Skip to main content

REVIEW article

Front. Immunol., 05 November 2015
Sec. T Cell Biology
This article is part of the Research Topic Setting the path – How qualitative and quantitative differences in TCR-pMHC interactions regulate T cell immune responses View all 12 articles

The Timing of T Cell Priming and Cycling

  • Institute for Immunology, Ludwig-Maximilians-University Munich, Munich, Germany

The proliferation of specific lymphocytes is the central tenet of the clonal selection paradigm. Antigen recognition by T cells triggers a series of events that produces expanded clones of differentiated effector cells. TCR signaling events are detectable within seconds and minutes and are likely to continue for hours and days in vivo. Here, I review the work done on the importance of TCR signals in the later part of the expansion phase of the primary T cell response, primarily regarding the regulation of the cell cycle in CD4+ and CD8+ cells. The results suggest a degree of programing by early signals for effector differentiation, particularly in the CD8+ T cell compartment, with optimal expansion supported by persistent antigen presentation later on. Differences to CD4+ T cell expansion and new avenues toward a molecular understanding of cell cycle regulation in lymphocytes are discussed.

Introduction

The priming of naive T cells, i.e., their activation following a primary recognition of specific peptide–MHC complexes, consists of a series of biophysical, biochemical, genetic, and proliferative events that lead to populations of expanded clones of differentiated effector cells, some of which have the potential to become long-lived memory cells. This clonal lymphocyte expansion is at the center of the still ruling clonal selection paradigm of adaptive immunity by Talmage, Lederberg (1, 2), and Burnet who hypothesized that lymphocyte “proliferation will be initiated of all those clones whose reactive sites correspond to the antigenic determinants on the antigen used” (3). Soon after its proposal, it was observed that indeed a proportion of small parental lymphocytes transferred into neonatal F1 hosts enlarge, become “pyroninophilic,” i.e., express large amounts of mostly ribosomal RNA, and incorporate tritiated thymidine, i.e., proliferate, before the recipients succumb to graft-versus-host disease 2–3 weeks later (4, 5). Quickly, such results were translated to in vitro proliferation assays for the detection of immunoresponsive cells among splenocytes from immunized animals (6) or human allo-reactive lymphocytes (7). The identification of the TCR proteins and genes and how to trigger T cell responses by monoclonal antibodies and second messenger agonists set the stage for closer analyses of the molecular events initiating proliferation (810) for which the mutagenesis of T cell tumor cell lines has been an especially fruitful approach (11). The adoptive transfer of TCR-transgenic T cells and the use of tracking dyes visualized cell populations expand and contract in vivo (12, 13). The detection of endogenous T cells of defined specificities by restimulation or tetramer assays also confirmed that the key driver of adaptive immunity is antigen (14, 15).

Research on the biophysics of TCR-peptide/MHC interaction (16, 17), the biochemistry of signal transducers (17, 18), transcription (19), and proliferation has naturally focused on different time frames of seconds, minutes, hours, and days, respectively, while an overall integration of these events, as noted recently, is still missing (20). The TCR signals convert a small metabolically inactive resting cell into an expanded group of descendants with newly acquired migratory and effector functions. While the dominant function of CD8+ T cells is the deletion of cells infected by intracellular pathogens like viruses and bacteria, their CD4+ counterparts differentiate following microenvironmental cues into discrete lineages profiled by cytokines expressed (21, 22). Over the following weeks, T cell numbers decline and a residual population survives as memory cells. Important in this context is that differentiation and proliferation coincide but are molecularly not necessarily coupled (23).

Here, we discuss the effects and requirements of TCR signaling for T cell proliferation in the primary response: What aspects of T cell differentiation follow analog versus digital logics of signal processing? Is T cell proliferation programed early on or is it maintained by continued antigen triggers? These questions have been addressed in a variety of experimental systems. The data further our understanding of the nature of immune responses to complex pathogens and our capability to design more immunogenic T cell vaccines (24).

Lessons Learned from Intravital Imaging

Experiments using intravital 2-photon microscopy showed that in the steady state, a dendritic cell (DC) interacts with 500–5000 T cells per hour, which migrate within lymph nodes with a velocity of 10–12 μm per min, thereby scanning uncounted peptide/MHC complexes (2527). Interestingly, CD4+ and CD8+ T cells employ different strategies of surveillance: The interaction times of CD4+ T cells with DCs depend on MHC class II (MHC-II) molecules while CD8+ T cells traverse a LN slower and regardless of self-peptide/MHC-I complexes. They scan 160–200 and 300 DCs per hour and, thus, stay in a lymph node approximately for half a day and a day, respectively (28). Considering natural precursor frequencies, it has been assessed that at least 85 antigen-presenting DCs per lymph node are necessary to initiate a CD4+ T cell response (29). When confronted with a DC presenting antigen, specific T cells stop migrating and stay in touch with an individual DC for around a day (3033). Within this period, the T cells undergo changes classically summarized as “blasting”: They increase in size, double their protein contents, increase their total RNA contents 30-fold, induce the expression of around 1300 mRNAs, and change their metabolism before proliferation ensues 24 h later (3440).The stable interaction with one DC can be preceded by a phase of transient interactions with several APCs, the length of which is inversely correlated with APC density and antigen dose (41). Interestingly, it has been shown that, for CD8+ T cells, the phase of stable pairing with one DC is not necessarily required for expansion of effector clones, while memory differentiation is affected. These findings indicated that the memory potential of CD8+ T cells can be programed within the first 24 h of priming (36). The data also supported the relevance of observations that T cells can “memorize” sequential sub-threshold interactions with different APCs and accumulate such signals over time, perhaps via AP1 or NFAT (4246).

Counting Precursors and Effectors to Assess the Level of Expansion

The end result of priming is a population of expanded clones, the numerology of which has recently been assessed with great precision. It turned out that the precursor frequency of specific cells in the naive repertoire is a critical parameter for the magnitude of a T cell response. Even for “strong” antigens with unusually high precursor frequencies like alloantigens, estimates using classical techniques like limiting dilution analysis have been notoriously variable by several orders of magnitude (47), before the actual frequency of around 10% could be clarified with cell-tracking dyes in vivo (48, 49). However, the much lower precursor frequencies to standard antigens could not be determined until peptide/MHC tetramers were employed for enriching the naive precursors from unimmunized animals. This technology is likely geared toward the detection of high-affinity T cells and will evolve further (5053). For CD8+ T cells, precursor frequencies were found between 1 and 100 cells per 106 cells, while the numbers were three to five times lower for CD4+ T cells in both mouse and man (54, 55), confirming previous findings based on single-cell transfers and comparisons with expanded cells derived from the endogenous repertoire (56). Such numbers are important as they can partially predict the T cells’ clonal burst size and an individual’s immunological potential, though other mechanisms like antigen presentation efficiency, peptide/MHC stability, kind of APC, inflammation, niche issues, and self-reactivity likely contribute (5765). Importantly, the traficking and microfluidics of the secondary lymphoid organs are arranged to make priming an efficient process so that all existing rare precursors are recruited into an immune response rapidly and efficiently (29, 33, 66). Tetramer and restimulation assays have allowed for a determination of the T cells’ clonal burst size and, thus, the determination of their proliferative capacity.

In responses to several epitopes derived from lymphocytic choriomeningitis virus (LCMV), an acute viral infection that triggers exceptionally strong T cell responses, CD8+ T cells divide about 15 times (14, 15) and 16–19 times in response to the Gram-positive intracellular bacterium Listeria monocytogenes (67). More recently, two studies on clones derived from individually labeled CD8+ T cells responding to Listeria showed, on average, 15 divisions, but also a wide, and mysterious, variation of 10–20 divisions that was independent of TCR specificity (58, 59). By contrast, tetramer studies determined the division numbers of CD4+ T cells to 7 on average, with different specificities falling between 4 and 10 divisions (54, 55, 68). This difference between CD4+ and CD8+ T cells confirms earlier studies in LCMV, vaccinia and Sendai virus infections where CD8+ T cells expand to a much greater extent than their CD4+ counterparts (6971) and an intrinsic difference between the cell types, regarding their proliferative potential, had been suggested early on (72, 73).

Autopilot Experiments: CD8s

Early immunizations of TCR-transgenic cd28-deleted animals had suggested that the costimulatory requirements of CD8+ T cell priming can be overcome by prolonged antigen presentation (74). However, since it was quickly seen that TCR-transgenic animals are no models of T cell expansion (75, 76), more physiological systems were developed. To investigate the role of antigen persistence in the process of T cell priming the TCR signal has to be interrupted experimentally. The first approach to do this was to prime naive CD4+ T cells in plates coated with peptide/MHC complexes and transfer the cells to uncoated wells at different times, with the effects of costimulation, APC, and responder cell types assessed as well. The authors concluded that sustained TCR signaling is the key parameter for the priming of naive CD4+ T cells (77).

These results provoked work on the antigen dependence of CD8+ T cell expansion that projected a different picture. Naive P14 TCR-transgenic T cells primed for 24 h continued their proliferation in antigen-free cultures and upon transfer into antigen-free recipients (78). Experiments with OT-1 T cells primed with fibroblast transfectants or antibodies showed that a priming period as short as 2 or 2.5 h was sufficient for the T cells to continue their divisions over the following days in vitro (79, 80). In a follow-up study, however, it was shown that a 20-h period of priming was necessary for the quantitative maintenance of the cells upon transfer into recipient animals, indicating that proliferation in vitro can reflect an incomplete level of activation that leaves the cells ill-equipped for survival in secondary lymphoid organs (81). Not entirely consistent was the later observation that the 4-h-primed cells survive sufficiently to protect recipients against an OVA-expressing tumor (82).

The difference between the 2001 and 2003 studies by the Schoenberger group done in vitro and partially in vivo (79, 81) was later shown to be based on a factor working in trans between T cells cultured at high densities, namely IL-2 (83). Importantly, the merely accessory role of IL-2 for primary T cell expansion in vivo has been demonstrated in several infectious, transplantation, and vaccination models (84). For antigen-independent proliferation of CD8+ T cells, IL-2 has a role in vitro, but not in vivo (80, 85). Though IL-2 has been used for culturing T cell lines and clones in the laboratory for decades, in vivo it mostly affects the maintenance of CD25+ regulatory T cells (8688), which are the only T cells where IL-2/STAT5 signaling can be detected following immunization (89), and the programing of an efficient secondary response (90, 91). However, the contribution of IL-2 signals to T cell expansion in vivo is not 0 and its actions may transmit signals from regulatory T cells and the inflammatory microenvironment (57, 9294).

The results of the 2001 “autopilot” publications had been foreshadowed by experiments using the timed application of ampicillin to remove intracellular antigen-delivering L. monocytogenes bacteria from infected animals. The expansion of CD8+ T cells specific for two Listeria-derived epitopes continued, despite the efficient removal of live bacteria 24 h p.i. (95). In hindsight, one may ask what the half-life of residual antigen presentation, which has been reported in several infection models since, might have been. However, the swift disappearance of T cell antigen from Listeria-infected animals within a day after antibiotic treatment was shown later by T cell transfers and in vitro assays and supported the initial conclusion that CD8+ T cell proliferation is programed in vivo within the initial 24 h of priming (96, 97). The observation that T cell contraction is programed as well strengthened this view (98, 99).

These observations are widely cited and considered seminal (100103). The behavior of CD8+ T cells to execute a program set in place within the first day of priming has been illustratively summarized as being “on autopilot” or “programed” (100, 102). However, there are a significant number of publications that do not fit the scheme. Curtsinger et al. showed that adoptively transferred OT-1 T cells primed in vitro 16 h earlier did not proliferate and accumulate in antigen-free hosts compared to previously vaccinated ones (104). It is possible that in these experiments the programing threshold at 24 h had not yet been reached. The paper also showed that inflammatory stimuli can affect CD8+ T cell expansion, in agreement with experiments using P14 TCR-transgenic cells whose cognate antigen Db/LCMV-GP33–41 has a short half-life in vivo. Comparing animals immunized with peptide, virus-like particles, and live virus, CD8+ T cell expansion correlated with antigen persistence, clearly arguing against an “autopilot” model (105). In the Listeria model, it was reported that ampicillin treatment between 24 and 60 h p.i. affected both endogenous CD4+ and CD8+ T cell expansion when read out in a peptide restimulation assay (106). The main difference to the earlier reports from four different laboratories was that they had used adoptive transfers of TCR-transgenic T cells, allowing for a clearly defined starting point of the expansion phase, and had visualized the proliferating T cells directly by fluorescent tracking dyes (78, 80, 85, 96, 97).

In immunizations with irradiated Plasmodium yoelii sporozoites, it was shown that a T cell antigen derived from them can persist for months. Using consecutive transfers of TCR-transgenic T cells specific for P. yoelii between recipients immunized with P. yoelii or Plasmodium falciparum, it was shown that antigen presentation beyond day 4 significantly contributes to the number of memory cells detectable a month later, implying that it supported a better primary response and revising an earlier interpretation (107, 108).

Another approach to limit the time of antigen exposure is the delayed T cell transfer of naive T cells into recipients infected days earlier. This procedure necessarily varies both parameters, dose and time, of priming antigen. In addition, the transferred cells face a rising competition of endogenous effector cells (109). All these parameters are likely to contribute to the result that the CD8+ T cells transferred at the peak of the response to vaccinia virus on day 7 exhibit overall inefficient priming with lower levels of division, survival, and memory differentiation (110).

The ablation of DCs from CD11c-diphtheria toxin receptor (DTR)/green fluorescent protein (GFP) transgenic animals is another way to interrupt the interactions between T cells and APCs (111, 112). Such animals express a fusion protein of the DTR and GFP on the surface of DCs. Mice are naturally resistant to diphtheria toxin as they lack a receptor for this toxin, so that only transgene-expressing DCs and some macrophages are depleted (113). Since the expression of such conventional transgenes is, like many others, mosaic, the depletion of DCs from such animals is incomplete. In addition, the fact that the depletion of large numbers of DCs in transgenics directly modifies lymphocyte homing to lymph nodes via high endothelial venules is an unintended consequence (114). Thus, the procedure employed by Prlic et al., who sorted DTR/GFP+ DCs, transferred them after peptide loading, and depleted them at different time points of a T cell response, avoided these potential pitfalls. Responding TCR-transgenic cells accumulated linearly with the delay of the depletion, while their functionality in primary and secondary responses remained unchanged (115). Similar results were presented in an elegant study that terminated TCR signals by turning off a tetracycline-inducible lck gene during the expansion phase of endogenous cells responding to vaccinia virus (116). Also, the expansion of cells primed in vitro was significantly supported only in recipients that underwent an infection with the cognate antigen (117). Thus, these three datasets supported the interpretation that the differentiation of CD8+ T cells is programed within the initial 24 h of priming, but not their quantitative accumulation.

Blocking antigen presentation in vivo with peptide/MHC-specific antibodies has been difficult as only few reagents with a sufficiently high affinity exist. Blair et al. used antibodies specific for Kb/Ova257–264 and Ab/Eα52–68 to block the priming of TCR-transgenic T cells with the respective specificities following systemic infections with modified VSV. The results showed that both CD4+ and CD8+ T cell responses were affected. Though the antibodies blocked T cell expansion only incompletely, the results argue against the autopilot mode of T cell expansion (118).

These discrepancies cannot easily be explained by experimental details and modes of infection. We are left with two groups of publications of about equal numbers, arguing for and against an antigen-independent phase of CD8+ T cell expansion. Perhaps one is left with the understanding that CD8+ T cell proliferation and differentiation is initially programed, but is not hardwired and can be modified by later TCR signals (21). Several more recent studies looking at the effects of viral antigen presented locally support this view.

In the first 2 days of a pulmonary influenza infection, alveolar DCs migrate to the draining lymph nodes for priming (119, 120). The remaining DCs and macrophages can be partially depleted by i.n. application of clodronate-containing liposomes. By using this technique, McGill et al. showed that CD8+ T cells expand in response to antigen presented by such APCs in the lung, suggesting a “two-hit model” of CD8+ T cell expansion (121, 122). The depletion of DCs from CD11c-DTR/GFP transgenics on day 6 p.i. also reduces the number of CD8+ effector cells in the lungs significantly (123). An additional role of antigen presentation following day 7 for the CD8+ memory cell functionality was shown by León et al. in the same system. This paper with insight showed that late antigen presentation depended on specific IgG and FcγR+ DCs, suggesting that immune complexes are cross-presented to CD8+ T cells late in the primary response and affect proper memory cell differentiation (124). A division of labor for effector and memory cell differentiation was also shown for two migratory DC types, one expressing CD103, the other CD11b, arguing for more nuanced T cell-APC interactions via costimulatory molecules, here CD24 (119). In the LCMV infection model, Kang et al. followed primed T cells at the peak of the response entering the CNS and found them synthesizing DNA, implying that they left the secondary lymphoid organs while actively cycling. In the CNS, the cells established long-lived interactions with local DCs and T cell transfers into MHC-I-negative recipients showed that it is local antigen presentation that supports additional divisions within the CNS (125). These data from infection models stress the role late antigen presentation has on CD8+ T cell proliferation.

Autopilot Experiments: CD4s

The initial in vitro experiments by Iezzi et al., reporting the necessity of antigen persistence for several days, were done with CD4+ T cells exposed to peptide/MHC complexes for limited periods of time and then recultured in new dishes (77). These findings were reproduced in experiments with antigen-loaded APCs by interrupting the TCR signals with MHC-specific antibodies (126128) and then in vivo in the Listeria model where the antigen removal by ampicillin affected the CD4+ T cell responses much more than those of the CD8+ T cells (96, 97). A careful study using a heterologous rechallenge extended these findings to secondary responses and established that CD4+ T memory cells do not acquire an “autopilot” phenotype (129). A direct demonstration of antigen dependence of the CD4+ T cell response was possible by using transgenic mice in which the presentation of an MHC-II restricted antigen by DCs could be controlled by doxycycline in vivo. Antigen withdrawal quickly arrested proliferation of adoptively transferred T cells (130). These findings were in agreement with results of Celli et al. who transferred antigen-presenting DCs consecutively and showed the enhanced expansion and effector differentiation of CD4+ T cells and, importantly, visualized the stable contacts between DCs and T cells at later stages in the priming process by intravital microscopy (131, 132). Also the residual antigen left behind following the resolution of an influenza infection and still visible to CD4+ T cells supports their memory differentiation (133).

A caveat to these studies was the high numbers of T cells transferred, a procedure that can affect the results (67, 134137). However, the antigen dependence of CD4+ T cell expansion was also demonstrated by transferring small numbers of transgenic T cells (39, 138). An additional in vivo study looking at anti-HY responses of polyclonal T cells showed that antigen persistence is necessary for CD4+ T cells to expand and license DCs for proper CD8+ priming in a system relying entirely on natural precursor frequencies (139).

This work supported the model that CD4+ T cells require contact with APCs and TCR signals to keep the cell cycle going for several rounds. There are, nevertheless, a number of experiments, mostly done in vitro, that show antigen-independent proliferation of CD4+ T cells, either supported by cytokines (140) or not (141, 142). Interestingly, some of them reported a dampening effect of antigen presented to CD4+ T cells repeatedly, which drives the cells into exhaustion (143), a finding with clear correlates in vivo (144) and in man (77, 145).

Comparing CD4s and CD8s

The few direct comparisons between CD4+ and CD8+ T cells that have been done mostly observe a higher degree of programing in the CD8+ compartment. This was found in experiments with Listeria removed by ampicillin (96, 97) and a more recent side-by-side comparison of in vitro-primed CD4+ and CD8+ T cells. Here, not just proliferation but also functionality and gene expression profiles supported the different consequences of interrupted TCR signals by the two subsets (39). These findings resonate with a number of observations that intrinsically differentiate the two subsets. First, CD4 has a higher affinity to the key signaling molecule lck than CD8 so that different proportions of coreceptors are bound by lck (146, 147). Second, it has been reported in several systems that CD4/CD8 lineage commitment in the thymus is regulated by the kinetics of the positively selecting signal (148150). Third, naive CD4+ T cells traverse lymph nodes faster than CD8+ T cells and do so in an MHC-dependent way (28). Fourth, tissue-resident CD4+ T memory cells recirculate via lymph and blood while their CD8+ counterparts stay put in the tissues and persist there in the absence of antigen (151). Fifth, the different importance of programing for proliferation may explain the different numbers of divisions observed in the two subsets mentioned before (54). And sixth, the higher degree of plasticity of CD4+ T cells is reflected in the multiplicity of lineages they can differentiate into under the guidance of cytokines and transcription factors (22).

In addition, there are also cell-extrinsic differences between the two subsets: MHC-I molecules are expressed by all nucleated cells, but MHC-II by professional APCs only. The substrates of antigen presentation via MHC-I may be short-lived defective ribosomal products rather than comparably stable proteins for the MHC-II pathway (152). Thus, one might speculate that the aspect of programing of CD8+ T cells reflects the more transient antigen presentation by MHC-I (39). Upon activation, the stability of MHC-II molecules on DCs is tightly regulated (153, 154) via ubiquitination (155, 156), presumably by MARCH family members (157, 158), while that of MHC-I molecules is not (39, 153). There is also increasing evidence that the two T cell subsets are primed by different DCs. Antigen targeting with the two antibodies DEC205 and 33D1 revealed that CD8+ T cells are primed by CD205+CD8+ DCs, while CD4+ T cells are engaged best by DCIR2+CD8 DCs (159). These pivotal observations were recently confirmed by genetic means: the two APC subsets are also differentiated by the transcription factors IRF4 and IRF8, respectively (160). Intravital 2-photon microscopy of animals systemically immunized with a non-replicative vaccinia virus has shown directly that the initial priming of CD4+ and CD8+ T cells occurs at spatially separate sites in the lymph node by different DCs. In later phases of the response, CD4+ and CD8+ T cells were seen clustered around the same XCR1+ DCs where T cell help is transmitted, which is evidenced by the finding that CD8+ T cells primed in XCR1+-DC-depleted animals display a “helpless” phenotype (161). A similar division of labor among priming DCs has been shown for local HSV infections where antigen presentation also involves migratory DCs (162). A key question in the future will be how this spatial separation is accomplished and what purpose it might serve.

Outlook: New Approaches to Interrupt TCR Signals and to Visualize the Cell Cycle of T Cells

The shortcomings of the techniques used to study the role of antigen for T cell proliferation so far has recently motivated new approaches. To quickly and specifically interrupt TCR signals at will have been notoriously difficult, especially in vivo [e.g., Ref. (39)]. Art Weiss’ laboratory has worked toward this goal by developing a variant of the indispensible signal transducer ZAP70 whose kinase domain is sensitive to the ATP analog 3-MB-PP1 (163, 164). The combination with a novel Nur77-GFP reporter (165) allowed for the visualization of TCR signals and their proliferative consequences with ZAP70 abruptly turned off at different time points following stimulation in vitro. The experiments demonstrated a clear temporal threshold of 24 h for both CD4+ and CD8+ T cells to commit to proliferation that then commences, mostly driven by IL-2. Both subsets also continued to divide several times following the termination of ZAP70 signaling (166). Experiments using such techniques in vivo to follow immune responses with signaling terminated abruptly will certainly open new windows in the future.

The fact that tracking dyes cannot be resolved by flow cytometry beyond the eighth division, before the expansion of CD8+ T cells is finished, calls for other ways to document the proliferative status of cell populations, especially late in the response. Classical and novel DNA dyes visualize the cell cycle status of heterogeneous cell populations (167, 168) and new antibodies against cell cycle components are available, at least for human cells (169). The nucleoside analog EdU whose detection is based on click chemistry will certainly assist future research on cell cycle control in lymphocytes (170). Particularly, the consecutive injections of BrdU and EdU allow the precise labeling of cells in S phase in vivo even of cells with a complex history: Nussenzweig and colleagues showed by this technique that follicular helper T cells accelerate the cell cycle of germinal center B cells (171, 172). The doxycycline-induced expression of a very stable GFP–H2B fusion protein that then gets diluted upon division has been used to visualize the proliferative history of cells without liberating them from their microenvironment, like hematopoietic stem cells and germinal center B cells (171, 173). Another novel approach is the use of a double transgenic mouse that expresses two different fluorescent dyes under cycle-dependent promoters to differentiate G0/G1 from S/G2/M phases (93, 172, 174176). In vitro, tracking family trees of divided cells by time lapse microscopy showed that the division program, called its division destiny, is programed prior to the first division and passed on for several generations (93). Following CD8+ TCR-transgenic cells through a primary response to influenza, Kinjyo et al. observed that the cells divide at least eight times in a homogeneous and fast manner. At the peak of the response at day 7, however, a small CD62Lhi subpopulation emerges that slows down its cell cycle and express gene sets similar to memory cells. By following division trees in vitro, the authors show that phenotype and cell cycle duration are inherited to daughter cells, suggesting a cell-intrinsic program to diversify the proliferative activity in the priming phase of CD8+ T cells along the split between short-lived and memory-precursor effector cells (176). These data indicate a direction for future research.

Most of the genes analyzed so far are general cell cycle genes active in many cell types: cyclin D3 (177), p27kip1 (178, 179), CDK5 (180), CDK6 (181), Bcl11b (182), FoxM1 (183), myc (23), and Geminin (184). However, the first exception is CTP synthase 1 that exclusively affects lymphocyte proliferation in individuals lacking a functional allele (185). Nevertheless, the molecular regulation of the cell cycle in lymphocytes, and, thus, the core of the clonal selection paradigm, is still a black box. Since the cell cycle of lymphocytes is four to five times faster than that of, e.g., Hela cells, one would assume a specific machinery or cell-type-specific components that run or control proliferation in the adaptive arm of the immune system.

Conflict of Interest Statement

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

This work was supported by GRC grant SFB1054-B07.

References

1. Talmage DW. Allergy and immunology. Annu Rev Med (1957) 8:239–56. doi: 10.1146/annurev.me.08.020157.001323

CrossRef Full Text | Google Scholar

2. Lederberg J. Genes and antibodies. Science (1959) 129:1649–53. doi:10.1126/science.129.3364.1649

CrossRef Full Text | Google Scholar

3. Burnet FM. A modification of Jerne’s theory of antibody production using the concept of clonal selection. Aust J Sci (1957) 20:67–9.

Google Scholar

4. Gowans JL. The fate of parental strain small lymphocytes in F1 hybrid rats. Ann N Y Acad Sci (1962) 99:432–55. doi:10.1111/j.1749-6632.1962.tb45326.x

CrossRef Full Text | Google Scholar

5. Porter KA, Cooper EH. Transformation of adult allogeneic small lymphocytes after transfusion into newborn rats. J Exp Med (1962) 115:997–1008. doi:10.1084/jem.115.5.997

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Dutton RW, Pearce JD. Antigen-dependent stimulation of synthesis of deoxyribonucleic acid in spleen cells from immunized rabbits. Nature (1962) 194:93–4. doi:10.1038/194093a0

CrossRef Full Text | Google Scholar

7. Bach FH, Voynow NK. One-way stimulation in mixed leukocyte cultures. Science (1966) 153:545–7. doi:10.1126/science.153.3735.545

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Oettgen HC, Terhorst C, Cantley LC, Rosoff PM. Stimulation of the T3-T cell receptor complex induces a membrane-potential-sensitive calcium influx. Cell (1985) 40:583–90. doi:10.1016/0092-8674(85)90206-5

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Truneh A, Albert F, Golstein P, Schmitt-Verhulst AM. Early steps of lymphocyte activation bypassed by synergy between calcium ionophores and phorbol ester. Nature (1985) 313:318–20. doi:10.1038/313318a0

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Weiss A, Wiskocil RL, Stobo JD. The role of T3 surface molecules in the activation of human T cells: a two-stimulus requirement for IL 2 production reflects events occurring at a pre-translational level. J Immunol (1984) 133:123–8.

PubMed Abstract | Google Scholar

11. Abraham RT, Weiss A. Jurkat T cells and development of the T-cell receptor signalling paradigm. Nat Rev Immunol (2004) 4:301–8. doi:10.1038/nri1330

CrossRef Full Text | Google Scholar

12. Kearney ER, Pape KA, Loh DY, Jenkins MK. Visualization of peptide-specific T cell immunity and peripheral tolerance induction in vivo. Immunity (1994) 1:327–39. doi:10.1016/1074-7613(94)90084-1

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Lyons AB, Parish CR. Determination of lymphocyte division by flow cytometry. J Immunol Methods (1994) 171:131–7. doi:10.1016/0022-1759(94)90236-4

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Butz EA, Bevan MJ. Massive expansion of antigen-specific CD8+ T cells during an acute virus infection. Immunity (1998) 8:167–75. doi:10.1016/S1074-7613(00)80469-0

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Murali-Krishna K, Altman JD, Suresh M, Sourdive DJ, Zajac AJ, Miller JD, et al. Counting antigen-specific CD8 T cells: a reevaluation of bystander activation during viral infection. Immunity (1998) 8:177–87. doi:10.1016/S1074-7613(00)80470-7

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Huppa JB, Davis MM. The interdisciplinary science of T-cell recognition. Adv Immunol (2013) 119:1–50. doi:10.1016/B978-0-12-407707-2.00001-1

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Malissen B, Bongrand P. Early T cell activation: integrating biochemical, structural, and biophysical cues. Annu Rev Immunol (2015) 33:539–61. doi:10.1146/annurev-immunol-032414-112158

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Chakraborty AK, Weiss A. Insights into the initiation of TCR signaling. Nat Immunol (2014) 15:798–807. doi:10.1038/ni.2940

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Kaech SM, Cui W. Transcriptional control of effector and memory CD8+ T cell differentiation. Nat Rev Immunol (2012) 12:749–61. doi:10.1038/nri3307

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Zikherman J, Au-Yeung B. The role of T cell receptor signaling thresholds in guiding T cell fate decisions. Curr Opin Immunol (2015) 33:43–8. doi:10.1016/j.coi.2015.01.012

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Zehn D, King C, Bevan MJ, Palmer E. TCR signaling requirements for activating T cells and for generating memory. Cell Mol Life Sci (2012) 69:1565–75. doi:10.1007/s00018-012-0965-x

PubMed Abstract | CrossRef Full Text | Google Scholar

22. O’Shea JJ, Paul WE. Mechanisms underlying lineage commitment and plasticity of helper CD4+ T cells. Science (2010) 327:1098–102. doi:10.1126/science.1178334

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Guy CS, Vignali KM, Temirov J, Bettini ML, Overacre AE, Smeltzer M, et al. Distinct TCR signaling pathways drive proliferation and cytokine production in T cells. Nat Immunol (2013) 14:262–70. doi:10.1038/ni.2538

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Burchill MA, Tamburini BA, Pennock ND, White JT, Kurche JS, Kedl RM. T cell vaccinology: exploring the known unknowns. Vaccine (2013) 31:297–305. doi:10.1016/j.vaccine.2012.10.096

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Cahalan MD, Parker I. Choreography of cell motility and interaction dynamics imaged by two-photon microscopy in lymphoid organs. Annu Rev Immunol (2008) 26:585–626. doi:10.1146/annurev.immunol.24.021605.090620

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Moreau HD, Bousso P. Visualizing how T cells collect activation signals in vivo. Curr Opin Immunol (2014) 26:56–62. doi:10.1016/j.coi.2013.10.013

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Qi H, Kastenmüller W, Germain RN. Spatiotemporal basis of innate and adaptive immunity in secondary lymphoid tissue. Annu Rev Cell Dev Biol (2014) 30:141–67. doi:10.1146/annurev-cellbio-100913-013254

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Mandl JN, Liou R, Klauschen F, Vrisekoop N, Monteiro JP, Yates AJ, et al. Quantification of lymph node transit times reveals differences in antigen surveillance strategies of naive CD4+ and CD8+ T cells. Proc Natl Acad Sci U S A (2012) 109:18036–41. doi:10.1073/pnas.1211717109

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Celli S, Day M, Muller AJ, Molina-Paris C, Lythe G, Bousso P. How many dendritic cells are required to initiate a T-cell response? Blood (2012) 120:3945–8. doi:10.1182/blood-2012-01-408260

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Bousso P, Robey E. Dynamics of CD8+ T cell priming by dendritic cells in intact lymph nodes. Nat Immunol (2003) 4:579–85. doi:10.1038/ni928

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Dustin ML, Bromley SK, Kan Z, Peterson DA, Unanue ER. Antigen receptor engagement delivers a stop signal to migrating T lymphocytes. Proc Natl Acad Sci U S A (1997) 94:3909–13. doi:10.1073/pnas.94.8.3909

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Mempel TR, Henrickson SE, von Andrian UH. T-cell priming by dendritic cells in lymph nodes occurs in three distinct phases. Nature (2004) 427:154–9. doi:10.1038/nature02238

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Miller MJ, Hejazi AS, Wei SH, Cahalan MD, Parker I. T cell repertoire scanning is promoted by dynamic dendritic cell behavior and random T cell motility in the lymph node. Proc Natl Acad Sci U S A (2004) 101:998–1003. doi:10.1073/pnas.0306407101

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Grumont R, Lock P, Mollinari M, Shannon FM, Moore A, Gerondakis S. The mitogen-induced increase in T cell size involves PKC and NFAT activation of Rel/NF-kappaB-dependent c-myc expression. Immunity (2004) 21:19–30. doi:10.1016/j.immuni.2004.06.004

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Best JA, Blair DA, Knell J, Yang E, Mayya V, Doedens A, et al. Transcriptional insights into the CD8+ T cell response to infection and memory T cell formation. Nat Immunol (2013) 14:404–12. doi:10.1038/ni.2536

CrossRef Full Text | Google Scholar

36. Henrickson SE, Perro M, Loughhead SM, Senman B, Stutte S, Quigley M, et al. Antigen availability determines CD8+ T cell-dendritic cell interaction kinetics and memory fate decisions. Immunity (2013) 39:496–507. doi:10.1016/j.immuni.2013.08.034

CrossRef Full Text | Google Scholar

37. MacIver NJ, Michalek RD, Rathmell JC. Metabolic regulation of T lymphocytes. Annu Rev Immunol (2013) 31:259–83. doi:10.1146/annurev-immunol-032712-095956

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Mingueneau M, Kreslavsky T, Gray D, Heng T, Cruse R, Ericson J, et al. The transcriptional landscape of αβ T cell differentiation. Nat Immunol (2013) 14:619–32. doi:10.1038/ni.2590

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Rabenstein H, Behrendt AC, Ellwart JW, Naumann R, Horsch M, Beckers J, et al. Differential kinetics of antigen dependency of CD4+ and CD8+ T cells. J Immunol (2014) 192:3507–17. doi:10.4049/jimmunol.1302725

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Buck MD, O’Sullivan D, Pearce EL. T cell metabolism drives immunity. J Exp Med (2015) 212:1345–60. doi:10.1084/jem.20151159

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Henrickson SE, Mempel TR, Mazo IB, Liu B, Artyomov MN, Zheng H, et al. T cell sensing of antigen dose governs interactive behavior with dendritic cells and sets a threshold for T cell activation. Nat Immunol (2008) 9:282–91. doi:10.1038/ni1559

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Gunzer M, Schafer A, Borgmann S, Grabbe S, Zanker KS, Brocker EB, et al. Antigen presentation in extracellular matrix: interactions of T cells with dendritic cells are dynamic, short lived, and sequential. Immunity (2000) 13:323–32. doi:10.1016/S1074-7613(00)00032-7

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Faroudi M, Zaru R, Paulet P, Muller S, Valitutti S. T lymphocyte activation by repeated immunological synapse formation and intermittent signaling. J Immunol (2003) 171:1128–32. doi:10.4049/jimmunol.171.3.1128

CrossRef Full Text | Google Scholar

44. Gunzer M, Weishaupt C, Hillmer A, Basoglu Y, Friedl P, Dittmar KE, et al. A spectrum of biophysical interaction modes between T cells and different antigen-presenting cells during priming in 3-D collagen and in vivo. Blood (2004) 104:2801–9. doi:10.1182/blood-2004-03-1193

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Clark CE, Hasan M, Bousso P. A role for the immediate early gene product c-fos in imprinting T cells with short-term memory for signal summation. PLoS One (2011) 6:e18916. doi:10.1371/journal.pone.0018916

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Marangoni F, Murooka TT, Manzo T, Kim EY, Carrizosa E, Elpek NM, et al. The transcription factor NFAT exhibits signal memory during serial T cell interactions with antigen-presenting cells. Immunity (2013) 38:237–49. doi:10.1016/j.immuni.2012.09.012

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Klein J. Nature of alloreactivity. Frequency of responding cells. In: Klein J. Natural History of the Major Histocompatibility Complex. New York, NY: Wiley-Interscience (1986). p. 417–21.

Google Scholar

48. Noorchashm H, Lieu YK, Rostami SY, Song HK, Greeley SA, Bazel S, et al. A direct method for the calculation of alloreactive CD4+ T cell precursor frequency. Transplantation (1999) 67:1281–4. doi:10.1097/00007890-199905150-00015

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Suchin EJ, Langmuir PB, Palmer E, Sayegh MH, Wells AD, Turka LA. Quantifying the frequency of alloreactive T cells in vivo: new answers to an old question. J Immunol (2001) 166:973–81. doi:10.4049/jimmunol.166.2.973

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Sabatino JJ Jr, Huang J, Zhu C, Evavold BD. High prevalence of low affinity peptide-MHC II tetramer-negative effectors during polyclonal CD4+ T cell responses. J Exp Med (2011) 208:81–90. doi:10.1084/jem.20101574

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Schmidt J, Dojcinovic D, Guillaume P, Luescher I. Analysis, isolation, and activation of antigen-specific CD4+ and CD8+ T cells by soluble MHC-peptide complexes. Front Immunol (2013) 4:218. doi:10.3389/fimmu.2013.00218

CrossRef Full Text | Google Scholar

52. Martinez RJ, Evavold BD. Lower affinity T cells are critical components and active participants of the immune response. Front Immunol (2015) 6:468. doi:10.3389/fimmu.2015.00468

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Tungatt K, Bianchi V, Crowther MD, Powell WE, Schauenburg AJ, Trimby A, et al. Antibody stabilization of peptide-MHC multimers reveals functional T cells bearing extremely low-affinity TCRs. J Immunol (2015) 194:463–74. doi:10.4049/jimmunol.1401785

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Jenkins MK, Moon JJ. The role of naive T cell precursor frequency and recruitment in dictating immune response magnitude. J Immunol (2012) 188:4135–40. doi:10.4049/jimmunol.1102661

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Nelson RW, Beisang D, Tubo NJ, Dileepan T, Wiesner DL, Nielsen K, et al. T cell receptor cross-reactivity between similar foreign and self peptides influences naive cell population size and autoimmunity. Immunity (2015) 42:95–107. doi:10.1016/j.immuni.2014.12.022

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Stemberger C, Huster KM, Koffler M, Anderl F, Schiemann M, Wagner H, et al. A single naive CD8+ T cell precursor can develop into diverse effector and memory subsets. Immunity (2007) 27:985–97. doi:10.1016/j.immuni.2007.10.012

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Haring JS, Badovinac VP, Harty JT. Inflaming the CD8+ T cell response. Immunity (2006) 25:19–29. doi:10.1016/j.immuni.2006.07.001

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Buchholz VR, Flossdorf M, Hensel I, Kretschmer L, Weissbrich B, Graf P, et al. Disparate individual fates compose robust CD8+ T cell immunity. Science (2013) 340:630–5. doi:10.1126/science.1235454

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Gerlach C, Rohr JC, Perié L, van Rooij N, van Heijst JWJ, Velds A, et al. Heterogeneous differentiation patterns of individual CD8+ T cells. Science (2013) 340:635–9. doi:10.1126/science.1235487

PubMed Abstract | CrossRef Full Text | Google Scholar

60. La Gruta NL, Rothwell WT, Cukalac T, Swan NG, Valkenburg SA, Kedzierska K, et al. Primary CTL response magnitude in mice is determined by the extent of naive T cell recruitment and subsequent clonal expansion. J Clin Invest (2010) 120:1885–94. doi:10.1172/JCI41538

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Sant AJ, Chaves FA, Leddon SA, Tung J. The control of the specificity of CD4 T cell responses: thresholds, breakpoints, and ceilings. Front Immunol (2013) 4:340. doi:10.3389/fimmu.2013.00340

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Kim A, Sadegh-Nasseri S. Determinants of immunodominance for CD4 T cells. Curr Opin Immunol (2015) 34:9–15. doi:10.1016/j.coi.2014.12.005

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Mandl JN, Monteiro JP, Vrisekoop N, Germain RN. T cell-positive selection uses self-ligand binding strength to optimize repertoire recognition of foreign antigens. Immunity (2013) 38:263–74. doi:10.1016/j.immuni.2012.09.011

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Persaud SP, Parker CR, Lo WL, Weber KS, Allen PM. Intrinsic CD4+ T cell sensitivity and response to a pathogen are set and sustained by avidity for thymic and peripheral complexes of self peptide and MHC. Nat Immunol (2014) 15:266–74. doi:10.1038/ni.2822

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Fulton RB, Hamilton SE, Xing Y, Best JA, Goldrath AW, Hogquist KA, et al. The TCR’s sensitivity to self peptide-MHC dictates the ability of naive CD8+ T cells to respond to foreign antigens. Nat Immunol (2015) 16:107–17. doi:10.1038/ni.3043

CrossRef Full Text | Google Scholar

66. van Heijst JWJ, Gerlach C, Swart E, Sie D, Nunes-Alves C, Kerkhoven RM, et al. Recruitment of antigen-specific CD8+ T cells in response to infection is markedly efficient. Science (2009) 325:1265–9. doi:10.1126/science.1175455

PubMed Abstract | CrossRef Full Text | Google Scholar

67. Badovinac VP, Haring JS, Harty JT. Initial T cell receptor transgenic cell precursor frequency dictates critical aspects of the CD8+ T cell response to infection. Immunity (2007) 26:827–41. doi:10.1016/j.immuni.2007.04.013

CrossRef Full Text | Google Scholar

68. Tubo NJ, Pagan AJ, Taylor JJ, Nelson RW, Linehan JL, Ertelt JM, et al. Single naive CD4+ T cells from a diverse repertoire produce different effector cell types during infection. Cell (2013) 153:785–96. doi:10.1016/j.cell.2013.04.007

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Cauley LS, Cookenham T, Miller TB, Adams PS, Vignali KM, Vignali DA, et al. Virus-specific CD4+ memory T cells in nonlymphoid tissues express a highly activated phenotype. J Immunol (2002) 169:6655–8. doi:10.4049/jimmunol.169.12.6655

CrossRef Full Text | Google Scholar

70. Harrington LE, van der Most RG, Whitton JL, Ahmed R. Recombinant vaccinia virus-induced T-cell immunity: quantitation of the response to the virus vector and the foreign epitope. J Virol (2002) 76:3329–37. doi:10.1128/JVI.76.7.3329-3337.2002

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Homann D, Teyton L, Oldstone MBA. Differential regulation of antiviral T-cell immunity results in stable CD8+ but declining CD4+ T-cell memory. Nat Med (2001) 7:913–9. doi:10.1038/90950

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Foulds KE, Zenewicz LA, Shedlock DJ, Jiang J, Troy AE, Shen H. CD4 and CD8 T cells are intrinsically different in their proliferative responses. J Immunol (2002) 168:1528–32. doi:10.4049/jimmunol.168.4.1528

CrossRef Full Text | Google Scholar

73. Seder RA, Ahmed R. Similarities and differences in CD4+ and CD8+ effector and memory T cell generation. Nat Immunol (2003) 4:835–42. doi:10.1038/ni969

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Kündig TM, Shahinian A, Kawai K, Mittrücker HW, Sebzda E, Bachmann MF, et al. Duration of TCR stimulation determines costimulatory requirement of T cells. Immunity (1996) 5:41–52. doi:10.1016/S1074-7613(00)80308-8

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Laouar Y, Crispe IN. Functional flexibility in T cells: independent regulation of CD4+ T cell proliferation and effector function in vivo. Immunity (2000) 13:291–301. doi:10.1016/S1074-7613(00)00029-7

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Sprent J. The power of dilution: using adoptive transfer to study TCR transgenic T cells. J Immunol (2013) 191:5325–6. doi:10.4049/jimmunol.1302679

CrossRef Full Text | Google Scholar

77. Iezzi G, Karjalainen K, Lanzavecchia A. The duration of antigenic stimulation determines the fate of naive and effector T cells. Immunity (1998) 8:89–95. doi:10.1016/S1074-7613(00)80461-6

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Kaech SM, Ahmed R. Memory CD8+ T cell differentiation: initial antigen encounter triggers a developmental program in naive cells. Nat Immunol (2001) 2:415–22. doi:10.1038/87720

CrossRef Full Text | Google Scholar

79. van Stipdonk MJ, Lemmens EE, Schoenberger SP. Naive CTLs require a single brief period of antigenic stimulation for clonal expansion and differentiation. Nat Immunol (2001) 2:423–9. doi:10.1038/87730

CrossRef Full Text | Google Scholar

80. Wong P, Pamer EG. Antigen-independent CD8 T cell proliferation. J Immunol (2001) 166:5864–8. doi:10.4049/jimmunol.166.10.5864

CrossRef Full Text | Google Scholar

81. van Stipdonk MJ, Hardenberg G, Bijker MS, Lemmens EE, Droin NM, Green DR, et al. Dynamic programming of CD8+ T lymphocyte responses. Nat Immunol (2003) 4:361–5. doi:10.1038/ni912

PubMed Abstract | CrossRef Full Text | Google Scholar

82. van Stipdonk MJ, Sluijter M, Han WG, Offringa R. Development of CTL memory despite arrested clonal expansion. Eur J Immunol (2008) 38:1839–46. doi:10.1002/eji.200737974

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Spierings DC, Lemmens EE, Grewal K, Schoenberger SP, Green DR. Duration of CTL activation regulates IL-2 production required for autonomous clonal expansion. Eur J Immunol (2006) 36:1707–17. doi:10.1002/eji.200635929

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Malek TR. The biology of interleukin-2. Annu Rev Immunol (2008) 26:453–79. doi:10.1146/annurev.immunol.26.021607.090357

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Wong P, Pamer EG. Disparate in vitro and in vivo requirements for IL-2 during antigen-independent CD8 T cell expansion. J Immunol (2004) 172:2171–6. doi:10.4049/jimmunol.172.4.2171

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Bayer AL, Yu A, Malek TR. Function of the IL-2R for thymic and peripheral CD4+CD25+Foxp3+ T regulatory cells. J Immunol (2007) 178:4062–71. doi:10.4049/jimmunol.178.7.4062

PubMed Abstract | CrossRef Full Text | Google Scholar

87. D’Cruz LM, Klein L. Development and function of agonist-induced CD25+Foxp3+ regulatory T cells in the absence of interleukin 2 signaling. Nat Immunol (2005) 6:1152–9. doi:10.1038/ni1264

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Fontenot JD, Rasmussen JP, Gavin MA, Rudensky AY. A function for interleukin 2 in Foxp3-expressing regulatory T cells. Nat Immunol (2005) 6:1142–51. doi:10.1038/ni1263

PubMed Abstract | CrossRef Full Text | Google Scholar

89. O’Gorman WE, Dooms H, Thorne SH, Kuswanto WF, Simonds EF, Krutzik PO, et al. The initial phase of an immune response functions to activate regulatory T cells. J Immunol (2009) 183:332–9. doi:10.4049/jimmunol.0900691

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Williams MA, Tyznik AJ, Bevan MJ. Interleukin-2 signals during priming are required for secondary expansion of CD8+ memory T cells. Nature (2006) 441:890–3. doi:10.1038/nature04790

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Feau S, Arens R, Togher S, Schoenberger SP. Autocrine IL-2 is required for secondary population expansion of CD8+ memory T cells. Nat Immunol (2011) 12:908–13. doi:10.1038/ni.2079

CrossRef Full Text | Google Scholar

92. Kastenmüller W, Gasteiger G, Subramanian N, Sparwasser T, Busch DH, Belkaid Y, et al. Regulatory T cells selectively control CD8+ T cell effector pool size via IL-2 restriction. J Immunol (2011) 187:3186–97. doi:10.4049/jimmunol.1101649

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Marchingo JM, Kan A, Sutherland RM, Duffy KR, Wellard CJ, Belz GT, et al. Antigen affinity, costimulation, and cytokine inputs sum linearly to amplify T cell expansion. Science (2014) 346:1123–7. doi:10.1126/science.1260044

CrossRef Full Text | Google Scholar

94. Starbeck-Miller GR, Xue HH, Harty JT. IL-12 and type I interferon prolong the division of activated CD8 T cells by maintaining high-affinity IL-2 signaling in vivo. J Exp Med (2014) 211:105–20. doi:10.1084/jem.20130901

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Mercado R, Vijh S, Allen SE, Kerksiek K, Pilip IM, Pamer EG. Early programming of T cell populations responding to bacterial infection. J Immunol (2000) 165:6833–9. doi:10.4049/jimmunol.165.12.6833

PubMed Abstract | CrossRef Full Text | Google Scholar

96. Corbin GA, Harty JT. Duration of infection and antigen display have minimal influence on the kinetics of the CD4+ T cell response to Listeria monocytogenes infection. J Immunol (2004) 173:5679–87. doi:10.4049/jimmunol.173.9.5679

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Williams MA, Bevan MJ. Shortening the infectious period does not alter expansion of CD8 T cells but diminishes their capacity to differentiate into memory cells. J Immunol (2004) 173:6694–702. doi:10.4049/jimmunol.173.11.6694

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Busch DH, Pilip IM, Vijh S, Pamer EG. Coordinate regulation of complex T cell populations responding to bacterial infection. Immunity (1998) 8:353–62. doi:10.1016/S1074-7613(00)80540-3

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Badovinac VP, Porter BB, Harty JT. Programmed contraction of CD8+ T cells after infection. Nat Immunol (2002) 3:619–26. doi:10.1038/ni804

CrossRef Full Text | Google Scholar

100. Bevan MJ, Fink PJ. The CD8 response on autopilot. Nat Immunol (2001) 2:381–2. doi:10.1038/87676

CrossRef Full Text | Google Scholar

101. Blattman JN, Cheng LE, Greenberg PD. CD8+ T cell responses: it’s all downhill after their prime. Nat Immunol (2002) 3:601–2. doi:10.1038/ni0702-601

CrossRef Full Text | Google Scholar

102. Masopust D, Kaech SM, Wherry EJ, Ahmed R. The role of programming in memory T-cell development. Curr Opin Immunol (2004) 16:217–25. doi:10.1016/j.coi.2004.02.005

PubMed Abstract | CrossRef Full Text | Google Scholar

103. Arens R, Schoenberger SP. Plasticity in programming of effector and memory CD8 T-cell formation. Immunol Rev (2010) 235:190–205. doi:10.1111/j.0105-2896.2010.00899.x

PubMed Abstract | CrossRef Full Text | Google Scholar

104. Curtsinger JM, Johnson CM, Mescher MF. CD8 T cell clonal expansion and development of effector function require prolonged exposure to antigen, costimulation, and signal 3 cytokine. J Immunol (2003) 171:5165–71. doi:10.4049/jimmunol.171.10.5165

PubMed Abstract | CrossRef Full Text | Google Scholar

105. Storni T, Ruedl C, Renner WA, Bachmann MF. Innate immunity together with duration of antigen persistence regulate effector T cell induction. J Immunol (2003) 171:795–801. doi:10.4049/jimmunol.171.2.795

PubMed Abstract | CrossRef Full Text | Google Scholar

106. Tseng KE, Chung CY, H’Ng WS, Wang SL. Early infection termination affects number of CD8+ memory T cells and protective capacities in Listeria monocytogenes-infected mice upon rechallenge. J Immunol (2009) 182:4590–600. doi:10.4049/jimmunol.0801125

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Cockburn IA, Chen YC, Overstreet MG, Lees JR, van Rooijen N, Farber DL, et al. Prolonged antigen presentation is required for optimal CD8+ T cell responses against malaria liver stage parasites. PLoS Pathog (2010) 6:e1000877. doi:10.1371/journal.ppat.1000877

PubMed Abstract | CrossRef Full Text | Google Scholar

108. Hafalla JC, Sano G, Carvalho LH, Morrot A, Zavala F. Short-term antigen presentation and single clonal burst limit the magnitude of the CD8+ T cell responses to malaria liver stages. Proc Natl Acad Sci U S A (2002) 99:11819–24. doi:10.1073/pnas.182189999

CrossRef Full Text | Google Scholar

109. Catron DM, Rusch LK, Hataye J, Itano AA, Jenkins MK. CD4+ T cells that enter the draining lymph nodes after antigen injection participate in the primary response and become central-memory cells. J Exp Med (2006) 203:1045–54. doi:10.1084/jem.20051954

PubMed Abstract | CrossRef Full Text | Google Scholar

110. Quigley M, Huang X, Yang Y. Extent of stimulation controls the formation of memory CD8 T cells. J Immunol (2007) 179:5768–77. doi:10.4049/jimmunol.179.9.5768

PubMed Abstract | CrossRef Full Text | Google Scholar

111. Jung S, Unutmaz D, Wong P, Sano G, De los Santos K, Sparwasser T. In vivo depletion of CD11c+ dendritic cells abrogates priming of CD8+ T cells by exogenous cell-associated antigens. Immunity (2002) 17:211–20. doi:10.1016/S1074-7613(02)00365-5

PubMed Abstract | CrossRef Full Text | Google Scholar

112. Hochweller K, Striegler J, Hämmerling GJ, Garbi N. A novel CD11c.DTR transgenic mouse for depletion of dendritic cells reveals their requirement for homeostatic proliferation of natural killer cells. Eur J Immunol (2008) 38:2776–83. doi:10.1002/eji.200838659

PubMed Abstract | CrossRef Full Text | Google Scholar

113. Bernhard CA, Ried C, Kochanek S, Brocker T. CD169+ macrophages are sufficient for priming of CTLs with specificities left out by cross-priming dendritic cells. Proc Natl Acad Sci U S A (2015) 112:5461–6. doi:10.1073/pnas.1423356112

PubMed Abstract | CrossRef Full Text | Google Scholar

114. Moussion C, Girard JP. Dendritic cells control lymphocyte entry to lymph nodes through high endothelial venules. Nature (2011) 479:542–6. doi:10.1038/nature10540

PubMed Abstract | CrossRef Full Text | Google Scholar

115. Prlic M, Hernandez-Hoyos G, Bevan MJ. Duration of the initial TCR stimulus controls the magnitude but not functionality of the CD8+ T cell response. J Exp Med (2006) 203:2135–43. doi:10.1084/jem.20060928

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Tewari K, Walent J, Svaren J, Zamoyska R, Suresh M. Differential requirement for Lck during primary and memory CD8+ T cell responses. Proc Natl Acad Sci U S A (2006) 103:16388–93. doi:10.1073/pnas.0602565103

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Shaulov A, Murali-Krishna K. CD8 T cell expansion and memory differentiation are facilitated by simultaneous and sustained exposure to antigenic and inflammatory milieu. J Immunol (2008) 180:1131–8. doi:10.4049/jimmunol.180.2.1131

PubMed Abstract | CrossRef Full Text | Google Scholar

118. Blair DA, Turner DL, Bose TO, Pham QM, Bouchard KR, Williams KJ, et al. Duration of antigen availability influences the expansion and memory differentiation of T cells. J Immunol (2011) 187:2310–21. doi:10.4049/jimmunol.1100363

PubMed Abstract | CrossRef Full Text | Google Scholar

119. Kim TS, Gorski SA, Hahn S, Murphy KM, Braciale TJ. Distinct dendritic cell subsets dictate the fate decision between effector and memory CD8+ T cell differentiation by a CD24-dependent mechanism. Immunity (2014) 40:400–13. doi:10.1016/j.immuni.2014.02.004

CrossRef Full Text | Google Scholar

120. Yoon H, Legge KL, Sung SS, Braciale TJ. Sequential activation of CD8+ T cells in the draining lymph nodes in response to pulmonary virus infection. J Immunol (2007) 179:391–9. doi:10.4049/jimmunol.179.1.391

PubMed Abstract | CrossRef Full Text | Google Scholar

121. McGill J, Legge KL. Contribution of lung-resident T cell proliferation to the overall magnitude of the antigen-specific CD8 T cell response in the lungs following murine influenza virus infection. J Immunol (2009) 183:4177–81. doi:10.4049/jimmunol.0901109

CrossRef Full Text | Google Scholar

122. McGill J, Van Rooijen N, Legge KL. Protective influenza-specific CD8 T cell responses require interactions with dendritic cells in the lungs. J Exp Med (2008) 205:1635–46. doi:10.1084/jem.20080314

PubMed Abstract | CrossRef Full Text | Google Scholar

123. Dolfi DV, Duttagupta PA, Boesteanu AC, Mueller YM, Oliai CH, Borowski AB, et al. Dendritic cells and CD28 costimulation are required to sustain virus-specific CD8+ T cell responses during the effector phase in vivo. J Immunol (2011) 186:4599–608. doi:10.4049/jimmunol.1001972

PubMed Abstract | CrossRef Full Text | Google Scholar

124. León B, Ballesteros-Tato A, Randall TD, Lund FE. Prolonged antigen presentation by immune complex-binding dendritic cells programs the proliferative capacity of memory CD8 T cells. J Exp Med (2014) 211:1637–55. doi:10.1084/jem.20131692

PubMed Abstract | CrossRef Full Text | Google Scholar

125. Kang SS, Herz J, Kim JV, Nayak D, Stewart-Hutchinson P, Dustin ML, et al. Migration of cytotoxic lymphocytes in cell cycle permits local MHC I-dependent control of division at sites of viral infection. J Exp Med (2011) 208:747–59. doi:10.1084/jem.20101295

PubMed Abstract | CrossRef Full Text | Google Scholar

126. Schrum AG, Turka LA. The proliferative capacity of individual naive CD4+ T cells is amplified by prolonged T cell antigen receptor triggering. J Exp Med (2002) 196:793–803. doi:10.1084/jem.20020158

CrossRef Full Text | Google Scholar

127. Gett AV, Sallusto F, Lanzavecchia A, Geginat J. T cell fitness determined by signal strength. Nat Immunol (2003) 4:355–60. doi:10.1038/ni908

PubMed Abstract | CrossRef Full Text | Google Scholar

128. Schrum AG, Palmer E, Turka LA. Distinct temporal programming of naive CD4+ T cells for cell division versus TCR-dependent death susceptibility by antigen-presenting macrophages. Eur J Immunol (2005) 35:449–59. doi:10.1002/eji.200425635

PubMed Abstract | CrossRef Full Text | Google Scholar

129. Ravkov EV, Williams MA. The magnitude of CD4+ T cell recall responses is controlled by the duration of the secondary stimulus. J Immunol (2009) 183:2382–9. doi:10.4049/jimmunol.0900319

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Obst R, van Santen HM, Mathis D, Benoist C. Antigen persistence is required throughout the expansion phase of a CD4+ T cell response. J Exp Med (2005) 201:1555–65. doi:10.1084/jem.20042521

CrossRef Full Text | Google Scholar

131. Celli S, Garcia Z, Bousso P. CD4 T cells integrate signals delivered during successive DC encounters in vivo. J Exp Med (2005) 202:1271–8. doi:10.1084/jem.20051018

PubMed Abstract | CrossRef Full Text | Google Scholar

132. Celli S, Lemaitre F, Bousso P. Real-time manipulation of T cell-dendritic cell interactions in vivo reveals the importance of prolonged contacts for CD4+ T cell activation. Immunity (2007) 27:625–34. doi:10.1016/j.immuni.2007.08.018

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Jelley-Gibbs DM, Brown DM, Dibble JP, Haynes L, Eaton SM, Swain SL. Unexpected prolonged presentation of influenza antigens promotes CD4 T cell memory generation. J Exp Med (2005) 202:697–706. doi:10.1084/jem.20050227

PubMed Abstract | CrossRef Full Text | Google Scholar

134. Marzo AL, Klonowski KD, Le Bon A, Borrow P, Tough DF, Lefrancois L. Initial T cell frequency dictates memory CD8+ T cell lineage commitment. Nat Immunol (2005) 6:793–9. doi:10.1038/ni1227

PubMed Abstract | CrossRef Full Text | Google Scholar

135. Blair DA, Lefrancois L. Increased competition for antigen during priming negatively impacts the generation of memory CD4 T cells. Proc Natl Acad Sci U S A (2007) 104:15045–50. doi:10.1073/pnas.0703767104

PubMed Abstract | CrossRef Full Text | Google Scholar

136. Whitmire JK, Benning N, Eam B, Whitton JL. Increasing the CD4+ T cell precursor frequency leads to competition for IFN-γ thereby degrading memory cell quantity and quality. J Immunol (2008) 180:6777–85. doi:10.4049/jimmunol.180.10.6777

CrossRef Full Text | Google Scholar

137. Weaver JM, Chaves FA, Sant AJ. Abortive activation of CD4 T cell responses during competitive priming in vivo. Proc Natl Acad Sci U S A (2009) 106:8647–52. doi:10.1073/pnas.0811584106

PubMed Abstract | CrossRef Full Text | Google Scholar

138. Yarke CA, Dalheimer SL, Zhang N, Catron DM, Jenkins MK, Mueller DL. Proliferating CD4+ T cells undergo immediate growth arrest upon cessation of TCR signaling in vivo. J Immunol (2008) 180:156–62. doi:10.4049/jimmunol.180.1.156

PubMed Abstract | CrossRef Full Text | Google Scholar

139. Jusforgues-Saklani H, Uhl M, Blachere N, Lemaitre F, Lantz O, Bousso P, et al. Antigen persistence is required for dendritic cell licensing and CD8+ T cell cross-priming. J Immunol (2008) 181:3067–76. doi:10.4049/jimmunol.181.5.3067

PubMed Abstract | CrossRef Full Text | Google Scholar

140. Jelley-Gibbs DM, Lepak NM, Yen M, Swain SL. Two distinct stages in the transition from naive CD4 T cells to effectors, early antigen-dependent and late cytokine-driven expansion and differentiation. J Immunol (2000) 165:5017–26. doi:10.4049/jimmunol.165.9.5017

PubMed Abstract | CrossRef Full Text | Google Scholar

141. Bajénoff M, Wurtz O, Guerder S. Repeated antigen exposure is necessary for the differentiation, but not the initial proliferation, of naive CD4+ T cells. J Immunol (2002) 168:1723–9. doi:10.4049/jimmunol.168.4.1723

CrossRef Full Text | Google Scholar

142. Lee WT, Pasos G, Cecchini L, Mittler JN. Continued antigen stimulation is not required during CD4+ T cell clonal expansion. J Immunol (2002) 168:1682–9. doi:10.4049/jimmunol.168.4.1682

CrossRef Full Text | Google Scholar

143. Jelley-Gibbs DM, Dibble JP, Filipson S, Haynes L, Kemp RA, Swain SL. Repeated stimulation of CD4 effector T cells can limit their protective function. J Exp Med (2005) 201:1101–12. doi:10.1084/jem.20041852

PubMed Abstract | CrossRef Full Text | Google Scholar

144. Han S, Asoyan A, Rabenstein H, Nakano N, Obst R. Role of antigen persistence and dose for CD4+ T-cell exhaustion and recovery. Proc Natl Acad Sci U S A (2010) 107:20453–8. doi:10.1073/pnas.1008437107

PubMed Abstract | CrossRef Full Text | Google Scholar

145. McKinney EF, Lee JC, Jayne DRW, Lyons PA, Smith KGC. T-cell exhaustion, co-stimulation and clinical outcome in autoimmunity and infection. Nature (2015) 523:612–6. doi:10.1038/nature14468

PubMed Abstract | CrossRef Full Text | Google Scholar

146. Erman B, Alag AS, Dahle O, van Laethem F, Sarafova SD, Guinter TI, et al. Coreceptor signal strength regulates positive selection but does not determine CD4/CD8 lineage choice in a physiologic in vivo model. J Immunol (2006) 177:6613–25. doi:10.4049/jimmunol.177.10.6613

PubMed Abstract | CrossRef Full Text | Google Scholar

147. Van Laethem F, Tikhonova AN, Pobezinsky LA, Tai X, Kimura MY, Le Saout C, et al. Lck availability during thymic selection determines the recognition specificity of the T cell repertoire. Cell (2013) 154:1326–41. doi:10.1016/j.cell.2013.08.009

PubMed Abstract | CrossRef Full Text | Google Scholar

148. Yasutomo K, Doyle C, Miele L, Fuchs C, Germain RN. The duration of antigen receptor signalling determines CD4+ versus CD8+ T-cell lineage fate. Nature (2000) 404:506–10. doi:10.1038/35006664

PubMed Abstract | CrossRef Full Text | Google Scholar

149. Liu X, Bosselut R. Duration of TCR signaling controls CD4-CD8 lineage differentiation in vivo. Nat Immunol (2004) 5:280–8. doi:10.1038/ni1040

PubMed Abstract | CrossRef Full Text | Google Scholar

150. Saini M, Sinclair C, Marshall D, Tolaini M, Sakaguchi S, Seddon B. Regulation of Zap70 expression during thymocyte development enables temporal separation of CD4 and CD8 repertoire selection at different signaling thresholds. Sci Signal (2010) 3:ra23. doi:10.1126/scisignal.2000702

PubMed Abstract | CrossRef Full Text | Google Scholar

151. Mueller SN, Gebhardt T, Carbone FR, Heath WR. Memory T cell subsets, migration patterns, and tissue residence. Annu Rev Immunol (2013) 31:137–61. doi:10.1146/annurev-immunol-032712-095954

PubMed Abstract | CrossRef Full Text | Google Scholar

152. Mackay LK, Long HM, Brooks JM, Taylor GS, Leung CS, Chen A, et al. T cell detection of a B-cell tropic virus infection: newly-synthesised versus mature viral proteins as antigen sources for CD4 and CD8 epitope display. PLoS Pathog (2009) 5:e1000699. doi:10.1371/journal.ppat.1000699

PubMed Abstract | CrossRef Full Text | Google Scholar

153. Cella M, Engering A, Pinet V, Pieters J, Lanzavecchia A. Inflammatory stimuli induce accumulation of MHC class II complexes on dendritic cells. Nature (1997) 388:782–7. doi:10.1038/42030

PubMed Abstract | CrossRef Full Text | Google Scholar

154. Pierre P, Turley SJ, Gatti E, Hull M, Meltzer J, Mirza A, et al. Developmental regulation of MHC class II transport in mouse dendritic cells. Nature (1997) 388:787–92. doi:10.1038/42039

PubMed Abstract | CrossRef Full Text | Google Scholar

155. Shin JS, Ebersold M, Pypaert M, Delamarre L, Hartley A, Mellman I. Surface expression of MHC class II in dendritic cells is controlled by regulated ubiquitination. Nature (2006) 444:115–8. doi:10.1038/nature05261

PubMed Abstract | CrossRef Full Text | Google Scholar

156. van Niel G, Wubbolts R, ten Broeke T, Buschow SI, Ossendorp FA, Melief CJ, et al. Dendritic cells regulate exposure of MHC class II at their plasma membrane by oligoubiquitination. Immunity (2006) 25:885–94. doi:10.1016/j.immuni.2006.11.001

PubMed Abstract | CrossRef Full Text | Google Scholar

157. De Gassart A, Camosseto V, Thibodeau J, Ceppi M, Catalan N, Pierre P, et al. MHC class II stabilization at the surface of human dendritic cells is the result of maturation-dependent MARCH I down-regulation. Proc Natl Acad Sci U S A (2008) 105:3491–6. doi:10.1073/pnas.0708874105

PubMed Abstract | CrossRef Full Text | Google Scholar

158. Cho KJ, Walseng E, Ishido S, Roche PA. Ubiquitination by March-I prevents MHC class II recycling and promotes MHC class II turnover in antigen-presenting cells. Proc Natl Acad Sci U S A (2015) 112:10449–54. doi:10.1073/pnas.1507981112

PubMed Abstract | CrossRef Full Text | Google Scholar

159. Dudziak D, Kamphorst AO, Heidkamp GF, Buchholz VR, Trumpfheller C, Yamazaki S, et al. Differential antigen processing by dendritic cell subsets in vivo. Science (2007) 315:107–11. doi:10.1126/science.1136080

PubMed Abstract | CrossRef Full Text | Google Scholar

160. Vander Lugt B, Khan AA, Hackney JA, Agrawal S, Lesch J, Zhou M, et al. Transcriptional programming of dendritic cells for enhanced MHC class II antigen presentation. Nat Immunol (2014) 15:161–7. doi:10.1038/ni.2795

PubMed Abstract | CrossRef Full Text | Google Scholar

161. Eickhoff S, Brewitz A, Gerner MY, Klauschen F, Komander K, Hemmi H, et al. Robust anti-viral immunity requires multiple distinct T cell-dendritic cell interactions. Cell (2015) 162:1322–37. doi:10.1016/j.cell.2015.08.004

PubMed Abstract | CrossRef Full Text | Google Scholar

162. Hor JL, Whitney PG, Zaid A, Brooks AG, Heath WR, Mueller SN. Spatiotemporally distinct interactions with dendritic cell subsets facilitates CD4+ and CD8+ T cell activation to localized viral infection. Immunity (2015) 43:554–65. doi:10.1016/j.immuni.2015.07.020

CrossRef Full Text | Google Scholar

163. Levin SE, Zhang C, Kadlecek TA, Shokat KM, Weiss A. Inhibition of ZAP-70 kinase activity via an analog-sensitive allele blocks T cell receptor and CD28 superagonist signaling. J Biol Chem (2008) 283:15419–30. doi:10.1074/jbc.M709000200

PubMed Abstract | CrossRef Full Text | Google Scholar

164. Au-Yeung BB, Levin SE, Zhang C, Hsu LY, Cheng DA, Killeen N, et al. A genetically selective inhibitor demonstrates a function for the kinase Zap70 in regulatory T cells independent of its catalytic activity. Nat Immunol (2010) 11:1085–U95. doi:10.1038/ni.1955

PubMed Abstract | CrossRef Full Text | Google Scholar

165. Zikherman J, Parameswaran R, Weiss A. Endogenous antigen tunes the responsiveness of naive B cells but not T cells. Nature (2012) 489:160–U85. doi:10.1038/nature11311

PubMed Abstract | CrossRef Full Text | Google Scholar

166. Au-Yeung BB, Zikherman J, Mueller JL, Ashouri JF, Matloubian M, Cheng DA, et al. A sharp T-cell antigen receptor signaling threshold for T-cell proliferation. Proc Natl Acad Sci U S A (2014) 111:E3679–88. doi:10.1073/pnas.1413726111

PubMed Abstract | CrossRef Full Text | Google Scholar

167. Darzynkiewicz Z, Crissman H, Jacobberger JW. Cytometry of the cell cycle: cycling through history. Cytometry (2004) 58:21–32. doi:10.1002/cyto.a.20003

CrossRef Full Text | Google Scholar

168. Kim KH, Sederstrom JM. Assaying cell cycle status using flow cytometry. Curr Protoc Mol Biol (2015) 111:28.6.1–28.6.11. doi:10.1002/0471142727.mb2806s111

PubMed Abstract | CrossRef Full Text | Google Scholar

169. Darzynkiewicz Z, Zhao H, Zhang S, Lee MYWT, Lee EYC, Zhang Z. Initiation and termination of DNA replication during S phase in relation to cyclins D1, E and A, p21WAF1, Cdt1 and the p12 subunit of DNA polymerase δ revealed in individual cells by cytometry. Oncotarget (2015) 6:11735–50. doi:10.18632/oncotarget.4149

PubMed Abstract | CrossRef Full Text | Google Scholar

170. Flomerfelt FA, Gress RE. Analysis of cell proliferation and homeostasis using EdU labeling. Methods Mol Biol (2016) 1323:211–20. doi:10.1007/978-1-4939-2809-5_18

PubMed Abstract | CrossRef Full Text | Google Scholar

171. Gitlin AD, Shulman Z, Nussenzweig MC. Clonal selection in the germinal centre by regulated proliferation and hypermutation. Nature (2014) 509:637–40. doi:10.1038/nature13300

PubMed Abstract | CrossRef Full Text | Google Scholar

172. Gitlin AD, Mayer CT, Oliveira TY, Shulman Z, Jones MJ, Koren A, et al. T cell help controls the speed of the cell cycle in germinal center B cells. Science (2015) 349:643–6. doi:10.1126/science.aac4919

CrossRef Full Text | Google Scholar

173. Wilson A, Laurenti E, Oser G, van der Wath RC, Blanco-Bose W, Jaworski M, et al. Hematopoietic stem cells reversibly switch from dormancy to self-renewal during homeostasis and repair. Cell (2008) 135:1118–29. doi:10.1016/j.cell.2008.10.048

PubMed Abstract | CrossRef Full Text | Google Scholar

174. Sakaue-Sawano A, Kurokawa H, Morimura T, Hanyu A, Hama H, Osawa H, et al. Visualizing spatiotemporal dynamics of multicellular cell-cycle progression. Cell (2008) 132:487–98. doi:10.1016/j.cell.2007.12.033

PubMed Abstract | CrossRef Full Text | Google Scholar

175. Dowling MR, Kan A, Heinzel S, Zhou JH, Marchingo JM, Wellard CJ, et al. Stretched cell cycle model for proliferating lymphocytes. Proc Natl Acad Sci U S A (2014) 111:6377–82. doi:10.1073/pnas.1322420111

PubMed Abstract | CrossRef Full Text | Google Scholar

176. Kinjyo I, Qin J, Tan SY, Wellard CJ, Mrass P, Ritchie W, et al. Real-time tracking of cell cycle progression during CD8+ effector and memory T-cell differentiation. Nat Commun (2015) 6:6301. doi:10.1038/ncomms7301

PubMed Abstract | CrossRef Full Text | Google Scholar

177. Sicinska E, Aifantis I, Le Cam L, Swat W, Borowski C, Yu Q, et al. Requirement for cyclin D3 in lymphocyte development and T cell leukemias. Cancer Cell (2003) 4:451–61. doi:10.1016/S1535-6108(03)00301-5

PubMed Abstract | CrossRef Full Text | Google Scholar

178. Rowell EA, Walsh MC, Wells AD. Opposing roles for the cyclin-dependent kinase inhibitor p27kip1 in the control of CD4+ T cell proliferation and effector function. J Immunol (2005) 174:3359–68. doi:10.4049/jimmunol.174.6.3359

PubMed Abstract | CrossRef Full Text | Google Scholar

179. Jatzek A, Marie Tejera M, Plisch EH, Fero ML, Suresh M. T-cell intrinsic and extrinsic mechanisms of p27Kip1 in the regulation of CD8 T-cell memory. Immunol Cell Biol (2013) 91:120–9. doi:10.1038/icb.2012.71

PubMed Abstract | CrossRef Full Text | Google Scholar

180. Pareek TK, Lam E, Zheng X, Askew D, Kulkarni AB, Chance MR, et al. Cyclin-dependent kinase 5 activity is required for T cell activation and induction of experimental autoimmune encephalomyelitis. J Exp Med (2010) 207:2507–19. doi:10.1084/jem.20100876

PubMed Abstract | CrossRef Full Text | Google Scholar

181. Veiga-Fernandes H, Rocha B. High expression of active CDK6 in the cytoplasm of CD8 memory cells favors rapid division. Nat Immunol (2004) 5:31–7. doi:10.1038/ni1015

PubMed Abstract | CrossRef Full Text | Google Scholar

182. Zhang S, Rozell M, Verma RK, Albu DI, Califano D, VanValkenburgh J, et al. Antigen-specific clonal expansion and cytolytic effector function of CD8+ T lymphocytes depend on the transcription factor Bcl11b. J Exp Med (2010) 207:1687–99. doi:10.1084/jem.20092136

PubMed Abstract | CrossRef Full Text | Google Scholar

183. Xue L, Chiang L, He B, Zhao YY, Winoto A. FoxM1, a forkhead transcription factor is a master cell cycle regulator for mouse mature T cells but not double positive thymocytes. PLoS One (2010) 5:e9229. doi:10.1371/journal.pone.0009229

PubMed Abstract | CrossRef Full Text | Google Scholar

184. Karamitros D, Kotantaki P, Lygerou Z, Veiga-Fernandes H, Pachnis V, Kioussis D, et al. Differential geminin requirement for proliferation of thymocytes and mature T cells. J Immunol (2010) 184:2432–41. doi:10.4049/jimmunol.0901983

PubMed Abstract | CrossRef Full Text | Google Scholar

185. Martin E, Palmic N, Sanquer S, Lenoir C, Hauck F, Mongellaz C, et al. CTP synthase 1 deficiency in humans reveals its central role in lymphocyte proliferation. Nature (2014) 510:288–92. doi:10.1038/nature13386

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: clonal selection, T cells, primary immune response, expansion, differentiation, cell cycle

Citation: Obst R (2015) The timing of T cell priming and cycling. Front. Immunol. 6:563. doi: 10.3389/fimmu.2015.00563

Received: 27 September 2015; Accepted: 22 October 2015;
Published: 05 November 2015

Edited by:

Eric Huseby, University of Massachusetts Medical School, USA

Reviewed by:

George Kassiotis, National Institute for Medical Research, UK
Megan K. L. MacLeod, University of Glasgow, UK

Copyright: © 2015 Obst. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Reinhard Obst, reinhard.obst@med.uni-muenchen.de

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.