6
Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi 1 , Kate Senger 1 , Tao Sai 1 , Yan Qu, Yonglian Sun, Kajal Hamidzadeh, Allen Nguyen, Zhaoyu Jin, Meijuan Zhou, Donghong Yan, Wei Yu Lin, Zhonghua Lin, Maria N. Lorenzo, Andrew Sebrell, Jiabing Ding, Min Xu, Patrick Caplazi, Cary D. Austin, Mercedesz Balazs, Merone Roose-Girma, Laura DeForge, Søren Warming, Wyne P. Lee, Vishva M. Dixit 2 , and Ali A. Zarrin 2 Genentech, Inc., South San Francisco, CA 94080 Edited by Frederick W. Alt, Program in Cellular and Molecular Medicine, Boston Childrens Hospital; and Howard Hughes Medical Institute, Harvard Medical School, Boston, MA, and approved August 9, 2013 (received for review December 18, 2012) Preceding antibody constant regions are switch (S) regions varying in length and repeat density that are targets of activation-induced cytidine deaminase. We asked how participating S regions in- uence each other to orchestrate rearrangements at the IgH locus by engineering mice in which the weakest S region, Se, is replaced with prominent recombination hotspot Sμ. These mice produce copious polyclonal IgE upon challenge, providing a platform to study IgE biology and therapeutic interventions. The insertion enhances e germ-line transcript levels, shows a preference for di- rect vs. sequential switching, and reduces intraswitch recombina- tion events at native Sμ. These results suggest that the sufciency of Sμ to mediate IgH rearrangements may be inuenced by context- dependent cues. immunoglobulin | AICDA | asthma | allergy | germline transcription S witch (S) regions are essential and specialized targets of ac- tivation-induced cytidine deaminase (AID) (13) that are ordered 5-Sμ-Sγ3-Sγ1-Sγ2b-Sγ2a-Se-Sα-3(4) in the mouse IgH locus (Fig. 1A). Joining of distant dsDNA breaks (DSBs) be- tween donor Sμ and any downstream S region constitutes class switch recombination (CSR). CSR to specic constant heavy- chain genes is subject to tight transcriptional regulation, which increases the accessibility of a given S region before CSR (5, 6). The primary role of S regions is to seed DSBs (7), which are repaired by nonhomologous end joining (8) predominately dur- ing the G1 phase of the cell cycle (9), whereas homologous re- combination is dispensable in CSR (8, 10, 11). AID initiates CSR by targeting cytidines in transcribed repeat-rich S regions (1214). Limited amounts of active AID in B cells can be inferred from recent studies that showed that AID heterozygous mice have reduced levels of somatic hypermutation (SHM) and CSR (1518). Epigenetic modications during CSR are emerging as an important regulatory mechanism to inuence CSR (19). Thus, complex regulatory mechanisms act in concert to ensure appro- priate AID regulation, likely to limit its off-target activity (20). S regions have acquired intrinsic properties to make them the ultimate substrate for AID within the genome (21). Ancient S regions resemble SHM substrates, except they have a higher density of hotspots. The density of hotspots in S regions is sig- nicantly higher than in V regions (4), potentially creating areas highly susceptible to DSBs (4). In mammals, S regions appear to have further diverged by incorporating features such as the ability to form R-loops, which are single-stranded DNA loops formed by the association of an RNA transcript with a DNA template (22) and G-quartets, which are four-stranded structures of guanine-rich DNA (23), to maximize them as targets for AID (24). S region length enhances CSR (25), and there is an inverse correlation between the distance of DSBs and recombination frequency (7). In mice, Se is one of the shortest and least re- petitive S regions, and, with the exception of Sα, it is the farthest from Sμ (4). CSR to Se involves sequential CSR between Sμ and Sγ1 before combining with Se (2629); however, a sequential pathway is not required, as direct CSR between Sμ and Se occurs when Sγ1 is genetically ablated (30, 31). It is possible for multiple DSBs to occur within a single S region, which leads to intraswitch recombination (ISR). This phenomenon is seen more frequently in Sμ (28, 32, 33) than in other S regions in the context of the normal IgH locus, possibly because it is enriched in AID target motif sites (4) or because context-dependent cues regulate the targeting of the donor Sμ region. ISR in downstream acceptor S regions are more abundant in Sμ /(28) or transcriptionally in- active Sμ mutants (34), suggesting acceptor S regions are able to mount ISR in the absence of Sμ. To generate a polyclonal hyper-IgEmouse model and to gain insights into how S regions work outside their native context, we created a mouse model in which the weakest S region, Se, was replaced with the strongest AID hotspot, Sμ.Sμ knock-in (SμKI) mice produce abundant IgE at the expense of other isotypes. SμKI IgE is antigen (Ag)-specic and produced in response to a variety of local and systemic stimuli. On a mechanistic level, the presence of Sμ in place of Se enhances e germ-line transcript (GLT), sug- gesting its presence inuences accessibility of the locus. Circle transcript studies reveal a preference for direct CSR vs. sequential in SμKI mice. The knocked-in switch also negatively affects ISR of endogenous Sμ. Taken together, these results suggest the Sμ se- quence has properties that are at least in part context-dependent. Results SμKI Modied e Allele Outcompetes Sγ1 to Produce High Amounts of IgE in Stimulated B Cells. An overview of the IgH locus is shown in Signicance Switch (S) regions are repetitive DNA sequences. During an immune response, one of several S regions recombine with a donor switch (Sμ) that is constitutively on,resulting in the production of antibodies with new functions. Donor Sμ is large and very repeat-rich, while another switch, Se, is less than half its size with a low density of repeats. We replaced Se with Sμ in mice. These mice switch to Se more effectively and produce high levels of IgE antibodies implicated in asthma, making this a useful model to study disease. In addition, placing Sμ outside of its native context revealed insights into how switches work. Author contributions: S.M., K.S., V.M.D., and A.A.Z. designed research; S.M., K.S., T.S., Y.Q., Y.S., K.H., A.N., Z.J., M.Z., D.Y., W.Y.L., Z.L., A.S., J.D., M.X., P.C., C.D.A., M.B., W.P.L., and A.A.Z. performed research; T.S., M.N.L., M.R.-G., and S.W. contributed new reagents/ analytic tools; S.M., K.S., T.S., Y.Q., Y.S., K.H., A.N., Z.J., W.Y.L., Z.L., J.D., M.X., P.C., M.B., L.D., W.P.L., V.M.D., and A.A.Z. analyzed data; and S.M., K.S., V.M.D., and A.A.Z. wrote the paper. The authors declare no conict of interest. This article is a PNAS Direct Submission. Freely available online through the PNAS open access option. 1 S.M., K.S., and T.S. contributed equally to this work. 2 To whom correspondence may be addressed. E-mail: [email protected] or zarrin. [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1221661110/-/DCSupplemental. 1577015775 | PNAS | September 24, 2013 | vol. 110 | no. 39 www.pnas.org/cgi/doi/10.1073/pnas.1221661110 Downloaded by guest on April 30, 2021

Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

Polyclonal hyper-IgE mouse model reveals mechanisticinsights into antibody class switch recombinationShahram Misaghi1, Kate Senger1, Tao Sai1, Yan Qu, Yonglian Sun, Kajal Hamidzadeh, Allen Nguyen, Zhaoyu Jin,Meijuan Zhou, Donghong Yan, Wei Yu Lin, Zhonghua Lin, Maria N. Lorenzo, Andrew Sebrell, Jiabing Ding, Min Xu,Patrick Caplazi, Cary D. Austin, Mercedesz Balazs, Merone Roose-Girma, Laura DeForge, Søren Warming, Wyne P. Lee,Vishva M. Dixit2, and Ali A. Zarrin2

Genentech, Inc., South San Francisco, CA 94080

Edited by Frederick W. Alt, Program in Cellular and Molecular Medicine, Boston Children’s Hospital; and Howard Hughes Medical Institute, Harvard MedicalSchool, Boston, MA, and approved August 9, 2013 (received for review December 18, 2012)

Preceding antibody constant regions are switch (S) regions varyingin length and repeat density that are targets of activation-inducedcytidine deaminase. We asked how participating S regions in-fluence each other to orchestrate rearrangements at the IgH locusby engineering mice in which the weakest S region, Se, is replacedwith prominent recombination hotspot Sμ. These mice producecopious polyclonal IgE upon challenge, providing a platform tostudy IgE biology and therapeutic interventions. The insertionenhances e germ-line transcript levels, shows a preference for di-rect vs. sequential switching, and reduces intraswitch recombina-tion events at native Sμ. These results suggest that the sufficiencyof Sμ to mediate IgH rearrangements may be influenced by context-dependent cues.

immunoglobulin | AICDA | asthma | allergy | germline transcription

Switch (S) regions are essential and specialized targets of ac-tivation-induced cytidine deaminase (AID) (1–3) that are

ordered 5′-Sμ-Sγ3-Sγ1-Sγ2b-Sγ2a-Se-Sα-3′ (4) in the mouse IgHlocus (Fig. 1A). Joining of distant dsDNA breaks (DSBs) be-tween donor Sμ and any downstream S region constitutes classswitch recombination (CSR). CSR to specific constant heavy-chain genes is subject to tight transcriptional regulation, whichincreases the accessibility of a given S region before CSR (5, 6).The primary role of S regions is to seed DSBs (7), which arerepaired by nonhomologous end joining (8) predominately dur-ing the G1 phase of the cell cycle (9), whereas homologous re-combination is dispensable in CSR (8, 10, 11). AID initiates CSRby targeting cytidines in transcribed repeat-rich S regions (12–14). Limited amounts of active AID in B cells can be inferredfrom recent studies that showed that AID heterozygous micehave reduced levels of somatic hypermutation (SHM) and CSR(15–18). Epigenetic modifications during CSR are emerging asan important regulatory mechanism to influence CSR (19). Thus,complex regulatory mechanisms act in concert to ensure appro-priate AID regulation, likely to limit its off-target activity (20).S regions have acquired intrinsic properties to make them the

ultimate substrate for AID within the genome (21). Ancient Sregions resemble SHM substrates, except they have a higherdensity of hotspots. The density of hotspots in S regions is sig-nificantly higher than in V regions (4), potentially creating areashighly susceptible to DSBs (4). In mammals, S regions appear tohave further diverged by incorporating features such as theability to form R-loops, which are single-stranded DNA loopsformed by the association of an RNA transcript with a DNAtemplate (22) and G-quartets, which are four-stranded structuresof guanine-rich DNA (23), to maximize them as targets for AID(24). S region length enhances CSR (25), and there is an inversecorrelation between the distance of DSBs and recombinationfrequency (7). In mice, Se is one of the shortest and least re-petitive S regions, and, with the exception of Sα, it is the farthestfrom Sμ (4). CSR to Se involves sequential CSR between Sμ andSγ1 before combining with Se (26–29); however, a sequentialpathway is not required, as direct CSR between Sμ and Se occurs

when Sγ1 is genetically ablated (30, 31). It is possible for multipleDSBs to occur within a single S region, which leads to intraswitchrecombination (ISR). This phenomenon is seen more frequentlyin Sμ (28, 32, 33) than in other S regions in the context of thenormal IgH locus, possibly because it is enriched in AID targetmotif sites (4) or because context-dependent cues regulate thetargeting of the donor Sμ region. ISR in downstream acceptor Sregions are more abundant in Sμ−/− (28) or transcriptionally in-active Sμ mutants (34), suggesting acceptor S regions are able tomount ISR in the absence of Sμ.To generate a polyclonal “hyper-IgE” mouse model and to gain

insights into how S regions work outside their native context, wecreated a mouse model in which the weakest S region, Se, wasreplaced with the strongest AID hotspot, Sμ. Sμ knock-in (SμKI)mice produce abundant IgE at the expense of other isotypes. SμKIIgE is antigen (Ag)-specific and produced in response to a varietyof local and systemic stimuli. On a mechanistic level, the presenceof Sμ in place of Se enhances e germ-line transcript (GLT), sug-gesting its presence influences accessibility of the locus. Circletranscript studies reveal a preference for direct CSR vs. sequentialin SμKI mice. The knocked-in switch also negatively affects ISR ofendogenous Sμ. Taken together, these results suggest the Sμ se-quence has properties that are at least in part context-dependent.

ResultsSμKI Modified e Allele Outcompetes Sγ1 to Produce High Amounts ofIgE in Stimulated B Cells. An overview of the IgH locus is shown in

Significance

Switch (S) regions are repetitive DNA sequences. During animmune response, one of several S regions recombine with adonor switch (Sμ) that is constitutively “on,” resulting in theproduction of antibodies with new functions. Donor Sμ is largeand very repeat-rich, while another switch, Se, is less than halfits size with a low density of repeats. We replaced Sewith Sμ inmice. These mice switch to Se more effectively and producehigh levels of IgE antibodies implicated in asthma, making thisa useful model to study disease. In addition, placing Sμ outsideof its native context revealed insights into how switches work.

Author contributions: S.M., K.S., V.M.D., and A.A.Z. designed research; S.M., K.S., T.S.,Y.Q., Y.S., K.H., A.N., Z.J., M.Z., D.Y., W.Y.L., Z.L., A.S., J.D., M.X., P.C., C.D.A., M.B., W.P.L.,and A.A.Z. performed research; T.S., M.N.L., M.R.-G., and S.W. contributed new reagents/analytic tools; S.M., K.S., T.S., Y.Q., Y.S., K.H., A.N., Z.J., W.Y.L., Z.L., J.D., M.X., P.C., M.B.,L.D., W.P.L., V.M.D., and A.A.Z. analyzed data; and S.M., K.S., V.M.D., and A.A.Z. wrotethe paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

Freely available online through the PNAS open access option.1S.M., K.S., and T.S. contributed equally to this work.2To whom correspondence may be addressed. E-mail: [email protected] or [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1221661110/-/DCSupplemental.

15770–15775 | PNAS | September 24, 2013 | vol. 110 | no. 39 www.pnas.org/cgi/doi/10.1073/pnas.1221661110

Dow

nloa

ded

by g

uest

on

Apr

il 30

, 202

1

Page 2: Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

Fig. 1A. A diagram of probes used in Southern blotting analysesis shown in Fig. 1B. Gene targeting was used to replace murineSe (∼2 kb) with Sμ (∼5 kb) in C57BL/6 ES cells (Fig. S1 A andB). Homozygous knock-in mice are referred to hereafter as SμKImice, and all animals were maintained in a C57BL/6 background.A phenotypic analysis of the SμKI mice did not reveal anydefects in lymphocyte development (Fig. S1C). The effect ofinserted Sμ on CSR was assessed by FACS and ELISA analysisof splenocytes stimulated with LPS plus IL-4 or LPS alone for4 d (Fig. 2). In cells stimulated with LPS and IL-4 (Fig. 2A), WTand SμKI cultures contained similar numbers of IgM+ B cells.However, in the SμKI cultures, IgG1+ B cells were decreased∼11-fold, whereas the level of IgE+ B cells in the SμKI culturesincreased approximately sixfold relative to WT. As expected, thiseffect was AID-dependent, as AID-deficient splenocytes (35)yielded negligible IgG1+ or IgE+ B cells. In response to LPSstimulation, WT and SμKI cells produced comparable percen-tages of IgM+ B cells (Fig. 2B). A ∼50% decrease in IgG3 levelswas seen in the SμKI, whereas an approximately fourfold in-crease in IgE+ cells was seen in SμKI cultures relative to WTcultures. ELISAs conducted 6 d after stimulation showed similartradeoffs in secreted IgG and IgE (Fig. 2 C andD). With LPS/IL-4stimulation (Fig. 2C), IgM levels were comparable among WT,heterozygous, and SμKI cultures; however, IgE was far moreabundant when cells carried an SμKI allele. With LPS treatment(Fig. 2D), a ∼50% decrease in secreted IgG3 was observed in theKI, as well as an enhanced level of secreted IgE.DNA rearrangements of spleen cell-derived IgH loci in

hybridomas reflect CSR events in normal splenic B cells ata single-cell level. To quantitate CSR, hybridomas were gener-ated from splenocytes stimulated for 2 and 4 d with LPS/IL-4(Table 1). On day 4, IgG1+ clones were reduced ∼10-fold in theKI relative to WT, and IgE+ clones increased approximatelysevenfold. Heterozygous hybridomas showed an intermediatephenotype, with threefold fewer IgG1+ clones and fivefold moreIgE+ clones relative to WT. The increased number of IgE+

clones in the KI does not appear to result from sequentialswitching, as more IgE+ B-cells are already evident in SμKIcompared with WT only 2 d after stimulation. PCR and se-quencing analysis confirmed that IgE+ hybridomas containedSμ–Se (WT) or Sμ–Sμ (SμKI) junctions (Figs. S2 and S3). Theseresults, together with the aforementioned FACS and ELISAdata, are consistent with AID targeting to e being enhancedsubstantially by substitution of Sμ for Se.

SμKI Mice Produce Copious IgE in Response to a Variety of Systemicand Local Challenges. We determined whether the IgE producedby SμKI mice is Ag-specific or merely reflects a nonspecific surgein CSR to the modified e locus by applying several systemic and

local challenges. Mice were immunized with T-cell–dependentantigen 2,4,6-trinitrophenyl ovalbumin (TNP-OVA) (31) and IgEtiter was measured. No TNP-specific IgE was present before im-munization (Fig. 3A), although slightly enhanced total serum IgEwas seen in SμKI (Fig. S4A). On days 14 and 28 postimmunization,TNP-specific IgE titers in SμKI mice were approximately fivefoldhigher than in WT mice. SμKI mice were also locally challengedwith various model allergens to see if they would produce serumIgE in excess. SμKI mice immunized with TNP-OVA were chal-lenged intranasally with aerosolized TNP-OVA to model asthma(Fig. 3B). Challenged SμKI mice produced 8- to 10-fold higherTNP-specific serum IgE than their WT counterparts, whereasbaseline TNP-specific IgE was undetectable. Remarkably, SμKImice accumulated significant levels of Ag-specific IgE (∼100,000ng/mL) within 4 d after challenge. Induction of IgE in SμKI micewas similarly robust following intranasal delivery of an Ag mix-ture in the absence of systemic sensitization (Fig. 3C and Fig.S4B), or following s.c. infection with Nippostrongylus brasiliensislarva (Fig. 3D). By contrast, IgE titers in WT mice postinfectionbarely reached the basal IgE titers seen in the SμKI mice. In sum,our in vivo data demonstrate that the SμKI allele supports robustIgE isotype switching in response to a variety of challenges.

A

B

Fig. 1. Targeting strategy and generation of SμKI mice. B, BamHI; E, EcoRI;H, HindIII. (A) Schematic representation of the IgH constant region in mice.(B) Schematic representation of Sμ, Se, and SμKI within the IgH locus. Sμ, Cμ,Ie, and Ce probes are depicted. Estimated sizes of Sμ vs. SμKI after HindIIIdigestion and SμKI vs. Se after BamHI digestion are indicated.

A

B

C

D

Fig. 2. In vitro stimulated SμKI B cells produce higher levels of IgE and lowerlevels of IgG1 antibodies compared with WT B cells. (A) Splenocytes werestimulated 4 d with LPS/IL-4 and the indicated antibody isotypes measured byFACS. B220 was used as a B-cell marker, and AID−/−mice were used as negativecontrol. (B) Splenocytes were stimulated 4 d with LPS, and the indicated an-tibody isotypes measured by FACS. B220 was used as a B-cell marker. (C)Splenocytes from WT, heterozygous, and SμKI mice (n = 5) were stimulatedwith LPS/IL-4 for 6 d, and the indicated antibody isotypes measured by ELISA.(D) Splenocytes fromWT, heterozygous, and SμKI mice (n = 5) were stimulatedwith LPS for 6 d, and indicated antibody isotypes measured by ELISA.

Misaghi et al. PNAS | September 24, 2013 | vol. 110 | no. 39 | 15771

IMMUNOLO

GY

Dow

nloa

ded

by g

uest

on

Apr

il 30

, 202

1

Page 3: Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

Replacement of Se with Sμ Enhances Epsilon GLT. To explore themechanistic basis for enhanced IgE production in SμKI mice, weanalyzed GLT levels under various stimuli. Germ-line transcriptsare non-coding RNAs that begin at a short exon upstream of each Sregion termed the I exon, and proceed through the constant regions.PurifiedB cells were activated 2 dwithLPSorLPSplus IL-4 (Fig. 4AandBandFig. S5A),LPSplus IFN-γ, or IL-4plusαCD-40andTGF-β(Fig. S5 B and C). The LPS-treated samples showed comparablelevels of GLT induction for μ, γ2b, and γ3 betweenWT and SμKI Bcells. As expected, WT cells did not produce e transcripts before orafter LPS treatment. In contrast, SμKI cells showed enhancedbasal e transcription without stimulation, and with LPS treatmentthis was slightly enhanced (Fig. 4A). Likewise, in independentstimulations that used LPS plus IL-4 (Fig. 4B), eGLT levels wereagain high in resting SμKI cells andwere induced by approximatelytwofold in response to activation. GLT levels for μ and γ1appeared unaltered by the Sμ insertion. Northern blot analysisconsistently showed enhanced eGLT (Fig. S5A) (36). GLT levelsfor μ, γ2a, and α appeared comparable between WT and SμKI(Fig. S5 B and C). The enhanced e GLT seen in the knock-insuggests that, in addition to providing a longer andmoremotif-richregion for CSR, the insertion alters e locus accessibility.

CSR to the SμKI Locus Is Largely Direct Instead of Sequential. Theenhanced CSR to the SμKI locus raises the question of whetherswitching occurs sequentially via Sγ1 or is direct. Sequentialswitching is a two-part mechanism in which CSR first occurs be-tween Sμ and Sγ1, which subsequently switches to Se, or betweenSγ1 and Se, which then switches to Sμ (27, 28). To address thisquestion, we looked for the presence of circle transcripts in LPS/IL-4 activated B cells over time (Fig. 5). Circle transcripts areproduced when the intervening genomic DNA between two con-necting S regions is looped out, forming an episome in which the Ipromoter of the downstream S region ends up driving transcrip-tion of the constant (C) region adjacent to the upstream S region.We analyzed RNA from resting and activated WT and SμKI B

cells by using RT-PCR and primer sets capable of detecting di-rect and sequential switch products (Fig. 5A). The product ofprimers IeFW and CμRV (Fig. 5B, lanes 1–5) displays switchingto Se, which could be direct or sequential via an Sμ/Sγ1 or Sγ1/Seintermediate. Visualized with a probe to Cμ, this product appearsafter 48 h and appears significantly more robust in SμKI cells,consistent with their enhanced IgE production. The combinationof primers IeFW and Cγ1RV identifies CSR between Sγ1 and Se,which is a unique sequential switch product (Fig. 5B, lanes 6–10).Highlighted by a Cγ1 probe, this product appears after 72 h inWT cells, becoming more abundant by 96 h. Its levels are sig-nificantly lower in SμKI cells, suggesting the cells favor a moredirect mechanism of CSR over a sequential pathway. Equivalent

RNA inputs were used in this experiment, as shown by β-actinRT-PCR (Fig. 5C).To determine what fraction of B cells undergo sequential

switching, we generated a panel of IgM+ and IgE+ hybridomasfrom WT and SμKI B cells stimulated with LPS/IL-4 (Fig. S6).Genomic DNA from the IgM+ hybridomas was digested withEcoRI to release a fragment containing Se/SμKI (Fig. S6 A andB; see also Fig. 1B) and hybridized to the Ce probe. In the eventof recombination between Sγ1 and Se/SμKI, the size of the EcoRIfragment would be altered (28). We did not detect changes inband size in hybridomas from 2- or 4-d stimulated B cells, sug-gesting the frequency of sequential switching is low in both gen-otypes. We also used a nested PCR strategy to amplify Sμ–Sejunctions from IgE+ hybridomas and probed for the presence ofSγ1 in Southern blots (Fig. S6 C and D). We detected seven Sμ-Sejunctions containing Sγ1 among 57 WT hybridomas (12.2%), vs. 1of 58 hybridomas containing Sγ1 for SμKI (1.7%), suggesting se-quential switching is less favored in SμKI B cells.

Endogenous Sμ ISR Is Reduced in the Presence of Knocked-In Sμ. Toview the impact of the SμKI on ISR, IgM+ hybridomas wereanalyzed for deletions within endogenous Sμ by Southern blot(Fig. 6 A and B). ISR was detected in 60% of WT hybridomas,which decreased to 38% and 25% of heterozygous and homozy-gous SμKI hybridomas, respectively. The observation that SμKInegatively impacts ISR at upstream endogenous Sμ implies thatthe insertion competes away factors necessary for ISR or thatSμKI channels endogenous Sμ region breaks into a productiveCSR reaction with higher frequency. It should be noted thatSouthern blotting is a low-resolution assay, and so may reflectonly a fraction of total events (32).We next examined whether SμKI could undergo ISR in the

context of the e locus. The Ie probe was used to analyze ISRwithin Se and inserted Sμ in hybridomas created from IgM+ Bcells (Fig. 6C). Among 96 hybridomas, Se did not display anyinternal recombination events and SμKI displayed only one event,

Table 1. Quantification of isotype switching by B-cellhybridomas

Day/hybridoma No. of clones IgM+, % IgG1+, % IgE+, %

Day 2WT 562 95 3 2KI 594 88 1 11

Day 4WT (fusion 1) 1,440 32 57 11WT (fusion 2) 946 21 68 11HET 1,379 20 21 59KI (fusion 1) 1,444 14 6 80KI (fusion 2) 1,141 18 4 78

At 2 or 4 d after LPS/IL-4 stimulation, splenocytes from WT and SμKI (i.e.,heterozygous or homozygous) mice were fused to the NS-1 host to createhybridomas. ELISAs were used to obtain percent IgM+, IgG1+, and IgE+

hybridomas.

A B

C D

Fig. 3. SμKI mice express higher levels of IgE compared with WT mice in vivo.(A) WT (n = 3), heterozygous (Het; n = 3), and SμKI (n = 7) mice were immu-nized with TNP-OVA. TNP-OVA–specific serum IgE levels were measured byELISA before and at 14 and 28 d postimmunization. (B) Mice (n = 5 per group)were immunized with TNP-OVA/alum and challenged with aerosolized 1%TNP-OVA. Baseline, prechallenge and postchallenge TNP-OVA–specific serumIgE levels were measured by ELISA. (C) WT, Het, or SμKI mice (n = 6 per group)were sensitized to multiple allergens as explained in Materials and Methods,and total serum IgE levels were measured by ELISA before and on days 21 and42 postsensitization. (D) Total serum IgE levels in WT, Het, and SμKI mice (n = 5per group) were measured by ELISA before and 4, 9, and 14 d postinfectionwith 500 N. brasiliensis larvae. P values were calculated between WT and SμKIsamples by using a t test assuming unequal variances.

15772 | www.pnas.org/cgi/doi/10.1073/pnas.1221661110 Misaghi et al.

Dow

nloa

ded

by g

uest

on

Apr

il 30

, 202

1

Page 4: Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

suggesting the primary sequence of Sμ is not efficient at ISR. In-stead, context-dependent cues such as the presence of regulatoryelements (e.g., enhancer Eμ or Iμ), rate of transcription, or localchromatin modifications may have influence.The differences in mechanism by which ISR, CSR, and SHM

occur are not well understood and the processes can be tem-porally distinct (37). We hypothesized that, if SμKI impacts otherS regions in the context of CSR, it might also affect SHM inSμKI B cells. We sorted IgM+ or IgG1+ germinal center B cellsin response to TNP-OVA to scan for mutations downstream ofJH4 and upstream of endogenous Sμ (Fig. S7A) (31). Our analysisrevealed that the rate of SHM in SμKI B cells was not significantlydifferent from WT. Rather, there was a slight trend toward lowermutation frequency in the SμKI allele. To document the impacton SHM directly in V regions, a panel of IgG1-producing TNP-OVA specific hybridomas was generated after TNP-OVA immu-nization, and heavy chain variable region (VH) gene sequenceswere aligned with germ-line counterparts to locate mutations (Fig.S7B). We did not detect a major difference in mutation frequencybetween WT and SμKI. Thus, alteration of the IgH locus in SμKIimpacts CSR and ISR, but does not appear to impact SHM.

DiscussionSμ is a constitutively transcribed donor S region that is unusuallydense with AID target hotspots (4). S region size linearly cor-relates with enhanced CSR (25). In our model, we replaced the2-kb Se with a larger ∼5-kb core Sμ. The GLT of the SμKI locusis enhanced relative to unmodified Se. Thus, increased CSR seenin the SμKI could be a result of the sequence of Sμ or the in-creased S region length combined with greater transcriptionalaccessibility. The result is a dramatic increase in the magnitudeof IgE isotype switching, even exceeding the IgG1 levels associ-ated with the largest S region, Sγ1.It is not understood to what degree the context of Sμ influ-

ences its ability to undergo CSR and ISR. We provide datasupporting the notion that Sμ inserted in place of Se is in-sufficient to carry out certain functions it would have in its en-dogenous location. The knocked-in sequence is able to reduceISR of endogenous Sμ from a distance, either by competing for

limited factors or through providing DSBs leading to productiveCSR with its upstream twin. However, despite the known re-combinogenicity of Sμ, the KI is insufficient to generate levelsof ISR comparable to its upstream counterpart, suggesting thesequence itself is insufficient or the e locus is not conducive toISR events (28, 31). This is also unusual given that ISR is foundfor endogenous Se (28, 31) as well as other acceptor S regions(28, 32, 33). The robust targeting of native Sμ might facilitateproductive CSR by making targeting of downstream S regionsrate-limiting, which ultimately may limit the duration and pro-duction of unnecessary DSBs (32). Although AID is capable ofinitiating DSBs in all S regions, AID-mediated ISR may be influ-enced by regional DNA architecture and recruited DNA repairmachinery. Our finding suggests that the frequent targeting ofendogenous Sμ is context-dependent.The mechanism of direct vs. sequential CSR to produce IgE is

not well understood, and this process might be developmentallyregulated (38). Our analysis of circle transcripts indicates thatdirect and sequential switching can mediate CSR. Sequentialswitching might involve Sμ/Sγ1 junctions joining to Se orswitching between Sγ1/Se before joining Sμ (27, 39). In our assay,we observed reduced levels of Sγ1/Se products in SμKI relative toWT. These data indicate that, in SμKI cells, switching is moredirect. In addition, our ISR data suggests that a prominent, ac-tive downstream S substrate prefers direct switching vs. indirectwith donor Sμ, in which there is a higher chance to generate twosimultaneous DSBs (7). This interpretation agrees with the factthe sequential switching is mainly observed for Se and not forother S regions (27, 39).It is intriguing that steady-state GLT at the e locus is positively

affected by the Sμ insertion. This observation could explain thelow but detectable levels of IgE observed in LPS-stimulatedSμKI B cells. It could be that the e locus is an inherently difficult

A

B

Fig. 4. GLT levels in WT vs. Sμ-KI mice. At least three independent miceper genotype were tested with similar results. (A) Semiquantitative RT-PCR with RNA from WT or Sμ-KI B cells that were freshly isolated orstimulated with LPS for 2 d. Starting amounts of RNA were 2 ng, 10 ng,and 50 ng. A fixed cycle number in the linear range for the specificproduct was used. Gel images are shown for μ (25 cycles) and β-actin (25cycles). Southern blot autoradiographs are shown for γ2b (30 cycles), γ3(30 cycles), and e (45 cycles for WT, 30 cycles for SμKI). (B) Similar to A butwith LPS/IL-4 stimulation. Gel images are shown for μ, γ1 (30 cycles), andβ-actin. A Southern blot autoradiograph is shown for e.

A

B

C

Fig. 5. Contribution of direct vs. sequential switching to IgE. (A) Diagram ofthe IgH locus and relevant features and primers indicated. (B) RNA wasisolated from WT or SμKI B cells at the indicated time points after LPS/IL-4stimulation. RT-PCR was performed by using primers specific to Ie and Cμ(result of circles formed from Sμ–Se junctions, lanes 1–5), and primers specificto Ie and Cγ1 (result of circles formed from Sγ1:Se junctions, lanes 6–10). Theresulting products were hybridized to probes corresponding to Cμ and Cγ1.(C) β-Actin at 20 cycles is shown as a loading control.

Misaghi et al. PNAS | September 24, 2013 | vol. 110 | no. 39 | 15773

IMMUNOLO

GY

Dow

nloa

ded

by g

uest

on

Apr

il 30

, 202

1

Page 5: Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

substrate to transcribe and easily influenced by perturbationsand/or the Sμ sequence somehow renders the locus transcrip-tionally open. In support of the latter, limited studies of marginalzone B-cell lymphomas suggest human Sμ sequence can act asa promoter to drive the expression of the Pax-5 gene in recur-ring (9:14) human chromosomal translocations (40). Sequenceamplifications of Sμ and surrounding regulatory regions havebeen reported in mouse and human B-cell tumors that involveCSR (41). On the contrary, removal of core Sμ has not beenshown to alter the GLT of endogenous Cμ (1), and substitutingSγ1 with Xenopus Sμ does not affect Iγ1 promoter activity in thecontext of the endogenous mouse IgH locus (3). Insertion of theheterologous promoter for Phosphoglycerate kinase 1 (Pgk) linkedto neomycin within the IgH locus has been implied to affect GLTby isolating germ-line I promoters from the 3′ regulatory region,or by competing for access to this region (42, 43). In the context of

the epsilon locus, replacement of Ce exons with a Pgk-neomycincassette in an antisense orientation can inhibit the GLT of Cγ3,Cγ2b, and Cγ2a genes (36). According to our study, however, theenhanced activity of the Ie promoter at the SμKI allele does notappear to appreciably affect the GLT of other heavy chain con-stant (CH) genes. In contrast to the Pgk-neomycin insertion model,the sense orientation of the endogenous I epsilon promoter in ourSμKI model is conserved and does not include additional sequen-ces such as a neomycin cassette. Our finding demands additionalstudies to elucidate the mechanism behind this observation.The SμKI hyper-IgE mouse model is unique for the study of IgE

biology and its associated pathology. Allergen-bound IgE activatesmast cells via the Fce receptor and releases a spectrum of in-flammatory mediators (44). Poor CSR potential and short IgE t1/2might have been selected during evolution to limit IgE abundance (4)and the danger of unwanted effector functions associated with ana-phylaxis and allergies (44). Not only are acute asthmatic patientssuccessfully treated with IgE neutralizing antibodies (45), new indi-cations for this therapeutic intervention include allergic rhinitis, foodallergy, atopic dermatitis, chronic urticaria, and allergic broncho-pulmonary aspergillosis (45). IgE transgenic mice have the majorshortcoming of being unable to mount a polyclonal Ag-specific IgEresponse (46–48). By altering de novo CSR, the resultant SμKI miceare able tomount diverse andpolyclonalAg-specific IgE responses tolocal or systemic challenges, thereby providing a unique experimentalmodel for the accelerated assessment of future therapies in a facilesystem that more closely mimics the human condition.

Materials and MethodsCSR/ISR/SHM Assays. Splenocytes were isolated and stimulated as previouslyexplained (31). ELISAs and surface staining of B cells for IgM, IgG1, IgG3, andB220 were performed as explained (31). To visualize IgE, cells were blockedwith anti-mouse IgE blocking antibody (eBioscience) and then stained withbiotin-isotype or anti-mouse IgE-biotin (eBioscience) followed by streptavidin–phycoerythrin (Pharmingen). Southern blot methods were described previously(31). Southern probes included Cμ (XbaI-BamHI, 0.8 kb), Ie (BamHI-NcoI, 1 kb),Sμ (HindIII, 5 kb), and Ce (PstI-PvuI, 1 kb; Fig. 1B). For GLT and circle transcripts,RT-PCR reactions were carried out by using the Access RT-PCR kit (Promega).Cycling parameters were 45 °C for 45 min and 94 °C for 2 min, then repeatingcycles of 94 °C for 30 s, 60 °C for 30 s, and 68 °C for 30 s. GLT amplificationprimers were reported elsewhere (24). Circle transcripts were detected withGLT primers except for CμRV 5′-AATGGTGCTGGGCAGGAAGT-3′.

Hybridomas. The procedure for generating hybridomas have been describedelsewhere (49). Briefly, splenocytes from 6- to 8-wk-old mice (WT, heterozygous,andKI) were stimulated in vitrowith LPS (20 ng/mL) and IL-4 (25 ng/mL). After 4 d,20 million B cells were fused with NS1 cells (American Type Culture Collection) ata 1:1 ratio by using the Cyto Pulse CEEF-50 apparatus (Cyto Pulse Sciences). After 7to 10 d selection in hypoxanthine-aminopterin-thymidine (HAT) media (StemCellTechnologies), the single hybridoma clones were screened for IgE-, IgM- or IgG1-specific hybridomas by Velocity 11 Biocel 1200 Automation System (AgilentTechnologies). TNP-OVA specific hybridomaswere generated from splenocytes atday 42 after repeated immunization (as described later) and screened by ELISA.

In Vivo Studies and Animal Care. TNP-OVA immunization (T-cell–dependentimmune response) was performed by i.p. injection of TNP-OVA/Alum adjuvant atday 0, followed by another injection of TNF-OVA at day 28. Blood sampleswere collected on days 3, 7, 14, 21, 28, 35, and 42 for antibody isotype meas-urements. For the asthma model, mice were sensitized with an i.p. injection ofTNP-OVA/Alum on day 0. Mice were challenged with aerosolized 1% TNP-OVAon day 35 after sensitization, followed by seven consecutive days of challeng-ing as indicated earlier. Blood samples for antibody isotype measurementswere collected at day 2 (baseline bleed), day 34 (prechallenge), and day 42(postchallenge). Mice infected with N. brasiliensis larvae s.c. were placed onpolymyxin B (110 mg/L; Calbiochem) and neomycin (1.1 g/L; Sigma-Aldrich)medicated water for 5 d. Serum samples were collected before and 4, 9, or 14d after infection for antibody isotype measurements. For the multiple Ag al-lergen model, mice were sensitized and exposed to multiple allergens in-tranasally twice per week for 6 wk. Each mouse was exposed intranasally toextracts of Dermatophagoides farinae (5 μg), Dermatophagoides pteronyssinus(5 μg), Ambrosia artemisiifolia (50 μg), and Aspergillus fumigatus (5 μg) (GreerLaboratories) mixed in 50 μL PBS solution. One week after the last exposure,total IgE and IgG1 were measured by ELISA. SμKI and WT mice were generated

A

B

C

Fig. 6. ISR frequency in B-cell hybridomas by Southern blot. Restriction frag-ments andprobesaredepicted inFig. 1B. (A) IgM+hybridomasderived fromLPS/IL-4 stimulatedB cellsweredigestedwithEcoRI andprobedwithCμ todetect ISRwithin endogenous Sμ. (B) Table summarizing thenumberof endogenous Sμ ISRevents in IgM+ hybridomas in WT, heterozygous (HET), and SμKI. (C) IgM+

hybridomas derived from LPS/IL-4–stimulated B cells were digested with BamHIand probed with Ie. Note that an ∼3-kb increase in size is observed when Se isreplacedwith Sμ (i.e., SμKI). ISR of Se or SμKIwas absent in all 48WT hybridomasscreened, andonly one ISR eventwas detected in the SμKI (asterisk). The Se bandin WT samples may originate from both the B-cell and NS-1 fusion partner,whereas, in KI samples, it stems only from the fusion partner.

15774 | www.pnas.org/cgi/doi/10.1073/pnas.1221661110 Misaghi et al.

Dow

nloa

ded

by g

uest

on

Apr

il 30

, 202

1

Page 6: Polyclonal hyper-IgE mouse model reveals mechanistic ...Polyclonal hyper-IgE mouse model reveals mechanistic insights into antibody class switch recombination Shahram Misaghi1, Kate

at Genentech and maintained in accordance with American Association ofLaboratory Animal Care guidelines. The experiments were conducted in com-pliance with National Institute of Health Guide for the Care and Use of Lab-oratory Animals and were approved by the Genentech Institutional AnimalCare and Use Committee.

ACKNOWLEDGMENTS. The authors thank Eric Pinaud and Ming Tian forproviding technical advice and insightful discussions; Mariela del Rio, BenGrellman, and Vida Asghari for managing the mouse colony; and KimNewton, Flavius Martin, Harinder Singh, and Menno Van LookerenCampagne for carefully reviewing the manuscript.

1. Khamlichi AA, et al. (2004) Immunoglobulin class-switch recombination in mice de-void of any S mu tandem repeat. Blood 103(10):3828–3836.

2. Luby TM, Schrader CE, Stavnezer J, Selsing E (2001) The mu switch region tandemrepeats are important, but not required, for antibody class switch recombination.J Exp Med 193(2):159–168.

3. Zarrin AA, et al. (2004) An evolutionarily conserved target motif for immunoglobulinclass-switch recombination. Nat Immunol 5(12):1275–1281.

4. Hackney JA, et al. (2009) DNA targets of AID evolutionary link between antibodysomatic hypermutation and class switch recombination. Adv Immunol 101:163–189.

5. Yancopoulos GD, et al. (1986) Secondary genomic rearrangement events in pre-Bcells: VHDJH replacement by a LINE-1 sequence and directed class switching. EMBO J5(12):3259–3266.

6. Stavnezer-Nordgren J, Sirlin S (1986) Specificity of immunoglobulin heavy chainswitch correlates with activity of germline heavy chain genes prior to switching.EMBO J 5(1):95–102.

7. Zarrin AA, et al. (2007) Antibody class switching mediated by yeast endonuclease-generated DNA breaks. Science 315(5810):377–381.

8. Yan CT, et al. (2007) IgH class switching and translocations use a robust non-classicalend-joining pathway. Nature 449(7161):478–482.

9. Schrader CE, Guikema JE, Linehan EK, Selsing E, Stavnezer J (2007) Activation-inducedcytidine deaminase-dependent DNA breaks in class switch recombination occur dur-ing G1 phase of the cell cycle and depend upon mismatch repair. J Immunol 179(9):6064–6071.

10. Robert I, Dantzer F, Reina-San-Martin B (2009) Parp1 facilitates alternative NHEJ,whereas Parp2 suppresses IgH/c-myc translocations during immunoglobulin classswitch recombination. J Exp Med 206(5):1047–1056.

11. Hasham MG, et al. (2012) Activation-induced cytidine deaminase-initiated off-targetDNA breaks are detected and resolved during S phase. J Immunol 189(5):2374–2382.

12. Petersen-Mahrt SK, Harris RS, Neuberger MS (2002) AID mutates E. coli suggestinga DNA deamination mechanism for antibody diversification. Nature 418(6893):99–103.

13. Chaudhuri J, et al. (2003) Transcription-targeted DNA deamination by the AID anti-body diversification enzyme. Nature 422(6933):726–730.

14. Bransteitter R, Pham P, Scharff MD, Goodman MF (2003) Activation-induced cytidinedeaminase deaminates deoxycytidine on single-stranded DNA but requires the actionof RNase. Proc Natl Acad Sci USA 100(7):4102–4107.

15. McBride KM, et al. (2006) Regulation of hypermutation by activation-induced cytidinedeaminase phosphorylation. Proc Natl Acad Sci USA 103(23):8798–8803.

16. Dorsett Y, et al. (2007) A role for AID in chromosome translocations between c-mycand the IgH variable region. J Exp Med 204(9):2225–2232.

17. Takizawa M, et al. (2008) AID expression levels determine the extent of cMyc onco-genic translocations and the incidence of B cell tumor development. J Exp Med 205(9):1949–1957.

18. Sernández IV, de Yébenes VG, Dorsett Y, Ramiro AR (2008) Haploinsufficiency ofactivation-induced deaminase for antibody diversification and chromosome trans-locations both in vitro and in vivo. PLoS ONE 3(12):e3927.

19. Daniel JA, et al. (2010) PTIP promotes chromatin changes critical for immunoglobulinclass switch recombination. Science 329(5994):917–923.

20. Vuong BQ, Chaudhuri J (2012) Combinatorial mechanisms regulating AID-dependentDNA deamination: Interacting proteins and post-translational modifications. SeminImmunol 24(4):264–272.

21. Zhu C, et al. (2012) Origin of immunoglobulin isotype switching. Curr Biol 22(10):872–880.

22. Yu K, Chedin F, Hsieh CL, Wilson TE, Lieber MR (2003) R-loops at immunoglobulin classswitch regions in the chromosomes of stimulated B cells. Nat Immunol 4(5):442–451.

23. Dempsey LA, Sun H, Hanakahi LA, Maizels N (1999) G4 DNA binding by LR1 and itssubunits, nucleolin and hnRNP D, A role for G-G pairing in immunoglobulin switchrecombination. J Biol Chem 274(2):1066–1071.

24. Muramatsu M, et al. (2000) Class switch recombination and hypermutation requireactivation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell102(5):553–563.

25. Zarrin AA, Tian M, Wang J, Borjeson T, Alt FW (2005) Influence of switch regionlength on immunoglobulin class switch recombination. Proc Natl Acad Sci USA 102(7):2466–2470.

26. Yoshida K, et al. (1990) Immunoglobulin switch circular DNA in the mouse infectedwith Nippostrongylus brasiliensis: Evidence for successive class switching from mu toepsilon via gamma 1. Proc Natl Acad Sci USA 87(20):7829–7833.

27. Mandler R, Finkelman FD, Levine AD, Snapper CM (1993) IL-4 induction of IgE classswitching by lipopolysaccharide-activated murine B cells occurs predominantlythrough sequential switching. J Immunol 150(2):407–418.

28. Zhang T, et al. (2010) Downstream class switching leads to IgE antibody production byB lymphocytes lacking IgM switch regions. Proc Natl Acad Sci USA 107(7):3040–3045.

29. Xiong H, Dolpady J, Wabl M, Curotto de Lafaille MA, Lafaille JJ (2012) Sequential classswitching is required for the generation of high affinity IgE antibodies. J Exp Med209(2):353–364.

30. Jung S, Siebenkotten G, Radbruch A (1994) Frequency of immunoglobulin E classswitching is autonomously determined and independent of prior switching to otherclasses. J Exp Med 179(6):2023–2026.

31. Misaghi S, et al. (2010) Increased targeting of donor switch region and IgE inSgamma1-deficient B cells. J Immunol 185(1):166–173.

32. Dudley DD, et al. (2002) Internal IgH class switch region deletions are position-independent and enhanced by AID expression. Proc Natl Acad Sci USA 99(15):9984–9989.

33. Reina-San-Martin B, Chen HT, Nussenzweig A, Nussenzweig MC (2004) ATM is re-quired for efficient recombination between immunoglobulin switch regions. J ExpMed 200(9):1103–1110.

34. Gu H, Zou YR, Rajewsky K (1993) Independent control of immunoglobulin switchrecombination at individual switch regions evidenced through Cre-loxP-mediatedgene targeting. Cell 73(6):1155–1164.

35. Sun Y, et al. (2013) Critical role of activation induced cytidine deaminase in experi-mental autoimmune encephalomyelitis. Autoimmunity 46(2):157–167.

36. Seidl KJ, et al. (1999) Position-dependent inhibition of class-switch recombination byPGK-neor cassettes inserted into the immunoglobulin heavy chain constant regionlocus. Proc Natl Acad Sci USA 96(6):3000–3005.

37. Di Noia JM, Neuberger MS (2007) Molecular mechanisms of antibody somatic hy-permutation. Annu Rev Biochem 76:1–22.

38. Wesemann DR, et al. (2011) Immature B cells preferentially switch to IgE with in-creased direct Sμ to Se recombination. J Exp Med 208(13):2733–2746.

39. Mills FC, Mitchell MP, Harindranath N, Max EE (1995) Human Ig S gamma regions andtheir participation in sequential switching to IgE. J Immunol 155(6):3021–3036.

40. Morrison AM, et al. (1998) Deregulated PAX-5 transcription from a translocated IgHpromoter in marginal zone lymphoma. Blood 92(10):3865–3878.

41. Janz S (2006) Myc translocations in B cell and plasma cell neoplasms. DNA Repair(Amst) 5(9-10):1213–1224.

42. Manis JP, Tian M, Alt FW (2002) Mechanism and control of class-switch recombination.Trends Immunol 23(1):31–39.

43. Pinaud E, et al. (2011) The IgH locus 3′ regulatory region: Pulling the strings frombehind. Adv Immunol 110:27–70.

44. Galli SJ, Tsai M, Piliponsky AM (2008) The development of allergic inflammation.Nature 454(7203):445–454.

45. Ben-Shoshan M (2010) Omalizumab for asthma: indications, off-label uses and futuredirections. Recent Pat Inflamm Allergy Drug Discov 4(3):183–192.

46. Adamczewski M, Köhler G, Lamers MC (1991) Expression and biological effects of highlevels of serum IgE in epsilon heavy chain transgenic mice. Eur J Immunol 21(3):617–626.

47. Curotto de Lafaille MA, et al. (2001) Hyper immunoglobulin E response in mice withmonoclonal populations of B and T lymphocytes. J Exp Med 194(9):1349–1359.

48. Matsuoka K, et al. (1999) Establishment of antigen-specific IgE transgenic mice to studypathological and immunobiological roles of IgE in vivo. Int Immunol 11(6):987–994.

49. Sun Y, et al. (2012) Evolutionarily conserved paired immunoglobulin-like receptor α(PILRα) domain mediates its interaction with diverse sialylated ligands. J Biol Chem287(19):15837–15850.

Misaghi et al. PNAS | September 24, 2013 | vol. 110 | no. 39 | 15775

IMMUNOLO

GY

Dow

nloa

ded

by g

uest

on

Apr

il 30

, 202

1