Skip to main page content
U.S. flag

An official website of the United States government

Dot gov

The .gov means it’s official.
Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

Https

The site is secure.
The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Clinical Trial
. 2008 Dec 2;6(12):2853-68.
doi: 10.1371/journal.pbio.0060301.

Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor

Affiliations
Clinical Trial

Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor

Jean-Philippe Coppé et al. PLoS Biol. .

Abstract

Cellular senescence suppresses cancer by arresting cell proliferation, essentially permanently, in response to oncogenic stimuli, including genotoxic stress. We modified the use of antibody arrays to provide a quantitative assessment of factors secreted by senescent cells. We show that human cells induced to senesce by genotoxic stress secrete myriad factors associated with inflammation and malignancy. This senescence-associated secretory phenotype (SASP) developed slowly over several days and only after DNA damage of sufficient magnitude to induce senescence. Remarkably similar SASPs developed in normal fibroblasts, normal epithelial cells, and epithelial tumor cells after genotoxic stress in culture, and in epithelial tumor cells in vivo after treatment of prostate cancer patients with DNA-damaging chemotherapy. In cultured premalignant epithelial cells, SASPs induced an epithelial-mesenchyme transition and invasiveness, hallmarks of malignancy, by a paracrine mechanism that depended largely on the SASP factors interleukin (IL)-6 and IL-8. Strikingly, two manipulations markedly amplified, and accelerated development of, the SASPs: oncogenic RAS expression, which causes genotoxic stress and senescence in normal cells, and functional loss of the p53 tumor suppressor protein. Both loss of p53 and gain of oncogenic RAS also exacerbated the promalignant paracrine activities of the SASPs. Our findings define a central feature of genotoxic stress-induced senescence. Moreover, they suggest a cell-nonautonomous mechanism by which p53 can restrain, and oncogenic RAS can promote, the development of age-related cancer by altering the tissue microenvironment.

PubMed Disclaimer

Conflict of interest statement

Competing interests. The authors have declared that no competing interests exist.

Figures

Figure 1
Figure 1. SASP of Human Fibroblasts
(A) Soluble factors secreted by the indicated cells were detected by antibody arrays and analyzed as described in the text, Materials and Methods; Text S2; and Datasets S1–S4. For each cell strain, the PRE and SEN signals were averaged and used as the baseline. Senescence (SEN) was induced by either repeatedly passaging the cells (REP, replicative exhaustion) or by exposing them to a relatively high dose (10 Gy) of ionizing radiation (XRA, X-irradiation); for simplification, XRA and REP signals were averaged as a single SEN signal (see also Figure 1C). Signals higher than baseline are displayed in yellow; signals below baseline are displayed in blue. The numbers on the heat map key (right) indicates log2-fold changes from the baseline. (B) Validation by immunostaining. PRE WI-38 cells were made senescent by REP or XRA, maintained in 10% serum until fixation, and immunostained for the cytokines IL-6 and IL-8 and the senescence marker p16INK4a. (C) Correlation between the SASPs of WI-38 cells induced to senescence by XRA or REP, compared to PRE cells, depicted as log2-fold changes. (D) WI-38 cells were X-irradiated at the indicated doses. CM was collected 2, 4, 7, or 10 d later, and soluble factors were analyzed by antibody arrays. PRE cells and cells irradiated with 0.5 Gy transiently arrested growth, but resumed proliferation 24–48 h after CM collection. Cells irradiated with 10 Gy became senescent, and therefore did proliferate during and beyond the course of the experiment. PRE and SEN (10 d) signals were averaged and used as the baseline. Signals higher than baseline are displayed in yellow; signals below baseline are displayed in blue. (E) PRE and SEN cells were analyzed for the indicated mRNAs by quantitative RT-PCR, and normalized to the corresponding PRE values (baselines). The senescence inducers (REP or XRA) are given in parentheses. Signals higher than baselines are shown in red, signals below baselines are displayed in green, and the fold changes in signals are given to the right of the heat map. The results are averages for four cell strains (WI-38, IMR90, HCA-2, and BJ), and the total number of samples for each condition are given below the heat map. An asterisk (*) indicates non-SASP factors (see Figure S4). (F) Correlation between mRNA and secreted protein levels for SASP (red) and non-SASP (blue) factors (see Figure S4). PRE and SEN values were averaged to create a baseline; all values were expressed as the log2-fold change relative to this baseline; PRE versus baseline and SEN versus baseline are shown.
Figure 2
Figure 2. SASP of Human Epithelial Cells
(A) Soluble factors secreted by the indicated normal cell type (epithelial vs. stromal) were detected by antibody arrays and analyzed as described in the text, Materials and Methods, and Datasets S5–S8. Normal prostate epithelial cells (PrECs) were induced to senesce by 10 Gy irradiation, and CM from the PRE and SEN cells were analyzed. The SASP of PrECs was compared side by side to the SASP of SEN(XRA) fibroblasts (WI-38, IMR-90, HCA-2, and BJ). The PRE values for each cell type served as the baseline. Signals higher than baseline are displayed in yellow; signals below baseline are displayed in blue. The heat map key (right) indicates log2-fold changes from the baseline. p-Values were calculated by the Student t-test, and are given to the right of the heat map. ns = not significant (p > 0.05) and defines non-SASP factors. (B) The log2-fold changes for all 120 proteins detected by the antibody arrays were plotted for SEN(XRA) PrECs and SEN(XRA) fibroblasts, relative to their PRE baseline. Seventy-nine secreted factors (66%) followed the same regulatory trend (in red). The remaining factors were not coregulated (depicted in blue). (C) Soluble factors produced by the indicated normal or transformed prostate epithelial cells were analyzed by antibody arrays and the results displayed as described for Figure 1A. For each cell strain or cell line, PRE and SEN signals were averaged and used as the baseline. Signals higher than the baseline are shown in yellow; signals below baseline are displayed in blue. An asterisk (*) indicates SASP factors that are conserved between all fibroblasts (Figure 1A) and all epithelial cells.
Figure 3
Figure 3. Chemotherapy-Induced SASP in Culture and In Vivo
(A) Overall correlation between XRA and mitoxantrone (MIT)-induced SASPs for all three prostate epithelial cancer cell lines (BPH-1, RWPE-1, and PC-3). Correlations for the individual cell lines are given in the table to the right. The senescence inducer (XRA or MIT) is given in parentheses. (B–E) Human tumor cells were isolated from biopsies obtained from the same patient before MIT chemotherapy and from prostate tissue following prostatectomy after MIT chemotherapy. Laser captured cells were analyzed by quantitative RT-PCR for the mRNAs encoding the indicated proteins, as described in Materials and Methods and Text S1. The results are displayed on a log10 scale indicating the values before (horizontal or x-axis) compared to after (vertical or y-axis) chemotherapy (top left panel in [B]). Each black dot in (B, C, and D) represents the results obtained from a single patient. The average values for all patients before versus after chemotherapy are indicated by a red dot (B–D); these values are also represented as a heat map in (E). (B) Values for mRNAs encoding proteins associated with senescence (p16 and p21) and proliferation (cyclin A, MCM-3, and PCNA). (C) Values for mRNAs encoding SASP-associated proteins (IL-6, IL-8, GM-CSF, GRO-α, IGFBP-2, and IL-1β). (D) Value for an mRNA encoding a non-SASP–associated protein (IL-2). (E) Averages for the values shown in (B–D). Overall p-values, determined by the Student t-test, and number of paired samples (or patients) analyzed for each mRNA are given to the right of the heat map. Signals higher than the prechemotherapy baseline are shown in red; signals below baseline are displayed in green.
Figure 4
Figure 4. Novel SASP Biological Activities and Key Factors
(A) T47D and ZR75.1 cells were incubated for 2 d with CM from PRE fibroblasts, or SEN fibroblasts induced by XRA. The cells were photographed under phase contrast, or analyzed for cluster size using an automated Cellomix imager and software. Smaller cluster or clump sizes (pixel2) indicate greater scattering. The senescence inducer is given in parentheses. Quadruple asterisks (****) indicate p < 0.001. Error bars indicate the standard deviation around the mean. (B) T47D and ZR75.1 cells were incubated with the indicated CM for 3 d and immunostained for the indicated EMT marker proteins. Induction of the mesenchymal marker vimentin by SEN (XRA) CM is shown by the western blot in Figure 7A. (C) Epithelial cell invasion was measured using Boyden chambers containing CM alone or CM plus IL-6 and IL-8 recombinant proteins or IL-6 and IL-8 blocking antibodies, as described in Materials and Methods. After 16–24 h, invasion was scored by counting the number of cells on the underside of the membrane. Invasion stimulated by PRE CM was given a value of one, and other conditions were normalized to this value. Invasion was significantly stimulated by recombinant protein and significantly inhibited by antibodies (p < 0.05). Error bars indicate the standard deviation around the mean.
Figure 5
Figure 5. Oncogenic RAS Amplifies the SASP
(A) Soluble factors produced by the indicated fibroblasts were analyzed by antibody arrays and displayed as described for Figure 1A, but in this case, PRE signals were used as the baseline. Therefore, color intensities represent log2-fold changes of SEN CM relative to PRE CM from cells of the same genotype under the same culture conditions. We pooled and averaged highly correlated data (see Figure 1C) from cells originating from a common tissue (WI-38, IMR-90 from embryonic lung; and HCA-2, BJ from neonatal foreskin), and from senescence induced by REP or XRA. Details of the data processing are provided in Datasets S9–S12. The senescence inducer is given in parentheses. Signals higher than baseline are shown in yellow; signals below baseline are displayed in blue. The numbers on the heat map key (right) indicates log2-fold changes from the baseline. (B) WI-38 cells induced to senesce by RAS were immunostained for the SASP proteins IL-6 and IL-8, and the senescence marker p16INK4a. (C) Correlations between SASPs induced by RAS versus REP or XRA in fibroblasts from embryonic lung (WI-38, IMR-90, left) or neonatal foreskin (HCA-2, right). Correlations for the individual cell strains and senescence inducers are given in the tables below the graphs. (D) Shown are the log2-fold values for factors that are significantly increased in CM from fibroblasts induced to senesce by RAS compared to fibroblasts induced to senesce by XRA or REP. Green indicates SEN WI-38 and IMR-90 cells induced to senesce by RAS versus XRA. Red indicates SEN WI-38 and IMR-90 induced to senesce by RAS versus REP. Gray indicates HCA2 cells induced to senesce by RAS versus XRA. Blue indicates HCA2 cells induced to senesce by RAS versus REP. (E) Shown are the log2-fold values for factors that are uniquely and significantly increased in CM from fibroblasts induced to senesce by RAS compared to PRE CM, but not significantly changed in CM from fibroblasts induced to senesce by XRA or REP. The color code is identical to (D). (F) Soluble factors from CM produced by the indicated epithelial cells were analyzed by antibody arrays and displayed as described for Figure 1A, using PRE CM as the baseline. Signals higher than baseline are shown in yellow; signals below baseline are in blue. Asterisks (*) indicate factors conserved between fibroblasts and epithelial cells. (G) Correlations between SASPs induced by RAS versus XRA in prostate epithelial cells.
Figure 6
Figure 6. p53 Restrains the SASP
(A) CM containing factors secreted by the indicted cells were analyzed by antibody arrays and displayed, using PRE CM as the baseline. We pooled data from cells of the same genotype (p53 wild type or p53 deficient) under the same culture conditions. SEN indicates pooled data from cells originating from the same tissue (WI-38, IMR-90 from embryonic lung; and HCA-2, BJ from neonatal foreskin) and induced to senesce by REP or XRA. Pooling and averaging of highly correlated samples was performed as described for Figure 5, and details of the data processing are provided in Datasets S13–S16. The top four rows are the same top four rows in Figure 5A and are included to serve as a visual reference. The senescence inducer is given in parentheses. p53 status is indicated as either wild type (wt) or deficient (d) owing to either GSE22 expression or expression of an shRNA against p53. Manipulations are indicated in sequence, separated by a greater than symbol (>). The heat map key (right) indicates the log2-fold changes. Signals higher than the baseline are shown in yellow; signals below baseline are displayed in blue. Comparison between rows is accurately illustrated in (B) and (C) in which each genetically manipulated cell type is compared to its appropriate control baseline. (B) Log2-fold values for SASP factors that are significantly increased, or significantly and uniquely (as indicated by double asterisks [**]) elevated, in CM from SEN cells made p53 deficient by GSE22, using untreated wild-type SEN values as the baseline. Green indicates WI-38 cells made senescent by XRA, after which p53 was subsequently inactivated by expressing GSE22 using a lentivirus; these cells do not resume proliferation (“unreverted”) upon p53 inactivation (see Figure S6). Blue indicates WI-38 in which p16 was inactivated by an shRNA, induced to senesce by XRA, then infected with the GSE22-expressing lentivirus; these cells do revert (REV) after p53 inactivation. Pink indicates HCA2 cells made SEN by XRA, then infected with GSE22 lentivirus; these cells also revert after p53 inactivation. (C) Log2-fold values for SASP factors that are significantly increased, or significantly and uniquely (as indicated by double asterisks [**]) elevated, in CM from cells made p53-deficient (by GSE22 expression), then induced to senesce by REP, XRA, or RAS. Red indicates WI-38 and IMR90 (averaged) cells expressing GSE22, then induced to senesce by XRA or REP, using cells made SEN by XRA or REP as a baseline. Gray indicates WI-38, IMR-90, and HCA2 (averaged) expressing GSE22, then made senescent by RAS, using SEN by RAS as a baseline. (D) WI-38 cells expressing GSE22 were induced to senesce by XRA and then immunostained for the SASP proteins IL-6 and IL-8, the senescence marker p16INK4a, and p53, which accumulates in the presence of GSE22. (E) Comparative graphical representation of the secretory profiles of cells made senescent by XRA or REP (dotted line), RAS (black line), p53 inactivation (GSE22) followed by XRA or REP (blue line), or p53 inactivation (GSE22) followed by RAS (red line). The increased slopes (as indicated by the arrow)indicate amplified SASPs. (F) Hierarchical cluster analysis of all the cells analyzed in (A), plus the SASP induced by RAS (see Figure 5). RAS status is indicated as either wild type (wt) or oncogenic (o) owing to expression of Ha-RASv12. (G) WI-38 cells with wild-type (wt) or inactive (GSE) p53 were irradiated or induced to express oncogenic RAS (RAS), and CM was collected 4 or 10 d later. Soluble factors were analyzed by antibody arrays and displayed as described in Figure 1D, using PRE CM as the baseline (black column on the left; see also Figure S5C for details). Signals higher than baseline are shown in yellow; signals below baseline are in blue. n/a, not applicable. (H) Log2-fold values for prostate epithelial cell SASP factors that are significantly or uniquely (as indicated by double asterisks [**]) elevated in CM from p53-deficient cancer cells (PC3, BPH1, and RWPE1) that were induced to senesce by XRA, compared to primary p53 wild-type cells (PrECs) that were induced to senesce by XRA.
Figure 7
Figure 7. Biological Activities of the Amplified SASP
(A) T47D and ZR75.1 cells were incubated with the indicated CM for 3 d and then analyzed for cell scattering, immunostained for the indicated proteins, and analyzed for vimentin and actin levels by western blotting. Controls for the immunofluorescence from the same individual experiment are shown in Figure 4B. The senescence inducer is given in parentheses. p53 status was either wild type or deficient (GSE). Manipulations are indicated in sequence, separated by >. (B) Epithelial cells were incubated with CM from the indicated WI-38 cells and assayed for invasion as described in Materials and Methods and Figure 3C. Double asterisks (**) indicate p < 0.02. Error bars indicate the standard deviation around the mean. (C) Epithelial cells were incubated with CM from the indicated fibroblasts and cell number was determined by cell counting, total protein, or green fluorescent protein (GFP) fluorescence using epithelial cells expressing GFP, as described in [80]. A single asterisks (*) indicates p < 0.05. Error bars indicate the standard deviation around the mean. (D) Model for the cell-nonautonomous activities of oncogenic RAS and the p53 tumor suppressor protein. Oncogenic and genotoxic stress of sufficient magnitude to cause senescence induce a SASP, whereby cells secrete inflammatory cytokines, chemokines, and growth factors that can alter the tissue microenvironment and stimulate malignant phenotypes in nearby cells. Thus, in addition to their well-known cell-autonomous effects, oncogenes can promote cancer cell-nonautonomously by inducing a SASP. Oncogenic and genotoxic stress also activate p53, which suppresses cancer by cell-autonomous mechanisms (promoting repair, or inducing apoptosis or senescence). In addition, p53 suppresses cancer by the cell-nonautonomous effects of suppressing the intensity of the SASP and its deleterious effects.

Similar articles

Cited by

  • Apoptosis signal-regulating kinase 1 promotes inflammation in senescence and aging.
    Odawara T, Yamauchi S, Ichijo H. Odawara T, et al. Commun Biol. 2024 Jun 5;7(1):691. doi: 10.1038/s42003-024-06386-0. Commun Biol. 2024. PMID: 38839869 Free PMC article.
  • Role of cellular senescence in inflammation and regeneration.
    Saito Y, Yamamoto S, Chikenji TS. Saito Y, et al. Inflamm Regen. 2024 Jun 3;44(1):28. doi: 10.1186/s41232-024-00342-5. Inflamm Regen. 2024. PMID: 38831382 Free PMC article. Review.
  • SenNet recommendations for detecting senescent cells in different tissues.
    Suryadevara V, Hudgins AD, Rajesh A, Pappalardo A, Karpova A, Dey AK, Hertzel A, Agudelo A, Rocha A, Soygur B, Schilling B, Carver CM, Aguayo-Mazzucato C, Baker DJ, Bernlohr DA, Jurk D, Mangarova DB, Quardokus EM, Enninga EAL, Schmidt EL, Chen F, Duncan FE, Cambuli F, Kaur G, Kuchel GA, Lee G, Daldrup-Link HE, Martini H, Phatnani H, Al-Naggar IM, Rahman I, Nie J, Passos JF, Silverstein JC, Campisi J, Wang J, Iwasaki K, Barbosa K, Metis K, Nernekli K, Niedernhofer LJ, Ding L, Wang L, Adams LC, Ruiyang L, Doolittle ML, Teneche MG, Schafer MJ, Xu M, Hajipour M, Boroumand M, Basisty N, Sloan N, Slavov N, Kuksenko O, Robson P, Gomez PT, Vasilikos P, Adams PD, Carapeto P, Zhu Q, Ramasamy R, Perez-Lorenzo R, Fan R, Dong R, Montgomery RR, Shaikh S, Vickovic S, Yin S, Kang S, Suvakov S, Khosla S, Garovic VD, Menon V, Xu Y, Song Y, Suh Y, Dou Z, Neretti N. Suryadevara V, et al. Nat Rev Mol Cell Biol. 2024 Jun 3. doi: 10.1038/s41580-024-00738-8. Online ahead of print. Nat Rev Mol Cell Biol. 2024. PMID: 38831121 Review.
  • Comprehensive analysis of gene signatures associated with aging in human aortic dissection.
    Yan J, Tan X. Yan J, et al. Heliyon. 2024 May 17;10(11):e31298. doi: 10.1016/j.heliyon.2024.e31298. eCollection 2024 Jun 15. Heliyon. 2024. PMID: 38828294 Free PMC article.
  • Association Between Atopic Dermatitis and Aging: Clinical Observations and Underlying Mechanisms.
    Chen PY, Shen M, Cai SQ, Tang ZW. Chen PY, et al. J Inflamm Res. 2024 May 28;17:3433-3448. doi: 10.2147/JIR.S467099. eCollection 2024. J Inflamm Res. 2024. PMID: 38828054 Free PMC article. Review.

References

    1. Hanahan D, Weinberg RA. The hallmarks of cancer. Cell. 2000;100:57–70. - PubMed
    1. Bissell MJ, Radisky D. Putting tumours in context. Nature Rev Cancer. 2001;1:46–54. - PMC - PubMed
    1. Coussens LM, Werb Z. Inflammation and cancer. Nature. 2002;420:860–867. - PMC - PubMed
    1. Campisi J, d'Adda di Fagagna F. Cellular senescence: when bad things happen to good cells. Nature Rev Molec Cell Biol. 2007;8:729–740. - PubMed
    1. Dimri GP. What has senescence got to do with cancer. Cancer Cell. 2005;7:505–512. - PMC - PubMed

Publication types

MeSH terms

-