Open access peer-reviewed chapter

Gene Expression during the Activation of Human B Cells

Written By

Youming Zhang

Submitted: 17 July 2017 Reviewed: 23 October 2017 Published: 21 February 2018

DOI: 10.5772/intechopen.71863

From the Edited Volume

Gene Expression and Regulation in Mammalian Cells - Transcription From General Aspects

Edited by Fumiaki Uchiumi

Chapter metrics overview

1,777 Chapter Downloads

View Full Metrics

Abstract

Human B lymphocytes not only play a critical role in the humoral immunity to generate antibodies, but also are equally important to cellular immunity as B lymphocytes can present antigens to T lymphocytes and can release a range of potential immune-regulating cytokines after stimulations. Human immunoglobulin class switch recombination (CSR) in activated B cells is an essential process in the humoral immunity and the process is complicated and tightly controlled by many regulators. The recent genomic and genetic approaches in CSR identified novel genes that were actively involved in the process. Understanding the roles of the novel genes in CSR will bring new insights into the mechanisms of the process and new potential therapeutic targets for immunoglobulin-related disorders such as allergic asthma and autoimmune diseases.

Keywords

  • B cells
  • immunoglobulin class switch recombination
  • gene expression
  • regulation

1. Introduction

Human lymphocytes include T lymphocytes, B lymphocytes and natural killer cells. T lymphocytes are majorly responsible for cell-mediated immunity. B lymphocytes (cells) play the critical roles in the humoral immunity to activate immune system by secreting antibodies. B lymphocytes are also equally critical to cellular immunity as they can also present antigens to T lymphocytes and can also release a range of potential immune-regulating cytokines [1, 2].

The “B” from B cells came from the name of the bursa of Fabricius, a lymphoid organ in birds, where B cells mature. It was first discovered by Chang and Glick [3]. B cells mature in the bone marrow in mammals. B cells express B cell receptors (BCRs) on their cell membrane and BCRs allow the cell to bind to a specific antigen and initiate an antibody response. Each B cell carries a unique receptor for antigen that is composed of the membrane-band form of its antibody. After antigen recognition by the membrane-bound receptor, the B cells can proliferate to increase their numbers and differentiate to secrete their antigen-specific antibodies.

There are three principle classes of B cells in humans according to their ontogeny and anatomic localization: B1 cells arise from fetal liver precursors and are enriched in mucosal tissues and the pleural and peritoneal cavities. B2 cells arise from bone marrow–derived precursors and are enriched in secondary lymphoid organs [4]. Marginal zone (MZ) and follicular (FO) B cells are differentiated from B2 cells in human spleen and lymph nodes [2]. B cells of each linage have distinct and overlapping functions in recognizing antigens via T-independent and T-dependent pathways to produce rapid IgM or long-lasting IgG antibody response [2]. Cytokines play a key role in the commitment of naïve B cells to B effector 1 (Be-1) and B effector 2 (Be-2) lineage. Be-2 differentiation is dependent on the engagement of IL4α on B cells [5], while Be-1 cell development is dependent on the activation of the transcription factor T-bet and the IFNγR on B cells [6].

The process of human B cell development is very complicated and is controlled by many transcription factors [7]. Human B cells are generated in bone marrow from progenitor cells that are committed to the B cell linages (pro-B cells). Each pro-B cell undergoes independent rearrangement of diverse variable (V), diversity (D) and joining (J) gene segments of the immunoglobulin heavy (H)-chain locus [8]. Rearrangement of the H-chain locus creates in each B cell a variable exon with a unique upstream of the immunoglobulin constant region exons and drives the expression of H-chain protein and then proliferate and differentiate to commence immunoglobulin light (L)-chain gene recombination. When a B cell expresses L-chain protein, it pairs with the previously arranged H chain and is expressed as membrane immunoglobulin on the cell surface [9]. Human immune system can generate a diversity of specific antibodies in response to antigen stimulation. This process is of fundamental importance to acquired immunity. The human constant H-chain genes are on chromosome 14 containing Cμ, Cδ and two repeated clusters each having two Cγ genes and Cε genes (Cγ3, Cγ1, pseudo-ε, Cα1 and Cγ2, Cγ4, Cε, Cα2).

In this chapter, I will briefly introduce the roles of germinal centers (GCs) and the steps of immunoglobulin class switch recombination (CSR) in human B cells in GCs. I will discuss the potential functional roles of the newly identified genes from the results of our experiments for global transcript profiling in CSR. I will also discuss the future direction of the researches on CSR in human B cells.

Advertisement

2. Germinal centers and immunoglobulin CSR

Germinal centers (GCs) are the sites within secondary lymphoid organs such as lymph nodes and the spleen where mature B cells can proliferate, differentiate and mutate their antibody genes and switch the class of their antibodies (e.g., from IgM to IgE) during a normal immune response to antigens [10]. In the GCs, naïve B cells can have clonal expansion, somatic hypermutation, affinity maturation, development of B cell memory and long-life plasma cells [11, 12, 13]. B cell activation is initiated in the follicle in GCs when it encounters specific antigen [14], and then the B cells are relocated to the periphery of the follicle [15]. The inter-follicular zone in GCs is the site where B cell and T follicular helper cell differential initiates [16]. They develop dynamically after the activation of follicular B cells by T-dependent antigen. B cells in GCs proliferate and can class-switch the BCR constant region from IgM/IgD to IgG, IgA and IgE (discussed later). Additionally, the IgV region genes of B cells in GCs can undergo somatic hypermutation to change the affinity of the encoded BCR for its cognate antigens, allowing subsequent antigen-driven selection and clonal expansion of high-affinity B cells [17]. Human B cells in GCs, in vitro–activated naïve B cells, and those with specific and rapid recall responses to previously encountered antigen express cell-surface CD27. B cells with CD27 expression correlated with greater cell sizes, proliferative capacity, antigen presentation capacity and differentiation into antibody secreting cells [18, 19, 20].

In order to generate antibodies, two somatic DNA recombination events of the genetic elements take place in B cells. Firstly, V(D)J recombination generates the functional variable regions of the Ig heavy-chain (IgH) and light-chain genes. Initiation of V(D)J recombination requires the products of recombination activating genes (RAG) 1 and 2 [21, 22]. Lymphoid-specific expression of RAG 1 and 2 limits V(D)J recombination to B and T lymphocytes. Following activation, mature B cells can undergo CSR, linking the IgH variable regions with one of the downstream CH genes, changing the effector function of the antibody [23]. CSR and the other main diversification event, somatic hypermutation (SHM), are both dependent on activation-induced cytidine deaminase (AID), a protein expressed only in activated germinal center B cells [24]. The basic steps of CSR include creating double-strand DNA breaks (DSBs) for CSR and joining donor and acceptor S regions. Class switch recombination occurs between switch (S) regions located upstream of each of the CH regions except Cδ and results in a change from IgM and IgD expression in naïve B cells to express one of the downstream isotypes such as IgG subclasses, IgA and IgE. AID plays a critical role in the vertebrate adaptive immune response [24, 25]. It initiates the conversion of several dC bases to dU bases in each S region, dU bases are then excised by uracil DNA glycosylase (UNG), and the resulting abasic sites are nicked by apurinic/apyrimidic endonuclease (APE), creating single strand breaks (SSBs), that can spontaneously form DSBs if they are near each other on opposite DNA strands. After formation of the DSB in the donor and acceptor S regions, the S regions are recombined by ubiquitous proteins that perform nonhomologous end-joining (NHEJ) [26]. VDJ recombination and early B cell development takes place in the bone morrow. Immature B cells expressing IgM on the surface migrate to peripheral lymphoid tissue in the spleen, lymph nodes and gut-associated lymphoid tissue. CSR and SHM happen in the germinal centers of secondary lymphoid tissues but also in germinal center–like structures in local (nonlymphoid) tissues [27]. CSR is induced by both T lymphocyte–dependent (TD) antigens and T lymphocyte–independent (TI) antigens. TD antigen stimulation can be mimicked in vitro by culturing B cells in the presence of anti-CD40 antibodies along with specific cytokines. IL-4 and anti-CD40 induce isotype switching to IgG1 and IgE [28]. Chromatin structure also contributes to the regulation of CSR. Ig heavy-chain constant genes and 3-regulatory regions are in an active chromatin conformation (acetylated H3 and H4 and lysine 4 trimethylation H3) in unstimulated human B cells, and these modifications can spread to the S region after cytokine stimulation [29]. AID is exclusively expressed in the germinal centers [21]. The basic AID-mediated mechanisms of CSR are quite well studied and defined, but the global regulation of the CSR, accompanying networks of AID and other well-known regulators, remains relatively unclear.

Advertisement

3. The global gene regulation of CSR in human B cells

In vitro, IL4 and anti-CD40 signals can mimic signals from T cell in GCs to induce a strong activation of NF-κB leading human B cells to a proliferative burst and CSR to IgE and IgG [30]. These costimulation signals were applied in the naïve B cells isolated from healthy tonsils and profiled the transcripts at 6 time points for 12 days (0, 12, 36, 72, 120 and 288 h). More than one thousand genes were observed to have significantly differentiated expression after IL4 and anti-CD40 stimulation [31]. The significantly differentiated genes can be formed in 4 cluster groups including 13 temporal profiles. Each cluster contains many new genes that were not known to have roles in CSR before.

3.1. Cluster A group

Cluster A group represented the gene expression on (Cluster A1) or off (Cluster A2) in naïve human B cells after IL4 and anti-CD40 stimulation.

3.1.1. Cluster A1

Cluster A1 was immediately upregulated after IL4 and anti-CD40 cosignal stimulation and the expressions did not change during the course of 12 days for the experiments. The cluster contains 153 genes. The analysis of transcription factor–binding sites for the cluster showed that genes from this cluster were enriched to transcription factors BACH1 and BACH2. BACH1 and BACH2 promote B cell development by repressing the myeloid program [32]. They belong to the basic region–leucine zipper family and are transcription repressors binding to Maf-recognition elements (MAREs) [33]. BACH2 has critical roles in both acquired immunity and innate immunity, including immunoglobulin CSR, the somatic hypermutation of immunoglobulin-encoding genes [34, 35]. BACH2 expression is activated by E2A, Foxo1 and Pax5 in pro-B cells. BACH2 may have a role in early B cell development [36]. BACH1 structure is closely related to BACH2, but its role in B cell development and hematopoiesis largely remains unclear [33]. BACH2 expression frequently preceded that of Ebf1 and Pax5 in the common lymphoid progenitors (CLPs). BACH factors directly repressed various myeloid genes in CLPs and this repression restricted the fate of CLPs to the B cell lineage [32].

In this cluster, chemokine genes CCL22 and CCL17 were the most significantly differentiated genes during naïve B cell activation after IL4 and anti-CD40 signal stimulation. CCL22 and CCL17 are both ligands for the chemokine receptor CCR4. CCR4 gene was also showed in Cluster C6 to have a transient induction after IL4 and anti-CD40 signal stimulation. All three transcripts were within the top 20 differentially expressed genes during the activation of immunoglobulin class switching in human activated B cells. CCL22 and CCL17 are NF-kappa B (NF-kB) target genes, indicating a central role for the NF-kB pathway in the activation of CSR stimulated by IL4 and anti-CD40. The top differentially expressed genes also contained another NF-kB target gene, the TNF receptor-associated factor (TRAF1) [37], which was also profiled in Cluster A1. There were many clinical reports that indicated both chemokines might be involved in human immunoglobulin class switching. A significantly higher increase in CCL17, CCL22 and IL-4 serum levels in grass pollen–exposed subjects was observed [38]. Sensitized children with allergic symptoms showed higher CCL17 and CCL22 levels and higher ratios between these Th2-associated chemokines and the Th1-associated chemokine CXCL10 than nonsensitized children without allergic symptoms [39]. Using human dendritic cells (DCs), in vitro exposure to house dust mite (HDM) of DCs from HDM-allergic patients but not healthy controls caused CCL17 and CCL22 release that resulted in chemoattraction of polarized human Th2 cells in a CCR4-dependent way [40]. Both CCL22 and CCL17 have been suggested as biomarkers for disease activity in atopic dermatitis (AD), and raised cord blood (CB) levels of CCL22 predict subsequent allergic sensitization, while raised CCL17 in GC predicts the later development of allergic symptoms, including asthma. Consistent with these observations, allergen exposure in sensitized individuals leads to a dynamic increase in CCL17 and CCL22 [38]. High-affinity neutral ligands have been developed for CCL22 and CCL17 and attenuate levels of CCL22, CCL17 and IgE in a mouse model of atopic dermatitis as well as improve skin inflammatory symptoms [31]. CCL17 was also shown to have exon retention during B cell activation [31]. All the evidence indicated both chemokines and their receptor CCR4 play important roles in immunoglobulin class switching.

3.1.2. Cluster A2

Cluster A2 was immediately downregulated after IL4 and anti-CD40 cosignal stimulation and the expressions did not change during the course of 12 days of CSR. Cluster A2 contains 83 genes that present downregulating genes during B cell activation. These genes expressed significantly lower in activated B cells than in naïve B cells during the time course of IL4 and anti-CD40 stimulation. The analysis of gene ontology indicated the genes in this cluster were involved with immune system process.

FOSB and FOS were the most significantly downregulated genes in Cluster A2 during CSR in naïve B cells after IL4 and anti-CD40 stimulation. FOS genes encode leucine zipper proteins that can dimerize with proteins of the JUN family and form transcription factor complex activating protein-1 (AP-1) [41]. The FOS family consists of 4 members: FOS, FOSB, FOSL1 and FOSL2. Activating protein-1 (AP-1) is a dimeric transcription factor composed of Jun, FOS or activating transcription factor (ATF) subunits that bind to a common DNA site, the AP-1–binding site [42]. The different AP-1 factors may regulate different target genes and thus execute distinct biological functions [43]. In addition to regulation by heterodimerization among Jun, FOS and ATF proteins, AP-1 activity is regulated through interactions with specific protein kinases and a variety of transcriptional coactivators [44, 45, 46]. Nitrogen oxide (NO) is the radical inhibiting IgE/Ag-induced IL-4, IL-6 and TNF production. It inhibits phosphorylation of phospholipase Cγ1 and the AP-1 transcription factor protein c-Jun. NO further completely abrogated IgE/Ag-induced DNA-binding activity of the nuclear AP-1 proteins FOS and Jun to regulate allergic inflammation [47]. FOS-interacting protein (FIP) is a transcription factor that binds to c-FOS. The aggregation of the mast cell’s high-affinity receptor for IgE induced the synthesis of FIP and increased its DNA-binding activity. Moreover, downregulation of the isoenzyme protein kinase C-β (PKC-β) resulted in profound inhibition of FIP-FOS DNA-binding activity [48].

3.2. Cluster B group

Cluster B group showed gradually sustained induction during CSR in B cells. Cluster B1 is the most interesting cluster that sustained induction earlier than Cluster B2.

3.2.1. Cluster B1

Cluster B1 was the first group to show gradually sustained expression after IL4 and anti-CD40 cosignal stimulation. Cluster B1 contains 126 genes and the analysis of gene ontology showed that genes in this cluster were majorly involved in the cellular amine metabolic process. The analysis of transcription factor–binding sites indicated the genes in this cluster were enriched to transcription factors RSRFC4 and STAT. Both transcription factors were involved with allergic and airway epithelia inflammations [49, 50]. RSRF-binding sites were found in the regulatory sequences of a number of growth factor–inducible and muscle-specific genes [51]. It was showed that engagement of the B cell antigen receptor could activate STAT through Lyn in a JAK-independent pathway [52].

There were several well-known genes to regulate B cell differentiation in germinal center including AICDA [24], IRF4 [53], XBP1 [54], BATF3 [55] and NFIL3 [56] in this cluster. The cluster showed other genes exhibiting synchronic, coordinated expression with the well-documented regulation genes. IL17RB and BHLHE40 genes were the most significantly differentiated in the cluster. IL17RB encodes a cytokine receptor that specifically binds to IL17B and IL17E but does not bind to IL17 and IL17C. IL17RB has been shown to mediate the activation of NF-κB [57]. IL17RB showed highly synchronic expression with AICDA in the cluster. IL17RB abundance has previously been shown to increase upon allergen challenge in patients with seasonal allergic rhinitis [58], IgE [59] and asthma [60]. The result indicated that the increase in IL17RB formed an early component of the transcriptional cascade that initiated the germinal center response in B cells. BHLHE40 encodes a basic helix-loop-helix protein expressed in various tissues and is an environmentally inducible moderator of circadian rhythms and cellular differentiation. BHLHE40 was profiled at its core of the B1 Cluster. BHLHE40 was recently shown to operate as a master regulator of germinal center activities, modulating the expression of more than 100 target genes [61]. Circadian oscillations in symptom severity are a prominent feature of atopic diseases including atopic dermatitis, asthma, chronic urticarial and allergic rhinitis [62, 63, 64]. The variation in IgE/mast cell allergic reactions was recently demonstrated to depend on the circadian clock in mice [65]. Mice deficient for the BHLHE40 ortholog display a variety of immune features including abnormal IgG1 and IgE levels and defective elimination of activated B cells, as well as exhibiting circadian rhythm phenomena [66]. Like BHLHE40, NFIL3 in this cluster also participates in signaling pathways relating to the circadian clock [67], and together, these data suggest there may be a circadian component to class switch recombination and that this may be of relevance to time-of-day phenomena in IgE-driven diseases.

3.2.2. Cluster B2

The genes in Cluster B2 were also gradually sustained induction but they come later in time than Cluster B1 during CSR in human B cell. The cluster contains 112 genes. The most significantly differentiated genes were EPHB1 and TNFSF4.

Erythropoietin-producing hepatocellular carcinoma (Eph) receptors are a subfamily of receptor tyrosine kinases (RTKs) [68, 69]. The receptors and their ligands, the ephrins, mediate numerous developmental processes, particularly in the nervous system [70]. Tyrosine phosphorylation of EphB1 requires presentation of ephrin-B1 in either clustered or membrane-attached forms [71]. Eph receptors and ephrin ligands have been shown to be differentially expressed on leucocytes. Ephrin-B3 binds to B lymphocytes, most likely via a nonclassical receptor, and induces migration of the memory B cell subpopulation [72]. NFSF4 encodes a cytokine of the tumor necrosis factor (TNF) ligand family. The encoded protein functions in T cell and antigen-presenting cell (APC) interact and mediate adhesion of activated T cells to endothelial cells. The tumor necrosis factor ligand superfamily member 4 gene (TNFSF4, OX40L), which encodes for the costimulatory molecule OX40 ligand, has been identified as a susceptibility gene for systemic lupus erythematosus (SLE) in multiple studies [73, 74].

3.3. Cluster C group

Cluster C group has six profiling clusters to show transient induction during CSR in B cells according to the time they were inducted.

3.3.1. Cluster C1

Cluster C1 was the first group to be inducted transcendently during CSR. It has 79 genes and the analysis of gene ontology indicated the genes in this cluster were involved in ribonucleoprotein complex biogenesis. Heterogeneous nuclear ribonucleoproteins (hnRNPs) are a large family of RNA-binding proteins that are important for multiple aspects of nucleic acid metabolism [75]. TFEC and RRP12 were the most significantly introduced genes in this cluster.

Transcription factor EC (TFEC) acts as a repressor or an activator. TFEC works as a transcriptional repressor on minimal promoter containing element F in an E-box sequence-specific manner [76]. It can act as a transcriptional transactivator on the proximal promoter region of the tartrate-resistant acid phosphatase (TRAP) E-box containing promoter. It also acts as a transcriptional repressor on minimal promoter containing mu E3 enhancer sequence [77]. Gain-of-function assays indicated that TFEC was capable of expanding hematopoietic stem cells–derived hematopoiesis. TFEC mutants were showed to reduce hematopoiesis in the caudal hematopoietic tissue, leading to anemia. It mediated these changes by increasing the expression of several cytokines in caudal endothelial cells [78]. Ribosomal RNA Processing 12 Homolog (RRP12) is a protein that may have a function to bind to poly(A) RNA. Rrp12 and the exportin Crm1 participate in late assembly events in the nucleolus during 40S ribosomal subunit biogenesis [79], but there is little knowledge of TFEC and RR12 regulating B cell growth.

3.3.2. Cluster C2

Cluster C2 was the second group to be inducted transcendently during CSR in human B cells and this cluster contains 112 genes. LMNB2 and B4GALT5 were the most significantly introduced in this cluster.

LMNB2 encodes a B-type nuclear lamin. The nuclear lamina consists of a two-dimensional matrix of proteins located next to the inner nuclear membrane. Lamin proteins are thought to be involved in nuclear stability, chromatin structure and gene expression. B-type lamins play a role in DNA replication, the formation of the mitotic spindle, chromatin organization and regulation of gene expression [80]. B4GALT5 encodes one of seven beta-1,4-galactosyltransferase. It is the type II membrane-bound glycoproteins that appear to have exclusive specificity for the donor substrate UDP-galactose; B4GALT5 was found to have a change in a statin-induced experiment in gene expression in EBV-transformed and native B cells [81].

3.3.3. Cluster C3

Cluster C3 was the third group to be inducted transcendently during CSR in B cells and the cluster has 105 genes. The analysis of gene ontology indicated the genes in this cluster were involved with DNA metabolic process. UHRF1 and CHEK1 were the most significantly introduced in this group.

UHRF1 gene encodes a member of a subfamily of RING-finger–type E3 ubiquitin ligases. The protein binds to specific DNA sequences and recruits a histone deacetylase to regulate gene expression. Its expression peaks at late G1 phase and continues during G2 and M phases of the cell cycle. Colonization of germ-free mice with gut microbiota showed increasing expression of Uhrf1 in Treg cells. Uhrf1 deficiency resulted in de-repression of the gene (Cdkn1a) [77]. CHEK1 encodes a protein belonging to the Ser/Thr protein kinase family. It is required for checkpoint-mediated cell cycle arrest in response to DNA damage or the presence of unreplicated DNA. Activated CHEK1 can phosphorylate and modulate the activity of a number of proteins including p53, providing a link between sensing of DNA damage and p53 checkpoint activity. BCL6 can directly bind to a DNA consensus element in the CHEK1 promoter and repress its expression in normal and malignant B cell [82].

3.3.4. Cluster C4

Cluster C4 was the fourth group to be inducted transcendently during CSR and it has 151 genes. The analysis of gene ontology showed that the genes in this cluster were majorly involved in the M phase. The analysis of transcription factor enrichment indicated the genes in this cluster were enriched to nuclear transcription factor Y (NF-Y). NF-Y in eukaryotes consists of three different subunits, NF-YA, NF-YB and NF-YC, which are all necessary for the formation of NF-Y complexes and binding to CCAAT boxes in promoters of their target genes. Recent studies demonstrated novel contributions of NF-Y to apoptosis and apoptosis-induced proliferation and in photoreceptor cell differentiation during the development of the Drosophila compound eye [83].

KIF14 and PRC1 were the most significantly differential expression genes in the cluster. KIF14 encodes a member of the kinesin-3 superfamily of microtubule motor proteins. These proteins are involved in numerous processes including vesicle transport, chromosome segregation, mitotic spindle formation and cytokinesis. Knockdown of this gene results in failed cytokinesis with endoreplication. This gene was identified as a likely oncogene in breast, lung and ovarian cancers, as well as in retinoblastomas and gliomas [84]. Protein regulator of cytokinesis 1 (PRC1) gene is a crucial regulator of cytokinesis [85]. Its suppression may result in mitotic failure and its involvement in various cancers [86]. PRC1 is a key regulator of cytokinesis that cross-links antiparallel microtubules. Multiple mitotic kinesins and microtubule-associated proteins (MAPs) act in concert to direct cytokinesis [87]. The MAP and microtubule-bundling protein PRC1 is one of the key molecules required for the integrity of this structure. Endogenous PRC1 can be interacted with KIF14. KIF14 targets the central spindle via its interaction with PRC1 and has an essential function in cytokinesis [88].

3.3.5. Cluster C5

Cluster C5 was the fifth group to be inducted transcendently and it has 99 genes. MCM10 and PCNA genes were the most differentially expressed in the cluster.

MCM10 encodes one of the highly conserved mini-chromosome maintenance proteins (MCM) that are involved in the initiation of eukaryotic genome replication. Human MCM10 regulates the catalytic subunit of DNA polymerase-α and prevents DNA damage during replication [89]. MCM10 interacts with RECQ4 (RecQ helicases 4) and is important for efficient replication [90, 91]. PCNA encodes a cofactor of DNA polymerase delta in nucleus. The protein acts as a homotrimer and helps increase the processivity of leading strand synthesis during DNA replication. PCNA was well studied in plants and had the ability to stimulate the activity of DNA polymerase δ and the ability to interact with p21, a regulator of the cell cycle [92].

3.3.6. Cluster C6

Cluster C6 was the sixth group to be inducted transcendently and it contains 128 genes. CCR4 and HIST1H1C genes were the most differentially expressed in the whole process during naïve B cell activation with IL4 and anti-CD40 signal stimulation.

CCR4 is the receptor of CCL17 and CCL22. It is later inducted, which means that the three may work in late stage of CSR. CCR4 was previously detected in nongerminal center cells. The possible functional roles in CSR were discussed in Section 3.1.1. Histone H1 has previously been shown to influence mast cell–mediated type-I hyperreactivity in mice [93].

3.4. Cluster D group

Cluster D group has three profiling clusters to show transient downregulation during CSR according to the time of downregulation.

3.4.1. Cluster D1

Cluster D1 was the first group to be inducted transcendently and it contains 99 genes. GPR18 and TP53INP1 genes were the most differentially expressed in the cluster.

GPR18 encodes G protein–coupled receptor 18. The activity of this receptor is mediated by G proteins, which inhibit adenylyl cyclase [94], and it contributes to regulation of the immune system. GPR18 also mediates NAGly-induced process of reorganization of actin filaments and induction of acrosomal exocytosis. Stimulation of human spermatozoa with the GPR18 ligand N-arachidonoylglycine induced the phosphorylation of 12 protein kinases. N-arachidonoylglycine affects the cytoskeleton by changing levels of F-actin and inducing the acrosome reaction in human spermatozoa in a concentration-dependent manner. GPR18 might be involved in physiological processes of human spermatozoa [95]. Tumor protein 53-induced nuclear protein 1 (TP53INP1) is a tumor suppressor. It was described as a p53 target gene involved in cell death, cell-cycle arrest and cellular migration [96]. TP53INP1 is also able to interact with ATG8-family proteins to induce autophagy-dependent cell death by caspase-dependent autophagy [97].

3.4.2. Cluster D2

Cluster D2 was the second group to be inducted transcendently during CSR in B cells. It contains 69 genes. RAB6B and PM20D1 genes were the most differentially expressed.

RAB6B (RAB6B, member RAS oncogene family) has the ability for GTP binding and myosin V binding. Members of the RAB subfamily of small GTPases play an important role in the regulation of intracellular transport routes [98]. RAB6B is predominantly expressed in brain and the neuroblastoma cell line SK-N-SH. In brain, RAB6B was found to be specifically expressed in microglia, pericytes and Purkinje cells. Endogenous RAB6B localizes to the Golgi apparatus and to ERGIC-53–positive vesicles. RAB6B displayed lower GTP-binding activities, and in overexpression studies, the protein is distributed over Golgi and ER membranes [99]. A secreted enzyme, peptidase M20 domain containing 1 (PM20D1), is enriched in UCP1(+) versus UCP1(−) adipocytes. These data identify an enzymatic node and a family of metabolites that regulate energy homeostasis [99].

3.4.3. Cluster D3

Cluster D3 was the third group to be inducted transcendently in CSR of human B cells and the cluster has 113 genes. IKZF2 and ADCY1 genes were the most differentially expressed.

IKZF2 encodes a member of the Ikaros family of zinc-finger proteins. Three members of this protein family (Ikaros, Aiolos and Helios) are hematopoietic-specific transcription factors involved in the regulation of lymphocyte development. This protein forms homo- or heterodimers with other Ikaros family members and has a function predominantly in early hematopoietic development. Helios is preferentially expressed at the mRNA level by regulatory T cells (Treg cells) and is potentially a specific marker of thymic-derived Treg cells. It raises the possibility that a significant percentage of Foxp3+ Treg cells are generated extrathymically [100]. ADCY1 gene encodes a member of the adenylate cyclase family that is primarily expressed in the brain. This protein is regulated by calcium/calmodulin concentration. Cyclic AMP (cAMP) production, which is important for mechanotransduction within the inner ear, is catalyzed by adenylate cyclases (AC). ADCY1 has an evolutionarily conserved role in hearing, and cAMP signaling is important to hair cell function within the inner ear [101].

The most significantly differential genes in each cluster and major pathways in each cluster are listed in Table 1 . The most interesting clusters are Cluster A1 and Cluster B1 and the full gene names of these two clusters are listed in Tables 2 and 3 .

Cluster Numbers Activation during CSR Most significant genes Pathways involved References
A1 153 Expression on CCL22; CCL17; TRAF; BCL2L1; MYB; VIM; TRIP10; FAS; PTGIR; EBI3; AHR; NCF2 ERK signaling, TRAF pathway; insulin pathway; NF-kB pathway [105, 106, 107, 108]
A2 83 Expression off MARCH1; FOSB; DUSP1; FOS; CR1; CR2; RGS2; PLD4; CCR6; RASGRP2; MOP-1; FCRLA Toll-like receptor signaling pathways, MAPK signaling pathway; innate immune system pathway [109, 110, 111]
B1 126 Sustained induction BHLHE40; IL17RB; NFIL3; HOMER2; AICDA; BATF3; ARID5A; DUSP4; CD80; TNFAIP2; XBP1; MTHFD2 Circadian rhythm pathway; IL-17 family signaling pathways; IL4-mediated signaling pathway [57, 112, 113]
B2 112 Sustained induction EPHB1; TNFSF4; DPYSL2; RPS6KA2; SLC41A1; AMICA1; MIIP; RGS9; CISH; LRRC32; AUH; SLC37A3 EPH-Ephrin signaling; TNF superfamily pathway; transport of glucose pathway [114, 115, 116]
C1 79 Transient induction TFEC; RRP12; SLC29A1; GPATCH4; SSRP1; BCL2A1; CYB561; NME1; TTLL12; FASN; NETO2; SLC27A4 C-MYB transcription factor network; apoptosis modulation and signaling [117, 118]
C2 112 Transient induction LMNB2; B4GALT5; SLC43A3; ESPL1; EZH2; PSMC3; SUV39H2; MREG; FSCN1; SRC; PHOSPHO2-KLHL23 Apoptosis pathway; glycosaminoglycan metabolism pathway [119, 120]
C3 105 Transient induction UHRF1; CHEK1; FANCI; CHAF1B; DTL; CDC6; EXO1; MCM6; CHAF1A; CDC45; TCF19 Chromatin regulation/acetylation pathway; DNA double-strand break repair pathway [121, 122]
C4 151 Transient induction KIF14; PRC1; NDC80; NUF2; HMMR; DEPDC1; AURKA; ARHGAP11B; BRCA1; FAM72B; HIST1H4L; DLGAP5; HIST1H1B Signaling by Rho GTPases; cell cycle pathway; DNA double-strand break repair pathway [85, 123, 124]
C5 99 Transient induction MCM10; PCNA; TCFL5; HELLS; ZC3HAV1L; PHF19; CARM1; VDR; LIMA1; MYH10; SEMA4A; TMOD1 Telomere C-strand synthesis pathway; apoptosis modulation and signaling; chromatin regulation/acetylation [125, 126, 127]
C6 128 Transient induction CCR4; HIST1H1C; CHRNA6; HIST1H3I; CCL1; GPR55; SYT11; PSTPIP2; KIAA1549L; HIST1H1D; PSAT1; TFDP2 Signaling by GPCR; apoptosis induced DNA fragmentation; nicotine pathway [128, 129, 130]
D1 69 Transient downregulation GPR18; TP53INP1; IFIT2; RNASET2; LBH; DOK3; FGD3; CD69; OAS1; ABCG1; PNOC; PARP15 Signaling by GPCR; p53 pathway, innate immune system; B cell development pathway [96, 131, 132, 133]
D2 69 Transient downregulation RAB6B; PM20D1; CYP2C19; CPNE4; TNFSF8; HIST1H2BD; METTL7A; ADHFE1; TMEM140; JMJD7; KLHL24; POU2AF1 Vesicle-mediated transport; drug metabolism; ERK signaling [134, 135, 136]
D3 113 Transient downregulation IKZF2; ADCY1; APOBEC3H; VAMP5; PDCD1LG2; CYP2C18; ILDR1; ADRB1; TM6SF1; GCSAM; CHAC1; ENPP3 mRNA editing—C to U conversion; NF-kB signaling; cytochrome P450 pathway [137, 138, 139]

Table 1.

The most significantly differential genes and pathways in each cluster during CSR.

Cluster A1: 153 genes
CCL22 CCL17 TRAF1 BCL2L1 MYB VIM TRIP10 FAS PTGIR EBI3
AHR NCF2 BCAT1 MGLL SEMA4C SIRPA TFPI2 ARNTL2 STAT5A HYOU1
ADA ICAM1 CFLAR MYO1C IGSF3 CCR7 SEPT11 WSB2 IL13RA1 CDK2AP1
CLIP2 NFKB2 PGD HIPK2 CD58 MARCKS SNX8 TNFAIP3 TNIP1 SLAMF7
SLC43A2 FADS1 FAM129A RAP2A YWHAG CYFIP1 LOC101929479 GNG8 IRF5 TMEM120A
FCER2 IPO7 CBX6 CDK6 BLVRA AP1S3 XPO5 PIK3R5 LSP1 FNBP1
TAGLN2 RBBP9 SPINT2 FYTTD1 GPR137B PEF1 PPA1 RANGAP1 LSM11 RHOF
WDR91 ATP1A1 MIR17HG SUPT16H USP14 CPNE8 MMP7 EEF2K PPME1 ANXA7
EHD4 CDR2 SRPK1 BLMH UBE2Z KDM2B PPP1R7 CAPRIN1 SF3B6 TALDO1
LMNA MCOLN2 SCD PCGF5 HERPUD1 NFKB1 CHURC1-FNTB ERH TJP2 BSG
VOPP1 KEAP1 CEP85 NAE1 UBTD2 TKT GOT2 SZRD1 APOL1 NSUN4
CHMP4B TRIM28 KCNN4 SLC16A1 SLC9A7 SLC20A1 MAP4 ZNF48 ECE1 ADCY3
STX11 LARP4 OTUD7B GPR183 PSMD8 ZAK SLC35F2 COA7 ANXA5 IDE
SLC39A1 MDFIC GSTP1 OGFOD1 SEC23B IMPDH2 HDLBP HTT KIF13B GTF3C4
ACSL4 MPC1 PITPNB TNIP2 EXOC3 MDH1 PLEKHA7 UBE2D4 JADE3 POR
PEA15 PIK3CD TP53BP2

Table 2.

The lists of Cluster A1.

Cluster B1: 126 genes
BHLHE40 IL17RB NFIL3 HOMER2 AICDA BATF3 ARID5A DUSP4 CD80 TNFAIP2
XBP1 MTHFD2 CD86 CD59 CAMK4 MFHAS1 SLC1A5 SRGN USP46 CHDH
HDGFR3 PIGX FLT1 RNF19B LTA NOD2 ZNF788 AARS ATXN1 RFC5
WARS PXDC1 PPP1R14A DENND5A QSOX1 STK38L PRR5L RGS10 SLC7A5 SCCPDH
RRAGD LY75-CD302 ADAMDEC1 YARS GPHN TRIM16L IRF4 NINJ1 SLC7A1 SOCS1
CD274 ECHDC3 NECAP2 TSPAN33 SEC11C LOXL3 AHRR RALB ARID3A RDX
CSF1 THG1L SLC39A8 SAMSN1 TXNL1 STK35 DARS TARS CLDND1 C12orf5
SEL1L CARS FAM162A VCL SEPHS1 XPOT ACSL1 GOT1 PFKM NSMCE1
RBM47 CEP19 ATXN2L DHRS3 RAB39B DCTN2 PABPC4 HIVEP1 CCDC126 ACADVL
MTX2 AEN TFG RBPJ SLC25A20 ETFDH COPA NR4A3 GPX4 ITFG3
DUSP22 CTNS IL2RA RAP1A TNFAIP1 PAM SLC37A1 DCTN6 AKAP2 RIPK2
RAB21 RPS23 KIAA1279 MARS ZNF267 CLCN5 NFKBID PRPSAP1 NEDD1 ZNF382
CDKN1A PRRT3 LYSMD1 NCK2 AZIN1 KIF5B

Table 3.

The lists of Cluster B1.

Advertisement

4. The future research on CSR in human B cells

A total of 1399 genes were shown to have differential expression during CSR in human B cells, and the novel genes have the roles in immune system process, cellular amine and DNA process and cell cycle phase or ribonucleoprotein biogenesis. Understanding the precisely functional roles of these novel genes in CSR in human B cells will bring new insights into the mechanisms of CSR and find potential therapeutic targets for human immune disorders such as allergic asthma and autoimmune diseases.

The next stage of research will also focus on determining how the naïve B cells turn into the specific IgE-, IgA- or IgG-releasing cells after T cell cytokines signal stimulation. The different stages of CSR in human B cell may contain unique transcription regulators for the destiny for each single cell. The development of single-cell sequencing provides a unique opportunity to explore the subsets of the human B cells to generate IgE, IgA and IgG. Obtaining high-quality single-cell sequencing data from B cells depends on efficient isolation of individual cells and amplifications of the genome or transcriptome of single cell to acquire sufficient materials for downstream analysis, identifying true variations from technological biases [102]. One of the major challenges of analyzing single-cell genomics data is to develop tools that differentiate technical artifacts and noise introduced during single-cell isolation, whole genome amplification, whole transcriptome amplification and sequencing from true biological variation. There are many factors that can influence the single-cell analysis. During single-cell isolation, the population of cells can be biased through the selection of cells based on size, viability or propensity to enter the cell cycle. Cells from cell lines as control may be problematic as they may not be diploid, and they can be highly aneuploid or even polyploidy. These will affect experimental performance [102]. Our understanding of human B cell function in CSR will derive from comparisons between healthy individuals and those with particular immunological diseases, and among groups of patients having the same disease with different clinical outcomes. For example, human SLE is clinically heterogeneous [103], making treatment decision challenging. It is important to know which B cell subsets are responsible for which functions in immune diseases, in addition to identifying how a “signature” profile for an individual subject’s collection of subsets may correlate with disease outcome that could eventually allow greater optimization of targeted therapies [104].

Advertisement

Abbreviation

ACadenylate cyclases
AIDactivation-induced cytidine deaminase
AP-1activating protein-1
APCantigen-presenting cell
APEapurinic/apyrimidic endonuclease
BCRsB cell receptors
Be-1B effector 1
Be-2B effector 2
CBcord blood
CLPscommon lymphoid progenitors
CSRclass switch recombination
DCsdendritic cells
DSBsdouble-strand DNA breaks
Epherythropoietin-producing hepatocellular
FIPFOS-interacting protein
FOfollicular
GCsgerminal centers
HDMhouse dust mite
hnRNPsheterogeneous nuclear ribonucleoproteins
IgHIg heavy-chain
MAPsmicrotubule-associated proteins
MAREsMaf-recognition elements
MCMmini-chromosome maintenance proteins
MZmarginal zone
NF-kBNF-kappa B
NHEJnonhomologous end-joining
NOnitrogen oxide
PKC-βprotein kinase C-β
PRC1protein regulator of cytokinesis 1
RAGrecombination activating gene
RTKsreceptor tyrosine kinases
SLEsystemic lupus erythematosus
TDT lymphocyte dependent
TFECtranscription factor EC
TIT lymphocyte independent
TNFtumor necrosis factor
TRAFTNF receptor associated factor
TRAPtartrate-resistant acid phosphatase
UNGuracil DNA glycosylase

References

  1. 1. Lund FE, Cytokine-producing B. Lymphocytes-key regulators of immunity. Current Opinion in Immunology. 2008;20(3):332-338
  2. 2. Hoffman W, Lakkis FG, Chalasani GB. Cells, antibodies, and more. Clinical Journal of the American Society of Nephrology: CJASN. 2016;11(1):137-154
  3. 3. Ribatti D, Crivellato E, Vacca A. The contribution of Bruce Glick to the definition of the role played by the bursa of Fabricius in the development of the B cell lineage. Clinical and Experimental Immunology. 2006;145(1):1-4
  4. 4. Welner RS, Pelayo R, Kincade PW. Evolving views on the genealogy of B cells. Nature Reviews Immunology. 2008;8(2):95-106
  5. 5. Harris DP, Goodrich S, Mohrs K, Mohrs M, Lund FE. Cutting edge: The development of IL-4-producing B cells (B effector 2 cells) is controlled by IL-4, IL-4 receptor alpha, and Th2 cells. Journal of Immunology. 2005;175(11):7103-7107
  6. 6. Harris DP, Goodrich S, Gerth AJ, Peng SL, Lund FE. Regulation of IFN-gamma production by B effector 1 cells: Essential roles for T-bet and the IFN-gamma receptor. Journal of Immunology. 2005;174(11):6781-6790
  7. 7. Lin YC, Jhunjhunwala S, Benner C, Heinz S, Welinder E, Mansson R, et al. A global network of transcription factors, involving E2A, EBF1 and Foxo1, that orchestrates B cell fate. Nature Immunology. 2010;11(7):635-643
  8. 8. Tonegawa S. Somatic generation of antibody diversity. Nature. 1983;302(5909):575-581
  9. 9. Nemazee D. Mechanisms of central tolerance for B cells. Nature Reviews Immunology. 2017;17(5):281-294
  10. 10. Pascual V, Liu YJ, Magalski A, de Bouteiller O, Banchereau J, Capra JD. Analysis of somatic mutation in five B cell subsets of human tonsil. The Journal of Experimental Medicine. 1994;180(1):329-339
  11. 11. Allen CD, Okada T, Cyster JG. Germinal-center organization and cellular dynamics. Immunity. 2007;27(2):190-202
  12. 12. Good-Jacobson KL, Shlomchik MJ. Plasticity and heterogeneity in the generation of memory B cells and long-lived plasma cells: The influence of germinal center interactions and dynamics. Journal of Immunology. 2010;185(6):3117-3125
  13. 13. Hauser AE, Junt T, Mempel TR, Sneddon MW, Kleinstein SH, Henrickson SE, et al. Definition of germinal-center B cell migration in vivo reveals predominant intrazonal circulation patterns. Immunity. 2007;26(5):655-667
  14. 14. Batista FD, Harwood NE. The who, how and where of antigen presentation to B cells. Nature Reviews Immunology. 2009;9(1):15-27
  15. 15. Okada T, Miller MJ, Parker I, Krummel MF, Neighbors M, Hartley SB, et al. Antigen-engaged B cells undergo chemotaxis toward the T zone and form motile conjugates with helper T cells. PLoS Biology. 2005;3(6):e150
  16. 16. Kerfoot SM, Yaari G, Patel JR, Johnson KL, Gonzalez DG, Kleinstein SH, et al. Germinal center B cell and T follicular helper cell development initiates in the interfollicular zone. Immunity. 2011;34(6):947-960
  17. 17. Kaminski DA, Wei C, Qian Y, Rosenberg AF, Sanz I. Advances in human B cell phenotypic profiling. Frontiers in Immunology. 2012;3:302
  18. 18. Wirths S, Lanzavecchia A. ABCB1 transporter discriminates human resting naive B cells from cycling transitional and memory B cells. European Journal of Immunology. 2005;35(12):3433-3441
  19. 19. Good KL, Avery DT, Tangye SG. Resting human memory B cells are intrinsically programmed for enhanced survival and responsiveness to diverse stimuli compared to naive B cells. Journal of Immunology. 2009;182(2):890-901
  20. 20. Tangye SG, Avery DT, Hodgkin PDA. Division-linked mechanism for the rapid generation of Ig-secreting cells from human memory B cells. Journal of Immunology. 2003;170(1):261-269
  21. 21. Oettinger MA, Schatz DG, Gorka C, Baltimore D. RAG-1 and RAG-2, adjacent genes that synergistically activate V(D)J recombination. Science. 1990;248(4962):1517-1523
  22. 22. Schatz DG, Oettinger MA, Baltimore D. The V(D)J recombination activating gene, RAG-1. Cell. 1989;59(6):1035-1048
  23. 23. Stavnezer J, Guikema JE, Schrader CE. Mechanism and regulation of class switch recombination. Annual Review of Immunology. 2008;26:261-292
  24. 24. Muramatsu M, Kinoshita K, Fagarasan S, Yamada S, Shinkai Y, Honjo T. Class switch recombination and hypermutation require activation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell. 2000;102(5):553-563
  25. 25. Okazaki IM, Kinoshita K, Muramatsu M, Yoshikawa K, Honjo T. The AID enzyme induces class switch recombination in fibroblasts. Nature. 2002;416(6878):340-345
  26. 26. Fear DJ. Mechanisms regulating the targeting and activity of activation induced cytidine deaminase. Current Opinion in Immunology. 2013;25(5):619-628
  27. 27. Dudley DD, Chaudhuri J, Bassing CH, Alt FW. Mechanism and control of V(D)J recombination versus class switch recombination: Similarities and differences. Advances in Immunology. 2005;86:43-112
  28. 28. Manis JP, Tian M, Alt FW. Mechanism and control of class-switch recombination. Trends in Immunology. 2002;23(1):31-39
  29. 29. Chowdhury M, Forouhi O, Dayal S, McCloskey N, Gould HJ, Felsenfeld G, et al. Analysis of intergenic transcription and histone modification across the human immunoglobulin heavy-chain locus. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(41):15872-15877
  30. 30. Jeppson JD, Patel HR, Sakata N, Domenico J, Terada N, Gelfand EW. Requirement for dual signals by anti-CD40 and IL-4 for the induction of nuclear factor-kappa B, IL-6, and IgE in human B lymphocytes. Journal of Immunology. 1998;161(4):1738-1742
  31. 31. Zhang Y, Fear DJ, Willis-Owen SA, Cookson WO, Moffatt MF. Global gene regulation during activation of immunoglobulin class switching in human B cells. Scientific Reports. 2016;6:37988
  32. 32. Itoh-Nakadai A, Hikota R, Muto A, Kometani K, Watanabe-Matsui M, Sato Y, et al. The transcription repressors Bach2 and Bach1 promote B cell development by repressing the myeloid program. Nature Immunology. 2014;15(12):1171-1180
  33. 33. Oyake T, Itoh K, Motohashi H, Hayashi N, Hoshino H, Nishizawa M, et al. Bach proteins belong to a novel family of BTB-basic leucine zipper transcription factors that interact with MafK and regulate transcription through the NF-E2 site. Molecular and Cellular Biology. 1996;16(11):6083-6095
  34. 34. Muto A, Tashiro S, Nakajima O, Hoshino H, Takahashi S, Sakoda E, et al. The transcriptional programme of antibody class switching involves the repressor Bach2. Nature. 2004;429(6991):566-571
  35. 35. Muto A, Ochiai K, Kimura Y, Itoh-Nakadai A, Calame KL, Ikebe D, et al. Bach2 represses plasma cell gene regulatory network in B cells to promote antibody class switch. The EMBO Journal. 2010;29(23):4048-4061
  36. 36. McManus S, Ebert A, Salvagiotto G, Medvedovic J, Sun Q, Tamir I, et al. The transcription factor Pax5 regulates its target genes by recruiting chromatin-modifying proteins in committed B cells. The EMBO Journal. 2011;30(12):2388-2404
  37. 37. Schwenzer R, Siemienski K, Liptay S, Schubert G, Peters N, Scheurich P, et al. The human tumor necrosis factor (TNF) receptor-associated factor 1 gene (TRAF1) is up-regulated by cytokines of the TNF ligand family and modulates TNF-induced activation of NF-kappaB and c-Jun N-terminal kinase. The Journal of Biological Chemistry. 1999;274(27):19368-19374
  38. 38. Werfel T, Heratizadeh A, Niebuhr M, Kapp A, Roesner LM, Karch A, et al. Exacerbation of atopic dermatitis on grass pollen exposure in an environmental challenge chamber. The Journal of Allergy and Clinical Immunology. 2015;136(1):96-103 e9
  39. 39. Abelius MS, Ernerudh J, Berg G, Matthiesen L, Nilsson LJ, Jenmalm MC. High cord blood levels of the T-helper 2-associated chemokines CCL17 and CCL22 precede allergy development during the first 6 years of life. Pediatric Research. 2011;70(5):495-500
  40. 40. Perros F, Hoogsteden HC, Coyle AJ, Lambrecht BN, Hammad H. Blockade of CCR4 in a humanized model of asthma reveals a critical role for DC-derived CCL17 and CCL22 in attracting Th2 cells and inducing airway inflammation. Allergy. 2009;64(7):995-1002
  41. 41. Hai T, Curran T. Cross-family dimerization of transcription factors Fos/Jun and ATF/CREB alters DNA binding specificity. Proceedings of the National Academy of Sciences of the United States of America. 1991;88(9):3720-3724
  42. 42. Lee W, Mitchell P, Tjian R. Purified transcription factor AP-1 interacts with TPA-inducible enhancer elements. Cell. 1987;49(6):741-752
  43. 43. Zhou H, Zarubin T, Ji Z, Min Z, Zhu W, Downey JS, et al. Frequency and distribution of AP-1 sites in the human genome. DNA Research: An International Journal for Rapid Publication of Reports on Genes and Genomes. 2005;12(2):139-150
  44. 44. Karin M, Liu Z, Zandi E. AP-1 function and regulation. Current Opinion in Cell Biology. 1997;9(2):240-246
  45. 45. Gustems M, Woellmer A, Rothbauer U, Eck SH, Wieland T, Lutter D, et al. c-Jun/c-Fos heterodimers regulate cellular genes via a newly identified class of methylated DNA sequence motifs. Nucleic Acids Research. 2014;42(5):3059-3072
  46. 46. Chiu R, Boyle WJ, Meek J, Smeal T, Hunter T, Karin M. The c-Fos protein interacts with c-Jun/AP-1 to stimulate transcription of AP-1 responsive genes. Cell. 1988;54(4):541-552
  47. 47. Davis BJ, Flanagan BF, Gilfillan AM, Metcalfe DD, Coleman JW. Nitric oxide inhibits IgE-dependent cytokine production and Fos and Jun activation in mast cells. Journal of Immunology. 2004;173(11):6914-6920
  48. 48. Lewin I, Jacob-Hirsch J, Zang ZC, Kupershtein V, Szallasi Z, Rivera J, et al. Aggregation of the Fc epsilon RI in mast cells induces the synthesis of Fos-interacting protein and increases its DNA binding-activity: The dependence on protein kinase C-beta. The Journal of Biological Chemistry. 1996;271(3):1514-1519
  49. 49. Zhao J, Harper R, Barchowsky A, Di YP. Identification of multiple MAPK-mediated transcription factors regulated by tobacco smoke in airway epithelial cells. American Journal of Physiology Lung Cellular and Molecular Physiology. 2007;293(2):L480-L490
  50. 50. Papin JA, Palsson BO. The JAK-STAT signaling network in the human B-cell: An extreme signaling pathway analysis. Biophysical Journal. 2004;87(1):37-46
  51. 51. Pollock R, Treisman R. Human SRF-related proteins: DNA-binding properties and potential regulatory targets. Genes & Development. 1991;5(12A):2327-2341
  52. 52. Wang L, Kurosaki T, Corey SJ. Engagement of the B-cell antigen receptor activates STAT through Lyn in a Jak-independent pathway. Oncogene. 2007;26(20):2851-2859
  53. 53. Mittrucker HW, Matsuyama T, Grossman A, Kundig TM, Potter J, Shahinian A, et al. Requirement for the transcription factor LSIRF/IRF4 for mature B and T lymphocyte function. Science. 1997;275(5299):540-543
  54. 54. Todd DJ, McHeyzer-Williams LJ, Kowal C, Lee AH, Volpe BT, Diamond B, et al. XBP1 governs late events in plasma cell differentiation and is not required for antigen-specific memory B cell development. The Journal of Experimental Medicine. 2009;206(10):2151-2159
  55. 55. Ise W, Kohyama M, Schraml BU, Zhang T, Schwer B, Basu U, et al. The transcription factor BATF controls the global regulators of class-switch recombination in both B cells and T cells. Nature Immunology. 2011;12(6):536-543
  56. 56. Kashiwada M, Levy DM, McKeag L, Murray K, Schroder AJ, Canfield SM, et al. IL-4-induced transcription factor NFIL3/E4BP4 controls IgE class switching. Proceedings of the National Academy of Sciences of the United States of America. 2010;107(2):821-826
  57. 57. Bie Q, Jin C, Zhang B, Dong H. IL-17B: A new area of study in the IL-17 family. Molecular Immunology. 2017;90:50-56
  58. 58. Wang H, Mobini R, Fang Y, Barrenas F, Zhang H, Xiang Z, et al. Allergen challenge of peripheral blood mononuclear cells from patients with seasonal allergic rhinitis increases IL-17RB, which regulates basophil apoptosis and degranulation. Clinical and Experimental Allergy: Journal of the British Society for Allergy and Clinical Immunology. 2010;40(8):1194-1202
  59. 59. Hunninghake GM, Chu JH, Sharma SS, Cho MH, Himes BE, Rogers AJ, et al. The CD4+ T-cell transcriptome and serum IgE in asthma: IL17RB and the role of sex. BMC Pulmonary Medicine. 2011;11:17
  60. 60. Jung JS, Park BL, Cheong HS, Bae JS, Kim JH, Chang HS, et al. Association of IL-17RB gene polymorphism with asthma. Chest. 2009;135(5):1173-1180
  61. 61. Lefebvre C, Rajbhandari P, Alvarez MJ, Bandaru P, Lim WK, Sato M, et al. A human B-cell interactome identifies MYB and FOXM1 as master regulators of proliferation in germinal centers. Molecular Systems Biology. 2010;6:377
  62. 62. Reinberg A, Gervais P, Levi F, Smolensky M, Del Cerro L, Ugolini C. Circadian and circannual rhythms of allergic rhinitis: An epidemiologic study involving chronobiologic methods. The Journal of Allergy and Clinical Immunology. 1988;81(1):51-62
  63. 63. Kelly EA, Houtman JJ, Jarjour NN. Inflammatory changes associated with circadian variation in pulmonary function in subjects with mild asthma. Clinical and Experimental Allergy: Journal of the British Society for Allergy and Clinical Immunology. 2004;34(2):227-233
  64. 64. Maurer M, Ortonne JP, Zuberbier T. Chronic urticaria: An internet survey of health behaviours, symptom patterns and treatment needs in European adult patients. The British Journal of Dermatology. 2009;160(3):633-641
  65. 65. Nakamura Y, Harama D, Shimokawa N, Hara M, Suzuki R, Tahara Y, et al. Circadian clock gene Period2 regulates a time-of-day-dependent variation in cutaneous anaphylactic reaction. The Journal of Allergy and Clinical Immunology. 2011;127(4):1038-45 e1-3
  66. 66. Sun H, Lu B, Li RQ, Flavell RA, Taneja R, Defective T. Cell activation and autoimmune disorder in Stra13-deficient mice. Nature Immunology. 2001;2(11):1040-1047
  67. 67. Yu X, Rollins D, Ruhn KA, Stubblefield JJ, Green CB, Kashiwada M, et al. TH17 cell differentiation is regulated by the circadian clock. Science. 2013;342(6159):727-730
  68. 68. Nakada M, Hayashi Y, Hamada J. Role of Eph/ephrin tyrosine kinase in malignant glioma. Neuro-Oncology. 2011;13(11):1163-1170
  69. 69. Gucciardo E, Sugiyama N, Lehti K. Eph- and ephrin-dependent mechanisms in tumor and stem cell dynamics. Cellular and Molecular Life Sciences: CMLS. 2014;71(19):3685-3710
  70. 70. Wei W, Wang H, Ji S. Paradoxes of the EphB1 receptor in malignant brain tumors. Cancer Cell International. 2017;17:21
  71. 71. Davis S, Gale NW, Aldrich TH, Maisonpierre PC, Lhotak V, Pawson T, et al. Ligands for EPH-related receptor tyrosine kinases that require membrane attachment or clustering for activity. Science. 1994;266(5186):816-819
  72. 72. Holen HL, Zernichow L, Fjelland KE, Evenroed IM, Tveit H, Aasheim HC. Ephrin-B3 binds specifically to B lymphocytes in blood and induces migration. Scandinavian Journal of Immunology. 2011;74(2):144-154
  73. 73. Cunninghame Graham DS, Graham RR, Manku H, Wong AK, Whittaker JC, Gaffney PM, et al. Polymorphism at the TNF superfamily gene TNFSF4 confers susceptibility to systemic lupus erythematosus. Nature Genetics. 2008;40(1):83-89
  74. 74. Gourh P, Arnett FC, Tan FK, Assassi S, Divecha D, Paz G, et al. Association of TNFSF4 (OX40L) polymorphisms with susceptibility to systemic sclerosis. Annals of the Rheumatic Diseases. 2010;69(3):550-555
  75. 75. Sun X, Haider Ali MSS, Moran M. The role of interactions of long non-coding RNAs and heterogeneous nuclear ribonucleoproteins in regulating cellular functions. The Biochemical Journal. 2017;474(17):2925-2935
  76. 76. Chung MC, Kim HK, Kawamoto S. TFEC can function as a transcriptional activator of the nonmuscle myosin II heavy chain—A gene in transfected cells. Biochemistry. 2001;40(30):8887-8897
  77. 77. Zhao GQ, Zhao Q, Zhou X, Mattei MG, de Crombrugghe B. TFEC, a basic helix-loop-helix protein, forms heterodimers with TFE3 and inhibits TFE3-dependent transcription activation. Molecular and Cellular Biology. 1993;13(8):4505-4512
  78. 78. Mahony CB, Fish RJ, Pasche C, Bertrand JY. tfec controls the hematopoietic stem cell vascular niche during zebrafish embryogenesis. Blood. 2016;128(10):1336-1345
  79. 79. Moriggi G, Nieto B, Dosil M. Rrp12 and the Exportin Crm1 participate in late assembly events in the nucleolus during 40S ribosomal subunit biogenesis. PLoS Genetics. 2014;10(12):e1004836
  80. 80. Yang SH, Jung HJ, Coffinier C, Fong LG, Young SG. Are B-type lamins essential in all mammalian cells? Nucleus. 2011;2(6):562-569
  81. 81. Bolotin E, Armendariz A, Kim K, Heo SJ, Boffelli D, Tantisira K, et al. Statin-induced changes in gene expression in EBV-transformed and native B-cells. Human Molecular Genetics. 2014;23(5):1202-1210
  82. 82. Ranuncolo SM, Polo JM, Melnick A. BCL6 represses CHEK1 and suppresses DNA damage pathways in normal and malignant B-cells. Blood Cells, Molecules & Diseases. 2008;41(1):95-99
  83. 83. Ly LL, Yoshida H, Yamaguchi M. Nuclear transcription factor Y and its roles in cellular processes related to human disease. American Journal of Cancer Research. 2013;3(4):339-346
  84. 84. Corson TW, Huang A, Tsao MS, Gallie BL. KIF14 is a candidate oncogene in the 1q minimal region of genomic gain in multiple cancers. Oncogene. 2005;24(30):4741-4753
  85. 85. Jiang W, Jimenez G, Wells NJ, Hope TJ, Wahl GM, Hunter T, et al. PRC1: A human mitotic spindle-associated CDK substrate protein required for cytokinesis. Molecular Cell. 1998;2(6):877-885
  86. 86. Tang H, Xiao G, Behrens C, Schiller J, Allen J, Chow CW, et al. A 12-gene set predicts survival benefits from adjuvant chemotherapy in non-small cell lung cancer patients. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2013;19(6):1577-1586
  87. 87. Glotzer M. The molecular requirements for cytokinesis. Science. 2005;307(5716):1735-1739
  88. 88. Gruneberg U, Neef R, Li X, Chan EH, Chalamalasetty RB, Nigg EA, et al. KIF14 and citron kinase act together to promote efficient cytokinesis. The Journal of Cell Biology. 2006;172(3):363-372
  89. 89. Chattopadhyay S, Bielinsky AK. Human Mcm10 regulates the catalytic subunit of DNA polymerase-alpha and prevents DNA damage during replication. Molecular Biology of the Cell. 2007;18(10):4085-4095
  90. 90. Xu X, Rochette PJ, Feyissa EA, TV S, Liu Y. MCM10 mediates RECQ4 association with MCM2–7 helicase complex during DNA replication. The EMBO Journal. 2009;28(19):3005-3014
  91. 91. Kliszczak M, Sedlackova H, Pitchai GP, Streicher WW, Krejci L, Hickson ID. Interaction of RECQ4 and MCM10 is important for efficient DNA replication origin firing in human cells. Oncotarget. 2015;6(38):40464-40479
  92. 92. Strzalka W, Ziemienowicz A. Proliferating cell nuclear antigen (PCNA): A key factor in DNA replication and cell cycle regulation. Annals of Botany. 2011;107(7):1127-1140
  93. 93. Nakano T, Kamei R, Fujimura T, Takaoka Y, Hori A, Lai CY, et al. Impact of histone H1 on the progression of allergic rhinitis and its suppression by neutralizing antibody in mice. PLoS One. 2016;11(4):e0153630
  94. 94. Kohno M, Hasegawa H, Inoue A, Muraoka M, Miyazaki T, Oka K, et al. Identification of N-arachidonylglycine as the endogenous ligand for orphan G-protein-coupled receptor GPR18. Biochemical and Biophysical Research Communications. 2006;347(3):827-832
  95. 95. Flegel C, Vogel F, Hofreuter A, Wojcik S, Schoeder C, Kiec-Kononowicz K, et al. Characterization of non-olfactory GPCRs in human sperm with a focus on GPR18. Scientific Reports. 2016;6:32255
  96. 96. Shahbazi J, Lock R, Liu T. Tumor protein 53-induced nuclear protein 1 enhances p53 function and represses tumorigenesis. Frontiers in Genetics. 2013;4:80
  97. 97. Seillier M, Peuget S, Gayet O, Gauthier C, N’Guessan P, Monte M, et al. TP53INP1, a tumor suppressor, interacts with LC3 and ATG8-family proteins through the LC3-interacting region (LIR) and promotes autophagy-dependent cell death. Cell Death and Differentiation. 2012;19(9):1525-1535
  98. 98. Del Nery E, Miserey-Lenkei S, Falguieres T, Nizak C, Johannes L, Perez F, et al. Rab6A and Rab6A’ GTPases play non-overlapping roles in membrane trafficking. Traffic. 2006;7(4):394-407
  99. 99. Opdam FJ, Echard A, Croes HJ, van den Hurk JA, van de Vorstenbosch RA, Ginsel LA, et al. The small GTPase Rab6B, a novel Rab6 subfamily member, is cell-type specifically expressed and localised to the Golgi apparatus. Journal of Cell Science. 2000;113(Pt 15):2725-2735
  100. 100. Thornton AM, Korty PE, Tran DQ, Wohlfert EA, Murray PE, Belkaid Y, et al. Expression of Helios, an Ikaros transcription factor family member, differentiates thymic-derived from peripherally induced Foxp3+ T regulatory cells. Journal of Immunology. 2010;184(7):3433-3441
  101. 101. Santos-Cortez RL, Lee K, Giese AP, Ansar M, Amin-Ud-Din M, Rehn K, et al. Adenylate cyclase 1 (ADCY1) mutations cause recessive hearing impairment in humans and defects in hair cell function and hearing in zebrafish. Human Molecular Genetics. 2014;23(12):3289-3298
  102. 102. Gawad C, Koh W, Quake SR. Single-cell genome sequencing: Current state of the science. Nature Reviews Genetics. 2016;17(3):175-188
  103. 103. Bertsias GK, Salmon JE, Boumpas DT. Therapeutic opportunities in systemic lupus erythematosus: State of the art and prospects for the new decade. Annals of the Rheumatic Diseases. 2010;69(9):1603-1611
  104. 104. Sanz I, Lee FEB. Cells as therapeutic targets in SLE. Nature Reviews Rheumatology. 2010;6(6):326-337
  105. 105. Inoue M, Yamada J, Aomatsu-Kikuchi E, Satoh K, Kondo H, Ishisaki A, et al. SCRG1 suppresses LPS-induced CCL22 production through ERK1/2 activation in mouse macrophage Raw264.7 cells. Molecular Medicine Reports. 2017;15(6):4069-4076
  106. 106. Oeckinghaus A, Hayden MS, Ghosh S. Crosstalk in NF-kappaB signaling pathways. Nature Immunology. 2011;12(8):695-708
  107. 107. Das UN, Rao AA. Gene expression profile in obesity and type 2 diabetes mellitus. Lipids in Health and Disease. 2007;6:35
  108. 108. Suhasini M, Pilz RB. Transcriptional elongation of c-myb is regulated by NF-kappaB (p50/RelB). Oncogene. 1999;18(51):7360-7369
  109. 109. Kawai T, Akira S. TLR signaling. Cell Death and Differentiation. 2006;13(5):816-825
  110. 110. Kondoh K, Nishida E. Regulation of MAP kinases by MAP kinase phosphatases. Biochimica et Biophysica Acta. 2007;1773(8):1227-1237
  111. 111. Dunkelberger JR, Song WC. Complement and its role in innate and adaptive immune responses. Cell Research. 2010;20(1):34-50
  112. 112. Li JZ, Bunney BG, Meng F, Hagenauer MH, Walsh DM, Vawter MP, et al. Circadian patterns of gene expression in the human brain and disruption in major depressive disorder. Proceedings of the National Academy of Sciences of the United States of America. 2013;110(24):9950-9955
  113. 113. Pone EJ, Zhang J, Mai T, White CA, Li G, Sakakura JK, et al. BCR-signalling synergizes with TLR-signalling for induction of AID and immunoglobulin class-switching through the non-canonical NF-kappaB pathway. Nature Communications. 2012;3:767
  114. 114. Coulthard MG, Morgan M, Woodruff TM, Arumugam TV, Taylor SM, Carpenter TC, et al. Eph/Ephrin signaling in injury and inflammation. The American Journal of Pathology. 2012;181(5):1493-1503
  115. 115. Lee SJ, Park SS, Cho YH, Park K, Kim EJ, Jung KH, et al. Activation of matrix metalloproteinase-9 by TNF-alpha in human urinary bladder cancer HT1376 cells: The role of MAP kinase signaling pathways. Oncology Reports. 2008;19(4):1007-1013
  116. 116. Mastrototaro L, Tietjen U, Sponder G, Vormann J, Aschenbach JR, Kolisek M. Insulin modulates the Na+/Mg2+ exchanger SLC41A1 and influences Mg2+ efflux from intracellular stores in transgenic HEK293 cells. The Journal of Nutrition. 2015;145(11):2440-2447
  117. 117. Miller AJ, Levy C, Davis IJ, Razin E, Fisher DE. Sumoylation of MITF and its related family members TFE3 and TFEB. The Journal of Biological Chemistry. 2005;280(1):146-155
  118. 118. Pecina-Slaus N. Wnt signal transduction pathway and apoptosis: A review. Cancer Cell International. 2010;10:22
  119. 119. Yang SH, Chang SY, Yin L, Tu Y, Hu Y, Yoshinaga Y, et al. An absence of both lamin B1 and lamin B2 in keratinocytes has no effect on cell proliferation or the development of skin and hair. Human Molecular Genetics. 2011;20(18):3537-3544
  120. 120. Shi YF, Fong CC, Zhang Q, Cheung PY, Tzang CH, RS W, et al. Hypoxia induces the activation of human hepatic stellate cells LX-2 through TGF-beta signaling pathway. FEBS Letters. 2007;581(2):203-210
  121. 121. Du J, Johnson LM, Jacobsen SE, Patel DJDNA. Methylation pathways and their crosstalk with histone methylation. Nature Reviews Molecular Cell Biology. 2015;16(9):519-532
  122. 122. Khanna KK, Jackson SP. DNA double-strand breaks: Signaling, repair and the cancer connection. Nature Genetics. 2001;27(3):247-254
  123. 123. Tseliou M, Al-Qahtani A, Alarifi S, Alkahtani SH, Stournaras C, Sourvinos G. The role of RhoA, RhoB and RhoC GTPases in cell morphology, proliferation and migration in human cytomegalovirus (HCMV) infected glioblastoma cells. Cellular Physiology and Biochemistry: International Journal of Experimental Cellular Physiology, Biochemistry, and Pharmacology. 2016;38(1):94-109
  124. 124. Katsha A, Belkhiri A, Goff L, El-Rifai W. Aurora kinase A in gastrointestinal cancers: Time to target. Molecular Cancer. 2015;14:106
  125. 125. Moldovan GL, Pfander B, Jentsch S. PCNA the maestro of the replication fork. Cell. 2007;129(4):665-679
  126. 126. Huen MS, Chen J. The DNA damage response pathways: At the crossroad of protein modifications. Cell Research. 2008;18(1):8-16
  127. 127. Mjelle R, Hegre SA, Aas PA, Slupphaug G, Drablos F, Saetrom P, et al. Cell cycle regulation of human DNA repair and chromatin remodeling genes. DNA Repair. 2015;30:53-67
  128. 128. Shi GX, Harrison K, Han SB, Moratz C, Kehrl JH. Toll-like receptor signaling alters the expression of regulator of G protein signaling proteins in dendritic cells: Implications for G protein-coupled receptor signaling. Journal of Immunology. 2004;172(9):5175-5184
  129. 129. Millan-Arino L, Islam AB, Izquierdo-Bouldstridge A, Mayor R, Terme JM, Luque N, et al. Mapping of six somatic linker histone H1 variants in human breast cancer cells uncovers specific features of H1.2. Nucleic Acids Research. 2014;42(7):4474-4493
  130. 130. Wen L, Yang Z, Cui W, Li MD. Crucial roles of the CHRNB3-CHRNA6 gene cluster on chromosome 8 in nicotine dependence: Update and subjects for future research. Translational Psychiatry. 2016;6(6):e843
  131. 131. Yin H, Chu A, Li W, Wang B, Shelton F, Otero F, et al. Lipid G protein-coupled receptor ligand identification using beta-arrestin PathHunter assay. The Journal of Biological Chemistry. 2009;284(18):12328-12338
  132. 132. Perwitasari O, Cho H, Diamond MS, Gale M Jr. Inhibitor of kappaB kinase epsilon (IKK(epsilon)), STAT1, and IFIT2 proteins define novel innate immune effector pathway against West Nile virus infection. The Journal of Biological Chemistry. 2011;286(52):44412-44423
  133. 133. Acquati F, Monti L, Lualdi M, Fabbri M, Sacco MG, Gribaldo L, et al. Molecular signature induced by RNASET2, a tumor antagonizing gene, in ovarian cancer cells. Oncotarget. 2011;2(6):477-484
  134. 134. Schlager MA, Kapitein LC, Grigoriev I, Burzynski GM, Wulf PS, Keijzer N, et al. Pericentrosomal targeting of Rab6 secretory vesicles by Bicaudal-D-related protein 1 (BICDR-1) regulates neuritogenesis. The EMBO Journal. 2010;29(10):1637-1651
  135. 135. Zanger UM, Schwab M. Cytochrome P450 enzymes in drug metabolism: Regulation of gene expression, enzyme activities, and impact of genetic variation. Pharmacology & Therapeutics. 2013;138(1):103-141
  136. 136. Ward-Kavanagh LK, Lin WW, Sedy JR, Ware CF. The TNF receptor superfamily in co-stimulating and co-inhibitory responses. Immunity. 2016;44(5):1005-1019
  137. 137. Li MM, Emerman M. Polymorphism in human APOBEC3H affects a phenotype dominant for subcellular localization and antiviral activity. Journal of Virology. 2011;85(16):8197-8207
  138. 138. Steidl C, Gascoyne RD. The molecular pathogenesis of primary mediastinal large B-cell lymphoma. Blood. 2011;118(10):2659-2669
  139. 139. Elliot DJ, Suharjono LBC, Gillam EM, Birkett DJ, Gross AS, et al. Identification of the human cytochromes P450 catalysing the rate-limiting pathways of gliclazide elimination. British Journal of Clinical Pharmacology. 2007;64(4):450-457

Written By

Youming Zhang

Submitted: 17 July 2017 Reviewed: 23 October 2017 Published: 21 February 2018