Pro-inflammatory human Th17 cells selectively express P-glycoprotein and are refractory to glucocorticoids

Radha Ramesh, Lina Kozhaya, Kelly McKevitt, Ivana M Djuretic, Thaddeus J Carlson, Maria A Quintero, Jacob L McCauley, Maria T Abreu, Derya Unutmaz, Mark S Sundrud, Radha Ramesh, Lina Kozhaya, Kelly McKevitt, Ivana M Djuretic, Thaddeus J Carlson, Maria A Quintero, Jacob L McCauley, Maria T Abreu, Derya Unutmaz, Mark S Sundrud

Abstract

IL-17A-expressing CD4(+) T cells (Th17 cells) are generally regarded as key effectors of autoimmune inflammation. However, not all Th17 cells are pro-inflammatory. Pathogenic Th17 cells that induce autoimmunity in mice are distinguished from nonpathogenic Th17 cells by a unique transcriptional signature, including high Il23r expression, and these cells require Il23r for their inflammatory function. In contrast, defining features of human pro-inflammatory Th17 cells are unknown. We show that pro-inflammatory human Th17 cells are restricted to a subset of CCR6(+)CXCR3(hi)CCR4(lo)CCR10(-)CD161(+) cells that transiently express c-Kit and stably express P-glycoprotein (P-gp)/multi-drug resistance type 1 (MDR1). In contrast to MDR1(-) Th1 or Th17 cells, MDR1(+) Th17 cells produce both Th17 (IL-17A, IL-17F, and IL-22) and Th1 (IFN-γ) cytokines upon TCR stimulation and do not express IL-10 or other anti-inflammatory molecules. These cells also display a transcriptional signature akin to pathogenic mouse Th17 cells and show heightened functional responses to IL-23 stimulation. In vivo, MDR1(+) Th17 cells are enriched and activated in the gut of Crohn's disease patients. Furthermore, MDR1(+) Th17 cells are refractory to several glucocorticoids used to treat clinical autoimmune disease. Thus, MDR1(+) Th17 cells may be important mediators of chronic inflammation, particularly in clinical settings of steroid resistant inflammatory disease.

Figures

Figure 1.
Figure 1.
Th17 cytokines and IL23R are independently regulated in human T cell subsets. (A) CD4+CD25− memory (CD45RO+) T cell subsets from healthy adult donor peripheral blood were analyzed by flow cytometry. CCR6+ or CCR6− cells were gated as CCR7hi (central memory; TCM), CCR7int (CCR7-intermediate), or CCR7lo (effector memory; TEM) cells, and CCR4 and CXCR3 expression was analyzed. Data represent >20 stains performed on individual donors or donor pools. (B) FACS-sorted TCM (CCR7hi) or TEM (CCR7lo) subsets were stimulated with PMA and ionomycin (P + I) and production of IFN-γ and IL-17A was determined by intracellular cytokine staining. Th1, CCR6−CCR4loCXCR3hi; Th2, CCR6−CCR4hiCXCR3lo; Th17, CCR6+CCR4hiCXCR3lo; Th17.1, CCR6+CCR4loCXCR3hi. Representative flow cytometry plots from 3 experiments performed on different pools of healthy adult donor blood are shown, and each donor pool contained blood from 2–4 individual donors. (C) IFN-γ and IL-17A production by FACS-sorted TCM or TEM cells was determined by intracellular cytokine staining as in B on 3 different pools of healthy adult donor peripheral blood. Each donor pool contained blood from 2–4 different donors. Individual and mean percentages of IL-17A+ (left), IFN-γ+ (middle), or IL-17A+/IFN-γ+ (right) T cells are shown, and data from each donor pool is color-coded. *, P < 0.05 by paired Student’s t test. (D) TEM Th1, Th2, Th17, or Th17.1 subsets were FACS-sorted as in B from two different pools of healthy donor peripheral blood. Both donor pools contained blood from 2–4 individual donors. Sorted cells were stimulated with anti-CD3/anti-CD28 for 72 h and expression of the indicated genes was measured by nanostring. Data are shown as fold change in gene expression within each donor pool; data from the two donor pools are color-coded. Horizontal bars in C and D represent the mean values. (E) Expression of pathogenic (red) or nonpathogenic (blue) murine Th17-signature genes (Lee et al., 2012) was determined in TEM Th17 and Th17.1 cells by nanostring as in D. Data are shown as mean relative (Log2 fold change) mRNA expression ± SD from 2 experiments on cells from different pools of healthy adult donor blood as in D. (F) TEM Th1, Th2, Th17, or Th17.1 cells (FACS-sorted as in B) were stimulated with anti-CD3/anti-CD28 and cultured for 6 d with or without IL-23. Cells were then restimulated with PMA and ionomycin and IFN-γ and IL-17A expression was determined by intracellular cytokine staining and FACS analysis. Data represent 3 experiments performed on independent donor pools, with each pool containing blood from 2–4 individual donors.
Figure 2.
Figure 2.
A novel subset of human Th17.1 cells is characterized by transient c-Kit and stable MDR1 expression. (A) TEM Th1, Th2, Th17, or Th17.1 cells were FACS-sorted from healthy adult donor peripheral blood as in Fig. 1 B. Sorted cells were stimulated with anti-CD3/anti-CD28 for 36 h and RNA was isolated for microarray analysis. Mean normalized raw gene expression values from two independent microarray experiments on cells sorted from different donor pools (each pool containing blood from 3–4 donors) were used to identify differentially expressed genes (1.8-fold cutoff). Data shown is a hierarchical clustering heatmap of all differentially expressed genes, with log2 transformation and row normalization. Red, high relative gene expression; dark blue, low relative gene expression. Representative gene symbols within each cluster are shown (a complete list of the genes and their absolute expression values within each cluster is provided in Table S1). (B) TEM Th1, Th2, Th17, or Th17.1 cells, FACS-sorted as in Fig. 1 B, were stimulated with anti-CD3/anti-CD28 for 72 h and expression of ABCB1 was determined by nanostring. Data are shown as ABCB1 mRNA expression (AU, arbitrary units) in two experiments performed on independent (color-coded) donor pools, with each pool containing blood from 2–4 individual donors. Horizontal bars represent the mean values. (C) Total CD4+CD25−CD45RO+ memory T cells isolated from healthy adult donor peripheral blood were labeled with rhodamine 123 (Rh123). After a 1-h efflux period at 37°C in the presence of vehicle (DMSO) or MDR1 inhibitors (CsA, cyclosporine A; Elacridar, selective MDR1 inhibitor), cells were stained with antibodies against CCR6, and Rh123 efflux and CCR6 expression was analyzed by FACS. Data shown are FACS plots from one experiment performed on cells from a healthy adult donor, and represent 3–4 independent experiments performed on cells isolated from different donors. (D) Rh123 efflux by CD4+ memory T cells was determined by FACS analysis as in C. After Rh123 efflux, cells were stained with antibodies against CCR6, CCR4, and CXCR3, and CCR4 and CXCR3 expression was analyzed on total CCR6− or CCR6+ cells, or on CCR6+ cells gated as Rh123lo (MDR1+) or Rh123hi (MDR1−). Data shown are representative FACS plots of >10 experiments performed on memory T cells isolated from independent donors. (E) CD4+ memory T cells were labeled with Rh123 and analyzed for Rh123 efflux as in C. After Rh123 efflux, cells were stained with antibodies against CD25, CCR6, CCR4, CXCR3, CCR7, c-Kit (CD117), and CD161. Cells were gated as TEM (CCR7lo) Th1 (CD25−CCR6−CCR4loCXCR3hi), Th2 (CD25−CCR6−CCR4hiCXCR3lo), Th17 (CD25−CCR6+CCR4hiCXCR3lo), Th17.1 (CD25−CCR6+CCR4loCXCR3hi), or T reg (CD25hi), and Rh123 efflux versus c-Kit (CD117; top) or CD161 (bottom) expression was analyzed in each subset. The FACS plots shown are representative of 5 independent experiments performed on memory T cells isolated from different donors. (F) The percentage of CD161+ cells within human TEM Th1, Th2, Th17, Th17.1 subsets and T reg cells was determined by FACS analysis as in E. Th17.1 cells were further gated into 3 subsets based on c-Kit expression and Rh123 efflux: c-Kit−MDR1−/Rh123hi, c-Kit−MDR1+/Rh123lo, and c-Kit+MDR1+/Rh123lo. Individual and mean percentages of CD161+ cells within each subset ± SD from 5 independent experiments performed on memory T cells isolated from individual (color-coded) donors is shown. (G) Naive CD4+ T cells were isolated from healthy adult donor peripheral blood, stimulated with anti-CD3/anti-CD28, and transduced with empty- (HDV) or RORC-containing (HDV.RORC) lentiviral particles that also contain a mouse HSA (heat stable antigen; a.k.a. CD24) expression cassette. Transduced T cells were expanded in IL-2–containing media for 7 d, and were then loaded with Rh123, incubated at 37°C for 1 h to allow for Rh123 efflux, and stained with antibodies against c-Kit or mouse HSA; Rh123 efflux and c-Kit expression was analyzed as a function of HSA expression in transduced T cells by FACS. FACS plots shown are representative of 3 independent experiments performed on naive T cells isolated from different donors. (H) CD4+ memory T cells isolated from healthy adult peripheral blood were FACS-sorted into CD25−CCR6+c-Kit+ (red) or CCR6− (blue) subsets. Cells were either left resting (no TCR), or were stimulated with anti-CD3/anti-CD28 and cultured for 6 d with or without IL-23. On day 6, cells were loaded with Rh123, stained with antibodies against c-Kit after Rh123 efflux, and analyzed by FACS. FACS plots shown are representative of 3 experiments using cells sorted from different donor pools, with each pool containing blood from 2–4 individual donors. (I) CD4+ memory T cells isolated from healthy adult peripheral blood were analyzed for Rh123 efflux as in E. CCR7 expression was assessed on c-Kit+ (red histogram) and c-Kit− (blue histogram) MDR1+/Rh123loCD25−CCR6+CCR4loCXCR3hi Th17.1 cells by FACS. The overlaid FACS histogram represents 4 experiments performed on memory T cells isolated from individual donors. CCR7 mean fluorescent intensity (MFI) is shown for c-Kit+ (red text) and c-Kit− (blue text) MDR1+ Th17.1 cells. (J) CCR7 MFI in c-Kit+ (red) and c-Kit− (blue) MDR1+ Th17.1 cells was determined by FACS analysis as in I. Data are shown as mean CCR7 MFI ± SD in c-Kit+ or c-Kit− MDR1+ Th17.1 cells from 4 individual donors. *, P < 0.05 by paired Student’s t test.
Figure 3.
Figure 3.
Unique pro-inflammatory characteristics of human MDR1+ Th17.1 cells. (A and B) Human CD4+ memory T cells from pooled healthy adult donor peripheral blood were FACS-sorted into TEM (CCR7lo) Th17.1 (CCR6+CCR4loCXCR3hi), Th17 (CCR6+CCR4hiCXCR3lo), or Th1 (CCR6−CCR4loCXCR3hi) cells. Th17.1 cells were sub-sorted into c-Kit+ or c-Kit− MDR1+ (Rh123lo), or c-Kit−MDR1− (Rh123hi) cells as indicated; Th17 and Th1 cells were sorted as c-Kit−MDR1− cells (see Fig. S1 for gating/sorting strategy). All cells were stimulated with anti-CD3/anti-CD28 for 72 h and expression of ABCB1 (MDR1), KIT (c-Kit), and KLRB1 (CD161; A), or IL17A, IL17F, IL22, CCL20, CSF2 (GM-CSF), and IFNG (B) was determined by nanostring. Data are shown as individual (color-coded) and mean normalized raw expression values (AU, arbitrary units) in cells sorted from 2–3 independent donor pools, with each pool containing blood from 2–4 individual donors. (C) Expression of pathogenic (red dots) or nonpathogenic (blue dots) murine Th17-signature genes (Lee et al., 2012) in TEM c-Kit+MDR1+, c-Kit−MDR1+, or c-Kit−MDR1− Th17.1 cells was determined by nanostring as in A and B. Mean normalized raw expression values from 3 independent experiments performed on different donor pools containing blood from 2–4 individual donors were used for fold change calculations. (D) Expression of IL23R and SOCS3 mRNA was determined in human TEM subsets by nanostring as in A and B. (E) FACS-sorted TEM subsets from healthy donor peripheral blood (as in A and B) were stimulated with anti-CD3/anti-CD28 and cultured for 3 d with or without IL-23. Stat3 phosphorylation (pY705) was determined on day 3 by phospho-intracellular staining and FACS analysis. Gray filled histograms, media alone; red traced histograms, plus IL-23. Representative FACS plots are shown and represent 3 independent experiments using cells isolated from different donor pools, with each pool containing blood from 2–4 individual donors. (F) Stat3 phosphorylation was determined by flow cytometry in TEM subsets cultured with or without IL-23 for 3 d as in E. Data are shown as mean fold change in Stat3 pY705 MFI (IL-23/media alone) ± SD from 3 experiments using cells isolated from different (color-coded) donor pools. *, P < 0.05 by paired Student’s t test. (G) FACS-sorted TEM subsets from healthy donor peripheral blood were stimulated with anti-CD3/anti-CD28 and cultured for 6 d with or without IL-23. On day 6, cells were restimulated with PMA and ionomycin and IL-17A and IFN-γ expression was determined by intracellular cytokine staining and FACS analysis. FACS plots shown are representative of 3 experiments using cells sorted from different donor pools, with each pool containing blood from 2–4 individual donors. Horizontal bars represent the mean values.
Figure 4.
Figure 4.
MDR1+ Th17.1 cells are enriched and activated in clinically inflamed tissue. (A) Mononuclear cells were isolated from CD patient peripheral blood (PBMC; left), uninvolved gut (middle), or involved gut (right), and were analyzed for expression of CD45RO and CCR7. Data shown are on CD3+CD4+CD25− gated T cells and represent 5 experiments on cells from different patients. (B) Mononuclear cells from CD patient PBMC, uninvolved gut, or involved gut tissue (as in A) were loaded with Rh123, stained with antibodies against CD3, CD4, CD25, CD45RO, c-Kit, and CD161 after a 1-h Rh123 efflux period at 37°C, and analyzed by FACS. Data shown are on CD3+CD4+CD25−CD45RO+ gated memory T cells. Rh123 efflux versus c-Kit (top) or CD161 (bottom) expression is shown. Data represent 3 (CD161 staining) or 4 (c-Kit staining) experiments on cells isolated from different patients. (C) Percentages of total memory (CD45RO+) cells (top left), MDR1+/Rh123lo memory cells (top right), c-Kit+MDR1+/Rh123lo memory T cells (bottom left), or CD161+MDR1+/Rh123lo memory cells (bottom right) were determined in CD patient PBMC, uninvolved gut, or involved gut tissue by FACS analysis as in B. Data are shown as mean percentages ± SD from 3–5 individual patients. *, P < 0.05; **, P < 0.01 by paired Student’s t test. (D) MDR1+ (Rh123lo) or MDR1− (Rh123hi) CD3+CD4+CD25−CD45RO+ memory T cells were FACS-sorted from the PBMC of one HC donor or from mononuclear cells isolated from the involved gut of one CD patient. Sorted cells were lysed directly ex vivo, and RNA was isolated for microarray analysis. Relative (fold change) expression of ABCB1 (MDR1), KIT (c-Kit), KLRB1 (CD161), or IL17A is shown for the T cell subsets as indicated. (E) Relative ex vivo expression (Log2 fold change) of pathogenic mouse Th17-signature genes (red; Lee et al., 2012), nonpathogenic Th17-signature genes (blue; Lee et al., 2012), or other notable (gray) genes was determined by microarray analysis of MDR1+ or MDR1− memory T cells sorted from involved CD patient gut tissue as in D. (F and G) Mononuclear cells isolated from involved CD patient gut tissue were FACS-sorted into MDR1+ or MDR1− CD3+CD4+CD25−CD45RO+ memory T cells, and expression of pathogenic (F) or nonpathogenic (G) mouse Th17-signature genes (Lee et al., 2012) was analyzed by nanostring. Data are shown as individual (color coded) and mean expression values (AU – arbitrary units) from 2 independent patients. Horizontal bars represent the mean values.
Figure 5.
Figure 5.
MDR1+ Th17.1 cells are refractory to glucocorticoid-mediated T cell suppression. (A) Total CD4+ T cells from healthy donor peripheral blood were labeled with Rh123, stained for CD45RO and c-Kit expression after Rh123 efflux, and analyzed for the frequency of c-Kit+ and c-Kit− MDR1+ (Rh123lo) memory T cells ex vivo (day 0) by flow cytometry. FACS plots show MDR1 (Rh123 efflux) activity versus CD45RO (left) or c-Kit (right) expression. Data represent 4 experiments on individual donors. (B) Total CD4+ T cells were stimulated with anti-CD3/anti-CD28 and were cultured for 5 d (top) or 12 d (bottom) in the presence of DMSO (vehicle), dexamethasone (Dex; 0.1 µM), prednisolone (Pred; 1 µM), or rapamycin (Rap; 0.1 µM). At days 5 and 12, cells were analyzed by FACS for Rh123 efflux and c-Kit expression as in A. FACS plots shown are representative of 4 experiments on individual donors. (C) Percentage of Rh123lo (MDR1+) cells within total CD4+ T cells cultured for 5 or 12 d with DMSO, Dex, Pred, or Rap was determined by FACS analysis as in B. Data are shown as individual (color coded) and mean percentages of Rh123lo cells from 4 independent donors. *, P < 0.05; **, P < 0.01 by paired Student’s t test. Horizontal bars represent the mean values. (D) Total CD4+ T cells were stimulated with anti-CD3/anti-CD28 and treated with titrating concentrations of dexamethasone (Dex; left) or prednisolone (Pred; right). At day 12, the frequency of Rh123lo cells was determined by Rh123 efflux and FACS analysis. Data are shown as mean percentages of Rh123lo (MDR1+) T cells ± SD from 4 independent experiments performed on cells from different donors. (E) Total CD4+ T cells were labeled with CellTrace violet, stimulated with anti-CD3/anti-CD28, and treated with DMSO, Dex, Pred, or Rap as in B. On day 5, cells were further labeled with Rh123 and were analyzed for Rh123 efflux and CellTrace violet dilution by FACS analysis. Overlaid histograms show CellTrace violet dilution in T cells gated as MDR1+/Rh123lo (red histogram) or MDR1−/Rh123hi (blue histogram); data represent 4 experiments performed on T cells isolated from individual donors. (F) Total CD4+ T cells were stimulated with anti-CD3/anti-CD28, treated with dexamethasone (Dex; 0.1 µM) or prednisolone (Pred; 1 µM), and cultured for 12–14 d. Cells were then loaded with Rh123, FACS-sorted into MDR1+/Rh123lo and MDR1−/Rh123hi cells, and RNA was isolated to determine IL23R gene expression by qPCR. Expression of IL23R was normalized to ACTB (β-actin). Data are shown as relative (fold change) IL23R expression in MDR1+ (Rh123lo) and MDR1− (Rh123hi) T cells sorted from Dex- or Pred-treated cultures. Individual (color-coded) and mean values are shown for 3 experiments performed on cells from different donors. *, P < 0.05 by paired Student’s t test. Horizontal bars represent the mean values. (G) FACS-sorted CCR6+MDR1+/Rh123lo, CCR6+MDR1−/Rh123hi, or CCR6−MDR1−/Rh123hi cells were stimulated with anti-CD3/anti-CD28, treated with DMSO, dexamethasone (Dex; 0.1 µM), prednisolone (Pred; 1 µM), or rapamycin (Rap; 0.1 µM), and cultured for 5 d. Cells were restimulated with PMA and ionomycin, and IFN-γ and IL-10 expression was determined by intracellular staining and flow cytometry. FACS plots are representative of 4 experiments using cells sorted from individual donors. (H) Total CD4+ T cells were stimulated with anti-CD3/anti-CD28, treated with DMSO, dexamethasone (Dex; 0.1 µM), or prednisolone (Pred; 1 µM), in the absence (DMSO) or presence of elacridar (0.1 µM). Cells were cultured for 12 d and were then analyzed for Rh123 efflux by FACS as in A, B, and D. Data are shown as percentage of MDR1+/Rh123lo T cells ± SD from 3 independent experiments using cells from different donors. (I) FACS-sorted MDR1− or MDR1+ (CCR6+) cells were stimulated with anti-CD3/anti-CD28 and cultured in the absence (DMSO) or presence of dexamethasone (Dex; 0.1 µM), or prednisolone (Pred; 1 µM) plus vehicle (DMSO; D) or 0.1 µM elacridar (E) for 5 d. Cells were restimulated with PMA and ionomycin, and IL-10 expression was determined by intracellular cytokine staining and FACS analysis. Data are shown as mean percentages of IL-10–expressing cells ± SD from 3 donors. *, P < 0.05 by paired Student’s t test.

References

    1. Acosta-Rodriguez E.V., Rivino L., Geginat J., Jarrossay D., Gattorno M., Lanzavecchia A., Sallusto F., Napolitani G. 2007. Surface phenotype and antigenic specificity of human interleukin 17-producing T helper memory cells. Nat. Immunol. 8:639–646 10.1038/ni1467
    1. Afzali B., Mitchell P.J., Edozie F.C., Povoleri G.A., Dowson S.E., Demandt L., Walter G., Canavan J.B., Scotta C., Menon B., et al. 2013. CD161 expression characterizes a subpopulation of human regulatory T cells that produces IL-17 in a STAT3-dependent manner. Eur. J. Immunol. 43:2043–2054 10.1002/eji.201243296
    1. Ahern P.P., Schiering C., Buonocore S., McGeachy M.J., Cua D.J., Maloy K.J., Powrie F. 2010. Interleukin-23 drives intestinal inflammation through direct activity on T cells. Immunity. 33:279–288 10.1016/j.immuni.2010.08.010
    1. Alcorn J.F., Crowe C.R., Kolls J.K. 2010. TH17 cells in asthma and COPD. Annu. Rev. Physiol. 72:495–516 10.1146/annurev-physiol-021909-135926
    1. Aller S.G., Yu J., Ward A., Weng Y., Chittaboina S., Zhuo R., Harrell P.M., Trinh Y.T., Zhang Q., Urbatsch I.L., Chang G. 2009. Structure of P-glycoprotein reveals a molecular basis for poly-specific drug binding. Science. 323:1718–1722 10.1126/science.1168750
    1. Ano S., Morishima Y., Ishii Y., Yoh K., Yageta Y., Ohtsuka S., Matsuyama M., Kawaguchi M., Takahashi S., Hizawa N. 2013. Transcription factors GATA-3 and RORγt are important for determining the phenotype of allergic airway inflammation in a murine model of asthma. J. Immunol. 190:1056–1065 10.4049/jimmunol.1202386
    1. Araki K., Ellebedy A.H., Ahmed R. 2011. TOR in the immune system. Curr. Opin. Cell Biol. 23:707–715 10.1016/j.ceb.2011.08.006
    1. Barnes P.J., Adcock I.M. 2009. Glucocorticoid resistance in inflammatory diseases. Lancet. 373:1905–1917 10.1016/S0140-6736(09)60326-3
    1. Barrat F.J., Cua D.J., Boonstra A., Richards D.F., Crain C., Savelkoul H.F., de Waal-Malefyt R., Coffman R.L., Hawrylowicz C.M., O’Garra A. 2002. In vitro generation of interleukin 10-producing regulatory CD4+ T cells is induced by immunosuppressive drugs and inhibited by T helper type 1 (Th1)- and Th2-inducing cytokines. J. Exp. Med. 195:603–616 10.1084/jem.20011629
    1. Chaudhary P.M., Roninson I.B. 1991. Expression and activity of P-glycoprotein, a multidrug efflux pump, in human hematopoietic stem cells. Cell. 66:85–94 10.1016/0092-8674(91)90141-K
    1. Chen Z., Laurence A., Kanno Y., Pacher-Zavisin M., Zhu B.M., Tato C., Yoshimura A., Hennighausen L., O’Shea J.J. 2006. Selective regulatory function of Socs3 in the formation of IL-17-secreting T cells. Proc. Natl. Acad. Sci. USA. 103:8137–8142 10.1073/pnas.0600666103
    1. Codarri L., Gyülvészi G., Tosevski V., Hesske L., Fontana A., Magnenat L., Suter T., Becher B. 2011. RORγt drives production of the cytokine GM-CSF in helper T cells, which is essential for the effector phase of autoimmune neuroinflammation. Nat. Immunol. 12:560–567 10.1038/ni.2027
    1. Cohen C.J., Crome S.Q., MacDonald K.G., Dai E.L., Mager D.L., Levings M.K. 2011. Human Th1 and Th17 cells exhibit epigenetic stability at signature cytokine and transcription factor loci. J. Immunol. 187:5615–5626 10.4049/jimmunol.1101058
    1. Cosmi L., De Palma R., Santarlasci V., Maggi L., Capone M., Frosali F., Rodolico G., Querci V., Abbate G., Angeli R., et al. 2008. Human interleukin 17–producing cells originate from a CD161+CD4+ T cell precursor. J. Exp. Med. 205:1903–1916 10.1084/jem.20080397
    1. Cosmi L., Maggi L., Santarlasci V., Capone M., Cardilicchia E., Frosali F., Querci V., Angeli R., Matucci A., Fambrini M., et al. 2010. Identification of a novel subset of human circulating memory CD4(+) T cells that produce both IL-17A and IL-4. J. Allergy Clin. Immunol. 125:222–230: e1–e4 10.1016/j.jaci.2009.10.012
    1. Cosmi L., Maggi L., Santarlasci V., Liotta F., Annunziato F. 2013. T helper cells plasticity in inflammation. Cytometry A.:n/a 10.1002/cyto.a.22348
    1. Crowe A., Tan A.M. 2012. Oral and inhaled corticosteroids: differences in P-glycoprotein (ABCB1) mediated efflux. Toxicol. Appl. Pharmacol. 260:294–302 10.1016/j.taap.2012.03.008
    1. Croxford A.L., Mair F., Becher B. 2012. IL-23: one cytokine in control of autoimmunity. Eur. J. Immunol. 42:2263–2273 10.1002/eji.201242598
    1. Duhen T., Geiger R., Jarrossay D., Lanzavecchia A., Sallusto F. 2009. Production of interleukin 22 but not interleukin 17 by a subset of human skin-homing memory T cells. Nat. Immunol. 10:857–863 10.1038/ni.1767
    1. Esplugues E., Huber S., Gagliani N., Hauser A.E., Town T., Wan Y.Y., O’Connor W., Jr, Rongvaux A., Van Rooijen N., Haberman A.M., et al. 2011. Control of TH17 cells occurs in the small intestine. Nature. 475:514–518 10.1038/nature10228
    1. Eyerich S., Eyerich K., Pennino D., Carbone T., Nasorri F., Pallotta S., Cianfarani F., Odorisio T., Traidl-Hoffmann C., Behrendt H., et al. 2009. Th22 cells represent a distinct human T cell subset involved in epidermal immunity and remodeling. J. Clin. Invest. 119:3573–3585
    1. Flammer J.R., Rogatsky I. 2011. Minireview: Glucocorticoids in autoimmunity: unexpected targets and mechanisms. Mol. Endocrinol. 25:1075–1086 10.1210/me.2011-0068
    1. Ghoreschi K., Laurence A., Yang X.P., Tato C.M., McGeachy M.J., Konkel J.E., Ramos H.L., Wei L., Davidson T.S., Bouladoux N., et al. 2010. Generation of pathogenic T(H)17 cells in the absence of TGF-β signalling. Nature. 467:967–971 10.1038/nature09447
    1. Gottesman M.M., Fojo T., Bates S.E. 2002. Multidrug resistance in cancer: role of ATP-dependent transporters. Nat. Rev. Cancer. 2:48–58 10.1038/nrc706
    1. Hirota K., Duarte J.H., Veldhoen M., Hornsby E., Li Y., Cua D.J., Ahlfors H., Wilhelm C., Tolaini M., Menzel U., et al. 2011. Fate mapping of IL-17-producing T cells in inflammatory responses. Nat. Immunol. 12:255–263 10.1038/ni.1993
    1. Hueber W., Sands B.E., Lewitzky S., Vandemeulebroecke M., Reinisch W., Higgins P.D., Wehkamp J., Feagan B.G., Yao M.D., Karczewski M., et al. , for the Secukinumab in Crohn’s Disease Study Group 2012. Secukinumab, a human anti-IL-17A monoclonal antibody, for moderate to severe Crohn’s disease: unexpected results of a randomised, double-blind placebo-controlled trial. Gut. 61:1693–1700 10.1136/gutjnl-2011-301668
    1. Hyafil F., Vergely C., Du Vignaud P., Grand-Perret T. 1993. In vitro and in vivo reversal of multidrug resistance by GF120918, an acridonecarboxamide derivative. Cancer Res. 53:4595–4602
    1. Josefowicz S.Z., Lu L.F., Rudensky A.Y. 2012. Regulatory T cells: mechanisms of differentiation and function. Annu. Rev. Immunol. 30:531–564 10.1146/annurev.immunol.25.022106.141623
    1. Kebir H., Ifergan I., Alvarez J.I., Bernard M., Poirier J., Arbour N., Duquette P., Prat A. 2009. Preferential recruitment of interferon-gamma-expressing TH17 cells in multiple sclerosis. Ann. Neurol. 66:390–402 10.1002/ana.21748
    1. Kleinschek M.A., Boniface K., Sadekova S., Grein J., Murphy E.E., Turner S.P., Raskin L., Desai B., Faubion W.A., de Waal Malefyt R., et al. 2009. Circulating and gut-resident human Th17 cells express CD161 and promote intestinal inflammation. J. Exp. Med. 206:525–534 10.1084/jem.20081712
    1. Kryczek I., Zhao E., Liu Y., Wang Y., Vatan L., Szeliga W., Moyer J., Klimczak A., Lange A., Zou W. 2011. Human TH17 cells are long-lived effector memory cells. Sci. Transl. Med. 3:104ra100
    1. Lee Y., Awasthi A., Yosef N., Quintana F.J., Xiao S., Peters A., Wu C., Kleinewietfeld M., Kunder S., Hafler D.A., et al. 2012. Induction and molecular signature of pathogenic TH17 cells. Nat. Immunol. 13:991–999 10.1038/ni.2416
    1. Ludescher C., Thaler J., Drach D., Drach J., Spitaler M., Gattringer C., Huber H., Hofmann J. 1992. Detection of activity of P-glycoprotein in human tumour samples using rhodamine 123. Br. J. Haematol. 82:161–168 10.1111/j.1365-2141.1992.tb04608.x
    1. Maggi L., Santarlasci V., Capone M., Rossi M.C., Querci V., Mazzoni A., Cimaz R., De Palma R., Liotta F., Maggi E., et al. 2012. Distinctive features of classic and nonclassic (Th17 derived) human Th1 cells. Eur. J. Immunol. 42:3180–3188 10.1002/eji.201242648
    1. Manel N., Unutmaz D., Littman D.R. 2008. The differentiation of human T(H)-17 cells requires transforming growth factor-beta and induction of the nuclear receptor RORgammat. Nat. Immunol. 9:641–649 10.1038/ni.1610
    1. McGeachy M.J., Bak-Jensen K.S., Chen Y., Tato C.M., Blumenschein W., McClanahan T., Cua D.J. 2007. TGF-beta and IL-6 drive the production of IL-17 and IL-10 by T cells and restrain T(H)-17 cell-mediated pathology. Nat. Immunol. 8:1390–1397 10.1038/ni1539
    1. McGeachy M.J., Chen Y., Tato C.M., Laurence A., Joyce-Shaikh B., Blumenschein W.M., McClanahan T.K., O’Shea J.J., Cua D.J. 2009. The interleukin 23 receptor is essential for the terminal differentiation of interleukin 17-producing effector T helper cells in vivo. Nat. Immunol. 10:314–324 10.1038/ni.1698
    1. McKinley L., Alcorn J.F., Peterson A., Dupont R.B., Kapadia S., Logar A., Henry A., Irvin C.G., Piganelli J.D., Ray A., Kolls J.K. 2008. TH17 cells mediate steroid-resistant airway inflammation and airway hyperresponsiveness in mice. J. Immunol. 181:4089–4097
    1. Miossec P., Korn T., Kuchroo V.K. 2009. Interleukin-17 and type 17 helper T cells. N. Engl. J. Med. 361:888–898 10.1056/NEJMra0707449
    1. Muranski P., Borman Z.A., Kerkar S.P., Klebanoff C.A., Ji Y., Sanchez-Perez L., Sukumar M., Reger R.N., Yu Z., Kern S.J., et al. 2011. Th17 cells are long lived and retain a stem cell-like molecular signature. Immunity. 35:972–985 10.1016/j.immuni.2011.09.019
    1. Mutalithas K., Guillen C., Day C., Brightling C.E., Pavord I.D., Wardlaw A.J. 2010. CRTH2 expression on T cells in asthma. Clin. Exp. Immunol. 161:34–40
    1. Nistala K., Adams S., Cambrook H., Ursu S., Olivito B., de Jager W., Evans J.G., Cimaz R., Bajaj-Elliott M., Wedderburn L.R. 2010. Th17 plasticity in human autoimmune arthritis is driven by the inflammatory environment. Proc. Natl. Acad. Sci. USA. 107:14751–14756 10.1073/pnas.1003852107
    1. Rivino L., Messi M., Jarrossay D., Lanzavecchia A., Sallusto F., Geginat J. 2004. Chemokine receptor expression identifies Pre-T helper (Th)1, Pre-Th2, and nonpolarized cells among human CD4+ central memory T cells. J. Exp. Med. 200:725–735 10.1084/jem.20040774
    1. Roberts M.M., Swart B.W., Simmons P.J., Basser R.L., Begley C.G., To L.B. 1999. Prolonged release and c-kit expression of haemopoietic precursor cells mobilized by stem cell factor and granulocyte colony stimulating factor. Br. J. Haematol. 104:778–784 10.1046/j.1365-2141.1999.01231.x
    1. Sallusto F., Lenig D., Förster R., Lipp M., Lanzavecchia A. 1999. Two subsets of memory T lymphocytes with distinct homing potentials and effector functions. Nature. 401:708–712 10.1038/44385
    1. Sallusto F., Zielinski C.E., Lanzavecchia A. 2012. Human Th17 subsets. Eur. J. Immunol. 42:2215–2220 10.1002/eji.201242741
    1. Schinkel A.H. 1997. The physiological function of drug-transporting P-glycoproteins. Semin. Cancer Biol. 8:161–170 10.1006/scbi.1997.0068
    1. Simmons P.J., Leavesley D.I., Levesque J.P., Swart B.W., Haylock D.N., To L.B., Ashman L.K., Juttner C.A. 1994. The mobilization of primitive hemopoietic progenitors into the peripheral blood. Stem Cells. 12:187–201, discussion :201–202
    1. Sincock P.M., Ashman L.K. 1997. Expression of c-Kit and functional drug efflux are correlated in de novo acute myeloid leukaemia. Leukemia. 11:1850–1857 10.1038/sj.leu.2400823
    1. So T., Lee S.W., Croft M. 2008. Immune regulation and control of regulatory T cells by OX40 and 4-1BB. Cytokine Growth Factor Rev. 19:253–262 10.1016/j.cytogfr.2008.04.003
    1. Strouse J.J., Ivnitski-Steele I., Waller A., Young S.M., Perez D., Evangelisti A.M., Ursu O., Bologa C.G., Carter M.B., Salas V.M., et al. 2013. Fluorescent substrates for flow cytometric evaluation of efflux inhibition in ABCB1, ABCC1, and ABCG2 transporters. Anal. Biochem. 437:77–87 10.1016/j.ab.2013.02.018
    1. Sundrud M.S., Grill S.M., Ni D., Nagata K., Alkan S.S., Subramaniam A., Unutmaz D. 2003. Genetic reprogramming of primary human T cells reveals functional plasticity in Th cell differentiation. J. Immunol. 171:3542–3549
    1. Sundrud M.S., Koralov S.B., Feuerer M., Calado D.P., Kozhaya A.E., Rhule-Smith A., Lefebvre R.E., Unutmaz D., Mazitschek R., Waldner H., et al. 2009. Halofuginone inhibits TH17 cell differentiation by activating the amino acid starvation response. Science. 324:1334–1338 10.1126/science.1172638
    1. Trifari S., Kaplan C.D., Tran E.H., Crellin N.K., Spits H. 2009. Identification of a human helper T cell population that has abundant production of interleukin 22 and is distinct from T(H)-17, T(H)1 and T(H)2 cells. Nat. Immunol. 10:864–871 10.1038/ni.1770
    1. Turtle C.J., Swanson H.M., Fujii N., Estey E.H., Riddell S.R. 2009. A distinct subset of self-renewing human memory CD8+ T cells survives cytotoxic chemotherapy. Immunity. 31:834–844 10.1016/j.immuni.2009.09.015
    1. Wan Q., Kozhaya L., ElHed A., Ramesh R., Carlson T.J., Djuretic I.M., Sundrud M.S., Unutmaz D. 2011. Cytokine signals through PI-3 kinase pathway modulate Th17 cytokine production by CCR6+ human memory T cells. J. Exp. Med. 208:1875–1887 10.1084/jem.20102516
    1. Weiss J., Haefeli W.E. 2010. Impact of ATP-binding cassette transporters on human immunodeficiency virus therapy. Int Rev Cell Mol Biol. 280:219–279 10.1016/S1937-6448(10)80005-X
    1. Zorzi F., Monteleone I., Sarra M., Calabrese E., Marafini I., Cretella M., Sedda S., Biancone L., Pallone F., Monteleone G. 2013. Distinct profiles of effector cytokines mark the different phases of Crohn’s disease. PLoS ONE. 8:e54562 10.1371/journal.pone.0054562

Source: PubMed

3
Předplatit