Skip to main content
Nature Portfolio logoLink to Nature Portfolio
. 2014 Feb 11;92(5):392–399. doi: 10.1038/icb.2014.8

Dissecting mammalian immunity through mutation

Owen M Siggs 1,*
PMCID: PMC4038135  PMID: 24518983

Abstract

Although mutation and natural selection have given rise to our immune system, a well-placed mutation can also cripple it, and within an expanding population we are recognizing more and more cases of single-gene mutations that compromise immunity. These mutations are an ideal tool for understanding human immunology, and there are more ways than ever to measure their physiological effects. There are also more ways to create mutations in the laboratory, and to use these resources to systematically define the function of every gene in our genome. This review focuses on the discovery and creation of mutations in the context of mammalian immunity, with an emphasis on the use of genome-wide chemical and CRISPR/Cas9 mutagenesis to reveal gene function.

Keywords: genetics, immunity, mutation

INTRODUCTION

The origins of immunity

Since the emergence of unicellular life, there have been few selective pressures more abundant than infection. The evolutionary legacy of infection is a diverse antimicrobial arsenal, known collectively as immunity, that can be found in all species from the simplest prokaryotes to the most complex mammals. Bacteria and archaea, for example, evolved restriction enzymes and CRISPR/Cas (clustered, regularly interspaced short palindromic repeat/CRISPR-associated) systems to combat both phage infection and plasmid conjugation.1 Similarly, many antibiotics used in the clinic today are products of prokaryotic competition.

Antibiotics also evolved in eukaryotic organisms, as did a system of sequence-specific RNA silencing. Yet, with the many benefits of eukaryotic multicellularity came vulnerability to new classes of microbes, and therefore a need for new forms of immunity. Receptors for microbial products emerged, exemplified by the Toll protein in Drosophila, that in some species has expanded to families of hundreds of paralogs.2 Additional microbial molecules were recognized by other receptor families, including the NOD-like, RIG-I-like and C-type lectin receptors.3

The crowning achievement of eukaryotic immunity arose in vertebrates not once but twice,4 with an anticipatory system of somatically recombined and hypermutated antigen receptors. Not only did this system have the capacity to detect any pathogen that already existed, it had the bandwidth to preempt any pathogen that would exist, ever. Although combinatorial diversity created the hazard of self-reactivity, the evolutionary advantage afforded by it was simply too great to be ignored, and mechanisms of self-tolerance emerged as a result. Combinatorial immunity has been refined even further since, with unique structural adaptations to reach even more microbial epitopes.5, 6

Human immunity and the shadows of selection

Just as microbial infection acts as a selective pressure against the host, host immunity exerts powerful selective pressure against the microbe. As quickly as new forms of immunity have emerged, rapid microbial proliferation ensures that they quickly develop ways to avoid them. This perpetual struggle between pathogen and host is reflected by our recent evolutionary history that reveals that immune genes continue to be the most strongly selected elements in our genome.7, 8

More recent signatures of selection can be found in human populations with endemic infectious diseases. One classic example is the high prevalence of null mutations in the Duffy antigen gene (DARC) in West Africa that confer resistance to the malarial parasite Plasmodium vivax. Other positively selected variants have a less obvious origin: the CCR5Δ32 allele bestows resistance to the modern human pathogen HIV, yet was most likely selected by a more ancient microbe. Certain variants provide such a crucial advantage that they eventually reach fixation in a population because of a selective sweep, and there is evidence that several immune genes fall into this category. A series of parallel selective sweeps is, after all, what separates one species from another, and among other things explains why mice do not become sick after HIV inoculation, or why fruit bats carry Ebola virus without developing hemorrhagic fever.

Other variants with much smaller effects have been uncovered by genome-wide association studies, some of which associate with susceptibility or resistance to infectious disease.9 Similar methods have revealed risk variants for autoimmune and inflammatory diseases,10 the persistence of which may be a testament to their antimicrobial benefit. A key illustration of this is that loss-of-function variants of IFIH1 (encoding the microbial RNA sensor MDA5) are associated with resistance to type I diabetes, whereas the more common, functional alleles confer susceptibility.11

Although improved hygiene, vaccination and antibiotic use have dramatically reduced the burden of infectious disease over the past 200 years, it remains a powerful selective agent with ∼25% of people ultimately dying from it,12 many of whom are young. New pathogens continue to cross from animals into humans, and the pathogens that were once subdued by antimicrobial drugs are rapidly developing resistance. At the same time, the human population is undergoing explosive growth, with new single-nucleotide variants emerging at a rate of ∼1.2 × 10−8 per generation.13 This is estimated to introduce some 1011 de novo variants per generation,14 but how do we determine which of these variants affect immunity?

Contemporary experiments of nature

By recent conservative estimates, each human genome carries ∼300 variants that affect protein function.15 More than 86% of these are thought to have arisen within the past 10 000 years and therefore have low population frequencies (<5%),16 yet because of rapid population growth, most have remained in a heterozygous state and hence have not been subjected to purifying selection.

Nevertheless, in some cases these variants can still cause inherited disease. Some may affect haploinsufficient genes, such as variants in RPSA that cause autosomal dominant congenital asplenia.17 Other variants act in a dominant manner, such as PLCG2 mutations in cold-induced urticaria,18 or PIK3CD variants in a subset of primary immunodeficient patients.19, 20 The remainder are either X-linked (such as variants of the T-cell magnesium transporter gene MAGT1),21 or disrupted at both alleles via consanguinity or compound heterozygosity (homozygous IRF8 mutations in mycobacterial susceptibility being an example of the former22). These experiments of nature have taught us a great deal about immunity, not only in humans but also in animals. Some of the biggest conceptual breakthroughs in the past 20 years of immunology have emerged from the study of spontaneous mutations, from the understanding of microbial sensing (Tlr4) and dominant tolerance (Foxp3) in mouse mutants,23, 24 to the discovery of multiple variants causing monogenic autoimmunity, immunodeficiency or sterile inflammatory disease in humans.

Making sense of nature's experiments

Revealing the genetic basis of an inherited immune disorder lays a foundation for hypothesis and further experimentation. Fortunately, many human immune cells are accessible enough to test these hypotheses, and in many cases a patient blood sample is the perfect starting point. Occasionally, the critical cell type may be unavailable or inaccessible, for example, in TLR3-associated viral encephalitis, where the central nervous system is suspected to be central to pathogenesis. An emerging solution is the use of patient-derived induced pluripotent cells that can be differentiated into the appropriate cell lineage and studied in vitro alongside controls.25 An even better alternative is to use isogenic human induced pluripotent cells, where experimental and control cell lines differ by only a single genetic variant. Sequence-specific genome editing tools have been invaluable in this context, with the CRISPR/Cas9 platform emerging as the most versatile.

For many purposes these in vitro platforms will be sufficient, yet for many physiological phenomena they are not. Model organisms, and the mouse in particular, provide the most powerful means available to study mammalian physiology, offering an environmentally and genetically controlled platform for the study of cellular interactions within the context of a whole organism. For the 99% of human genes that have a direct counterpart,7 the mouse has been the dominant model for understanding their function. The tools available for engineering the mouse genome are without equal, with transgenesis, homologous recombination and sequence-specific nucleases used to create a spectrum of strains that have become a cornerstone of modern immunology. The speed and ease with which the CRISPR/Cas9 platform can create germline mutations makes the mouse an even more accessible platform for examining the effects of human genetic variation.26

Laboratory mice are nevertheless an imperfect model of human immune disease,27 as might be expected after 65 million years of genetic divergence. Although invaluable experimental interpretation and consistency, the widespread use of inbred strains of mice (particularly C57BL/6) fails to capture the genetic and phenotypic diversity of outbred human populations. A more suitable experimental model of immunological disease is not immediately obvious, although it may one day involve some combination of genetically humanized mice with transplanted human cells.

SYSTEMATICALLY CONNECTING GENOTYPE TO IMMUNE PHENOTYPE

Of the ∼20 000 protein coding genes and several thousand noncoding RNA genes in our genome, we know remarkably little about the normal function of most of them. At the same time, the catalog of human genetic variation is expanding exponentially, yet our understanding of its biological significance is languishing in its wake. The vast majority of our understanding of mammalian gene function has come from single-gene mutations in mice, with upwards of 7000 genes having at least one annotated phenotype.28 Mendelian human diseases follow closely behind, with approximately half as many individual genes linked to an inherited phenotype.29 Several efforts are now underway to systematically dissect the biological function of every human gene, and the relative strengths of each will be discussed below.

Systematic clinical phenotyping and genome sequencing

The most relevant way to understand the function of a human gene is to study mutations in humans themselves. Given the germline mutation rate and a population of >7 billion, it is reasonable to expect that multiple null alleles of every human gene already exist, some of which may be in a compound heterozygous or homozygous state. Our immune systems are also constantly challenged by infection, by immunization and (in some cases) by immune-modifying drugs. Together with clinicians, we are constantly monitoring the function of our immune systems, and those individuals with the most severe defects are quickly noticed. Many of these severe immune disorders are caused by mutations in a single gene, with more than 200 such genes now defined.30 This ability to connect individual human genes to immune function is immensely valuable, and the advent of massively parallel sequencing has meant that dozens more are discovered each year.

Functional variants not associated with clinical disease will be more difficult to identify. Some individuals may not respond to a given vaccine, for example, yet remain healthy because of herd immunity. Similarly, the highest responders will also be healthy, but may harbor valuable genetic clues for enhancing immunity in those with less capable immune systems. Identifying these individuals will require robust and comprehensive methods of immune phenotyping,31 and many hundreds or thousands of cases and controls to extract reliable genetic associations.

Targeted germline mutagenesis and phenotyping

A definitive way of determining mammalian gene function has been the creation of targeted gene knockouts in mice. Candidate genes are typically selected with an explicit hypothesis in mind: one that is often based upon previous in vitro experiments, homology to known immune genes or by comparative gene expression analysis. The creation and phenotyping of mutant mice is an efficient way to understand the function of families of genes, but usually ignores genes that are not suspected by prior hypothesis, or ignores phenotypes that are beyond a laboratory's sphere of expertise.

Previously created by specialist laboratories on a case-by-case basis, gene-targeted stem cells can now be ordered directly from large public repositories32 (Figure 1a). These archives include alleles made by both random gene trap insertions and by targeted homologous recombination (and, inevitably, by CRISPR/Cas9), and are on track to create null alleles of every gene in the mouse genome. The bottleneck now lies in the conversion of these embryonic stem (ES) cell clones into mutant mice: blastocyst injections and at least two generations of breeding are required to obtain the desired homozygous mutants.

Figure 1.

Figure 1

Contemporary methods for mouse germline mutagenesis. (a) Standard design of a lacZ-tagged conditional allele generated by the European Conditional Mouse Mutatgenesis (EUCOMM) targeting pipeline.32 In its default state, the allele is inactivated by splicing upstream exons to a splice acceptor in the targeting cassette. Treatment with Flp recombinase can delete the selection cassette to create a loxP-flanked conditional allele. βact::neo, β-actin promoter-driven neomycin resistance cassette; βgal, β-galactosidase; En2-SA, mouse En2 splice acceptor; FRT, FRT recombination site; IRES, internal ribosome entry site; loxP, loxP site; pA, SV40 polyadenylation signal. (b) CRISPR/Cas9-mediated generation of targeted mutations. Mammalian codon-optimized Cas9 mRNA and target-specific single-guide RNA(s) are injected into zygotes, leading to the creation of indels as a consequence of double-stranded break repair26 (see text for more detail). Multiple alleles and multiple genes may be targeted simultaneously. DNA oligos or vectors with target homology may also be coinjected, allowing the introduction of point mutations or exogenous DNA by homology-mediated repair.65 (c) Pedigree structure for the generation of ENU-induced homozygous germline mutations. Male C57BL/6J pedigree founders are treated with ENU, and bred with either a wild-type female, or with the female offspring of an ENU-mutagenized male (G0′). After two successive generations of brother–sister mating, recessive mutations can be brought to homozygosity, and both recessive and dominant phenotypes may be identified in third-generation (G3) mice. Examples of visible phenotypes with recessive (prune (Hr), pam gray (Hps3), wavedX (Adam17)) or dominant (Dalmatian (Sox10)) inheritance are shown.

A comparatively rapid alternative to targeting in mouse embryonic stem cells is to create mutations directly in zygotes. Sequence-specific nucleases (including zinc-finger nucleases and transactivator-like effector nucleases) have been used to generate biallelic null mutations within a single generation. Yet, neither zinc-finger nucleases nor transactivator-like effector nucleases are as simple to engineer as the CRISPR/Cas9 system that was recently adapted from Streptococcus pyogenes for genome modification in mammals.33, 34 Coinjection of mRNA for the Cas9 endonuclease and a single-guide RNA (sgRNA, containing 20 bp of complementarity to the target) is sufficient to obtain high frequencies of biallelic null mutations in the target of choice, with few off-target effects and few restrictions on the location of the sgRNA target26 (Figure 1b).

Along with archives of targeted mouse mutants, the establishment of high-throughput phenotyping pipelines has meant that diverse phenotypes can be explored systematically in a single facility, rather than sequentially over many years in several different labs. This has already led to some unexpected discoveries, with new phenotypes detected for 56 of the first 250 genes analyzed.28 Many of these genes had been examined by specialist laboratories in the past, but certain physiological functions had never been suspected and thus were never examined (or reported). Although it requires an enormous amount of coordinated effort and resources, broad phenotyping can nonetheless expose entirely unexpected gene functions, in addition to eliminating others. In the domain of immunity, current assays include complete blood count, lymphocyte flow cytometry and infection with Salmonella typhimurium.28 More specialized immune pipelines have also been launched, including more comprehensive flow cytometric analysis of primary and secondary lymphoid organs, detection of autoreactive antibodies and the challenge of intestinal integrity (A Hayday et al., personal communication). Far from exhaustive, these pipelines are designed to provide a standardized public resource of broad phenotypic data, with all mutant lines and data made publicly available for further analysis. The majority of these alleles can also be converted to a conditional state,32 allowing investigators to study genes in a cell- and tissue-specific manner: an especially useful feature for the ∼30% of genes that are homozygous lethal when mutated.28

Random germline mutagenesis

A key limitation of large-scale phenotyping efforts is the depth of phenomena that can be tested. Immunology is awash with phenomena: specialized subsets of leukocytes are constantly being described, and new pathways for microbial recognition continue to emerge. There are simply too many to monitor in a single pipeline, and therefore many important phenotypes will inevitably be missed. Almost a third of null alleles are also embryonic lethal in homozygous form, yet missense alleles of many of the same genes can be viable (and sometimes even associated with disease35).

Random germline mutagenesis provides a way to circumvent some of these issues. Radiation, chemicals, retroviruses and transposons can all create germline mutations at random in vivo, although the most widely used mouse germline mutagen is the alkylating agent N-ethyl-N-nitrosourea (ENU). At an optimal dose, ENU creates single-nucleotide variants in mouse spermatogonial stem cells every 600 000 bp;36 ∼500 times higher than the de novo missense mutation rate in humans. Optimally treated male founders transmit ∼60 genetic variants that alter the proteins that they encode, and with an appropriate breeding scheme (Figure 1c) ∼6 such mutations can be examined in a homozygous state in every third-generation (G3) mouse, with an additional 39 that are heterozygous.37 A pipeline such as this can be established within a year and requires no technical expertise beyond intraperitoneal injection and mouse husbandry, and comes at a substantially lower per-gene cost than creating the equivalent number of targeted mutations. Unlike a well-designed knockout allele, not all ENU mutations will disrupt protein function, although current estimates suggest that ∼15% of them do.38 Some of these variants create null alleles, others create a partial loss of function (where full loss of function is often lethal) and some can even lead to gains of function. Even where a knockout allele has been studied in depth, the phenotypes resulting from missense alleles can sometimes also have unexpected effects,39 and therefore reveal even more about the proteins and pathways they affect.

Phenotyping assays applied to an ENU screen should have high specificity and sensitivity, as in many cases only a single homozygous mutant (or a few heterozygous mutants) will be phenotyped (as opposed to several in the case of targeted mutagenesis pipelines). The most effective immunological screens to date have focused on cellular phenotypes measured by flow cytometry40 (Table 1), complete blood count41 or Toll-like receptor-induced cytokine production,42, 43 although others measuring systemic phenotypes such as antigen-specific antibody production,44 intestinal integrity45 and susceptibility to infection46, 47 have also been successful. A low false positive frequency is crucial for saving time and resources, and high sensitivity is important to ensure that genuine phenotypic mutations are not missed.

Table 1. Examples of immunological gene function revealed by flow cytometric screening of mouse pedigrees with ENU-induced germline mutations.

Gene Affected lineage Human Mendelian disease variants? Reference
Atp11c B   67, 68
Card11 T, B Yes 69
Cd83 T, DC   70
Coro1a T Yes 71
Fnip1 B   72
Gfi1 Neu Yes 73
Gon4l B   74
Hnrnpll T   75
Ikzf1 T, B   76
Il7 T, B   77, 78
Itgb2 NK Yes 79
Itpkb T   80
Lig4 T, B Yes 81
Lyn B   82
Myb LT-HSC   83
Ncaph2 T   84
Nckap1l T, B   85
Nfkb2 B   86
Ptprc T Yes 87
Rltpr T   88
Sppl2a B, DC   89, 90
Tap1 T Yes 91
Tap2 T Yes 91
Tbx21 NK   92
Themis T   93, 94
Zap70 T Yes 39, 95
Zbtb1 T, B, NK   96
Zbtb7b T, NKT   97

Abbreviations: B, B cell; DC, dendritic cell; ENU, N-ethyl-N-nitrosourea; LT-HSC, long-term hematopoietic stem cell; Neu, neutrophil; NK, natural killer cell; T, T cell.

Once a phenotype is identified, whole-genome or whole-exome sequencing can quickly establish a list of potential causative mutations. Sequencing only a single affected mouse will reveal dozens of variants, and therefore each variant should be genotyped across multiple affected individuals to test for phenotypic concordance. This often points to a probable causative variant, although transgenesis or allelism testing (or some other form of complementation) is necessary to formally assign cause and effect. The vast majority of causative variants are those that change protein coding sense (with very few exceptions48), making exome sequencing a popular platform for ENU mutation discovery. In comparison, an advantage of genome sequencing (when performed across multiple samples) is that it can also define the chromosomal boundaries within which the causative mutation must lie36—a valuable asset when an obvious candidate is not found on the first pass.

Haploid cell mutagenesis

Not all immunological phenotypes require whole organisms to study, and indeed many can be more efficiently studied in individual cells. Although biallelic mutations have been rather difficult to create in diploid mammalian cells, random mutagenesis of cancer cell lines has nonetheless revealed several key components of T-cell receptor, interferon and nuclear factor-κB signaling pathways.49, 50, 51

The development of haploid human cell lines (namely, the KBM-7 chronic myelogenous leukemia line) has allowed the hurdle of diploidy to be bypassed, and the advent of massively parallel sequencing means that mutations can be found with increasing ease. This has led to a resurgence of interest in cellular mutagenesis,52 and millions of null mutations can now be screened simultaneously within pools of cells, for example, by selecting for resistance to viruses or bacterial toxins52, 53, 54 or by selecting for the absence of key cell surface molecules.55, 56

Haploid mouse ES cells are potentially even more versatile,57 given their capacity to differentiate into a wide variety of lineages. Although not yet as stable as their immortalized counterparts, ES cell haploidy can be maintained by regular cell sorting for haploid DNA content. Mutagenizing cells immediately after enrichment ensures that mutations will persist in either a hemizygous (if haploid) or homozygous (if diploid) state, and can therefore be used directly for recessive screens.57, 58 By combining this approach with directed differentiation of ES cells into other lineages, the spectrum of available phenotypes expands even further. Human haploid ES cells could also serve as a valuable in vitro platform, and although they are yet to be reported, recent success with macaque embryos suggest that human haploid ES cell lines may not be far away.59

CRISPR/Cas9 cellular mutagenesis

Perhaps the most versatile mutagenesis platform of all has emerged with the development of mammalian CRISPR/Cas9 mutagenesis. Recently adapted from the bacterium S. pyogenes, the CRISPR/Cas9 system requires coordinated expression of two components: a mammalian codon-optimized Cas9 endonuclease, and a chimeric sgRNA consisting of CRISPR RNA (crRNA) and trans-acting crRNA elements.60 Each crRNA contains a 20-bp guide sequence that directs the sgRNA/Cas9 complex to its complementary DNA target, and repair of the ensuing double-stranded break ultimately results in small deletion (or insertion) mutations at the target site. The sgRNA/Cas9 complex will continue to cut its target until disrupted by an indel, and therefore diploidy and polyploidy present no particular hurdle.

sgRNA/Cas9 target sites can be selected across the genome: the only requirement is that they precede an NGG trinucleotide sequence (also known as the protospacer adjacent motif). Different Cas proteins have different protospacer adjacent motif requirements, although Cas9 is currently the dominant nuclease used in mammalian cells and can potentially target ∼40% of all human exons.34 However, not all crRNAs are equally effective (nor are they all specific), and hence individual genes should ideally be targeted independently with multiple sgRNAs. As the sgRNA guide sequence itself is only 20 bp, genome-wide libraries can be readily synthesized and cloned into transducible vectors. Lentiviral vectors are especially appealing for this purpose, given their ability to transduce a wide array of cell types from many different species (Figure 2a).

Figure 2.

Figure 2

Genome-wide CRISPR/Cas9 mutagenesis. (a) Design of hCas9 expression vector and sgRNA lentiviral vector for genome-wide CRISPR/Cas9 mutagenesis screens (adapted from Koike-Yusa et al.61). CMV, CMV promoter; EF1α, elongation factor-1α promoter; PB, piggyBac repeats; PGK, mouse Pgk1 promoter; Puro-2A-hCas9, puromycin resistance cassette and humanized Cas9 cDNA separated by the T2A self-cleaving peptide; Puro-2A-BFP, puromycin resistance cassette and blue fluorescent protein cDNA separated by the T2A self-cleaving peptide; RU5, 5′ long terminal repeat; sgRNA, single-guide RNA; T, U6 terminator; tracrRNA, trans-activated crRNA; U6, U6 RNA polymerase III promoter; ΔU3RU5, RU5 long terminal repeat lacking U3 region. (b) Target enrichment screens, which enrich for cells with the mutation of interest, can be performed with a wide variety of selective agents. Similar screens have successfully been performed in haploid cell lines using transposon-mediated mutagenesis.52, 55, 56, 57, 66 In the context of genome-wide CRISPR/Cas mutagenesis, an approach validated by Koike-Yusa et al.61 begins with the transduction of hCas9-expressing cells with a lentiviral genome-wide library of sgRNAs. Transduced cells can be selected by sorting for BFP+ cells, in which sgRNA/Cas9-mediated double-stranded breaks lead to the introduction of indel mutations. These transduced cells can then be subjected to an array of selective pressures, including drug- or pathogen-induced cell death, fluorescence or magnetic cell sorting, or the induction of a stimulus-responsive selection marker. sgRNA sequences can then be sequenced in the remaining cells, and used to identify target genes that are critical for the phenomenon of interest. (c) Target depletion screens, also known as ‘dropout' screens, use subtractive methods to identify genes critical for a given phenomenon. Cells are first transduced with a drug-inducible hCas9 construct (along with the TetR transcriptional repressor, in the case of a tetracycline-inducible system), then stable lines are transduced with a genome-wide library of sgRNAs. The pool of BFP+ cells is then split into two, with hCas9 expression induced in one but not the other. hCas9 expression leads to the creation of sgRNA-directed indels, a subset of which will be detrimental to the survival of the cell (or its proliferation or differentiation). Following selection for the phenotype of interest (for example, survival), sgRNA sequences are amplified from the remaining induced and noninduced populations and sequenced. Those sgRNAs that are present in the noninduced population, but absent from the induced population, must therefore be suspected to target a critical gene. These screens are conceptually similar to pooled short hairpin RNA (shRNA) screens, in which barcoded inducible shRNAs are transduced into target cells before selection.64

As with all cellular screens, CRISPR/Cas9 screens can be performed clonally with individual vectors, yet the most efficient screens are those that can be done with a pooled vector library on large populations of cells.61, 62, 63 An inherent benefit of pooled sgRNA screens is that each sgRNA sequence doubles as a molecular barcode, such that the depletion or enrichment of individual sgRNAs can easily be linked back to the gene it targets by deep sequencing. This makes it possible to perform saturation mutagenesis screens for essentially any cell-intrinsic phenotype, so long as a reliable means of phenotypic selection is available (Figures 2b and c). A similar approach is commonly used with lentiviral short hairpin RNA libraries,64 and the same screening principle can potentially be applied to any transducible cell from any organism: a published genome sequence being the only prerequisite for library design.

There are dozens of immunological phenomena that might be subjected to a genome-wide cellular screen. For instance, how does a pathogen infect its target cell, and how does that cell react? How does a hematopoietic stem cell give rise to a diverse range of lineages, and what allows a malignant lymphocyte to escape these boundaries of proliferation? These questions and many others may all be addressed with cellular CRISPR/Cas9 mutagenesis.

CONCLUSION

Mutation has made our immune system what it is today, but a single mutation can also reveal its most obscure secrets. Billions of new genetic variants are entering the human gene pool in every generation, and those that compromise immunity will inevitably end up in the clinic. In many cases, finding the causative mutation is now easier than ever before, and there are new tools to measure the immunological effects of genetic variation in healthy individuals. It is also easier to create mutations in an experimental setting, be they in a mouse or a human cell, and public resources of mutant cells and organisms are more accessible than ever. Applying these resources to systematically understand immunity requires coordinated effort, but genome-wide mutagenesis platforms are becoming accessible even to the modestly funded laboratory.

The field of immunology is already faced with an abundance of questions, and single-gene mutations in humans, animals or cells have been some of the very best tools to answer them. But new questions are emerging at every turn: only 20 years ago were we oblivious to the molecular basis of microbial recognition, regulatory T cells and innate lymphoid cells. The acquired immunodeficiency and severe acute respiratory syndromes (and the pathogens that cause them) are also relatively new. For all of these phenomena, and indeed for those that might emerge next, there is much to be gained from the continued application of mutation.

Acknowledgments

I thank B Beutler, C Goodnow, R Cornall, A Bradley and K Yusa for discussions. This study was supported by the Wellcome Trust (100083/Z/12/Z).

The author declares no conflict of interest.

References

  1. Marraffini LA, Sontheimer EJ. CRISPR interference: RNA-directed adaptive immunity in bacteria and archaea. Nat Rev Genet. 2010;11:181–190. doi: 10.1038/nrg2749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Ronald PC, Beutler B. Plant and animal sensors of conserved microbial signatures. Science. 2010;330:1061–1064. doi: 10.1126/science.1189468. [DOI] [PubMed] [Google Scholar]
  3. Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell. 2010;140:805–820. doi: 10.1016/j.cell.2010.01.022. [DOI] [PubMed] [Google Scholar]
  4. Pancer Z, Amemiya CT, Ehrhardt GRA, Ceitlin J, Gartland GL, Cooper MD.Somatic diversification of variable lymphocyte receptors in the agnathan sea lamprey Nature 2004430174–180.15241406 [Google Scholar]
  5. Wang F, Ekiert DC, Ahmad I, Yu W, Zhang Y, Bazirgan O, et al. Reshaping antibody diversity. Cell. 2013;153:1379–1393. doi: 10.1016/j.cell.2013.04.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Muyldermans S. Nanobodies: natural single-domain antibodies. Annu Rev Biochem. 2013;82:775–797. doi: 10.1146/annurev-biochem-063011-092449. [DOI] [PubMed] [Google Scholar]
  7. Waterston RH, Lindblad-Toh K, Birney E, Rogers J, Abril JF, et al. Mouse Genome Sequencing Consortium. Initial sequencing and comparative analysis of the mouse genome. Nature. 2002;420:520–562. doi: 10.1038/nature01262. [DOI] [PubMed] [Google Scholar]
  8. Nielsen R, Bustamante C, Clark AG, Glanowski S, Sackton TB, Hubisz MJ, et al. A scan for positively selected genes in the genomes of humans and chimpanzees. PLoS Biol. 2005;3:e170. doi: 10.1371/journal.pbio.0030170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Chapman SJ, Hill AVS. Human genetic susceptibility to infectious disease. Nat Rev Genet. 2012;13:175–188. doi: 10.1038/nrg3114. [DOI] [PubMed] [Google Scholar]
  10. Wellcome Trust Case Control Consortium Genome-wide association study of 14 000 cases of seven common diseases and 3000 shared controls. Nature. 2007;447:661–678. doi: 10.1038/nature05911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Nejentsev S, Walker N, Riches D, Egholm M, Todd JA. Rare variants of IFIH1, a gene implicated in antiviral responses, protect against type 1 diabetes. Science. 2009;324:387–389. doi: 10.1126/science.1167728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Morens DM, Folkers GK, Fauci AS. The challenge of emerging and re-emerging infectious diseases. Nature. 2004;430:242–249. doi: 10.1038/nature02759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Kong A, Frigge ML, Masson G, Besenbacher S, Sulem P, Magnusson G, et al. Rate of de novo mutations and the importance of father's age to disease risk. Nature. 2012;488:471–475. doi: 10.1038/nature11396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Fu W, Akey JM. Selection and adaptation in the human genome. Annu Rev Genomics Hum Genet. 2013;14:467–489. doi: 10.1146/annurev-genom-091212-153509. [DOI] [PubMed] [Google Scholar]
  15. Tennessen JA, Bigham AW, O'Connor TD, Fu W, Kenny EE, Gravel S, et al. Evolution and functional impact of rare coding variation from deep sequencing of human exomes. Science. 2012;337:64–69. doi: 10.1126/science.1219240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Fu W, O'Connor TD, Jun G, Kang HM, Abecasis G, Leal SM, et al. Analysis of 6515 exomes reveals the recent origin of most human protein-coding variants. Nature. 2013;493:216–220. doi: 10.1038/nature11690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bolze A, Mahlaoui N, Byun M, Turner B, Trede N, Ellis SR, et al. Ribosomal protein SA haploinsufficiency in humans with isolated congenital asplenia. Science. 2013;340:976–978. doi: 10.1126/science.1234864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Ombrello MJ, Remmers EF, Sun G, Freeman AF, Datta S, Torabi-Parizi P, et al. Cold urticaria, immunodeficiency, and autoimmunity related to PLCG2 deletions. New Engl J Med. 2012;366:330–338. doi: 10.1056/NEJMoa1102140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Angulo I, Vadas O, Garçon F, Banham-Hall E, Plagnol V, Leahy TR, et al. Phosphoinositide 3-kinase δ gene mutation predisposes to respiratory infection and airway damage. Science. 2013;342:866–871. doi: 10.1126/science.1243292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Lucas CL, Kuehn HS, Zhao F, Niemela JE, Deenick EK, Palendira U, et al. Dominant-activating germline mutations in the gene encoding the PI(3)K catalytic subunit p110δ result in T cell senescence and human immunodeficiency. Nat Immunol. 2014;15:88–97. doi: 10.1038/ni.2771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Li F-Y, Chaigne-Delalande B, Kanellopoulou C, Davis JC, Matthews HF, Douek DC, et al. Second messenger role for Mg2+ revealed by human T-cell immunodeficiency. Nature. 2011;475:471–476. doi: 10.1038/nature10246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Hambleton S, Salem S, Bustamante J, Bigley V, Boisson-Dupuis S, Azevedo J, et al. IRF8 mutations and human dendritic-cell immunodeficiency. New Engl J Med. 2011;365:127–138. doi: 10.1056/NEJMoa1100066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Poltorak A, He X, Smirnova I, Liu MY, van Huffel C, Du X, et al. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science. 1998;282:2085–2088. doi: 10.1126/science.282.5396.2085. [DOI] [PubMed] [Google Scholar]
  24. Brunkow ME, Jeffery EW, Hjerrild KA, Paeper B, Clark LB, Yasayko S-A, et al. Disruption of a new forkhead/winged-helix protein, scurfin, results in the fatal lymphoproliferative disorder of the scurfy mouse. Nat Genet. 2001;27:68–73. doi: 10.1038/83784. [DOI] [PubMed] [Google Scholar]
  25. Lafaille FG, Pessach IM, Zhang S-Y, Ciancanelli MJ, Herman M, Abhyankar A, et al. Impaired intrinsic immunity to HSV-1 in human iPSC-derived TLR3-deficient CNS cells. Nature. 2012;491:769–773. doi: 10.1038/nature11583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Wang H, Yang H, Shivalila CS, Dawlaty MM, Cheng AW, Zhang F, et al. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell. 2013;153:910–918. doi: 10.1016/j.cell.2013.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Davis MM. A prescription for human immunology. Immunity. 2008;29:835–838. doi: 10.1016/j.immuni.2008.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. White JK, Gerdin A-K, Karp NA, Ryder E, Buljan M, Bussell JN, et al. Genome-wide generation and systematic phenotyping of knockout mice reveals new roles for many genes. Cell. 2013;154:452–464. doi: 10.1016/j.cell.2013.06.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Bamshad MJ, Ng SB, Bigham AW, Tabor HK, Emond MJ, Nickerson DA, et al. Exome sequencing as a tool for Mendelian disease gene discovery. Nat Rev Genet. 2011;12:745–755. doi: 10.1038/nrg3031. [DOI] [PubMed] [Google Scholar]
  30. Bousfiha AA, Jeddane L, Ailal F, Al-Herz W, Conley ME, Cunningham-Rundles C, et al. A phenotypic approach for IUIS PID classification and diagnosis: guidelines for clinicians at the bedside. J Clin Immunol. 2013;33:1078–1087. doi: 10.1007/s10875-013-9901-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Bendall SC, Simonds EF, Qiu P, Amir E-AD, Krutzik PO, Finck R, et al. Single-cell mass cytometry of differential immune and drug responses across a human hematopoietic continuum. Science. 2011;332:687–696. doi: 10.1126/science.1198704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Skarnes WC, Rosen B, West AP, Koutsourakis M, Bushell W, Iyer V, et al. A conditional knockout resource for the genome-wide study of mouse gene function. Nature. 2011;474:337–342. doi: 10.1038/nature10163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, et al. Multiplex genome engineering using CRISPR/Cas systems. Science. 2013;339:819–823. doi: 10.1126/science.1231143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Mali P, Yang L, Esvelt KM, Aach J, Guell M, DiCarlo JE, et al. RNA-guided human genome engineering via Cas9. Science. 2013;339:823–826. doi: 10.1126/science.1232033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Siggs OM, Berger M, Krebs P, Arnold CN, Eidenschenk C, Huber C, et al. A mutation of Ikbkg causes immune deficiency without impairing degradation of IkappaB alpha. Proc Natl Acad Sci USA. 2010;107:3046–3051. doi: 10.1073/pnas.0915098107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Bull KR, Rimmer AJ, Siggs OM, Miosge LA, Roots CM, Enders A, et al. Unlocking the bottleneck in forward genetics using whole-genome sequencing and identity by descent to isolate causative mutations. PLoS Genet. 2013;9:e1003219. doi: 10.1371/journal.pgen.1003219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Moresco EMY, Li X, Beutler B. Going forward with genetics: recent technological advances and forward genetics in mice. Am J Pathol. 2013;182:1462–1473. doi: 10.1016/j.ajpath.2013.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Arnold CN, Barnes MJ, Berger M, Blasius AL, Brandl K, Croker B, et al. ENU-induced phenovariance in mice: inferences from 587 mutations. BMC Res Notes. 2012;5:577. doi: 10.1186/1756-0500-5-577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Siggs OM, Miosge LA, Yates AL, Kucharska EM, Sheahan D, Brdicka T, et al. Opposing functions of the T cell receptor kinase ZAP-70 in immunity and tolerance differentially titrate in response to nucleotide substitutions. Immunity. 2007;27:912–926. doi: 10.1016/j.immuni.2007.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Nelms KA, Goodnow CC. Genome-wide ENU mutagenesis to reveal immune regulators. Immunity. 2001;15:409–418. doi: 10.1016/s1074-7613(01)00199-6. [DOI] [PubMed] [Google Scholar]
  41. Loughran SJ, Kruse EA, Hacking DF, de Graaf CA, Hyland CD, Willson TA, et al. The transcription factor Erg is essential for definitive hematopoiesis and the function of adult hematopoietic stem cells. Nat Immunol. 2008;9:810–819. doi: 10.1038/ni.1617. [DOI] [PubMed] [Google Scholar]
  42. Hoebe K, Du X, Georgel P, Janssen E, Tabeta K, Kim SO, et al. Identification of Lps2 as a key transducer of MyD88-independent TIR signalling. Nature. 2003;424:743–748. doi: 10.1038/nature01889. [DOI] [PubMed] [Google Scholar]
  43. Blasius AL, Arnold CN, Georgel P, Rutschmann S, Xia Y, Lin P, et al. Slc15a4, AP-3, and Hermansky-Pudlak syndrome proteins are required for Toll-like receptor signaling in plasmacytoid dendritic cells. Proc Natl Acad Sci USA. 2010;107:19973–19978. doi: 10.1073/pnas.1014051107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Arnold CN, Pirie E, Dosenovic P, McInerney GM, Xia Y, Wang N, et al. A forward genetic screen reveals roles for Nfkbid, Zeb1, and Ruvbl2 in humoral immunity. Proc Natl Acad Sci USA. 2012;109:12286–12293. doi: 10.1073/pnas.1209134109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Brandl K, Tomisato W, Li X, Neppl C, Pirie E, Falk W, et al. Yip1 domain family, member 6 (Yipf6) mutation induces spontaneous intestinal inflammation in mice. Proc Natl Acad Sci USA. 2012;109:12650–12655. doi: 10.1073/pnas.1210366109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Crozat K, Georgel P, Rutschmann S, Mann N, Du X, Hoebe K, et al. Analysis of the MCMV resistome by ENU mutagenesis. Mamm Genome. 2006;17:398–406. doi: 10.1007/s00335-005-0164-2. [DOI] [PubMed] [Google Scholar]
  47. Vidal SM, Malo D, Marquis J-F, Gros P. Forward genetic dissection of immunity to infection in the mouse. Annu Rev Immunol. 2008;26:81–132. doi: 10.1146/annurev.immunol.26.021607.090304. [DOI] [PubMed] [Google Scholar]
  48. Lewis MA, Quint E, Glazier AM, Fuchs H, de Angelis MH, Langford C, et al. An ENU-induced mutation of miR-96 associated with progressive hearing loss in mice. Nat Genet. 2009;41:614–618. doi: 10.1038/ng.369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Velazquez L, Fellous M, Stark GR, Pellegrini S. A protein tyrosine kinase in the interferon alpha/beta signaling pathway. Cell. 1992;70:313–322. doi: 10.1016/0092-8674(92)90105-l. [DOI] [PubMed] [Google Scholar]
  50. Straus DB, Weiss A. Genetic evidence for the involvement of the lck tyrosine kinase in signal transduction through the T cell antigen receptor. Cell. 1992;70:585–593. doi: 10.1016/0092-8674(92)90428-f. [DOI] [PubMed] [Google Scholar]
  51. Yamaoka S, Courtois G, Bessia C, Whiteside ST, Weil R, Agou F, et al. Complementation cloning of NEMO, a component of the IkappaB kinase complex essential for NF-kappaB activation. Cell. 1998;93:1231–1240. doi: 10.1016/s0092-8674(00)81466-x. [DOI] [PubMed] [Google Scholar]
  52. Carette JE, Guimaraes CP, Varadarajan M, Park AS, Wuethrich I, Godarova A, et al. Haploid genetic screens in human cells identify host factors used by pathogens. Science. 2009;326:1231–1235. doi: 10.1126/science.1178955. [DOI] [PubMed] [Google Scholar]
  53. Carette JE, Guimaraes CP, Wuethrich I, Blomen VA, Varadarajan M, Sun C, et al. Global gene disruption in human cells to assign genes to phenotypes by deep sequencing. Nat Biotechnol. 2011;29:542–546. doi: 10.1038/nbt.1857. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Carette JE, Raaben M, Wong AC, Herbert AS, Obernosterer G, Mulherkar N, et al. Ebola virus entry requires the cholesterol transporter Niemann-Pick C1. Nature. 2011;477:340–343. doi: 10.1038/nature10348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Duncan LM, Timms RT, Zavodszky E, Cano F, Dougan G, Randow F, et al. Fluorescence-based phenotypic selection allows forward genetic screens in haploid human cells. PLoS One. 2012;7:e39651. doi: 10.1371/journal.pone.0039651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Jae LT, Raaben M, Riemersma M, van Beusekom E, Blomen VA, Velds A, et al. Deciphering the glycosylome of dystroglycanopathies using haploid screens for lassa virus entry. Science. 2013;340:479–483. doi: 10.1126/science.1233675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Leeb M, Wutz A. Derivation of haploid embryonic stem cells from mouse embryos. Nature. 2011;479:131–134. doi: 10.1038/nature10448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Elling U, Taubenschmid J, Wirnsberger G, O'Malley R, Demers S-P, Vanhaelen Q, et al. Forward and reverse genetics through derivation of haploid mouse embryonic stem cells. Cell Stem Cell. 2011;9:563–574. doi: 10.1016/j.stem.2011.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Yang H, Liu Z, Ma Y, Zhong C, Yin Q, Zhou C, et al. Generation of haploid embryonic stem cells from Macaca fascicularis monkey parthenotes. Cell Res. 2013;23:1187–1200. doi: 10.1038/cr.2013.93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science. 2012;337:816–821. doi: 10.1126/science.1225829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Koike-Yusa H, Li Y, Tan E-P, Del Castillo Velasco-Herrera M, Yusa K. Genome-wide recessive genetic screening in mammalian cells with a lentiviral CRISPR-guide RNA library. Nat Biotechnol. 2014;32:267–273. doi: 10.1038/nbt.2800. [DOI] [PubMed] [Google Scholar]
  62. Shalem O, Sanjana NE, Hartenian E, Shi X, Scott DA, Mikkelsen TS, et al. Genome-scale CRISPR-Cas9 knockout screening in human cells. Science. 2014;343:84–87. doi: 10.1126/science.1247005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Wang T, Wei JJ, Sabatini DM, Lander ES. Genetic screens in human cells using the CRISPR-Cas9 system. Science. 2014;343:80–84. doi: 10.1126/science.1246981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Ngo VN, Davis RE, Lamy L, Yu X, Zhao H, Lenz G, et al. A loss-of-function RNA interference screen for molecular targets in cancer. Nature. 2006;441:106–110. doi: 10.1038/nature04687. [DOI] [PubMed] [Google Scholar]
  65. Yang H, Wang H, Shivalila CS, Cheng AW, Shi L, Jaenisch R. One-step generation of mice carrying reporter and conditional alleles by CRISPR/Cas-mediated genome engineering. Cell. 2013;154:1370–1379. doi: 10.1016/j.cell.2013.08.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Lee CC, Carette JE, Brummelkamp TR, Ploegh HL. A. Reporter screen in a human haploid cell line identifies CYLD as a constitutive inhibitor of NF-κB. PLoS One. 2013;8:e70339. doi: 10.1371/journal.pone.0070339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Siggs OM, Arnold CN, Huber C, Pirie E, Xia Y, Lin P, et al. The P4-type ATPase ATP11C is essential for B lymphopoiesis in adult bone marrow. Nat Immunol. 2011;12:434–440. doi: 10.1038/ni.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Yabas M, Teh CE, Frankenreiter S, Lal D, Roots CM, Whittle B, et al. ATP11C is critical for the internalization of phosphatidylserine and differentiation of B lymphocytes. Nat Immunol. 2011;12:441–449. doi: 10.1038/ni.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Jun JE, Wilson LE, Vinuesa CG, Lesage S, Blery M, Miosge LA, et al. Identifying the MAGUK protein Carma-1 as a central regulator of humoral immune responses and atopy by genome-wide mouse mutagenesis. Immunity. 2003;18:751–762. doi: 10.1016/s1074-7613(03)00141-9. [DOI] [PubMed] [Google Scholar]
  70. Tze LE, Horikawa K, Domaschenz H, Howard DR, Roots CM, Rigby RJ, et al. CD83 increases MHC II and CD86 on dendritic cells by opposing IL-10-driven MARCH1-mediated ubiquitination and degradation. J Exp Med. 2011;208:149–165. doi: 10.1084/jem.20092203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Shiow LR, Roadcap DW, Paris K, Watson SR, Grigorova IL, Lebet T, et al. The actin regulator coronin 1A is mutant in a thymic egress-deficient mouse strain and in a patient with severe combined immunodeficiency. Nat Immunol. 2008;9:1307–1315. doi: 10.1038/ni.1662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Park H, Staehling K, Tsang M, Appleby MW, Brunkow ME, Margineantu D, et al. Disruption of fnip1 reveals a metabolic checkpoint controlling B lymphocyte development. Immunity. 2012;36:769–781. doi: 10.1016/j.immuni.2012.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Ordoñez-Rueda D, Jönsson F, Mancardi DA, Zhao W, Malzac A, Liang Y, et al. A hypomorphic mutation in the Gfi1 transcriptional repressor results in a novel form of neutropenia. Eur J Immunol. 2012;42:2395–2408. doi: 10.1002/eji.201242589. [DOI] [PubMed] [Google Scholar]
  74. Lu P, Hankel IL, Knisz J, Marquardt A, Chiang M-Y, Grosse J, et al. The Justy mutation identifies Gon4-like as a gene that is essential for B lymphopoiesis. J Exp Med. 2010;207:1359–1367. doi: 10.1084/jem.20100147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Wu Z, Jia X, La Cruz de L, Su X-C, Marzolf B, Troisch P, et al. Memory T cell RNA rearrangement programmed by heterogeneous nuclear ribonucleoprotein hnRNPLL. Immunity. 2008;29:863–875. doi: 10.1016/j.immuni.2008.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Papathanasiou P, Perkins AC, Cobb BS, Ferrini R, Sridharan R, Hoyne GF, et al. Widespread failure of hematolymphoid differentiation caused by a recessive niche-filling allele of the Ikaros transcription factor. Immunity. 2003;19:131–144. doi: 10.1016/s1074-7613(03)00168-7. [DOI] [PubMed] [Google Scholar]
  77. Feng J, Wang H, Morse HC. Functional deficiency in IL-7 caused by an N-ethyl-N-nitrosourea-induced point mutation. Genetics. 2007;175:545–551. doi: 10.1534/genetics.106.066043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Huang H-W, Chiang Y-J, Hung S-I, Li C-L, Yen JJ-Y. An IL-7 splicing-defect lymphopenia mouse model revealed by genome-wide mutagenesis. J Biomed Sci. 2007;14:169–181. doi: 10.1007/s11373-006-9135-1. [DOI] [PubMed] [Google Scholar]
  79. Crozat K, Eidenschenk C, Jaeger BN, Krebs P, Guia S, Beutler B, et al. Impact of β2 integrin deficiency on mouse natural killer cell development and function. Blood. 2011;117:2874–2882. doi: 10.1182/blood-2010-10-315457. [DOI] [PubMed] [Google Scholar]
  80. Wen BG, Pletcher MT, Warashina M, Choe SH, Ziaee N, Wiltshire T, et al. Inositol (1,4,5) trisphosphate 3 kinase B controls positive selection of T cells and modulates Erk activity. Proc Natl Acad Sci USA. 2004;101:5604–5609. doi: 10.1073/pnas.0306907101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Nijnik A, Woodbine L, Marchetti C, Dawson S, Lambe T, Liu C, et al. DNA repair is limiting for haematopoietic stem cells during ageing. Nature. 2007;447:686–690. doi: 10.1038/nature05875. [DOI] [PubMed] [Google Scholar]
  82. Barouch-Bentov R, Che J, Lee CC, Yang Y, Herman A, Jia Y, et al. A conserved salt bridge in the G loop of multiple protein kinases is important for catalysis and for in vivo Lyn function. Mol Cell. 2009;33:43–52. doi: 10.1016/j.molcel.2008.12.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Papathanasiou P, Tunningley R, Pattabiraman DR, Ye P, Gonda TJ, Whittle B, et al. A recessive screen for genes regulating hematopoietic stem cells. Blood. 2010;116:5849–5858. doi: 10.1182/blood-2010-04-269951. [DOI] [PubMed] [Google Scholar]
  84. Gosling KM, Makaroff LE, Theodoratos A, Kim Y-H, Whittle B, Rui L, et al. A mutation in a chromosome condensin II subunit, kleisin beta, specifically disrupts T cell development. Proc Natl Acad Sci USA. 2007;104:12445–12450. doi: 10.1073/pnas.0704870104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Park H, Staehling-Hampton K, Appleby MW, Brunkow ME, Habib T, Zhang Y, et al. A point mutation in the murine Hem1 gene reveals an essential role for Hematopoietic protein 1 in lymphopoiesis and innate immunity. J Exp Med. 2008;205:2899–2913. doi: 10.1084/jem.20080340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Miosge LA, Blasioli J, Blery M, Goodnow CC. Analysis of an ethylnitrosourea-generated mouse mutation defines a cell intrinsic role of nuclear factor kappaB2 in regulating circulating B cell numbers. J Exp Med. 2002;196:1113–1119. doi: 10.1084/jem.20020959. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Wu Z, Yates AL, Hoyne GF, Goodnow CC. Consequences of increased CD45RA and RC isoforms for TCR signaling and peripheral T cell deficiency resulting from heterogeneous nuclear ribonucleoprotein L-like mutation. J Immunol. 2010;185:231–238. doi: 10.4049/jimmunol.0903625. [DOI] [PubMed] [Google Scholar]
  88. Liang Y, Cucchetti M, Roncagalli R, Yokosuka T, Malzac A, Bertosio E, et al. The lymphoid lineage-specific actin-uncapping protein Rltpr is essential for costimulation via CD28 and the development of regulatory T cells. Nat Immunol. 2013;14:858–866. doi: 10.1038/ni.2634. [DOI] [PubMed] [Google Scholar]
  89. Beisner DR, Langerak P, Parker AE, Dahlberg C, Otero FJ, Sutton SE, et al. The intramembrane protease Sppl2a is required for B cell and DC development and survival via cleavage of the invariant chain. J Exp Med. 2013;210:23–30. doi: 10.1084/jem.20121072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Bergmann H, Yabas M, Short A, Miosge L, Barthel N, Teh CE, et al. B cell survival, surface BCR and BAFFR expression, CD74 metabolism, and CD8- dendritic cells require the intramembrane endopeptidase SPPL2A. J Exp Med. 2013;210:31–40. doi: 10.1084/jem.20121076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Theodoratos A, Whittle B, Enders A, Tscharke DC, Roots CM, Goodnow CC, et al. Mouse strains with point mutations in TAP1 and TAP2. Immunol Cell Biol. 2010;88:72–78. doi: 10.1038/icb.2009.61. [DOI] [PubMed] [Google Scholar]
  92. Jenne CN, Enders A, Rivera R, Watson SR, Bankovich AJ, Pereira JP, et al. T-bet-dependent S1P5 expression in NK cells promotes egress from lymph nodes and bone marrow. J Exp Med. 2009;206:2469–2481. doi: 10.1084/jem.20090525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Johnson AL, Aravind L, Shulzhenko N, Morgun A, Choi S-Y, Crockford TL, et al. Themis is a member of a new metazoan gene family and is required for the completion of thymocyte positive selection. Nat Immunol. 2009;10:831–839. doi: 10.1038/ni.1769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Kakugawa K, Yasuda T, Miura I, Kobayashi A, Fukiage H, Satoh R, et al. A novel gene essential for the development of single positive thymocytes. Mol Cell Biol. 2009;29:5128–5135. doi: 10.1128/MCB.00793-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Jakob T, Köllisch GV, Howaldt M, Bewersdorff M, Rathkolb B, Müller ML, et al. Novel mouse mutants with primary cellular immunodeficiencies generated by genome-wide mutagenesis J Allergy Clin Immunol 2008121179–184.e7. [DOI] [PubMed] [Google Scholar]
  96. Siggs OM, Li X, Xia Y, Beutler B. ZBTB1 is a determinant of lymphoid development. J Exp Med. 2012;209:19–27. doi: 10.1084/jem.20112084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Enders A, Stankovic S, Teh C, Uldrich AP, Yabas M, Juelich T, et al. ZBTB7B (Th-POK) regulates the development of IL-17-producing CD1d-restricted mouse NKT cells. J Immunol. 2012;189:5240–5249. doi: 10.4049/jimmunol.1201486. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Immunology and Cell Biology are provided here courtesy of Nature Publishing Group

RESOURCES