[Previous]

1. Résumé en français


1.1 Les molécules du complexe majeur d'histocompatibilité classe II

      Les molécules du complexe majeur d'histocompatibilité de classe II (CMH-II) (1), aussi appelées antigènes leucocytaires humains (HLA), et antigènes Ia ou H2 chez la souris, sont codées par des gènes regroupés dans une seule région sur le chromosome 6 chez l'homme (Fig.3, p.20) et sur le chromosome 17 chez la souris. Il y a trois différents isotypes de molécules CMH-II chez l'homme (HLA-DR, HLA-DP et HLA-DQ) (2-5), et deux isotypes chez la souris (I-A et I-E) (6, 7). Les molécules CMH-II jouent un rôle essentiel dans les réponses immunitaires spécifiques chez les vertébrés. Elles sont exprimées à la surface des cellules présentatrices d'antigènes pour présenter des peptides antigéniques aux lymphocytes T auxiliaires CD4+ (8, 9). Si un antigène est reconnu par le récepteur T d'un lymphocyte T, ce lymphocyte est activé et déclenche une réponse immunitaire spécifique, notamment en sécrétant des cytokines qui stimuleront la production d'anticorps par les cellules B et activeront les cellules T cytotoxiques CD8+.

      Les molécules CMH-II sont constituées par l'association non covalente de deux chaînes glycoprotéiques trans-membranaires, a et b (Fig.1, p.18; Fig.2, p.19). Les dimères ab s'assemblent dans le réticulum endoplasmique et sont stabilisés par la chaîne invariante (10) (Fig.4, p.22). La chaîne invariante a deux fonctions: elle facilite le transport du réticulum endoplasmique à travers l'appareil de Golgi vers l'endosome et elle empêche la liaison précoce des peptides antigéniques aux molécules CMH-II. Dans l'endosome(11), la chaîne invariante est clivée par des protéases (12, 13). Un peptide de la chaine invariante, CLIP, reste attaché au site de liaison des molécules CMH-II. Catalysé par les molécules HLA-DM, CLIP est ensuite dissocié et remplacé par des peptides antigéniques, qui sont issus de la protéolyse des antigènes exogènes (14-16). Les complexes CMH-II sont finalement transporté à la surface cellulaire.


1.2 L'expression des gènes CMH-II

      L'expression des gènes CMH-II est strictement régulée et est restreinte dans la plupart des situations physiologiques à quelques types cellulaires hautement spécialisés. Ceux-ci comprennent les cellules présentatrices d'antigènes, plus précisément les cellules dendritiques, les cellules B, les macrophages et les cellules T activées chez l'homme (17-21). De plus, un compartiment thymic spécialisé dans la sélection positive des lymphocytes T exprime les molécules CMH-II (22, 23). La plupart des autres types cellulaires n'expriment pas les molécules CMH-II de manière constitutive mais ces molécules peuvent être induites par l'interféron-g (IFN-g) (17-21).

      Les expressions constitutive et inductible des gènes CMH-II sont principalement régulées au niveau transcriptionnel par des séquences régulatrices communes situées à proximité du site d'initiation de la transcription (18-21, 24-27). Elles sont composées de quatre éléments, la boîte W, la boîte X1, la boîte X2 et la boîte Y. La séquence, l'orientation, l'ordre et l'espacement de ces éléments sont conservées dans tous les gènes CMH-II ainsi qu'entre différentes espèces (18-21, 24-27). La plus part des facteurs se liant à ces éléments ont pu être identifiés (Fig.6a, p.33). Le complexe RFX, composé de RFX5, RFXAP et RFXANK, se lie à la boîte X1 (28-33). La boîte X2 est contactée par le facteur X2BP (34, 35). Finalement, le facteur NF-Y, composé des trois sous-unités NF-YA, NF-YB et NF-YC se lie à la boîte Y (36). L'ensemble des trois facteurs RFX, X2BP et NF-Y fixé à l'ADN forme une plateforme qui sert de site de recrutement d'un coactivateur essentiel nommé CIITA (37, 38).

      CIITA et les trois sous-unités de RFX ont été isolés grâce à l'existence de patients atteints par le syndrome de déficience en molécules CMH-II (18, 25-27, 39-43). Cette maladie, appelée aussi BLS pour "Bare Lymphocyte Syndrome" est caractérisé par une absence d'expression des molécules CMH-II à la surface des cellules présentatrices d'antigène et des autres types cellulaires après stimulation par l'IFN-g. Ceci entraîne une immunodéficience humorale et cellulaire sévère et abouti à la mort des patients au cours de l'enfance. L'absence globale d'expression des molécules CMH-II est la conséquence d'un défaut de transcription des gènes CMH-II. Elle concerne tous les gènes a et b codants pour HLA-DR, HLA-DP et HLA-DQ. Des expériences de fusion cellulaire ont permis de classer les patients dans quatre groupes de complémentation génétique distincts. Un gène différent est muté dans chaque groupe: MHC2TA dans le groupe A (38), RFXANK dans le groupe B (29, 30), RFX5 dans le groupe C (32) et RFXAP dans le groupe D (28) (Fig.5, p.28).

      RFX5 est la plus grande sous-unité du complexe RFX (32) (Fig.7C, p.35). Elle appartient à la famille RFX qui regroupe des protéines possédant toutes un domaine de liaison à l'ADN typique et spécifique de cette famille (44, 45). La structure du domaine de liaison à l'ADN à été définie pour un des membres de la famille RFX (RFX1) (46). Elle fait partie de la sous-famille "winged helix" qui elle fait partie de la famille hélice-boucle-hélice. La région N-terminale de la protéine RFX5 est impliquée dans l'interaction avec RFXANK et RFXAP (47, 48). En revanche, la partie C-terminale interagit avec CIITA et NF-Y (48).

      RFXANK porte son nom parce qu'il contient un domaine d'interaction entre protéines composé de quatre répétés ankyrine (29, 30) (Fig.7B, p.35). Ce domaine interagit avec CIITA, RFX5 et RFXAP. En plus, RFXANK touche l'ADN dans la partie 3' de la boîte X1.

      RFXAP contient deux domaines d'activation potentiels, l'un riche en acides aminés acides et l'autre riche en résidus glutamine (28) (Fig.7D, p.35). Seul le domaine riche en résidus glutamine est essentiel pour la fonction de la protéine (47, 49).

      La liaison du complexe RFX à la boîte X1 est fortement stabilisée par l'interaction synergique avec les facteurs X2BP et NF-Y. L'ensemble ces trois facteurs RFX, X2BP et NF-Y liés à l'ADN forment l'« enhanceosome » (37). Ce complexe est extrêmement stable et occupe les promoteurs des gènes CMH-II avec une grande spécificité (50). L'absence du facteur RFX conduit à la perte de l'occupation du promoteur comme peut être observé dans les cellules de groupes de complémentation B à D (51-53) (Fig.6C, p.33). L'« enhanceosome » est essentiel mais pas suffisant pour l'expression des gènes CMH-II (Fig.6B, p.33). C'est le coactivateur CIITA qui est indispensable pour l'expression de manière constitutive et inductible (38, 54-60). Il détermine le niveau et la spécificité de l'expression des gènes CMH-II (18, 61). En effet, CIITA est exprimé uniquement dans les cellules exprimant les molécules CMH-II. La transfection des cellules qui n'expriment pas les molécules CMH-II avec du CIITA suffit pour activer l'expression des molécules CMH-II (55, 56, 62, 63). CIITA régule également l'expression des molécules CMH-II au cours de la différenciation des cellules B (56, 64). Dans les plasmocytes, le stade finale de différenciation des lymphocytes B, l'extinction des gènes CMH-II est la conséquence directe de la répression du gène MHC2TA. Enfin, CIITA régule aussi l'expression des molécules CMH-II de manière quantitative (65).

      CIITA contient deux domaines riches en acides aminés acides et riches en résidus proline/sérine/thréonine dans la partie N-terminale, un domaine de liaison au GTP situé au centre de la protéine et des répétés riche en leucine situés dans la partie C-terminale (18, 61) (Fig.7A, p.35). Le contact avec l'« enhanceosome » se fait avec une grande partie de la région C-terminale par de multiples contacts protéine-protéine (37, 66). Cette région C-terminale peut fonctionner comme mutant dominant négatif (67-71). CIITA active la transcription par le biais du domaine d'activation dans sa partie N-terminale (68, 72, 73). Ce domaine contacte plusieurs facteurs généraux de la transcription (TAFII32, TAFII70, TFIIB), des facteurs impliqués dans l'élongation de la transcription (TFIIH, pTEFb) et des facteurs qui modifient la chromatine (CBP, pCAF) (69, 74-77) (Fig.10, p.49). Pour son activité, CIITA s'associe avec soi-même (78-80).

      Le contrôle de l'expression des gènes CMH-II par CIITA se fait principalement au niveau de la transcription du gène MHC2TA (81). MHC2TA et son homologue Mhc2ta chez la souris sont localisés tous deux sur le chromosome 16. L'expression de MHC2TA est contrôlée par quatre promoteurs différents (82) (Fig.8, p.41). Les promoteurs pI, pIII et pIV sont conservés entre l'homme et la souris, le promoteur pII n'existe que chez l'homme et n'a pas été étudié en détail. Chaque promoteur précède un premier exon différent. Ce premier exon est épissé de manière alternative au deuxième exon commun. Ceci produit trois types de transcrits (type I, type III et type IV), chaqu'un possédant une extrémité différente en 5' (82). Une protéine de 121 kD est traduite à partir d'un site d'initiation localisé dans le deuxième exon commun à tous les transcrits. Les transcrits de type I et III contiennent un site d'initiation de plus dans leur premier exon. L'utilisation de ces sites d'initiation alternatifs mène à des isoformes de 132 et 124 kD. Les extensions N-terminales uniques pour ces isoformes confèrent possiblement un aspect spécifique à ces protéines. L'isoforme de 132 kD par exemple semble avoir une demi-vie inférieure aux autres isoformes.

      Les différents types de CIITA sont exprimés dans différents types cellulaires (82). C'est le contrôle strict de la transcription du gène MHC2TA qui dicte l'expression restreinte et inductible des gènes CMH-II. Le promoteur pI de CIITA est exprimé spécifiquement dans les cellules dendritiques et dans les macrophages de souris induits par l'IFN-g (-). La régulation de ce promoteur n'a pas encore été étudiée. Le promoteur pIII est utilisé dans les cellules B, dans les cellules T humaines activées et dans certains types de cellules dendritiques (82, 83, 86, 87). Une région promotrice de 320 paires de bases à été trouvée suffisante pour l'expression spécifique (82, 86) (Fig.9B, p.44). Il contient des éléments liés par des facteurs de transcription inconnus. Le pIII peut être activé par l'IFN-g dans certains types cellulaires par le biais d'un élément régulateur situé cinq kilobases en amont du site d'initiation de la transcription (88). L'activité du pIII est éteinte au cours de la différenciation terminale des cellules B (56, 64). Le facteur PRDI-BF1 a été proposé d'agir comme répresseur du pIII dans les plasmocytes (89, 90). Le promoteur pIV est inductible par l'IFN-g dans la plus part des types cellulaires (82, 88, 91-94). Un fragment promoteur de 300 paires de bases est suffisant pour la réponse à l'IFN-g (82, 93, 94) (Fig.9C, p.44). Il contient un élément GAS, une boîte E et un site de liaison pour le facteur IRF-1. L'élément GAS et la boîte E sont liés par les facteurs STAT-1 et USF-1 de manière coopérative. L'expression du facteur IRF-1 dépend elle-même de l'induction par l'IFN-g. L'expression de CIITA induite par l'IFN-g peut être supprimée par plusieurs cytokines, notamment le TGF-b, l'IL-1, l'IL-4 et l'IL-10 (91, 95-97). Dans les souris auxquelles il manque le pIV CIITA les gènes CMH-II ne peuvent plus être induits par l'IFN-g sauf par les cellules présentatrices d'antigène telles que la microglie ou les macrophages (85). C'est le pI qui est activé par l'IFN-g dans ces cellules par un mécanisme inconnu (85). De plus, les cellules épithéliales du cortex thymique perdent aussi l'expression du CMH-II ce qui produit un déficit grave dans la sélection positive des lymphocytes T CD4+ (85). Le pIV est alors indispensable pour la réponse à l'IFN-g dans toutes les cellules d'origine non-hematopoiëtiques.


1.3 Les cellules dendritiques

      Les cellules dendritiques représentent des cellules présentatrices d'antigène uniques parce qu'elles sont capables d'activer des cellules T naïves (98-100). Elles existent en deux états de différentiation: immature et mature (Fig.11, p.66). En état immature, les cellules dendritiques sont localisées dans les tissus périphériques à l'interface avec l'environnement où elles absorbent des antigènes par micropinocytose, phagocytose et endocytose (101, 102). Les cellules dendritiques immatures expriment peu de molécules CMH-II à leur surface, mais elles les accumulent dans leur compartiment endosomale (103, 104). Des antigènes pathogénique et des cytokines inflammatoires sont capables d'activer les cellules dendritiques et d'induire leur programme de maturation (98). Ceci implique des changements majeurs dans la morphologie et la fonction des cellules. Les cellules dendritiques induites à la maturation dégradent les antigènes qu'elles ont endocytosés auparavant et elles les chargent sur leurs molécules CMH-II qui sont stabilisées par cela (103, 104). Elles migrent dans les organes lymphatiques secondaires (105, 106) pour présenter les peptides antigéniques dans le contexte des molécules CMH-II aux cellules T CD4+. Les cellules T CD4+ activées par l'engagement du recepteur T déclenchent une réponse immunitaire en activant des cellules T cytotoxiques et les cellules B (98).

      Les cellules dendritiques sont capables de présenter des antigènes endogènes et exogènes. Les antigènes endogènes sont dégradés par le protéasome et transférés par le transporteur TAP dans le réticulum endoplasmique (Fig.12, p.69). Dans ce compartiment ils sont chargés sur les molécules CMH classe I (CMH-I). Finalement, les complexes CMH-I-peptide sont transportés à la surface cellulaire. Pendant la maturation, le protéasome est modifié de telle manière qu'il favorise la génération de peptides qui s'adaptent bien au site de liaison des molécules CMH-I (107-109). Les antigènes exogènes sont dégradés dans le compartiment lysosomale et présentées par les molécules CMH-II (Fig.13, p.70). La formation des complexes CMH-II-peptide ne se fait qu'après l'induction de la maturation des cellules dendritiques (103, 110-113). Dans les cellules dendritiques immatures, les molécules CMH-II sont retenues dans le compartiment endosomale tardif et elles sont associées avec la chaîne invariante. La chaîne invariante est dégradée après l'induction de la maturation (110). Les peptides antigéniques peuvent s'attacher alors au site de liaison des molécules CMH-II et les complexe CMH-II-peptide sont transportés à la surface cellulaire. Les antigènes exogènes peuvent aussi être présentés par les molécules CMH-I et activer des cellules T CD8+ par le biais de la cross-présentation (114-116) (Fig.14, p.71). La cross-présentation permet l'initiation de réponses immunitaires cytotoxiques contre des antigènes qui ne sont pas synthétisés par les cellules dendritiques elles-mêmes. Cela implique donc le transfert d'antigène de cellules infectées aux cellules dendritiques.


1.4 L'expression du gène MHC2TA dans les cellules dendritiques

      CMH-II est une molécule clé pour le fonctionnement des cellules dendritiques. Son expression et sa distribution cellulaire sont liées strictement à l'état de maturation dans ces cellules. Plusieurs études ont analysé le transport intracellulaire et le chargement des molécules CMH-II dans les cellules dendritiques (103, 111, 112, 117). Cependant, les mécanismes qui contrôlent les changements dans la biosynthèse des molécules CMH-II durant la maturation des cellules dendritiques sont mal connus. Cet aspect a donc été étudié pendant cette thèse. Les résultats obtenus ont fait objet d'un article qui a été publié dans le « Journal of Experimental Medecine » (83).

      En utilisant des cellules dendritiques obtenues par différenciation de monocytes humains du sang péripherique, nous avons observé que le niveau des ARN messagers codant pour les protéines CMH-II diminue pendant la maturation induite par le lipopolysaccharide (LPS), jusqu'à 23% de celui qui est observé dans les cellules dendritiques immatures (Fig.1C dans la publication, p.88[382]). Une évolution inverse a été observé pour les taux de protéines CMH-II présentées à la surface cellulaire durant cette même maturation par le LPS (Fig.1A et B dans la publication, p.88[382]).

      Pour comprendre l'origine de la diminution d'ARN messager CMH-II, nous avons analysé le coactivateur CIITA qui est essentiel pour la transcription des gènes CMH-II. Les trois isoformes de CIITA ont été détecté dans les cellules dendritiques immatures par immuno-précipitation (Fig.2C dans la publication, p.88[382]). Le taux de CIITA est réduit de huit fois pendant la maturation, et seule l'isoforme de 121 kD reste détectable dans les cellules matures.

      En ce qui concerne les ARN messagers de CIITA, nous avons observé que les cellules dendritiques immatures expriment des grandes quantités des transcrits de type I et de type III (Fig.3 dans la publication, p.xx). Le messager de type IV n'est guère exprimé dans ces cellules. Peu après l'induction de la maturation, le niveau des transcrits de type I et de type III est fortement réduit (Fig.3 dans la publication, p.88[384]). Un niveau basal à 5% du niveau de départ pour le type I et à 20% pour le type III est atteint après seulement 4 heures de stimulation avec le LPS. La réduction de l'expression des ARN messagers CIITA est considérablement plus rapide que celle des ARN messagers CMH-II. Cette différence de cinétique et au moins partiellement due à la différence de demi-vie des deux ARN messagers (plus de 24 heures pour CMH-II et ~1 heure pour CIITA). Les donnés sur l'extinction de CIITA pendant la maturation des cellules dendritiques humaines peuvent être étendues sur d'autres types de cellules dendritiques et sur d'autres espèces. Les cellules dendritiques dérivées de la moëlle osseuse de souris expriment l'ARN messager CIITA de type I et plus faiblement de type III et IV (Fig.4B dans la publication, p.88[385]). Avec l'induction de la maturation avec du LPS, cette expression est éteinte.

      Afin de déterminer si l'extinction de CIITA observée dans les cellules dendritiques est spécifique à la stimulation par le LPS ou représente une conséquence de la maturation en général, nous avons mesuré les taux d'ARN messager CIITA dans les cellules dendritiques humaines après exposition au TNF-a, au CD40 ligand, à l'IFN-a, et après infection avec la souche bactérienne Salmonella typhimurium ou le virus de Sendai (Fig.5A dans la publication, p.88[385]). Tous ces stimuli ont induits une réduction profonde de CIITA malgré des cinétiques différentes. L'expression de CIITA a aussi été étudiée in vivo pendant le cours d'une maladie inflammatoire (Fig.4C et D dans la publication, p.88[385]). CIITA est réduite dans la rate de souris atteintes par l'encéphalite autoimmune expérimentale (EAE) aiguë comparé à des souris contrôles. Cette réduction est plus marquée quand l'analyse est faite avec des cellules dendritiques isolées à partir de la rate.

      Afin de déterminer si la réduction de l'ARN messager CIITA est due à une déstabilisation de l'ARN messager, nous avons comparé la stabilité des transcrits CIITA dans les cellules dendritiques immatures et matures. La demi-vie de l'ARN messager est très courte (~1 heure) dans les deux cas et aucune différence entre les deux n'a pu être mis en évidence (Fig.6A dans la publication, p.88[386]). Ces donnés suggèrent qu'un arrêt de la transcription est responsable de la diminution de l'ARN messager CIITA pendant la maturation des cellules dendritiques. Pour confirmer cette hypothèse nous avons mesuré la vitesse de transcription en quantifiant les transcrits naissants liés à la chromatine (Fig.6B dans la publication, p.88[386]). Nous avons détecté facilement les transcrits naissants de CIITA de type I et de type III dans les cellules dendritiques immatures. Dans les cellules dendritiques matures par contre, la synthèse de ces transcrits naissants est éteinte. Comme pour les taux d'ARN messager, la réduction est plus forte pour le type I que pour le type III de CIITA. En conclusion, ces résultats indiquent que la réduction de l'expression de CIITA est contrôlée par l'extinction de la transcription du gène MHC2TA.

      Des changements dans l'activité transcriptionelle sont souvent dus à des altérations d'occupation des séquences régulatrices ce qui peut être détecté par des expériences de « in vivo footprint ». Nous avons donc comparé l'occupation des promoteurs pI et pIII entre les cellules dendritiques immatures et matures (Fig.7 dans la publication, p.88[387]). La région proximale du pI et caractérisée par des protections sur les deux brins de l'ADN dans les cellules immatures. La maturation ne s'accompagne pas d'une modification de ces protections. La région proximale du pIII est également occupée dans les cellules immatures et montre un profile similaire à celui observé dans les cellules B. Comme pour le pI, il n'y a aucun changement dans l'occupation du promoteur pendant la maturation.

      Trois arguments indiquent que les promoteurs pI et pIII du gène MHC2TA sont contrôlés par un mécanisme de répression globale, peut-être au niveau de la chromatine, pendant la maturation des cellules dendritiques. 1) pI et pIII sont espacés de plus que 12 kilobases. 2) Ils ne contiennent pas d'élément de régulation commun. 3) Les taux d'occupation du pI et pIII ne changent pas malgré l'extinction de ces promoteurs pendant la maturation. Au niveau de la chromatine, la répression de la transcription est souvent associée à la déacétylation des résidus lysine dans les queues des histones. Afin de savoir si ce mécanisme joue un rôle dans la répression du gène MHC2TA dans les cellules dendritiques, nous avons induit leur maturation avec du LPS en présence de TSA, un inhibiteur de déacétylases. TSA a effectivement empêché l'extinction de CIITA induit par le LPS (Fig.8A dans la publication, p.88[388]). Cette donnée suggère que la répression de CIITA est effectuée par un mécanisme qui mène à la déacétylation des histones associés à la région régulatrice du gène MHC2TA. Nous avons pu prouver cette hypothèse avec des expériences d'immuno-précipitation de la chromatine, utilisant des anticorps contre les histones H3 et H4 acétylés (Fig.8B dans la publication, p.88[388]). La région régulatrice du gène MHCIITA, qui s'étale sur plus que 12 kilobases, est acétylée entièrement dans les cellules dendritiques immatures. Après maturation par contre, cette acétylation est perdue. Le mécanisme qui à conduit l'extinction du gène MHC2TA implique donc la déacétylation d'une grande région régulatrice.


1.5 L'analyse moléculaire du promoteur I de CIITA

      Les cellules dendritiques immatures expriment CIITA de type I et type III à des niveaux élevés. Les mécanismes qui contrôlent cette expression dans ces cellules est toujours inconnu. Nous avons donc étudié la régulation des promoteurs pI et pIII en détail.

      Les cellules dendritiques n'étant pas faciles à transfecter par des méthodes classiques, nous avons choisi un système de transduction utilisant des vecteurs lentiviraux. Ce système permet d'intégrer des transgènes de manière stable dans le génome des cellules qui ne se divisent pas. Les cellules dendritiques peuvent être transduites avec une efficacité de plus que 50% avec le vecteur pWPTS qui code pour la GFP sous le contrôle du promoteur EF-1a (Fig.18A, p.94). Ce système de transduction n'engendre pas la maturation des cellules dendritiques (Fig.18B, p.94). Pour définir les régions fontionelles de pI, nous avons construit une série de délétions progressives en 5' du promoteur, fusionées au gène GFP dans le vecteur pWPTS. Le fragment promoteur contenant 390 pairs de bases (pI-390) est suffisant pour une activité transcriptionelle (Fig.19, p.95). Les fragments plus longs par contre sont très faiblement ou pas du tout actifs. pI-390 est un promoteur faible comparé à EF-1a (Fig.20, p.96). Son activité est aussi légèrement en dessous de l'activité du promoteur minimal de pIII, pIII-322. Par contre, le niveau d'expression de pI-390 est clairement plus élevé que le niveau basal d'un vecteur qui ne contient aucun promoteur (p0). pI-390 a été testé pour sa spécificité cellulaire. Il est exprimé uniquement dans les cellules dendritiques et des lignées promonocytiques U937 et THP1 (Fig.21, p.98). pI-390 n'est donc pas strictement restreint aux cellules dendritiques mais semble être spécifique pour les cellules d'origine myéloide. En plus, le promoteur pIII-322 montre une activitée transcriptionelle élevée dans toutes les cellules testées, pas seulement dans les cellules B et dendritiques. Ces résultats sont inattendus et différent de ceux obtenus avec un système de tranfection transitoire (82). Aucune diminution de l'activité transcriptionelle de pI-390 et pIII-322 n'a pu être observée pendant la maturation des cellules dendritiques (Fig.22, p.99), malgré la forte répression des promoteurs pI et pIII endogènes. Les éléments régulateurs impliqués dans cette répression sont probablement localisés en dehors des séquences promotrices minimales et loin des sites d'initiation de la transcription. Cette hypothèse est en accord avec le mécanisme global qui est résponsable de l'extinction de CIITA pendant la maturation des cellules dendritiques.


1.6 Conclusions

      Les cellules dendritiques immatures synthétisent des molécules CMH-II en grande quantité, mais ce n'est qu'après la maturation que ces molécules sont chargées et transportées à la surface. L'expression de surface élevée soutient les cellules dendritiques matures dans leur rôle de présentation des peptides antigéniques aux cellules T CD4+. A l'inverse de l'augmentation des molécules CMH-II à la surface, la biosynthèse de CMH-II est réduite pendant la maturation. Cet arrêt de synthèse garantit que les complexes CMH-II-peptide ne sont pas dilués et remplacés par de nouveaux complexes produits pendant la maturation. Ceci est cohérent avec le fait que les cellules matures perdent leur capacité d'intérnaliser des antigènes. La réduction de la synthèse des molécules CMH-II pendant la maturation est due à l'extinction du gène MHC2TA. Cette réduction est observée dans les cellules dendritiques de l'homme et de la souris en réponse à toute une série de stimulis, et même chez des souris souffrant d'une maladie inflammatoire. Les deux promoteurs pI et pIII sont réprimés de manière coordinée pendant la maturation des cellules dendritiques. pI et pIII sont distants de 12 kb et ils ne contiennent aucune séquence régulatrice en commun. Ces données suggèrent que l'extinction est médiée par un mécanisme régulateur global qui concerne tous les promoteurs du gène MHC2TA simultanément. En effet, ce mécanisme implique la déacétylation de la région promotrice entière du gène MHC2TA. C'est donc un remaniement de la chromatine et non pas une répression restreinte aux promoteurs individuels qui mène à l'extinction de MHC2TA pendant la maturation des cellules dendritiques.

      Le mécanisme de la régulation de pI n'a pas été étudié pendant de nombreuses années par manque d'une lignée cellulaire qui exprime ce promoteur et qui se comporte comme des cellules dendritiques primaires. Les cellules dendritiques primaires ne peuvent pas être transfectées par des méthodes classiques. C'est grace à un système de transduction utilisant des vecteurs lentiviraux que nous avons pu réaliser l'analyse de pI dans ces cellules. Un fragment promoteur de 390 pairs de bases est suffisant pour dirigé l'expression dans les cellules dendritiques. Son activité est restreinte aux cellules d'origine myéloide. En plus, l'état de maturation des cellules n'influence pas l'expression de pI-390. Ces charactéristiques rendent le promoteur pI très intéressant pour des stratégies d'immunothérapie ou de vaccination.

      En conclusion, ce travail de thèse a permis de mieux comprendre les mécanismes qui contrôlent l'expression de CMH-II dans les cellules dendritiques immatures et matures. En particulier, il éclaircit le rôle de CIITA, le régulateur clé des gènes CMH-II, pendant la maturation. En plus, la régulation du promoteur de CIITA qui est spécifique pour les cellules dendritiques a été examinée.


2. Introduction


2.1 MHCII

      Major histocompatibility complex (MHC) molecules, also called human leukocyte antigens (HLA) in human and Ia or H-2 antigens in mouse, were first detected as cell surface antigens that varied between different strains of mice and that were encoded by several genes responsible for the rejection of transplanted tissues (118). Later, the genes controlling immune responses in transplantation and specificity of antigen recognition by T cells were assigned to the same MHC loci (119-121). Since then, much has been learned about the structure, diversity and function of MHC molecules.


2.1.1 MHCI versus MHCII molecules

      There are two classes of MHC molecules, class I (MHCI) and class II (MHCII). MHCI and MHCII molecules are both transmembrane glycoproteins belonging to the immunoglobulin supergene family, and their structures are very similar. However, their polypeptide chain composition, patterns of expression and functions in the immune system are very different (Fig.1).

      MHCI molecules are composed of a transmembrane a chain associated non-covalently with the b2-microglobulin (b2m) chain. They are expressed on essentially all nucleated cell types. MHCI molecules are specialized for the presentation of peptides derived from endogenous proteins to the T cell receptor (TCR) of CD8+ T cells.

      MHCII molecules are composed of two non-covalently linked transmembrane chains, called a and b. In contrast to the ubiquitous expression of MHCI, MHCII molecules are expressed only on a subset of cells. They are specialized for the presentation of extracellular antigens to the TCR of CD4+ T cells.

      

Fig. 1 MHCI and MHCII molecules.

      A. Interaction of a peptide-loaded MHCI molecule on any cell type with the TCR of a CD8+ T cell. B. Interaction of a peptide-loaded MHCII molecule on an antigen presenting cell (APC) with the TCR of a CD4+ T cell.


2.1.2 The role of MHCII molecules in the immune response

      MHCII is one of the key molecules for the development of a specific immune response to a pathogen (8, 9). MHCII-peptide complexes displayed at the surface of antigen presenting cells (APCs) are recognized by the TCR and the CD4 co-receptor on CD4+ T cells. This triggers activation and proliferation of the T cells and thus elicits an immune response specific for the antigen from which the MHCII bound peptides were derived.

      MHCII molecules are also crucial for selection and maturation of CD4+ T cells in the thymus. Positive selection, which ensures the survival of T cells that carry TCRs capable of recognizing self-MHC molecules, is driven by MHCII positive epithelial cells in the thymic cortex (cTECs) (23). On the other hand, elimination of autoreactive T cells by negative selection is driven by MHCII positive thymic dendritic cells in the medulla (122, 123).


2.1.3 The structure of MHCII molecules

      MHCII molecules belong to the immunoglobulin gene family. They are heterodimeric glycoproteins consisting of two non-covalently linked transmembrane chains called a (33 kD) and b (29 kD). The difference in size of the two chains is mainly attributed to differences in N-linked glycosylation. The a and b chains have the same overall conformation, each consisting of two extracellular domains, a1 and a2, and b1 and b2, respectively. The membrane-distal domains combine to form a single peptide binding groove composed of two antiparallel a-helical loops supported by a platform of eight antiparallel b strands (124) (Fig.2). The groove is capable of binding a wide range of peptides. Peptides binding to MHCII molecules are at least 13 amino acids long and can be much longer. Their N- and C-termini may extend beyond the ends of the groove. The peptides are held in the peptide-binding groove both by peptide side chains (anchor positions) that protrude into polymorphic pockets lined by residues that vary between MHCII molecules, and by interactions between the peptide backbone and the side chains of residues that are conserved in all MHCII peptide-binding grooves.

      

Fig. 2 Three-dimensional structure of MHCII.

      A. X-ray structure of HLA-DR1 with the a chain in blue and the b chain in red. B. Complex between HLA-DR3 and the CLIP peptide (in red). CLIP and antigenic peptides bind in an almost identical manner (125).


2.1.4 The diversity of MHCII genes and their genomic organization

      Three classical MHCII molecules exist in man: HLA-DP (3), HLA-DQ (4) and HLA-DR (2, 5). Mice only express proteins orthologous to the last two, I-A (6) and I-E (7), respectively. In addition, both species encode so-called nonclassical molecules, namely HLA-DM (126) and HLA-DO (127, 128) in human, and H2-M (129) and H2-O (130) in mouse. The a and b chains of each MHCII molecule are encoded by separate genes in the class II region of the MHC (1) (Fig.3). The human MHC is located on the short arm of chromosome 6, while the mouse MHC is situated on chromosome 17. In all cases, except for HLA-DO, the pairs of genes are encoded adjacently. Remarkably, the class II region of the MHC is full of pseudogenes. These may be involved in generating new alleles by gene conversion. The class II region appears to have been subjected to several duplications generating novel gene family members, which have then diverged into new functions. The duplications must have taken place at several different periods throughout evolution of the class II gene family. HLA-DM, for instance, is only weakly related to other class II sequences, thus resulting from an ancient gene duplication, whereas HLA-DO shares about 60% identity with HLA-DR, thus representing a more recent duplication.

      The class II region of the human and mouse MHC harbors a small number of genes involved in antigen presentation by MHCI. These encode the TAP transporter (TAP1, TAP2) as well as subunits of the immunoproteasome (LMP2, LMP7). In addition, many loci exist in the MHC that do not play any role in the immune system (131).

      

Fig. 3 The class II region of the human MHC (131).

      The class II region of the human MHC extends over 1'000 kb from its centromeric to telomeric end (from right to left) of the short arm of chromosome 6. The approximate positions and transcriptional orientations of all identified genes (in red) and pseudogenes (in gray) are shown. The expressed immune loci are marked with an asterix. The classical class II region is highlighted by a blue background, the extended class II region by a pink background.


2.1.5 Polymorphism

      Classical MHCII genes exhibit an extraordinary degree of allelic polymorphism. There are more than 200 alleles of some MHCII loci. The polymorphic residues are primarily those that interact with the TCR and determine the shape of the peptide-binding groove. They consequently influence both the recognition of the MHCII-peptide complex by the TCR and the peptide specificities of different MHCII alleles. In addition, MHCII polymorphism contributes to the susceptibility to autoimmune diseases, has important functional implications in clinical organ and bone marrow transplantation, and provides a very useful set of markers for the field of human population genetics.

      The nonclassical class II molecules are relatively invariant. Some alleles of both HLA-DM and HLA-DO have been described, but these vary by small numbers of amino acids and, so far, have no known functional significance. In addition, the human DRa chain has also been shown to be monomorphic.


2.1.6 Formation of MHCII-peptide complexes

      MHCII molecules are synthesized in the endoplasmatic reticulum (ER). In the ER, three a/b MHCII dimers associate with a trimer of invariant chains (Ii) (10) (Fig.4). A number of functional consequences of the association of MHCII with Ii trimers are known. First, Ii serves as a scaffold to facilitate protein folding and assembly of MHCII molecules. Second, Ii occupies the class II peptide binding cleft, thereby blocking premature MHCII-peptide association. Third, Ii-association is critical for normal intracellular trafficking of MHCII molecules, since MHCII molecules synthesized in the absence of Ii aggregate and do not exit the ER (132, 133). The Ii-associated MHCII molecules exit the ER and are transported via the Golgi to the endocytic pathway, more precisely to a late endosomal antigen-processing compartment called MIIC (11). This process is regulated by two di-leucine endocytic signals present in the Ii cytosolic domain (134). Once in the MIIC, the Ii-associated MHCII complexes meet an acidic, protease rich environment, where the Ii chain is degraded by the stepwise action of aspartyl proteases and cathepsins (12, 13). This generates the intermediates Iip23 and Iip10 and finally CLIP (class II associated Ii peptide). CLIP is derived from Ii amino acid residues 81-104 and binds to the antigen-binding groove of MHCII molecules (Fig.2). MHCII molecules then become competent to bind antigenic peptides that come from the degradation of exogenous proteins. Efficient exchange of CLIP for antigenic peptides is catalyzed by HLA-DM. HLA-DM binds transiently to MHCII-CLIP and stabilizes an intermediate state where CLIP is released, allowing other peptides to bind (14-16). HLA-DM dependent peptide loading occurs within the context of other associated molecules, such as HLA-DO (135, 136) and tetraspanins (137). HLA-DO is a B-cell specific and pH-dependent modulator of HLA-DM. It effectively blocks HLA-DM function at the endosomal pH, while allowing HLA-DM action at the more acidic pH of the MIICs (135, 138, 139). In this way, HLA-DO action may skew the class II-peptide loading process toward acidic compartments. Once stable complexes of MHCII and antigenic peptides are formed, they are released from HLA-DM and transported to the cell surface.

      

Fig. 4 Formation of MHCII-peptide complexes.

      Newly synthesized ab dimers associate with Ii in the ER and are transported through the Golgi apparatus. The Ii-associated MHCII molecules are diverted from the constitutive seccretory pathway into the endocytic pathway. This can occur following a direct pathway from the Golgi to a specialized antigen-processing compartment (MIIC) or indirectly after transit through to the cell surface and endocytosis. In the MIIC, Ii is degraded leaving behind the CLIP peptide. Exogenous antigens are derived from proteins that are endocytosed and processed by proteases. The peptides bind to MHCII molecules replacing CLIP in an HLA-DM dependent manner, and the MHCII-peptide complexes are then transported to the plasma membrane.


2.1.7 Expression of MHCII molecules

      The pattern of MHCII expression is complex and very tightly regulated. Under normal conditions, most classical and non-classical MHCII genes are regulated in a coordinated fashion. Two general modes of MHCII expression can be distinguished: constitutive and inducible (17-21).

      Constitutive MHCII expression is largely restricted to APCs, such as dendritic cells (DCs), B cells and macrophages. In addition, certain epithelial and endothelial cells, most importantly cTECs, and activated human T cells express MHCII constitutively. Expression within an APC lineage can vary dramatically as a function of developmental stage. For instance, immature pre-B cells are MHCII negative, but become positive as they mature into B cells. Upon terminal differentiation of B cells into plasmocytes, MHCII expression is silenced again. In immature DCs, only few MHCII molecules are displayed at the cell surface, but they are upregulated dramatically with the induction of DC maturation. In many MHCII-negative cell types, MHCII expression can be induced by various stimuli of which IFN-g is the most potent and well known. Constitutive and inducible MHCII expression can be modulated by a large number of stimuli (17-21). For example, IL-4, IL-10 and IL-13 enhance MHCII expression in B cells, whereas glucocorticoids and prostaglandins diminish it. IFN-g induced MHCII expression is positively regulated by IL-4 and TNF-a, but negatively by TGF-b, IL-10, CSF-1, and type 1 interferons.


2.2 Regulation of MHCII expression

      The molecular mechanisms regulating MHCII expression are the subject of a review that has been accepted for publication.

      "CIITA and the MHCII enhanceosome in the regulation of MHCII expression"

      Salomé Landmann, Jean-Marc Waldburger, Krzysztof Masternak, Annick Mühlethaler-Mottet and Walter Reith.

      Current Genomics (2002), in press.


Summary

      MHCII molecules direct the development, activation and homeostasis of CD4+ T cells. Given these key functions it is not surprising that the absence of MHCII expression results in a severe primary immunodeficiency disease called MHCII deficiency or the Bare Lymphocyte syndrome (BLS). The genetic defects responsible for BLS lie in genes encoding transcription factors required for MHCII expression. Four different MHCII regulatory genes encoding RFXANK, RFX5, RFXAP and CIITA have been identified. The first three are subunits of RFX, a ubiquitously expressed factor that binds cooperatively with other proteins to MHCII and related promoters to form a highly stable macromolecular nucleoprotein complex referred to as the MHCII enhanceosome. This enhanceosome serves as a landing pad for the MHCII transactivator CIITA. CIITA is a non-DNA binding coactivator that serves as the master control factor for MHCII expression. The highly regulated expression pattern of CIITA ultimately dictates the cells type specificity, induction and level of MHCII expression. The enhanceosome and CIITA collaborate in activating transcription by promoting histone hyperacetylation and by recruiting components of the general transcription machinery.

      CIITA AND THE MHCII ENHANCEOSOME IN THE REGULATION OF MHCII EXPRESSION

      S.Landmann, J.-M..Waldburger, K.Masternak, A.Muhlethaler-Mottet and W.Reith

      Department of Genetics and Microbiology, University of Geneva Medical School


Abstract

      Major histocompatibility complex class II (MHCII) molecules direct the development, activation and homeostasis of CD4+ T cells. Given these key functions it is not surprising that the absence of MHCII expression results in a severe primary immunodeficiency disease called MHCII deficiency or the Bare Lymphocyte Syndrome (BLS). The genetic defects responsible for BLS lie in genes encoding transcription factors required for MHCII expression. Four different MHCII regulatory genes encoding RFXANK, RFX5, RFXAP and CIITA have been identified. The first three are subunits of RFX, a ubiquitously expressed factor that binds cooperatively with other proteins to MHCII and related promoters to form a highly stable macromolecular nucleoprotein complex referred to as the MHCII enhanceosome. This enhanceosome serves as a landing pad for the MHCII transactivator CIITA. CIITA is a non-DNA binding coactivator that serves as the master control factor for MHCII expression. The highly regulated expression pattern of CIITA ultimately dictates the cell type specificity, induction and level of MHCII expression. The enhanceosome and CIITA collaborate in activating transcription by promoting histone hyperacetylation and by recruiting components of the general transcription machinery. In this review we summarize what is known about the molecular basis of BLS and what this has taught us about the mechanisms regulating transcription of MHCII and related genes. Particular attention is devoted to the structure, function and mode of action of the MHCII enhanceosome and CIITA. In addition, we focus on the highly regulated and cell type specific expression of CIITA.


Introduction

      MHCII molecules are heterodimeric (a chain - b chain) transmembrane glycoproteins displayed at the surface of specialized cells of the immune system. In humans, there are three MHCII isotypes designated HLA-DR, HLA-DQ and HLA-DP. All three isotypes serve the same function, namely the presentation of peptides to the TCR of CD4+ T helper cells. This is crucial for numerous aspects of the adaptive immune system, including the selection, activation and survival of CD4+ T cells [1,2]. Two general modes of MHCII expression exist: constitutive and inducible [3-9]. Constitutive expression is the hallmark of professional APCs. These include B lymphocytes, cells of the monocyte/macrophage lineage and DCs. cTECs and activated human T cells are also MHCII positive. Most other cell types do usually not express MHCII molecules but can be induced to do so in response to various stimuli of which IFN-g is the most potent and well known. Both constitutive and induced MHCII expression can be further modulated by additional signals. Constitutive expression in B cells and DCs is for instance regulated as a function of developmental stage. MHCII expression is extinguished upon differentiation of B cells into plasmocytes. In DCs, maturation is accompanied by an increase in cell surface MHCII expression. Finally, IFN-g induced expression can be modified by various stimuli. For example TGF-b, IFN-a and IL-4 inhibit induction of MHCII by IFN-g.

      The central importance of correctly regulated MHCII expression is underlined by the fact that deregulation of this expression leads to disease. Aberrant or inappropriate MHCII expression has been incriminated in the pathology of certain CD4+ T cell-mediated autoimmune diseases [10]. On the other hand, the lack of MHCII expression severely cripples the immune system and leads to a life-threatening immunodeficiency syndrome [7-9,11-14].A detailed understanding of the molecular mechanisms that control MHCII expression thus represents an important contribution to both molecular immunology and immunopathology.


The Bare Lymphocyte Syndrome


A disease resulting from the absence of MHCII expression

      The absence of MHCII expression is the cause of a primary immunodeficiency syndrome called the Bare Lymphocyte Syndrome (BLS) [7-9,11-16]. BLS is a rare autosomal recessive disease and has a high incidence of consanguinity [17]. Since its first description in the late 1970s and early 1980s [18-23], only about 70 patients from 50 unrelated families have been reported.

      Detailed descriptions of the clinical and immunopathological characteristics of the disease have been published previously [11-13,24], and we will thus restrict ourselves here to a brief description of the most salient features. Symptoms start within the first year of life, and comprise primarily severe and repeated infections, protracted diarrhea, malabsorption, and failure to thrive. The patients are particularly prone to infections of the gastrointestinal, pulmonary, upper respiratory and urinary tracts. The type of infectious agent is not pathognomic, and they can be of viral, bacterial, fungal or protozoan origin. Few children reach puberty. The majority dies between the age of 6 months and 5 years.

      The mainstay of diagnosis is the absence of constitutive and inducible expression of MHCII genes [11-14,24]. Moreover, expression of the HLA-DM genes is suppressed and expression of the invariant chain Ii is reduced [25,26]. The level of MHCI expression is also reduced to a variable extent in many BLS patients. However, the clinical and immunopathological manifestations of the disease are thought to result mainly from the defect in MHCII expression.

      Total numbers of circulating T and B lymphocytes are normal. However, in the majority of patients, the ratio of CD4+ to CD8+ T cells is reduced [11-13,24]. There is a reduction in the absolute number of CD4+ T cells and a concomitant increase in CD8+ T cells. This presumably reflects a deficiency in positive selection of CD4+ thymocytes, and is a consequence of the lack of MHCII expression on cTECs. Albeit reduced in numbers, the remaining CD4+ T cells in the patients do not exhibit major phenotypic abnormalities [27]. Although the residual CD4+ T cells in BLS patients may be functional, the absence of MHCII on APCs and the resulting inability to present antigens to CD4+ T cells leads to a severely compromised immune system [11-13,24]. Humoral immune responses to immunizations and to infectious agents are absent or strongly reduced. Cellular immune responses are also defective.

      In all patients, the lack of cell surface MHCII expression is a consequence of the fact that the corresponding genes are not transcribed [28]. Several lines of evidence have shown that the defects responsible for the disease do not reside in the MHCII locus itself, but lie in transacting regulatory genes controlling MHCII expression [12,17,29-34] (Fig.5).

      

Fig. 5 BLS is a disease of gene regulation.

      The MHC2TA, RFXANK, RFX5 and RFXAP genes are mutated in the BLS complementation groups A, B, C and D, respectively. These four genes are essential for the expression of MHCII and related genes. These include the genes coding for HLA-DR, HLA-DP, HLA-DQ, HLA-DM, HLA-DO and the invariant chain Ii.


Genetic heterogeneity in the cause of BLS

      BLS is considered to be a single phenotypic entity. However, the disease is genetically heterogeneous [17,30-32]. Four different BLS complementation groups (A to D), reflecting the existence of mutations in four distinct MHCII regulatory genes, have been defined [17,30-32]. The regulatory gene affected in each complementation group has now been identified. The four affected genes are called MHC2TA (group A), RFXANK (group B), RFX5 (group C) and RFXAP (group D) [35-39]. The patient-derived cell lines, together with several experimentally generated MHCII negative mutants, have thus allowed the isolation and characterization of four key MHCII specific transcription factors and have hence constituted a unique tool for the investigation of the regulation of MHCII gene expression.

      The nature of the genetic defect does not correlate with the course of the disease. There is no obvious genotype-phenotype relationship between the regulatory gene that is affected and either the immunological features, the precise clinical picture, the severity of the disease or the prognosis. Despite the genetic heterogeneity, BLS is therefore clinically homogenous.

      Certain BLS patients display only mild symptoms or are even assymptomatic [40]. A patient exhibiting a very late onset of the disease (in his late twenties) has also been described [41]. Interestingly, in contrast to the situation observed in the majority of classical patients, the mutations identified in late onset patients do not destabilize the affected factor [40]. As the other 'classical' patients, these individuals completely lack MHCII expression or display only very low residual levels of MHCII molecules at the surface of their cells. Moreover, they have been assigned to one of the four known complementation groups. It is thus clear, that some patients display an attenuated clinical phenotype despite their profound defect in MHCII expression. This has two important implications. First, there are likely to be additional - as yet undefined - 'modifier' loci that have a major impact on the severity and course of the disease. Second, the frequency of the inherited MHCII deficiency disease is likely to be underestimated.


Lessons from mouse models of BLS

      There are no spontaneous animal models for BLS. However, identification of the genes affected in the human disease has permitted the generation of two mouse models. Knockout mice reproducing the molecular defects of BLS patients in groups A and C have been constructed by deletion of the Mhc2ta and Rfx5 genes [42-45]. The major immunopathological characteristics of the human disease are reproduced in these mice. Both models show a strong reduction of constitutive MHCII expression on professional APCs (B cells, DCs and macrophages). IFN-g induced expression is also abolished. As in humans, this loss of MHCII expression results in a severely compromised immune system.

      The absence of CD4+ T cell-dependent immune responses in the mouse models is a consequence both of the inability to present antigen via MHCII molecules and of a severe CD4+ T cell deficiency. The CD4+ T cell population in Mhc2ta and Rfx5 knockout mice is decreased over tenfold [42-45]. This strong reduction is a consequence of severely impaired positive selection, which results from an essentially complete loss of MHCII expression on epithelial cells in the thymus. Surprisingly, the reduction in CD4+ T cells in BLS patients is rarely greater than two to three fold [11,12] and is thus considerably less pronounced than in the knockout mice. At least two explanations could account for this difference between the human and mouse phenotypes. First, the relatively mild decrease in peripheral CD4+ T cells in BLS patients suggests that positive selection may be compromised only partially. This could be explained by the retention of a sufficient level of residual MHCII expression in the thymus. Whether or not this is indeed the case remains to be determined because MHCII expression patterns in the thymus have only been examined in a few isolated cases [46]. It may be relevant that residual expression of MHCII in the thymus has been observed in one patient [46]. A second interpretation could be that the CD4+ T cells in BLS patients have escaped the normal selection processes in the thymus. They may for example have been selected on ligands other than MHCII molecules. In MHCII deficient mice, for instance, a large proportion of the residual CD4+ cells are CD1-restricted [47]. Such alternative selection pathways could be more prominent in BLS patients. Interestingly, an analysis of the T cell repertoire in BLS patients has revealed minor alterations, suggesting that the CD4+ T cells in these patients may indeed have been subjected to an unusual selection mechanism [48,49].

      Both the Mhc2ta-/- and Rfx5-/- mice exhibit residual MHCII expression in certain tissues and cell types. The precise pattern of residual expression differs between the two mice. This implies that specific cellular compartments possess alternative pathways for MHCII expression that can partially bypass the strict requirement for RFX5 and/or CIITA [42-44]. Rfx5-/- mice retain relatively strong MHCII expression in the thymic medulla and significant, albeit weaker, expression on a fraction of splenic and bone marrow derived dendritic cells. Low MHCII expression is also induced on B cells from Rfx5-/- mice following activation in vitro with lipopolysaccharide (LPS) and/or IL-4. In contrast, residual MHCII expression in the Mhc2ta-/- mice is mainly restricted to DCs in the paracortex of lymph nodes, B cells in the germinal centers of the spleen and lymph nodes, and a subset of cTECs. This difference in the residual expression pattern is surprising because the human disease is considered to be phenotypically homogeneous. Leaky expression has been observed in cells from certain BLS patients, but no characteristic residual expression pattern distinguishing RFX5 deficient patients from those with defects in CIITA have been described [11,12,24]. This discrepancy may reflect species-specific differences in the dependence on the two MHCII regulatory genes. However, it is also possible that the phenotypic differences observed in the mouse system exist in the human disease as well, but have escaped attention until now. Due to the rarity and severity of the disease, only relatively few patients from defined complementation groups have been studied in detail with respect to residual MHCII expression.


Therapeutic strategies


Carrier detection and prenatal diagnosis for BLS

      Considering the rarity of BLS, prenatal/postnatal diagnosis or carrier detection on a population-wide scale is not justified. On the other hand, these procedures could be valuable if restricted to families that already have affected children, or when a consanguineous union is envisaged in a high risk population. Previously, prenatal diagnosis relied on the analysis of MHCII expression on fetal leukocytes obtained by an umbilical vein puncture [50]. With our current knowledge of the molecular defects in BLS, mutated alleles of MHC2TA, RFXANK, RFX5 or RFXAP can now be screened for by using flanking polymorphic markers or by a direct search for known mutations. The greater sensitivity of these molecular techniques will allow the development of prenatal diagnosis by less invasive procedures, such as choriocentesis or amniocentesis. In addition, thanks to these techniques healthy carriers can now be identified rapidly and given accurate genetic counseling.


Gene therapy for BLS

      Allogeneic bone marrow transplantation (BMT) is currently the only curative treatment available for BLS [51]. A cord blood transplantation from an HLA-identical sibling has also been successful in the case of one patient with an allogeneic BMT graft failure [52]. The overall success rate of BMT in this disease has been reported to be relatively poor as compared to other immunodeficiency syndromes. This does not appear to be a peculiarity resulting from the immunological phenotype of BLS. Instead, it is likely to be due largely to other problems, such as the fact that the disease is often diagnosed at a rather late age, when recurrent illness will have compromised the success of BMT [51,53]. Now that the affected genes have been identified, gene therapy becomes a potential alternative to BMT. Introduction of the wild type MHC2TA, RFXANK, RFX5 or RFXAP genes into hematopoietic stem cells (HSCs) of BLS patients in complementation groups A, B, C or D, respectively, would represent a logical therapeutic strategy. The Mhc2ta and Rfx5 knockout mice will be valuable for developing gene therapy for BLS.

      Abnormal selection of CD4+ T cells, resulting from the absence of MHCII expression on thymic epithelial cells, is unlikely to represent an obstacle for gene therapy in BLS. The fact that classical BMT can cure BLS suggests that restoring MHCII expression on thymic epithelial cells is not essential. This is consistent with the fact that successful BMT does not require a functional thymus; the T cell pool can be reconstituted through peripheral expansion of passenger donor T cells present in the graft [54]. In this respect it is also worth mentioning that in a MHCI deficient mouse model, MHCI expression on bone marrow derived cells was found to be sufficient to compensate for the defect in thymic selection of CD8+ T cells [55].

      One potential concern is that the therapeutic transgene could induce ectopic or non-physiological levels of MHCII expression, which could have deleterious consequences and result in tissue destruction or autoimmunity. This is unlikely to be a major problem in the case of RFX5, RFXANK and RFXAP, which are expressed ubiquitously in all cell types. The MHC2TA gene, on the other hand, is tightly regulated and it will be difficult to obtain a pattern of transgene expression that mimics that of the endogenous gene.


RFX and CIITA as targets for novel immunomodulators

      MHCII expression is completely lost or severely reduced in BLS patients. This is also the case in most cell types in the two mouse models for the human disease. This suggests that CIITA, RFXANK, RFX5 and RFXAP are essential and that no bypass or alternative pathways can compensate efficiently for their absence. Moreover, no other major systems are as critically dependent on these regulatory factors. These features imply that inhibitors specific for CIITA, RFXANK, RFXAP and RFX5 should induce a highly selective downregulation of MHCII expression. CIITA, RFXANK, RFXAP and RFX5 may thus represent prime targets for novel immunomodulatory drugs having wide applications in situations such as organ transplantation and autoimmune diseases, where inhibition of MHCII expression might be desirable or beneficial.


Tumor immunotherapy.

      Tumors frequently have reduced immunogenicity because they lack MHCII molecules. The immunogenicity and rejection of such tumors can be increased by the expression of MHCII alone or in combination with costimulatory molecules [56-58]. This has of course raised hopes that the introduction of CIITA to activate MHCII expression might enhance tumor immunogenicity and thus contribute to the success of tumor therapy. Experiments of this type have been performed [59], although with limited success so far.


Regulation of MHCII expression


Overview

      MHCII expression is regulated primarily at the level of transcription [3-9,14]. The promoter proximal region of MHCII and related genes contains the conserved cis-regulatory elements referred to as W (or S), X, X2 and Y 'boxes' (Fig.6). These elements are highly conserved in their sequence, orientation, order and spacing relative to each other, and they function together as a single composite enhancer unit [3-9,14].

      

Fig. 6 The promoter proximal region of MHCII and related genes is bound by the MHCII enhanceosome, which serves as a landing pad for the recruitment of CIITA.

      a) The promoter proximal regions of MHCII and related genes contain the conserved W, X, X2 and Y cis-regulatory elements. These elements are bound by the RFX, X2BP and NF-Y complexes and a putative W binding protein. These factors function together as a composite unit called the MHCII enhanceosome. The latter recruits the master regulator CIITA to the promoters. b) CIITA is mutated in BLS complementation group A. In the absence of CIITA, MHCII and related promoters are occupied by the enhanceosome but are not transcribed. c) In BLS complementation groups B, C and D one of the subunits of the RFX complex is mutated. In the absence of an intact RFX complex, the enhanceosome can not assemble, CIITA is not recruited, and MHCII and related promoters remain silent.

      The X box is bound by the trimeric factor RFX, which is composed of RFX5, RFXANK and RFXAP. The X2 box is recognized by X2BP [60,61]. Finally, the trimeric complex NF-Y, composed of NF-YA, NF-YB and NF-YC, binds to the Y box [62]. Several proteins have been described to bind to the W box, but none has been definitely demonstrated to be the functionally relevant W box binding protein. All of the factors binding to the cis-regulatory elements are required for MHCII gene expression. However, they are ubiquitously expressed and fail to account for either constitutive or IFN-g inducible MHCII gene expression. Instead, they bind cooperatively to the promoter to form a landing pad for an additional factor that plays an essential regulatory role [63]. This factor is the class II transactivator CIITA [35]. The latter ultimately determines the cell type specificity, inducibility and expression level of MHCII genes. CIITA is now widely accepted as being the master regulator of MHCII expression [9,64].


The RFX complex

      The RFX (regulatory factor X) complex is composed of three subunits RFXANK, RFX5 and RFXAP. These factors are defective in BLS complementation groups B, C and D, respectively (Fig.6C). The three subunits of the RFX complex are unrelated and share no sequence homology. The gene encoding the 75 kD RFX5 subunit was identified by means of a genetic complementation approach using a cell line derived from a group C patient [36]. RFX5 was the fifth member of the RFX family of DNA binding proteins to be identified [65]. All members of this family share a characteristic DNA binding domain (DBD) referred to as the RFX motif [65,66]. The three dimensional structure of the RFX motif of one family member (RFX1) has been solved and was found to belong to the winged helix subfamily of helix-turn-helix proteins [67]. The DBD of RFX5 lies near its N-terminus [68] (Fig.7C). Outside of the DBD, RFX5 contains no obvious functional motifs. There is a centrally placed proline rich region, of which the function remains unknown.

      A biochemical approach was used to isolate the gene encoding the 36 kD RFXAP (RFX associated protein) subunit. This approach relied on affinity purification of the RFX complex and sequencing of peptides derived from its 36 kD subunit [37]. RFXAP contains an acidic domain, a glutamine rich segment and a basic region resembling a bipartite nuclear localization signal (NLS) [69,70] (Fig.7D).

      

Fig. 7 The domain structure and function of CIITA, RFXANK, RFX5 and RFXAP.

      a) The 1130 amino acid CIITA protein contains an N-terminal acidic domain (DE), an adjacent proline/serine/threonine rich (PST) region, a central GTP binding domain and a series of leucine rich repeats (LRR) at the C-terminus. At least three nuclear localization signals (NLS 1 to 3) are found in the protein. The N-terminus functions as a transcription activation domain and can interact with the indicated general transcription factors and coactivators. The central and C-terminal domains are required for recruitment of CIITA to the MHCII enhanceosome and function as a dominant negative mutant in the absence of the N-terminal activation domain. Domains believed to be involved in self-association of CIITA are indicated. b) The hallmark of the 269 amino acid RFXANK protein is a C-terminal ankyrin repeat domain. This domain is sufficient for complementation of cell lines from BLS complementation group B. The N-terminal part of the protein contains an acidic (DE) domain. c) The 616 amino acid RFX5 protein is the only subunit of the RFX complex containing a DNA binding domain (DBD). This DBD lies in an N-terminal region that is sufficient for assembly with RFXANK and RFXAP. For complementation of RFX5 defective cell lines, a central region containing a proline rich domain (P) is required in addition to the DBD. d) The C-terminus of RFXAP is sufficient for complementation of BLS cells lines in group D. Expression of HLA-DR only requires a short segment spanning the glutamine rich region (Q). Expression of HLA-DP and HLA-DQ required a larger C-terminal region. The acidic region (DE) and NLS-like sequence are not essential for the function of the protein.

      RFXANK, the gene encoding the smallest 33 kD subunit of RFX was isolated by a biochemical approach similar to that used to isolate the RFXAP gene [38]. The gene was named RFXANK because it encodes a factor that contains a protein-protein interaction domain consisting of ankyrin repeats (Fig.7B). The same gene was isolated independently by another laboratory [39]. They called the encoded factor RFXB to indicate that this factor is mutated in BLS complementation group B. The N-terminal region of RFXANK is rich in acidic amino acids [39,69].

      RFX5, RFXAP and RFXANK assemble to form the RFX complex that binds to the X box of MHCII promoters. The complex pre-assembles in solution before binding to its target site. The RFX complex can be faithfully reconstituted with the three recombinant proteins, indicating that the complex does not contain additional subunits [68-71]. However, the stoichiometry of the three subunits in the RFX complex remains to be determined. The absence of any one of the three subunits destroys the RFX complex and eliminates its ability to bind to DNA. This is illustrated by cell lines from BLS complementation groups B, C and D, which all lack RFX binding activity [28,72,73]. Experiments with recombinant RFX proteins have also shown that the individual subunits or combinations of two of the subunits are not capable of binding specifically to DNA in band shift assays [68-70]. RFX5 is the only RFX subunit containing a well defined DNA binding domain, although DNA-protein crosslink experiments have shown that all three subunits contact the DNA at the X box [74].

      Specific domains that are necessary for complex formation and DNA binding activity have been identified in each of the RFX subunits (Fig.7B-D). In RFX5, the N-terminus (amino acids 39 to 194) comprising the DBD is sufficient for interaction with RFXAP, RFXANK and X box DNA [68,69]. In RFXAP, the 49 C terminal amino acids spanning the highly conserved glutamine rich region are sufficient for complex formation and binding to the X box [69,70]. Finally, the ankyrin repeat region of RFXANK (amino acids 84 to 269) meets all the requirements needed to form an efficient DNA bound complex with RFXAP and RFX5 [69] [our unpublished data]. The minimal domain of RFXANK allowing RFX complex formation and DNA binding is sufficient to restore MHCII expression in cell lines from RFXANK deficient BLS cells [69]. In the case of RFXAP, the minimal 49 amino acid C-terminal region is sufficient to restore HLA-DR expression in BLS cells lacking RFXAP [70]. However, a longer segment of RFXAP is required to restore expression of HLA-DQ and HLA-DP [70]. The domain of RFX5 mediating complex formation and DNA binding (amino acids 39 to 194), is not sufficient for complementation of RFX5 deficient cells [68]. An additional domain located between amino acids 410 to 515 is required [68]. This region has been shown to mediate cooperative DNA binding between RFX and NF-Y [68].

      Outside of the essential domains defined above, several features within RFX appear to be dispensable. An acidic region and a putative nuclear localization signal present in RFXAP can be removed or mutated without affecting the ability of the protein to restore MHCII expression in cells from group D [69,70]. The N-terminal acidic region of RFXANK is not essential because it can be removed without eliminating the ability to complement cells from group B [69] [our unpublished data].


The MHCII enhanceosome

      The trimeric RFX complex binds with relatively low affinity to the X box of MHCII promoters. However, this binding is greatly stabilized by synergistic interactions with two other factors binding to MHCII promoters, namely with the trimeric NF-Y complex [68,71,75] and the dimeric protein X2BP [61,76,77]. X2BP has recently been proposed to contain CREB [78]. The combination of these cooperative interactions results in stable occupation of MHCII promoters.

      The quaternary complex composed of RFX, X2BP (CREB), NF-Y and MHCII promoter DNA is extremely stable with a half-life greatly exceeding that of any of the possible secondary or ternary complexes [77]. RFX plays a crucial role in promoting this cooperativity. This important role of the RFX complex is strongly emphasized by the analysis of promoter occupation in RFX deficient BLS cells (complementation groups B, C and D). MHCII promoters in these cell lines are unoccupied and lack the DNaseI hypersensitive sites typically observed in RFX positive cells [79-81].

      The stable higher order nucleoprotein complex that assembles at MHCII promoters has been coined the 'MHCII enhanceosome' [63]. The enhanceosome can be isolated from cell extracts by promoter pull down assays, and has been shown to contain all of the known subunits of RFX, X2BP (CREB) and NF-Y [63]. It may also contain additional components but these remain to be characterized.

      Formation of the enhanceosome requires the presence and proper arrangement of all cis-regulatory elements, that is to say the W, X, X2 and Y boxes. The spacing between the X and Y elements is highly conserved at approximately two helical turns [82-84]. Correct stereoalignment is essential for promoter activation, probably because the X and Y box binding factors need to be assembled on the same side of the DNA helix. It was also observed that the spacing between the W and X boxes is critical for promoter activation [83,84] [A.Muhlethaler-Mottet, unpublished results].

      In addition to the crucial role of the MHCII enhanceosome in enhancing the stability and specificity of promoter occupation, it constitutes a platform onto which the transcriptional coactivator CIITA is recruited. CIITA has been directly implicated as a key protein mediating transcriptional activation of MHCII genes (see below). However, it has been recognized recently that the enhanceosome also contributes directly to the process of transcription activation independently of CIITA [85]. The relative contributions of the enhanceosome and CIITA to transcription activation depends on the MHCII promoter [85]. For instance, the enhanceosome has a dominant role in histone acetylation and recruitment of general transcription factors (GTF) at the DMB promoter. At the DRA promoter on the other hand the dominant role in these processes is played by CIITA. Promoters such as DPB depend equally on both the enhanceosome and CIITA. The example of the DMB promoter demonstrates that the enhanceosome is not merely a landing pad for CIITA, but can actually support CIITA independent events contributing directly to transcription initiation.

      B cells display constitutively occupied MHCII promoters, even in the absence of CIITA (cells in group A) [80,81]. However, CIITA is clearly required for promoter occupancy in IFN-g induced or CIITA transfected fibroblasts [86-88]. It is not clear how CIITA induces promoter occupation in these cells. One possibility is that binding of CIITA favors assembly and/or stability of the enhanceosome by interacting with many of its individual components (see below). Alternatively, CIITA may facilitate in vivo occupation indirectly by changing accessibility of the promoter DNA.


The master regulator CIITA

      The enhanceosome is essential but not sufficient for appropriate MHCII gene expression. This is demonstrated by the study of B cell lines from BLS patients in complementation group A. Enhanceosome assembly, promoter occupation and the presence of DNase hypersensitive sites are all normal at MHCII promoters in these cells (Fig.6B). Yet, the MHCII genes are not transcribed. The defect in this complementation group lies in the MHC2TA gene encoding the regulatory factor CIITA [35]. MHC2TA was isolated by the same approach employed for isolation of the RFX5 gene, namely by complementation cloning using the in vitro generated cell line RJ2.2.5 (complementation group A) [35].

      CIITA expression is a nearly absolute prerequisite for both the constitutive or inducible expression of MHCII genes [35,42,89-94].The expression profile of CIITA is the major determinant dictating the tightly controlled pattern of MHCII expression [9,64]. The following findings emphasize this key regulatory function of CIITA. 1) A quantitative correlation between MHC2TA and MHCII expression levels has been revealed by the analysis of a large number of human and mouse tissues and cell lines [95]. In addition, an experimental setup using a tetracycline inducible system has shown that the level of MHC2TA expression directly determines the level of MHCII expression [95]. 2) Most cell types do not express MHC2TA and are thus MHCII negative. Transfection of these cells with a CIITA expression plasmid is generally sufficient to render them MHCII positive [89,90,96,97]. 3) IFN-g induces expression of the MHC2TA gene, and this mediates activation of MHCII genes [89,96]. 4) Induction of MHCII expression upon activation of human T cells is mediated by upregulation of MHC2TA expression [98] [our unpublished data]. 5) B cells loose MHCII expression during their differentiation into plasmocytes because the MHC2TA gene is repressed [90,99]. Exogenous expression of CIITA is sufficient to re-establish MHCII expression in plasmocytes [90,99]. 6) Silencing of MHCII expression in trophoblast cells is caused by inhibition of MHC2TA expression [94,100]. 7) Cytokines modulating IFN-g induced MHCII expression, such as TGF-b, IL-1, IL-4 and IL-10, do so by modulating MHC2TA expression [101-105]. 8) Pathogens such as cytomegalovirus, varicella-zoster virus, Mycobacterium bovis and Chlamydia can downregulate MHCII expression by interfering with transcription of the MHC2TA gene [106-109]. In summary, CIITA exerts a strict qualitative and quantitative control over MHCII expression and thus deserves the title of the master regulator of MHCII expression [9,64].

      The control of CIITA over MHCII expression is achieved mainly via the level of transcription of the MHC2TA gene. However, a few examples of other levels of control are known. The inhibitory effect of IFN-b and TNF-a on IFN-g induced MHCII expression is exerted downstream of MHC2TA transcription [110-112]. The transactivating capacity of CIITA can be modulated by posttranslational modifications. protein kinase A (PKA) can inactivate CIITA by phosphorylating it at serine residues 834 and 1050 [113]. Phosphorylation of CIITA by PKA is thought to be involved in the inhibition of MHCII expression by prostaglandin E (PGE), since PKA is a downstream effector in the PGE/cAMP cascade.

      A few examples of CIITA independent MHCII expression have been reported. Isotype specific MHCII expression has been observed in certain mutant human cell lines devoid of functional CIITA [114]. MHCII expression has also been detected in certain tissues of CIITA-/- mice, although the levels are significantly lower than in control animals [42,44]. For the moment, these cases of CIITA independent expression remain unresolved exceptions.


Regulation of MHC2TA expression


The MHC2TA gene is controlled by multiple promoters

      The MHC2TA gene resides in a locus known as AIR-1 on chromosome 16 (16p13) [115]. Mhc2ta, the homologous mouse gene, is localized in a syntenic region on mouse chromosome 16. The genomic organization of the entire mouse gene has been determined. It consists of 19 exons spanning 42 kb of DNA [our unpublished data].

      Expression of MHC2TA is controlled by four different promoters (pI to pIV) [116] (Fig.8). Three of these promoters are highly conserved between the human and mouse genes (pI, pIII and pIV). pII has only been found in the human gene. It displays only a very low transcriptional activity and its significance is not known. pII will thus not be discussed further. The different promoters do not share any sequence homology and are not co-regulated. They span a large (> 12 kb) genomic region. Each promoter precedes a distinct first exon that is spliced alternatively to a shared second exon. This leads to the production of three types of transcript (type I, type III and type IV), each possessing a different 5' end [116] (Fig.8).

      The shared second exon contains an AUG translation initiation codon that can be used in all three types of transcript to give rise to a 1106 amino acid protein. However, the first exons of the type I and type III transcripts each contain an additional in-frame translation initiation codon. Usage of these alternative start codons leads to the synthesis of protein isoforms of 1207 and 1130 amino acids, respectively. The three CIITA isoforms have apparent molecular weights of 132 kD, 124 kD and 121 kD (Fig.8). All three protein variants exist in vivo. The first CIITA cDNA clone to be isolated corresponded to a CIITA type III mRNA. Therefore, nucleotide and amino acid numbering are generally given with respect to the transcription initiation site of type III mRNA and the translational initiation codon of the 1130 kD isoform.

      

Fig. 8 Expression of the MHC2TA gene is controlled by four different promoters (pI to pIV).

      pI is active in dendritic cells and in IFN-g activated monocytes and macrophages. pII is only found in humans and its significance is unknown. pIII is used in B cells, in activated human T cells and in some types of dendritic cells. pIV is responsive to IFN-g and is constitutively expressed in cTECs. The three types of mRNAs (types I, III and IV) initiated at pI, pIII and pIV encode three different protein isoforms (121, 124 and 132 kD). These proteins differ only at their N-terminal end. Gray bars represent exons. white bars represent mRNAs and black bars represent proteins encoded by these mRNAs. The boundary between the alternative first exons and the shared downstream exons is indicated by a vertical line. The positions of translation initiation codons are indicated.

      The N-terminal extensions encoded by the unique first exons of CIITA type I and type III mRNAs may confer specific properties on the corresponding 132 kD and 124 kD CIITA isoforms. It has been suggested that the 132 kD isoform of CIITA, which is mainly expressed in DCs, has an enhanced capacity to activate MHCII transcription compared to the smaller two isoforms [117] [Luc Otten, unpublished results]. In addition, this isoform seems to have an extremely short half-life [our unpublished observations]. It is likely that these features are due to an intrinsic property of the N-terminus unique to this CIITA isoform. A recent study has proposed that the N-terminal extension of the 132 kD isoform displays homology to caspase recruitment domains (CARD) [117], which are found in components of the apoptosis and NF-kB signaling pathways [118,119]. The significance of this is not clear because CIITA is not known to be involved in either of these pathways.

      Cell type specific and modulated expression of the MHC2TA gene is controlled by the differential activities of the three promoters. It is thus the sophisticated transcriptional control of the MHC2TA gene that dictates the cell type specific and inducible expression of MHCII genes.


pI and expression of MHC2TA in DC

      MHC2TA pI is highly specific for DCs. CIITA type I transcripts always represent a preponderant fraction in various DC preparations including ex vivo mouse DCs, mouse bone marrow-derived DCs, long-term mouse DC cultures and human monocyte-derived DCs [116,120-122] [our unpublished data]. However, CIITA type III transcripts are also found in significant amounts in certain DC preparations [121]. Moreover, DC specificity of pI is not as absolute as first thought. pI can for instance also be induced by IFN-g in microglia and peritoneal macrophages [120,122].

      A recent report from our laboratory has characterized in detail the activity of pI (and pIII) during the course of maturation of human monocyte-derived DCs [121]. In response to a variety of stimuli, such as infection by bacteria or viruses, immature DCs are induced to undergo profound changes in their morphology and function. Changes in the expression of MHCII represent a key aspect of this maturation process. The number of MHCII molecules expressed at the cell surface is increased as a result of changes in the intracellular localization and stability of preexisting MHCII protein. In contrast, de novo synthesis of MHCII molecules is shut down. This reduction in MHCII synthesis during DC maturation is a consequence of rapid silencing of the MHC2TA gene. The arrest in CIITA expression is the result of a transcriptional inactivation of the MHC2TA gene. This is mediated by a global repression mechanism implicating histone deacetylation over a large domain spanning the entire MHC2TA regulatory region, including the two promoters (pI and pIII) that are active in DCs.


pIII and expression of MHC2TA in B cells

      MHC2TA pIII is used primarily in B cells [116,123], activated human T cells [98] [our unpublished observations] and to a lesser extent in some DC subsets [116,121]. Aberrant CIITA expression in melanoma cells was also reported to be initiated at pIII [124,125]. A 320 bp promoter proximal regulatory region is sufficient for B cell specific activity [116,123]. This region contains five sequence elements that can be shown to be occupied in vivo in B cells in genomic footprinting experiments [126] (Fig.9B). At least two of these elements are critical for proper activity, namely a TEF-2 like element and a binding site for a novel transcription factor [126].

      A 5' flanking sequence situated approximately five kb upstream of the transcription initiation site of pIII of the human MHC2TA gene has been reported to confer IFN-g responsiveness [103]. The functional relevance of this sequence is supported by the presence of a DNaseI hypersensitive site at this position in vivo [127]. The putative regulatory region acts as a STAT-1 dependent and IRF-1 independent enhancer [127]. In contrast to these findings in human cells, pIII does not seem to be inducible by IFN-g in primary rat astrocytes [112] and in mouse macrophages [122]. There may thus be a species specific difference in the IFN-g responsiveness of pIII.

      Expression of the MHC2TA gene is actively silenced during terminal differentiation of B cells into plasmocytes [90,99]. The sequence elements of pIII that are occupied in normal B cells are completely bare in plasmocytes [90,126]. The human positive regulatory domain I binding factor-1 (PRDI-BF1) [128] and its mouse homologue B lymphocyte-induced maturation protein-1 (Blimp-1) [129] have been proposed to play a crucial role in the repression of pIII in plasmocytes [130,131]. PRDI-BF1/Blimp-1 is upregulated when B cells differentiate into plasmocytes [132]. It has also been shown to drive, at least partially, the final differentiation of B cells into plasmocytes if expressed ectopically in BCL1 lymphoma cells [132]. PRDI-BF1/Blimp-1 can bind specifically to a sequence in the promoter proximal region of pIII [130,131]. However, no occupation of this site is seen in in vivo footprint experiments performed with plasmocytes [126]. The mechanism, by which PRDI-BF1/Blimp-1 silences the MHC2TA gene, is thus not well understood.

      

Fig. 9 The promoter proximal regions of CIITA pI, pIII and pIV contain several cis-regulatory elements.

      a) The regulation of pI is not well understood. The depicted elements represent only potential consensus binding sites. The arrowheads represent protections and enhancements observed in in vivo footprint experiments at the sense (downwards) and antisense (upwards) strands in human dendritic cells. b) pIII contains five cis-regulatory elements defined by in vivo footprint experiments (arrowheads). They are called site A, site B, ARE-2, ARE-1 and site C. In vitro binding studies have revealed that ARE-1 may be bound by AML2 in B and T cells, ARE-2 by a member of the CREB/ATF protein family in B and T cells, site A by NF-1 in B cells and site C by PRDI-BF1 in plasmocytes. c) The regulation of IFN-g induced pIV expression is well understood. Binding of IFN-g to its receptor induces JAK-1 and JAK-2 activation and thus STAT-1 phosphorylation, dimerization and translocation to the nucleus. STAT-1 binds cooperatively with USF-1 to a composite GAS/E box motif. STAT-1 also activates expression of IRF-1, which then binds to its cognate site in pIV. The arrowheads depict protections and enhancement observed in in vivo footprint experiments. The mechanism mediating expression of pIV in cTECs is not known.


pIV and IFN-g induced MHC2TA expression

      MHC2TA pIV is induced by IFN-g in most cell types, including cells of the monocyte/macrophage lineage, endothelial cells, fibroblasts, astrocytes and cTECs [103,105,110,116,127,133]. A 300 bp promoter proximal region is sufficient for the IFN-g response [116,127,133]. This region contains a GAS element, an E box and an IRF-1 binding site (Fig.9C), all three of which are required for induction in most cell types [103,110,116,133,134]. The former two are bound cooperatively by STAT-1 and upstream regulatory factor-1 (USF-1) [103,110,133]. STAT-1 is activated and translocated to the nucleus by the classical IFN-g signal transduction pathway. USF-1 is a constitutively expressed member of the basic helix-loop-helix/leucine zipper family. In rat astrocytes, STAT-1 is dispensable for IFN-g induced activation of pIV [112]. This may represent a species or cell type specific exception to the general requirement for STAT-1 in IFN-g induced pIV activity. The IRF-1 binding site of pIV is occupied by IRF-1. IRF-1 synthesis itself is induced by IFN-g. Thus, the dependence of CIITA on IRF-1 explains the delayed kinetics of CIITA induction by IFN-g. [89]. One group has proposed that IRF-1 and IRF-2 bind cooperatively to pIV [135], and found a reduction of CIITA expression in IRF-2 knockout mice [136]. IRF-2 may thus also be involved in activation of pIV.

      IFN-g activated expression of CIITA can be suppressed by a number of different stimuli. TGF-b markedly attenuates IFN-g induced CIITA expression. The inhibitory mechanism involves suppression at the level of MHC2TA transcription [101,102,104]. Surprisingly, TGF-b does not affect IFN-g induced phosphorylation of JAK-1, JAK-2 and STAT-1. Nor does it interfere with binding of STAT-1, USF-1 or IRF-1 to pIV of the MHC2TA gene [137,138]. Moreover, TGF-b even inhibits basal non-induced expression levels of pIV [103,139]. Finally ,activity of the putative IFN-g response element situated upstream of pIII is also inhibited by TGF-b [7]. A recent report has implicated Smad-3 in the inhibitory effect of TGF-b on CIITA expression [138]. IL-1, IL-4 and IL-10 have also been shown to exert an inhibitory effect similar to TGF-b on CIITA transcription in human astrocytes (IL-1) [105] and in mouse microglia cells (IL-4, IL-10) [104]. The role of IL-4 mediated CIITA inhibition seems to be cell type dependent.

      IFN-g induced gene activation is generally a transient event. Suppressor of cytokine signaling-1 (SOCS-1) has been shown to be induced by IFN-g and this protein negatively regulates the IFN-g signal transduction pathway by binding to JAK-2 and inhibiting its kinase activity [140,141]. SOCS-1 can thus also suppress IFN-g activated expression of pIV of the MHC2TA gene [134].

      Trophoblasts lack MHCII and CIITA expression in order to evade maternal immune recognition. They do not respond to IFN-g, although MHCII can be induced by ectopic expression of CIITA in these cells [94,100]. Recently, two laboratories have reported that suppression of CIITA expression in trophoblasts is due to methylation of pIV [142,143].


Lessons from pIV knockout mice

      Mhc2ta pIV knockout mice have been generated in our laboratory [122]. These mice allowed us to define precisely the cell type specificity of pIV in vivo. pIV -/- mice exhibit selective abrogation of IFN-g induced MHCII expression on a wide variety of non-hematopoietic cells, including endothelia, epithelia, astrocytes and fibroblasts. The cTECs of these mice are also MHCII negative. In contrast, constitutive and inducible MHCII expression is unaffected on professional APCs including B cells, DCs and IFN-g activated cells of the monocyte/macrophage lineage.

      The phenotype of the pIV -/- mice demonstrates that professional APCs do no depend on pIV for IFN-g inducible expression of CIITA. Ex vivo macrophages from control mice use both, pI and pIV in response to IFN-g. Macrophages from pIV deficient animals express similar amounts of total CIITA mRNA in response to IFN-g but transcripts are derived exclusively from pI. It remains unknown, how IFN-g activates pI. No IFN-g responsive sequences have been identified in the vicinity of pI. It is thus possible that IFN-g may affect pI indirectly as a consequence of macrophage activation. The answer to this question awaits further dissection of the regulatory mechanism controlling pI.

      The finding that constitutive MHCII expression on cTECs is abolished was completely unexpected. The loss of MHCII molecules on cTECs in the pIV knockout mice results in the abrogation of positive selection of CD4+ T cells in the thymus. CD4+ T cell counts in these mice are strongly reduced in the thymus and in the periphery. pIV is thus essential for positive selection of CD4+ T cells.

      Isolated cTECs loose MHCII expression if they are not maintained in three-dimensional re-aggregate thymus organ cultures [144,145]. This indicates that a stimulus must be required to maintain constitutive MHCII expression on cTECs. This stimulus could be provided either by cell-cell interactions between the cTECs or by a short acting soluble factor produced by the cTECs. The precise nature of the signal required for MHCII expression in cTECs as well as its mode of action is still unknown. The phenotype of the pIV-/- mice indicates that this signal functions by activating pIV of the MHC2TA gene. However, it must achieve this via a mechanism that is independent of the IFN-g signaling pathway, because knockout mice lacking key components of this pathway, such as the IFN-g receptor, STAT-1 and IRF-1, have normal MHCII expression on cTEC and normal positive selection of CD4+ T cells.

      The Mhc2ta pIV knockout mice have provided definitive evidence that differential CIITA promoter usage does indeed play an important physiological role. They have also allowed us to draw a more accurate picture of the differential promoter usage. Cells of myeloid origin (DCs, monocytes/macrophages) use mainly pI for constitutive or IFN-g induced CIITA expression. Constitutive CIITA expression in cells of lymphoid origin (B cells, T cells) is driven by pIII. Finally, pIV is indispensable for IFN-g activated expression in non-hematopoietic cells and for expression in cTECs.


Structure and function of CIITA


Structure of CIITA

      Given the key role of CIITA as the master regulator of MHCII expression, a considerable amount of effort has been devoted to the elucidation of its structure and mode of action. Analysis of the primary sequence of CIITA has revealed four major features of interest, namely an N-terminal acidic domain, an adjacent proline/serine/threonine rich (PST) region, a central GTP binding domain and a series of leucine rich repeats (LRR) at the C-terminus [9,64] (Fig.7A). Several other proteins have been found to share a similar domain structure. In particular, the nucleotide binding domain and the leucine rich repeats are conserved in a number of proteins [64]. Intriguingly, these proteins have functions very distinct from that of CIITA. They include for instance Nod-1, which induces activation of caspase-9 and NF-kB, and certain plant resistance proteins [119,146]. The similarity in the domain structure between CIITA and these unrelated proteins may indicate a yet unknown analogy in their mode of action.


Recruitment of CIITA to the MHCII enhanceosome

      In vivo association of CIITA with MHCII and related promoters was recently demonstrated by chromatin immunoprecipitation experiments [63,147]. Moreover, DNA dependent co-immunoprecipitation assays and pull-down experiments using immobilized promoter templates have demonstrated a direct physical interaction between CIITA and the MHCII enhanceosome [63]. This interaction requires the integrity of the enhanceosome and depends on multiple, synergistic protein-protein interactions with the multiple factors constituting the enhanceosome.

      CIITA has been demonstrated to interact with several components of the MHCII enhanceosome. An interaction with RFX5 was first revealed in a yeast-two-hybrid system [148]. In vitro interactions were subsequently also demonstrated with RFX5, RFXANK, NF-YB, NF-YC and CREB [69,84,149]. The regions of CIITA that are required for recruiting CIITA to the enhanceosome [63,149] and for co-immunoprecipitation with these proteins [84] have been partially mapped.


Activation of transcription by CIITA

      CIITA activates transcription of MHCII genes when recruited to the enhanceosome. The regions of the protein that are essential for this transactivation have been mapped to the N-terminal acidic and PST domains [150-152]. These domains are believed to constitute transcription activation domains and they resemble the activation domains found in other transcription factors. When fused to a GAL4 DNA binding domain, the N-terminus of CIITA can activate transcription of a reporter gene [150]. Moreover, the transcription activation domains of VP16 and HSV1a transducing factor can substitute for the N-terminus of CIITA [150,151]. This substitution is only partial, indicating that the N-terminus of CIITA may also support additional functions.

      Deletion of the acidic and (or) PST domains results in dominant negative mutants that repress the activity of wild type CIITA in B cells and fibroblasts [152-156]. In vitro promoter pull down experiments have shown that these dominant negative mutants of CIITA are recruited more efficiently to MHCII promoters than wild type CIITA [63]. This suggests that their dominant negative phenotype is due at least in part to an increased affinity for the enhanceosome.

      Various general transcription factors or coactivators can interact with the N-terminus of CIITA (Fig.10). The TAFII32 and TAFII70 subunits of the TFIID complex, TFIIB, TFIIH, pTEFb, CBP and pCAF have been shown to be able to interact with this region [154,157-160]. This suggests that the N-terminus of CIITA acts as a transcriptional activation domain recruiting factors involved in transcription initiation (TFIID, TFIIB), promoter clearance and transcription elongation (TFIIH, pTEFb), or chromatin remodeling (CBP, pCAF). The interactions of CIITA with CBP and pCAF have been addressed in greater detail. First, CBP and pCAF can both increase, synergistically with CIITA, the expression of a HLA-DRA reporter gene [158,161,162]. Surprisingly, this cooperativity has been shown to be independent of the histone acetyltransferase (HAT) activity of CBP and pCAF [162]. Second, the synergy between CIITA and CBP is inhibited by a dominant negative variant of CBP [158]. Finally, when overexpressed, CIITA can sequester CBP and thus downregulate other CBP dependent genes [158] (see below) [163].

      

Fig. 10 CIITA is a non-DNA binding transcriptional coactivator that functions via multiple protein-protein interactions.

      Interactions with several enhanceosome components tether CIITA to MHCII and related promoters. Once tethered to the promoter, CIITA is believed to activate transcription by recruiting other factors including TFIIB, TFIID, TFIIH, pTEFb, CBP and pCAF. These factors are involved in transcription initiation (TFIIB, TFIID), promoter clearance and transcription elongation (TFIIH, pTEFb) and chromatin remodeling (CBP, pCAF). Protein-protein interactions are indicated by double headed arrows.

      Acetylation of lysines in the N-terminal tails of histones H3 and H4 is generally associated with increased transcriptional activity of eukaryotic genes [164-166]. In MHCII and related genes, increased acetylation of histones H3 and H4 correlates with binding of CIITA to the enhanceosome in various cell types [85,147]. The high acetylation status observed in the wild type B cell line Raji is strongly reduced in the CIITA deficient cell line RJ2.2.5 [85,147]. In IFN-g induced HeLa cells, the kinetics of the increase in H3 and H4 acetylation correlates well with the binding of CIITA to the HLA-DRA promoter [147]. Both processes are very dynamic. Binding of CIITA and acetylation start to increase within a few hours after induction by IFN-g, are maximal after 24 hours and then decrease. Acetylation of the HLA-DRA promoter is pronounced at the promoter proximal region [85,147] but is also seen as far as 15 kb upstream of the transcription initiation site [K.Masternak, unpublished results] Multiple HAT activities may be involved in histone acetylation at MHCII promoters. These include the CBP and pCAF co-activators that can bind to CIITA. It has also been proposed that CIITA itself contains an intrinsic acetyl transferase activity [167].


The GTP binding domain and the leucine rich repeat region

      CIITA is the first transcription factor known to have a GTP binding domain. This GTP binding domain (amino acids 421 to 561) is composed of a Walker A motif (also called P loop) involved in phosphate binding, a magnesium binding site and a guanine coordination site. This tripartite motif is essential for GTP binding and CIITA function [152,168]. Mutants of the GTP binding domain fail to translocate to the nucleus (see below). Mice bearing a deletion spanning the GTP binding domain of the Mhc2ta gene display a phenotype similar to that of other CIITA knockout mice [45].

      LRR domains are known to mediate protein-protein interactions and are found in many different classes of proteins. The LRRs found at the C terminus of CIITA (amino acids 988 to 1097) are essential for its function [149,169]. Partial deletion of these LRRs generates a CIITA mutant displaying a dominant negative phenotype [169].


Dimerization of CIITA

      Several laboratories have reported that CIITA can bind to itself via intra- or intermolecular contacts [170-172]. However, there is no uniform consensus on the precise regions involved in this self-association. The central region encompassing the GTP binding site may contact the acidic and PST region [170,172] or may instead bind to the C-terminus [171,172]. LXXLL motifs residing in the GTP binding domain may play a crucial role for self-association [172]. In vitro translated CIITA does not dimerize, indicating that additional factors or modifications may be involved [170-172]. Finally, the precise role of self-association of CIITA is not certain. It may play a role in subcellular localization of CIITA [170]. It should be mentioned here that many of the results described above have been obtained under conditions of overexpression of CIITA and should therefore be interpreted with caution.


Nuclear localization of CIITA

      Depending on the cell type and the method of detection, CIITA is either found in both the nucleus and cytoplasm or found predominantly in the nucleus [149,160,168,170,173,174]. As mentioned above for dimerization, nuclear localization studies have been done mainly with overexpressed CIITA. Three potential NLS's have been identified in CIITA (Fig.7A). NLS1 (amino acids 141 to 159) resembles a bipartite NLS [160]. Nuclear localization mediated by NLS1 is affected positively by CBP and pCAF, but this is independent of their HAT activity [160,162]. NLS2 has been localized between amino acids 405 and 414 [174]. It displays similarity to the classical SV40 NLS. Deletion of NLS2 results in cytoplasmic localization and the loss of transactivation by CIITA [174]. Finally, NLS3 was identified thanks to a BLS patient expressing a CIITA mutant that lacks an exon spanning amino acids 940 to 963 [173]. This exon contains a RFLKK motif that functions as a NLS. The loss of this exon abolishes nuclear import of CIITA [173]. In addition to the three NLS's, GTP binding and integrity of the LRRs have been found to be crucial for proper nuclear import of CIITA [149,168,175]. Finally, nuclear export sequences (NES's) have also been found to influence the subcellular distribution of CIITA. Two regions in CIITA have been found to be crucial for nuclear export and can bind to the export receptor CRM1 [170]. The presence of both NLS's and NES's suggests that CIITA shuttles back and forth between the nuclear and cytoplasmic compartments.


The role of CIITA and the enhanceosome in MHCI expression

      In addition to their crucial role in MHCII gene expression, both CIITA and the MHCII enhanceosome contribute to the regulation of MHCI genes [63,176-178]. The W, X, X2 and Y sequence elements characteristic of MHCII promoters are also present in the promoter proximal regions of the MHCI and b2m genes. The X2 element in MHCI promoters corresponds to a sequence that was first called site a. The latter was defined as a sequence that has homology to the cAMP response element (CRE) and is bound by members of the ATF/CREB family. The Y box of MHCI promoters was initially called enhancer B. CIITA and the RFX complex contribute to the expression of MHCI genes by interacting with the W-X-X2-Y region of MHCI promoters. This has been demonstrated by chromatin immunoprecipitation experiments and by transactivation assays [63,176-178]. The contribution of CIITA and RFX to the activation of MHCI promoters explains the finding that BLS patients frequently have reduced levels of MHCI expression [11,12].

      Compared to their essential role at MHCII promoters, CIITA and RFX contribute only relatively weakly to the expression of MHCI genes. This is evident from the phenotype of BLS patients, which exhibit a complete absence of MHCII expression but retain strong, albeit reduced, levels of MHCI expression [11,12]. The same is also true for mouse models of the human disease [42-45]. CIITA deficient mice do not display significantly reduced MHCI expression. In RFX5 deficient mice MHCI expression is reduced two to ten fold in certain cell types (B and T lymphocytes) [our unpublished data]. The fact that RFX and CIITA are less important for MHCI expression is readily explained by the presence of additional key regulatory sequences present in the promoters of MHCI and b2m genes.


Additional targets of CIITA

      Although MHCII and related genes are undoubtedly the major target genes of CIITA, it has recently also been implicated in other systems. For several genes it has been proposed that CIITA functions as a negative regulator. These include the IL-4 gene, the Fas ligand gene and the collagen a2 (I) gene [163,179,180]. In all three examples, sequestration of CBP by CIITA has been proposed to account for the repression. The same mechanism may also be implicated in downregulation of the thymidine kinase and cyclin D1 genes [163].

      IL-4, a Th2-type cytokine, was found to be aberrantly expressed in Th1 cells derived from CIITA knock-out mice [179]. Moreover, ectopic expression of CIITA in Th2 cells from CIITA knockout mice was found to suppress the expression of Th2 cytokines [179]. Taken together, these results suggested that CIITA can inhibit expression of the IL-4 gene. The negative effect of CIITA was proposed to be due to a competition between NF-AT, an essential IL-4 gene transcription factor, and CIITA for binding to CBP [181]. Whether repression of IL-4 expression by CIITA actually occurs in vivo remains to be demonstrated. In this respect it should be mentioned that ectopic expression of CIITA at physiological levels in transgenic mice does not affect IL-4 expression in Th2 cells [Luc Otten, unpublished data].

      NF-AT is also an important transcription factor for Fas ligand expression. As for the IL-4 gene, the activity of NF-AT on the FasL promoter was shown to be repressed by overexpression of CIITA [180].

      The collagen a2(I) gene was recently recognized as a new target for repression by CIITA [163]. The mechanism proposed for downregulation of this gene again involves binding of CIITA to CBP.

      The promoter of the HIV provirus recently attracted attention as a novel target of transcriptional regulation by CIITA. Saiffudin et al. reported that expression of CIITA increases LTR promoter activity and HIV replication in fibroblast and T cell lines [182,183]. This was speculated to be relevant because CIITA is expressed in activated human T cells and macrophages, both of which are primary targets of HIV infection. However, in other studies HIV infection and LTR activity were found to be reduced in CIITA positive B and T cell lines [184,185]. It is therefore presently not clear exactly how CIITA affects the level of transcription from the HIV promoter under physiological conditions.

      The viral transactivator Tat was reported to be able to downregulate MHCII expression in HIV infected or Tat transfected fibroblasts and T cell lines [159,184,186]. MHCII protein and mRNA levels were reduced in these cells while CIITA expression levels were not affected. Conversely, overexpression of CIITA was found to inhibit Tat-mediated transactivation of the HIV-1 LTR [184]. Tat and CIITA can both bind to the same surface in cyclin T1, one of the subunits of the elongation factor pTEFb. It was therefore proposed that CIITA and Tat compete for binding to pTEFb [159,184]. In addition, a mechanism independent of pTEFb has also been implicated in inhibition of CIITA activity by Tat [187]. Taken together, these findings suggest that the actions of Tat and CIITA may be mutually antagonistic. It is likely that the precise outcome of the reciprocal actions of Tat and CIITA will depend on the relative concentrations of the two proteins and possibly on the cell type in which they are expressed. It remains to be determined if Tat is inhibited by or inhibits CIITA in vivo during the course of infection by HIV.


CIITA as a target of pathogens

      Pathogens have developed a wide variety of strategies to escape immune surveillance by their hosts. In order to inhibit the establishment of a protective immune response, several bacteria and viruses have acquired the ability to downregulate MHCII expression and thus prevent activation of specific CD4+ T cells. They achieve this by interfering with the function or expression of CIITA. The intracellular bacterium Chlamydia downregulates CIITA expression by inducing degradation of USF-1, an ubiquitous factor that is required for the activation of pIV of the MHC2TA gene [108]. Infection with Mycobacterium bovis bacillus Calmette-Guérin also downregulates CIITA expression, but the precise mechanisms that is involved remains unknown [107]. Varizella zoster virus inhibits IFN- g induced transcription of the MHC2TA gene by interfering with the IFN-g signaling pathway. It blocks STAT1a and JAK2 protein expression and therefore transcription of the downstream IRF-1 and CIITA genes [109]. Human Cytomegalovirus has also developed mechanisms to modulate IFN-g induced CIITA expression. Infection with this virus can direct JAK-1 protein to the proteasome for degradation or affect the IFN-g signaling pathway downstream of STAT phosphorylation and nuclear translocation [106,188]. Finally, as mentioned above, the HIV Tat protein can inhibit MHCII expression by interfering with the function of CIITA [159,184,186]. Taken together, these findings demonstrate that many pathogens have acquired a means to target CIITA in order to escape immune recognition and establish persistence. It will be of great interest to further our understanding of the strategies used by pathogens to interfere with CIITA and MHCII expression. This will contribute to the design of new approaches for fighting these pathogens.


Acknowledgments

      The authors would like to thank all past and current members of the laboratory for their contributions to the work reviewed here. The work performed in the laboratory of the authors was supported by the Swiss National Science Foundation, the NCCR-NEURO, the Gabriella Giorgi-Cavaglieri Foundation, the Ernst and Lucie Schmidheiny Foundation, the Roch Research Foundation and the Novartis Stiftung.


References

  1. tructrucCresswell P: Assembly, transport, and function of MHC class II molecules. Annu.Rev.Immunol. 1994, 12: 259-293.
  2. tructrucViret C and Janeway CAJ: MHC and T cell development. Reviews in Immunogenetics 1999, 1: 91-104.
  3. tructrucBenoist C and Mathis D: Regulation of major histocompatibility complex class II genes: X, Y and other letters of the alphabet. Ann.Rev.Immunol. 1990, 8: 681-715.
  4. tructrucGlimcher LH and Kara CJ: Sequences and factors: A guide to MHC class-II transcription. Annu.Rev.Immunol. 1992, 10: 13-49.
  5. tructrucTing JP and Baldwin AS: Regulation of MHC gene expression. Curr.Opin.Immunol. 1993, 5: 8-16.
  6. tructrucBoss JM: Regulation of transcription of MHC class II genes. Curr.Opin.Immunol. 1997, 9: 107-113.
  7. tructrucMasternak K, Muhlethaler-Mottet A, Villard J, Peretti M, Reith W: Molecular genetics of the Bare lymphocyte syndrome. Rev.Immunogenet. 2000, 2: 267-282.
  8. tructrucWaldburger JM, Masternak K, Muhlethaler-Mottet A, Villard J, Peretti M, Landmann S, Reith W: Lessons from the bare lymphocyte syndrome: molecular mechanisms regulating MHC class II expression. Immunol.Rev. 2000, 178: 148-165.
  9. tructrucReith W and Mach B: The bare lymphocyte syndrome and the regulation of mhc expression. Annu.Rev.Immunol. 2001, 19: 331-373.
  10. tructrucGuardiola J and Maffei A: Control of MHC class II gene expression in autoimmune, infectious, and neoplastic diseases. Crit.Rev Immunol. 1993, 13: 247-268.
  11. tructrucKlein C, Lisowska-Grospierre B, LeDeist F, Fischer A, Griscelli C: Major histocompatibility complex class II deficiency: clinical manifestations, immunologic features, and outcome. J.Pediatr. 1993, 123: 921-928.
  12. tructrucGriscelli C, Lisowska-Grospierre B, Mach B: Combined immunodeficiency with defective expression in MHC class II genes. In Immunodeficiencies. F.S.Rosen and M.Seligman, editors. Harwood Academic Publishers, Chur, Switzerland, 1993, 141-154.
  13. tructrucElhasid R and Etzioni A: Major histocompatibility complex class II deficiency: a clinical review. Blood Rev. 1996, 10: 242-248.
  14. tructrucMach B, Steimle V, Martinez-Soria E, Reith W: Regulation of MHC class II genes: Lessons from a disease. Annu.Rev.Immunol. 1996, 14: 301-331.
  15. tructrucTouraine JL, Marseglia GL, Betuel H, Souillet G, Gebuhrer L: The bare lymphocyte syndrome. Bone Marrow Transplant. 1992, 9 Suppl 1: 54-56.
  16. tructrucPrimary immunodeficiency diseases: report of a WHO scientific group. Immunodefic.Rev. 1992, 3: 195-236.
  17. tructrucLisowska-Grospierre B , Fondaneche MC, Rols MP, Griscelli C, Fischer A: Two complementation groups account for most cases of inherited MHC class II deficiency. Hum.Mol.Genet. 1994, 3: 953-958.
  18. tructrucTouraine J L, Betuel H, Souillet G: Combined immunodeficiency disease associated with absence of cell surface HLA A and B antigen. J.Pediatr. 1978, 93: 47-51.
  19. tructrucSchuurmann RKB, Van Rood JJ, Vossen JM, Schellekens PTA, Feltkamp-Vroom TM, Doyer E, Gmelig-Meyling F, Visser HKA: Failure of Lymphocyte-Membrane HLA A and B Expression in Two Siblings With Combined Immunodeficiency. Clin.Immunol.Immunopathol. 1979, 14: 418-434.
  20. tructrucGriscelli C, Durandy A, Virelizier JL, Hors J, Lepage V, Colombani J: Impaired cell-to-cell interaction in partial combined immunodeficiency with variable expression of HLA antigens. In Primary Immunodeficiencies. M.Seligman and W.H.Hitzig, editors. Elsevier/North Holland, Amsterdam, 1980, 499-503.
  21. tructrucKuis W, Roord J, Zegers BJM, Schuurmann RKB, Heijnen CJ, Baldwin WM, Goulmy E, Claas F, van de Griend RJ, Rijkers GT, Van Rood JJ, Vossen JM, Ballieux RE, Stoop RJ: Clinical and Immunological Studies in a Patient With the 'Bare Lymphocyte' Syndrome. In Bone Marrow Transplantation in Europe. J.L.Touraine, E.Gluckman, and C.Griscelli, editors. Amsterdam:Excerpta Medica, 1981, 201-208.
  22. tructrucLisowska-Grospierre B, Durandy A, Virelizier JL, Fischer A, Griscelli C: Combined Immunodeficiency with Defective Expression of HLA: Modulation of an Abnormal HLA Synthesis and Functional Studies. Birth Defects 1983, 19: 87-92.
  23. tructrucHadam MR, Dopfer R, Dammer G, Peter HH, Schlesier M, Müller C, Niethammer D: Defective expression of HLA-D-region determinants in children with congenital agammaglobulinemia and malabsorption: a new syndrome. In Histocompatibility Testing. E.D.Albert, M.P.Baur, and W.R.Mayr, editors. Springer-Verlag, Berlin Heidelberg, 1984, 645-650.
  24. tructrucReith W, Steimle V, Lisowska-Grospierre B, Fischer A, Mach B: Molecular Basis of Major Histocompatibility Class II Deficiency. In Primary Immunodeficiency Diseases, a Molecular and Genetic Approach. H.Ochs, J.Puck, and E.Smith, editors. Oxford University Press, New York, 1999, 167-180.
  25. tructrucNocera A, Barocci S, Depalma R, Gorski J: Analysis of Transcripts of Genes Located Within the HLA-D Region in B-Cells from an HLA-Severe Combined Immunodeficiency Individual. Hum.Immunol 1993, 38: 231-234.
  26. tructrucKern I, Steimle V, Siegrist CA, and Mach B: The two novel MHC class II transactivators RFX5 and CIITA both control expression of HLA-DM genes. Int.Immunol 1995, 7: 1295-1299.
  27. tructrucRieux Laucat F, Le Deist F, Selz F, Fischer A, de Villartay JP: Normal T cell receptor V beta usage in a primary immunodeficiency associated with HLA class II deficiency. Eur.J.Immunol. 1993, 23: 928-934.
  28. tructrucReith W, Satola S, Herrero Sanchez C, Amaldi I, Lisowska-Grospierre B, Griscelli C, Hadam MR, Mach B: Congenital immunodeficiency with a regulatory defect in MHC class II gene expression lacks a specific HLA-DR promoter binding protein, RF-X. Cell 1988, 53: 897-906.
  29. tructrucde Preval C, Lisowska-Grospierre B, Loche M, Griscelli C, Mach B: A trans-acting class II regulatory gene unlinked to the MHC controls expression of HLA class II genes. Nature 1985, 318: 291-293.
  30. tructrucHume CR and Lee JS: Congenital immunodeficiencies associated with absence of HLA class II antigens on lymphocytes result from distinct mutations in trans-acting factors. Hum.Immunol. 1989, 26: 288-309.
  31. tructrucBenichou B and Strominger JL: Class II-antigen-negative patient and mutant B-cell lines represent at least three, and probably four, distinct genetic defects defined by complementation analysis. Proc.Natl.Acad.Sci.USA 1991, 88: 4285-4288.
  32. tructrucSeidl C, Saraiya C, Osterweil Z, Fu YP, Lee JS: Genetic Complexity of Regulatory Mutants Defective for HLA Class- II Gene-Expression. J.Immunol. 1992, 148: 1576-1584.
  33. tructrucHasegawa SL, Riley JL, Sloan III JH, Boss JM: Protease treatment of nuclear extracts distinguishes between class II MHC X1 box DNA-binding proteins in wild-type and class II-deficient B cells. J.Immunol. 1993, 150: 1781-1793.
  34. tructrucRiley JL and Boss JM: Class II MHC transcriptional mutants are defective in higher order complex formation. J.Immunol. 1993, 151: 6942-6953.
  35. tructrucSteimle V, Otten LA, Zufferey M, Mach B: Complementation cloning of an MHC class II transactivator mutated in hereditary MHC class II deficiency. Cell 1993, 75: 135-146.
  36. tructrucSteimle V, Durand B, Barras E, Zufferey M, Hadam MR, Mach B, Reith W: A novel DNA binding regulatory factor is mutated in primary MHC class II deficiency (Bare Lymphocyte Syndrome). Genes & Dev. 1995, 9: 1021-1032.
  37. tructrucDurand B, Sperisen P, Emery P, Barras E, Zufferey M, Mach B, Reith W: RFXAP, a novel subunit of the RFX DNA binding complex is mutated in MHC class II deficiency. EMBO J 1997, 16: 1045-1055.
  38. tructrucMasternak K, Barras E, Zufferey M, Conrad B, Corthals G, Aebersold R, Sanchez JC, Hochstrasser DF, Mach B, Reith W: A gene encoding a novel RFX-associated transactivator is mutated in the majority of MHC class II deficiency patients. Nat.Genet 1998, 20: 273-277.
  39. tructrucNagarajan UM, Louis-Plence P, DeSandro A, Nilsen R, Bushey A, Boss JM: RFX-B is the gene responsible for the most common cause of the bare lymphocyte syndrome, an MHC class II immunodeficiency. Immunity 1999, 10: 153-162.
  40. *tructrucWiszniewski W, Fondaneche MC, Le Deist F, Kanariou M, Selz F, Brousse N, Steimle V, Barbieri G, Alcaide-Loridan C, Charron D, Fischer A, Lisowska-Grospierre B: Mutation in the class II trans-activator leading to a mild immunodeficiency. J.Immunol. 2001, 167: 1787-1794.
    The group A BLS patients described in this paper display an attenuated clinical phenotype or even an asymptomatic clinical course despite their profound defect in the expression of MHCII genes. More and more patients with such an attenuated phenotype are being identified indicating that the BLS disease may be more frequent than previously thought.
  41. tructrucQuan V, Towey M, Sacks S, Kelly AP: Absence of MHC class II gene expression in a patient with a single amino acid substitution in the class II transactivator protein CIITA. Immunogenetics 1999, 49: 957-963.
  42. tructrucChang CH, Guerder S, Hong SC, van Ewijk W, Flavell R: Mice lacking the MHC class II transactivator (CIITA) show tissue- specific impairment of MHC class II expression. Immunity 1996, 4: 167-178.
  43. tructrucClausen BE, Waldburger JM, Schwenk F, Barras E, Mach B, Rajewsky K, Forster I, Reith W: Residual MHC class II expression on mature dendritic cells and activated B cells in RFX5-deficient mice. Immunity 1998, 8: 143-155.
  44. tructrucWilliams GS, Malin M, Vremec D, Chang CH, Boyd R, Benoist C, Mathis D: Mice lacking the transcription factor CIITA--a second look. Int Immunol. 1998; 10: 1957-1967.
  45. tructrucItoh-Lindstrom Y, Piskurich JF, Felix NJ, Wang Y, Brickey WJ, Platt JL, Koller BH, Ting JP: Reduced IL-4-, lipopolysaccharide-, and IFN-gamma-induced MHC class II expression in mice lacking class II transactivator due to targeted deletion of the GTP-binding domain. J.Immunol. 1999, 163: 2425-2431.
  46. tructrucSchuurman HJ, van de Wijngaert FP, Huber JP, Schuurman RK, Zegers BJ, Roord JJ, Kater L: The thymus in 'bare lymphocyte' syndrome: significance of expression of major histocompatibility complex antigens on thymic epithelial cells in intrathymic T-cell maturation. Hum.Immunol. 1985, 13: 69-82.
  47. tructrucCardell S, Tangri S, Chan S, Kronenberg M, Benoist C, Mathis D:: CD1-restricted CD4+ T cells in major histocompatibility complex class II-deficient mice. J.Exp.Med. 1995, 182: 993-1004.
  48. tructrucvan Eggermond MC, Rijkers GT, Kuis W, Zegers BJ, van den Elsen PJ: T cell development in a major histocompatibility complex class II- deficient patient. Eur.J.Immunol. 1993, 23: 2585-2591.
  49. tructrucHenwood J, van Eggermond MCJA, van Boxel-Dezaire AHH, Schipper R, den Hoedt M, Peijnenburg A, Sanal Ö, Ersoy F, Rijkers GT, Zegers BJM, Vossen JM, van Tol MJD, van den Elsen PJ: Human T Cell Repertoire Generation in the Absence of MHC Class II Expression Results in a Circulating CD4+CD8- Population with Altered Physicochemical Properties of Complementarity-Determining Region 3. J.Immunol. 1996, 156: 895-906.
  50. tructrucDurandy A, Cerf-Bensussan N, Dumez Y, Griscelli C: Prenatal Diagnosis of Severe Combined Immunodeficiency With Defective Synthesis of HLA Molecules. Prenatal Diagnosis 1987, 7: 27-34.
  51. tructrucKlein C, Cavazzana-Calvo M, Le Deist F, Jabado N, Benkerrou M, Blanche S, Lisowska-Grospierre B, Griscelli C: Bone marrow transplantation in major histocompatibility complex class II deficiency: a single-center study of 19 patients. Blood 1995, 85: 580-587.
  52. tructrucBonduel M, Pozo A, Zelazko M, Raslawski E, Delfino S, Rossi J, Figueroa C, Sackmann MF: Successful related umbilical cord blood transplantation for graft failure following T cell-depleted non-identical bone marrow transplantation in a child with major histocompatibility complex class II deficiency. Bone Marrow Transplant. 1999, 24: 437-440.
  53. tructrucCanioni D, Patey N, Cuenod B, Benkerrou M, Brousse N: Major histocompatibility complex class II deficiency needs an early diagnosis: report of a case. Pediatr.Pathol.Lab.Med. 1997, 17: 645-651.
  54. tructrucHaynes BF, Markert ML, Sempowski GD, Patel DD, Hale LP: The role of the thymus in immune reconstitution in aging, bone marrow transplantation, and HIV-1 infection. Annu.Rev.Immunol. 2000, 18: 529-560.
  55. tructrucZinkernagel RM and Althage A: On the role of thymic epithelium vs. bone marrow-derived cells in repertoire selection of T cells. Proc.Natl.Acad.Sci.USA 1999, 96: 8092-8097.
  56. tructrucOstrand-Rosenberg S, Clements VK, Thakur A, Cole GA: Transfection of major histocompatibility complex class I and class II genes causes tumour rejection. J.Immunogenet. 1989, 16: 343-349.
  57. tructrucBaskar S, Clements VK, Glimcher LH, Nabavi N, Ostrand-Rosenberg S: Rejection of MHC class II-transfected tumor cells requires induction of tumor-encoded B7-1 and/or B7-2 costimulatory molecules. J.Immunol. 1996, 156: 3821-3827.
  58. tructrucPanelli MC, Wang E, Shen S, Schluter SF, Bernstein RM, Hersh EM, Stopeck A, Gangavalli R, Barber J, Jolly D, Akporiaye ET: Interferon gamma (IFNgamma) gene transfer of an EMT6 tumor that is poorly responsive to IFNgamma stimulation: increase in tumor immunogenicity is accompanied by induction of a mouse class II transactivator and class II MHC. Cancer Immunol.Immunother. 1996, 42: 99-107.
  59. tructrucMartin BK, Frelinger JG, Ting JP: Combination gene therapy with CD86 and the MHC class II transactivator in the control of lung tumor growth. J.Immunol. 1999, 162: 6663-6670.
  60. tructrucHasegawa SL and Boss JM: Two B cell factors bind the HLA-DRA X box region and recognize different subsets of HLA class II promoters. Nucleic Acids Res. 1991, 19: 6269-6276.
  61. tructrucMoreno CS, Emery P, West JE, Durand B, Reith W, Mach B, Boss JM: Purified X2 binding protein (X2BP) cooperatively binds the class II MHC X box region in the presence of purified RFX, the X box factor deficient in the bare lymphocyte syndrome. J.Immunol. 1995, 155: 4313-4321.
  62. tructrucMantovani R: The molecular biology of the CCAAT-binding factor NF-Y. Gene 1999, 239: 15-27.
  63. tructrucMasternak K, Muhlethaler-Mottet A, Villard J, Zufferey M, Steimle V, Reith W: CIITA is a transcriptional coactivator that is recruited to MHC class II promoters by multiple synergistic interactions with an enhanceosome complex. Genes Dev. 2000, 14: 1156-1166.
  64. tructrucHarton JA and Ting JP: Class II transactivator: mastering the art of major histocompatibility complex expression. Mol.Cell Biol. 2000, 20: 6185-6194.
  65. tructrucEmery P, Durand B, Mach B, Reith W: RFX proteins, a novel family of DNA binding proteins conserved in the eukaryotic kingdom. Nucl.Acids Res. 1996, 24: 803-807.
  66. tructrucEmery P, Strubin M, Hofmann K, Bucher P, Mach B, Reith W: A consensus motif in the RFX DNA binding domain and binding domain mutants with altered specifity. Mol.Cell.Biol. 1996, 16: 4486-4494.
  67. tructrucGajiwala KS, Chen H, Cornille F, Roques BP, Reith W, Mach B, Burley SK: Structure of the winged-helix protein hRFX1 reveals a new mode of DNA binding. Nature 2000, 403: 916-921.
  68. tructrucVillard J, Peretti M, Masternak K, Barras E, Caretti G, Mantovani R, Reith W: A functionally essential domain of RFX5 mediates activation of major histocompatibility complex class II promoters by promoting cooperative binding between RFX and NF-Y. Mol.Cell Biol. 2000, 20: 3364-3376.
  69. tructrucDeSandro AM, Nagarajan UM, Boss JM: Associations and interactions between bare lymphocyte syndrome factors. Mol.Cell.Biol. 2000, 20: 6587-6599.
  70. tructrucPeretti M, Villard J, Barras E, Zufferey M, Reith W: Expression of the three human major histocompatibility complex class II isotypes exhibits a differential dependence on the transcription factor RFXAP. Mol.Cell Biol. 2001, 21: 5699-5709.
  71. tructrucCaretti G, Cocchiarella F, Sidoli C, Villard J, Peretti M, Reith W, Mantovani R: Dissection of functional NF-Y-RFX cooperative interactions on the MHC class II E alpha promoter. J.Biol.Chem. 2000, 302: 539-552.
  72. tructrucHerrero Sanchez C, Reith W, Silacci P, Mach B: The DNA-binding defect observed in major histocompatibility complex class II regulatory mutants concerns only one member of a family of complexes binding to the X boxes of class II promoters. Mol.Cell Biol. 1992, 12: 4076-4083.
  73. tructrucMoreno CS, Rogers EM, Brown JA, Boss JM: Regulatory factor X, a bare lymphocyte syndrome transcription factor, is a multimeric phosphoprotein complex. J.Immunol. 1997, 158: 5841-5848.
  74. tructrucWesterheide SD and Boss JM: Orientation and positional mapping of the subunits of the multicomponent transcription factors RFX and X2BP to the major histocompatibility complex class II transcriptional enhancer. Nucleic Acids Res. 1999, 27: 1635-1641.
  75. tructrucReith W, Siegrist CA, Durand B, Barras E, Mach B: Function of major histocompatibility complex class II promoters requires cooperative binding between factors RFX and NF-Y. Proc.Natl.Acad.Sci.U.S.A. 1994, 91: 554-558.
  76. tructrucReith W, Kobr M, Emery P, Durand B, Siegrist CA, Mach B: Cooperative binding between factors RFX and X2bp to the X and X2 boxes of MHC class II promoters. J.Biol.Chem. 1994, 269: 20020-20025.
  77. tructrucLouis-Plence P, Moreno CS, Boss JM: Formation of a regulatory factor X/X2 box-binding protein/nuclear factor-Y multiprotein complex on the conserved regulatory regions of HLA class II genes. J.Immunol. 1997, 159: 3899-3909.
  78. tructrucMoreno CS, Beresford GW, Louis-Plence P, Morris AC, Boss JM: CREB regulates MHC class II expression in a CIITA-dependent manner. Immunity 1999, 10: 143-151.
  79. tructrucGönczy P, Reith W, Barras E, Lisowska-Grospierre B, Griscelli C, Hadam MR, Mach B: Inherited Immunodeficiency with a defect in a Major Histocompatibility Complex class II promoter-binding protein differs in the chromatin structure of the HLA-DRA gene. Mol.Cell.Biol. 1989, 9: 296-302.
  80. tructrucKara CJ and Glimcher LH: In vivo footprinting of MHC class II genes: bare promoters in the bare lymphocyte syndrome. Science 1991, 252: 709-712.
  81. tructrucKara CJ and Glimcher LH: Three in vivo promoter phenotypes in MHC class II deficient combined immunodeficiency. Immunogenetics 1993, 37: 227-230.
  82. tructrucVilen BJ, Cogswell JP, Ting JP: Stereospecific alignment of the X and Y elements is required for major histocompatibility complex class II DRA promoter function. Mol.Cell.Biol. 1991, 11: 2406-2415.
  83. tructrucVilen BJ, Penta JF, Ting JP: Structural constraints within a trimeric transcriptional regulatory region: Constitutive and Interferon-g inducible expression of the HLA-DRA gene. J.Biol.Chem. 1992, 267: 23728-23734.
  84. tructrucZhu XS, Linhoff MW, Li G, Chin KC, Maity SN, Ting JP: Transcriptional scaffold: CIITA interacts with NF-Y, RFX, and CREB To cause stereospecific regulation of the class II major histocompatibility complex promoter. Mol.Cell.Biol. 2000, 20: 6051-6061.
  85. **tructrucMasternak K and Reith W: Promoter-specific functions of CIITA and the MHC class II enhanceosome in transcriptional activation. EMBO J. 2002, 21: 1379-1388.
    This in vivo study demonstrates that the MHCII enhanceosome contributes at some promoters to histone acetylation and transcription factor recruitment in a CIITA-independent manner.
  86. tructrucRigaud G, De Lerma BA, Nicolis M, Cestari T, Ramarli D, Riviera AP, Accolla RS: Induction of CIITA and modification of in vivo HLA-DR promoter occupancy in normal thymic epithelial cells treated with IFN-gamma: similarities and distinctions with respect to HLA-DR-constitutive B cells. J.Immunol. 1996, 156: 4254-4258.
  87. tructrucWright KL, Chin KC, Linhoff M, Skinner C, Brown JA, Boss JM, Stark GR, Ting JP: CIITA stimulation of transcription factor binding to major histocompatibility complex class II and associated promoters in vivo. Proc.Natl.Acad.Sci.U.S.A 1998, 95: 6267-6272.
  88. tructrucVillard J, Muhlethaler-Mottet A, Bontron S, Mach B, Reith W: CIITA-induced occupation of MHC class II promoters is independent of the cooperative stabilization of the promoter-bound multi-protein complexes. Int.Immunol. 1999, 11: 461-469.
  89. tructrucSteimle V, Siegrist CA, Mottet A, Lisowska-Grospierre B, Mach B: Regulation of MHC class II expression by Interferon-gamma mediated by the transactivator gene CIITA. Science 1994, 265: 106-109.
  90. tructrucSilacci P, Mottet A, Steimle V, Reith W, Mach B: Developmental extinction of major histocompatibility complex class II gene expression in plasmocytes is mediated by silencing of the transactivator gene CIITA. J.Exp.Med. 1994, 180: 1329-1336.
  91. tructrucChin KC, Mao C, Skinner C, Riley LJ, Wright KL, Moreno CS, Stark GR, Boss JM, Ting JP: Molecular analysis of G1B and G3A IFN-gamma mutants reveals that defects in CIITA or RFX result in defective class II MHC and Ii gene induction. Immunity 1994, 1: 687-697.
  92. tructrucChang CH and Flavell RA: Class II transactivator regulates the expression of multiple genes involved in antigen presentation. J.Exp.Med. 1995, 181: 765-767.
  93. tructrucRamassar V, Goes N, Hobart M, Halloran PF: Evidence for the in vivo role of class II transactivator in basal and IFN-gamma induced class II expression in mouse tissue. Transplantation. 1996, 62: 1901-1907.
  94. tructrucMorris AC, Riley JL, Fleming WH, Boss JM: MHC class II gene silencing in trophoblast cells is caused by inhibition of CIITA expression. Am.J.Reprod.Immunol. 1998, 40: 385-394.
  95. tructrucOtten LA, Steimle V, Bontron S, Mach B: Quantitative control of MHC class II expression by the transactivator CIITA. Eur.J.Immunol. 1998, 28: 473-478.
  96. tructrucChang CH, Fontes JD, Peterlin M, Flavell RA: Class II transactivator (CIITA) is sufficient for the inducible expression of major histocompatibility complex class II genes. J.Exp.Med. 1994, 180: 1367-1374.
  97. tructrucSartoris S, Valle MT, Barbaro AL, Tosi G, Cestari T, D'Agostino A, Megiovanni AM, Manca F, Accolla RS: HLA class II expression in uninducible hepatocarcinoma cells after transfection of AIR-1 gene product CIITA: acquisition of antigen processing and presentation capacity. J.Immunol. 1998, 161: 814-820.
  98. tructrucHolling TM, van der Stoep N, Quinten E, van den Elsen PJ: Activated Human T Cells Accomplish MHC Class II Expression Through T Cell-Specific Occupation of Class II Transactivator Promoter III. J.Immunol. 2002, 168: 763-770.
  99. tructrucSartoris S, Tosi G, De Lerma B, Cestari T, Accolla RS: Active suppression of the class II transactivator-encoding AIR-1 locus is responsible for the lack of major histocompatibility complex class II gene expression observed during differentiation from B cells to plasma cells. Eur.J.Immunol. 1996, 26: 2456-2460.
  100. tructrucMurphy SP and Tomasi TB: Absence of MHC class II antigen expression in trophoblast cells results from a lack of class II transactivator (CIITA) gene expression. Mol.Reprod.Dev. 1998, 51: 1-12.
  101. tructrucLee YJ, Han Y, Lu HT, Nguyen V, Qin H, Howe PH, Hocevar BA, Boss JM, Ransohoff RM, Benveniste EN: TGF-beta suppresses IFN-gamma induction of class II MHC gene expression by inhibiting class II transactivator messenger RNA expression . J.Immunol. 1997, 158: 2065-2075.
  102. tructrucNandan D and Reiner NE: TGF-beta attenuates the class II transactivator and reveals an accessory pathway of IFN-gamma action. J.Immunol. 1997, 158: 1095-1101.
  103. tructrucPiskurich JF, Wang Y, Linhoff, MW, White LC, Ting JP: Identification of distinct regions of 5' flanking DNA that mediate constitutive, IFN-gamma, STAT1, and TGF-beta-regulated expression of the class II transactivator gene. J.Immunol. 1998, 160: 233-240.
  104. tructrucO'Keefe GM, Nguyen VT, Benveniste EN: Class II transactivator and class II MHC gene expression in microglia: modulation by the cytokines TGF-beta, IL-4, IL-13 and IL-10. Eur.J.Immunol. 1999, 29: 1275-1285.
  105. tructrucRohn W, Tang LP, Dong Y, Benveniste EN: IL-1 beta inhibits IFN-gamma-induced class II MHC expression by suppressing transcription of the class II transactivator gene. J.Immunol. 1999, 162: 886-896.
  106. tructrucMiller DM, Rahill BM, Boss JM, Lairmore MD, Durbin JE, Waldman JW, Sedmak DD: Human cytomegalovirus inhibits major histocompatibility complex class II expression by disruption of the Jak/Stat pathway. J.Exp.Med. 1998, 187: 675-683.
  107. tructrucWojciechowski W, DeSanctis J, Skamene E, Radzioch D: Attenuation of MHC class II expression in macrophages infected with Mycobacterium bovis bacillus Calmette-Guerin involves class II transactivator and depends on the Nramp1 gene. J.Immunol. 1999, 163: 2688-2696.
  108. tructrucZhong G, Fan T, Liu L: Chlamydia inhibits interferon gamma-inducible major histocompatibility complex class II expression by degradation of upstream stimulatory factor 1. J.Exp.Med. 1999, 189: 1931-1938.
  109. tructrucAbendroth A, Slobedman B, Lee E, Mellins E, Wallace M, Arvin AM: Modulation of major histocompatibility class II protein expression by varicella-zoster virus. J.Virol. 2000, 74: 1900-1907.
  110. tructrucDong Y, Rohn WM, Benveniste EN: IFN-gamma regulation of the type IV class II transactivator promoter in astrocytes. J.Immunol. 1999, 162: 4731-4739.
  111. tructrucHan Y, Zhou ZH, Ransohoff RM: TNF-alpha suppresses IFN-gamma-induced MHC class II expression in HT1080 cells by destabilizing class II trans-activator mRNA. J.Immunol. 1999, 163: 1435-1440.
  112. tructrucNikcevich KM, Piskurich JF, Hellendall RP, Wang Y, Ting JP: Differential selectivity of CIITA promoter activation by IFN-gamma and IRF-1 in astrocytes and macrophages: CIITA promoter activation is not affected by TNF-alpha. J.Neuroimmunol. 1999, 99: 195-204.
  113. tructrucLi G, Harton JA, Zhu X, Ting JP: Downregulation of CIITA function by protein kinase a (PKA)-mediated phosphorylation: mechanism of prostaglandin E, cyclic AMP, and PKA inhibition of class II major histocompatibility complex expression in monocytic lines. Mol.Cell Biol. 2001, 21: 4626-4635.
  114. tructrucDouhan J, Lieberson R, Knoll JHM, Zhou H, Glimcher LH: An isotype-specific activator of major histocompatibility complex (MHC) class II genes that is independent of class II transactivator. J.Exp.Med. 1997, 185: 1885-1895.
  115. tructrucAccolla RS, Jotterand-Bellomo M, Scarpellino L, Maffei A, Carra G, Guardiola J: aIr-1, a newly found locus on mouse chromosome 16 encoding a trans-acting activator factor for MHC class II gene expression. J.Exp.Med. 1986, 164: 369-374.
  116. tructrucMuhlethaler-Mottet A, Otten LA, Steimle V, Mach B: Expression of MHC class II molecules in different cellular and functional compartments is controlled by differential usage of multiple promoters of the transactivator CIITA. EMBO J. 1997, 16: 2851-2860.
  117. tructrucNickerson K, Sisk TJ, Inohara N, Yee CS, Kennell J, Cho MC, Yannie PJ, Nunez G, Chang CH: Dendritic cell-specific MHC class II transactivator contains a caspase recruitment domain that confers potent transactivation activity. J.Biol.Chem. 2001, 276: 19089-19093.
  118. tructrucHofmann K, Bucher P, Tschopp J: The CARD domain: a new apoptotic signalling motif. Trends Biochem.Sci. 1997, 22: 155-156.
  119. tructrucInohara N, Koseki T, del Peso L, Hu Y, Yee C, Chen S, Carrio R, Merino J, Liu D, Ni J, Nunez G. Nod1, an Apaf-1-like activator of caspase-9 and nuclear factor-kappaB. J.Biol.Chem. 1999, 274: 14560-14567.
  120. tructrucSuter T, Malipiero U, Otten L, Ludewig B, Muelethaler-Mottet A, Mach B, Reith W, Fontana A: Dendritic cells and differential usage of the MHC class II transactivator promoters in the central nervous system in experimental autoimmune encephalitis. Eur.J.Immunol. 2000, 30: 794-802.
  121. *tructrucLandmann S, Muhlethaler-Mottet A, Bernasconi L, Suter T, Waldburger JM, Masternak K,Arrighi JF, Hauser C, Fontana A, Reith W: Maturation of dendritic cells is accompanied by rapid transcriptional silencing of class ii transactivator (ciita) expression. J.Exp.Med. 2001, 194: 379-392.
    This paper represents a detailed analysis of the expression and promoter usage of the CIITA gene in immature DCs and during their maturation.
  122. **tructrucWaldburger JM, Suter T, Fontana A, Acha-Orbea H, Reith W: Selective abrogation of major histocompatibility complex class II expression on extrahematopoietic cells in mice lacking promoter IV of the class II transactivator gene. J.Exp.Med. 2001, 194: 393-406.
    Mice lacking CIITA promoter IV exhibit selective abrogation of IFN-g induced MHCII expression on non-bone marrow-derived cells, but not on bone marrow-derived cells. Constitutive MHCII expression on cTECs, and thus positive selection of CD4+ T cells in the thymus are also abolished in these mice.
  123. tructrucLennon AM, Ottone C, Rigaud G, Deaven LL, Longmire J, Fellous M, Bono R, Alcaide-Loridan C: Isolation of a B-cell-specific promoter for the human class II transactivator. Immunogenetics 1997, 45: 266-273.
  124. tructrucGoodwin BL, Xi H, Tejiram R, Eason DD, Ghosh N, Wright KL, Nagarajan U, Boss JM, Blanck G: Varying functions of specific major histocompatibility class II transactivator promoter III and IV elements in melanoma cell lines. Cell Growth Differ. 2001, 12: 327-335.
  125. tructrucDeffrennes V, Vedrenne J, Stolzenberg MC, Piskurich JF, Barbieri G, Ting JP, Charron D, Alcaide-Loridan C: Constitutive expression of MHC class II genes in melanoma cell lines results from the transcription of class II transactivator abnormally initiated from its B cell-specific promoter. J.Immunol. 2001, 167: 98-106.
  126. tructrucGhosh N, Piskurich JF, Wright G, Hassani K, Ting JP, Wright KL: A novel element and a TEF-2-like element activate the major histocompatibility complex class II transactivator in B-lymphocytes. J.Biol.Chem. 1999, 274: 32342-32350.
  127. tructrucPiskurich JF, Linhoff MW, Wang Y, Ting JP: Two distinct gamma interferon-inducible promoters of the major histocompatibility complex class II transactivator gene are differentially regulated by STAT1, interferon regulatory factor 1, and transforming growth factor beta. Mol.Cell Biol. 1999, 19: 431-440.
  128. tructrucKeller AD and Maniatis T: Identification and characterization of a novel repressor of beta- interferon gene expression. Genes Dev. 1991, 5: 868-879.
  129. tructrucKakkis E, Riggs KJ, Gillespie W, Calame K: A transcriptional repressor of c-myc. Nature 1989, 339: 718-721.
  130. *tructrucPiskurich JF, Lin KI, Lin Y, Wang Y, Ting JP, Calame K: BLIMP-I mediates extinction of major histocompatibility class II transactivator expression in plasma cells. Nat.Immunol. 2000, 1: 526-532.
  131. *tructrucGhosh N, Gyory I, Wright G, Wood J, Wright KL: Positive regulatory domain I binding factor 1 silences class II transactivator expression in multiple myeloma cells. J.Biol.Chem. 2001, 276: 15264-15268.
    MHCII expression is silenced during differentiation of B cells into plasma cells. Together with reference 130, this paper suggests that PRDI-BF1 / Blimp-1, a factor induced during differentiation into plasma cells, can repress expression of the CIITA gene.
  132. tructrucTurner CA Jr., Mack DH, Davis MM: Blimp-1, a novel zinc finger-containing protein that can drive the maturation of B lymphocytes into immunoglobulin-secreting cells. Cell 1994, 77: 297-306.
  133. tructrucMuhlethaler-Mottet A, Di Berardino W, Otten LA, Mach B: Activation of the MHC class II transactivator CIITA by interferon-g requires cooperative interaction between Stat1 and USF-1. Immunity 1998, 8: 157-166.
  134. tructrucO'Keefe GM, Nguyen VT, Ping Tang LL, Benveniste EN: IFN-gamma regulation of class II transactivator promoter IV in macrophages and microglia: involvement of the suppressors of cytokine signaling-1 protein. J.Immunol. 2001, 166: 2260-2269.
  135. tructrucXi H, Eason DD, Ghosh D, Dovhey S, Wright KL, Blanck G: Co-occupancy of the interferon regulatory element of the class II transactivator (CIITA) type IV promoter by interferon regulatory factors 1 and 2. Oncogene 1999, 18: 5889-5903.
  136. tructrucXi H, Goodwin B, Shepherd AT, Blanck G: Impaired class II transactivator expression in mice lacking interferon regulatory factor-2. Oncogene 2001, 20: 4219-4227.
  137. tructrucPanek RB, Lee YJ, Benveniste EN: TGF-beta suppression of IFN-gamma-induced class II MHC gene expression does not involve inhibition of phosphorylation of JAK1, JAK2, or signal transducers and activators of transcription, or modification of IFN-gamma enhanced factor X expression. J.Immunol. 1995, 154: 610-619.
  138. tructrucDong Y, Tang L, Letterio JJ, Benveniste EN: The Smad3 protein is involved in TGF-beta inhibition of class II transactivator and class II MHC expression. J.Immunol. 2001, 167: 311-319.
  139. tructrucNavarrete SA, Kehlen A, Schutte W, Langner J, Riemann D: Regulation by transforming growth factor-beta1 of class II mRNA and protein expression in fibroblast-like synoviocytes from patients with rheumatoid arthritis. Int Immunol. 1998, 10: 601-607.
  140. tructrucEndo TA, Masuhara M, Yokouchi M, Suzuki R, Sakamoto H, Mitsui K, Matsumoto A, Tanimura S, Ohtsubo M, Misawa H, Miyazaki T, Leonor N, Taniguchi T, Fujita T, Kanakura Y, Komiya S, Yoshimura A: A new protein containing an SH2 domain that inhibits JAK kinases. Nature 1997, 387: 921-924.
  141. tructrucYasukawa H, Misawa H, Sakamoto H, Masuhara M, Sasaki A, Wakioka T, Ohtsuka S, Imaizumi T, Matsuda T, Ihle JN, Yoshimura A: The JAK-binding protein JAB inhibits Janus tyrosine kinase activity through binding in the activation loop. EMBO J. 1999, 18: 1309-1320.
  142. tructrucMorris AC, Spangler WE, Boss JM: Methylation of class II trans-activator promoter IV: a novel mechanism of MHC class II gene control. J.Immunol. 2000, 164: 4143-4149.
  143. tructrucvan den Elsen PJ, van der Stoep N, Vietor HE, Wilson L, van Zutphen M, Gobin SJ: Lack of CIITA expression is central to the absence of antigen presentation functions of trophoblast cells and is caused by methylation of the IFN-gamma inducible promoter (PIV) of CIITA. Hum.Immunol. 2000, 61: 850-862.
  144. tructrucAnderson G, Jenkinson EJ, Moore NC, Owen JJ: MHC class II-positive epithelium and mesenchyme cells are both required for T-cell development in the thymus. Nature 1993, 362: 70-73.
  145. tructrucHoney K and Rudensky A: The pIV-otal class II transactivator promoter regulates major histocompatibility complex class II expression in the thymus. J.Exp.Med. 2001, 194: F15-F18.
  146. tructrucErickson FL, Dinesh-Kumar SP, Holzberg S, Ustach CV, Dutton M, Handley V, Corr C, Baker BJ: Interactions between tobacco mosaic virus and the tobacco N gene. Philos.Trans.R.Soc.Lond.B.Biol.Sci. 1999, 354: 653-658.
  147. **tructrucBeresford GW and Boss JM: CIITA coordinates multiple histone acetylation modifications at the HLA- DRA promoter. Nat.Immunol. 2001, 2: 652-657.
    This in vivo study shows that acetylation and deacetylation of histone H3 and H4 at the HLA-DR promoter occurs with a time course that depends on CIITA expression and binding.
  148. tructrucScholl T, Mahanta SK, Strominger JL: Specific complex formation between the type II bare lymphocyte syndrome- associated transactivators CIITA and RFX5. Proc.Natl.Acad.Sci.USA 1997, 94: 6330-6334.
  149. tructrucHake SB, Masternak K, Kammerbauer C, Janzen C, Reith W, Steimle V: CIITA leucine-rich repeats control nuclear localization, In vivo recruitment to the major histocompatibility complex (MHC) class II enhanceosome, and MHC class II gene transactivation. Mol.Cell Biol. 2000, 20: 7716-7725.
  150. tructrucRiley JL, Westerheide SD, Price JA, Brown JA, Boss JM: Activation of class II MHC genes requires both the X box and the class II transactivator (CIITA). Immunity 1995, 2: 533-543.
  151. tructrucZhou H and Glimcher LH: Human MHC class II gene transcription directed by the carboxyl terminus of CIITA, one of the defective genes in type II MHC combined immune deficiency. Immunity 1995, 2: 545-553.
  152. tructrucChin KC, Li GG, Ting JP: Importance of acidic, proline/serine/threonine-rich, and GTP- binding regions in the major histocompatibility complex class II transactivator: generation of transdominant-negative mutants. Proc.Natl.Acad.Sci.USA 1997, 94: 2501-2506.
  153. tructrucBontron S, Ucla C, Mach B, Steimle V: Efficient Repression of Endogenous Major Histocompatibility Complex Class II Expression through Dominant Negative CIITA Mutants Isolated by a Functional Selection Strategy. Mol.Cell.Biol. 1997, 17: 4249-4258.
  154. tructrucFontes JD, Jiang B, Peterlin BM: The class II trans- activator CIITA interacts with the TBP- associated factor TAF II 32. Nucleic Acids Res. 1997, 25: 2522-2528.
  155. tructrucYun S, Gustafsson K, Fabre JW: Suppression of MHC class II expression by human class II trans- activator constructs lacking the N-terminal domain. Int Immunol. 1997, 9: 1545-1553.
  156. tructrucZhou H, Su HS, Zhang X, Douhan J 3rd, Glimcher LH: CIITA-dependent and -independent class II MHC expression revealed by a dominant negative mutant. J.Immunol. 1997, 158: 4741-4749.
  157. tructrucMahanta SK, Scholl T, Yang FC, Strominger JL: Transactivation by CIITA, the type II bare lymphocyte syndrome- associated factor, requires participation of multiple regions of the TATA box binding protein. Proc.Natl.Acad.Sci.U.S.A. 1997, 94: 6324-6329.
  158. tructrucFontes JD, Kanazawa S, Jean D, Peterlin BM: Interactions between the class II transactivator and CREB binding protein increase transcription of major histocompatibility complex class II genes. Mol.Cell Biol. 1999, 19: 941-947.
  159. tructrucKanazawa S, Okamoto T, Peterlin BM: Tat competes with CIITA for the binding to P-TEFb and blocks the expression of MHC class II genes in HIV infection. Immunity 2000, 12: 61-70.
  160. tructrucSpilianakis C, Papamatheakis J, Kretsovali A: Acetylation by PCAF enhances CIITA nuclear accumulation and transactivation of major histocompatibility complex class II genes. Mol.Cell Biol. 2000, 20: 8489-8498.
  161. tructrucKretsovali A, Agalioti T, Spilianakis C, Tzortzakaki E, Merika M, Papamatheakis J: Involvement of CREB binding protein in expression of major histocompatibility complex class II genes via interaction with the class II transactivator. Mol.Cell Biol. 1998, 18: 6777-6783.
  162. tructrucHarton JA, Zika E, Ting JP: The histone acetyltransferase domains of CBP and pCAF are not necessary for cooperativity with CIITA. J.Biol.Chem. 2001, 276: 38715-38720
  163. *tructrucZhu XS and Ting JP: A 36-Amino-Acid Region of CIITA Is an Effective Inhibitor of CBP: Novel Mechanism of Gamma Interferon-Mediated Suppression of Collagen alpha(2)(I) and Other Promoters. Mol.Cell Biol. 2001, 21: 7078-7088.
    Together with references 179, 180 and 181, this paper suggests that CIITA may inhibit the expression of certain genes by sequestering the general coactivator CBP.
  164. tructrucLee TI and Young RA: Transcription of eukaryotic protein-coding genes. Annu.Rev.Genet. 2000, 34: 77-137.
  165. tructrucStrahl BD and Allis CD: The language of covalent histone modifications. Nature 2000, 403: 41-45.
  166. tructrucRoth SY, Denu JM, Allis CD: Histone acetyltransferases. Annu.Rev.Biochem. 2001, 70: 81-120.
  167. *tructrucRaval A, Howcroft TK, Weissman JD, Kirshner S, Zhu XS, Yokoyama K, Ting JP, Singer DS: Transcriptional coactivator, CIITA, is an acetyltransferase that bypasses a promoter requirement for TAF(II)250. Mol.Cell 2001, 7: 105-115.
    This study suggests that CIITA has an endogenous acteyltransferase activity.
  168. tructrucHarton JA, Cressman DE, Chin KC, Der CJ, Ting JP: GTP binding by class II transactivator: role in nuclear import. Science 1999, 285: 1402-1405.
  169. tructrucChin KC, Li G, Ting JP: Activation and transdominant suppression of MHC class II and HLA-DMB promoters by a series of C-terminal class II transactivator deletion mutants. J.Immunol. 1997, 159: 2789-2794.
  170. **tructrucKretsovali A, Spilianakis C, Dimakopoulos A, Makatounakis T, Papamatheakis J: Self-association of class II transactivator correlates with its intracellular localization and transactivation. J.Biol.Chem. 2001, 276: 32191-32197.
  171. **tructrucLinhoff MW, Harton JA, Cressman DE, Martin BK, Ting JP: Two distinct domains within CIITA mediate self-association: involvement of the GTP-binding and leucine-rich repeat domains. Mol.Cell Biol. 2001, 21: 3001-3011.
  172. **tructrucSisk TJ, Roys S, Chang CH: Self-association of CIITA and its transactivation potential. Mol.Cell Biol. 2001, 21: 4919-4928.
    **References 170 to 172 all demonstrate that CIITA requires self-association for its function, although there is some controversy concerning the precise domains that are involved.
  173. tructrucCressman DE, Chin KC, Taxman DJ, Ting JP: A defect in the nuclear translocation of CIITA causes a form of type II bare lymphocyte syndrome. Immunity 1999, 10: 163-171.
  174. tructrucCressman DE, O'Connor WJ, Greer SF, Zhu XS, Ting JP: Mechanisms of nuclear import and export that control the subcellular localization of class ii transactivator. J.Immunol. 2001, 167: 3626-3634.
  175. tructrucTowey M and Kelly AP: Nuclear localisation of CIITA is controlled by a carboxy terminal leucine-rich repeat region. Mol.Immunol. 2002, 38: 627-634.
  176. tructrucGobin SJ, Peijnenburg A, Keijsers V, van den Elsen PJ: Site alpha is crucial for two routes of IFN gamma-induced MHC class I transactivation: the ISRE-mediated route and a novel pathway involving CIITA. Immunity 1997, 6: 601-611.
  177. tructrucMartin BK, Chin KC, Olsen JC, Skinner CA, Dey A, Ozato K, Ting JP: Induction of MHC class I expression by the MHC class II transactivator CIITA. Immunity 1997, 6: 591-600.
  178. tructrucGobin SJ, Peijnenburg A, Van Eggermond M, van Zutphen M, van den Berg R, van den Elsen PJ: The RFX complex is crucial for the constitutive and CIITA-mediated transactivation of MHC class I and beta2-microglobulin genes. Immunity 1998, 9: 531-541.
  179. tructrucGourley T, Roys S, Lukacs NW, Kunkel SL, Flavell RA, Chang CH: A novel role for the major histocompatibility complex class II transactivator CIITA in the repression of IL-4 production. Immunity 1999, 10: 377-386.
  180. tructrucGourley TS and Chang CH: Cutting edge: the class II transactivator prevents activation-induced cell death by inhibiting Fas ligand gene expression. J.Immunol. 2001, 166: 2917-2921.
  181. tructrucSisk TJ, Gourley T, Roys S, Chang CH: MHC class II transactivator inhibits IL-4 gene transcription by competing with NF-AT to bind the coactivator CREB binding protein (CBP)/p300. J.Immunol. 2000, 165: 2511-2517.
  182. tructrucSaifuddin M, Roebuck KA, Chang C, Ting JP, Spear GT: Cutting edge: activation of HIV-1 transcription by the MHC class II transactivator. J.Immunol. 2000, 164: 3941-3945.
  183. tructrucSaifuddin M, Spear GT, Chang C, Roebuck KA: Expression of MHC class II in T cells is associated with increased HIV-1 expression. Clin.Exp.Immunol. 2000, 121: 324-331.
  184. tructrucOkamoto H, Asamitsu K, Nishimura H, Kamatani N, Okamoto T: Reciprocal modulation of transcriptional activities between HIV-1 Tat and MHC class II transactivator CIITA. Biochem.Biophys.Res.Commun. 2000, 279 :494-499.
  185. **tructrucAccolla RS, De Lerma BA, Mazza A, Casoli C, De Maria A, Tosi G: The MHC class II transactivator: prey and hunter in infectious diseases. Trends Immunol. 2001, 22: 560-563.
    This review discusses a series of results (references 159, 182, 184 and 186) addressing the interaction between CIITA and pathogens such as HIV.
  186. tructrucTosi G, De Lerma BA, D'Agostino A, Valle MT, Megiovanni AM, Manca F, Caputo A, Barbanti-Brodano G, Accolla RS: HIV-1 Tat mutants in the cysteine-rich region downregulate HLA class II expression in T lymphocytic and macrophage cell lines. Eur.J.Immunol. 2000, 30: 19-28.
  187. tructrucMudhasani R and Fontes JD: Inhibition of class II trans-activator function by HIV-1 tat in mouse cells is independent of competition for binding to cyclin T1. Mol.Immunol. 2002, 38: 539-546.
  188. tructrucLe Roy E, Muhlethaler-Mottet A, Davrinche C, Mach B, Davignon JL: Escape of human cytomegalovirus from HLA-DR-restricted CD4(+) T-cell response is mediated by repression of gamma interferon-induced class II transactivator expression. J.Virol. 1999, 73: 6582-6589.

2.3 Dendritic cells


2.3.1 The role of dendritic cells in the immune system

      Host defense relies on a concerted action of both antigen-nonspecific innate immunity and antigen-specific adaptive immunity (140, 141). Key features of the innate immune system include the ability to rapidly recognize pathogens and/or tissue injury and the ability to signal the presence of danger to cells of the adaptive immune system (142). The innate system includes phagocytic cells, natural killer (NK) cells, complement and interferons (IFNs). Cells of the innate system use a variety of germline-encoded pattern recognition receptors to recognize patterns shared between pathogens, for instance bacterial LPS, carbohydrates, and double-stranded viral RNA (143-145). Evolutionary pressure has led to the development of adaptive immunity, the key features of which are the ability to rearrange genes of the immunoglobulin family, permitting creation of a large diversity of antigen-specific clones and immunological memory. Yet this highly sophisticated and potent system needs to be instructed and regulated by APCs. DCs are unique APCs because they are the only ones that are able to induce primary immune responses, thus permitting establishment of immunological memory (98-100).

      DCs represent a heterogeneous cell population, residing in most peripheral tissues, particularly at sites of interface with the environment (skin, mucosae) (98). They display a high phagocytic capacity. Following tissue damage, DCs process the captured antigens, load them on MHC molecules and migrate to the lymphoid organs, where they present the antigenic peptides in the context of MHC molecules to T cells, thereby initiating adaptive immune responses (146, 147). DCs activate antigen-specific CD4+ T helper cells, which in turn regulate the immune effectors, including antigen-specific CD8+ cytotoxic T cells (CTLs) and B cells, as well as non-antigen-specific macrophages, eosinophils and NK cells. Moreover, DCs educate effector cells to home to the site of injury (98). Four stages of DC development have been delineated, including 1) bone marrow progenitors; 2) precursor DCs that are patrolling through blood and lymphatics as well as lymphoid tissues, and that upon pathogen recognition, release large amounts of cytokines, e.g. IFN-a, thereby limiting the spread of infection; 3) tissue-residing immature DCs, which possess high endocytic and phagocytic capacity permitting antigen capture; and 4) mature DCs, present within secondary lymphoid organs, that express high levels of costimulatory molecules permitting antigen presentation (Fig.11).

      

Fig. 11 The life cycle of DCs.

      Circulating precursor DCs enter tissues as immature DCs. They may directly encounter pathogens that induce secretion of cytokines (e.g. IFN-a). Immature DCs reside at strategically important sites in the periphery to encounter pathogens. After antigen capture, immature DCs migrate to lymphoid organs where, after maturation, they display peptide-MHC complexes, which allows selection of rare circulating antigen-specific lymphocytes. These activated T cells help DCs in terminal maturation, which allows lymphocyte expansion and differentiation. Activated T lymphocytes migrate and reach the injured tissues. Helper T cells secrete cytokines, which permit activation of macrophages, NK cells and eosinophils. Cytotoxic T cells eventually lyse the infected cells. B cells become activated after contact with T cells and DCs and then mature into antibody-producing plasma cells. It is believed that, after interaction with lymphocytes, DCs die by apoptosis.


2.3.2 Immature dendritic cells

      Immature DCs reside in peripheral tissues at sentinel positions where they sample self and non-self antigens. However, they are not able to present the antigens. Immature DCs accumulate MHC molecules intracellularly and present only a small fraction at the cell surface (103, 104).

      Three types of antigen uptake are known: macropinocytosis, phagocytosis and clathrin-mediated endocytosis (101, 102). Different types of antigens are internalized via these different routes. The constitutive and cytoskeleton-dependent process of macropinocytosis allows rapid and non-specific sampling of large amounts of surrounding fluid and results in the formation of large intracellular vacuoles. Phagocytosis is a receptor-mediated process dependent on regulated actin assembly. In general, the same receptors mediating phagocytosis of pathogens are also engaged in clathrin-dependent endocytosis. The latter allows uptake of macromolecules through specialized regions of the plasma membrane, termed coated pits (148). A large number of endocytic receptors are expressed on immature DCs, namely C-type lectins (101, 149, 150), receptors for the Fc portion of immunoglobulins (FcRs) (151, 152), complement receptors (153), receptors for heat shock proteins (154, 155), and scavenger receptors (156).


2.3.3 Dendritic cell activation and maturation

      A signal from pathogens, often referred to as a danger signal (157), induces DCs to enter a developmental program, called maturation, which transforms DCs from sentinels into efficient APCs and T cell stimulators (98). The danger signal can be a bacterial or viral product as well as an inflammatory cytokine. Danger signals are recognized through specific pattern-recognition receptors, such as Toll-like receptors, FcRs and cytokine receptors (151, 158).

      With the induction of maturation, DCs undergo massive morphological and functional changes. Antigen internalization is downmodulated due to repression of most antigen-receptors and to an overall reduction in phagocytosis and macropinocytosis (101). In addition, MHC molecules are redistributed from intracellular endocytic compartments to the cell surface (103, 104). Peptide loading as well as the half-life of MHC molecules is increased (103, 104). Finally, the surface expression of T cell costimulatory molecules also rises.

      Concomitant with the modifications of their antigen presentation abilities, maturation induces migration of DCs out of peripheral tissues (105). Modifications in the expression of chemokine receptors and adhesion molecules, as well as profound changes of the cytoskeleton organization, contribute to the migration of DCs, through the lymph, towards secondary lymphoid organs (106). By linking antigen uptake, peptide loading, and cell migration to the encounter of a danger signal, DCs restrict antigen presentation to those antigens internalized during maturation, thus favoring the stimulation of T cells specific for potentially pathogenic antigens.


2.3.4 Antigen presentation by dendritic cells

      CD8+ and CD4+ T cells express clonally distributed TCRs that recognize antigenic peptides associated with MHCI and MHCII molecules, respectively. A strict compartmentalization of MHCI and MHCII biogenesis results in the loading of exogenous antigens on MHCII molecules in the endocytic pathway, and the selective loading of endogenous antigens on MHCI molecules in the ER. This model accounts for the selective killing by MHCI-restricted CD8+ T cells of virus-infected cells (expressing endogenous viral antigens), but not of neighboring cells that have internalized inactive virus or apoptotic cells.

      Peptides to be loaded onto MHCI molecules are generated by proteasome-mediated degradation of newly synthesized ubiquitinylated proteins (Fig.12). The immunoproteasome, which is expressed at low levels in immature DCs, becomes the main type of proteasome in mature DCs (107-109). The peptides resulting from proteasome mediated degradation are transferred to the ER by a specialized peptide transporter, TAP, and loaded onto newly synthesized MHCI molecules under the control of several ER resident chaperones (159). MHCI-peptide complexes are then rapidly transferred through the Golgi apparatus to the plasma membrane. MHCI synthesis and half-life are slightly increased during DC maturation (103, 104).

      Fig. 12 MHCI restricted antigen presentation.

      

      Pathogen-derived or self-proteins within the cytosol of APCs are enzymatically digested into peptides, mainly by the cytosolic protease called proteasome, and are then transported by TAP molecules into the ER. In the ER lumen, peptides bind to MHCI molecules, which are subsequently transported via the Golgi to the plasma membrane.

      The formation of MHCII-peptide complexes in DCs is induced by maturation. This is different from the situation in other APCs, where this process is constitutive. In immature DCs, most MHCII molecules are retained in late-endosomal and lysosomal compartments (Fig.13). They are persistently associated with the invariant chain Ii, which is poorly degraded in these cells (110, 111) and thus prevents peptide loading. However, upon induction of maturation, Ii is degraded by cathepsin S and MHCII-peptide complexes are formed and transported to the surface (110). Additional mechanisms for the regulation of MHCII-peptide complex formation and presentation exist in DCs. First, internalized antigens are processed into peptides only once the cell starts to mature (112). Second, de novo MHCII synthesis is transiently upregulated during DC maturation (103, 104). Finally, the reduction in endocytosis slows down the recycling of the MHCII molecules at the cell surface and thus contributes to their increased half-life.

      

Fig. 13 MHCII restricted antigen presentation.

      Exogenous antigens are derived from proteins that are endocytosed and processed by proteases. Newly synthesized MHCII molecules are retained in lysosomal compartments. Peptide loading occurs upon DC maturation in specialized antigen-processing vesicles, MIIC. The peptide-MHCII complexes are then externalized to the plasma membrane.

      In contrast to the model discussed above, exogenous antigens can also be presented by MHCI molecules, thereby leading to CD8+ T cell activation (114-116). This phenomenon is called cross-presentation. It is crucial for the initiation of cytotoxic immune responses against antigens that are not synthesized by the DCs themselves, but by other cells. Normally, antigen-bearing cells do not directly initiate immune responses, since they are not capable of stimulating resting naive T cells. Most immune responses thus require antigen transfer to professional APCs, most probably DCs (160, 161), which then stimulate naive T cells. Apoptotic cells seem to be favorite targets for cross-presentation (161-163). Two intracellular pathways can result in cross-presentation: TAP-independent endocytotic or TAP-dependent ER MHCI peptide loading (164) (Fig.14). The latter implicates the existence of an endosome-to-cytosol transport system (165).

      

Fig. 14 Cross presentation.

      a) Exogenous antigens can access MHCI molecules within endosomes that contain MHCI molecules that have been recycled from the membrane. b) Alternatively, antigen is transported from the endosome to the cytosol and processed similarly to endogenously derived antigens.


2.3.5 T cell stimulation and tolerization by dendritic cells

      DCs have the unique ability to prime naïve CD4+ and CD8+ T cells. This has been shown in in vivo, in vitro and in situ experiments (166-175). In most cases, CD4+ T cell help is required for efficient CTL activation. In traditional models of CTL activation, the CD4+ and CD8+ T cells were thought to recognize antigen on the same APC. However, in the current model (176-178), the APCs are licensed to activate CTLs by T helper cells via upregulation of CD40 ligand (CD40L) on the DCs. Thus, a conditioned DC becomes a temporal bridge between a CD4+ T helper cell and a CTL. In addition to priming, DCs appear essential to maintain survival of naive CD4+ T cells (123) and immune T cell memory (179). Importantly, DCs are also involved in the tolerization of the T cell repertoire to self-antigen. This occurs in the thymus (central tolerance) by deletion of developing T cells (123), and in lymphoid organs (peripheral tolerance) probably by the induction of anergy or deletion of mature T cells.

      DCs play apparently contradictory roles by stimulating as well as tolerizing T cells. This inconsistency may be attributed either to distinct DC lineages endowed with unique T cell stimulatory capacity or to a single DC type, which is instructed by environmental stimuli to perform different functions. Indeed, two major subtypes of DCs, that exhibit different functions, exist: myeloid and lymphoid (98). In addition, the different DC functions can be altered by the cytokine environment.


3. Materials and methods


3.1 Preparation of monocytes and dendritic cells

      PBMCs were prepared from buffy coat fractions (Blood Transfusion Center, Geneva) by isolation over Ficoll-Paque (Pharmacia). Monocytes were obtained by adhesion of PBMCs in 10 9cm tissue culture plates in RPMI 2% FCS, 100 U/ml penicillin, 0.1 mg/ml streptomycin, 2 mM L-glutamine at 37°C for 1 hour. The non-adhering cells were removed by 4 washes and monocytes were incubated over night in RPMI 2% FCS, 100 U/ml penicillin, 0.1 mg/ml streptomycin, 2 mM L-glutamine at 37°C. Contaminating B cells were depleted using anti-CD19 antibody coupled magnetic beads (Dynal). The monocytes, > 90% pure as assessed by flow cytometry, were either cultured at 2x106 cells/ml in RPMI 10% FCS, 100 U/ml penicillin, 0.1 mg/ml streptomycin, 2 mM L-glutamine supplemented with 2-mercaptoethanol (50 mM, Gibco) and 100 U/ml M-CSF (R&D Systems) or differentiated into DCs. For differentiation into DCs, the monocytes were cultured at 1,2x106 cells/ml in RPMI 10% FCS, 100 U/ml penicillin, 0.1 mg/ml streptomycin, 2 mM L-glutamine supplemented with 2-mercaptoethanol (50 mM), GM-CSF (600 U/ml, Leukomax, Essex Chemie AG) and IL-4 (750 U/ml, R&D Systems). Every 2 days, half of the medium was replaced by fresh medium containing a two-fold concentration of cytokines. After 6 days of differentiation, the cells exhibited an immature DC phenotype (CD14-, CD1a+, MHCIIlow, CD86low, CD40low, CD80-, CD83-).

      DC maturation was induced by stimulating the immature DCs with LPS (from Salmonella abortus equi, 10 ng/ml, Sigma) for 24 to 48 hours. Mature DCs can be identified by high levels of surface MHCII, costimulatory molecules (CD86, CD80 and CD40) and the DC-restricted marker CD83. In certain experiments, DCs were stimulated with TNF-a (75 ng/ml, R&D systems), IFN-a (1000 U/ml, kindly provided by C. Weissmann, Zürich) or IFN-g (1000 U/ml, Life Technologies). For stimulation with CD40L, immature DCs were co-cultured at a ratio of 1:5 with CD40L-expressing J558L cells (a gift from E. Padovan, Basel). For infection with Salmonella typhimurium (strain 14028 phoQ24), 5x106 cfu were added per 2.5x105 immature DCs and the mixture was incubated for 30 minutes at 37°C. Extracellular bacteria were then killed by the addition of Gentamycin (50 mg/ml). For infection with Sendai (strain M, SeVM), the virus was added to immature DCs at a multiplicity of infection (MOI) of 20.

      Trichostatin A (TSA, Sigma) was co-administered with LPS at 100 ng/ml.


3.2 Transduction of dendritic cells with lentiviral vectors

      Production of the HIV-derived vectors pseudotyped with the vesicular stomatitis virus (VSV)-G envelope was achieved by transient co-transfection of three plasmids (see below) into the 293T cell line using calcium phosphate DNA precipitation. The supernatants containing viral particles were collected 48 hours after transfection, filtered through a 45 mm pore size filter, and concentrated by ultracentrifugation at 26'000 rpm for 90 minutes. The pellet was resuspended in 1/100 volume of DMEM. Viral stocks were stored at -70°C and titers determined by transduction and FACS analysis (see below) of GFP expression in HeLa cells. Titers were usually between 5x107 and 2x108 transducing units (TU) per ml. In addition, the integrated vector copies were quantified by real time PCR (see below) on genomic DNA isolated from transduced HeLa cells. A pair of probe and primers specific for the gag gene (present in the vector plasmid) was used together with a pair specific for mitochondrial DNA serving as an internal control. The HIV-derived packaging construct used was pCMVR8.91, which encodes the HIV-1 Gag and Pol precursors as well as the regulatory proteins Tat and Rev (180). VSV-G was expressed from pMD.G. The HIV vector plasmid pWPTS (kindly provided by Didier Trono, Geneva) was a derivative of the original pHR' backbone (181), with the following modifications. Self-inactivating (SIN) vectors contain a deletion in the U3 region of the 3' long terminal repeat (LTR) from nucleotide -418 to -18, removing all the transcriptionally active sequences (182). The woodchuck hepatitis virus posttranscriptional regulatory element (WPRE) was inserted between the GFP gene and the 3' LTR (183). A central polypurine tract was inserted upstream of the transgenic promoter to increase the nuclear import of the proviral DNA. The original pWPTS plasmid contained an EF-1a promoter derived from the pEF-BOS plasmid (184). The EF-1a promoter was either removed in order to create pWPTS-p0 or replaced by various CIITA promoter fragments to create pWPTS-pIII-322 (nucleotides -322 to +101 of pIII), pWPTS-pI-390 (nucleotides -390 to +92 of pI), pWPTS-pI-1042 (nucleotides -1042 to +92 of pI), pWPTS-pI-1987 (nucleotides -1987 to +92 of pI) and pWPTS-pI-2846 (nucleotides -2846 to +92 of pI).

      106 human peripheral blood monocytes were transduced with lentiviral vectors at a MOI between 2 and 10. The following day, fresh medium containing GM-CSF and IL-4 was added and the cells were differentiated into immature DCs during 6 days and eventually induced for maturation with LPS as described above.


3.3 Cell lines

      Raji (Burkitts's lymphoma, EBV positive), RJ2.2.5 (CIITA deficient, derived from Raji), U937 (promonocyte), THP1 (promonocyte) and Me67.8 (melanoma) were grown in RPMI 10% FCS, 100 U/ml penicillin, 0.1 mg/ml streptomycin, 2 mM L-glutamine. HeLa and 293T were grown in DMEM 10% FCS, 100 U/ml penicillin, 0.1 mg/ml streptomycin, 2 mM L-glutamine.

      RJ2.2.5 was transfected with an episomal expression vector containing a CIITA type I cDNA under the control of the SRa promoter. Transfected cells were selected with hygromycin for 10 days and checked for cell surface MHCII expression by FACS.


3.4 Embryonic stem cells

      E14.1 (129Sv) embryonic stem (ES) cells were cultured on gelatinized plates on a layer of irradiated feeder cells (SNL 76/7, (185)) in DMEM 20% FCS, 4 mM L-Glutamine, 1 mM Na-Pyruvate, 0.1 mM non-essential amino acids, 100 mM 2-mercaptoethanol, 50 mg/ml gentamycin, 2000 U/ml leukemia-inhibitory factor (LIF). ES cells were transfected by electroporation (3 mF, 0.8 V) with 40 mg of the linearized targeting plasmid in 137 mM NaCl, 5 mM KCl, 0.7 mM Na2HPO4, 6 mM glucose, 28 mM 2-mercaptoethanol, 20 mM HEPES pH 7.0. G418-resistant clones were selected as described (186). Homologous recombinants were screened by PCR using a primer upstream of the 5' end of the targeting construct (5'-GTGGGGTCAGTGTTGAC-3') and an internal primer lying in the nitroreductase gene (5'-ATCTGCTCGGCCTGTTC-3'). Positive clones were confirmed by Southern blotting using EcoRI in combination with a 5' external and a 3' external probe (Fig.24, p.103). For deletion of the neo gene, one targeted ES cell clone was transfected with 40 mg of an expression plasmid coding for the Flp recombinase under the control of a CMV enhancer/b-actin promoter and the puromycin gene (kindly provided by Francis Stewart, EMBL, Heidelberg). The transfected cells were selected with puromycin (1 mg/ml) for three days. neo-deleted (G418 sensitive) clones were screened by PCR using a primer in the b-galactosidase gene (5'-GCAACTCTGGCTCACAG-3') and a primer downstream of the neo gene (5'-TCCACTGTACACCTGAACAG-3'). Positive clones were confirmed by Southern blotting using PstI in combination with a probe downstream of the neo gene (Fig.25, p.104).


3.5 Cytofluorometry

      250'000 cells were blocked with mouse IgG (2 mg/ml) prior to staining. Anti-human antibodies used were FITC conjugated anti-HLA-DR (clone G46-6, Pharmingen), anti-CD80 (clone BB1, Pharmingen) and anti-CD40 (clone 5C3, Pharmingen); PE conjugated anti-CD1a (clone BL6, Immunotech), anti-CD83 (clone HB15a, Immunotech) and anti-CD86 (clone IT2.2, Pharmingen); and biotinylated anti-CD14 (clone UCHM1, Ancell) followed by Allophycocyanin-conjugated Streptavidin (Pharmingen). Dead cells were excluded by loading with 7-Aminoactinomycin D (Sigma). Staining with isotype matched antibodies was performed in parallel.


3.6 RNA isolation

      Total RNA was prepared with Trizol (Life Technologies). 106 cells were lysed in 0.5 ml Trizol reagent for 5 minutes at room temperature and were eventually stored at -70°C until used. 100 ml chloroform was added. The samples were mixed vigorously and centrifuged for 15 minutes at 12'000g at 4°C. The aqueous phase was transferred to a new tube and precipitated with 250 ml isopropanol and 5 mg glycogen. After centrifugation for 10 minutes at 12'000 g at 4°C, the pellet was washed with 75% ethanol and resuspended in 20 ml TE pH 7.5.


3.7 RNase protection assay

      The RNA probes were generated under standard conditions by incubating 1 mg linearized plasmid with 75 mCi a32P-UTP in a final volume of 20 ml for 30 minutes at 37°C. After DNaseI treatment for 15 minutes at 37°C, the probes were extracted, precipitated with ethanol and resuspended in FA buffer (90% formamide, 1 mM EDTA, xylene cyanol, bromophenol blue). They were purified on a 6% denaturing polypolyacrylamide gel and eluted in 200 ml 0.5 M NH4Ac, 0.1% SDS, 0.2 mg/ml tRNA in TE for 90 minutes at 4°C. The eluates were precipitated and resuspended at 500'000 cpm/ml in 80% formamid / 20% 5x HB (200 mM PIPES pH 6.4, 2 M NaCl, 5 mM EDTA). The probes were as follows:

      
RPA probe description linearization transcription non-digested probe protected probe
hCIITAI nu 107 exon1 -> ; nu 226 exon2 XhoI T7 RNA polymerase 399 nu 333 nu / 226 nu (non-I)
hCIITAIII nu 162 exon1 -> ; nu 226 exon2 EcoRI T7 RNA polymerase 454 nu 388 nu / 226 nu (non-III)
hCIITAIV nu 46 exon1 -> ; nu 226 exon2 XhoI T7 RNA polymerase 350 nu 272 nu / 226 nu (non-IV)
total hCIITA nu 1169 -> ; nu 1361 XhoI T7 RNA polymerase 230 nu 193 nu
HLA-DR nu 182 -> ; nu 491 HaeIII Sp6 RNA polymerase 323 nu 309 nu
hGAPDH nu 67 -> ; nu 166 EcoRI T7 RNA polymerase 139 nu 100 nu

      2-10 mg RNA was mixed with 10 ml 5x HB, lyophilized and resuspended in H2O + total yeast RNA to obtain a final amount of 50 mg total RNA in 10 ml. 40 ml formamide was added, and samples were heated at 85°C for 5 minutes and incubated at 50°C over night. 350 ml RNase buffer (10 mM Tris pH 7.5, 0.3 M NaCl, 5 mM EDTA, 40 mg/ml RNase A, 6.7 U/ml RNase T1) was added, incubated at 30°C for 30 minutes and inactivated with 20 ml 10% SDS and 50 mg proteinase K for 15 minutes at 37°C. After extraction and ethanol precipitation, the samples were resuspended in 10 ml FA buffer, heated at 85°C for 5 minutes, run on a 6% denaturing polyacrylamide gel and visualized by autoradiography. The results were quantified by PhosphorImager analysis using the ImageQuant program.


3.8 Real time PCR and RT-PCR.

      cDNA was synthesized from total RNA using random hexamers and Superscript II reverse transcriptase (Life Technologies). 300 ng genomic DNA or cDNA from 40 ng of total RNA (supplemented with 0.8 mg yeast tRNA or 0.5 mg glycogen) was used per PCR reaction. Real time PCR was performed with the TaqMan sequence detection system (Applied Biosystems). Alternatively, the SYBR green method was used in certain experiments. The PrimerExpress software was used to design the primers and TaqMan probes for detection of mRNAs, non-spliced nascent transcripts and genomic sequences. The sequences of the primers and probes are as follows.

      
Name forward primer (5'-> ; 3') reverse primer (5'-> ; 3') probe (5'-> ; 3')
total CIITA CCTGCTGTTCGGGACCTAAA GGATCCGCACCAGTTTGG AGGGCCCAGCGCAAACTCCAGT
hCIITA I (spliced) CTAGAGAAAGGAGACCTGGATTTG TCATAGAAGTGGTAGAGGCACAGG CTGGAGCTTCTTAACAGCGATGCTGACC
hCIITA III (spliced) TGGGATTCCTACACAATGCGT GGGTCAGCATCGCTGTTAAGA CAGAGCCCCAAGGCAGCTCACAGT
hCIITA IV (spliced) CGGGGAACAGCGGCA TCATAGAAGTGGTAGAGGCACAGG CTGGAGCTTCTTAACAGCGATGCTGACC
hCIITA I (non-spliced) GCCCCGGCCACAGTG TCCCATTGACTTCCCTTTCAGA TGATTAAAAGTGATGCCAACTTACCACCATGG
hCIITA III (non-spliced) TGCTGGGTCCTACCTGTCAGA CAGGACCAGCTGAGACTGCAC CTTTCCCGGCCTTTTTACCTTGGGG
HLA-DRA GCCAACCTGGAAATCATGACA agggctgttcgtgagcaca caactatactccgatcaccaatgtacctccagag
GFP CTGCTGCCCGACAACCAC ACCATGTGATCGCGCTTCTC CCAGTCCGCCCTGAGCAAAGACC
hGAPDH GAAGGTGAAGGTCGGAGTC GAAGATGGTGATGGGATTTC CAAGCTTCCCGTTCTCAGCC
18S rRNA sequence unknown, purchased from Applied Biosystems    
gag GGAGCTAGAACGATTCGCAGTTA GGTTGTAGCTGTCCCAGTATTTGTC ACAGCCTTCTGATGTTTCTAACAGGCCAGG
mito ACCCACTCCCTCTTAGCCAATATT GTAGGGCTAGGCCCACCG CTAGTCTTTGCCGCCTGCGAAGCA

      Samples were quantified using relative standard curves for each amplification. All results were normalized with respect to the internal control, and expressed relative to the levels found in immature DCs.


3.9 Measurement of CIITA mRNA stability

      Immature DCs or DCs stimulated with LPS for 2.5 hours were supplemented with 50 mM 5,6-dichloro-1-b-ribofuranosyl benzamidazole (DRB, Sigma) for 0, 1, 2, 3 and 4 hours. Total RNA was then prepared and the remaining CIITA mRNA was quantified by real time RT-PCR.


3.10 Isolation of nascent RNAs

      107 cells were washed 3 times with cold PBS and resuspended in 1 cell volume HB 0.3 (0.3 M sucrose, 10% glycerol, 60 mM KCl, 15 mM NaCl, 15 mM HEPES pH 7.9, 0.5 mM EDTA, 0.15 mM Spermine, 0.5 mM Spermidine). 1 cell volume HB 0.3 supplemented with 0.8% NP40 was added, gently mixed and incubated on ice for 10 minutes. The lysed cells were layered over 600 ml HB 0.9 (same as HB 0.3 but containing 0.9 M sucrose) and centrifuged for 10 minutes at 1500g at 4°C. The pellet was resuspended in 400 ml NSB (20 mM Tris pH 7.9, 75 mM NaCl, 0.5 mM EDTA, 50% glycerol, 0.85 mM DTT, 0.125 mM PMSF, 0.1 mg/ml tRNA). After addition of 3 ml nuclear lysis buffer ( 20 mM HEPES pH 7.6, 0.3 M NaCl, 1 M urea, 7.5 mM CaCl2, 0.2 mM EDTA, 1 mM DTT, 1% NP-40, 0.1 mg/ml tRNA) and incubation on ice for 10 minutes, the samples were centrifuged for 10 minutes at 15'000 g at 4°C. The supernatant containing the free nuclear transcripts was supplemented with SDS to 0.1% final, extracted, ethanol precipitated and resuspended in 50 ml 10 mM Tris pH 7.5. The chromatin pellets containing the nascent transcripts were resuspended in 2 ml chromatin buffer (50 mM NaAc pH 5, 50 mM NaCl, 0.5% SDS, 0.1 mg/ml tRNA), extracted 3 times with hot phenol (non buffered and supplemented with 1/10 volume of supplement mix (50 mM NaAc pH 5, 50 mM NaCl)). The NaCl concentration was adjusted to 0.15 M and the samples were ethanol precipitated and resuspended in 50 ml 10 mM Tris pH 7.5. The free nuclear RNAs and the nascent transcripts were treated with RNase-free DNaseI (Roche), reverse transcribed and analyzed by real time RT-PCR.


3.11 In vivo genomic footprinting

      107 cells were resuspended in 10 ml RPMI containing 0.1% DMS and incubated for 10 minutes at 37°C. The cells were washed 3 times and resuspended in 1.5 ml harvest buffer (1 mM Tris pH 7.5, 0.4 M NaCl, 2 mM EDTA, 0.2% SDS, 0.2 mg/ml proteinase K). After 3 hours of incubation at 37°C, the lysate was extracted and precipitated with 1/2 volume NH4Ac 7.5 M and 1/2 volume isopropanol. The pellet was resuspended in 1 ml H2O, precipitated with 1/10 volume NaAc 3 M and 3 volumes ethanol and resuspended in 200 ml H2O. The in vitro treated DNA was isolated as described above except that the incubation in 0.1% DMS was omitted. The DMS treatment was instead performed after the second precipitation by adding 0.9 ml DMS to the DNA and incubating for 30 seconds at room temperature. In vivo and in vitro treated samples were then precipitated with 50 ml ice-cooled stop buffer ( 1.5 M NaAc pH 7, 1 M 2-mercaptoethanol, 0.1 mg/ml tRNA) and 750 ml ethanol. The genomic DNAs were cut at the DMS-modified G nucleotides by incubation in 200 ml 1 M piperidine for 30 minutes at 90°C followed by lyophylization. The samples were resuspended in 200 ml HO and precipitated twice with 1/10 volume NaAc 3 M and 2.5 volumes ethanol.

      The ligation-mediated PCR (LM-PCR) was done with Vent DNA polymerase (New England Biolabs) for all steps. For the first extension, 1-2 mg DNA were mixed with 6 ml 1st strand buffer 5x (200 mM NaCl, 50 mM Tris pH 8.9, 25 mM MgSO4, 0.05% gelatine), 1 ml primer 1 0.3 mM, 1 ml dNTP 6.25 mM and 0.5 U Vent polymerase in a final volume of 30 ml. The following cycle was performed: 5 minutes at 95°C, 30 minutes at Tm+2°C, 10 minutes at 76°C. 20 ml ligation solution (110 mM Tris pH 7.5, 18 mM MgCl2, 50 mM DTT, 0.0125% BSA), 15.5 ml ligation mix (10 mM MgCl2, 20 mM DTT, 3 mM ATP, 0.005% BSA), 5 ml annealed linker (20 mM in 250 mM Tris pH 7.7) and 4.5 ml T4 DNA ligase (1 U/ml) were added. After over night incubation at 16°C, the samples were precipitated with 10 mg tRNA, 1/10 volume NaAc 3 M and 2.5 volumes ethanol and resuspended in 71 ml H2O. For amplification by PCR, 20 ml vent buffer 5x (200 mM NaCl, 100 mM Tris pH 8.9, 25 mM MgSO4, 0,045% gelatine, 0.45% Triton X-100), 3.2 ml dNTP 6.25 mM, 1 ml primer 2 (10 mM), 1 ml linker 1 (10 mM) and 1 U vent polymerase were added and the following cycle was performed: 5 minutes at 95°C, then 25 cycles with 1 minute at 95°C, 2 minutes at Tm+2°C, 1 minute at 73°C. For the last extension, 1 ml vent buffer 5x, 1.6 ml dNTP 6.25 mM, 2 ml primer 3 (labeled with g32P-ATP, 3 mM) and 1 U vent polymerase were added and extended for 5 minutes at 95°C, 2 minutes at Tm+2°C, 10 minutes at 76°C. The reaction was stopped by the addition of 295 ml stop mix ( 10 mM Tris pH 7.5, 260 mM NaAc, 4 mM EDTA, 20 mg tRNA), extracted and precipitated. The samples were resuspended in 20 ml formamide loading buffer, run on a sequencing gel and visualized by autoradiography.

      
primer pair primer 1 (5' -> ; 3') primer 2 (5' -> ; 3') primer 3 (5' -> ; 3')
hCIITA pI forward ATTGGCTCCAACAGAAGGCTG CAGAAGGCTGTGGGCTTCTCTG TGGGCTTCTCTGGCACATGCACCTG
hCIITA pI reverse CTGGCCAGTGCCTGGAATC GTGCCTGGAATCTCCGCTCAC TCCGCTCACCCAGCATGCAGCATC
hCIITA pIII reverse AGAAGCACACAGCCTCATCACTA CACTAGCCTCATCACTAACCAGTCA TAACCAGTCACCAGTTGGGAGCCCG
linker 1 GCGGTGACCCGGGAGATCTGAATTC    
linker 2 GAATTCAGATC    


3.12 Protein extracts

      20x106 cells were chilled on ice, washed twice with cold PBS and resuspended in two cell volumes of X-400 buffer (50 mM HEPES pH 7.9, 400 mM KCl, 1 mM EDTA, 1 mM EGTA, 20% glycerol, 2 mM DTT, 1x CompleteTM, 5 mg/ml leupeptin, 1 mM PMSF, 0.5 mM NaF, 0.5 mM Na3VO4). After three cycles of snap-freezing in liquid nitrogen and thawing on ice, the extracts were cleared by centrifugation. The extract concentration was determined using the Bradford method.


3.13 Immunoprecipitation and immunoblotting

      For immunoprecipitation, 5 mg a-CIITA-C antibody (specific for the C-terminal part of CIITA) was covalently coupled with dimethyl pimelinediimidate (DMP, Fluka) in 0.2 M sodium borate pH 9.0 to 32 ml protein-A Sepharose beads. The coupling was stopped by addition of glacial acetic acid and quenched in 0.2 M ethanolamine pH 8.0. After three washes with storage buffer (1x PBS , 0.01% merthiolate, 1 mg/ml BSA), the coupled antibody was stored in this same buffer.

      800 mg extract (corresponding to 5x106 cells) was mixed with 1 volume dilution buffer (20 mM HEPES pH 7.9, 0.1 M EDTA, 0.01% NP-40, 1 mM DTT, 1x CompleteTM) and cleared by centrifugation. The coupled a-CIITA-C antibody was washed in dilution buffer and added to the cleared extract. After 3 hours of incubation with end-over-end rotation at 4°C, the beads were washed 3 times with washing buffer (20 mM HEPES pH 7.9, 10% glycerol, 1 mM EDTA, 0.01% NP-40, 100 mM KCl). The precipitated proteins were then eluted in 11 ml elution buffer (20 mM Tris pH 6.8, 6% glycerol, 2.5% SDS) by incubation at 90°C for 2 minutes. The supernatant was supplemented with 1,2 ml of a mix containing 40% 2-mercaptoethanol, 0.3 M KCl and 0.2% bromphenolblue. The samples were denatured for 2 minutes at 90°C and separated by SDS-PAGE (3.9% stacking and 8% resolving gel).

      The proteins were transferred to a PVDF membrane (ImmobiolonTM-P, Millipore) using a semi-dry transfer cell (Biorad). The membrane was rinsed in TBS-T (100 mM Tris pH 7.9, 0.9% NaCl, 0.1% Tween-20) and blocked in 1% blocking buffer (Roche) over night at 4°C. The a-CIITA-N antibody (specific for the N-terminus of CIITA) diluted 1000 times in 0.5% blocking buffer was added and incubated for 1 hour at room temperature. After 1 wash in TBS-T and 3 washes in 1% milk in TBS-T for 5 minutes each, the secondary antibody coupled to POD (horseradish peroxidase) diluted 2000 times in 0.5% blocking buffer was added and incubated for 30 minutes at room temperature. The membrane was then washed for 5 minutes in TBS-T and for 10 minutes in 1% milk in TBS-T. Revelation was done with Lumi-Light Western Blotting substrate (Roche).


3.14 Chromatin immunoprecipitation and quantification

      40x106 cells were fixed in 40 ml PBS supplemented with 1% formaldehyde, 10 mM NaCl, 5 mM HEPES pH 7.9 for 8 minutes at room temperature. The fixation was stopped by the addition of glycine to a final concentration of 0.18 M. The cells were then washed in cold PBS and stored at -70°C. For isolation of the chromatin, the cell pellet was resuspended in TE containing 1 mM PMSF, 5 mM benzamidine, 1 mg/ml leupeptine, 5 mg/ml aprotinin and 0.5% NP-40 and centrifuged for 5 minutes at 2'500 rpm. The pellet was then resuspended in 50 ml TE containing 1 mM PMSF, 5 mM benzamidine, 1 mg/ml leupeptine, 5 mg/ml aprotinin, 1% Triton X-100, 0.5% Na-deoxycolate and 0.5 M NaCl. After another centrifugation, the pellet was washed in PBS and stored at -70°C. For sonication, the chromatin were resuspended in 1.5 ml TEN (10 mM Tris pH 8.0, 1 mM EDTA, 0.1 M NaCl) and fragmented 12 times for 15 seconds at maximal power. After clearing by centrifugation, the chromatin was stored at -70°C. The size of the sonicated fragments and the concentration of the chromatin were estimated on a 1% agarose gel after reversing the crosslink (see below).

      4 mg chromatin per immunoprecipitation reaction were diluted with 5 volumes of dilution/incubation buffer (20 mM HEPES pH 7.9, 0.2 M NaCl, 2 mM EDTA, 0.1% Na-deoxycolate, 0.1% SDS ,200 mg/ml herring sperm DNA, 1 mg/ml BSA, 1x CompleteTM) and cleared by centrifugation. The supernatant was added to protein A Sepharose CL-4B beads (20 ml per reaction) and incubated for 30 minutes with end-over-end rotation at room temperature. The supernatant was then collected, cleared by centrifugation and diluted with 650 ml dilution/incubation buffer. After addition of the antibodies (specific for acetylated histones H3 and H4 (Upstate Biotechnology) and pre-immune serum), the reactions were incubated at 4°C over night with very slow end-over-end rotation. The immunoprecipitations were then cleared by centrifugation, added to protein A Sepharose beads (12 ml per reaction) in 400 ml dilution/incubation buffer and incubated for 1 hour at room temperature with medium-fast end-over-end rotation. 7 washes were performed with the following buffers: wash 1 and 2: 20 mM HEPES pH 7.9, 0.2 M NaCl, 2 mM EDTA, 0.1% Na-deoxycolate, 0.1% SDS; wash 3 and 4: 20 mM HEPES pH 7.9, 0.5 M NaCl, 2 mM EDTA, 0.1% Na-deoxycolate, 0.1% SDS; wash 5 and 6: 20 mM Tris pH 8.0, 0.25 M LiCl, 2 mM EDTA, 0.5% NP-40, 0.5% Na-deoxycolate; wash 7: TE supplemented with 0.1% NP-40. After the last wash, 500 ml elution buffer (111 mM Tris pH 8.0, 1.11% SDS) were added and incubated at 65°C for 10 minutes. At this step, a chromatin sample (4 mg, "total standard"), that has not been immunoprecipitated, was included and treated as the immunoprecipitated reactions. 14 ml proteinase K (10 mg/ml) and 11 ml NaCl (5 M) were added. After centrifugation of the samples, the supernatant was retrieved and incubated at 42°C for 2 hours and at 67°C over night to reverse the crosslink. The samples were then extracted, precipitated with 20 mg glycogen, 66 ml NaAc (3 M, pH 5.1) and 920 ml isopropanol and resuspended in 50 ml TE plus 50 ml sample dilution buffer (0.25x TE, 30 mM NaCl, 0.2 mg/ml glycogen).

      The immunoprecipitated DNA fragments were amplified by PCR and analyzed on 4% agarose gels or quantified with the SYBR green real time PCR method (Applied Biosystems). The primers for each amplicon were as follows.

      
primer forward (5' -> ; 3') reverse (5' -> ; 3')
A (pI-1.5kb) ccccagctgagagatggtaatc GCACAAAACAGAGGATTTGCATAG
B (pI-0kb) AAAAGCCAATATCCATCCGTTC GCATCCAAAACATGAAGTGAAAAC
C (pII-3.7kb) CCAGGCTGCTTGCGAAC TGCATTTCAGAAGGAGATGGAAT
D (pII-0kb) TGGGGCCCCAGACAATATC GCCCATGTGCCAGTTCAAC
E (pIII-3kb) TCCTGGCTGGGTACCACTG CTCTGAGCAGAGCAAGTGACATC
F (pIII-0kb) AGAAACAGAAATCTGACCGCTTG TCATCACTAACCAGTCACCAGTTG
G (pIV-0kb) CCACTGTGAGGAACCGACTG TGGAGCAACCAAGCACCTAC
DRA ATTTTTCTGATTGGCCAAAGAGTAATT AAAAGAAAAGAGAATGTGGGGTGTAA
GAPDH CCCGGCTACTAGCGGTTTTAC CTGCGGGCTCAATTTATAGAAAC


3.15 Yeast-two-hybrid screening

      The yeast-two-hybrid screen was performed with the DupLEX-A system (OriGene). The 303 base pair fragment coding for the N-terminal extension of the 132 kD CIITA isoform were cloned into the two bait plasmids pEG202 (selectable marker HIS3, C-terminal fusion) and pNLexA (selectable marker HIS3, N-terminal fusion). Yeast strain EGY48 (MATa trp1 his3 ura3 leu2::6 LexAop-LEU2) was transformed with either pEG202-CIITA and the reporter plasmid pJK103 (selectable marker URA3, 2 LexAop-lacZ, medium sensitivity) or pNLexA-CIITA and the reporter plasmid pSH18-34 (selectable marker URA3, 8 LexAop-lacZ, high sensitivity) according to standard protocols (187). The resulting parental strains were tested for autoactivation by the bait of the lacZ and the LEU2 reporter genes. For monitoring lacZ expression, the yeast cells were plated on selective media supplemented with 80 mg/ml X-gal and the color of the colonies was observed. For monitoring LEU2 expression, yeast colonies were tested for growth on leucine-selective plates. The parental strains were transformed with a cDNA library derived from a human B cell line and cloned into the pRS314 vector (selectable marker TRP1) downstream of a VP16 activation domain and under the control of a galactose inducible promoter (188). The transformants were grown over night at 30°C in glucose selective medium. After centrifugation, the cells were resuspended in galactose selective medium and incubated over night at 30°C to induce expression of the library. Transformants were then plated on galactose selective plates lacking leucine. The cDNA expression plasmids were rescued from potential positive transformants according to standard protocols and sequenced.


4. Results

      Parts of the results obtained during this thesis have been the subject of a publication. This publication constitutes the first part of the results chapter of this thesis. In addition, some non-published results are included.


4.1 Expression and promoter usage of the MHC2TA gene in dendritic cells


4.1.1 Publication

      "Maturation of Dendritic cells Is Accompanied by Rapid Transcriptional Silencing of Class II Transactivator (CIITA) Expression"

      Salomé Landmann, Annick Mühlethaler-Mottet, Luca Bernasconi, Tobias Suter, Jean-Marc Waldburger, Krzysztof Masternak, Jean-François Arrighi, Conrad Hauser, Adriano Fontana and Walter Reith.

      Journal of Experimental Medecine (2001), 194 (4): 379-391.

      Summary

      Cell surface expression of MHCII molecules is increased during the maturation of DCs. In contrast to the increased cell surface MHCII expression, de novo biosynthesis of MHCII mRNA is turned off during DC maturation. This paper shows that this is due to a remarkably rapid reduction in the synthesis of class II transactivator (CIITA) mRNA and protein. This reduction in CIITA expression occurs in human monocyte-derived DCs and mouse bone marrow-derived DCs, and is triggered by a variety of different maturation stimuli, including LPS, TNF-a, CD40 ligand, IFN-a, and infection with Salmonella typhimurium or Sendai virus. It is also observed in vivo in splenic DCs in acute myelin oligodendrocyte glycoprotein induced experimental autoimmune encephalitis. The arrest in CIITA expression is the result of a transcriptional inactivation of the MHC2TA gene. This is mediated by a global repression mechanism implicating histone deacetylation over a large domain spanning the entire MHC2TA regulatory region.


4.1.2 Expression of CIITA during the differentiation of monocytes into dendritic cells

      Immature human monocyte-derived DCs express high levels of CIITA mRNA. CIITA type I and type III mRNAs are represented at similar levels, whereas type IV mRNAs are scarcely detectable (Fig.3 in the publication, p.88[384]).

      We were interested to determine how CIITA expression is regulated during the differentiation of human monocytes into DCs. RNA was thus isolated from monocytes and immature DCs and all types of CIITA mRNA were quantified by RNase protection. Monocytes were found to express about ten times less total CIITA mRNA than immature DCs. In contrast to immature DCs, CIITA type III mRNA is the only type expressed at significant levels in monocytes. Type I and type IV mRNAs are both expressed very weakly or not at all in these cells (Fig.15A). In a time course analysis of the differentiation from monocytes into immature DCs over six days, a progressive increase of type I and type III CIITA mRNA expression was observed (Fig.15B). The time course for type I and type III mRNA induction is similar.

      

Fig. 15 CIITA expression during the differentiation from monocytes into immature DCs.

      A. Expression of all three types of CIITA mRNA was analyzed by RNase protection and quantified by PosphoImager in monocytes and immature DCs. The results are standardized with respect to endogenous GAPDH mRNA. B. The expression of CIITA type I and type III mRNA was quantified by real time RT-PCR during the differentiation of monocytes into immature DCs (day 0 to day 6). The values are normalized with respect to 18S rRNA and are given as the percentage relative to the levels found in immature DCs (day 6).


4.1.3 Occupation of the CIITA type I promoter in monocytes and immature dendritic cells

      CIITA type I mRNA expression is strongly induced during the differentiation of monocytes into immature DCs, most likely as the result of activation of transcription. Changes in transcriptional activity are frequently reflected by alterations in promoter occupation that can be visualized in living cells by means of in vivo genomic footprint experiments (189). We thus used this technique to study whether the differentiation of monocytes into immature DCs is accompanied by modifications in the occupation of CIITA pI. The region that was analyzed is situated immediately upstream of the transcription initiation site of pI and contains a short 120 base pair sequence displaying a high degree of homology between the human and the mouse gene.

      The footprints found in immature DCs were mostly the same in monocytes (Fig.16, residues -41, -86 and -90 in the lower strand and residues -49 and -118 in the upper strand). In addition, guanine residues -33 and -75 in the upper strand are occupied in monocytes, but not in immature DCs. It may be speculated that proteins bound to these residues could be transcriptional repressors that bind to pI in monocytes and are released during the differentiation into DCs. The existence and identity of such factors requires further investigations.

      

Fig. 16 Differences in occupation of CIITA pI in monocytes and immature DCs.

      The occupation of pI in monocytes and immature DCs was analyzed by in vivo genomic footprinting. The first and second lane in each panel show the pattern obtained for monocytes and immature DCs, respectively. The third lane in each panel shows the pattern obtained in vitro with naked control DNA. Open arrowheads represent residues occupied in monocytes and immature DCs, filled arrowheads represent residues occupied in monocytes only. The bent arrow indicates the site of transcription initiation.


4.1.4 CIITA pI is not responsive to IFN-g in human peripheral monocytes

      Mouse peritoneal macrophages express high levels of CIITA type I mRNA after exposure to IFN-g (). They do so in both the presence and absence of a functional pIV. This represents the first, and up to now unique, example of pI activation in response to IFN-g.

      In order to determine whether cells of the human monocyte/macrophage lineage display a similar pI usage in response to IFN-g, we analyzed human peripheral blood monocytes after activation with IFN-g. Interestingly, the situation is different in human and mouse cells. Human monocytes strongly upregulate CIITA pIV (and slightly pIII) after exposure to 1000 U/ml IFN-g for 48 hours (Fig.17). However, CIITA pI stays at uninduced levels.

      CIITA promoter usage may thus be regulated in a species-specific manner. An alternative explanation for the above observations may be that the peritoneal macrophages used in the mouse experiments and the peripheral blood monocytes used in the human experiments are not of the same sublineage and can thus not be compared in their respective responsiveness to IFN-g.

      

Fig. 17 CIITA expression of human monocytes in response to IFN-g.

      Human peripheral blood monocytes were stimulated with 1000 U/ml IFN-g for 48 hours or left untreated, and total RNA was isolated. The different CIITA mRNA types were analyzed by real time RT-PCR. The values are normalized with respect to 18S rRNA and are given as the fold induction.


4.2 Molecular dissection of CIITA promoter I


4.2.1 Introduction

      Immature DCs express high levels of CIITA. Similar quantities of CIITA type I and type III transcripts are detectable (Fig.3 in the publication, p.88[384]). The mechanisms controlling the expression of these two types of CIITA in immature DCs are so far unknown. In vivo genomic footprint experiments have revealed occupations in the proximal region of pI within or near sequence motifs representing potential binding sites for known transcription factors (Fig.7 in the publication, p.88[387]). However, none of these motifs or the factors binding to them have been confirmed by functional studies. Concerning pIII, cis-regulatory elements and several factors binding to them have been identified in B cells (190), in which pIII is highly expressed. However, the role of these sequence elements and factors in immature DCs is not clear. In vivo genomic footprint patterns of pIII look similar in B cells and immature DCs (190) (Fig.7 in the publication, p.88[387]).

      We have now investigated the regulation of pI and to a lesser extent of pIII in immature DCs. Different fragments of pI were fused to a reporter gene. To test the activity of these constructs it was necessary to develop a gene transfer system that could introduce exogenous DNA efficiently into DCs.


4.2.2 Gene transfer into dendritic cells

      Primary DCs cannot be transfected easily by classical gene transfer technologies. We have thus developed a transduction system with lentiviral vectors. Lentiviral vectors allow efficient transduction of non-dividing cells and stable integration of the transgene into the genome. DCs can be transduced efficiently with human immunodeficiency virus (HIV)-based vectors (191-193) (Fig.18A). The currently available vectors are optimized in terms of biosafety requirements and efficiency of transgene expression. They contain a 400 base pair deletion in the 3' long terminal repeat (LTR) leading to self-inactivation (SIN) (182). The deletion, which includes the TATA box, abolishes LTR promoter activity and thus allows transgenes to be expressed from tissue-specific or regulated promoters. Transduction of immature DCs with lentiviral vectors does not lead to maturation of these cells as demonstrated by comparing the expression levels of the maturation marker CD83 before and after transduction (Fig 18B).

      

Fig. 18 Transduction of human monocyte-derived DCs with lentiviral vectors.

      A. Immature DCs were transduced with a lentiviral vector expressing GFP under the control of the EF1a promoter at an MOI of 0, 0.6, 3 and 6. Six days after transduction, GFP expression (fl1) was analyzed by FACS. B. Transduced and non-transduced DCs were analyzed by FACS for expression of the maturation marker CD83 (fl2).


4.2.3 The first 390 base pairs of pI are sufficient for transcriptional activity in immature DCs

      To define the functional regions of pI, we constructed a series of progressive 5' deletions of pI fused to the GFP reporter gene in the lentiviral vector plasmid pWPTS (Fig.19D). All constructs contained the first 92 base pairs of the 5' untranslated sequence of CIITA type I. Immature DCs were transduced with the lentiviral vectors and GFP expression was analyzed by FACS. The shortest construct containing only 390 base pairs of the promoter (pI-390) displayed significant GFP expression (Fig.19A). Additional upstream sequences spanning 1042, 1987 and 2846 base pairs of pI (pI-1042, pI-1987 and pI-2846) did not further enhance transcriptional activity, but rather reduced it (Fig.19A). The activity of the different promoter fragments was also analyzed at the mRNA level. Total RNA was isolated from the transduced DCs and GFP mRNA levels were quantified by real time RT-PCR (Fig.19B). The relative quantities of GFP transcripts correlated well with the levels of GFP proteins detected at the cell surface. In order to verify that the reduced level of activity of the pI-1042, pI-1987 and pI-2846 constructs was not due to an unwanted splicing event affecting the integrity of the promoter fragments, we checked their integrity in the transduced cells by southern blot analysis. As shown in Fig 19C, all promoter fragments were of the expected length and should thus be functional.

      

Fig. 19 The first 390 base pairs of pI are sufficient for transcriptional activity in immature DCs.

      A. Four pI fragments of different length (pI-390, pI-1042, pI-1987 and pI-2846) were fused to the GFP reporter gene in the lentiviral vector plasmid pWPTS. Immature DCs were transduced with the vectors and GFP expression was analyzed by FACS six days after transduction. The mean fluorescence intensity is indicated for each sample. n.t., non-transduced. B. Total RNA was isolated from the transduced cells described in A. and GFP mRNA was quantified by real time RT-PCR. The values are normalized with respect to GAPDH mRNA. C. The integrity of the different promoter fragments was checked by Southern blot analysis. Genomic DNA was isolated from HeLa cells transduced with the vectors described in A. and digested with the enzyme AccI. The probe hybridized to fragments of 1394 base pairs (pI-390), 2046 base pairs (pI-1042), 2991 base pairs (pI-1987) and 3850 base pairs (pI-2846). D. Scheme of the pWPTS vector plasmid and the different pI promoter fragments. The restriction sites for cloning (ClaI, BamHI, SalI) and for southern blot analysis (AccI) are indicated.

      In a subsequent set of experiments, the transcriptional activity of the pI-390 promoter fragment was compared to several other promoters. The promoter fragment in the pWPTS vector plasmid was either removed in order to obtain a negative control vector lacking a promoter (p0) or was replaced by the promoter of the elongation factor EF-1a (EF-1a) gene which displays a strong activity in immature DCs (194). In addition, a pWPTS vector plasmid containing the first 322 base pairs of pIII, which has been shown to be sufficient for activity in B cells, was constructed (pIII-322). Immature DCs transduced with the vector containing the EF-1a promoter expressed GFP at high levels (Fig.20A).

      

Fig. 20 pI-390 displays a weaker transcriptional activity than EF-1a.

      A. Immature DCs were transduced with lentiviral vectors expressing GFP under the control of pI-390, pIII-322 or EF-1a. Non-specific background expression was monitored with the p0 construct lacking any promoter. Transduced cells were analyzed for GFP expression by FACS. The mean fluorescence intensity is indicated for each sample. n.t., non-transduced. B. Total RNA was isolated from the transduced cells in A. and GFP mRNA was quantified by real time RT-PCR. The values are normalized with respect to GAPDH mRNA.

      The transcriptional activities of pIII-322 and pI-390 were much weaker than the one of EF-1a, but they were both clearly above the background of p0. Expression driven by pIII-322 was slightly higher than for pI-390 (Fig.20A). These results were confirmed by measuring GFP expression at the mRNA level (Fig.20B). As in the previous experiment, the GFP mRNA levels agree well with the GFP levels observed at the cell surface. The relatively high mRNA background observed with p0 is most probably due to some transcription initiation events from sites in and around the 5' LTR giving rise to functional and non-functional transcripts which cannot be discriminated by our RT-PCR measurement.


4.2.4 The minimal promoter region pI-390 is specific for cells of myeloid origin

      In order to define the cell type specificity of the pI-390 promoter fragment, we tested its activity in different cell types, namely in two epithelial cell lines (HeLa, 293T), in a melanoma cell line (Me67.8), in a B cell line (Raji) and in two promonocytic cell lines (U937, THP1) (Fig.21). The level of background GFP expression (in the cells transduced with the p0 vector) varied from one cell type to another. This may be explained by a strong cell type dependence of transcription initiation from cryptic promoter sequences in and around the 5' LTR. GFP expression levels from p0 were therefore accepted as background for each cell type. The pI-390 promoter fragment proved to be silent in HeLa, 293T, Me67.8 and Raji cell lines. However, it displayed some transcriptional activity in the promonocytic cell lines U937 and THP1. It may thus be speculated that pI-390 expression is not absolutely restricted to DCs but may instead be specific for cells of the myeloid lineage such as monocytes and monocyte-derived DCs. Surprisingly, the pIII-322 promoter fragment displays a high transcriptional activity not only in immature DCs and the B cell line Raji but in all cell types tested. This finding contrasts with the results of earlier studies in which the pIII-322 promoter fragment was analyzed in a transitory transfection system and found to be inactive in the Me67.8 melanoma cell line (82).

      

Fig. 21 pI-390 is specifically active in myeloid cells.

      Immature DCs and HeLa, 293T, Me67.8, Raji, U937 and THP1 cells lines were transduced with lentiviral vectors encoding the GFP reporter gene under the control of the promoters p0, pI-390, pIII-322 and EF-1a. GFP expression was analyzed by FACS six days after transduction. The mean fluorescence intensity is indicated for each sample. n.t., non-transduced.


4.2.5 The transcriptional activity of pI-390 does not change with DC maturation

      Expression of endogenous CIITA in DCs is rapidly silenced upon induction of maturation (Fig.3 in the publication, p.88[384]). The mechanism responsible for this silencing appears to involve a global repression of the entire regulatory region of the MHC2TA gene (Fig.8 in the publication, p.88[388]). Sequences and factors controlling the silencing have not been identified so far. It can thus not be excluded that the proximal promoter regions are involved. In order to test this, we analyzed the activity of the minimal promoter regions of pI and pIII during maturation of DCs. Immature DCs were transduced with the lentiviral vectors encoding GFP under the control of pI-390, pIII-322 or the control promoters p0 and EF-1a. The cells were either left untreated or induced to mature with LPS and GFP expression was then analyzed by FACS. As expected, the activity of the EF-1a promoter, as well as the background activity of p0, were not influenced by maturation of DCs (Fig.22A). GFP expression levels driven by pI-390 and pIII-322 did also not change during maturation (Fig.22A). To confirm these findings we analyzed GFP expression at the mRNA level and did again not find any difference between immature and mature DCs (Fig.22B). Due to the likely implication of a global silencing mechanism of pI and pIII during DC maturation, it is not unexpected that repression cannot be seen with the minimal promoter fragments. Potential regulatory elements involved in repression may lie far upstream or downstream of the transcription initiation sites.

      

Fig. 22 The activities of pI-390 and pIII-322 are not influenced by maturation of DCs.

      A. Immature DCs were transduced with lentiviral vectors encoding GFP under the control of p0, pI-390, pIII-322 or EF-1a as described. Cells were then either left untreated (left column) or induced to mature with 10 ng/ml LPS for 48 hours (right column). GFP surface expression was analyzed by FACS. n.t., non-transduced. B. Total RNA was isolated from the cells in A. and GFP mRNA was quantified by real time RT-PCR. The values are normalized with respect to GAPDH mRNA.

      A potential repression of pI-390 and pIII-322 during DC maturation could have been masked in the previous experiment because of the long half-life of the GFP protein (> 24 hours). A short-lived GFP variant (d4GFP) with a half-life of only four hours also exists. In order to try to reveal a possible repression of pI-390 and pIII-322 during DC maturation, we replaced the long-lived GFP in our lentiviral vector plasmids with the d4GFP variant and transduced immature DCs with the modified vectors. Expression of d4GFP from the EF-1a promoter was weak in both immature and mature DCs (Fig.23). d4GFP was scarcely detectable when placed under the control of pI-390 or pIII-322 and no difference in d4GFP expression could be seen between immature and mature DCs (Fig.23). The short-lived d4GFP reporter gene is thus not suitable for the analysis of the weak pI-390 and pIII-322 promoters.

      

Fig. 23 The short-lived d4GFP is too weak for our system.

      The usual long-lived GFP was replaced in all lentiviral vector plasmids with the short-lived variant d4GFP. Immature DCs were transduced and either left untreated (left column) or induced for maturation with 10 ng/ml LPS for 48 hours (right column). d4GFP expression was analyzed by FACS.


4.2.6 Summary

      We have used a lentiviral gene transfer system to perform reporter gene experiments with CIITA promoter I. We could show that a short 390 base pair fragment (pI-390) is sufficient for expression of the GFP reporter in immature DCs. Larger promoter fragments containing additional upstream sequences did not further enhance the transcriptional activity of pI-390, but rather reduced it. Such an effect of large promoter fragments is frequently observed in reporter gene assays and is not easily explained. The upstream sequences may for instance contain some repression elements that inhibit promoter activity. pI-390 is a relatively weak promoter compared to the minimal promoter region of CIITA pIII (pIII-322) or the EF-1a promoter. However, pI-390 activity is clearly above the background expression of the promoter-less construct p0. pI-390 is transcriptionaly active in two promonocytic cell lines in addition to monocyte-derived DCs. It thus displays specificity for cells of myeloid origin rather than a strict DC specificity. It should be interesting to challenge this hypothesis by testing the activity of pI-390 in primary monocytes and macrophages. We could further show that pI-390 activity is not influenced by maturation of the DCs.


4.3 CIITA pI knockout mice

      Expression of the MHC2TA gene is controlled by a complex regulatory region containing the three independent promoters pI, pIII and pIV (Fig.8, p.41). In the past, several in vitro studies have established the specificities of these promoters. pI drives constitutive expression in DCs, pIII does so in B cells and in certain types of human DCs, while pIV is activated by IFN-g in other cell types. Recently, pIV knockout mice have been generated (85). They have contributed largely to our current understanding of the complicated regulation of CIITA expression. Deletion of pIV selectively abrogates IFN-g-induced MHCII expression in non-bone marrow-derived cells and constitutive MHCII expression in cTECs. However, MHCII expression is unaffected in bone marrow-derived APCs, including DCs, B cells, and IFN-g-activated cells of the macrophage lineage. Macrophages have been found to activate CIITA expression mainly by means of pI (85). In addition, a recent study shows that in wild type mice, macrophages use both, pI and pIV, for CIITA expression in response to IFN-g (). pIV activity declines after a short period of time (6 hours), whereas pI stays active up to 72 hours after IFN-g stimulation.

      Inspired by these new findings on pI in mice and because of the important role of pI for CIITA expression in DCs, we decided to investigate further the role and specificity of pI in vivo. We decided to tackle this by the generation of pI knockout mice using the gene targeting method (196, 197). The targeting construct was designed to replace the coding region of CIITA exon I by a loxP-flanked nitroreductase (NTR) gene, a b-galactosidase (lacZ) gene and a frt-flanked neo gene (Fig.24A). The E.coli NTR converts the prodrug CB1954 ((5-aziridin-1-yl)-2,4-dinitro-benzamide) into a cytotoxic DNA interstrand cross-linking agent leading to DNA crosslinking and DNA strand breakage (198). It thus allows rapid and selective killing of the cells that express NTR, in our case cells that express pI, primarily DCs. Mice devoid of DCs represent an interesting and new tool to study the importance of this cell type under normal conditions and during the course of infections. They may be strongly affected in their capacity to activate their adaptive immune system. NTR-mediated inducible cell ablation offers a number of advantages over the use of HSV1 thymidine kinase for the selective killing of cells in vivo. The former does not require cell proliferation, displays only little bystander activity and does not lead to male sterility. The NTR gene can be excised by means of Cre recombinase, which recognizes the loxP sites flanking the NTR gene in our targeting construct. Excision of the NTR gene allows expression of the downstream lacZ from pI. The lacZ gene serves to determine the pI expression pattern in vivo. lacZ expression can be visualized by in situ staining or by flow cytometry using the substrate di-b-D-galactopyranoside (FDG) (199). The neo gene present in the targeting construct is under the control of the PGK promoter and serves for selection of clones that have integrated the targeting construct. It can be deleted by Flp-mediated recombination to ensure that the in vivo pattern of pI activity is not perturbed.

      ES cells were culture as described in the materials and methods section. ES cell clones in which the targeting construct was integrated by homologous recombination were isolated by a PCR strategy and the presence of the mutated locus was confirmed by Southern blot experiments (Fig.24B).

      

Fig. 24 Generation of heterozygous Mhc2ta pI knockout ES cells.

      A. The regulatory region of the wild type Mhc2ta locus, the targeting construct and the targeted locus are depicted. Exons I and III are represented by open boxes, loxP and frt sites by filled triangles. The EcoRI sites (RI) and probes used for Southern blotting are indicated. B. The blots for three positive (19, 31 and 148, encircled) and one negative (151) ES cell clones probed with the 5'probe (left) and 3' probe (right) are shown. Both probes hybridize to a 12.3 kb fragment in the wild type locus. In the targeted locus, the 5' probe hybridizes to a 3.5 kb fragment and the 3' probe to an 8.5 kb fragment.

      The neo gene was deleted in the positive ES cell clone 148 by transfecting it with a Flp expression vector (Fig.25A). neo-deleted clones were identified by a PCR strategy and confirmed by Southern blot experiments (Fig.25B). The clones in which the neo gene was deleted have been amplified and the cells are currently being injected into blastocysts in collaboration with the group of Ari Waisman in Cologne.

      

Fig. 25 Flp-mediated deletion of the targeted locus.

      A. The regulatory region of the wild type Mhc2ta locus, the targeted locus and the neo-deleted locus are depicted. Exons I and III are represented by open boxes, loxP and frt sites by filled triangles. The PstI sites and the probe used for Southern blotting are indicated. B. The blot for four clones in which the neo gene was deleted (148.1, 148.23, 148.40 and 148.56, encircled) and five clones in which the neo gene was not deleted (148.4, 148.7, 148.14, 148.15 and 148.18) is shown. The probe hybridizes to a 3.5 kb fragment in the wild type locus, to a 3.9 kb fragment in the targeted locus and to a 12.2 kb fragment in the deleted locus.


4.4 Interaction partners of the N-terminus of dendritic cell-specific CIITA

      As a consequence of the existence of three different CIITA promoters, there are three CIITA isoforms differing in their N-terminus (Fig.8, p.41). The largest 132 kD isoform, which is mainly expressed in DCs, results from initiation at the first AUG initiation codon in the CIITA type I mRNA. It contains a 101 amino acid N-terminal extension not present in the other isoforms and shows enhanced capacity to activate MHCII transcription compared to the other CIITA isoforms (200) (Luc Otten, unpublished results). In addition, the 132 kD isoform seems to have an extremely short half-life (our unpublished observations). It is likely that these features are due to an intrinsic property of the N-terminus unique to this CIITA isoform. We were thus interested in investigating the role, structure and function of this N-terminal extension. A yeast-two-hybrid approach was chosen to find possible interaction partners. This method allows the identification of proteins that can bind to a protein of interest.

      The N-terminal 101 amino acids of the 132 kD CIITA isoform (bait) were fused to the bacterial LexA DNA binding domain. The VP16 activation domain was placed upstream of a human B cell cDNA library controlled by the inducible GAL1 promoter. Two independent reporter genes were used: the b-galactosidase gene under the control of two LexA operator sites and the chromosomal LEU2 gene under the control of six LexA operator sites. 3x107 independent yeast transformants were screened for expression of the LEU2 gene. 140 LEU2 positive colonies were selected and tested for expression of the b-galactosidase gene. The cDNA expression plasmids were recovered from 13 strong blue colonies and transformed back into the parental strains. Four of them could be confirmed in a second round of selection. Three out of the four clones interact strongly with the bait and contain the same 2.24 kb insert coding for a guanine nucleotide-exchange protein (p200, BIG1) (201-203). The fourth clone interacts only weakly with the bait and contains a 0.54 kb insert coding for a homologue of the mouse WW domain binding protein 3 (204). Interestingly, the mutation L27Q of the bait, which was found to reduce the transactivation potential of CIITA type I (compared to the wild type protein) in mammalian cells (200), abolishes the interaction with all of the four clones isolated in the yeast system.

      The BIG1 guanine nucleotide exchange protein (202, 203) catalyzes the replacement of GDP with GTP on ADP-ribosylation factors (ARFs), which play an important role in intracellular vesicular trafficking. BIG1 is expressed ubiquitously in eukaryotes and localizes to the Golgi apparatus (203). It contains a Sec7 domain, which confers sensitivity to the drug brefeldin A (205). However, it is not clear how BIG1 interacts with the N-terminal extension of the 132 kD CIITA isoform and what the role of this interaction could be. The WW domain binding protein 3 is a novel protein containing a proline-rich region including the motif PPLP (204). It was isolated as a ligand of the WW domain of the formin-binding protein 11 (204). Several WW domain containing proteins have been localized in the nucleus and have been documented to participate in co-activation of transcription and modulation of RNA polymerase II activity (206). The precise link between proteins containing a WW domain, WW domain binding proteins and CIITA awaits further investigation. The N-terminus of CIITA itself however does not contain a WW domain.


5. Discussion : conclusions and perspectives

      CIITA is the master regulator of MHCII expression. The work presented in this thesis has shed light on how CIITA controls MHCII expression in DCs. In particular, it describes the constitutive expression of CIITA type I and type III in immature DCs and the rapid transcriptional silencing of the CIITA gene during DC maturation. Furthermore, the generation of CIITA pI knockout mice was initiated. This discussion focuses on the non-published results. The published results have been extensively discussed in the publication and only some aspects will be addressed again here.


CIITA silencing in dendritic cells during maturation

      Cell surface expression of MHCII molecules is increased during DC maturation, while de novo biosynthesis is turned off. We have shown that the reduction in MHCII synthesis is due to silencing of the MHC2TA gene during DC maturation. Reduction of CIITA expression was observed in response to a variety of different DC maturation stimuli as well as in vivo in splenic DCs during MOG induced EAE. Our results demonstrate further that inactivation of the MHC2TA gene during DC maturation is mediated by a global repression mechanism implicating histone deacetylation over a large domain.

      The changes in CIITA levels during the course of EAE are remarkable (Fig.4D in the publication, p.88[385]). CIITA expression is at least 20 times lower in purified splenic DCs from EAE mice as compared to normal mice. All three types of CIITA (I, III and IV) are affected. CIITA type I is however most relevant because it is the only type to be expressed at high levels in immature mouse DCs. The almost complete extinction of CIITA expression in splenic DCs of EAE mice suggests that the entire DC population is induced to maturate. Immunization of mice with MOG in adjuvant triggers a massive systemic inflammation. One may expect that the observed CIITA shutdown and DC maturation are consequences of the systemic inflammation rather than a specific response to MOG immunization. It will be very interesting to compare the situation observed in EAE with other inflammatory and autoimmune diseases (i.e. lupus or arteriosclerosis). During minor local inflammations, not enough DCs may be activated and matured to detect any CIITA reduction in response to DC maturation in the population of total splenic DCs. Progress in understanding the role of DCs in the course of MOG-induced EAE should be made by the detailed characterization of splenic DCs from EAE mice. DCs are a heterogeneous population and at least three DC subpopulations have been identified in the mouse spleen (207). They differ in their phenotype, localization and function. Lymphoid CD4-CD8+ DCs reside in the T cell areas of the periarterial lymphoid sheath (PALS), whereas myeloid CD4+CD8- and CD4-CD8- DCs are present in the marginal zones. CD8+ CDs tend to induce a Th1-biased T cell response, whereas CD8- DCs tend to induce a Th2-biased response (208-210). In addition, CD8+ DCs were reported to induce tolerization (211). However, this conflicts with their role in promoting Th1 responses. The characterization of the phenotype of the DCs from EAE mice will possibly give some indications on their function, be it immunogenic or tolerogenic. Interestingly, recent papers have reported that some tolerogenic subtypes of DCs are able to prevent or relieve of EAE (212-214).

      Along with transcriptional inactivation of the MHC2TA gene, a large regulatory region is deacetylated in mature DCs (Fig.8B and C in the publication, p.88[388]). The region extending from 1.5 kb upstream of pI down to pIV (15 kb) has been analyzed in two to three kb intervals. It will be interesting to investigate how large the deacetylated region is. Does it cover the coding region of MHC2TA ? How far does it extend upstream of pI ? The boundaries of the deacetylated region and fluctuations in the acetylation levels may give some hints on the control of deacetylation and the repression mechanism.

      Almost 40 years ago, histone deacetylation was already correlated with transcriptional repression (215). Under normal conditions, up to 13 of the 30 lysine residues in the tails of the core histones are acetylated (216). This steady-state level of acetylation is maintained by the opposing actions of histone acetyltransferases (HATs) and histone deacetylases (HDACs) (217, 218). It is believed that targeting of HATs and HDACs to promoter regions creates specific patterns of hyper- and hypoacetylation in a background of global acetylation that correlates with transcription activation and repression, respectively (219, 220). How does histone acetylation regulate transcription of specific genes ? Although histone acetylation does not disrupt individual nucleosomes, moderate levels of acetylation can destabilize the higher order folding of nucleosomal arrays (221), and the acetylation of specific lysine residues can regulate the chromatin binding or enzymatic activity of other nonhistone-proteins (222-224). Many HATs and HDACs controlling the acetylation state of chromatin have been identified. They always seem to act as a part of large complexes. They are unable to access their histone substrates unless they are targeted there by DNA bound activators or repressors. Recently, it has been recognized that different HDACs deacetylate preferentially very specific lysine residues in different chromatin regions (225). Some have diverse promoters as their targets whereas others are specialized for heterochromatin or ribosomal DNA. There is thus only a small degree of functional overlap among different HDACs. It should be interesting to try to determine which HDACs and HATs are implicated in modulating the acetylation status of the MHC2TA gene in DCs.

      Acetylation is not the only type of covalent modification found in histones. Histone tails can also be methylated, phosphorylated and ubiquitinated. Methylation occurs on different lysine and arginine residues of histone H3 and H4 and the transcriptional consequences can be either negative or positive. Methylation of lysine 4 (H3) and arginine 17 (H3) for instance are linked to gene activation (226, 227), whereas methylation of lysine 9 (H3) results in gene repression (228). Numerous methyltransferases targeting specific lysine or arginine residues have been identified so far. When all possible histone modifications are taken into account, one can hypothesize that certain combinations of modifications in one or more histone tails act sequentially or concomitantly to form a histone code (229, 230). Different patterns of modifications correlate with specific transcriptional states. An inactive state can for instance be characterized by deacetylation at lysine 14 (H3), which precedes methylation of lysine 9 (H3). Further work in our system should attempt to establish the histone modification code at the MHC2TA gene in immature DCs, and how it changes during DC maturation.

      In summary, our knowledge on histone modifications and transcription regulation is growing constantly. The increasing number of known histone modifications and histone modifying enzymes provides new tools and should allow us to characterize further the mechanism leading to silencing of the MHC2TA gene during DC maturation. First, the exact lysine residues that are deacetylated need to be identified. Chromatin immunoprecipitation experiments using antibodies specific for single acetyllysine residues should be helpful. When histone methylation and eventually phosphorylation and ubiquitination are also taken into account, an entire pattern of histone modifications may be recognized. This should then allow us to identify the corresponding HDACs and histone methyltransferases involved in the repression process and eventually to situate them in the context of the signal transduction events involved in DC maturation.


MHC2TA promoter usage

      Expression of CIITA is controlled by three alternative promoters (Fig.8, p.xx). They are used differentially in a cell type specific and inducible manner. According to the classical model, pI is active constitutively in DCs, pIII is used in B cells and pIV is inducible by IFN-g. Recent results have changed this view of promoter usage to some extent. Monocytes and macrophages use pI in addition to pIV for IFN-g induced CIITA expression (85). Differences in the kinetics of pI versus pIV CIITA expression induced by IFN-g could be evidenced (195). Similar levels of CIITA type I and type IV are present after six hours of IFN-g stimulation. However, after 72 hours of exposure to IFN-g, CIITA type IV declines while CIITA type I remains constant. Consistent with this, mice lacking pIV don't lose IFN-g inducible CIITA expression in monocytes and macrophages, but selectively abrogate it in non-hematopoietic cells. A modified model of promoter usage may thus be proposed (Fig.8, p.41). Cells of myeloid origin (DCs, monocytes, macrophages) require pI for constitutive or IFN-g induced CIITA expression. Constitutive CIITA expression in cells of lymphoid origin (B cells, T cells) is driven by pIII. Finally, pIV is indispensable for expression in non-hematopoietic cells (fibroblasts, astrocytes, cTECs). Some species-specific differences have also been recognized. In humans, pI does not respond to IFN-g, either in monocytes or in monocyte-derived DCs (Fig.5D of the publication, p.88[385] ; Fig.17 p.92). In addition, human monocyte-derived DCs express not only CIITA type I but also type III (Fig.3 in the publication, p.88[384]), while mouse DCs use essentially only pI. The promoter usage in other human DC subtypes, for instance in plasmocytoid DCs or thymic DCs, has not yet been studied.

      In order to confirm and refine our current knowledge about CIITA promoter usage, we have initiated the project of creating pI knockout mice. These mice should lack CIITA, and thus MHCII, expression on their DCs. It is not clear however, if all DC subtypes will be equally affected. If thymic DCs lack MHCII expression in the pI knockout mice, negative selection of thymocytes will most likely be strongly impaired leading to a strong susceptibility for autoimmune diseases. Furthermore, it is an open question how monocytes and macrophages will respond to IFN-g in these mice. What will be the consequences of the changed MHCII expression for the induction of primary immune responses ?

      The pI knockout mice were designed such that the first exon of CIITA type I will be replaced by the NTR gene. NTR will thus be expressed in all cells in which pI is active, most likely DCs. These cells will be killed selectively after administration of the prodrug CB1954. Mice devoid of DCs should represent an interesting and new tool to assess the physiological role of DCs in the immune response and in thymocyte differentiation. They should also help to understand the consequences of DC dysfunction. The NTR gene in the pI knockout mice will be flanked by loxP sites and can thus be eliminated by means of the Cre recombinase. This will lead to pI-driven expression of the lacZ gene, which was introduced downstream of the loxP-flanked NTR gene. Expression of lacZ under the control of pI should help to identify all cell types that express pI in vivo under normal or pathological conditions.

      Identification of the alternative CIITA promoters raises the question of whether the three resulting CIITA transcripts and protein isoforms have different functions. Each of the three CIITA transcripts has a different 5' end (Fig.8, p.41). They give rise to three protein isoforms, each possessing a different N-terminus. The role of the different N-terminal extensions is not yet understood. The extension of the 132 kD isoform resulting from the CIITA type I transcript consists of 101 N-terminal amino acids. It may confer an enhanced transactivation capacity to the 132 kD isoform compared to the other two isoforms (200) (Luc Otten, unpublished results). However, this hypothesis awaits further experimental support. Our attempt to identify interaction partners of the 101 amino acid N-terminal extension by means of a yeast-two-hybrid screen resulted in the isolation of two candidates, a guanine nucleotide-exchange protein and a homologue of the mouse WW domain binding protein 3. The relevance of these two proteins and their interaction with CIITA are not yet clear. The new technique of RNA interference (RNAi) should be suitable for analyzing their role in CIITA mediated MHCII transcription. RNAi is a process of sequence-specific post-transcriptional gene silencing in animals and plants. It is initiated by a double-stranded RNA that is homologous in sequence to the silenced gene. The mediators of this process are 21 to 22 nucleotide small interfering RNA (siRNA) duplexes that cause efficient and specific downregulation of gene expression, resulting in functional inactivation of the targeted genes. Targeting siRNAs to the candidate genes isolated in our yeast-two-hybrid system should give important clues concerning the relevance and role of these genes in modulating the activity of CIITA.


Molecular dissection of CIITA promoter I

      The regulatory mechanisms controlling two of the three CIITA promoters (pIII and pIV) have been addressed several years ago (82, 86, 190). In this respect, pI is a straggler. The molecular analysis of pI has been severely hampered by the necessity of working with primary DCs, which are difficult to obtain in large numbers and to transfect by classical transfection methods. Stable DC cell lines have been established (231-234), but they are not suitable because they either do not express significant levels of CIITA and MHCII, or do not respond to maturation stimuli as primary DCs do. We therefore decided to investigate the regulation of pI in primary DCs by means of a lentiviral transduction system. Lentiviral vectors allow very efficient transduction of non-dividing cells. However, the production of the vectors is variable in efficiency and very time-consuming.

      We have determined the minimal promoter region of pI, which encompasses 390 base pairs. Compared to EF-1a, pI-390 is a rather weak promoter. Its activity is also only about half of that of the minimal promoter region of pIII, pIII-322. The activity of pI-390 is restricted to DCs and promonocytic cell lines. However, its activity in primary monocytes and macrophages remains to be determined. It is thus possible that pI-390 expression is more specific for DCs in vivo than suggested by our reporter gene experiments.

      Due to its restricted expression, pI-390 attracts attention with regards to possible applications in the field of immunotherapy. DCs represent an ideal adjuvant for immunization against pathogens or tumor antigens. The availability of a DC-specific promoter offers great advantages over constitutive promoters. It ensures DC-specific expression of the antigen even if a heterogeneous population of cells (containing DCs or DC precursors) is transduced, such that isolation of pure DCs is not required. In the context of immunotherapy it is worth mentioning that the activity of pI-390 does not change during the maturation of DCs. This allows the constitutive supply and presentation of antigen, which should assist the induction of potent CTL responses and protective immunity.

      Our genomic footprint experiments (Fig.7 in the publication, p.88[387] ; Fig.16, p.91) have pinpointed a number of sites in pI that are occupied in DCs in vivo. This represents a strong indication that these sequences constitute important cis-acting regulatory elements for the activity of pI in DCs. Another indication of important promoter regions is generally provided by the presence of sites that display strong homology between different species. In the case of pI, the sites that are occupied in the in vivo footprint experiments do not always match the most conserved stretches of sequence. It will thus be a good idea to mutate both types of sequences, the occupied and the conserved ones, in the context of pI-390 in order to analyze their role in the regulation of pI. These experiments may be performed with the recently developed nucleofection technology (235) instead of the lentiviral transduction system. Nucleofection is an improvement of the classical electroporation technology and allows highly efficient and non-viral transient transfection of primary cells. DCs have been transfected this way with an efficiency of over 50% (data not shown). The mutated pI-390 promoter fragments can be fused to the luciferase reporter gene and thus be analyzed by a classical luciferase assay. This method is highly quantitative and appropriate for the comparison of different promoter mutants. Once relevant cis-acting regulatory elements have been identified, the task of identifying the transcription factors that bind to these sequences will have to be tackled.

      In conclusion, this thesis has contributed to our understanding of the mechanism controlling MHCII biosynthesis in immature and mature DCs. In particular, it provides insights into the role of the master regulator CIITA during DC maturation. Furthermore, the regulation of the DC-specific CIITA promoter was investigated. The results have opened interesting perspectives for anti-tumor immunotherapy and vaccination.


[Previous] [Next]