out in different detergents showed no evidence of differential CD3
interactions with TCRCD3 components from E6.1, J79 or J79r58 cells. The suggested CD3
displacement by
2 homodimers could not be demonstrated despite several attempts. On the contrary, CD3
seems to be mostly involved in the control of CD3
CD3
and TCR
CD3
interactions like BIP (25). As these interactions seem normal in J79 cells, it could be expected that CD3
is not a possible candidate for the mutated chaperone in J79r58 cells. Nevertheless, the above described experiments were carried out due to the data published by Neisig et al. (27) and due to the observation that in J79r58 cell lysates, anti-CD3
mAb co-precipitates more CD3
compared to wild-type Jurkat T cells (not shown).
The
chain is an important molecule for TCRCD3 membrane expression, for signal transduction and for thymocyte differentiation (2,17,54,55). The amount of intracellular
chains seems to be an essential control point for TCRCD3 complex activities. Intracellular cytometry analysis on saponin-treated E6.1, J79 or J79r58 cells, as well as quantitative immunoprecipitation studies, showed similar amounts of
chains in all three cell lines. The concentration of
chains appears to be regulated by two mechanisms: limited
chain biosynthesis (17,54) and
chain recycling from the surface membrane (40,56). Thus, whereas TCRCD3 complexes seem stable at the T cell surface,
chains have a half-life of <2 h at the surface membrane. However, dissociation of biotinylated
chains from TCRCD3 complexes on E6.1 or J79r58 cells was similar (Fig. 8
). Therefore, TCRCD3 hexamers in J79r58 cells appear to assemble and to be transported by
2 homodimers to the cell surface in the same way as in E6.1 cells. In addition, it seems likely that
2 homodimers recycle normally in J79 cells compared to E6.1 or J79r58 cells. This means that
2 homodimers are expressed autonomously at the T cell surface even without TCRCD3 complexes (40).
The conclusion from the experimental data with J79r58 cells is (i) TCR
FVßCD3 complexes can interact physically with
2 homodimers, (ii) the quantity and recycling function of
chains is normal, and (iii) TCR
FVßCD3/
complexes on the cell surface are functional. The J79-TCRCD3 function can be reconstituted by transfection with wild-type TCR
cDNA, and
chain turnover and function is normal in J79 cells.
Thus, without changing the cDNA sequences of any of the
, ß,
,
,
or
transcripts in J79r58 cells, these cells express functional TCRCD3 complexes at the cell surface membrane. It would appear logical then to suggest that TCR
FVßCD3 complexes and
2 homodimers can interact physically in J79 cells but they never get a chance to do so, because the two partners are present in two different intracellular compartments. Preliminary confocal and electron microscopy experiments indicate that TCRCD3 hexamers and
chains are separate in J79 cells; in contrast, TCRCD3 hexamers and
chains overlap in E6.1 cells, J79r58 cells and J79 cells transfected with TCR
cDNA (in preparation). Therefore, we suggest that the compensatory mutation in J79r58 cells has happened in an intracellular molecule that (i) controls the tertiary structure of TCRCD3 complexes before interaction with
chains, (ii) catalyzes the TCRCD3
2 homodimer interaction or (iii) transports TCRCD3 complexes to the
2 recycling compartment.
In summary, we have described a phenotypic revertant T cell clone J79r58, which has a compensatory mutation in a molecule behaving as an intracellular chaperone. This new chaperone-like function, which catalyzes the TCRCD3
2 homodimer interaction, could not be attributed to calnexin, to CD3
or to
2 homodimer recycling. We are now attempting to clone the mutated chaperone cDNA by transfection of an expression cDNA library from J79r58 cells into J79 variant Jurkat T cells.
 |
Acknowledgments
|
---|
The collaboration of Drs F. Lenfant, M. Brenner and J. Kuhlmann during parts of this work is gratefully acknowledged. Many thanks are sent to Dr W. R. Clark (Los Angeles) for comments to and for reading the manuscript. We want to express our gratitude for help with cell sorting from H. Brun and G. Cassar. Finally, we thank our colleagues Drs B. Alarcon, M. Brenner, R. Kubo, M. Owen, T. Plesner and C. Terhorst for providing mAb or mAb-producing hybridomas. The present work was supported by the CNRS, l'Université Paul Sabatier, l'ARC (no. 6223), la Ligue Régionale contre le Cancer (Toulouse) and Immunotech (Marseille).
 |
Abbreviations
|
---|
C | constant region |
EMS | ethylmethylsulfonate |
ER | endoplasmic reticulum |
FV | phenylalanine valine exchange |
MFI | mean fluorescence intensity |
PBL | peripheral blood lymphocyte |
PE | phycoerythrin |
PHA | phytohemagglutinin A |
PMA | phorbol myristate acetate |
TM | transmembrane |
TCRM | TCR heterodimer with either TCR or ß chains mutated in the TCR C -Phe195 equivalent position |
V | variable region |
 |
Notes
|
---|
Transmitting editor: M. Reth
Received 4 November 1998,
accepted 9 March 1999.
 |
References
|
---|
-
Clevers, H., Alarcon, B., Wileman, T. and Terhorst, C. 1988. The T cell receptorCD3 complex: a dynamic protein ensemble. Annu. Rev. Immunol. 6:629.[ISI][Medline]
-
Wegener, A. M., Letourneur, F., Hoeveler, A., Brocker, T., Luton, F. and Malissen, B. 1992. The T cell receptorCD3 complex is composed of at least two autonomous transduction modules. Cell68:83.[ISI][Medline]
-
Weiss, A. and Littman, D. R. 1994. Signal transduction by lymphocyte antigen receptors. Cell 76:263.[ISI][Medline]
-
Klausner, R. D., Lippincott-Schwartz, J. and Bonifacino, J. S. 1990. The T cell antigen receptor: insights into organelle biology. Annu. Rev. Cell Biol. 6:403.[ISI]
-
Alarcon, B., Berkhout, B., Breitmeyer, J. and Terhorst, C. 1988. Assembly of the human T cell receptorCD3 complex takes place in the endoplasmic reticulum and involves intermediary complexes between the CD3-gamma.delta.epsilon core and single T cell receptor alpha or beta chains. J. Biol. Chem. 263:2953.[Abstract/Free Full Text]
-
Lippincott-Schwartz, J., Bonifacino, J. S., Yuan, L. C. and Klausner, R. D. 1988. Degradation from the endoplasmic reticulum: disposing of newly synthesized proteins. Cell 54:209.[ISI][Medline]
-
Mallabiabarrena, A., Fresno, M. and Alarcon, B. 1992. An endoplasmic reticulum retention signal in the CD3 epsilon chain of the T-cell receptor. Nature 357:593.[ISI][Medline]
-
Mallabiabarrena, A., Jimenez, M. A., Rico, M. and Alarcon, B. 1995. A tyrosine-containing motif mediates ER retention of CD3-epsilon and adopts a helix-turn structure. EMBO J. 14:2257.[Abstract]
-
Wileman, T., Carson, G. R., Shih, F. F., Concino, M. F. and Terhorst, C. 1990. The transmembrane anchor of the T-cell antigen receptor beta chain contains a structural determinant of pre-Golgi proteolysis. Cell Reg. 1:907.[ISI][Medline]
-
Wileman, T., Kanet, L., Carson, G. and Terhorst, C. 1991. Depletion of cellular calcium accelerates protein degradation in the endoplasmic reticulum. J. Biol. Chem. 266:4500.[Abstract/Free Full Text]
-
Manolios, N., Letourneur, F., Bonifacino, J. S. and Klausner, R. D. 1991. Pairwise, cooperative and inhibitory interactions describe the assembly and probable structure of the T-cell antigen receptor. EMBO J. 10:1643.[Abstract]
-
Kearse, K. P., Roberts, J. L. and Singer, A. 1995. TCR alphaCD3 delta epsilon association is the initial step in alpha beta dimer formation in murine T cells and is limiting in immature CD4+ CD8+ thymocytes. Immunity 2:391.[ISI][Medline]
-
Arnaud, J., Chenu, C., Huchenq, A., Gouaillard, C., Kuhlmann, J. and Rubin, B. 1996. Defective interactions between TCR chains and CD3 heterodimers prevent membrane expression of TCR-alpha beta in human T cells. J. Immunol. 156:2155[Abstract]
-
Manolios, N., Kemp, O. and Li, Z. G. 1994. The T cell antigen receptor alpha and beta chains interact via distinct regions with CD3 chains. Eur. J. Immunol. 24:84.[ISI][Medline]
-
Arnaud, J., Cayrou, C., Llobera, R. and Rubin, B. 1997. Mutation in splicing consensus sequences causes lack of TCR membrane expression due to exon excision. Immunogenetics 45:311.[ISI][Medline]
-
Arnaud, J., Huchenq, A., Vernhes, M. C., Casparbauguil, S., Lenfant, F., Sancho, J., Terhorst, C. and Rubin, B. 1997. The interchain disulfide bond between TCR-
ß heterodimers on human T cells is not required for TCRCD3 membrane expression and signal transduction. Int. Immunol. 9:615.[Abstract]
-
Sussman, J. J., Bonifacino, J. S., Lippincott-Schwartz, J., Weissman, A. M., Saito, T., Klausner, R. D. and Ashwell, J. D. 1988. Failure to synthesize the T cell CD3-zeta chain: structure and function of a partial T cell receptor complex. Cell 52:85.[ISI][Medline]
-
Geisler, C., Rubin, B., Caspar-Bauguil, S., Champagne, E., Vangsted, A., Hou, X. and Gajhede, M. 1992. Structural mutations of C-domains in members of the Ig superfamily. Consequences for the interactions between the T cell antigen receptor and the zeta 2 homodimer. J. Immunol. 148:3469.[Abstract/Free Full Text]
-
Caspar-Bauguil, S., Arnaud, J. Gouaillard, C. Hou, X. Geisler, C. and Rubin, B. 1994. Functionally important amino acids in the TCR revealed by immunoselection of membrane TCR-negative T cells. J. Immunol. 152:5288.[Abstract/Free Full Text]
-
Rubin, B., Arnaud, J., Caspar-Bauguil, S., Conte, F. and Huchenq, A. 1994. Biological function of the extracellular domain of the T-cell receptor constant region. Scand. . J Immunol. 39:517.[ISI][Medline]
-
Caspar-Bauguil, S., Arnaud, J., Huchenq, A., Hein, W. R., Geisler, C. and Rubin, B. 1994. A highly conserved phenylalanine in the alpha, beta-T cell receptor (TCR) constant region determines the integrity of TCRCD3 complexes. Scand. J. Immunol. 40:323.[ISI][Medline]
-
Kuhlmann, J., Caspar-Bauguil, S., Geisler, C. and Rubin, B. 1993. Characterization of T cell receptor assembly and expression in a Ti gamma delta-positive cell line. Eur. J. Immunol. 23:487.[ISI][Medline]
-
Lippincott-Schwartz, J., Yuan, L. C., Bonifacino, J. S. and Klausner, R. D. 1989. Rapid redistribution of Golgi proteins into the ER in cells treated with brefeldin A: evidence for membrane cycling from Golgi to ER. Cell 56:801.[ISI][Medline]
-
Pettey, C. L., Alarcon, B., Malin, R., Weinberg, K. and Terhorst, C. 1987. T3-p28 is a protein associated with the delta and epsilon chains of the T cell receptorT3 antigen complex during biosynthesis. J. Biol. Chem. 262:4854.[Abstract/Free Full Text]
-
Suzuki, C. K., Bonifacino, J. S., Lin, A. Y., Davis, M. M. and Klausner, R. D. 1991. Regulating the retention of T-cell receptor alpha chain variants within the endoplasmic reticulum: Ca(2+)-dependent association with BiP. J. Cell Biol.
114:189.[Abstract]
-
David, V., Hochstenbach, F., Rajagopalan, S. and Brenner, M. B. 1993. Interaction with newly synthesized and retained proteins in the endoplasmic reticulum suggests a chaperone function for human integral membrane protein IP90 (calnexin). J. Biol. Chem. 268:9585.[Abstract/Free Full Text]
-
Neisig, A., Vangsted, A., Zeuthen, J. and Geisler, C. 1993. Assembly of the T-cell antigen receptor. Participation of the CD3 omega chain. J. Immunol. 151:870.[Abstract/Free Full Text]
-
Rajagopalan, S., Xu, Y. and Brenner, M. B. 1994. Retention of unassembled components of integral membrane proteins by calnexin. Science 263:387.[ISI][Medline]
-
Melnick, J. and Argon, Y. 1995
. Molecular chaperones and the biosynthesis of antigen receptors. Immunol. Today 16:243.[ISI][Medline]
-
van Leeuwen, J. E. and Kearse, K. P. 1996. Calnexin associates exclusively with individual CD3 delta and T cell antigen receptor (TCR) alpha proteins containing incompletely trimmed glycans that are not assembled into multisubunit TCR complexes. J. Biol. Chem. 271:9660.[Abstract/Free Full Text]
-
van Leeuwen, J. E. M. and Kearse, K. P. 1996. The related molecular chaperones calnexin and calreticulin differentially associate with nascent T cell antigen receptor proteins within the endoplasmic reticulum. J. Biol. Chem. 271:25345.[Abstract/Free Full Text]
-
Helenius, A., Trombetta, E. S., Hebert, D. N. and Simons, J. F. 1997. Calnexin, calreticulin and the folding of glycoproteins. Trends Cell Biol. 7:193.[ISI]
-
Geisler, C., Kuhlmann, J. and Rubin, B. 1989. Assembly, intracellular processing, and expression at the cell surface of the human alpha beta T cell receptorCD3 complex. Function of the CD3-zeta chain. J. Immunol. 143:4069.[Abstract/Free Full Text]
-
Triebel, F., Breathnach, R., Graziani, M., Hercend, T. and Debre, P. 1988. Evidence for expression of two distinct T cell receptor beta-chain transcripts in a human diphtheria toxoid-specific T cell clone. J. Immunol. 140:300.[Abstract/Free Full Text]
-
Hahn, W. C., Menzin, E., Saito, T., Germain, R. N. and Bierer, B. E. 1993. The complete sequences of plasmids pFNeo and pMH-Neo: convenient expression vectors for high-level expression of eukaryotic genes in hematopoietic cell lines. Gene 127:267.[ISI][Medline]
-
Dietrich, J., Neisig, A., Hou, X., Wegener, A. M., Gajhede, M. and Geisler, C. 1996. Role of CD3 gamma in T cell receptor assembly. J. Cell Biol. 132:299.[Abstract]
-
Wegener, A. M., Holm, B., Geisler, C. and Rubin, B. 1990. Cellular and molecular characteristics of transformed T cells from an antigen-specific T-cell line. Scand. J. Immunol. 31:645.[ISI][Medline]
-
Hochstenbach, F., David, V., Watkins, S. and Brenner, M. B. 1992. Endoplasmic reticulum resident protein of 90 kilodaltons associates with the T- and B-cell antigen receptors and major histocompatibility complex antigens during their assembly. Proc. Natl Acad. Sci. USA 89:4734.[Abstract]
-
Rodriguez, A. M., Mallet, V., Lenfant, F., Arnaud, J., Girr, M., Urlinger, S., Bensussan, A. and Lebouteiller, P. 1997. Interferon-gamma rescues Hla Class Ia cell surface expression in term villous trophoblast cells by inducing synthesis of Tap proteins. Eur. J. Immunol. 27:45.[ISI][Medline]
-
Ono, S., Ohno, H. and Saito, T. 1995. Rapid turnover of the CD3 zeta chain independent of the TCRCD3 complex in normal T cells. Immunity 2:639
.[ISI][Medline]
-
Chirgwin, J. M., Przybyla, A. E., MacDonald, R. J. and Rutter, W. J. 1979. Isolation of biologically active ribonucleic acid from sources enriched in ribonuclease. Biochemistry 18:5294.[ISI][Medline]
-
Huchenq, A., Alibaud, L., Bouchouata, C., Gouaillard, C., Llobera, R., Alcover, A., Arnaud, J. and Rubin, B. 1999. Genetic and biochemical analysis of the role of CD3
in human
ß T cell receptor assembly and ligand-induced internalisation. Eur. J. Immunol., submitted.
-
Bergeron, J. J., Brenner, M. B., Thomas, D. Y. and Williams, D. B. 1994. Calnexin: a membrane-bound chaperone of the endoplasmic reticulum. Trends Biochem. Sci, 19:124.[ISI][Medline]
-
Yang, M., Omura, S., Bonifacino, J. and Weissman, A. 1998. Novel aspects of degradation of T cell receptor subunits from the endoplasmic reticulum (ER) in T cells: importance of oligosaccharide processing, ubiquitination, and proteasome-dependent removal from ER membranes. J. Exp. Med. 187:835.[Abstract/Free Full Text]
-
Alarcon, B., Ley, S. C., Sanchez-Madrid, F., Blumberg, R. S., Ju, S. T., Fresno, M. and Terhorst, C. 1991. The CD3-gamma and CD3-delta subunits of the T cell antigen receptor can be expressed within distinct functional TCRCD3 complexes. EMBO J. 10:903.[Abstract]
-
Champagne, E., Huchenq, A., Sevin, J., Casteran N. and Rubin, B. 1993. Analysis of the T cell receptor Vß gene repertoire: clustering of Vß subfamilies responsive to staphylococcal enterotoxins B and E. Mol. Immunol. 30: 877.[ISI][Medline]
-
Niedergang, F., Sanjose, E., Rubin, B., Alarcon, B., Dautry-Varsat, A. and Alcover, A. 1997. Differential cytosolic tail dependence and intracellular fate of T-cell receptors internalized upon activation with superantigen or phorbol ester. Res. Immunol. 148:231.[ISI][Medline]
-
Kearse, K. P., Roberts, J. L., Munitz, T. I., Wiest, D. L., Nakayama, T. and Singer, A. 1994. Developmental regulation of alpha beta T cell antigen receptor expression results from differential stability of nascent TCR alpha proteins within the endoplasmic reticulum of immature and mature T cells. EMBO J. 13:4504.[Abstract]
-
Stewart, S. J., and Levy, R. 1989. Components of the human T lymphocyte antigen receptor complex, CD3-p28 and CD3-gamma, are biochemically distinct. J. Lab. Clin. Med. 114:394.[ISI][Medline]
-
Antusch, D., Bonifacino, J. S., Burgess, W. H. and Klausner, R. D. 1990. The T cell receptor-associated protein is proteolytically cleaved in a pre-Golgi compartment. J. Immunol. 145:885.[Abstract/Free Full Text]
-
Garboczi, D. N., Ghosh, P., Utz, U., Fan, Q. R., Biddison, W. E. and Wiley, D. C. 1996. Structure of the complex between human T-cell receptor, viral peptide and HLA-A2. Nature 384:134.[ISI][Medline]
-
Garcia, K. C., Degano, M., Stanfield, R. L., Brunmark, A., Jackson, M. R., Peterson, P. A., Teyton, L. and Wilson, I. A. 1996. An alphabeta T cell receptor structure at 2.5 A and its orientation in the TCRMHC complex. Science 274:209.[Abstract/Free Full Text]
-
Housset, D., Mazza, G., Gregoire, C., Piras, C., Malissen, B. and Fontecillacamps, J. C. 1997. The three-dimensional structure of a T-cell antigen receptor V-alphaV-beta heterodimer reveals a novel arrangement of the V-beta domain. EMBO J. 16:4205.[Abstract/Free Full Text]
-
Weissman, A. M., Frank, S. J., Orloff, D. G., Mercep, M., Ashwell, J. D. and Klausner, R. D. 1989. Role of the zeta chain in the expression of the T cell antigen receptor: genetic reconstitution studies. EMBO J. 8:3651.[Abstract]
-
Shores, E. W., Huang, K., Tran, T., Lee, E., Grinberg, A. and Love, P. E. 1994. Role of TCR zeta chain in T cell development and selection. Science 266:1047.[ISI][Medline]
-
Aoe, T., Goto, S., Ohno, H. and Saito, T. 1994
. Different cytoplasmic structure of the CD3
family dimer modulates the activation signal and function of T cells. Int. Immunol. 6:1671.[Abstract]