The regulation of IgA class switching.
Journal: 2008/June - Nature Reviews Immunology
ISSN: 1474-1741
Abstract:
IgA class switching is the process whereby B cells acquire the expression of IgA, the most abundant antibody isotype in mucosal secretions. IgA class switching occurs via both T-cell-dependent and T-cell-independent pathways, and the antibody targets both pathogenic and commensal microorganisms. This Review describes recent advances indicating that innate immune recognition of microbial signatures at the epithelial-cell barrier is central to the selective induction of mucosal IgA class switching. In addition, the mechanisms of IgA class switching at follicular and extrafollicular sites within the mucosal environment are summarized. A better understanding of these mechanisms may help in the development of more effective mucosal vaccines.
Relations:
Content
Citations
(170)
References
(140)
Pathways
(1)
Chemicals
(1)
Organisms
(2)
Processes
(3)
Affiliates
(2)
Similar articles
Articles by the same authors
Discussion board
Nat Rev Immunol 8(6): 421-434

The regulation of IgA class switching

Function of IgA class switching

Antibody diversification is essential for the immune system to mount protective humoral responses. B cells diversify their antibody repertoire through three main genetic alterations that occur in two distinct phases of B-cell development. In the antigen-independent phase, B-cell precursors lodged in the bone marrow generate antigen recognition diversity by assembling the exons that encode immunoglobulin heavy (H) and light (L) chain variable regions from individual variable (V), diversity (D) and joining (J) gene segments through V(D)J gene recombination4. This process is initiated by a lymphoid-cell- and sequence-specific RAG1 (recombination- activating gene 1)–RAG2 endonuclease complex and is completed by the non-homologous end-joining machinery4. Productive assembly of VHDJH and VLJL exons allows the expression of IgH and IgL chains as cell-surface IgM by newly generated B cells4. After further differentiation and expression of IgD, B cells emerging from the bone marrow migrate to secondary lymphoid organs, where they initiate the antigen-dependent phase of B-cell development.

In the presence of antigen, mature B cells diversify their antibody repertoire through somatic hypermutation (SHM) and class switching56. These processes take place in the germinal centres of secondary lymphoid follicles7 and require the DNA-editing enzyme activation-induced cytidine deaminase (AID)8. The process of SHM introduces point mutations at high rates into VHDJH and VLJL exons, thereby providing the structural correlate for selection by antigen of high-affinity immunoglobulin variants5. Class switching substitutes the IgH constant region μ (Cμ) and Cδ genes encoding primary IgM and IgD isotypes with Cγ, Cα or Cε genes through a process known as class-switch recombination (CSR)9. This molecular event generates secondary IgG, IgA and IgE isotypes that have the same antigen specificity as IgM and IgD, but different effector functions10. Indeed, secondary isotypes can activate multiple innate immune effector cells, including phagocytes, by binding to specific Fc receptors10. Together with post-IgA CSR modifications, IgA CSR generates multiple forms of membrane-bound IgA and of secreted IgA (sIgA), each characterized by a distinct location in the body and by distinct functions. Remarkably, some of these forms of IgA substantially differ in humans and mice (TABLE 1). Unlike mouse IgA, which comprises only one class, human IgA comprises two subclasses, IgA1 and IgA2, the latter being more abundant in the intestinal and genitourinary tracts. In addition, circulating IgA is predominantly monomeric in humans, but largely dimeric and oligomeric in mice.

Table 1

Differences in IgA class switching and production between mice and humans

ParameterMiceHumansReferences
Organization of the Cα locusOne Cα locusTwo Cα loci6,8,9
Modes of IgA CSRNo sequential IgA1-to-IgA2 CSRSequential IgA1-to-IgA2 CSR6,8,9,99
B cells undergoing IgA CSRB-1-and B-2-cell subsetsNo canonical B-1-cell subset1,103,127
Requirements for IgA CSRLPS induces IgA CSR via TLR4 in B cellsB cells lack TLR4 and are unresponsive to LPS9,84
Requirements for IgA secretionIL-5 increases IgA secretionIL-5 does not influence IgA secretion9
Types of IgA producedOne IgA class onlyTwo subclasses, IgA1 and IgA21,9,2123
Form of systemic IgA antibodiesMostly IgA oligomersMostly IgA1 monomers1,21
Form of mucosal IgA antibodiesIgA oligomersMostly IgA1 and IgA2 oligomers1,21
Effector functions of IgA antibodiesNo FcαRI expressionFcαRI expressed by innate immune cells18

CSR, class-switch recombination; FcαRI, type I Fc receptor for IgA; IL-5, interleukin-5; LPS, lipopolysaccharide; TLR4, Toll-like receptor 4.

Mucosal IgA

Mucosal secretions contain IgA dimers and oligomers in both mice and humans1. These IgA polymers originate from the interaction of IgA monomers with the J chain, a polypeptide synthesized by antibody-secreting cells11. In addition to assembling monomeric IgA, the J chain interacts with the polymeric immunoglobulin receptor (pIgR), an antibody-transporting protein expressed on the basolateral surface of mucosal epithelial cells12. The pIgR shuttles IgA across epithelial cells through a transcytotic process that culminates in the translocation of sIgA complexes to the mucosal surface1. These complexes comprise a secretory component that originates from the endocytic cleavage of pIgR and that confers mucophilic properties to sIgA1314. Remarkably, sIgA neutralizes toxins and pathogens without causing inflammation because of its inability to fix and activate the complement cascade1. In addition, sIgA anchors commensal bacteria to the mucus, thereby impeding their entry to the underlying intestinal mucosa1. Furthermore, sIgA promotes the establishment of a mutualistic host–microbe relationship by down-modulating the expression of pro-inflammatory epitopes by commensal bacteria15. Moreover, sIgA neutralizes microbial compounds with pro-inflammatory activity, such as lipopolysaccharide (LPS)16, and facilitates the formation of a biofilm that favours the growth of commensals while attenuating that of pathogens17. This property might depend on the ability of sIgA to agglutinate intestinal bacteria through carbohydrates associated with the Fcα region1. The secretory component would further increase the contributing capacity of sIgA to biofilm formation by anchoring intestinal bacteria to the mucus layer lining the epithelial-cell surface14. Finally, sIgA enhances mucosal immunity by delivering bacterial cargo to M cells (microfold or membrane cells)18, a specialized epithelial-cell type that is found adjacent to intestinal Peyer’s patches19.

Systemic IgA

Circulating IgA is largely present as a monomer in humans, although circulating IgA polymers are also present1. By contrast, circulating IgA is largely polymeric in mice1. Systemic IgA binds to various receptors expressed by granulocytes, monocytes, macrophages, DCs, eosinophils, follicular DCs, hepatocytes, hepatic Kupffer cells and renal mesangial cells, including the myeloid-cell-specific type I Fc receptor for IgA (FcαRI; also known as CD89), the Fcα/Fcμ receptor, the asialoglycoprotein receptor and the transferrin receptor20. The effector functions of these receptors remain poorly understood, although growing evidence indicates that FcαRI provides a second line of defence against intestinal bacteria that invade the portal venous system21. In particular, FcαRI might facilitate the internalization of IgA-opsonized bacteria by hepatic Kupffer cells and other phagocytic cells in a non-inflammatory context1922. Indeed, FcαRI engagement by IgA triggers the recruitment of SHP1 (SRC-homology-2-domain-containing protein tyrosine phosphatase 1), a crucial negative regulator of multiple pro-inflammatory receptors20. Finally, it must be noted that, as mice do not express FcαRI, most of the data documenting the in vivo function of FcαRI were derived from mice expressing a human FCAR transgene (TABLE 1).

Human IgA subclasses

The human IgA1 and IgA2 subclasses are encoded by two distinct Cα1 and Cα2 genes and possess a seemingly identical receptor-binding profile, but a different distribution in the body110. Indeed, the circulating IgA pool is comprised mostly of IgA1, whereas the mucosal IgA pool contains both IgA1 and IgA2 (REF. 23). IgA2 is particularly abundant at sites colonized by a large microbiota, including the distal intestinal tract and the urogenital tract2426. Another difference between IgA1 and IgA2 relates to the fact that IgA1 has a longer hinge region than IgA2 (REF. 27). This feature renders IgA1 more susceptible to degradation by bacterial proteases that target the hinge region of IgA2328. Furthermore, compared to IgA1 antibodies, IgA2 antibodies seem to have superior Fcα-mediated, mannose-dependent agglutinating properties against enteric microorganisms and exhibit more VH-mediated reactivity against LPS, a key component of Gram-negative bacteria residing in the distal gut2325.

Mucosal IgA

Mucosal secretions contain IgA dimers and oligomers in both mice and humans1. These IgA polymers originate from the interaction of IgA monomers with the J chain, a polypeptide synthesized by antibody-secreting cells11. In addition to assembling monomeric IgA, the J chain interacts with the polymeric immunoglobulin receptor (pIgR), an antibody-transporting protein expressed on the basolateral surface of mucosal epithelial cells12. The pIgR shuttles IgA across epithelial cells through a transcytotic process that culminates in the translocation of sIgA complexes to the mucosal surface1. These complexes comprise a secretory component that originates from the endocytic cleavage of pIgR and that confers mucophilic properties to sIgA1314. Remarkably, sIgA neutralizes toxins and pathogens without causing inflammation because of its inability to fix and activate the complement cascade1. In addition, sIgA anchors commensal bacteria to the mucus, thereby impeding their entry to the underlying intestinal mucosa1. Furthermore, sIgA promotes the establishment of a mutualistic host–microbe relationship by down-modulating the expression of pro-inflammatory epitopes by commensal bacteria15. Moreover, sIgA neutralizes microbial compounds with pro-inflammatory activity, such as lipopolysaccharide (LPS)16, and facilitates the formation of a biofilm that favours the growth of commensals while attenuating that of pathogens17. This property might depend on the ability of sIgA to agglutinate intestinal bacteria through carbohydrates associated with the Fcα region1. The secretory component would further increase the contributing capacity of sIgA to biofilm formation by anchoring intestinal bacteria to the mucus layer lining the epithelial-cell surface14. Finally, sIgA enhances mucosal immunity by delivering bacterial cargo to M cells (microfold or membrane cells)18, a specialized epithelial-cell type that is found adjacent to intestinal Peyer’s patches19.

Systemic IgA

Circulating IgA is largely present as a monomer in humans, although circulating IgA polymers are also present1. By contrast, circulating IgA is largely polymeric in mice1. Systemic IgA binds to various receptors expressed by granulocytes, monocytes, macrophages, DCs, eosinophils, follicular DCs, hepatocytes, hepatic Kupffer cells and renal mesangial cells, including the myeloid-cell-specific type I Fc receptor for IgA (FcαRI; also known as CD89), the Fcα/Fcμ receptor, the asialoglycoprotein receptor and the transferrin receptor20. The effector functions of these receptors remain poorly understood, although growing evidence indicates that FcαRI provides a second line of defence against intestinal bacteria that invade the portal venous system21. In particular, FcαRI might facilitate the internalization of IgA-opsonized bacteria by hepatic Kupffer cells and other phagocytic cells in a non-inflammatory context1922. Indeed, FcαRI engagement by IgA triggers the recruitment of SHP1 (SRC-homology-2-domain-containing protein tyrosine phosphatase 1), a crucial negative regulator of multiple pro-inflammatory receptors20. Finally, it must be noted that, as mice do not express FcαRI, most of the data documenting the in vivo function of FcαRI were derived from mice expressing a human FCAR transgene (TABLE 1).

Human IgA subclasses

The human IgA1 and IgA2 subclasses are encoded by two distinct Cα1 and Cα2 genes and possess a seemingly identical receptor-binding profile, but a different distribution in the body110. Indeed, the circulating IgA pool is comprised mostly of IgA1, whereas the mucosal IgA pool contains both IgA1 and IgA2 (REF. 23). IgA2 is particularly abundant at sites colonized by a large microbiota, including the distal intestinal tract and the urogenital tract2426. Another difference between IgA1 and IgA2 relates to the fact that IgA1 has a longer hinge region than IgA2 (REF. 27). This feature renders IgA1 more susceptible to degradation by bacterial proteases that target the hinge region of IgA2328. Furthermore, compared to IgA1 antibodies, IgA2 antibodies seem to have superior Fcα-mediated, mannose-dependent agglutinating properties against enteric microorganisms and exhibit more VH-mediated reactivity against LPS, a key component of Gram-negative bacteria residing in the distal gut2325.

Mechanism of IgA class switching

Mature B cells acquire IgA expression by undergoing CSR of Cμ to Cα (FIG. 1). CSR involves an exchange of upstream donor Cμ and Cδ genes with a downstream acceptor CH gene through a recombinatorial process that is guided by switch (S) regions10. S regions are located upstream of each CH gene, except Cδ, and consist of highly repetitive 1–12 kilobase sequences with G-rich non-template strands9. Each S region is preceded by a short intronic (I) exon and a promoter that initiates germline CH gene transcription when the B cell is exposed to activating stimuli9. Germline transcription is crucial for CSR as it renders the S region a substrate for AID, an inducible APOBEC (apolipoprotein B mRNA-editing enzyme, catalytic component 1) family member encoded by AICDA9. AID is essential for CSR, as Aicda-knockout mice or patients with AICDA mutations develop hyper-IgM type 2 syndrome (HIGM2) and fail to generate class-switched antibodies, including IgA2931.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f1.jpg
Recombinatorial and transcriptional events underlying IgA class switching

The immunoglobulin heavy chain (IgH) locus of mature B cells contains a rearranged variable (V) diversity (D) joining (J) exon encoding the antigen-binding domain of an immunoglobulin. Following rearrangement of the light chain, B cells produce intact IgM and IgD through a transcriptional process driven by a promoter (P) upstream of the VDJ exon. Production of downstream IgG, IgA or IgE with identical antigen specificity occurs through class-switch recombination (CSR). Appropriate stimuli induce germline transcription of the constant heavy chain α (Cα) gene from the promoter (Pα) of the intronic α (Iα) exon through the switch α (Sα) region between Iα and Cα exons. In addition to yielding a sterile Iα–Cα mRNA, germline transcription renders the Cα gene substrate for activation-induced cytidine deaminase (AID), an essential component of the CSR machinery. By generating and repairing DNA breaks at Sμ and Sα, the CSR machinery rearranges the IgH locus, thereby yielding a deletional recombination product known as the switch circle. This episomal DNA transcribes a chimeric Iα–Cμ mRNA under the influence of signals that activate Pα. Post-switch transcription of the IgH locus generates mRNAs for both secreted IgA and membrane IgA. Cα 1–3, exons that encode the Cα chain of IgA; S, 3′ portion of Cα3 encoding the tailpiece of secreted IgA; M, exon encoding the transmembrane and cytoplasmic portions of membrane-bound IgA; αs, polyadenylation site for secreted IgA mRNA; αm, polyadenylation site for membrane-bound IgA mRNA.

Germline Cα gene transcription yields a primary Iα–Sα–Cα transcript that is later spliced to form a non-coding germline Iα–Cα transcript10. The primary transcript physically associates with the template strand of the DNA to form a stable DNA–RNA hybrid32. This structure generates R loops in which the displaced non-template strand exists as a G-rich single-stranded DNA33. AID deaminates cytosine residues on both strands of the S-region DNA, thereby generating multiple DNA lesions that are ultimately processed into double-stranded DNA breaks69. Fusion of double-stranded DNA breaks at Sα and Sμ through the non-homologous end-joining pathway induces looping-out deletion of the intervening DNA, thereby juxtaposing VHDJH to Cα (REFS 69). This process yields a chromosomal VHDJH–Cα sequence, which encodes the IgA protein, and an extrachromosomal switch circle, which encodes a chimeric Iα–Cμ switch circle transcript34. Together with AID transcripts and switch circles, switch circle Iα–Cμ transcripts have a short half-life and therefore their detection indicates ongoing CSR3435.

T-cell-dependent IgA class switching

Most antigens, including microbial proteins, initiate protective humoral responses in germinal centres, which are specialized follicular environments that foster B-cell proliferation, AID expression and antibody gene diversification through CSR and SHM78. In general, germinal-centre reactions are highly dependent on cognate interactions between antigen-specific B cells and CD4 T cells that express CD40 ligand (CD40L; also known as CD154), a tumour-necrosis factor (TNF) family member that engages CD40 on B cells36. Antigen exposed on the surface of follicular DCs selects germinal-centre B cells expressing a high-affinity B-cell receptor (BCR), selected B cells thereafter differentiate into long-lived memory B cells and antibody-secreting plasma cells37. In the gut, T-cell-dependent antibody responses are strongly biased towards IgA and involve activation of B cells by antigen in the organized lymphoid tissue of Peyer’s patches, mesenteric lymph nodes and isolated lymphoid follicles12.

In addition to inducing AID, B-cell-activating signals induce germline CH gene transcription9. This process confers specificity to CSR38, because the promoter upstream of each CH gene responds only to a specific set of signal-induced transcription factors10. Promoters upstream of Cα genes, or Iα promoters, become activated in response to transforming growth factor-β1 (TGFβ1)3941, a cytokine that is secreted by many cell types, including various subsets of CD4 T cells4243. Together with CD40L, TGFβ1 is essential for the induction of T-cell-dependent IgA class switching10.

Role of TGFβ1 in germline Cα gene transcription

TGFβ1 is a pleiotropic cytokine that belongs to the TGFβ superfamily42. TGFβ1 is derived from the proteolytic cleavage of a pro-region, known as LAP (latency associated peptide), which undergoes dimerization and thereby forms an active molecule of 25 kDa42. Active TGFβ1 engages a heterotetrameric TGFβ receptor (TGFβR) complex composed of two type I and two type II transmembrane proteins that have serine/threonine kinase activity42. Signals emanating from the TGFβR induce both activating and inhibitory effects in a broad range of target cells. In B cells, low concentrations of TGFβ1 initiate Cα gene transcription (FIG. 2), whereas high concentrations suppress B-cell proliferation and differentiation, including antibody secretion44. This dual role is important for the homeostasis of the immune system. Indeed, mice with defective TGFβ1 signalling develop inflammatory and autoimmune disorders42.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f2.jpg
Signalling events leading to T-cell-dependent IgA class switching

CD4 T cells release the active transforming growth factor-β1 (TGFβ1) after processing of a latency-associated peptide (LAP). TGFβ1 forms a heteromeric TGFβ receptor (TGFβR) complex on B cells comprising TGFβRII and TGFβRI subunits. TGFβR undergoes degradation on binding by I-SMAD (inhibitory SMAD (mothers against decapentaplegic homologue)) proteins, such as SMAD7, which recruits ubiquitin ligases of the SMURF (SMAD ubiquitylation regulatory factor) family to TGFβRI. Alternatively, the TGFβR remains on the B-cell surface to activate SMAD proteins. In the presence of TGFβ1, TGFβRII kinases phosphorylate TGFβRI, leading to the activation of TGFβRI kinases. These kinases induce the phosphorylation of receptor-regulated SMAD (R-SMAD) proteins, including SMAD2 and SMAD3, thereby releasing them from the plasma membrane-anchoring protein SARA (SMAD anchor for receptor activation). After forming homo-oligomeric complexes, as well as hetero- oligomeric complexes with SMAD4 — a co-mediator SMAD (Co-SMAD) protein — R-SMAD proteins translocate to the nucleus, where they bind to SMAD-binding elements (SBEs) on target gene promoters, including constant heavy chain α (Cα) gene promoters. These SMAD complexes further associate with constitutive and TGFβR-induced co-factors, including runt-related transcription factor 3 (RUNX3), which binds to RUNX-binding elements (RBEs), cyclic AMP response element binding protein (CREB), which binds to a cyclic AMP response element (CRE), and Ets-like factor 1 (ELF1), which binds to an Ets-binding site. In addition to TGFβ, CD4 T cells express CD40 ligand (CD40L), which elicits oligomerization of CD40 on B cells, recruitment of tumour-necrosis-factor-receptor-associated factors (TRAFs) to CD40, activation of the IκB kinase (IKK) complex, phosphorylation of IκB (inhibitor of nuclear factor-κB (NF-κB)), and IκB degradation. IκB-free NF-κB translocates to the nucleus to induce the activation-induced cytidine deaminase (AICDA) gene promoter. Although NF-κB binds to an NF-κB-binding (κB) site on the Cα promoter, it has a marginal role in the transcription of the Cα gene.

TGFβR signals through mothers against decapentaplegic homologue 2 (SMAD2), SMAD3 and SMAD4 proteins, which form homo- and hetero-oligomeric complexes that bind to SMAD-binding elements (SBEs) on the promoters of target genes424547. This signalling pathway is negatively regulated by SMAD7 (REF. 42). The activation of Iα promoters requires the cooperation of SMAD proteins with runt-related transcription factor 3 (RUNX3; also known as CBFα3)454748, a TGFβ1-inducible member of the RUNX family of proteins with a DNA-binding runt domain49. SMAD proteins and RUNX3 bind to a direct repeat unit on Iα promoters that is known as the TGFβ1 responsive element (TGFβRE)50. This conserved cis-regulatory DNA region contains two tandemly arrayed RUNX-binding elements (RBEs) that are adjacent to SBEs454748. The proximity of SBEs and RBEs provides a structural correlate for the physical interaction of SMAD proteins with RUNX3, which in turn is central to the activation of Iα promoters by TGFβ1 (REFS 454748). This cytokine is important not only for the initiation of Cα gene transcription in vitro, but also for the induction of IgA class switching in vivo. Indeed, B-cell-conditional TGFβRII-deficient mice, SMAD2-deficient and SMAD3-deficient mice produce less IgA under steady-state and immunizing conditions in both systemic and mucosal sites, including the Peyer’s patches5153. Conversely, B cells from mice lacking SMAD7 show increased IgA CSR in response to TGFβ1 (REF. 54).

In addition to SBEs and RBEs, both mouse and human Iα promoters contain a cyclic AMP response element (CRE) associated with the TGFβRE50. This CRE site binds CRE binding protein (CREB), a TGFβ1-inducible factor that cooperates with SMAD and RUNX3 proteins46. Downstream of the CRE site, the mouse Iα promoter contains an Ets site, which binds ELF1 (Ets-like factor 1) and PU.1 (REF. 55). These Ets family members cooperate with SMAD and CREB proteins to activate Iα (REF. 55). Finally, near its main transcription initiation site, the mouse Iα promoter has a second CRE site and a site for PAX5 (paired box protein 5; also known as BSAP), which represses Cα gene transcription under basal conditions56.

Role of CD40L in CSR to Cα

CD40L is an essential requirement for T-cell-dependent class switching, including IgA class switching (FIG. 2), and CD40L cooperates with TGFβ1 to induce IgA CSR in vitro5761. Furthermore, CD40L- and CD40-deficient mice exhibit impaired systemic and intestinal IgA responses to T-cell-dependent stimuli in vivo6263. A significant systemic IgA deficiency can also be observed in humans affected with CD40L or CD40 signalling defects due to HIGM1 and HIGM3 syndromes, respectively6465.

CD40L expressed on the surface of antigen-activated CD4 T cells activates B cells by engaging the CD40 receptor36. By recruiting TNF-receptor-associated factor (TRAF) adaptor proteins36, CD40 forms a signalling platform that activates the IκB kinase (IKK) enzymatic complex66. This IKK complex mediates phosphorylation of inhibitor of NF-κB (IκB) proteins, which retain nuclear factor-κB (NF-κB) in a cytoplasmic inactive form66. Phosphorylation of IκBα by the IKK complex is followed by ubiquitylation and proteasome-dependent degradation of IκBα, which allows NF-κB to translocate into the nucleus66. Here, NF-κB binds to cis-regulatory κB sites, thereby determining the activation of promoters located upstream of key B-cell genes, including Iγ and Iε (REF. 10).

In addition to triggering IgG and IgE class switching10, NF-κB has an important role in IgA class switching, as this process is impaired in B cells from NF-κB-deficient mice67. Yet, Iα promoters contain only one κB site located downstream of the Ets site55. This κB site neither induces nor enhances the activation of Iα promoters55, suggesting that NF-κB regulates IgA CSR at a level other than germline Cα gene transcription. Most likely, NF-κB mediates the induction of AID expression68, an essential requirement for IgA CSR, in addition to germline Cα gene transcription9.

Remarkably, CD40L can induce IgA class switching in combination with cytokines other than TGFβ1, including interleukin-2 (IL-2), IL-4, IL-5, IL-6, IL-10 and VIP (vasoactive intestinal peptide)586061676970. These cytokines may enhance the production of endogenous TGFβ1 by B cells exposed to CD40L, thereby triggering IgA CSR through an autocrine TGFβ1-dependent loop61. In addition, they may augment the proliferation and plasma-cell differentiation of the B cells that have switched to IgA in response to CD40L and autocrine TGFβ1 (REF. 10).

Role of TGFβ1 in germline Cα gene transcription

TGFβ1 is a pleiotropic cytokine that belongs to the TGFβ superfamily42. TGFβ1 is derived from the proteolytic cleavage of a pro-region, known as LAP (latency associated peptide), which undergoes dimerization and thereby forms an active molecule of 25 kDa42. Active TGFβ1 engages a heterotetrameric TGFβ receptor (TGFβR) complex composed of two type I and two type II transmembrane proteins that have serine/threonine kinase activity42. Signals emanating from the TGFβR induce both activating and inhibitory effects in a broad range of target cells. In B cells, low concentrations of TGFβ1 initiate Cα gene transcription (FIG. 2), whereas high concentrations suppress B-cell proliferation and differentiation, including antibody secretion44. This dual role is important for the homeostasis of the immune system. Indeed, mice with defective TGFβ1 signalling develop inflammatory and autoimmune disorders42.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f2.jpg
Signalling events leading to T-cell-dependent IgA class switching

CD4 T cells release the active transforming growth factor-β1 (TGFβ1) after processing of a latency-associated peptide (LAP). TGFβ1 forms a heteromeric TGFβ receptor (TGFβR) complex on B cells comprising TGFβRII and TGFβRI subunits. TGFβR undergoes degradation on binding by I-SMAD (inhibitory SMAD (mothers against decapentaplegic homologue)) proteins, such as SMAD7, which recruits ubiquitin ligases of the SMURF (SMAD ubiquitylation regulatory factor) family to TGFβRI. Alternatively, the TGFβR remains on the B-cell surface to activate SMAD proteins. In the presence of TGFβ1, TGFβRII kinases phosphorylate TGFβRI, leading to the activation of TGFβRI kinases. These kinases induce the phosphorylation of receptor-regulated SMAD (R-SMAD) proteins, including SMAD2 and SMAD3, thereby releasing them from the plasma membrane-anchoring protein SARA (SMAD anchor for receptor activation). After forming homo-oligomeric complexes, as well as hetero- oligomeric complexes with SMAD4 — a co-mediator SMAD (Co-SMAD) protein — R-SMAD proteins translocate to the nucleus, where they bind to SMAD-binding elements (SBEs) on target gene promoters, including constant heavy chain α (Cα) gene promoters. These SMAD complexes further associate with constitutive and TGFβR-induced co-factors, including runt-related transcription factor 3 (RUNX3), which binds to RUNX-binding elements (RBEs), cyclic AMP response element binding protein (CREB), which binds to a cyclic AMP response element (CRE), and Ets-like factor 1 (ELF1), which binds to an Ets-binding site. In addition to TGFβ, CD4 T cells express CD40 ligand (CD40L), which elicits oligomerization of CD40 on B cells, recruitment of tumour-necrosis-factor-receptor-associated factors (TRAFs) to CD40, activation of the IκB kinase (IKK) complex, phosphorylation of IκB (inhibitor of nuclear factor-κB (NF-κB)), and IκB degradation. IκB-free NF-κB translocates to the nucleus to induce the activation-induced cytidine deaminase (AICDA) gene promoter. Although NF-κB binds to an NF-κB-binding (κB) site on the Cα promoter, it has a marginal role in the transcription of the Cα gene.

TGFβR signals through mothers against decapentaplegic homologue 2 (SMAD2), SMAD3 and SMAD4 proteins, which form homo- and hetero-oligomeric complexes that bind to SMAD-binding elements (SBEs) on the promoters of target genes424547. This signalling pathway is negatively regulated by SMAD7 (REF. 42). The activation of Iα promoters requires the cooperation of SMAD proteins with runt-related transcription factor 3 (RUNX3; also known as CBFα3)454748, a TGFβ1-inducible member of the RUNX family of proteins with a DNA-binding runt domain49. SMAD proteins and RUNX3 bind to a direct repeat unit on Iα promoters that is known as the TGFβ1 responsive element (TGFβRE)50. This conserved cis-regulatory DNA region contains two tandemly arrayed RUNX-binding elements (RBEs) that are adjacent to SBEs454748. The proximity of SBEs and RBEs provides a structural correlate for the physical interaction of SMAD proteins with RUNX3, which in turn is central to the activation of Iα promoters by TGFβ1 (REFS 454748). This cytokine is important not only for the initiation of Cα gene transcription in vitro, but also for the induction of IgA class switching in vivo. Indeed, B-cell-conditional TGFβRII-deficient mice, SMAD2-deficient and SMAD3-deficient mice produce less IgA under steady-state and immunizing conditions in both systemic and mucosal sites, including the Peyer’s patches5153. Conversely, B cells from mice lacking SMAD7 show increased IgA CSR in response to TGFβ1 (REF. 54).

In addition to SBEs and RBEs, both mouse and human Iα promoters contain a cyclic AMP response element (CRE) associated with the TGFβRE50. This CRE site binds CRE binding protein (CREB), a TGFβ1-inducible factor that cooperates with SMAD and RUNX3 proteins46. Downstream of the CRE site, the mouse Iα promoter contains an Ets site, which binds ELF1 (Ets-like factor 1) and PU.1 (REF. 55). These Ets family members cooperate with SMAD and CREB proteins to activate Iα (REF. 55). Finally, near its main transcription initiation site, the mouse Iα promoter has a second CRE site and a site for PAX5 (paired box protein 5; also known as BSAP), which represses Cα gene transcription under basal conditions56.

Role of CD40L in CSR to Cα

CD40L is an essential requirement for T-cell-dependent class switching, including IgA class switching (FIG. 2), and CD40L cooperates with TGFβ1 to induce IgA CSR in vitro5761. Furthermore, CD40L- and CD40-deficient mice exhibit impaired systemic and intestinal IgA responses to T-cell-dependent stimuli in vivo6263. A significant systemic IgA deficiency can also be observed in humans affected with CD40L or CD40 signalling defects due to HIGM1 and HIGM3 syndromes, respectively6465.

CD40L expressed on the surface of antigen-activated CD4 T cells activates B cells by engaging the CD40 receptor36. By recruiting TNF-receptor-associated factor (TRAF) adaptor proteins36, CD40 forms a signalling platform that activates the IκB kinase (IKK) enzymatic complex66. This IKK complex mediates phosphorylation of inhibitor of NF-κB (IκB) proteins, which retain nuclear factor-κB (NF-κB) in a cytoplasmic inactive form66. Phosphorylation of IκBα by the IKK complex is followed by ubiquitylation and proteasome-dependent degradation of IκBα, which allows NF-κB to translocate into the nucleus66. Here, NF-κB binds to cis-regulatory κB sites, thereby determining the activation of promoters located upstream of key B-cell genes, including Iγ and Iε (REF. 10).

In addition to triggering IgG and IgE class switching10, NF-κB has an important role in IgA class switching, as this process is impaired in B cells from NF-κB-deficient mice67. Yet, Iα promoters contain only one κB site located downstream of the Ets site55. This κB site neither induces nor enhances the activation of Iα promoters55, suggesting that NF-κB regulates IgA CSR at a level other than germline Cα gene transcription. Most likely, NF-κB mediates the induction of AID expression68, an essential requirement for IgA CSR, in addition to germline Cα gene transcription9.

Remarkably, CD40L can induce IgA class switching in combination with cytokines other than TGFβ1, including interleukin-2 (IL-2), IL-4, IL-5, IL-6, IL-10 and VIP (vasoactive intestinal peptide)586061676970. These cytokines may enhance the production of endogenous TGFβ1 by B cells exposed to CD40L, thereby triggering IgA CSR through an autocrine TGFβ1-dependent loop61. In addition, they may augment the proliferation and plasma-cell differentiation of the B cells that have switched to IgA in response to CD40L and autocrine TGFβ1 (REF. 10).

T-cell-independent IgA class switching

T-cell-dependent antibody responses take at least 5 to 7 days to develop, which is too much of a delay to neutralize pathogens that replicate quickly, commensal bacteria and dietary antigens. To compensate for this limitation, specialized B-cell subsets can rapidly produce IgM as well as class-switched IgG and IgA in a CD4 T-cell-and CD40L-independent manner71. In mice, intestinal T-cell-independent IgA responses rely on a peritoneal B-1-cell subset, which has ontogenic, phenotypic and genotypic features that are distinct from conventional (or B-2) B cells7274. Indeed, B-1 cells express unmutated IgA antibodies (that is, they have not been subjected to SHM) that recognize multiple specificities with low affinity12. These antibodies mediate immune exclusion of commensals and provide limited protection against some pathogens, including rotaviruses and Salmonella typhimurium7577. Systemic T-cell-independent IgA responses also exist and these appear to require B cells in the marginal zone of the spleen178. Similar to B-1 cells, mouse marginal-zone B cells express polyreactive IgA (and IgM) antibodies that may provide a second line of defence against commensal bacteria that breach the epithelial-cell barrier171.

T-cell-independent IgA responses are also present in humans, as patients with severe CD4 T-cell deficiency due to HIV infection, as well as patients lacking CD40 retain intestinal IgA class switching26. Although humans seem to lack B-1 cells, they have additional B-cell subsets that might be involved in T-cell-independent IgA responses, including IgM memory B cells79. These B cells can be detected in the circulation and in the marginal zone of the spleen, express mutated V(D)J genes, and undergo CD40-independent IgM and IgG production in response to bacterial polysaccharides, a canonical T-cell-independent antigen7980. An additional human B-cell subset that is possibly involved in T-cell-independent IgA responses is the transitional B-cell subset, which expresses polyreactive antibodies encoded by unmutated V(D)J genes8183.

T-cell-independent antigens initiate IgA class switching by linking B cells with multiple innate immune pathways. Whereas some T-cell-independent antigens, such as LPS, activate B cells through Toll-like receptors (TLRs)84, others, such as polysaccharides, activate B cells through their BCR85. T-cell-independent antigens can also provide additional B-cell-stimulating signals through DCs. Positioned as sentinels throughout the body, DCs sample T-cell-independent antigens from the environment and thereafter convey them to a non-degradative endocytic pathway86. Subsequent recycling of the endocytosed antigen to the plasma membrane is followed by its presentation to B cells8690. During this process, DCs release soluble class-switch-inducing factors related to CD40L, including B-cell activating factor (BAFF; also known as BLyS) and a proliferation-inducing ligand (APRIL)9193 (see later).

Role of microbial TLR ligands in CSR to Cα

The TLR4 ligand LPS, along with TGFβ1, can also initiate germline Cα gene transcription and CSR from Cμ to Cα in mouse B cells8494 (FIG. 3); however, the mechanism by which TLRs trigger IgA CSR in B cells remains unclear. TLRs activate NF-κB by recruiting various adaptor proteins, including myeloid differentiation primary-response protein 88 (MyD88), to their cytoplasmic tail95. MyD88 forms a signalling complex with multiple downstream elements, including IL-1-receptor-associated kinases (IRAKs) and TRAF6, thereby causing IKK activation and subsequent nuclear translocation of NF-κB96. Surprisingly, NF-κB is not required for the activation of Iα promoters, which is highly dependent on other transcription factors45485597. Instead, NF-κB may be required by TLRs to induce the expression of AID26349899.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f3.jpg
Signalling events leading to T-cell-independent IgA class switching

Dendritic cells (DCs) activate transforming growth factor-β1 (TGFβ1) by inducing the processing of a latency-associated peptide (LAP). TGFβ1 activates the constant heavy chain α (Cα) gene promoter (as shown in Figure 2). DCs also present bacterial products to B cells, thereby activating Toll-like receptors (TLRs). By recruiting myeloid differentiation primary-response protein 88 (MyD88), interleukin-1-receptor-associated kinase 1 (IRAK1) and IRAK4, TLRs induce activation of the IκB (inhibitor of nuclear factor-κB (NF-κB)) kinase (IKK) complex, phosphorylation and degradation of IκB. IκB-free NF-κB translocates to the nucleus to induce the promoter of the activation-induced cytidine deaminase (AICDA) gene. DCs further activate B cells by engaging transmembrane activator and calcium-modulating cyclophilin-ligand interactor (TACI) through B-cell-activating factor (BAFF) and a proliferation-inducing ligand (APRIL). TACI activates NF-κB after recruiting tumour-necrosis-factor-receptor-associated factors (TRAFs) to its cytoplasmic domain, thereby triggering AICDA gene expression. It is unknown whether TLRs and TACI also activate the Cα promoter. Co-SMAD, co-mediator SMAD; CRE, cyclic AMP response element; CREB, CRE-binding protein; ELF1, Ets-like factor 1; κB, NF-κB-binding site; RUNX3, runt-related transcription factor 3; RBE, RUNX-binding element; SARA, SMAD anchor for receptor activation; SBE, SMAD-binding element; SMAD, mothers against decapentaplegic homologue; TGFβR, TGFβ1 receptor.

Although sufficient to initiate IgA CSR41, TGFβ1 and LPS require additional signals, such as BCR engagement by dextran-conjugated IgD-specific antibodies, to induce significant IgA expression and secretion5878. These additional signals may be needed to optimize the expansion and differentiation of the B cells that have undergone IgA CSR in response to TGFβ1 and LPS. However, they may also be necessary to introduce crucial epigenetic changes rendering the Sα region more accessible to the CSR machinery. Indeed, mouse B cells undergo increased histone 3 acetylation at the Sα region on exposure to a range of stimuli including TGFβ1, LPS and a BCR ligand100. In addition, IgA CSR and production are enhanced by histone H3 methyltransferase Suv39h1 both in vitro and in vivo101. Suv39h1 might sequester an Sα-specific repressor by inducing relocation of proteins associated with heterochromatin, such as the polycomb group proteins HPC2 and HP1β (REF. 101). Alternatively, Suv39h1 might induce transcriptional inhibition of an Sα-specific repressor. A candidate for such a repressor is LSF (late SV40/CP2 factor), a protein that binds Sα (and Sμ) segments and inhibits IgA CSR102.

Interestingly, B-1 and marginal-zone B cells undergo IgA class switching more effectively than B-2 cells in response to T-cell-independent stimuli. This circumstance possibly reflects the unique antigen recognition profile of B-1 and marginal-zone B cells, which express both germline gene-encoded (that is, TLRs) and somatically recombined (that is, BCRs) antigen receptors103. When exposed to BAFF, LPS and TGFβ1, B-1 and marginal-zone B cells switch to IgA expression more readily than B-2 cells78. Such distinctive responsiveness may reflect the stimuli available in the microenvironments in which these B-cell subsets usually operate. Similar to B-1 and marginal-zone B cells, human IgM memory and transitional B cells exhibit robust antibody responses to microbial TLR and BCR ligands83104. Yet, the contribution of these human B-cell subsets to T-cell-independent IgA class switching is presently not known.

Role of BAFF and APRIL in CSR to Cα

In addition to engaging TLRs on B cells, microbial products stimulate the release of BAFF and APRIL by DCs (FIG. 3). These two molecules are soluble B-cell-stimulating factors that are structurally and functionally related to CD40L269199. In the presence of other cytokines, BAFF and APRIL induce germline Cα gene expression, AID expression and IgA class switching in a CD40-independent manner269199105. This effect depends on expression of the transmembrane activator and calcium-modulating cyclophilin-ligand interactor (TACI) receptor by B cells, as B cells lacking TACI do not express AID or undergo CSR in response to BAFF or APRIL105. In agreement with these data, TACI-deficient mice exhibit decreased steady-state serum IgA levels and make less IgA in response to T-cell-independent (but not T-cell-dependent) antigens106. The key role of TACI in T-cell-independent IgA responses is further indicated by its elevated expression by B-1 cells107. Similar to mice, humans develop selective IgA deficiency when TNFRSF13b108, the gene encoding TACI, is mutated (BOX 1). TNFRSF13b mutations also cause common variable immunodeficiency, which is associated with pan-hypogammaglobulinaemia108109. This more pervasive phenotype may reflect the ability of TACI to enhance CD40-dependent antibody secretion in addition to promoting CD40-independent class switching110111.

Box 1Lessons learned from IgA deficiencies

Selective IgA deficiency and common variable immunodeficiency (CVID) are the most common forms of primary immunodeficiency137. Selective IgA deficiency causes an isolated IgA defect, whereas CVID impairs IgM, IgG and IgA responses108109137. Although often asymptomatic, individuals with selective IgA deficiency and CVID can suffer from respiratory and gastrointestinal infections, respond poorly to T-cell- independent immunogens and develop B-cell lymphoproliferative disorders and autoimmunity137. In addition, some individuals with CVID develop chronic intestinal disorders, including inflammatory bowel disease, coeliac disease and nodular lymphoid hyperplasia (NLH), a follicular B-cell hyperplasia of the proximal intestine2137. These nodules are similar to isolated lymphoid follicles and can trigger bleeding and intussusception, a telescoping prolapse of the intestine into an immediate adjacent segment. Hyperstimulation of isolated lymphoid follicles by an altered microflora may have a role in NLH, as mice lacking activation-induced cytidine deaminase (AID), an enzyme essential for IgA class switching8, develop NLH-like lesions, as well as systemic and mucosal B-cell hyperactivation due to an abnormal expansion of commensal anaerobic bacteria in proximal intestinal segments31126. B-cell hyperactivation is also present in mice lacking transmembrane activator and calcium-modulating cyclophilin-ligand interactor (TACI), a class-switch-inducing receptor involved in T-cell-independent IgA responses105106. Similar to individuals with selective IgA deficiency or CVID, TACI-deficient mice develop autoimmune and lymphoproliferative disorders and produce less IgM, IgA and IgG in response to T-cell-independent antigens106. Remarkably, TACI is defective in a subset of individuals with selective IgA deficiency and CVID108109. Overall, the clinical manifestations of IgA deficiencies reflect recent evidence indicating that IgA class switching is important not only to protect against pathogens77117, but also to preserve a mutualistic host–microorganism relationship via modulation of the gene-expression profile of commensal bacteria15. By preserving the homeostasis of the gut microflora, IgA may prevent the overstimulation of mucosal and systemic B cells and the subsequent expansion of autoreactive and neoplastic clones.

BAFF and APRIL also bind to BAFF receptor (BAFFR; also known as BR3), B-cell maturation antigen (BCMA) and heparan-sulphate proteoglycans (HSPGs)9293. Engagement of BAFFR by BAFF delivers survival signals and, to some extent, CSR-inducing signals to peripheral B cells105112. By contrast, engagement of BCMA by BAFF or APRIL delivers survival signals to plasma cells113, but has no effect on CSR. Similarly, engagement of HSPGs by APRIL conveys survival and differentiation, but not CSR signals to plasma cells110. Of note, HSPGs form highly efficient TACI and BCMA signalling platforms by generating APRIL oligomers through the binding of their glycos-aminoglycan side chains to a basic QKQKKQ amino acid sequence that is proximal to the amino terminus of APRIL93.

The mechanism by which TACI triggers IgA CSR remains unclear. Similar to CD40, TACI is thought to induce TRAF-dependent activation of the IKK complex, followed by nuclear translocation of NF-κB114. This pathway may be crucial to induce AID expression68, but is unlikely to have a key role in germline Cα gene transcription, which usually requires signals from TGFβ1 (REF. 55). Yet, this cytokine is not absolutely required by BAFF and APRIL to induce IgA CSR, at least in murine B cells105, raising the possibility that TACI activates Iα promoters in a TGFβR-independent manner. Although BAFF and APRIL are sufficient to induce IgA secretion in mouse B cells105, human B cells require additional stimuli, including BCR or TLR engagement and IL-10 (REFS 269199), which enhances IgA class switching and production by functioning in synergy with paracrine or autocrine TGFβ1 (REFS 5761115). In summary, the experimental evidence presently available clearly points to the existence of an important innate pathway to IgA class switching that is highly dependent on the engagement of TACI on B cells by BAFF and APRIL. Yet, more studies are needed to elucidate the co-stimuli and signalling events required by TACI to initiate IgA CSR and production.

Role of microbial TLR ligands in CSR to Cα

The TLR4 ligand LPS, along with TGFβ1, can also initiate germline Cα gene transcription and CSR from Cμ to Cα in mouse B cells8494 (FIG. 3); however, the mechanism by which TLRs trigger IgA CSR in B cells remains unclear. TLRs activate NF-κB by recruiting various adaptor proteins, including myeloid differentiation primary-response protein 88 (MyD88), to their cytoplasmic tail95. MyD88 forms a signalling complex with multiple downstream elements, including IL-1-receptor-associated kinases (IRAKs) and TRAF6, thereby causing IKK activation and subsequent nuclear translocation of NF-κB96. Surprisingly, NF-κB is not required for the activation of Iα promoters, which is highly dependent on other transcription factors45485597. Instead, NF-κB may be required by TLRs to induce the expression of AID26349899.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f3.jpg
Signalling events leading to T-cell-independent IgA class switching

Dendritic cells (DCs) activate transforming growth factor-β1 (TGFβ1) by inducing the processing of a latency-associated peptide (LAP). TGFβ1 activates the constant heavy chain α (Cα) gene promoter (as shown in Figure 2). DCs also present bacterial products to B cells, thereby activating Toll-like receptors (TLRs). By recruiting myeloid differentiation primary-response protein 88 (MyD88), interleukin-1-receptor-associated kinase 1 (IRAK1) and IRAK4, TLRs induce activation of the IκB (inhibitor of nuclear factor-κB (NF-κB)) kinase (IKK) complex, phosphorylation and degradation of IκB. IκB-free NF-κB translocates to the nucleus to induce the promoter of the activation-induced cytidine deaminase (AICDA) gene. DCs further activate B cells by engaging transmembrane activator and calcium-modulating cyclophilin-ligand interactor (TACI) through B-cell-activating factor (BAFF) and a proliferation-inducing ligand (APRIL). TACI activates NF-κB after recruiting tumour-necrosis-factor-receptor-associated factors (TRAFs) to its cytoplasmic domain, thereby triggering AICDA gene expression. It is unknown whether TLRs and TACI also activate the Cα promoter. Co-SMAD, co-mediator SMAD; CRE, cyclic AMP response element; CREB, CRE-binding protein; ELF1, Ets-like factor 1; κB, NF-κB-binding site; RUNX3, runt-related transcription factor 3; RBE, RUNX-binding element; SARA, SMAD anchor for receptor activation; SBE, SMAD-binding element; SMAD, mothers against decapentaplegic homologue; TGFβR, TGFβ1 receptor.

Although sufficient to initiate IgA CSR41, TGFβ1 and LPS require additional signals, such as BCR engagement by dextran-conjugated IgD-specific antibodies, to induce significant IgA expression and secretion5878. These additional signals may be needed to optimize the expansion and differentiation of the B cells that have undergone IgA CSR in response to TGFβ1 and LPS. However, they may also be necessary to introduce crucial epigenetic changes rendering the Sα region more accessible to the CSR machinery. Indeed, mouse B cells undergo increased histone 3 acetylation at the Sα region on exposure to a range of stimuli including TGFβ1, LPS and a BCR ligand100. In addition, IgA CSR and production are enhanced by histone H3 methyltransferase Suv39h1 both in vitro and in vivo101. Suv39h1 might sequester an Sα-specific repressor by inducing relocation of proteins associated with heterochromatin, such as the polycomb group proteins HPC2 and HP1β (REF. 101). Alternatively, Suv39h1 might induce transcriptional inhibition of an Sα-specific repressor. A candidate for such a repressor is LSF (late SV40/CP2 factor), a protein that binds Sα (and Sμ) segments and inhibits IgA CSR102.

Interestingly, B-1 and marginal-zone B cells undergo IgA class switching more effectively than B-2 cells in response to T-cell-independent stimuli. This circumstance possibly reflects the unique antigen recognition profile of B-1 and marginal-zone B cells, which express both germline gene-encoded (that is, TLRs) and somatically recombined (that is, BCRs) antigen receptors103. When exposed to BAFF, LPS and TGFβ1, B-1 and marginal-zone B cells switch to IgA expression more readily than B-2 cells78. Such distinctive responsiveness may reflect the stimuli available in the microenvironments in which these B-cell subsets usually operate. Similar to B-1 and marginal-zone B cells, human IgM memory and transitional B cells exhibit robust antibody responses to microbial TLR and BCR ligands83104. Yet, the contribution of these human B-cell subsets to T-cell-independent IgA class switching is presently not known.

Role of BAFF and APRIL in CSR to Cα

In addition to engaging TLRs on B cells, microbial products stimulate the release of BAFF and APRIL by DCs (FIG. 3). These two molecules are soluble B-cell-stimulating factors that are structurally and functionally related to CD40L269199. In the presence of other cytokines, BAFF and APRIL induce germline Cα gene expression, AID expression and IgA class switching in a CD40-independent manner269199105. This effect depends on expression of the transmembrane activator and calcium-modulating cyclophilin-ligand interactor (TACI) receptor by B cells, as B cells lacking TACI do not express AID or undergo CSR in response to BAFF or APRIL105. In agreement with these data, TACI-deficient mice exhibit decreased steady-state serum IgA levels and make less IgA in response to T-cell-independent (but not T-cell-dependent) antigens106. The key role of TACI in T-cell-independent IgA responses is further indicated by its elevated expression by B-1 cells107. Similar to mice, humans develop selective IgA deficiency when TNFRSF13b108, the gene encoding TACI, is mutated (BOX 1). TNFRSF13b mutations also cause common variable immunodeficiency, which is associated with pan-hypogammaglobulinaemia108109. This more pervasive phenotype may reflect the ability of TACI to enhance CD40-dependent antibody secretion in addition to promoting CD40-independent class switching110111.

Box 1Lessons learned from IgA deficiencies

Selective IgA deficiency and common variable immunodeficiency (CVID) are the most common forms of primary immunodeficiency137. Selective IgA deficiency causes an isolated IgA defect, whereas CVID impairs IgM, IgG and IgA responses108109137. Although often asymptomatic, individuals with selective IgA deficiency and CVID can suffer from respiratory and gastrointestinal infections, respond poorly to T-cell- independent immunogens and develop B-cell lymphoproliferative disorders and autoimmunity137. In addition, some individuals with CVID develop chronic intestinal disorders, including inflammatory bowel disease, coeliac disease and nodular lymphoid hyperplasia (NLH), a follicular B-cell hyperplasia of the proximal intestine2137. These nodules are similar to isolated lymphoid follicles and can trigger bleeding and intussusception, a telescoping prolapse of the intestine into an immediate adjacent segment. Hyperstimulation of isolated lymphoid follicles by an altered microflora may have a role in NLH, as mice lacking activation-induced cytidine deaminase (AID), an enzyme essential for IgA class switching8, develop NLH-like lesions, as well as systemic and mucosal B-cell hyperactivation due to an abnormal expansion of commensal anaerobic bacteria in proximal intestinal segments31126. B-cell hyperactivation is also present in mice lacking transmembrane activator and calcium-modulating cyclophilin-ligand interactor (TACI), a class-switch-inducing receptor involved in T-cell-independent IgA responses105106. Similar to individuals with selective IgA deficiency or CVID, TACI-deficient mice develop autoimmune and lymphoproliferative disorders and produce less IgM, IgA and IgG in response to T-cell-independent antigens106. Remarkably, TACI is defective in a subset of individuals with selective IgA deficiency and CVID108109. Overall, the clinical manifestations of IgA deficiencies reflect recent evidence indicating that IgA class switching is important not only to protect against pathogens77117, but also to preserve a mutualistic host–microorganism relationship via modulation of the gene-expression profile of commensal bacteria15. By preserving the homeostasis of the gut microflora, IgA may prevent the overstimulation of mucosal and systemic B cells and the subsequent expansion of autoreactive and neoplastic clones.

BAFF and APRIL also bind to BAFF receptor (BAFFR; also known as BR3), B-cell maturation antigen (BCMA) and heparan-sulphate proteoglycans (HSPGs)9293. Engagement of BAFFR by BAFF delivers survival signals and, to some extent, CSR-inducing signals to peripheral B cells105112. By contrast, engagement of BCMA by BAFF or APRIL delivers survival signals to plasma cells113, but has no effect on CSR. Similarly, engagement of HSPGs by APRIL conveys survival and differentiation, but not CSR signals to plasma cells110. Of note, HSPGs form highly efficient TACI and BCMA signalling platforms by generating APRIL oligomers through the binding of their glycos-aminoglycan side chains to a basic QKQKKQ amino acid sequence that is proximal to the amino terminus of APRIL93.

The mechanism by which TACI triggers IgA CSR remains unclear. Similar to CD40, TACI is thought to induce TRAF-dependent activation of the IKK complex, followed by nuclear translocation of NF-κB114. This pathway may be crucial to induce AID expression68, but is unlikely to have a key role in germline Cα gene transcription, which usually requires signals from TGFβ1 (REF. 55). Yet, this cytokine is not absolutely required by BAFF and APRIL to induce IgA CSR, at least in murine B cells105, raising the possibility that TACI activates Iα promoters in a TGFβR-independent manner. Although BAFF and APRIL are sufficient to induce IgA secretion in mouse B cells105, human B cells require additional stimuli, including BCR or TLR engagement and IL-10 (REFS 269199), which enhances IgA class switching and production by functioning in synergy with paracrine or autocrine TGFβ1 (REFS 5761115). In summary, the experimental evidence presently available clearly points to the existence of an important innate pathway to IgA class switching that is highly dependent on the engagement of TACI on B cells by BAFF and APRIL. Yet, more studies are needed to elucidate the co-stimuli and signalling events required by TACI to initiate IgA CSR and production.

Geography of IgA class switching

By ensuring the establishment of complex commensal and symbiotic relationships, co-evolution of mammals and bacteria over the past 200 million years has contributed to the development of multiple follicular and extrafollicular layers of protection in the intestinal mucosa (FIG. 4). These layers encompass both T-cell-dependent and T-cell-independent pathways for IgA class switching and production.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f4.jpg
Map of IgA class switching in the gut

Dendritic cells (DCs) in the subepithelial dome (SED) of the Peyer’s patches capture antigen by interacting with microfold (M) cells or by extending transepithelial projections into the lumen. During this process, DCs are induced to express tumour-necrosis factor (TNF) and inducible nitric oxide synthase (iNOS) (and are therefore referred to as tiDCs), which present antigen to perifollicular CD4 T cells, thereby inducing them to differentiate into effector T cells releasing IgA-inducing cytokines. T cells also interact with antigen-specific IgMIgD naive B cells. Together with follicular dendritic cells (FDCs), this interaction fosters a germinal centre (GC) reaction that includes somatic hypermutation (SHM) and IgA class-switch recombination (CSR). The resulting IgA effector B cells home to the gut lamina propria, where they differentiate into plasma cells that secrete high-affinity IgA. Human IgA1 effector B cells can also undergo sequential IgA2 CSR on receiving T-cell-independent signals from bacteria-activated epithelial cells, DCs and tiDCs. Similar signals trigger direct IgA CSR in various B-cell subsets, including unmutated IgMIgD B-1 cells from the peritoneum and mutated IgMIgD effector B cells from Peyer’s patches. These local CSR events generate plasma cells secreting low- or high-affinity IgA. FAE, follicle-associated epithelium; FM, follicular mantle; HEV, high endothelial venule; sIgA, secreted IgA.

T-cell-dependent IgA CSR in Peyer’s patches

Prior studies have confirmed that the intestinal IgA inductive sites are the Peyer’s patches, isolated lymphoid follicles and mesenteric lymph nodes12116. The key role of these organized lymphoid structures in the induction of intestinal IgA was demonstrated in mice that have abnormal follicles as a result of a genetic manipulation or experimental disruption of the signalling pathways required for lymphoid organogenesis12117. In general, Peyer’s patches are thought to generate high-affinity IgA antibodies to toxins and pathogens through a canonical T-cell-dependent pathway that is orchestrated by DCs. Positioned in the subepithelial dome of Peyer’s patches (FIG. 5), DCs capture antigen from the intestinal lumen either directly by extending transepithelial projections or indirectly via M cells19. After migrating into the perifollicular area of the Peyer’s patches, DCs present the captured antigen to CD4 T cells, thereby inducing effector T cells that release IgA-inducing cytokines, including TGFβ1, IL-4, IL-6 and IL-10 (REFS 117119). This response is enhanced by thymic stromal lymphopoietin (TSLP), a DC-conditioning IL-7-like cytokine derived from epithelial cells that promotes the formation of non-inflammatory T cells with IgA-inducing functions119120. Ultimately, these non-inflammatory effector T cells trigger CSR from Cμ to Cα through a CD40-dependent pathway involving cognate T-cell–B-cell interactions in the germinal centre of mucosa-associated lymphoid follicles2.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f5.jpg
Cellular interactions causing IgA class switching in the gut

a | While capturing antigen, intestinal dendritic cells (DCs) are exposed to microbial Toll-like receptor (TLR) ligands and epithelial-cell-derived cytokines, including thymic stromal lymphopoietin (TSLP). These signals promote the generation of tiDCs, which are DCs that express tumour-necrosis factor and inducible nitric oxide synthase; these cells present antigen to CD4 T cells in the perifollicular area of Peyer’s patches. In addition, tiDCs transfer antigen to follicular IgMIgD naive B cells and induce them to upregulate the expression of TGFβ1 receptor (TGFβR) through nitric oxide (NO). During cognate interactions with CD4 T cells, B cells undergo IgA class-switch recombination (CSR) in response to CD40 ligand (CD40L) and transforming growth factor-β1 (TGFβ1) from activated T cells. IgA expression requires interleukin-5 (IL-5), IL-6 and IL-10 from activated T cells, as well as B-cell-activating factor (BAFF) and a proliferation-inducing ligand (APRIL) from tiDCs. After being imprinted by retinoic acid (RA) from DCs, IgA effector B cells migrate to the lamina propria, where they differentiate into IgA-secreting plasma cells. This differentiation is enhanced by APRIL secreted by epithelial cells, DCs and tiDCs. b | After sensing microorganisms via TLRs, epithelial cells from intestinal villi release APRIL, thereby triggering direct IgA CSR in lamina-propria IgM B cells and sequential IgA2 CSR in lamina-propria IgA1 B cells in a T-cell-independent manner. This pathway may also involve APRIL, BAFF and TGFβ1 from DCs and tiDCs exposed to microbial TLR ligands, NO and TSLP. Epithelial cells, DCs and tiDCs promote plasma-cell differentiation via APRIL, BAFF, IL-6 and IL-10. BCMA, B-cell maturation antigen; FAE, follicle-associated epithelium; M cell, microfold cell; pIgR, polymeric immunoglobulin receptor; SED, subepithelial dome; sIgA, secreted IgA; TACI, transmembrane activator and calcium-modulating cyclophilin-ligand interactor.

In addition to triggering the inductive phase of a T-cell-dependent IgA response, mucosal DCs initiate the effector phase of the mucosal humoral response by releasing retinoic acid, a compound derived from vitamin A that upregulates the expression of gut-homing receptors, including α4β7-integrin and CC-chemokine receptor 9 (CCR9), by IgA class-switched B cells121. This upregulation enables IgA B cells to migrate from Peyer’s patches to the gut lamina propria via the draining mesenteric lymph nodes, the thoracic duct and blood121. During this migration, IgA B cells develop into IgA-secreting plasma cells under the influence of cytokines released by lymphoid, stromal and epithelial cells and by DCs. The reason why class switching is biased towards IgA in the Peyer’s patches remains unclear, although the release of IL-6, IL-10, TGFβ1, retinoic acid and nitric oxide by gut DCs seems to have an important role119121123.

Peyer’s patches can also induce IgA class switching through an alternative T-cell-dependent pathway that does not require BCR signals from canonical cognate T-cell–B-cell interactions, but that rather relies on TLR signals from commensal bacteria124. This observation would explain why a fraction of B-2 cells from Peyer’s patches expresses a restricted IgA repertoire that does not show the sequential acquisition of somatic hypermutations that would be expected in a canonical germinal-centre reaction125. By using both canonical and alternative T-cell-dependent pathways, Peyer’s patches may generate IgA antibodies to control both commensal and pathogenic bacteria1117126.

T-cell-independent IgA CSR in the lamina propria

In mice, T-cell-independent IgA production requires B-1 cells and preferentially targets commensal bacteria3. Indeed, experiments in which radiation chimaeras have been generated with allotypic markers to distinguish the IgA derived from B-1 and B-2 cells indicate that up to 50% of the intestinal IgA originates from B-1 cells73127, although this point remains controversial128. Most of this IgA is produced in a T-cell-independent manner, because MHC-class-II-deficient mice (lacking cognate T-cell–B-cell interactions) or mice deficient for the T-cell receptor β- and γ-chains (lacking T cells) retain B-1-cell-mediated IgA responses to commensal bacteria7476. Recent data indicate that also the human intestine supports T-cell-independent IgA production26.

But where do B cells undergo T-cell-independent IgA class switching? One probable place is the intestinal lamina propria, as IgA B cells from this site contain hallmarks of ongoing IgA CSR, including AICDA gene transcripts, AID protein and Iα–Cμ switch circle transcripts226129. In humans, these indicators are still present in the absence of CD4 T cells, CD40 signalling and/or germinal centres26, suggesting that lamina-propria B cells can undergo IgA CSR in situ in a T-cell-independent manner. Consistent with this possibility, lamina-propria B cells retain IgA CSR in an AID mouse reporter strain that lacks germinal centres as a result of a deletion of the gene for the transcription factor OCA-B (Oct co-activator from B cells; also known as OBF1)129. Although still debated130, the presence of IgA CSR in the lamina propria is in agreement with compelling evidence showing that CSR is not restricted to lymphoid follicles, but also occurs in extrafollicular lymphoid areas131, including subepithelial areas99. In both mice and humans, lamina-propria-derived IgA CSR-inducing signals may target multiple IgM B-cell subsets that originate from various sites, including the peritoneum, mucosal follicles and bone marrow263173129132. A full characterization of these IgM B-cell subsets is made more difficult by the fact that they may rapidly modify their phenotype on entering the microenvironment of the lamina propria.

Role of DCs in CSR to Cα

How do lamina-propria B cells undergo T-cell-independent IgA CSR in response to T-cell-independent antigens? In humans, DCs acquire B-cell-licensing functions, including IgA-inducing activity, following stimulation by microbial TLR ligands (FIG. 5), such as LPS91115. In mice, intestinal DCs initiate T-cell-independent IgA responses by activating B-1 cells after loading commensal bacteria7476. Consistent with these findings, DCs in the intestinal lamina propria continuously sample antigens from the intestinal lumen through transepithelial projections133134. This sampling activity probably results in the induction of IgA CSR-inducing factors, such as BAFF and APRIL269199, and in the presentation of T-cell-independent antigens to B cells8690. Of note, bacterial TLR ligands have recently been shown to generate a mucosal DC subset that expresses TNF and inducible nitric oxide synthase (iNOS) and has potent IgA-inducing functions123. Nitric oxide can enhance T-cell-independent IgA class switching by stimulating BAFF and APRIL production by lamina-propria DCs and enhance T-cell-dependent IgA class switching by upregulating TGFβR expression by B cells from the Peyer’s patches123. In this way, TNFiNOS DCs would contribute to the striking IgA bias observed in the gut.

Role of epithelial cells in CSR to Cα

DCs are unlikely to be the only inducers of T-cell-independent antibody responses in the gut135. Positioned at the interface between the antigen-rich intestinal lumen and the B-cell-rich lamina propria, epithelial cells produce numerous mediators with IgA-inducing function, including IL-10 and TGFβ1. Similar to respiratory epithelial cells99136, human colonocytes also express BAFF and APRIL (FIG. 5) and further upregulate this expression on sensing bacteria via TLRs through a mechanism that involves DC activation by TSLP2699. Together with IL-10, APRIL triggers IgA2 CSR in B cells26, suggesting that epithelial cells are central to the induction of IgA2 at mucosal sites colonized by a large microbiota, such as the colon2425. At these sites, IgA2 may be more beneficial than IgA1, perhaps because IgA2 is more resistant than IgA1 to enzymatic digestion by bacterial proteases2528.

The presence of local IgA2 CSR in the human colonic lamina propria is consistent with several observations. First, the colonic lamina propria contains higher levels of IgA2 than IgA1, whereas colonic lymphoid follicles contain more IgA1 than IgA2 (REF. 26), suggesting that both inductive and migratory events contribute to IgA2 production in the colonic lamina propria. Second, the colonic lamina propria retains IgA2, as well as hallmarks of ongoing CSR, in the absence of functional CD4 T cells, CD40 signals and germinal centres26, which indicates that IgA2 responses proceed locally in a T-cell-independent manner. Third, the colonic lamina propria contains traces of both direct IgM-to-IgA2 and sequential IgA1-to-IgA2 CSR events, which are both highly dependent on APRIL26. Although direct IgA2 CSR can take place in unmutated or mutated lamina-propria IgM B cells originating from systemic and mucosal sites, sequential IgA2 CSR would occur in mutated lamina-propria IgA1 B cells arriving from mucosal follicles. In this way, epithelial cells would maximize the diversity of the IgA2 repertoire released onto the mucosal surface. Finally, APRIL may enable epithelial cells to optimize mucosal IgA2 secretion by enhancing plasma-cell survival and differentiation110113.

Altogether, the available evidence indicates that human intestinal epithelial cells use APRIL to trigger T-cell-independent production of protease-resistant IgA2 antibodies, which may be more suited than IgA1 antibodies to cope with the dense microbial community of the large intestine. Consistent with its ability to induce IgA1 in addition to IgA2 class switching26, APRIL is also expressed by epithelial cells at mucosal sites with predominant IgA1 production, such as the respiratory tract2326136. This implies that APRIL influences the IgA1–IgA2 balance in the context of other unknown factors that are probably specific to each mucosal site.

T-cell-dependent IgA CSR in Peyer’s patches

Prior studies have confirmed that the intestinal IgA inductive sites are the Peyer’s patches, isolated lymphoid follicles and mesenteric lymph nodes12116. The key role of these organized lymphoid structures in the induction of intestinal IgA was demonstrated in mice that have abnormal follicles as a result of a genetic manipulation or experimental disruption of the signalling pathways required for lymphoid organogenesis12117. In general, Peyer’s patches are thought to generate high-affinity IgA antibodies to toxins and pathogens through a canonical T-cell-dependent pathway that is orchestrated by DCs. Positioned in the subepithelial dome of Peyer’s patches (FIG. 5), DCs capture antigen from the intestinal lumen either directly by extending transepithelial projections or indirectly via M cells19. After migrating into the perifollicular area of the Peyer’s patches, DCs present the captured antigen to CD4 T cells, thereby inducing effector T cells that release IgA-inducing cytokines, including TGFβ1, IL-4, IL-6 and IL-10 (REFS 117119). This response is enhanced by thymic stromal lymphopoietin (TSLP), a DC-conditioning IL-7-like cytokine derived from epithelial cells that promotes the formation of non-inflammatory T cells with IgA-inducing functions119120. Ultimately, these non-inflammatory effector T cells trigger CSR from Cμ to Cα through a CD40-dependent pathway involving cognate T-cell–B-cell interactions in the germinal centre of mucosa-associated lymphoid follicles2.

An external file that holds a picture, illustration, etc.
Object name is nihms276719f5.jpg
Cellular interactions causing IgA class switching in the gut

a | While capturing antigen, intestinal dendritic cells (DCs) are exposed to microbial Toll-like receptor (TLR) ligands and epithelial-cell-derived cytokines, including thymic stromal lymphopoietin (TSLP). These signals promote the generation of tiDCs, which are DCs that express tumour-necrosis factor and inducible nitric oxide synthase; these cells present antigen to CD4 T cells in the perifollicular area of Peyer’s patches. In addition, tiDCs transfer antigen to follicular IgMIgD naive B cells and induce them to upregulate the expression of TGFβ1 receptor (TGFβR) through nitric oxide (NO). During cognate interactions with CD4 T cells, B cells undergo IgA class-switch recombination (CSR) in response to CD40 ligand (CD40L) and transforming growth factor-β1 (TGFβ1) from activated T cells. IgA expression requires interleukin-5 (IL-5), IL-6 and IL-10 from activated T cells, as well as B-cell-activating factor (BAFF) and a proliferation-inducing ligand (APRIL) from tiDCs. After being imprinted by retinoic acid (RA) from DCs, IgA effector B cells migrate to the lamina propria, where they differentiate into IgA-secreting plasma cells. This differentiation is enhanced by APRIL secreted by epithelial cells, DCs and tiDCs. b | After sensing microorganisms via TLRs, epithelial cells from intestinal villi release APRIL, thereby triggering direct IgA CSR in lamina-propria IgM B cells and sequential IgA2 CSR in lamina-propria IgA1 B cells in a T-cell-independent manner. This pathway may also involve APRIL, BAFF and TGFβ1 from DCs and tiDCs exposed to microbial TLR ligands, NO and TSLP. Epithelial cells, DCs and tiDCs promote plasma-cell differentiation via APRIL, BAFF, IL-6 and IL-10. BCMA, B-cell maturation antigen; FAE, follicle-associated epithelium; M cell, microfold cell; pIgR, polymeric immunoglobulin receptor; SED, subepithelial dome; sIgA, secreted IgA; TACI, transmembrane activator and calcium-modulating cyclophilin-ligand interactor.

In addition to triggering the inductive phase of a T-cell-dependent IgA response, mucosal DCs initiate the effector phase of the mucosal humoral response by releasing retinoic acid, a compound derived from vitamin A that upregulates the expression of gut-homing receptors, including α4β7-integrin and CC-chemokine receptor 9 (CCR9), by IgA class-switched B cells121. This upregulation enables IgA B cells to migrate from Peyer’s patches to the gut lamina propria via the draining mesenteric lymph nodes, the thoracic duct and blood121. During this migration, IgA B cells develop into IgA-secreting plasma cells under the influence of cytokines released by lymphoid, stromal and epithelial cells and by DCs. The reason why class switching is biased towards IgA in the Peyer’s patches remains unclear, although the release of IL-6, IL-10, TGFβ1, retinoic acid and nitric oxide by gut DCs seems to have an important role119121123.

Peyer’s patches can also induce IgA class switching through an alternative T-cell-dependent pathway that does not require BCR signals from canonical cognate T-cell–B-cell interactions, but that rather relies on TLR signals from commensal bacteria124. This observation would explain why a fraction of B-2 cells from Peyer’s patches expresses a restricted IgA repertoire that does not show the sequential acquisition of somatic hypermutations that would be expected in a canonical germinal-centre reaction125. By using both canonical and alternative T-cell-dependent pathways, Peyer’s patches may generate IgA antibodies to control both commensal and pathogenic bacteria1117126.

T-cell-independent IgA CSR in the lamina propria

In mice, T-cell-independent IgA production requires B-1 cells and preferentially targets commensal bacteria3. Indeed, experiments in which radiation chimaeras have been generated with allotypic markers to distinguish the IgA derived from B-1 and B-2 cells indicate that up to 50% of the intestinal IgA originates from B-1 cells73127, although this point remains controversial128. Most of this IgA is produced in a T-cell-independent manner, because MHC-class-II-deficient mice (lacking cognate T-cell–B-cell interactions) or mice deficient for the T-cell receptor β- and γ-chains (lacking T cells) retain B-1-cell-mediated IgA responses to commensal bacteria7476. Recent data indicate that also the human intestine supports T-cell-independent IgA production26.

But where do B cells undergo T-cell-independent IgA class switching? One probable place is the intestinal lamina propria, as IgA B cells from this site contain hallmarks of ongoing IgA CSR, including AICDA gene transcripts, AID protein and Iα–Cμ switch circle transcripts226129. In humans, these indicators are still present in the absence of CD4 T cells, CD40 signalling and/or germinal centres26, suggesting that lamina-propria B cells can undergo IgA CSR in situ in a T-cell-independent manner. Consistent with this possibility, lamina-propria B cells retain IgA CSR in an AID mouse reporter strain that lacks germinal centres as a result of a deletion of the gene for the transcription factor OCA-B (Oct co-activator from B cells; also known as OBF1)129. Although still debated130, the presence of IgA CSR in the lamina propria is in agreement with compelling evidence showing that CSR is not restricted to lymphoid follicles, but also occurs in extrafollicular lymphoid areas131, including subepithelial areas99. In both mice and humans, lamina-propria-derived IgA CSR-inducing signals may target multiple IgM B-cell subsets that originate from various sites, including the peritoneum, mucosal follicles and bone marrow263173129132. A full characterization of these IgM B-cell subsets is made more difficult by the fact that they may rapidly modify their phenotype on entering the microenvironment of the lamina propria.

Role of DCs in CSR to Cα

How do lamina-propria B cells undergo T-cell-independent IgA CSR in response to T-cell-independent antigens? In humans, DCs acquire B-cell-licensing functions, including IgA-inducing activity, following stimulation by microbial TLR ligands (FIG. 5), such as LPS91115. In mice, intestinal DCs initiate T-cell-independent IgA responses by activating B-1 cells after loading commensal bacteria7476. Consistent with these findings, DCs in the intestinal lamina propria continuously sample antigens from the intestinal lumen through transepithelial projections133134. This sampling activity probably results in the induction of IgA CSR-inducing factors, such as BAFF and APRIL269199, and in the presentation of T-cell-independent antigens to B cells8690. Of note, bacterial TLR ligands have recently been shown to generate a mucosal DC subset that expresses TNF and inducible nitric oxide synthase (iNOS) and has potent IgA-inducing functions123. Nitric oxide can enhance T-cell-independent IgA class switching by stimulating BAFF and APRIL production by lamina-propria DCs and enhance T-cell-dependent IgA class switching by upregulating TGFβR expression by B cells from the Peyer’s patches123. In this way, TNFiNOS DCs would contribute to the striking IgA bias observed in the gut.

Role of epithelial cells in CSR to Cα

DCs are unlikely to be the only inducers of T-cell-independent antibody responses in the gut135. Positioned at the interface between the antigen-rich intestinal lumen and the B-cell-rich lamina propria, epithelial cells produce numerous mediators with IgA-inducing function, including IL-10 and TGFβ1. Similar to respiratory epithelial cells99136, human colonocytes also express BAFF and APRIL (FIG. 5) and further upregulate this expression on sensing bacteria via TLRs through a mechanism that involves DC activation by TSLP2699. Together with IL-10, APRIL triggers IgA2 CSR in B cells26, suggesting that epithelial cells are central to the induction of IgA2 at mucosal sites colonized by a large microbiota, such as the colon2425. At these sites, IgA2 may be more beneficial than IgA1, perhaps because IgA2 is more resistant than IgA1 to enzymatic digestion by bacterial proteases2528.

The presence of local IgA2 CSR in the human colonic lamina propria is consistent with several observations. First, the colonic lamina propria contains higher levels of IgA2 than IgA1, whereas colonic lymphoid follicles contain more IgA1 than IgA2 (REF. 26), suggesting that both inductive and migratory events contribute to IgA2 production in the colonic lamina propria. Second, the colonic lamina propria retains IgA2, as well as hallmarks of ongoing CSR, in the absence of functional CD4 T cells, CD40 signals and germinal centres26, which indicates that IgA2 responses proceed locally in a T-cell-independent manner. Third, the colonic lamina propria contains traces of both direct IgM-to-IgA2 and sequential IgA1-to-IgA2 CSR events, which are both highly dependent on APRIL26. Although direct IgA2 CSR can take place in unmutated or mutated lamina-propria IgM B cells originating from systemic and mucosal sites, sequential IgA2 CSR would occur in mutated lamina-propria IgA1 B cells arriving from mucosal follicles. In this way, epithelial cells would maximize the diversity of the IgA2 repertoire released onto the mucosal surface. Finally, APRIL may enable epithelial cells to optimize mucosal IgA2 secretion by enhancing plasma-cell survival and differentiation110113.

Altogether, the available evidence indicates that human intestinal epithelial cells use APRIL to trigger T-cell-independent production of protease-resistant IgA2 antibodies, which may be more suited than IgA1 antibodies to cope with the dense microbial community of the large intestine. Consistent with its ability to induce IgA1 in addition to IgA2 class switching26, APRIL is also expressed by epithelial cells at mucosal sites with predominant IgA1 production, such as the respiratory tract2326136. This implies that APRIL influences the IgA1–IgA2 balance in the context of other unknown factors that are probably specific to each mucosal site.

Concluding remarks

Despite recent advances, the functions and mechanisms of mucosal IgA class switching remain unclear. Commensal communities are strikingly different at distinct mucosal sites and may influence the subclass and binding properties of IgA class-switched antibodies in a site-specific manner. However, little is known about the role of individual bacterial species in the induction of IgA class switching and about the mechanisms by which class-switched IgA antibodies modify the composition and mutualistic relationship between the commensal microbiota and the host. We also need more information on the phenotype, IgA class-switch-inductive properties, and activation requirements of individual DC and B-cell subsets at distinct mucosal sites. Furthermore, more data are needed on the follicular and extrafollicular pathways that mediate mucosal IgA class switching and on the role of these pathways in the preservation of mucosal homeostasis and in the generation of immune protection against mucosal pathogens (BOX 2). Addressing these questions will not only help to develop more efficient mucosal vaccines, but also to prevent and treat mucosal inflammatory disorders.

Box 2Mucosal vaccines

Mucosal immune responses provide a first line of defence against infectious agents that use mucosal surfaces as a portal site of entry. Growing evidence indicates that the generation of mucosal immunity requires topical immunization19138. Indeed, intramuscular or subcutaneous vaccines induce poor mucosal immune protection compared to oral, nasal, vaginal or rectal vaccines. Unfortunately, only a few mucosal vaccines are currently available and these include oral vaccines against poliovirus, Salmonella typhimurium, Vibrio cholerae and rotavirus, as well as an inhalable vaccine against influenza virus19138. The development of mucosal vaccines has lagged behind that of other vaccines owing to a limited knowledge of mucosal immunity and to the technical challenges associated with the measurement of correlates of mucosal protection, including IgA antibodies in mucosal secretions. Additional challenges relate to the fact that mucosal vaccines are diluted by mucosal secretions, captured by mucus gels, attacked by enzymes and excluded by epithelial-cell barriers. Furthermore, in the absence of adequate adjuvants, mucosal antigens tend to induce tolerance rather than immunity19138. Hence, an effective mucosal vaccine against infectious agents should avoid physical elimination and enzymatic digestion, target mucosal inductive sites and stimulate the innate immune system to generate effective adaptive immune responses19138. Stimulated by the threat posed by HIV, current research efforts aim at identifying new mucosal delivery systems and adjuvants. Whether mucosally transmitted or injected, HIV preferentially replicates in mucosal districts, where it establishes viral reservoirs. Therefore, consensus is growing for the notion that effective HIV vaccines should elicit both cellular and humoral immune responses in mucosal regions19138. Secreted IgA, with or without broad neutralizing activity, may prevent HIV from contacting mucosal surfaces, adhering to epithelial cells or crossing the epithelial-cell barrier139. Indeed, there is some evidence from highly exposed, uninfected individuals that IgA in mucosal secretions increases resistance to sexually transmitted HIV infection140.

Acknowledgments

The author is supported by US National Institutes of Health (NIH) research grants AI057653, {"type":"entrez-nucleotide","attrs":{"text":"AI057653","term_id":"3331519","term_text":"AI057653"}}AI057653-S1 and {"type":"entrez-nucleotide","attrs":{"text":"AI074378","term_id":"3401022","term_text":"AI074378"}}AI074378, and by an Irma T. Hirschl Career Scientist Award.

Andrea Cerutti, Department of Pathology and Laboratory Medicine, Weill Medical College of Cornell University, and Weill Graduate School of Medical Sciences of Cornell University, 1300 York Avenue, New York, New York 10021, USA;
Andrea Cerutti: ude.llenroc.dem@ittureca
Andrea Cerutti: ude.llenroc.dem@ittureca

Abstract

IgA class switching is the process whereby B cells acquire the expression of IgA, the most abundant antibody isotype in mucosal secretions. IgA class switching occurs via both T-cell-dependent and T-cell-independent pathways, and the antibody targets both pathogenic and commensal microorganisms. This Review describes recent advances indicating that innate immune recognition of microbial signatures at the epithelial-cell barrier is central to the selective induction of mucosal IgA class switching. In addition, the mechanisms of IgA class switching at follicular and extrafollicular sites within the mucosal environment are summarized. A better understanding of these mechanisms may help in the development of more effective mucosal vaccines.

Abstract

IgA has been selected throughout evolution to provide a first line of immune protection at mucosal surfaces — vulnerable frontline sites that are exposed to potentially harmful commensal, airborne, ingested and sexually transmitted agents. Growing evidence indicates that IgA uses a high-affinity binding system to neutralize microbial toxins and pathogens, and a low-affinity binding system to prevent commensal bacteria from breaching the mucosal surface1. This latter process is known as immune exclusion and has a fundamental role in the intestine, which is home to a number of commensal bacteria exceeding that of human cells by an estimated order of magnitude2.

Remarkably, intestinal IgA achieves both immune protection and immune exclusion in a non-inflammatory manner, thereby promoting the establishment of a sustainable host–microbial mutualism3. The complex relationship between IgA and the intestinal microbiota is further exemplified by the fact that IgA responses are highly dependent on intestinal colonization by commensal microorganisms. Indeed, the number of IgA-secreting B cells is dramatically reduced in the intestine of germ-free animals and these cells are virtually absent in neonates before their exposure to bacteria3.

In this Review, I summarize recent advances in our understanding of the function, regulation and geography of IgA class switching. In addition to analysing the signalling pathways underlying IgA class switching, I discuss new evidence indicating that commensal bacteria regulate intestinal IgA responses by promoting the crosstalk between B cells and multiple components of the mucosal innate immune system, including epithelial cells and dendritic cells (DCs).

Footnotes

Competing interests statement

The author declares no competing financial interests.

DATABASES

Entrez Gene:http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?db=gene

AID|APRIL|BAFF|BAFFR|BCMA|CD40L|TACI|TGFβ1|TSLP

FURTHER INFORMATION

Andrea Cerutti’s homepage:http://www.med.cornell.edu/research/acerutti/index.html

ALL LINKS ARE ACTIVE IN THE ONLINE PDF

Footnotes

References

  • 1. Macpherson AJ, McKoy KD, Johansen FE, Brandtzaeg PThe immune geography of IgA induction and function. Mucosal Immunol. 2008;1:11–22. An overview of the function and topography of IgA responses and the mechanisms of their induction. [[PubMed][Google Scholar]
  • 2. Fagarasan S, Honjo TIntestinal IgA synthesis: regulation of front-line body defences. Nature Rev Immunol. 2003;3:63–72.[PubMed][Google Scholar]
  • 3. Macpherson AJ, Harris NLInteractions between commensal intestinal bacteria and the immune system. Nature Rev Immunol. 2004;4:478–485.[PubMed][Google Scholar]
  • 4. Schlissel MSRegulating antigen-receptor gene assembly. Nature Rev Immunol. 2003;3:890–899.[PubMed][Google Scholar]
  • 5. Odegard VH, Schatz DGTargeting of somatic hypermutation. Nature Rev Immunol. 2006;6:573–583.[PubMed][Google Scholar]
  • 6. Stavnezer J, Guikema JE, Schrader CEMechanism and regulation of class switch recombination. Annu Rev Immunol. 2008;26:261–292. An updated overview of the molecular pathways of CSR. [Google Scholar]
  • 7. MacLennan ICGerminal centers. Annu Rev Immunol. 1994;12:117–139.[PubMed][Google Scholar]
  • 8. Muramatsu M, Nagaoka H, Shinkura R, Begum NA, Honjo TDiscovery of activation-induced cytidine deaminase, the engraver of antibody memory. Adv Immunol. 2007;94:1–36.[PubMed][Google Scholar]
  • 9. Chaudhuri J, Alt FWClass-switch recombination: interplay of transcription, DNA deamination and DNA repair. Nature Rev Immunol. 2004;4:541–552.[PubMed][Google Scholar]
  • 10. Stavnezer JAntibody class switching. Adv Immunol. 1996;61:79–146.[PubMed][Google Scholar]
  • 11. Mestecky J, Zikan J, Butler WTImmunoglobulin M and secretory immunoglobulin A: presence of a common polypeptide chain different from light chains. Science. 1971;171:1163–1165.[PubMed][Google Scholar]
  • 12. Mostov KE, Deitcher DLPolymeric immunoglobulin receptor expressed in MDCK cells transcytoses IgA. Cell. 1986;46:613–621.[PubMed][Google Scholar]
  • 13. Brandtzaeg PMucosal and glandular distribution of immunoglobulin components: differential localization of free and bound SC in secretory epithelial cells. J Immunol. 1974;112:1553–1559.[PubMed][Google Scholar]
  • 14. Phalipon A, et al Secretory component: a new role in secretory IgA-mediated immune exclusion in vivo. Immunity. 2002;17:107–115. An elegant study showing that the secretory component facilitates the exclusion of commensal bacteria from the mucosal surface by increasing the mucophilic properties of class-switched IgA antibodies. [[PubMed][Google Scholar]
  • 15. Peterson DA, McNulty NP, Guruge JL, Gordon JIIgA response to symbiotic bacteria as a mediator of gut homeostasis. Cell Host Microbe. 2007;2:328–339.[PubMed][Google Scholar]
  • 16. Fernandez MI, et al Anti-inflammatory role for intracellular dimeric immunoglobulin A by neutralization of lipopolysaccharide in epithelial cells. Immunity. 2003;18:739–749.[PubMed][Google Scholar]
  • 17. Bollinger RR, et al Secretory IgA and mucin-mediated biofilm formation by environmental strains of Escherichia coli: role of type 1 pili. Mol Immunol. 2006;43:378–387.[PubMed][Google Scholar]
  • 18. Favre L, Spertini F, Corthesy BSecretory IgA possesses intrinsic modulatory properties stimulating mucosal and systemic immune responses. J Immunol. 2005;175:2793–2800. References 15–18 describe some of the mechanisms through which sIgA antibodies promote gut homeostasis and immune protection. [[PubMed][Google Scholar]
  • 19. Neutra MR, Kozlowski PAMucosal vaccines: the promise and the challenge. Nature Rev Immunol. 2006;6:148–158.[PubMed][Google Scholar]
  • 20. Monteiro RC, Van De Winkel JGIgA Fc receptors. Annu Rev Immunol. 2003;21:177–204.[PubMed][Google Scholar]
  • 21. van Egmond M, et al FcαRI-positive liver Kupffer cells: reappraisal of the function of immunoglobulin A in immunity. Nature Med. 2000;6:680–685.[PubMed][Google Scholar]
  • 22. Pasquier B, et al Identification of FcαRI as an inhibitory receptor that controls inflammation: dual role of FcRγ ITAM. Immunity. 2005;22:31–42. References 21 and 22 show that systemic IgA proteins provide a second line of defence against intestinal bacteria by binding to the FcαRI on phagocytes. [[PubMed][Google Scholar]
  • 23. Woof JM, Mestecky JMucosal immunoglobulins. Immunol Rev. 2005;206:64–82.[PubMed][Google Scholar]
  • 24. Kett K, Brandtzaeg P, Radl J, Haaijman JJDifferent subclass distribution of IgA-producing cells in human lymphoid organs and various secretory tissues. J Immunol. 1986;136:3631–3635.[PubMed][Google Scholar]
  • 25. Kett K, et al Intestinal B-cell isotype response in relation to local bacterial load: evidence for immunoglobulin A subclass adaptation. Gastroenterology. 1995;109:819–825.[PubMed][Google Scholar]
  • 26. He B, et al Intestinal bacteria trigger T cell-independent immunoglobulin A2 class switching by inducing epithelial-cell secretion of the cytokine APRIL. Immunity. 2007;26:812–826. This study links intestinal IgA2 class switching with epithelial-cell and DC production of APRIL in response to commensal bacteria. [[PubMed][Google Scholar]
  • 27. Flanagan JG, Lefranc MP, Rabbitts THMechanisms of divergence and convergence of the human immunoglobulin α1 and α2 constant region gene sequences. Cell. 1984;36:681–688.[PubMed][Google Scholar]
  • 28. Plaut AG, Wistar R, Jr, Capra JDDifferential susceptibility of human IgA immunoglobulins to streptococcal IgA protease. J Clin Invest. 1974;54:1295–1300.[Google Scholar]
  • 29. Muramatsu M, et al Class switch recombination and hypermutation require activation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell. 2000;102:553–563. The first demonstration that AID is essential for the induction of CSR in mouse and human B cells. [[PubMed][Google Scholar]
  • 30. Revy P, et al Activation-induced cytidine deaminase (AID) deficiency causes the autosomal recessive form of the Hyper-IgM syndrome (HIGM2) Cell. 2000;102:565–575.[PubMed][Google Scholar]
  • 31. Fagarasan S, Kinoshita K, Muramatsu M, Ikuta K, Honjo T. In situ class switching and differentiation to IgA-producing cells in the gut lamina propria. Nature. 2001;413:639–643. The first paper showing that IgA class switching is not restricted to canonical intestinal inductive sites, such as the Peyer’s patches, but can also occur in the lamina propria. [[PubMed]
  • 32. Reaban ME, Griffin JAInduction of RNA-stabilized DNA conformers by transcription of an immunoglobulin switch region. Nature. 1990;348:342–344.[PubMed][Google Scholar]
  • 33. Yu K, Chedin F, Hsieh CL, Wilson TE, Lieber MRR-loops at immunoglobulin class switch regions in the chromosomes of stimulated B cells. Nature Immunol. 2003;4:442–451.[PubMed][Google Scholar]
  • 34. Kinoshita K, Harigai M, Fagarasan S, Muramatsu M, Honjo TA hallmark of active class switch recombination: transcripts directed by I promoters on looped-out circular DNAs. Proc Natl Acad Sci USA. 2001;98:12620–12623.[Google Scholar]
  • 35. Cerutti A, et al Ongoing in vivo immunoglobulin class switch DNA recombination in chronic lymphocytic leukemia B cells. J Immunol. 2002;169:6594–6603.[Google Scholar]
  • 36. Quezada SA, Jarvinen LZ, Lind EF, Noelle RJCD40/CD154 interactions at the interface of tolerance and immunity. Annu Rev Immunol. 2004;22:307–328.[PubMed][Google Scholar]
  • 37. McHeyzer-Williams MG, Ahmed RB cell memory and the long-lived plasma cell. Curr Opin Immunol. 1999;11:172–179.[PubMed][Google Scholar]
  • 38. Stavnezer-Nordgren J, Sirlin SSpecificity of immunoglobulin heavy chain switch correlates with activity of germline heavy chain genes prior to switching. EMBO J. 1986;5:95–102.[Google Scholar]
  • 39. Islam KB, Nilsson L, Sideras P, Hammarstrom L, Smith CITGF-β1 induces germ-line transcripts of both IgA subclasses in human B lymphocytes. Int Immunol. 1991;3:1099–1106.[PubMed][Google Scholar]
  • 40. Nilsson L, et al Structure of TGF-β1-induced human immunoglobulin Cα1 and Cα2 germ-line transcripts. Int Immunol. 1991;3:1107–1115.[PubMed][Google Scholar]
  • 41. Shockett P, Stavnezer J. Effect of cytokines on switching to IgA and α germline transcripts in the B lymphoma I.29μ. Transforming growth factor-β activates transcription of the unrearranged Cα gene. J Immunol. 1991;147:4374–4383. References 38–41 demonstrate that germline transcription confers specificity to CSR and show that TGFβ1 is required for the transcriptional activation of the Cα gene. [[PubMed]
  • 42. Rubtsov YP, Rudensky AYTGFβ signalling in control of T–cell–mediated self-reactivity. Nature Rev Immunol. 2007;7:443–453.[PubMed][Google Scholar]
  • 43. Dahan S, Roth-Walter F, Arnaboldi P, Agarwal S, Mayer LEpithelia: lymphocyte interactions in the gut. Immunol Rev. 2007;215:243–253.[Google Scholar]
  • 44. Stavnezer JRegulation of antibody production and class switching by TGF-β J Immunol. 1995;155:1647–1651.[PubMed][Google Scholar]
  • 45. Pardali E, et al Smad and AML proteins synergistically confer transforming growth factor β1 responsiveness to human germ-line IgA genes. J Biol Chem. 2000;275:3552–3560.[PubMed][Google Scholar]
  • 46. Zhang Y, Derynck RTranscriptional regulation of the transforming growth factor-β-inducible mouse germ line Igα constant region gene by functional cooperation of Smad, CREB, and AML family members. J Biol Chem. 2000;275:16979–16785.[PubMed][Google Scholar]
  • 47. Park SR, Lee JH, Kim PHSmad3 and Smad4 mediate transforming growth factor-β1-induced IgA expression in murine B lymphocytes. Eur J Immunol. 2001;31:1706–1517.[PubMed][Google Scholar]
  • 48. Shi MJ, Stavnezer JCBFα3 (AML2) is induced by TGF-β1 to bind and activate the mouse germline Igα promoter. J Immunol. 1998;161:6751–6760.[PubMed][Google Scholar]
  • 49. Miyazono K, Maeda S, Imamura TCoordinate regulation of cell growth and differentiation by TGF-β superfamily and Runx proteins. Oncogene. 2004;23:4232–4237.[PubMed][Google Scholar]
  • 50. Lin YC, Stavnezer JRegulation of transcription of the germ-line Igα constant region gene by an ATF element and by novel transforming growth factor-β1-responsive elements. J Immunol. 1992;149:2914–2925.[PubMed][Google Scholar]
  • 51. Yang X, et al Targeted disruption of SMAD3 results in impaired mucosal immunity and diminished T cell responsiveness to TGF-β EMBO J. 1999;18:1280–1291.[Google Scholar]
  • 52. Cazac BB, Roes JTGF-β receptor controls B cell responsiveness and induction of IgA in vivo. Immunity. 2000;13:443–451. The first in vivo demonstration that signals emanating from TGFβ1 are essential for the induction of IgA class switching and production in B cells. [[PubMed][Google Scholar]
  • 53. Klein J, et al B cell-specific deficiency for Smad2 in vivo leads to defects in TGF-β-directed IgA switching and changes in B cell fate. J Immunol. 2006;176:2389–2396.[PubMed][Google Scholar]
  • 54. Li R, et al Deletion of exon I of SMAD7 in mice results in altered B cell responses. J Immunol. 2006;176:6777–6784.[PubMed][Google Scholar]
  • 55. Shi MJ, Park SR, Kim PH, Stavnezer JRoles of Ets proteins, NF-κB and nocodazole in regulating induction of transcription of mouse germline Igα RNA by transforming growth factor-β1. Int Immunol. 2001;13:733–746.[PubMed][Google Scholar]
  • 56. Qiu G, Stavnezer JOverexpression of BSAP/Pax-5 inhibits switching to IgA and enhances switching to IgE in the I.29μ B cell line. J Immunol. 1998;161:2906–2918.[PubMed][Google Scholar]
  • 57. Defrance T, et al Interleukin 10 and transforming growth factor β cooperate to induce anti-CD40-activated naive human B cells to secrete immunoglobulin A. J Exp Med. 1992;175:671–682.[Google Scholar]
  • 58. McIntyre TM, Kehry MR, Snapper CMNovel in vitro model for high-rate IgA class switching. J Immunol. 1995;154:3156–3161.[PubMed][Google Scholar]
  • 59. Nakamura M, et al High frequency class switching of an IgM B lymphoma clone CH12F3 to IgA cells. Int Immunol. 1996;8:193–201.[PubMed][Google Scholar]
  • 60. Cerutti A, et al CD40 ligand and appropriate cytokines induce switching to IgG, IgA, and IgE and coordinated germinal center-like phenotype differentiation in a human monoclonal IgMIgD B cell line. J Immunol. 1998;160:2145–2157.[Google Scholar]
  • 61. Zan H, Cerutti A, Dramitinos P, Schaffer A, Casali P. CD40 engagement triggers switching to IgA1 and IgA2 in human B cells through induction of endogenous TGF-β. Evidence for TGF-β-dependent but not IL-10-dependent direct Sμ-Sα and sequential Sμ-Sγ, Sγ-Sα DNA recombination. J Immunol. 1998;162:5217–5225.
  • 62. Kawabe T, et al The immune responses in CD40-deficient mice: impaired immunoglobulin class switching and germinal center formation. Immunity. 1994;1:167–178.[PubMed][Google Scholar]
  • 63. Bergqvist P, Gardby E, Stensson A, Bemark M, Lycke NYGut IgA class switch recombination in the absence of CD40 does not occur in the lamina propria and is independent of germinal centers. J Immunol. 2006;177:7772–7783.[PubMed][Google Scholar]
  • 64. Allen RC, et al CD40 ligand gene defects responsible for X-linked hyper-IgM syndrome. Science. 1993;259:990–993.[PubMed][Google Scholar]
  • 65. Ferrari S, et al Mutations of CD40 gene cause an autosomal recessive form of immunodeficiency with hyper IgM. Proc Natl Acad Sci USA. 2001;98:12614–12619.[Google Scholar]
  • 66. Siebenlist U, Brown K, Claudio EControl of lymphocyte development by nuclear factor-κB. Nature Rev Immunol. 2005;5:435–445.[PubMed][Google Scholar]
  • 67. Snapper CM, et al B cells from p50/NF-κB knockout mice have selective defects in proliferation, differentiation, germ-line CH transcription, and Ig class switching. J Immunol. 1996;156:183–191.[PubMed][Google Scholar]
  • 68. Dedeoglu F, Horwitz B, Chaudhuri J, Alt FW, Geha RSInduction of activation-induced cytidine deaminase gene expression by IL-4 and CD40 ligation is dependent on STAT6 and NFκB. Int Immunol. 2004;16:395–404.[PubMed][Google Scholar]
  • 69. Zhang K, Mills FC, Saxon A. Switch circles from IL–4–directed ε class switching from human B lymphocytes. Evidence for direct, sequential, and multiple step sequential switch from μ to ε Ig heavy chain gene. J Immunol. 1994;152:3427–3435.[PubMed]
  • 70. Fujieda S, Waschek JA, Zhang K, Saxon AVasoactive intestinal peptide induces Sα/Sμ switch circular DNA in human B cells. J Clin Invest. 1996;98:1527–1532.[Google Scholar]
  • 71. Fagarasan S, Honjo TT-Independent immune response: new aspects of B cell biology. Science. 2000;290:89–92.[PubMed][Google Scholar]
  • 72. Hayakawa K, Hardy RRNormal, autoimmune, and malignant CD5 B cells: the Ly-1 B lineage? Annu Rev Immunol. 1988;6:197–218.[PubMed][Google Scholar]
  • 73. Kroese FG, et al Many of the IgA producing plasma cells in murine gut are derived from self-replenishing precursors in the peritoneal cavity. Int Immunol. 1989;1:75–84.[PubMed][Google Scholar]
  • 74. Macpherson AJ, et al A primitive T cell-independent mechanism of intestinal mucosal IgA responses to commensal bacteria. Science. 2000;288:2222–2226. The authors demonstrate that B-1 cells produce antibodies to commensal bacteria through a mechanism that does not require follicular structures and CD4 T-cell help to B cells. [[PubMed][Google Scholar]
  • 75. Franco MA, Greenberg HBImmunity to rotavirus in T cell deficient mice. Virology. 1997;238:169–179.[PubMed][Google Scholar]
  • 76. Macpherson AJ, Uhr TInduction of protective IgA by intestinal dendritic cells carrying commensal bacteria. Science. 2004;303:1662–1665. Evidence that T-cell-independent IgA responses to commensal bacteria require DCs and are confined to the intestinal mucosa. [[PubMed][Google Scholar]
  • 77. Wijburg OL, et al Innate secretory antibodies protect against natural Salmonella typhimurium infection. J Exp Med. 2006;203:21–26.[Google Scholar]
  • 78. Kaminski DA, Stavnezer JEnhanced IgA class switching in marginal zone and B1 B cells relative to follicular/B2 B cells. J Immunol. 2006;177:6025–6029.[PubMed][Google Scholar]
  • 79. Weller S, et al Human blood IgM ‘memory’ B cells are circulating splenic marginal zone B cells harboring a prediversified immunoglobulin repertoire. Blood. 2004;104:3647–3654.[Google Scholar]
  • 80. Weller S, et al CD40-CD40L independent Ig gene hypermutation suggests a second B cell diversification pathway in humans. Proc Natl Acad Sci USA. 2001;98:1166–1170.[Google Scholar]
  • 81. Wardemann H, et al Predominant autoantibody production by early human B cell precursors. Science. 2003;301:1374–1377.[PubMed][Google Scholar]
  • 82. Sims GP, et al Identification and characterization of circulating human transitional B cells. Blood. 2005;105:4390–4398.[Google Scholar]
  • 83. Capolunghi F, et al CpG drives human transitional B cells to terminal differentiation and production of natural antibodies. J Immunol. 2008;180:800–808. References 79–83 describe natural IgM memory B cells and transitional B cells in humans and their possible involvement in T-cell-independent antibody responses. [[PubMed][Google Scholar]
  • 84. Peng SLSignaling in B cells via Toll-like receptors. Curr Opin Immunol. 2005;17:230–236.[PubMed][Google Scholar]
  • 85. Mond JJ, Vos Q, Lees A, Snapper CMT cell independent antigens. Curr Opin Immunol. 1995;7:349–354.[PubMed][Google Scholar]
  • 86. Bergtold A, Desai DD, Gavhane A, Clynes RCell surface recycling of internalized antigen permits dendritic cell priming of B cells. Immunity. 2005;23:503–514.[PubMed][Google Scholar]
  • 87. Kushnir N, Liu L, MacPherson GGDendritic cells and resting B cells form clusters in vitro and in vivo: T cell independence, partial LFA-1 dependence, and regulation by cross-linking surface molecules. J Immunol. 1998;160:1774–1781.[PubMed][Google Scholar]
  • 88. Batista FD, Iber D, Neuberger MSB cells acquire antigen from target cells after synapse formation. Nature. 2001;411:489–494.[PubMed][Google Scholar]
  • 89. Balazs M, Martin F, Zhou T, Kearney JBlood dendritic cells interact with splenic marginal zone B cells to initiate T-independent-immune responses. Immunity. 2002;17:341–352. The first in vivo demonstration that DCs initiate T-cell independent antibody responses by activating B cells through BAFF and APRIL. [[PubMed][Google Scholar]
  • 90. Qi H, Egen JG, Huang AY, Germain RNExtrafollicular activation of lymph node B cells by antigen-bearing dendritic cells. Science. 2006;312:1672–1676.[PubMed][Google Scholar]
  • 91. Litinskiy MB, et al DCs induce CD40-independent immunoglobulin class switching through BLyS and APRIL. Nature Immunol. 2002;3:822–829. The first in vitro demonstration that B cells can undergo CD40-independent CSR in response to BAFF and APRIL provided by DCs. [Google Scholar]
  • 92. Schneider PThe role of APRIL and BAFF in lymphocyte activation. Curr Opin Immunol. 2005;17:282–289.[PubMed][Google Scholar]
  • 93. Dillon SR, Gross JA, Ansell SM, Novak AJAn APRIL to remember: novel TNF ligands as therapeutic targets. Nature Rev Drug Discov. 2006;5:235–246.[PubMed][Google Scholar]
  • 94. Coffman RL, Lebman DA, Shrader BTransforming growth factor β specifically enhances IgA production by lipopolysaccharide-stimulated murine B lymphocytes. J Exp Med. 1989;170:1039–1044.[Google Scholar]
  • 95. Takeda K, Kaisho T, Akira SToll-like receptors. Annu Rev Immunol. 2003;21:335–376.[PubMed][Google Scholar]
  • 96. O’Neill LA, Bowie AGThe family of five: TIR–domain–containing adaptors in Toll-like receptor signalling. Nature Rev Immunol. 2007;7:353–364.[PubMed][Google Scholar]
  • 97. Xie XQ, Pardali E, Holm M, Sideras P, Grundstrom TAML and Ets proteins regulate the Iα1 germ-line promoter. Eur J Immunol. 1999;29:488–498.[PubMed][Google Scholar]
  • 98. He B, Qiao X, Cerutti ACpG DNA induces IgG class switch DNA recombination by activating human B cells through an innate pathway that requires TLR9 and cooperates with IL-10. J Immunol. 2004;173:4479–4491.[PubMed][Google Scholar]
  • 99. Xu W, et al Epithelial cells trigger frontline immunoglobulin class switching through a pathway regulated by the inhibitor SLPI. Nature Immunol. 2007;8:294–303. The first demonstration that epithelial cells can initiate T-cell-independent IgG and IgA class switching after sensing microorganisms through TLRs. [[PubMed][Google Scholar]
  • 100. Kaminski DA, Stavnezer JStimuli that enhance IgA class switching increase histone 3 acetylation at Sα, but poorly stimulate sequential switching from IgG2b. Eur J Immunol. 2007;37:240–251.[PubMed][Google Scholar]
  • 101. Bradley SP, Kaminski DA, Peters AH, Jenuwein T, Stavnezer JThe histone methyltransferase Suv39h1 increases class switch recombination specifically to IgA. J Immunol. 2006;177:1179–1188.[PubMed][Google Scholar]
  • 102. Drouin EE, Schrader CE, Stavnezer J, Hansen UThe ubiquitously expressed DNA-binding protein late SV40 factor binds Ig switch regions and represses class switching to IgA. J Immunol. 2002;168:2847–2856.[PubMed][Google Scholar]
  • 103. Bendelac A, Bonneville M, Kearney JFAutoreactivity by design: innate B and T lymphocytes. Nature Rev Immunol. 2001;1:177–186.[PubMed][Google Scholar]
  • 104. Bernasconi NL, Traggiai E, Lanzavecchia AMaintenance of serological memory by polyclonal activation of human memory B cells. Science. 2002;298:2199–2202.[PubMed][Google Scholar]
  • 105. Castigli E, et al TACI and BAFF-R mediate isotype switching in B cells. J Exp Med. 2005;201:35–39. Identification of TACI as the major IgA class-switch-inducing receptor in B cells exposed to BAFF or APRIL. [Google Scholar]
  • 106. von Bulow GU, van Deursen JM, Bram RJRegulation of the T-independent humoral response by TACI. Immunity. 2001;14:573–582.[PubMed][Google Scholar]
  • 107. Ng LG, et al BAFF costimulation of Toll-like receptor-activated B-1 cells. Eur J Immunol. 2006;36:1837–1846.[PubMed][Google Scholar]
  • 108. Castigli E, et al TACI is mutant in common variable immunodeficiency and IgA deficiency. Nature Genet. 2005;37:829–834.[PubMed][Google Scholar]
  • 109. Salzer U, et al Mutations in TNFRSF13B encoding TACI are associated with common variable immunodeficiency in humans. Nature Genet. 2005;37:820–828. References 108 and 109 describe IgA deficiency in humans with TACI signalling defects. [[PubMed][Google Scholar]
  • 110. Sakurai D, et al TACI regulates IgA production by APRIL in collaboration with HSPG. Blood. 2007;109:2961–2967.[PubMed][Google Scholar]
  • 111. Castigli E, et al Transmembrane activator and calcium modulator and cyclophilin ligand interactor enhances CD40-driven plasma cell differentiation. J Allergy Clin Immunol. 2007;120:885–891.[Google Scholar]
  • 112. Thompson JS, et al BAFF-R, a newly identified TNF receptor that specifically interacts with BAFF. Science. 2001;293:2108–2111.[PubMed][Google Scholar]
  • 113. O’Connor BP, et al BCMA is essential for the survival of long-lived bone marrow plasma cells. J Exp Med. 2004;199:91–98.[Google Scholar]
  • 114. Xia XZ, et al TACI is a TRAF-interacting receptor for TALL-1, a tumor necrosis factor family member involved in B cell regulation. J Exp Med. 2000;192:137–143.[Google Scholar]
  • 115. Fayette J, et al Human dendritic cells skew isotype switching of CD40-activated naive B cells towards IgA1 and IgA2. J Exp Med. 1997;185:1909–1918.[Google Scholar]
  • 116. Craig SW, Cebra JJPeyer’s patches: an enriched source of precursors for IgA-producing immunocytes in the rabbit. J Exp Med. 1971;134:188–200.[Google Scholar]
  • 117. Martinoli C, Chiavelli A, Rescigno MEntry route of Salmonella typhimurium directs the type of induced immune response. Immunity. 2007;27:975–984. This study shows that the IgA response to invasive bacteria takes place in Peyer’s patches. [[PubMed][Google Scholar]
  • 118. Iwasaki A, Kelsall BLFreshly isolated Peyer’s patch, but not spleen, dendritic cells produce interleukin 10 and induce the differentiation of T helper type 2 cells. J Exp Med. 1999;190:229–239.[Google Scholar]
  • 119. Rimoldi M, et al Intestinal immune homeostasis is regulated by the crosstalk between epithelial cells and dendritic cells. Nature Immunol. 2005;6:507–514.[PubMed][Google Scholar]
  • 120. Ziegler SF, Liu YJThymic stromal lymphopoietin in normal and pathogenic T cell development and function. Nature Immunol. 2006;7:709–714.[PubMed][Google Scholar]
  • 121. Mora JR, et al Generation of gut-homing IgA-secreting B cells by intestinal dendritic cells. Science. 2006;314:1157–1160. This study shows that DCs confer gut-homing properties to IgA class-switched B cells through a process involving retinoic acid, a metabolite of vitamin A. [[PubMed][Google Scholar]
  • 122. Sato A, et al CD11b Peyer’s patch dendritic cells secrete IL-6 and induce IgA secretion from naive B cells. J Immunol. 2003;171:3684–3690.[PubMed][Google Scholar]
  • 123. Tezuka H, et al Regulation of IgA production by naturally occurring TNF/iNOS-producing dendritic cells. Nature. 2007;448:929–933. Recent evidence showing the involvement of bacteria-induced iNOSTNF DCs in the initiation of intestinal IgA responses, including IgA class switching. [[PubMed][Google Scholar]
  • 124. Casola S, et al B cell receptor signal strength determines B cell fate. Nature Immunol. 2004;5:317–327. This paper reports that Peyer’s patch B cells can undergo T-cell-dependent IgA class switching and production in the absence of canonical cognate T-cell–B-cell interactions. [[PubMed][Google Scholar]
  • 125. Stoel M, et al Restricted IgA repertoire in both B-1 and B-2 cell-derived gut plasmablasts. J Immunol. 2005;174:1046–1054.[PubMed][Google Scholar]
  • 126. Fagarasan S, et al Critical roles of activation-induced cytidine deaminase in the homeostasis of gut flora. Science. 2002;298:1424–1427.[PubMed][Google Scholar]
  • 127. Hayakawa K, Hardy RR, Stall AM, Herzenberg LA, Herzenberg LAImmunoglobulin-bearing B cells reconstitute and maintain the murine Ly-1 B cell lineage. Eur J Immunol. 1986;16:1313–1316.[PubMed][Google Scholar]
  • 128. Thurnheer MC, Zuercher AW, Cebra JJ, Bos NAB1 cells contribute to serum IgM, but not to intestinal IgA, production in gnotobiotic Ig allotype chimeric mice. J Immunol. 2003;170:4564–4571.[PubMed][Google Scholar]
  • 129. Crouch EE, et al Regulation of AID expression in the immune response. J Exp Med. 2007;204:1145–1156. Elegant work recapitulating AID expression during immune responses at different sites, including the intestinal lamina propria, in an AID reporter mouse. [Google Scholar]
  • 130. Cerutti ALocation, location, location: B-cell differentiation in the gut lamina propria. Mucosal Immunol. 2008;1:8–10.[PubMed][Google Scholar]
  • 131. MacLennan IC, et al Extrafollicular antibody responses. Immunol Rev. 2003;194:8–18.[PubMed][Google Scholar]
  • 132. Suzuki K, Meek B, Doi Y, Honjo T, Fagarasan STwo distinctive pathways for recruitment of naive and primed IgM B cells to the gut lamina propria. Proc Natl Acad Sci USA. 2005;102:2482–2486.[Google Scholar]
  • 133. Rescigno M, et al Dendritic cells express tight junction proteins and penetrate gut epithelial monolayers to sample bacteria. Nature Immunol. 2001;2:361–367.[PubMed][Google Scholar]
  • 134. Chieppa M, Rescigno M, Huang AY, Germain RNDynamic imaging of dendritic cell extension into the small bowel lumen in response to epithelial cell TLR engagement. J Exp Med. 2006;203:2841–2852.[Google Scholar]
  • 135. Hebel K, et al Plasma cell differentiation in T-independent type 2 immune responses is independent of CD11c dendritic cells. Eur J Immunol. 2006;36:2912–2919.[PubMed][Google Scholar]
  • 136. Kato A, Truong-Tran AQ, Scott AL, Matsumoto K, Schleimer RPAirway epithelial cells produce B cell-activating factor of TNF family by an IFN-γ-dependent mechanism. J Immunol. 2006;177:7164–7172.[Google Scholar]
  • 137. Cunningham-Rundles C, Knight AKCommon variable immune deficiency: reviews, continued puzzles, and a new registry. Immunol Res. 2007;38:78–86.[PubMed][Google Scholar]
  • 138. Holmgren J, Czerkinsky CMucosal immunity and vaccines. Nature Med. 2005;11:S45–S53.[PubMed][Google Scholar]
  • 139. Burton DRAntibodies, viruses and vaccines. Nature Rev Immunol. 2002;2:706–713.[PubMed][Google Scholar]
  • 140. Mazzoli S, et al HIV-specific mucosal and cellular immunity in HIV-seronegative partners of HIV-seropositive individuals. Nature Med. 1997;3:1250–1257.[PubMed][Google Scholar]
Collaboration tool especially designed for Life Science professionals.Drag-and-drop any entity to your messages.