Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Review Article
  • Published: 10 September 2021

Cell cycle control in cancer

  • Helen K. Matthews 1 , 2 ,
  • Cosetta Bertoli   ORCID: orcid.org/0000-0001-5684-4630 1 &
  • Robertus A. M. de Bruin   ORCID: orcid.org/0000-0001-9957-1409 1 , 3  

Nature Reviews Molecular Cell Biology volume  23 ,  pages 74–88 ( 2022 ) Cite this article

56k Accesses

705 Citations

104 Altmetric

Metrics details

  • Cell division
  • Checkpoints
  • DNA damage checkpoints
  • Tumour-suppressor proteins

Cancer is a group of diseases in which cells divide continuously and excessively. Cell division is tightly regulated by multiple evolutionarily conserved cell cycle control mechanisms, to ensure the production of two genetically identical cells. Cell cycle checkpoints operate as DNA surveillance mechanisms that prevent the accumulation and propagation of genetic errors during cell division. Checkpoints can delay cell cycle progression or, in response to irreparable DNA damage, induce cell cycle exit or cell death. Cancer-associated mutations that perturb cell cycle control allow continuous cell division chiefly by compromising the ability of cells to exit the cell cycle. Continuous rounds of division, however, create increased reliance on other cell cycle control mechanisms to prevent catastrophic levels of damage and maintain cell viability. New detailed insights into cell cycle control mechanisms and their role in cancer reveal how these dependencies can be best exploited in cancer treatment.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 12 print issues and online access

176,64 € per year

only 14,72 € per issue

Buy this article

  • Purchase on Springer Link
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

cancer cell research studies

Similar content being viewed by others

cancer cell research studies

Cyclers’ kinases in cell division: from molecules to cancer therapy

cancer cell research studies

The Cyclin-dependent kinase 1: more than a cell cycle regulator

cancer cell research studies

The two sides of chromosomal instability: drivers and brakes in cancer

Kops, G. J. P. L., Foltz, D. R. & Cleveland, D. W. Lethality to human cancer cells through massive chromosome loss by inhibition of the mitotic checkpoint. Proc. Natl Acad. Sci. USA 101 , 8699–8704 (2004).

CAS   PubMed   PubMed Central   Google Scholar  

Pennycook, B. R. & Barr, A. R. Restriction point regulation at the crossroads between quiescence and cell proliferation. FEBS Lett. 54 , 2046–2060 (2020).

Google Scholar  

Rubin, S. M., Sage, J. & Skotheim, J. M. Integrating old and new paradigms of G1/S control. Mol. Cell 80 , 183–192 (2020).

Técher, H., Koundrioukoff, S., Nicolas, A. & Debatisse, M. The impact of replication stress on replication dynamics and DNA damage in vertebrate cells. Nat. Rev. Genet. 18 , 535–550 (2017).

PubMed   Google Scholar  

Scully, R., Panday, A., Elango, R. & Willis, N. A. DNA double-strand break repair-pathway choice in somatic mammalian cells. Nat. Rev. Mol. Cell Biol. 20 , 698–714 (2019).

Malumbres, M. Cyclin-dependent kinases. Genome Biol. 15 , 122 (2014).

PubMed   PubMed Central   Google Scholar  

Musacchio, A. The molecular biology of spindle assembly checkpoint signaling dynamics. Curr. Biol. 2 , R1002–R1018 (2015).

Fisher, R. P. The CDK network: linking cycles of cell division and gene expression. Genes Cancer 3 , 731–738 (2012).

Simmons Kovacs, L. A., Orlando, D. A. & Haase, S. B. Transcription networks and cyclin/CDKs: the yin and yang of cell cycle oscillators. Cell Cycle 7 , 2626–2629 (2008).

Morgan, D. O. The Cell Cycle: Principles of Control (New Science Press, 2007).

Bertoli, C., Skotheim, J. M. & De Bruin, R. A. M. Control of cell cycle transcription during G1 and S phases. Nat. Rev. Mol. Cell Biol. 14 , 518–528 (2013).

Johnson, A. & Skotheim, J. M. Start and the restriction point. Curr. Opin. Cell Biol. 25 , 717–723 (2013).

CAS   PubMed   Google Scholar  

Bertoli, C. & De Bruin, R. A. M. Turning cell cycle entry on its head. eLife 2014 , e03475 (2014).

Caillot, M. et al. Cyclin D1 targets hexokinase 2 to control aerobic glycolysis in myeloma cells. Oncogenesis 9 , 68 (2020).

Sanidas, I. et al. A code of mono-phosphorylation modulates the function of RB. Mol. Cell 73 , 985–1000.e6 (2019).

Narasimha, A. M. et al. Cyclin D activates the Rb tumor suppressor by mono-phosphorylation. eLife 3 , e02872 (2014). Narashima et al. show that cyclin D–CDK4/6 monophosphorylates RB and contrary to the dogma this does not compromise its inhibitory function of E2F-dependent transcription to initiate S phase entry, but it maintains cells in a cell cycle state, preventing their exit .

PubMed Central   Google Scholar  

Topacio, B. R. et al. Cyclin D-Cdk4,6 drives cell-cycle progression via the retinoblastoma protein’s C-terminal helix. Mol. Cell 74 , 758–770.e4 (2019).

Zerjatke, T. et al. Quantitative cell cycle analysis based on an endogenous all-in-one reporter for cell tracking and classification. Cell Rep. 19 , 1953–1966 (2017).

Chung, M. et al. Transient hysteresis in CDK4/6 activity underlies passage of the restriction point in G1. Mol. Cell 76 , 562–573.e4 (2019).

Yang, H. W. et al. Stress-mediated exit to quiescence restricted by increasing persistence in cdk4/6 activation. eLife 9 , e44571 (2020).

Cappell, S. D. et al. EMI1 switches from being a substrate to an inhibitor of APC/CCDH1 to start the cell cycle. Nature 558 , 313–317 (2018).

Barr, A. R., Heldt, F. S., Zhang, T., Bakal, C. & Novák, B. A dynamical framework for the all-or-none G1/S transition. Cell Syst. 2 , 27–37 (2016).

Crncec, A. & Hochegger, H. Triggering mitosis. FEBS Lett. 593 , 2868–2888 (2019).

Hégarat, N. et al. Cyclin a triggers mitosis either via the greatwall kinase pathway or cyclin B. EMBO J. 39 , e104419 (2020).

Ha, S. H. & Ferrell, J. E. Thresholds and ultrasensitivity from negative cooperativity. Science 352 , 990–993 (2016).

Dephoure, N. et al. A quantitative atlas of mitotic phosphorylation. Proc. Natl Acad. Sci. USA 105 , 10762–10767 (2008).

Blethrow, J. D., Glavy, J. S., Morgan, D. O. & Shokat, K. M. Covalent capture of kinase-specific phosphopeptides reveals Cdk1-cyclin B substrates. Proc. Natl Acad. Sci. USA 105 , 1442–1447 (2008).

Joukov, V. & De Nicolo, A. Aurora-PLK1 cascades as key signaling modules in the regulation of mitosis. Sci. Signal. 11 , eaar4195 (2018).

Kettenbach, A. N. et al. Quantitative phosphoproteomics identifies substrates and functional modules of aurora and polo-like kinase activities in mitotic cells. Sci. Signal. 4 , rs5 (2011).

Gavet, O. & Pines, J. Progressive activation of cyclinB1-Cdk1 coordinates entry to mitosis. Dev. Cell 18 , 533–543 (2010).

Taubenberger, A. V., Baum, B. & Matthews, H. K. The mechanics of mitotic cell rounding. Front. Cell Dev. Biol. 8 , 687 (2020).

Agircan, F. G., Schiebel, E. & Mardin, B. R. Separate to operate: control of centrosome positioning and separation. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 369 , 20130461 (2014).

Gavet, O. & Pines, J. Activation of cyclin B1-Cdk1 synchronizes events in the nucleus and the cytoplasm at mitosis. J. Cell Biol. 189 , 247–259 (2010).

Santos, S. D. M., Wollman, R., Meyer, T. & Ferrell, J. E. Spatial positive feedback at the onset of mitosis. Cell 149 , 1500–1513 (2012).

Pines, J. & Hunter, T. Human cyclins A and B1 are differentially located in the cell and undergo cell cycle-dependent nuclear transport. J. Cell Biol. 115 , 1–17 (1991).

Hara, M. & Fukagawa, T. Kinetochore assembly and disassembly during mitotic entry and exit. Curr. Opin. Cell Biol. 52 , 73–81 (2018).

Foley, E. A. & Kapoor, T. M. Microtubule attachment and spindle assembly checkpoint signalling at the kinetochore. Nat. Rev. Mol. Cell Biol. 14 , 25–37 (2013).

Geley, S. et al. Anaphase-promoting complex/cyclosome-dependent proteolysis of human cyclin A starts at the beginning of mitosis and is not subject to the spindle assembly checkpoint. J. Cell Biol. 153 , 137–147 (2001).

Elzen, N. Den & Pines, J. Cyclin A is destroyed in prometaphase and can delay chromosome alignment and anaphase. J. Cell Biol. 153 , 121–136 (2001).

Yamano, H. APC/C: current understanding and future perspectives. F1000Res. https://doi.org/10.12688/f1000research.18582.1 (2019).

Article   PubMed   PubMed Central   Google Scholar  

Holder, J., Poser, E. & Barr, F. A. Getting out of mitosis: spatial and temporal control of mitotic exit and cytokinesis by PP1 and PP2A. FEBS Lett. 593 , 2908–2924 (2019).

Luo, S. & Tong, L. Structural biology of the separase–securin complex with crucial roles in chromosome segregation. Curr. Opin. Struct. Biol. 49 , 114–122 (2018).

Vukušić, K., Buđa, R. & Tolić, I. M. Force-generating mechanisms of anaphase in human cells. J. Cell Sci. 132 , jcs231985 (2019).

Green, A. R., Paluch, E. & Oegema, K. Cytokinesis in animal cells. Annu. Rev. Cell Dev. Biol. 28 , 29–58 (2012).

Huang, R.-X. & Zhou, P.-K. DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer. Signal. Transduct. Target. Ther. 5 , 60 (2020).

Matsuoka, S. et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316 , 1160–1166 (2007).

Bensimon, A. et al. ATM-dependent and -independent dynamics of the nuclear phosphoproteome after DNA damage. Sci. Signal. 3 , rs3 (2010).

Shieh, S. Y., Ahn, J., Tamai, K., Taya, Y. & Prives, C. The human homologs of checkpoint kinases Chk1 and Cds1 (Chk2) phosphorylate p53 at multiple DNA damage-inducible sites. Genes Dev. 14 , 289–300 (2000).

Hirao, A. et al. DNA damage-induced activation of p53 by the checkpoint kinase Chk2. Science 287 , 1824–1827 (2000).

Chehab, N. H., Malikzay, A., Appel, M. & Halazonetis, T. D. Chk2/hCds1 functions as a DNA damage checkpoint in G(1) by stabilizing p53. Genes Dev. 14 , 278–288 (2000).

Reinhardt, H. C. & Yaffe, M. B. Kinases that control the cell cycle in response to DNA damage: Chk1, Chk2, and MK2. Curr. Opin. Cell Biol. 21 , 245–255 (2009).

Brandsma, I. & Gent, D. C. Pathway choice in DNA double strand break repair: observations of a balancing act. Genome Integr. 3 , 9 (2012).

Janssen, A. & Medema, R. H. Genetic instability: tipping the balance. Oncogene 32 , 4459–4470 (2013).

Hafner, A., Bulyk, M. L., Jambhekar, A. & Lahav, G. The multiple mechanisms that regulate p53 activity and cell fate. Nat. Rev. Mol. Cell Biol. 20 , 199–210 (2019).

Zeman, M. K. & Cimprich, K. A. Causes and consequences of replication stress. Nat. Cell Biol. 16 , 2–9 (2014).

Herlihy, A. E. & de Bruin, R. A. M. The role of the transcriptional response to DNA replication stress. Genes 8 , 92 (2017).

Kotsantis, P., Petermann, E. & Boulton, S. J. Mechanisms of oncogene-induced replication stress: jigsaw falling into place. Cancer Discov. 8 , 537–555 (2018).

Abraham, R. T. Cell cycle checkpoint signaling through the ATM and ATR kinases. Genes Dev. 15 , 2177–2196 (2001).

Jackson, S. P. & Bartek, J. The DNA-damage response in human biology and disease. Nature 461 , 1071–1078 (2009).

McGowan, C. H. & Russell, P. The DNA damage response: sensing and signaling. Curr. Opin. Cell Biol. 16 , 629–633 (2004).

Zhang, Y. & Hunter, T. Roles of Chk1 in cell biology and cancer therapy. Int. J. Cancer 134 , 1013–1023 (2014).

Rothblum-Oviatt, C. J., Ryan, C. E. & Piwnica-Worms, H. 14-3-3 binding regulates catalytic activity of human Wee1 kinase. Cell Growth Differ. 12 , 581–589 (2001).

Lee, J., Kumagai, A. & Dunphy, W. G. Positive regulation of Wee1 by Chk1 and 14-3-3 proteins. Mol. Biol. Cell 12 , 551–563 (2001).

Lopes, M. et al. The DNA replication checkpoint response stabilizes stalled replication forks. Nature 412 , 557–561 (2001).

Feijoo, C. et al. Activation of mammalian Chk1 during DNA replication arrest: a role for Chk1 in the intra-S phase checkpoint monitoring replication origin firing. J. Cell Biol. 154 , 913–924 (2001).

Tercero, J. A. & Diffley, J. F. X. Regulation of DNA replication fork progression through damaged DNA by the Mec1/Rad53 checkpoint. Nature 412 , 553–557 (2001).

Zachos, G., Rainey, M. D. & Gillespie, D. A. F. Chk1-deficient tumour cells are viable but exhibit multiple checkpoint and survival defects. EMBO J. 22 , 713–723 (2003).

Petermann, E. et al. Chk1 requirement for high global rates of replication fork progression during normal vertebrate S phase. Mol. Cell. Biol. 26 , 3319–3326 (2006).

Katsuno, Y. et al. Cyclin A–Cdk1 regulates the origin firing program in mammalian cells. Proc. Natl Acad. Sci. USA 106 , 3184–3189 (2009).

Petermann, E., Woodcock, M. & Helleday, T. Chk1 promotes replication fork progression by controlling replication initiation. Proc. Natl Acad. Sci. USA 107 , 16090–16095 (2010).

Conti, C., Seiler, J. A. & Pommier, Y. The mammalian DNA replication elongation checkpoint: implication of Chk1 and relationship with origin firing as determined by single DNA molecule and single cell analyses. Cell Cycle 6 , 2760–2767 (2007).

Dimitrova, D. S. & Gilbert, D. M. Temporally coordinated assembly and disassembly of replication factories in the absence of DNA synthesis. Nat. Cell Biol. 2 , 686–694 (2000).

Yan, S. & Michael, W. M. TopBP1 and DNA polymerase α-mediated recruitment of the 9-1-1 complex to stalled replication forks: implications for a replication restart-based mechanism for ATR checkpoint activation. Cell Cycle 8 , 2877–2884 (2009).

El-shemerly, M., Hess, D., Pyakurel, A. K., Moselhy, S. & Ferrari, S. ATR-dependent pathways control hEXO1 stability in response to stalled forks. Nucleic Acids Res. 36 , 511–519 (2008).

Casper, A. M., Nghiem, P., Arlt, M. F. & Glover, T. W. ATR regulates fragile site stability. Cell 111 , 779–789 (2002).

Trenz, K., Smith, E., Smith, S. & Costanzo, V. ATM and ATR promote Mre11 dependent restart of collapsed replication forks and prevent accumulation of DNA breaks. EMBO J. https://doi.org/10.1038/sj.emboj.7601045 (2006).

Hayward, D., Alfonso-Pérez, T. & Gruneberg, U. Orchestration of the spindle assembly checkpoint by CDK1-cyclin B1. FEBS Lett. 593 , 2889–2907 (2019).

Grieco, D. & Serpico, A. F. Recent advances in understanding the role of Cdk1 in the spindle assembly checkpoint. F1000Res. https://doi.org/10.12688/f1000research.21185.1 (2020).

Yamaguchi, M. et al. Cryo-EM of mitotic checkpoint complex-bound APC/C reveals reciprocal and conformational regulation of ubiquitin ligation. Mol. Cell 63 , 593–607 (2016).

Alfieri, C. et al. Molecular basis of APC/C regulation by the spindle assembly checkpoint. Nature 536 , 431–436 (2016).

Sudakin, V., Chan, G. K. T. & Yen, T. J. Checkpoint inhibition of the APC/C in HeLa cells is mediated by a complex of BUBR1, BUB3, CDC20, and MAD2. J. Cell Biol. 154 , 925–936 (2001).

Liu, S.-T. & Zhang, H. The mitotic checkpoint complex (MCC): looking back and forth after 15 years. AIMS Mol. Sci. 3 , 597–634 (2016).

Kapanidou, M., Curtis, N. L. & Bolanos-Garcia, V. M. Cdc20: at the crossroads between chromosome segregation and mitotic exit. Trends Biochem. Sci. 42 , 193–205 (2017).

Liu, D., Vader, G., Vromans, M. J. M., Lampson, M. A. & Lens, S. M. A. Sensing chromosome bi-orientation by spatial separation of Aurora B kinase from kinetochore substrates. Science 323 , 1350–1353 (2009).

Rieder, C. L., Cole, R. W., Khodjakov, A. & Sluder, G. The checkpoint delaying anaphase in response to chromosome monoorientation is mediated by an inhibitory signal produced by unattached kinetochores. J. Cell Biol. 130 , 941–948 (1995).

Gascoigne, K. E. & Taylor, S. S. Cancer cells display profound intra- and interline variation following prolonged exposure to antimitotic drugs. Cancer Cell 14 , 111–122 (2008).

Vitale, I., Manic, G., Castedo, M. & Kroemer, G. Caspase 2 in mitotic catastrophe: the terminator of aneuploid and tetraploid cells. Mol. Cell. Oncol. 4 , e1299274 (2017).

Cheng, B. & Crasta, K. Consequences of mitotic slippage for antimicrotubule drug therapy. Endocr. Relat. Cancer 24 , T97–T106 (2017).

Brito, D. A. & Rieder, C. L. Mitotic checkpoint slippage in humans occurs via cyclin b destruction in the presence of an active checkpoint. Curr. Biol. 16 , 1194–1200 (2006).

Fujiwara, T. et al. Cytokinesis failure generating tetraploids promotes tumorigenesis in p53-null cells. Nature 437 , 1043–1047 (2005).

Thompson, S. L. & Compton, D. A. Proliferation of aneuploid human cells is limited by a p53-dependent mechanism. J. Cell Biol. 188 , 369–381 (2010).

Sanchez-Vega, F. et al. Oncogenic signaling pathways in the cancer genome atlas. Cell 173 , 321–337.e10 (2018).

Lecona, E. & Fernandez-Capetillo, O. Targeting ATR in cancer. Nat. Rev. Cancer 18 , 586–595 (2018).

Ghelli Luserna Di Rorà, A., Cerchione, C., Martinelli, G. & Simonetti, G. A WEE1 family business: regulation of mitosis, cancer progression, and therapeutic target. J. Hematol. Oncol. 13 , 1–17 (2020).

Peyressatre, M., Prével, C., Pellerano, M. & Morris, M. C. Targeting cyclin-dependent kinases in human cancers: from small molecules to peptide inhibitors. Cancers 7 , 179–237 (2015).

Liu, K. et al. The role of CDC25C in cell cycle regulation and clinical cancer therapy: a systematic review. Cancer Cell Int. 20 , 1–16 (2020).

Pérez de Castro, I., de Cárcer, G. & Malumbres, M. A census of mitotic cancer genes: new insights into tumor cell biology and cancer therapy. Carcinogenesis 28 , 899–912 (2007).

Bates, M. et al. Too MAD or not MAD enough: the duplicitous role of the spindle assembly checkpoint protein MAD2 in cancer. Cancer Lett. 469 , 11–21 (2020).

Wang, L. et al. Targeting Cdc20 as a novel cancer therapeutic strategy. Pharmacol. Ther. 151 , 141–151 (2015).

Xie, Y. et al. Mps1/TTK: a novel target and biomarker for cancer. J. Drug Target. 25 , 112–118 (2017).

Borah, N. A. & Reddy, M. M. Aurora kinase B inhibition: a potential therapeutic strategy for cancer. Molecules 26 , 1981 (2021).

Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144 , 646–674 (2011).

Zilfou, J. T. & Lowe, S. W. Tumor suppressive functions of p53. Cold Spring Harb. Perspect. Biol. 1 , a001883 (2009).

Chen, J. The cell-cycle arrest and apoptotic functions of p53 in tumor initiation and progression. Cold Spring Harb. Perspect. Med. 6 , a026104 (2016).

Chen, H. Z., Tsai, S. Y. & Leone, G. Emerging roles of E2Fs in cancer: an exit from cell cycle control. Nat. Rev. Cancer 9 , 785–797 (2009).

Di Micco, R. et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444 , 638–642 (2006). Di Micco et al. show that deregulated DNA replication, due to oncogene activation, triggers DNA damage response and cell cycle exit through senescence, establishing the concept of oncogene-induced replication stress driving cancer initiation .

Bartkova, J. et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444 , 633–637 (2006).

Jones, R. M. et al. Increased replication initiation and conflicts with transcription underlie cyclin E-induced replication stress. Oncogene 32 , 3744–3753 (2013).

Bester, A. C. et al. Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145 , 435–446 (2011). Bester et al. show that replication stress is an early event in cancer development and fuels genome instability, establishing the role of oncogene-induced replication stress in cancer initiation .

Dominguez-Sola, D. et al. Non-transcriptional control of DNA replication by c-Myc. Nature 448 , 445–451 (2007).

Swanton, C. et al. Chromosomal instability determines taxane response. Proc. Natl Acad. Sci. USA 106 , 8671–8676 (2009).

Lee, A. J. X. et al. Chromosomal instability confers intrinsic multidrug resistance. Cancer Res. 71 , 1858–1870 (2011).

Gerlinger, M. et al. Intratumor heterogeneity and branched evolution revealed by multiregion sequencing. N. Engl. J. Med. 366 , 883–892 (2012).

Burrell, R. A. et al. Replication stress links structural and numerical cancer chromosomal instability. Nature 494 , 492–496 (2013). Burrell et al. show that replication stress can cause CIN in cancer cells .

Lecona, E. & Fernández-Capetillo, O. Replication stress and cancer: it takes two to tango. Exp. Cell Res. 329 , 26–34 (2014).

Wilhelm, T., Said, M. & Naim, V. DNA replication stress and chromosomal instability: dangerous liaisons. Genes (Basel) 11 , 642 (2020).

CAS   Google Scholar  

Bartek, J., Bartkova, J. & Lukas, J. DNA damage signalling guards against activated oncogenes and tumour progression. Oncogene 26 , 7773–7779 (2007).

Bartkova, J. et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434 , 864–870 (2005).

Gorgoulis, V. G. et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434 , 907–913 (2005).

Taylor, A. M. et al. Genomic and functional approaches to understanding cancer aneuploidy. Cancer Cell 33 , 676–689.e3 (2018).

Boveri, T. Concerning the origin of malignant tumours by Theodor Boveri. Translated and annotated by Henry Harris. J. Cell Sci. 121 , 1–84 (2008).

Ganem, N. J. & Pellman, D. Linking abnormal mitosis to the acquisition of DNA damage. J. Cell Biol. 199 , 871–881 (2012).

Cahill, D. P. et al. Mutations of mitotic checkpoint genes in human cancers. Nature 392 , 300–303 (1998).

Ryan, S. D. et al. Up-regulation of the mitotic checkpoint component Mad1 causes chromosomal instability and resistance to microtubule poisons. Proc. Natl Acad. Sci. USA 109 , E2205–E2214 (2012).

Hernando, E. et al. Rb inactivation promotes genomic instability by uncoupling cell cycle progression from mitotic control. Nature 430 , 797–802 (2004).

Yuan, B. et al. Increased expression of mitotic checkpoint genes in breast cancer cells with chromosomal instability. Clin. Cancer Res. 12 , 405–410 (2006).

Sarkar, S. et al. Mitotic checkpoint defects: en route to cancer and drug resistance. Chromosom. Res. 29 , 131–144 (2021).

Sisken, J. E., Bonner, S. V., Grasch, S. D., Powell, D. E. & Donaldson, E. S. Alterations in metaphase durations in cells derived from human tumours. Cell Prolif. 18 , 137–146 (1985).

Therman, E., Buchler, D. A., Nieminen, U. & Timonen, S. Mitotic modifications and aberrations in human cervical cancer. Cancer Genet. Cytogenet. 11 , 185–197 (1984).

Kwiatkowski, N. et al. Small-molecule kinase inhibitors provide insight into Mps1 cell cycle function. Nat. Chem. Biol. 6 , 359–368 (2010).

Stolz, A. et al. Pharmacologic abrogation of the mitotic spindle checkpoint by an indolocarbazole discovered by cellular screening efficiently kills cancer cells. Cancer Res. 69 , 3874–3883 (2009).

Siri, S. O., Martino, J. & Gottifredi, V. Structural chromosome instability: types, origins, consequences, and therapeutic opportunities. Cancers (Basel) 13 , 3056 (2021).

Janssen, A., van der Burg, M., Szuhai, K., Kops, G. J. P. L. & Medema, R. H. Chromosome segregation errors as a cause of DNA damage and structural chromosome aberrations. Science 333 , 1895–1898 (2011).

Bakhoum, S. F., Kabeche, L., Murnane, J. P., Zaki, B. I. & Compton, D. A. DNA-damage response during mitosis induces whole-chromosome missegregation. Cancer Discov. 4 , 1281–1289 (2014).

Funk, L. C., Zasadil, L. M. & Weaver, B. A. Living in CIN: mitotic infidelity and its consequences for tumor promotion and suppression. Dev. Cell 39 , 638–652 (2016).

Sansregret, L. & Swanton, C. The role of aneuploidy in cancer evolution. Cold Spring Harb. Perspect. Med. 7 , 1–18 (2017).

Silk, A. D. et al. Chromosome missegregation rate predicts whether aneuploidy will promote or suppress tumors. Proc. Natl Acad. Sci. USA 110 , E4134–E4141 (2013).

Zasadil, L. M. et al. High rates of chromosome missegregation suppress tumor progression but do not inhibit tumor initiation. Mol. Biol. Cell 27 , 1981–1989 (2016).

Birkbak, N. J. et al. Paradoxical relationship between chromosomal instability and survival outcome in cancer. Cancer Res. 71 , 3447–3452 (2011).

Roylance, R. et al. Relationship of extreme chromosomal instability with long-term survival in a retrospective analysis of primary breast cancer. Cancer Epidemiol. Biomarkers Prev. 20 , 2183–2194 (2011).

Sansregret, L. et al. APC/C dysfunction limits excessive cancer chromosomal instability. Cancer Discov. 7 , 218–233 (2017). Sansregret et al. show that strengthening the SAC (by APC/C partial depletion) and prolonging mitosis prevents chromosome segregation errors in mitosis and guards against excessive genome instability in cancer cells .

Pearl, L. H., Schierz, A. C., Ward, S. E., Al-Lazikani, B. & Pearl, F. M. G. Therapeutic opportunities within the DNA damage response. Nat. Rev. Cancer 15 , 166–180 (2015).

Mukherjee, S. The Emperor of All Maladies: A Biography of Cancer (Scribner, 2010).

Malumbres, M. & Barbacid, M. Cell cycle, CDKs and cancer: a changing paradigm. Nat. Rev. Cancer 9 , 153–166 (2009).

Wagner, V. & Gil, J. Senescence as a therapeutically relevant response to CDK4/6 inhibitors. Oncogene 39 , 5165–5176 (2020).

Wang, C. et al. Inducing and exploiting vulnerabilities for the treatment of liver cancer. Nature 574 , 268–272 (2019).

Whittaker, S. R., Mallinger, A., Workman, P. & Clarke, P. A. Inhibitors of cyclin-dependent kinases as cancer therapeutics. Pharmacol. Ther. 173 , 83–105 (2017).

Sánchez-Martínez, C., Lallena, M. J., Sanfeliciano, S. G. & de Dios, A. Cyclin dependent kinase (CDK) inhibitors as anticancer drugs: recent advances (2015–2019). Bioorg. Med.Chem. Lett. 29 , 126637 (2019).

Choi, Y. J. & Anders, L. Signaling through cyclin D-dependent kinases. Oncogene 33 , 1890–1903 (2014).

Álvarez-Fernández, M. & Malumbres, M. Mechanisms of sensitivity and resistance to CDK4/6 inhibition. Cancer Cell 37 , 514–529 (2020).

Spring, L. M. et al. Cyclin-dependent kinase 4 and 6 inhibitors for hormone receptor-positive breast cancer: past, present, and future. Lancet 395 , 817–827 (2020).

Guiley, K. Z. et al. p27 allosterically activates cyclin-dependent kinase 4 and antagonizes palbociclib inhibition. Science 366 , eaaw2106 (2019).

Persky, N. S. et al. Defining the landscape of ATP-competitive inhibitor resistance residues in protein kinases. Nat. Struct. Mol. Biol. 27 , 92–104 (2020).

Schade, A. E., Oser, M. G., Nicholson, H. E. & DeCaprio, J. A. Cyclin D-CDK4 relieves cooperative repression of proliferation and cell cycle gene expression by DREAM and RB. Oncogene 38 , 4962–4976 (2019).

Ruscetti, M. et al. NK cell–mediated cytotoxicity contributes to tumor control by a cytostatic drug combination. Science 362 , 1416–1422 (2018).

Ruscetti, M. et al. Senescence-induced vascular remodeling creates therapeutic vulnerabilities in pancreas cancer. Cell 181 , 424–441.e21 (2020).

Christensen, C. L. et al. Targeting transcriptional addictions in small cell lung cancer with a covalent CDK7 inhibitor. Cancer Cell 26 , 909–922 (2014).

Patel, H. et al. ICEC0942, an orally bioavailable selective inhibitor of CDK7 for cancer treatment. Mol. Cancer Ther. 17 , 1156 LP–1151166 (2018).

Greenall, S. A. et al. Cyclin-dependent kinase 7 is a therapeutic target in high-grade glioma. Oncogenesis 6 , e336 (2017).

Eliades, P. et al. High MITF expression is associated with super-enhancers and suppressed by CDK7 inhibition in melanoma. J. Invest. Dermatol. 138 , 1582–1590 (2018).

Zhong, L., Yang, S., Jia, Y. & Lei, K. Inhibition of cyclin-dependent kinase 7 suppresses human hepatocellular carcinoma by inducing apoptosis. J. Cell. Biochem. 119 , 9742–9751 (2018).

Cao, X. et al. Targeting super-enhancer-driven oncogenic transcription by CDK7 inhibition in anaplastic thyroid carcinoma. Thyroid 29 , 809–823 (2019).

Lu, P. et al. THZ1 reveals CDK7-dependent transcriptional addictions in pancreatic cancer. Oncogene 38 , 3932–3945 (2019).

Zhong, S., Zhang, Y., Yin, X. & Di, W. CDK7 inhibitor suppresses tumor progression through blocking the cell cycle at the G2/M phase and inhibiting transcriptional activity in cervical cancer. Onco. Targets. Ther. 12 , 2137–2147 (2019).

Zhang, Y. et al. The covalent CDK7 inhibitor THZ1 potently induces apoptosis in multiple myeloma cells in vitro and in vivo. Clin. Cancer Res. 25 , 6195–6205 (2019).

Schachter, M. M. & Fisher, R. P. The CDK-activating kinase Cdk7. Cell Cycle 12 , 3239–3240 (2013).

Sava, G. P., Fan, H., Coombes, R. C., Buluwela, L. & Ali, S. CDK7 inhibitors as anticancer drugs. Cancer Metastasis Rev. 39 , 805–823 (2020).

Barr, A. R. et al. DNA damage during S-phase mediates the proliferation-quiescence decision in the subsequent G1 via p21 expression. Nat. Commun. 8 , 14728 (2017).

Shiohara, M., Koike, K., Komiyama, A. & Koeffler, H. P. p21WAF1 mutations and human malignancies. Leuk. Lymphoma 26 , 35–41 (1997).

Hirai, H. et al. Small-molecule inhibition of Wee1 kinase by MK-1775 selectively sensitizes p53-deficient tumor cells to DNA-damaging agents. Mol. Cancer Ther. 8 , 2992–3000 (2009).

Matheson, C. J., Backos, D. S. & Reigan, P. Targeting WEE1 kinase in cancer. Trends Pharmacol. Sci. 37 , 872–881 (2016).

Moens, S. et al. The mitotic checkpoint is a targetable vulnerability of carboplatin-resistant triple negative breast cancers. Sci. Rep. 11 , 3176 (2021).

Shamloo, B. & Usluer, S. p21 in cancer research. Cancers (Basel) 11 , 1178 (2019).

Macheret, M. & Halazonetis, T. D. DNA replication stress as a hallmark of cancer. Annu. Rev. Pathol. 10 , 425–448 (2015).

Murga, M. et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat. Struct. Mol. Biol. 18 , 1331–1335 (2011). Murga et al. show that inhibition of ATR and CHK1 can kill cells with oncogene-induced replication stress, establishing that targeting the replication stress response can selectively kill cancer .

Toledo, L. I., Murga, M. & Fernandez-Capetillo, O. Targeting ATR and Chk1 kinases for cancer treatment: a new model for new (and old) drugs. Mol. Oncol. 5 , 368–373 (2011).

Toledo, L. I. et al. A cell-based screen identifies ATR inhibitors with synthetic lethal properties for cancer-associated mutations. Nat. Struct. Mol. Biol. 18 , 721–727 (2011).

León, T. E. et al. EZH2-deficient T-cell acute lymphoblastic leukemia is sensitized to CHK1 Inhibition through enhanced replication stress. Cancer Discov. https://doi.org/10.1158/2159-8290.CD-19-0789 (2020).

Rogers, R. F. et al. CHK1 inhibition is synthetically lethal with loss of B-family DNA polymerase function in human lung and colorectal cancer cells. Cancer Res. 80 , 1735–1747 (2020).

Qiu, Z., Oleinick, N. L. & Zhang, J. ATR/CHK1 inhibitors and cancer therapy. Radiother . Oncol. J. Eur. Soc. Ther. Radiol. Oncol. 126 , 450–464 (2018).

Forment, J. V. & O’Connor, M. J. Targeting the replication stress response in cancer. Pharmacol. Ther. 188 , 155–167 (2018).

Eykelenboom, J. K. et al. ATR Activates the S-M checkpoint during unperturbed growth to ensure sufficient replication prior to mitotic onset. Cell Rep. 5 , 1095–1107 (2013).

Bertoli, C., Herlihy, A. E., Pennycook, B. R., Kriston-Vizi, J. & De Bruin, R. A. M. Sustained E2F-dependent transcription is a key mechanism to prevent replication-stress-induced DNA damage. Cell Rep. 15 , 1412–1422 (2016).

Penna, L. S., Henriques, J. A. P. & Bonatto, D. Anti-mitotic agents: are they emerging molecules for cancer treatment? Pharmacol. Ther. 173 , 67–82 (2017).

Weaver, B. A. How taxol/paclitaxel kills cancer cells. Mol. Biol. Cell 25 , 2677–2681 (2014).

Zasadil, L. M. et al. Cytotoxicity of paclitaxel in breast cancer is due to chromosome missegregation on multipolar spindles. Sci. Transl. Med. 6 , 1–10 (2014). Zasadil et al. show that the cytotoxic effects of paclitaxel at clinically relevant concentrations (as measured in human tumours) are not due to prolonged SAC activation and mitotic arrest but rather are due to chromosome segregation defects that result in unviable karyotypes and cell death .

Tischer, J. & Gergely, F. Anti-mitotic therapies in cancer. J. Cell Biol. 218 , 10–11 (2019).

Yan, V. C. et al. Why great mitotic inhibitors make poor cancer drugs. Trends Cancer 6 , 924–941 (2020).

Myers, S. M. & Collins, I. Recent findings and future directions for interpolar mitotic kinesin inhibitors in cancer therapy. Future Med. Chem. 8 , 463–489 (2016).

Borisa, A. C. & Bhatt, H. G. A comprehensive review on Aurora kinase: small molecule inhibitors and clinical trial studies. Eur. J. Med. Chem. 140 , 1–19 (2017).

Gutteridge, R. E. A., Ndiaye, M. A., Liu, X. & Ahmad, N. Plk1 inhibitors in cancer therapy: from laboratory to clinics. Mol. Cancer Ther. 15 , 1427–1435 (2016).

El-Arabey, A. A., Salama, S. A. & Abd-Allah, A. R. CENP-E as a target for cancer therapy: where are we now? Life Sci. 208 , 192–200 (2018).

Rowald, K. et al. Negative selection and chromosome instability induced by Mad2 overexpression delay breast cancer but facilitate oncogene-independent outgrowth. Cell Rep. 15 , 2679–2691 (2016).

Maia, A. R. R. et al. Mps1 inhibitors synergise with low doses of taxanes in promoting tumour cell death by enhancement of errors in cell division. Br. J. Cancer 118 , 1586–1595 (2018).

Cohen-Sharir, Y. et al. Aneuploidy renders cancer cells vulnerable to mitotic checkpoint inhibition. Nature 590 , 486–491 (2021). Cohen-Sharir et al. show that aneuploid cells are more sensitive to SAC inhibition than diploid cells. Aneuploid cells were found to depend on an intact SAC for accurate chromosome segregation and long-term survival, implicating the SAC as a potential therapeutic target in aneuploid cancers .

Simon Serrano, S. et al. Inhibition of mitotic kinase Mps1 promotes cell death in neuroblastoma. Sci. Rep. 10 , 11997 (2020).

Wengner, A. M. et al. Novel Mps1 kinase inhibitors with potent antitumor activity. Mol. Cancer Ther. 15 , 583–592 (2016).

Alimova, I. et al. MPS1 kinase as a potential therapeutic target in medulloblastoma. Oncol. Rep. 36 , 2633–2640 (2016).

Siemeister, G. et al. Inhibition of BUB1 kinase by Bay 1816032 sensitizes tumor cells toward taxanes, ATR, and PARP inhibitors in vitro and in vivo. Clin. Cancer Res. 25 , 1404–1414 (2019).

Silva, P. M. A. et al. Suppression of spindly delays mitotic exit and exacerbates cell death response of cancer cells treated with low doses of paclitaxel. Cancer Lett. 394 , 33–42 (2017).

Finn, R. S., Aleshin, A. & Slamon, D. J. Targeting the cyclin-dependent kinases (CDK) 4/6 in estrogen receptor-positive breast cancers. Breast Cancer Res. 18 , 17 (2016).

Fry, D. W. et al. Specific inhibition of cyclin-dependent kinase 4/6 by PD 0332991 and associated antitumor activity in human tumor xenografts. Mol. Cancer Ther. 3 , 1427–1438 (2004).

Hortobagyi, G. N. et al. Ribociclib as first-line therapy for HR-positive, advanced breast cancer. N. Engl. J. Med. 375 , 1738–1748 (2016).

Rader, J. et al. Dual CDK4/CDK6 inhibition induces cell-cycle arrest and senescence in neuroblastoma. Clin. Cancer Res. 19 , 6173–6182 (2013).

Tripathy, D. et al. Ribociclib plus endocrine therapy for premenopausal women with hormone-receptor-positive, advanced breast cancer (MONALEESA-7): a randomised phase 3 trial. Lancet Oncol. 19 , 904–915 (2018).

Sledge, G. W. Jr. Curing metastatic breast cancer. J. Oncol. Pract. 12 , 6–10 (2016).

Dickler, M. N. et al. MONARCH 1, A phase II study of abemaciclib, a CDK4 and CDK6 inhibitor, as a single agent, in patients with refractory HR + /HER2 − metastatic breast cancer. Clin. Cancer Res. 23 , 5218–5224 (2017).

Gelbert, L. M. et al. Preclinical characterization of the CDK4/6 inhibitor LY2835219: in-vivo cell cycle-dependent/independent anti-tumor activities alone/in combination with gemcitabine. Invest. N. Drugs 32 , 825–837 (2014).

Schulze, V. K. et al. Treating cancer by spindle assembly checkpoint abrogation: discovery of two clinical candidates, BAY 1161909 and BAY 1217389, targeting MPS1 kinase. J. Med. Chem. 63 , 8025–8042 (2020).

DePamphilis, M. L. Genome duplication at the beginning of mammalian development. Curr. Top. Dev. Biol. 120 , 55–102 (2016).

Arias, E. E. & Walter, J. C. Strength in numbers: preventing rereplication via multiple mechanisms in eukaryotic cells. Genes Dev. 21 , 497–518 (2007).

Davidson, I. F., Li, A. & Blow, J. J. Deregulated replication licensing causes DNA fragmentation consistent with head-to-tail fork collision. Mol. Cell 24 , 433–443 (2006).

Neelsen, K. J. et al. Deregulated origin licensing leads to chromosomal breaks by rereplication of a gapped DNA template. Genes Dev. 27 , 2537–2542 (2013).

Fragkos, M., Ganier, O., Coulombe, P. & Méchali, M. DNA replication origin activation in space and time. Nat. Rev. Mol. Cell Biol. 16 , 360–374 (2015).

Download references

Acknowledgements

R.A.M.d.B. and C.B. are supported by core funding from the MRC–UCL University Unit (reference MC_EX_G0800785) and R.A.M.d.B.’s Cancer Research UK Programme Foundation Award. H.K.M received funding from a CRUK–EPSRC Multidisciplinary Project Award (C1529/A23335). The authors thank J. Pines and J. Downs for helpful discussions and A. Barr and the peer reviewers for critical reading of the manuscript. They apologize to colleagues whose work could be cited only indirectly.

Author information

Authors and affiliations.

MRC Laboratory for Molecular Cell Biology, University College London, London, UK

Helen K. Matthews, Cosetta Bertoli & Robertus A. M. de Bruin

Department of Biomedical Science, University of Sheffield, Sheffield, UK

Helen K. Matthews

UCL Cancer Institute, University College London, London, UK

Robertus A. M. de Bruin

You can also search for this author in PubMed   Google Scholar

Contributions

The authors contributed equally to the writing and revisions of the article.

Corresponding author

Correspondence to Robertus A. M. de Bruin .

Ethics declarations

Competing interests.

The authors declare no competing financial interests.

Additional information

Peer review information.

Nature Reviews Molecular Cell Biology thanks Fred Dick, Tobias Meyer and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

(CDKs). CDKs depend on cyclins for their kinase activity. The levels of cyclins increase during the cell cycle and cyclin–CDK complex activity drives cell cycle progression by phosphorylating protein targets.

A reversible cellular state, outside G1 phase, from which cells can re-enter the cell cycle.

(APC/C). Ubiquitin ligase complex activity that is restricted to mitosis and G1 phase and is required to initiate exit from mitosis and indirectly for DNA replication.

Removal of nucleotides by exonucleases involved in repairing DNA, exposing tracks of single-stranded DNA at sites of double-strand break repair. It is required for homologous recombination.

DNA double-strand break repair mechanism based on the juxtaposition of two pieces of DNA.

Mechanism of DNA double-strand break repair requiring the presence of duplicated chromatids, occurring only in S and G2 phases. It requires the resection of DNA ends at the break site.

A non-reversible cellular state, outside G1 phase, from which cells cannot re-enter the cell cycle.

Stochastic acquisition of genetic change over many cell divisions that can result in mutations and chromosomal rearrangements and aneuploidy.

(CIN). Type of genome instability involving structural and/or numerical chromosome aberrations.

Abnormal number of chromosomes in a cell.

Compounds that selectively kill cells that are in a state of senescence.

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Matthews, H.K., Bertoli, C. & de Bruin, R.A.M. Cell cycle control in cancer. Nat Rev Mol Cell Biol 23 , 74–88 (2022). https://doi.org/10.1038/s41580-021-00404-3

Download citation

Accepted : 21 July 2021

Published : 10 September 2021

Issue Date : January 2022

DOI : https://doi.org/10.1038/s41580-021-00404-3

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Esco2’s oncogenic role in human tumors: a pan-cancer analysis and experimental validation.

  • Dapeng Chen
  • Jianqiang Zhu

BMC Cancer (2024)

Synthesis, design, biological evaluation, and computational analysis of some novel uracil-azole derivatives as cytotoxic agents

  • Leila Emami
  • Fateme Zare
  • Razieh Sabet

BMC Chemistry (2024)

Targeted inhibition of the PI3K/AKT/mTOR pathway by (+)-anthrabenzoxocinone induces cell cycle arrest, apoptosis, and autophagy in non-small cell lung cancer

  • Xiao-Qian Li
  • Xiao-Ju Cheng

Cellular & Molecular Biology Letters (2024)

R-loop and diseases: the cell cycle matters

Molecular Cancer (2024)

CCDC58 is a potential biomarker for diagnosis, prognosis, immunity, and genomic heterogeneity in pan-cancer

  • Ruohan Zhang

Scientific Reports (2024)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

cancer cell research studies

  • Open access
  • Published: 14 January 2024

Scaffold-based 3D cell culture models in cancer research

  • Waad H. Abuwatfa 1 , 2 ,
  • William G. Pitt 3 &
  • Ghaleb A. Husseini   ORCID: orcid.org/0000-0002-7244-3105 1 , 2  

Journal of Biomedical Science volume  31 , Article number:  7 ( 2024 ) Cite this article

7163 Accesses

14 Citations

Metrics details

Three-dimensional (3D) cell cultures have emerged as valuable tools in cancer research, offering significant advantages over traditional two-dimensional (2D) cell culture systems. In 3D cell cultures, cancer cells are grown in an environment that more closely mimics the 3D architecture and complexity of in vivo tumors. This approach has revolutionized cancer research by providing a more accurate representation of the tumor microenvironment (TME) and enabling the study of tumor behavior and response to therapies in a more physiologically relevant context. One of the key benefits of 3D cell culture in cancer research is the ability to recapitulate the complex interactions between cancer cells and their surrounding stroma. Tumors consist not only of cancer cells but also various other cell types, including stromal cells, immune cells, and blood vessels. These models bridge traditional 2D cell cultures and animal models, offering a cost-effective, scalable, and ethical alternative for preclinical research. As the field advances, 3D cell cultures are poised to play a pivotal role in understanding cancer biology and accelerating the development of effective anticancer therapies. This review article highlights the key advantages of 3D cell cultures, progress in the most common scaffold-based culturing techniques, pertinent literature on their applications in cancer research, and the ongoing challenges.

Graphical Abstract

cancer cell research studies

Introduction

Cancer is a group of diseases characterized by the uncontrolled growth and spread of abnormal cells in the body. It is one of the leading causes of death worldwide, with millions of new cases and deaths each year [ 1 ]. Despite significant advances in cancer research and treatment over the years, the disease remains a major public health challenge and a substantial burden on patients, families, and society. Cancer research is crucial to develop new and effective treatments, improve patient outcomes, and find cures for this disease. As understanding of cancer biology and genetics continues to evolve, so do the approaches used to diagnose, treat, and prevent the disease. However, there is still much to learn about the complex mechanisms underlying cancer development and progression and the unique challenges posed by different types of cancers [ 2 ]. In addition, there is a need to develop more personalized and targeted therapies that can improve patient outcomes and minimize side effects. As such, cancer research must continue to innovate and advance to keep pace with the evolving understanding of the disease. This includes exploring new treatment modalities, developing more sophisticated diagnostic tools, and understanding the genetic and molecular mechanisms involved in its development and progression [ 3 ].

Two-dimensional (2D) cell culture is a commonly used technique to grow and maintain cells in the laboratory. Cancer research extensively uses it to study cells under controlled conditions, where they are grown on a flat surface supplied with a nutrient-rich liquid medium that provides the necessary nutrients for cell growth and survival. The growth medium used in cell culture varies depending on the type of cancer being studied and the desired goals of the study. One of the most critical aspects of cell culture for cancer research is maintaining cell viability and function, as cancer cells are highly susceptible to environmental changes [ 4 ]. Another challenge facing cell culture for cancer research is the ability to accurately model the complexity of human tumors. These are typically highly heterogeneous, comprising different cell types, including cancer, stromal, and immune cells. Understandably, 2D cell culture does not accurately mimic tumors’ three-dimensional (3D) environment [ 5 ]. The architecture and organization of cells in a 3D environment differ from those in a 2D environment, which can affect cell behavior and drug response. Recreating this complexity in a laboratory setting is difficult, as it requires the development of culture conditions that promote the growth and interaction of multiple cell types in a multifaceted environment [ 6 ]. Therefore, 3D cell culture models were developed as they offer sophisticated platforms that mirror the structural and functional complexities of in vivo tissues, providing valuable insights for cancer research and drug development. This review article highlights the key advantages of 3D cell cultures, the most common scaffold-based 3D culturing techniques, pertinent literature about applications in cancer research, and the challenges associated with these culturing techniques. Due to the topic’s vastness, this paper focuses on examining scaffold-based models of 3D cell cultures.

Physiological relevance of 3D cell cultures to the ECM

Tumors are complex structures composed of cancer cells, non-cancerous cells (i.e., immune cells, fibroblasts, endothelial cells, etc.), and various extracellular matrix (ECM) components. The ECM plays a crucial role and contributes to the hallmarks of cancer in tumor progression, metastasis, and response to therapy [ 7 , 8 ]. The ECM can (1) secrete growth factors and cytokines that promote cell proliferation and survival [ 9 ], (2) modulate the expression of genes involved in cell cycle regulation and apoptosis [ 10 ], (3) control the expression of telomerase, an enzyme that extends the telomeres of chromosomes, (4) secrete angiogenic factors that promote the formation of new blood vessels, thereby providing the tumor with the nutrients and oxygen it needs to grow [ 11 ], (5) promote the epithelial-to-mesenchymal transition (EMT), a process by which epithelial cells acquire the ability to migrate and invade other tissues [ 12 ], and (6) temper the immune response by influencing the recruitment and function of immune cells in the TME [ 13 ]. Romero-López and colleagues [ 14 ] tested how the ECM derived from normal and tumor tissues impacted blood vessels and tumor growth using reconstituted ECM. Tumor tissue obtained from liver metastases of colon tumors was subjected to hematoxylin and eosin (H&E) staining to confirm the successful decellularization of both colon and tumor tissues. Subsequently, significantly distinct protein composition and stiffness were observed among the reconstituted matrices, leading to notable variations in vascular network formation and tumor growth in both in vitro and in vivo. Fluorescence Lifetime Imaging Microscopy was employed to evaluate the free/bound ratios of the nicotinamide adenine dinucleotide (NADH) cofactor in tumor and endothelial cells to indicate cellular metabolic state. Notably, cells seeded in tumor ECM exhibited elevated levels of free NADH, indicating an increased glycolytic rate compared to those seeded in normal ECM. These findings underscore the substantial influence of ECM on cancer cell growth and the accompanying vasculature (e.g., increased vessel length, increased vascular heterogeneity). Alterations in the composition of tumor ECM, such as augmented deposition and crosslinking of collagen fibers, can be attributed to communication between tumor cells and tumor-associated stromal cells.

Every tissue type has a distinct ECM composition, topology, and organization [ 15 ]. These factors play a significant role in controlling cell function, behavior, and interactions with the microenvironment, as they generate spatial gradients of biochemicals and metabolites that, in turn, may elicit distinctive cell-mediated responses (e.g., differentiation, migration) [ 16 ]. Langhans [ 17 ] analyzed the chemical components of ECM and reported that it contains water, carbohydrates, and proteins, such as fibrous matrix proteins, glycoproteins, proteoglycans, glycosaminoglycans, growth factors, protease inhibitors, and proteolytic enzymes. Thus, ECM organization can influence cell genotypes and phenotypes, where such effects can be explored through 3D cell cultures [ 16 , 18 ]. For example, variations in the gene and protein expression and activity of the epidermal growth factor receptors (EFGR), phosphorylated protein kinase B (phospho-AKT), and p42/44 mitogen-activated protein kinases (phospho-MAPK) in colorectal cancer cell lines (e.g., HT-29, CACO-2, DLD-1) affected the genotype and phenotype of cells in 3D cultures, as compared to 2D monolayers [ 19 , 20 ]. Moreover, the ECM can influence cell morphology and expression of chemokine receptors. Kiss et al. [ 21 ] showed that 3D cultured prostate cancer cells (e.g., LNCaP, PC3) exhibited a high level of interaction between the cells and ECM, which resulted in the upregulation and overexpression of the CXCR7 and CXCR4 chemokine receptors. While 2D cell culture has been the mainstay of laboratory cancer research, it has become increasingly clear that this approach is inadequate in replicating the in vivo conditions that cells experience in the human body. As a result, researchers have been turning to 3D cell culture as a more physiologically relevant model for studying cellular processes and disease. A key advantage of 3D models for cancer research is that they can better mimic the complex microenvironment of tumors, including tumor morphology and topography, upregulation of pro-angiogenic proteins, dispersion of biological and chemical factors, cell–cell and cell–matrix interactions, gradients of oxygen and nutrients, and a more realistic ECM composition [ 6 , 22 , 23 ]. Necrotic, hypoxic, quiescent, apoptotic, and proliferative cells are often found in spheroid cell clusters at different phases of development [ 24 ]. Since the outer layer of the spheroid has greater exposure to the nutrient-supported medium, it contains a higher number of proliferating cells. Cells in the spheroid core are hypoxic and often quiescent as they receive less oxygen, growth agents, and nutrients from the media. This results in more physiologically relevant gradients in tissue composition that can better inform drug discovery and development [ 24 ]. Furthermore, 3D cell culture accurately depicts the cellular response to drugs and other therapeutic agents. Such a model’s spatial and physical characteristics influence the transmission of signals between cells, which alters gene expression and cell behavior [ 25 ]. Loessner et al. [ 26 ] demonstrated a flexible 3D culture method where a synthetic hydrogel matrix with crucial biomimetic properties provided a system for studying cell–matrix dynamics related to tumorigenesis. The 3D cultured cells overexpressed mRNA for receptors on their surface (e.g., protease, α3, α5, β1 integrins) compared to 2D cultured cells. Moreover, spheroid progression and proliferation depended on the cells’ ability to proteolytically transform their ECM and cell-integrin interactions. Consequently, the 3D spheroids showed higher survival rates in contrast to 2D monolayers after exposure to the chemotherapeutic agent paclitaxel, which indicates that it better stimulates in vivo chemosensitivity and pathophysiological events. Table 1 below summarizes studies using different 3D models to investigate different types of cancers.

Figure  1 summarizes the main characteristics of 2D and 3D cell cultures. The shift to 3D cell culture is a significant advancement in laboratory research, as it provides a more physiologically relevant model for studying cellular processes and disease. While some challenges remain to be addressed, the advantages of 3D culture outweigh the limitations of 2D culture. As technology continues to evolve, 3D culture is likely to become an increasingly crucial tool in cancer research and other fields of biomedical science.

figure 1

Characteristics of 2D and 3D cell cultures

Table 2 below provides a comprehensive overview of 2D, 3D, and other model systems employed in cancer research. Besides 2D and 3D cell cultures, tissues and organs present structural and functional intricacies, capturing organ-specific responses but posing challenges in maintenance and accessibility. Furthermore, model animals mimic in vivo systemic responses, yet ethical concerns, high costs, and species differences limit their utility. While clinically relevant, patient-derived samples present challenges in experimental control and sample heterogeneity [ 43 ]. It is noteworthy to highlight the difference between spheroids and organoids as both are commonly used terms within the scope of 3D cell cultures [ 44 ]. Organoids and spheroids are different 3D cell culture models that can be cultured with different techniques. Organoids, characterized by intricate structures replicating real organs or tissues, are composed of multiple cell types that self-organize to mirror tissue-like architecture, deriving from stem cells or tissue-specific progenitors. Due to their high biological relevance, they find applications in disease modeling, drug testing, and understanding organ development. Beyond organoids, tumoroids (i.e., tumor-like organoids), derived from patient cancer tissues containing tumor and stroma cells of the TME, are becoming advanced 3D culture platforms for personalized drug evaluation and development. In contrast, spheroids are simpler spherical cellular aggregates lacking the distinct organ-like structures of organoids. Comprising one or multiple cell types, spheroids are used to study fundamental cellular behaviors and drug responses in a 3D environment. While both contribute to 3D cell culture studies, organoids closely resemble real organs compared to the simpler cellular aggregates represented by spheroids [ 44 ]. Patient models are valuable tools that aim to replicate the complexities of human tumors, providing insights into disease mechanisms, therapeutic responses, and personalized treatment strategies. They can be utilized in Patient-Derived Xenografts (PDX), organoids and 3D cultures, patient-derived cell lines, liquid biopsies, and clinical trials [ 45 ].

Cell sources and 3D culture heterogeneity

In 3D cell culture, achieving an optimal balance between homogeneity and heterogeneity is intricately linked to the cellular source, considering stem cells, induced Pluripotent Stem Cells (iPSCs), or mixed primary cells derived from tissues [ 46 ]. Stem cells in in vitro cell culture encompass embryonic stem cells (ESCs), induced pluripotent stem cells (iPSCs), and adult or somatic stem cells. Embryonic stem cells exhibit high pluripotency, capable of differentiating into any cell type, but their use raises ethical concerns due to their origin from embryos. iPSCs are generated from somatic cells (e.g., skin or blood cells) through reprogramming, reverting them to an embryonic-like pluripotent state, but face reprogramming efficiency and potential tumorigenicity challenges. This transformation creates an extensive and diverse reservoir of human cells, capable of developing into any cell type required for therapeutic applications. Human-induced pluripotent Stem Cells (HiPSCs) are particularly relevant in cancer research (Table  3 ) [ 47 ]. Thus, the reprogramming process pioneered by Shinya Yamanaka has opened new avenues for advancing cancer biology, drug discovery, and regenerative medicine in cancer treatment. Lastly, adult or somatic stem cells are tissue-specific, mirroring the characteristics of their origin, and present fewer ethical concerns as they are derived from adult tissues. However, they have limited differentiation potential and a finite lifespan in culture. The selection of the cell source significantly influences the composition and behavior of the 3D culture. Stem cells and iPSCs, known for their pluripotency, introduce an inherent heterogeneity due to their ability to differentiate into various cell types [ 45 , 46 ].

Furthermore, primary cells, derived directly from living organisms, possess unique characteristics that make them invaluable for in vitro studies. Maintaining biological relevance, these cells closely mimic the tissue or organ from which they are isolated, reflecting the intricacies of in vivo conditions. With donor-specific variability, primary cells allow researchers to explore genetic diversity’s impact on cell behavior, disease susceptibility, and drug responses. Retaining tissue-specific functions, differentiated primary cells are crucial for studying specific physiological processes and diseases associated with particular tissues [ 46 ]. However, these cells have challenges, including a limited lifespan and sensitivity to culture conditions. The finite replicative capacity and sensitivity contribute to the heterogeneity observed in 3D cell cultures, emphasizing the importance of carefully considering culture conditions and donor-specific variations to accurately represent in vivo scenarios. Despite these challenges, primary cells are vital in advancing our understanding of cell biology, disease mechanisms, and therapeutic development. Similarly, using mixed primary cells derived from tissues can contribute to a more heterogeneous cellular composition, resembling the complexity found in native tissues. Striking the right balance is crucial, as an excessive degree of heterogeneity may obscure specific responses, while too much homogeneity might oversimplify the representation of the tissue microenvironment. Therefore, a nuanced understanding of the cellular source is essential for tailoring 3D cell culture models to accurately reflect the intricacies of actual tissues and organs.

Scaffold-based techniques for 3D cell culture

As explained above, developing 3D cell culture techniques that more accurately model the TME is a major area of focus in cancer research [ 6 , 59 ]. Different approaches for 3D cell cultures exist and can be generally divided into scaffold-based and scaffold-free methods. Scaffold-free 3D cell culture refers to a cell culture technique in which cells are cultured and assembled into 3D structures without external scaffold material. Instead of being embedded within a supportive matrix, the cells self-assemble and interact with neighboring cells to form 3D tissue-like structures. Such cultures allow for more accurate cell–cell interactions, spatial organization, and physiological responses, making them valuable tools for various applications, including drug testing. They also usually have higher cell densities than scaffold-based models, which can influence cellular behavior, gene expression, and cellular functions. Lastly, non-scaffold models offer versatility and customizability in terms of cell types, culture conditions, and experimental designs. However, it is essential to consider that scaffold-free approaches might have limitations in providing mechanical support, shape control, and reproducibility compared to scaffold-based 3D cell culture methods [ 60 ]. As such, researchers often select the appropriate 3D cell culture method based on their specific research goals and the tissue or organ system they aim to model or engineer. Due to the topic’s vastness, the paper’s purviews' are limited to the examination of scaffold-based models of 3D cell cultures. Scaffolds are essential components in 3D cell culture systems, as they provide a 3D environment for cells to grow and interact with each other and their surroundings [ 61 , 62 ]. Biomaterials employed in such models can be categorized into the following primary groups: polymer scaffolds, hydrogels, decellularized tissue scaffolds, and hybrid scaffolds (e.g., incorporating microfluidic devices). Tables 4 and 5 summarize the advantages and limitations of commonly used scaffold-free and scaffold-based 3D cell culture techniques, respectively.

Polymer-based scaffolds

Polymer scaffolds revolutionize 3D cell culture by providing a biomimetic environment imitating the natural ECM, fostering cell proliferation and differentiation, often with remarkable efficiency and precision. These scaffolds offer a versatile platform for studying complex cell behaviors and hold immense promise in cancer research applications. They can be generally classified as natural or synthetic-derived (see Fig.  2 ). Natural polymer scaffolds are made from naturally occurring polymers. They can be processed into various forms, including fibers, films, or porous structures. They can be further classified into two main categories: protein-based and polysaccharide-based scaffolds. Protein-based scaffolds are derived from large molecules composed of amino acids (e.g., collagen, silk, gelatin, fibronectin [ 93 , 94 ]). Due to their bioactive properties, these scaffolds provide cell adhesion sites and can regulate cell behavior and tissue development. A 3D cell culture platform using collagen scaffolds was developed to investigate the tumorigenicity of cancer stem cells (CSCs) in breast cancer [ 95 ]. The study revealed that the 3D cell culture system demonstrated increased expression of pro-angiogenic growth factors, indicating a potential role in promoting blood vessel formation. Moreover, the overexpression of CSC markers such as OCT4A and SOX2, as well as breast cancer stem cell markers including SOX4 and JAG1, was observed in the 3D scaffolds, suggesting that the 3D model successfully replicated the molecular characteristics associated with CSCs. In terms of behavior, the 3D model more closely mimicked the characteristics of CSCs compared to an in vivo model, indicating its effectiveness in capturing the tumorigenic properties of CSCs. Therefore, the collagen scaffold-based 3D cell culture platform provided a valuable tool for studying CSC tumorigenicity in breast cancer, demonstrating the upregulation of pro-angiogenic growth factors, the overexpression of CSC and breast cancer stem cell markers, and a close resemblance to CSC behavior when compared to an in vivo model. Another study by McGrath et al. [ 96 ] used a 3D collagen matrix (GELFOAM™) to create an endosteal bone niche (EN) model, referred to as 3D-EN, for studying breast cancer cells’ quiescence and dormancy behaviors. The 3D-EN model effectively facilitated the identification of several genes associated with dormancy-reactivation processes, where among the tested cell lines, only MDA-MB-231 cells exhibited dormancy behavior, suggesting that they have a propensity for entering a dormant state in the simulated physiological conditions.

figure 2

Classification of polymers used for fabricating polymer-based 3D cell culture scaffolds

On the other hand, polysaccharide-based scaffolds are composed of long chains of sugar molecules (e.g., chitosan and hyaluronic acid). They are biocompatible, biodegradable, and can often be modified to adjust their physical and biological properties. Arya et al. [ 97 ] developed a 3D cell culture model using a chitosan scaffold, a natural polymer derived from chitin, to study breast cancer behavior. The scaffold was cross-linked with genipin, a natural cross-linker, to enhance its stability. The study found that the chitosan–gelatin (GC) scaffold provided a suitable environment for the growth of MCF-7 breast cancer cells, with the cells showing good adhesion and proliferation. The scaffold also supported the formation of cell clusters, which are more representative of in vivo tumor conditions compared to 2D cultures. The study concluded that the chitosan/gelatin scaffold could be useful for studying breast cancer in vitro, providing a more physiologically relevant model than traditional 2D cultures. GC scaffolds have been shown to support the formation of tumoroids that mimic tumors grown in vivo , making them an improved in vitro tumor model. These scaffolds have been successfully used to study lung cancer, as well as other types of cancer, such as breast, cervix, and bone [ 98 ]. These scaffolds have demonstrated gene-expression profiles similar to tumors grown in vivo, indicating their potential for studying cancer progression and drug screening for solid tumors [ 99 ]. The GC scaffolds have also been shown to improve the predictivity of preclinical studies and enhance the clinical translation of therapies [ 100 ]. Overall, the GC scaffolds provide a valuable tool for studying tumor development and evaluating the efficacy of anti-cancer drugs in an in vitro setting.

Synthetic polymer scaffolds (e.g., polylactic acid (PLA), polyglycolic acid (PGA), and polycaprolactone (PCL) can be tailored to have specific mechanical and biochemical properties. However, they can be less biocompatible than natural polymers and may require surface modifications to promote cell attachment and growth [ 60 ]. Palomeras et al. [ 101 ] tested the efficiency of 3D-printed PCL scaffolds for the culture of MCF7 breast cancer cells. The researchers found that the scaffold’s design, specifically the deposition angle, significantly influenced cell attachment and growth. Scaffolds with a deposition angle of 60° showed the highest cell counting after treatment with trypsin. Furthermore, the study found that the 3D culture in PCL scaffolds enriched the cancer stem cell (CSC) population compared to 2D culture control, increasing their Mammosphere Forming Index (MFI). The study concluded that 3D PCL scaffold culture could spur MCF7 cells to generate a cell population with CSC properties. This suggests its potential for studying CSC properties and screening new therapeutic agents targeting CSC populations. These efforts highlight the potential of natural polymer scaffolds in creating more physiologically relevant 3D cell culture models for cancer research. Using these scaffolds can enhance the understanding of cancer cell behavior and potentially lead to the discovery of more effective therapeutic strategies. Similarly, Rijal et al. [ 88 ] utilized modified gas foaming-based synthetic polymer scaffolds from poly(lactic-co-glycolic) acid (PLGA) and PCL for conducting 3D tissue cultures and animal models in breast cancer research. The research group investigated the response of MDA-MB-231 cells to anticancer drugs, their viability, morphology, proliferation, receptor expression, and ability to develop in vivo tumors using the 3D scaffolds. MDA-MB-231 cells were cultured on PLGA-coated 2D microscopic glass slides and in 3D-porous PLGA scaffolds to examine cancer cells’ survival on the polymeric substrata. The number of dead cells detected on the PLGA-coated glass slides and PLGA 3D scaffolds was negligible on Day 1. However, a significant increase in the number of dead cells was observed on the PLGA-coated glass slides compared to the 3D scaffolds on day 14. Additionally, the expression of ECM proteins and cell surface receptors on the synthetic polymers was investigated, where strong staining signals of type I collagen and integrin α2 were detected in both cell types using immunofluorescence (IF) microscopy. It is worth noting that integrin α2β1, which acts as a primary receptor for type I collagen, displayed a basal expression level in the 3D model. This expression pattern may promote breast cancer cell migration and tumor growth, as high levels of the integrin receptor tend to inhibit cancer cell migration. Notably, integrin α2 receptors showed a prominent colocalization with type I collagen, particularly around the cell edges, suggesting local deposition of type I collagen and subsequent binding of integrin α2 receptors, facilitating cell attachment and migration. Lastly, to evaluate the tumor formation capabilities of the polymeric porous scaffolds in mice, MDA-MB-231 cells were coated onto porous PLGA scaffolds and implanted into the mammary fat pads of NOD/SCID mice. Blank scaffolds without cells served as the negative control. As anticipated, the proliferating cell nuclear antigen biomarker Ki-67 was not detected in the blank scaffold implants. At the same time, its expression was significantly high within the tumors derived from the MDA-MB-231 cell-laden PLGA scaffolds. This finding suggested that the cancer cell population within the scaffolds exhibited rapid proliferation when embedded in the native breast tissues.

Hydrogel scaffolds

Hydrogels are 3D networks of hydrophilic polymers (can be natural, synthetic, or hybrid), that can absorb large amounts of water or biological fluids while maintaining their structural integrity [ 102 ]. Figure  3 shows common techniques for culturing with hydrogel scaffolds. In the dome technique (see Fig.  3 A), cells are mixed with temperature-sensitive hydrogels and then seeded as droplets within a cell culture vessel. This technique relies on careful temperature control to allow the hydrogel to polymerize and form a dome structure. Once the hydrogel has polymerized and the cell-hydrogel droplet is stabilized, it is delicately covered with cell culture media. This allows for a localized 3D cell culture in a larger vessel and can create multiple individual cell clusters or spheroids in a single plate. However, the maintenance of dome integrity can be challenging over time and might be affected by changes in temperature or physical disturbance. Also, it may not be suitable for long-term culture or cells requiring complex structural support due to the relatively simple and isolated 3D structure. Figure  3 B illustrates the insert wells technique, which consists of porous inserts to hold the cell-hydrogel mixture while cell culture media is added to the well surrounding the insert. This separation creates a differential environment, allowing for nutrient exchange while maintaining a distinct 3D culture within the insert. Heterogeneous spheroids will eventually form on the insert bottom due to gravitational pull and cell–cell interactions. Such a model can be used to study cell invasion or migration by placing the cell-hydrogel mixture on one side of a permeable membrane and chemo-attractants on the other. The gel-bottom support method (see Fig.  3 C) involves creating a thick layer of hydrogel at the bottom of a culture well, on top of which the cell suspension is placed. For instance, this method can be used for embedding cells within macroporous hydrogel scaffolds, such as AlgiMatrix ® (Thermo Fisher Scientific/Life Technologies, Carlsbad, USA)—an ionically gelled and dried scaffold that is conveniently provided in sterile pre-loaded disc format in standard cell culture well plates [ 103 , 104 ]. To initiate the cell culture, a concentrated cell suspension in culture media is seeded on top of the hydrogel, where it is subsequently absorbed, resulting in the entrapment of the cells within the porous structure of the hydrogel. Lastly, in the embedding technique (see Fig.  3 D), the cells are mixed with a hydrogel and directly placed at the bottom of a culture vessel, followed by a layer of culture media, allowing the cells to grow within the matrix of the hydrogel, thereby more accurately mimicking the in vivo 3D environment. This technique is beneficial for studying cell–cell and cell–matrix interactions, invasion, migration, and drug responses. However, it can be more technically challenging to embed cells evenly throughout the hydrogel; retrieving cells from the matrix for downstream analysis can be challenging. The permeability of the hydrogel to nutrients, gases, and wastes may need careful optimization to avoid creating a hypoxic environment or nutrient deprivation for cells located in the interior of the gel. Each of these methods must be selected based on the needs of the specific experiment and the type of cells being cultured. Additionally, the hydrogel composition and mechanical properties should be tuned according to the native ECM properties of the cell type of interest.

figure 3

Common methods of hydrogel 3D cultures: A the dome technique: cells are mixed with temperature-sensitive hydrogels then seeded as droplets in the cell culture vessel, then carefully covered with media, B insert wells: media is added in the well whereas cell suspension (cell in hydrogel mix) is placed in the insert, then covered with another layer of media. Heterogeneous spheroids will form on the insert bottom, C gel-bottom support: the bottom of the well is covered with a thick layer of hydrogel, on top of which the cell suspension is placed, and D embedding technique: cells mixed with hydrogel are placed on the bottom and then covered with a layer of media to support spheroid growth in the matrix

Due to their adjustable properties, synthetic hydrogels offer notable benefits in 3D cell culture. The RADA16-I peptide is a self-assembling peptide derived from a segment of Zuotin, a left-handed Z-DNA-binding protein originally discovered in yeast. This peptide has emerged as a novel nano-biomaterial due to its ability to form nanofiber scaffolds. Consequently, these scaffolds provide a supportive framework that promotes cell growth and fosters a conducive 3D milieu for cell culture. The peptide sequence can be modified to incorporate specific functional groups, thus fine-tuning the mechanical, chemical, and biological attributes of the resultant scaffold. This remarkable flexibility enables customization to align precisely with the unique demands of the cultured cells or the intended experimental objectives. These scaffolds, which are about 10 nm in diameter, are driven by positively and negatively charged residues through complementary ionic interactions. When dissolved in water, the RADA16-I peptide forms a stable hydrogel (nanofiber networks with pore sizes of about 5–200 nm) with extremely high water content at concentrations of 1–5 mg/mL, which closely mimics the porosity and gross structure of ECMs, making it suitable for the fabrication of artificial cell niches for applications in tumor biology. Yang and Zhao [ 105 ] prepared a RADA16-I peptide hydrogel that provided an elaborate 3D microenvironment for ovarian cancer cells in response to the surrounding topography. The 3D cell cultures exhibited a two to five-fold increase in drug resistance (paclitaxel, curcumin, and fluorouracil) compared to the 2D monolayers, which showed a good representation of the primary tumor and were likely to simulate the in vivo biological characteristics of ovarian cancer cells. Similarly, Song et al. [ 106 ] also proved that RADA16-I hydrogels can provide prominent and dynamic nanofiber frameworks to sustain robust cell growth and vitality. HO-8910PM cells, metastatic human ovarian cancer cells, were cultured in three hydrogel biomaterials, namely RADA16-I hydrogel, Matrigel, and collagen I. The specially designed RADA16-I peptide exhibited a well-defined nanofiber network structure within the hydrogel, providing a nanofiber-based cellular microenvironment similar to Matrigel and collagen I. Notably, the HO-8910PM cells exhibited distinctive growth patterns within the three matrices, including cell aggregates, colonies, clusters, strips, and multicellular tumor spheroids (MCTS). Moreover, the molecular expression of integrin β1, E-cadherin, and N-cadherin in 3D-cultured MCTS of HO-8910PM cells was elevated, and their chemosensitivity was reduced to cisplatin and paclitaxel in comparison to the 2D cell culture, evidenced by IC 50 values and inhibition rates.

Furthermore, polyvalent hyaluronic acid (HA) hydrogels are considered synthetic, as they are typically created through chemical modification of HA molecules, introducing crosslinking agents or functional groups that enable the formation of a gel-like structure. This modification allows for control over the physical and mechanical properties of the hydrogel, such as its stiffness, degradation rate, and bioactivity. Suo et al. [ 107 ] engineered an ECM-mimicking hydrogel scaffold to replicate the native breast cancer microenvironment to provide an effective in vitro model for studying breast cancer progression. HA hydrogels from polyvalent HA derivatives were prepared through an innovative dual crosslinking process involving hydrazone and photo-crosslinking reactions. Hydrazone crosslinking is a versatile, reversible process that allows for rapid gelation, while photo-crosslinking stabilizes the formed hydrogel. Using this approach, they could efficiently produce HA hydrogels in under 120 s. It was found that the developed HA hydrogels closely resembled the topography and mechanical properties of breast tumors, and their characteristics (i.e., rigidity and porosity) could be fine-tuned by adjusting the amount of aldehyde-HA in the hydrogel formulation. This ability to modulate the mechanical properties of the hydrogels opens up possibilities for modeling different stages of tumor progression or different types of tumors. Moreover, a critical feature of the developed HA hydrogels was their dual-responsive degradation behavior, which was found to be responsive to glutathione and hyaluronidase. The glutathione responsiveness allows for degradation in response to the redox environment, which is often disturbed in cancer cells. Meanwhile, responsiveness of hyaluronidase makes the hydrogels sensitive to an enzyme that is typically upregulated in invasive cancer cells. Significantly, the HA hydrogel-cultured MCF-7 cells displayed upregulated expression of vascular endothelial growth factor (VEGF), interleukin-8 (IL-8), and basic fibroblast growth factor (bFGF) compared to their 2D cultured counterparts. These molecules are key mediators of angiogenesis and inflammation in cancer, suggesting that the HA hydrogel environment better replicates the conditions that promote these processes in tumors. Besides, the hydrogel-cultured cells exhibited enhanced migration and invasion abilities, which are key hallmarks of aggressive cancer cells. In vivo studies supported these results and confirmed the superior tumorigenic capacity of the MCF-7 cells cultured in HA hydrogels compared to those cultured in 2D. The outcomes of this research are anticipated to have far-reaching implications for both the in vitro study of breast cancer and the development of effective therapeutic strategies.

Another investigation by Wang et al. [ 108 ] supported that the level of methacrylation significantly influenced the hydrogel’s microstructure, mechanical characteristics, and capacity for liquid absorption and degradation. The refined hydrogel, synthesized through the photocrosslinking of methacrylated HA, displayed a highly porous structure, a high equilibrium swelling ratio, appropriate mechanical properties, and a degradation process responsive to hyaluronidase. It was found that the HA hydrogel promoted the growth and proliferation of MCF-7 cells, which formed aggregates within the hydrogel. In addition, 3D-cultured MCF-7 cells showed an increased expression of VEGF, bFGF, and interleukin-8, and enhanced invasion and tumorigenesis capabilities compared to their 2D-cultured counterparts. As such, the HA hydrogel has proven to be a dependable alternative for constructing tumor models. Gelatin methacryloyl (GelMA) is another commonly used natural biomaterial for 3D hydrogel scaffolds in cancer research. GelMA is derived from gelatin, a natural protein obtained from collagen-rich sources. It is modified by adding methacryloyl groups that enable it to undergo photocrosslinking when exposed to ultraviolet (UV) light. This property allows GelMA to form stable hydrogel networks, making it suitable for creating 3D scaffolds that mimic the tumor microenvironment (TME). The tunable mechanical and biochemical properties of GelMA hydrogels, biocompatibility, and ability to support cell growth make them valuable tools for studying cancer cell behavior, tumor invasion, drug screening, and other aspects of cancer research. Kim et al. [ 109 ] developed a 3D cell culture model for the bladder by employing a novel acellular matrix and bioreactor. GelMA was utilized as a 3D scaffold for the bladder cancer cell culture, with an optimal scaffold height of 0.08 mm and a crosslinking time of 120 s [ 110 ]. Subsequently, 5637 and T24 cells were cultured in 2D and 3D environments and subjected to rapamycin and Bacillus Calmette-Guérin (BCG) drug treatments. It was found that the 3D bladder cancer cell culture model exhibited a faster establishment process and greater stability when compared to the 2D model. Moreover, the 3D-cultured cancer cells demonstrated heightened drug resistance and reduced sensitivity compared to the 2D-cultured cells. Additionally, the researchers observed cell-to-cell interaction and basal activity in the 3D model, closely resembling the in vivo environment.

Along the same lines, Arya et al. [ 111 ] investigated the suitability of GelMA hydrogels as in vitro 3D culture systems for modeling key characteristics of metastatic progression in breast cancer, specifically invasiveness and chemo-responsiveness. The mechanical and morphological properties of the hydrogels were tuned by varying the percentage of GelMA used. Compression testing revealed that the stiffness of 10% GelMA hydrogels was within the range reported for breast tissue, making them suitable matrices for mimicking the breast viscoelasticity in vitro, as cells cultured on 10% GelMA hydrogels exhibited a higher proliferation rate compared to 15% GelMA in both cell lines tested, making them robust systems for long-term cell culture. Furthermore, proliferation studies showed that the GelMA hydrogels could sustain breast cancer cells longer than 2D cultures. Overexpression of genes associated with invasiveness was also observed in 3D cultured breast cancer cells, suggesting potential changes important for metastatic progression. The response to chemotherapeutic drugs was evaluated, and it was observed that 3D spheroids of breast cancer cells cultured on GelMA hydrogels exhibited decreased sensitivity to taxane drugs like paclitaxel. The study highlighted the importance of an adequate matrix pore size for cell penetration, migration, proliferation, exchanging oxygen, nutrients, and waste materials in and out of the 3D culture scaffolds. Significantly, these studies emphasized the importance of the 3D cancer cell culture model in establishing a patient-like model. Utilizing such models can achieve a more precise evaluation of drug responses, potentially leading to advancements in cancer treatment and other diseases.

Cells are known to respond to their mechanical environment in a process known as mechano-transduction, where they transmute mechanical stimuli into biochemical signals, subsequently prompting alterations in cellular behavior and functional operations. Curtis et al. [ 112 ] investigated the influence of mechanical stimuli on the cell proliferation, growth, and protein expression of 4T1 breast cancer cells, serving as a model for cells that metastasize to bone. The researchers used 4T1 breast cancer cells and implanted them in gelatin-mTGase hydrogels that mimicked the mechanical properties of bone marrow. The hydrogels had different moduli of either 1 or 2.7 kPa. The cells were cultured under different conditions, including static culture, perfusion of media through the hydrogel, and combined perfusion with cyclic mechanical compression for 1 h per day for 4 days. Control samples were cultured under free-swelling conditions. Immunostaining techniques were used to analyze the protein expression within the cell spheroids formed during the culture. The study found that mechanical stimuli significantly influenced the behavior of the 4T1 breast cancer cells. The cells formed spheroids during the culture period, with larger spheroids observed in statically cultured constructs than those exposed to perfusion or compression. In the stiffer gelatin, compressed constructs resulted in smaller spheroids compared to perfusion alone, while compression had no significant effect in the softer gelatin. The immunostaining revealed the expression of proteins associated with bone metastasis within the spheroids, including osteopontin, parathyroid hormone-related protein, and fibronectin. The proliferative marker Ki67 was present in all spheroids on day 4. The intensity of Ki67 staining varied depending on the culture conditions and gelatin stiffness. It highlighted the mechanical sensitivity of 4T1 breast cancer cells and demonstrated how mechanical stimuli can impact their proliferation and protein expression within soft materials that mimic the mechanical properties of bone marrow. The findings emphasized the role of the mechanical environment in the bone for both in vivo and in vitro models of cancer metastasis.

Understanding the influence of mechanical factors on cancer cell behavior is crucial for developing effective strategies to prevent and treat metastasis to bone, potentially leading to improved clinical outcomes for patients with advanced cancer. Similarly, Cavo et al. [ 113 ] investigated the impact of substrate elasticity on breast adenocarcinoma cell activity using mechanically tuned alginate hydrogels. The study evaluated the viability, proliferation rates, and cluster organization of MCF-7 breast cancer cells in 3D alginate hydrogels compared to standard 2D environments. The elastic moduli of the different alginate hydrogels were measured using atomic force microscopy (AFM). The results demonstrated that substrate stiffness directly influenced cell fate in 2D and 3D environments. In the 3D hydrogels with an elastic modulus of 150–200 kPa, the MCF-7 cells exhibited uninhibited proliferation, forming cell clusters with 100 μm and 300 μm diameters after 1 and 2 weeks, respectively. This unimpeded cell growth observed in softer hydrogels mimicked the initial stages of solid tumor pre-vascularization and growth. Furthermore, the multicellular, cluster-like conformation observed in the 3D hydrogels closely resembled the in vivo organization of solid tumors, demonstrating the advantage of 3D cancer models for replicating cell–cell and cell–matrix interactions. The study also highlighted the influence of microenvironment dimensionality on cellular morphology, as cells displayed a flat shape in 2D cultures while adopting a round shape in the 3D environment. Cell proliferation in the 3D setting depended highly on substrate stiffness, which impacted nutrient diffusion and intracellular signaling through a mechano-transduction mechanism. The findings underscore the importance of considering substrate stiffness in the design of 3D cancer models, as it directly affects cell viability, proliferation, and organization. By understanding the relationship between substrate stiffness and cellular behavior, researchers can develop more realistic in vitro models that better mimic the microenvironment of solid tumors. These models can advance our understanding of cancer development and aid in the development of targeted therapies by allowing for the investigation of cell–cell and cell–matrix interactions in a more accurate setting.

Decellularized tissue scaffolds

Decellularized tissues have had their cellular components removed, leaving behind the ECM. Decellularized tissues can be used as scaffolds for 3D cell culture, providing a natural environment for cells to grow and interact [ 114 ]. The use of decellularized tissues as 3D cell culture scaffolds offers several advantages. Firstly, they retain the intricate ECM composition, including structural proteins, growth factors, and signaling molecules, which play critical roles in cell behavior and tissue organization. This enables cancer cells to interact with the ECM more akin to in vivo conditions, influencing their adhesion, migration, invasion, and differentiation. Moreover, decellularized tissues offer spatial organization and architectural cues that guide cellular behavior. Preserving tissue-specific topography, such as vasculature, allows for studying angiogenesis and vascularization processes in cancer progression. These scaffolds also provide mechanical support and stiffness that influence cellular mechanotransduction, impacting cell morphology, proliferation, and gene expression patterns. They can be derived from various sources, including solid organs, such as the liver or lung, or specific tissue compartments, such as the ECM-rich decellularized basement membrane (see Fig.  4 ).

figure 4

Preparation methods, characterization techniques, and sources of decellularized tissues used as scaffolds for 3D cell culture. SEM: scanning electron microscopy; AFM: atomic force microscopy; FTIR: Fourier-transform infrared spectroscopy

Landberg et al. [ 115 ] hypothesized that using a pre-clinical platform based on decellularized patient-derived scaffolds as growth substrates to account for hidden clinically relevant information and aid in modeling the individualized properties of microenvironments could be optimized for personalized treatment planning. Different decellularization techniques, such as chemical, physical, or enzymatic methods, remove cellular components while preserving the ECM integrity (see Table  6 ) [ 116 ]. The choice of decellularization method depends on the tissue type, desired scaffold characteristics, and the specific requirements of the study. Combinations of different techniques may also be employed to achieve optimal decellularization outcomes. However, challenges remain in the field. The immunogenicity and biocompatibility of decellularized tissues must be carefully considered to prevent adverse reactions when introducing foreign matrices into cell culture systems. Standardization and reproducibility of decellularization protocols are also crucial to ensure consistency across studies and facilitate comparison of results. Integration with advanced technologies, such as microfluidics or organ-on-a-chip systems, can further enhance the functionality and relevance of decellularized tissue models.

D’Angelo et al. [ 117 ] developed a more representative 3D model of colorectal cancer liver metastasis using patient-derived scaffolds. These scaffolds, created by decellularizing tissue-specific ECM, retain the metastatic microenvironment’s biological properties and structural characteristics. The HT-29 CRC cell line was cultured within these scaffolds, obtained explicitly from cancer patients. The study observed increased cell proliferation and migration in the cancer-derived scaffolds, highlighting their ability to provide a more conducive environment for tumor cell growth and spreading. Furthermore, the 3D culture system demonstrated a reduced response to chemotherapy. HT-29 cells cultured in the cancer-specific 3D microenvironments showed decreased sensitivity to treatment with 5-fluorouracil and a combination of 5-fluorouracil with Irinotecan, when used at standard IC50 concentrations. The use of patient-derived scaffolds allows for the study of colorectal cancer metastasis progression and the assessment of their response to chemotherapy agents, to develop new therapeutic strategies and personalized treatments. Additionally, it provides an opportunity to identify potential prognostic biomarkers and therapeutic targets specific to peritoneal metastasis. Varinelli et al. [ 118 ] conducted a study that employed a tissue-engineered model for investigating peritoneal metastases (PM) in vitro, yielding similar conclusions. The model involved seeding PM-derived organoids onto decellularized extracellular matrices (dECMs) sourced from the peritoneum, enabling the exploration of intricate interactions between neoplastic cells and the ECM in the PM system. Both neoplastic peritoneum and corresponding normal peritoneum tissues were utilized to generate 3D-dECMs. Utilizing confocal reflection and polarized light microscopy techniques, the study observed disparities in tissue texture and the distribution and integrity of individual collagen fibers between normal and neoplastic-derived tissues obtained from three distinct PM patients. The results demonstrated that 3D-dECMs derived from neoplastic peritoneum exhibited a notably higher proportion of Ki-67-positive cells after 5 and 12 days. Furthermore, expression levels of specific genes critical for tissue architecture, stiffness, ECM remodeling, fibril generation, epithelial cell differentiation, resistance to compression, and regulation of angiogenesis were found to be elevated in 3D-dECMs generated from neoplastic tissue compared to those from normal tissue or Matrigel-based models. In summary, by utilizing patient-derived scaffolds and cutting-edge techniques, the researchers successfully developed more physiologically relevant models that significantly contribute to our comprehension of colorectal cancer and PM biology. These models, alongside others [ 119 , 120 , 121 , 122 ], offer valuable insights into the intricate interplay between tumor cells and the ECM, paving the way for the potential discovery of novel therapeutic targets and the development of personalized treatment strategies for peritoneal metastases.

Furthermore, decellularized tissue scaffolds provide an efficient platform to study the interactions between different components abundantly found in the ECM, like macrophages and endothelial cells. Macrophages and endothelial cells are known for their involvement in cancer progression in the context of the ECM within solid tumors, as they are often found in large numbers [ 123 ]. Macrophages within the tumor (often referred to as tumor-associated macrophages or TAMs) can be “hijacked” by cancer cells and reprogrammed to support tumor growth and progression. For example, they can promote cancer cell proliferation, enhance blood vessel formation (angiogenesis), assist in tissue remodeling, and suppress the immune response against the tumor. Pinto et al. [ 123 ] investigated how human colorectal tumor matrices influence macrophage polarization and their subsequent role in cancer cell invasion. To facilitate this, a novel 3D-organotypic model was utilized using decellularized tissues from surgical resections of colorectal cancer patients. This model preserved native tissue characteristics, including major ECM components, architecture, and mechanical properties, while removing DNA and other cellular components. The study found that macrophages within tumor matrices displayed an M2-like anti-inflammatory phenotype, characterized by higher expression of IL-10, TGF-β, and CCL18, and lower expression of CCR7 and TNF. Furthermore, it was observed that tumor ECM-educated macrophages effectively promoted cancer cell invasion through a mechanism involving CCL18, as demonstrated by Matrigel invasion assays. The high expression of CCL18 at the invasive front of human colorectal tumors correlates with advanced tumor staging, underscoring its clinical significance. The findings highlight the potential of using tumor-decellularized matrices as exceptional scaffolds for recreating complex microenvironments, thereby enabling a more comprehensive understanding of cancer progression mechanisms and therapeutic resistance.

Besides TAMs, endothelial cells express various adhesion molecules and chemokines, such as selectins, integrins, and members of the immunoglobulin superfamily, which can interact with ligands on cancer cells, facilitating their adhesion to the endothelial cell layer. This adhesion is a critical step in the extravasation process, where cancer cells exit the bloodstream and invade surrounding tissues to form metastases. Moreover, endothelial cells can signal and recruit macrophages and other immune cells to the tumor site. Once there, macrophages can be “educated” by the tumor to adopt a pro-tumor phenotype, suppressing the immune response and promoting tumor growth. Therefore, decellularized matrices are suitable for studying such interactions as they closely resemble the natural tumor environment, including native adhesion sites, signaling molecules, and mechanical cues. Helal-Neto et al. [ 124 ] examined the influence of dECM produced by a highly metastatic human melanoma cell line (MV3) on the activation of endothelial cells and their intracellular cell differentiation signaling pathways. The researchers studied the differences in the ultrastructural organization and composition of melanocyte-derived ECM (NGM-ECM) and melanoma-derived (MV3-ECM). Higher levels of tenascin-C and laminin and lower fibronectin expression were detected in MV3-ECM. Moreover, endothelial cells cultured in the MV3-ECM underwent morphological transformations and exhibited increased adhesion, mobility, growth, and tubulogenesis. The interaction between the endothelial cells and decellularized matrix induced integrin signaling activation, resulting in focal adhesion kinase (FAK) phosphorylation and its association with Src (a non-receptor tyrosine kinase protein). Src activation, in turn, stimulated the activation of vascular endothelial growth factor receptor 2 (VEGFR2), enhancing the receptor’s response to VEGF. The activation of VEGF and the association between FAK and Src was inhibited by blocking the αvβ3 integrin, which reduced tubulogenesis. In conclusion, the findings suggested that the interaction of endothelial cells with melanoma-ECM triggered integrin-dependent signaling, which led to the activation of the Src pathway that sequentially potentiated VEGFR2 activation and enhanced angiogenesis. Thus, progress in cancer biology relies on understanding the specific cellular responses influenced by the matrix signals within the ECM, as its nature inherently imposes spatial variations on cellular signaling, composition, topography, and biochemical factors. Table 7 summarizes some studies using hydrogel and decellularized tissue scaffolds for 3D cell cultures.

Hybrid scaffolds

Integrating multiple scaffold types offers the potential to create 3D cell culture systems that closely mimic the physiological conditions of living tissues. This approach enables researchers to develop more accurate and biologically relevant models for studying cellular behavior, disease progression, and therapeutic responses. By combining different scaffold materials, such as natural and synthetic polymers or hydrogels, researchers can replicate the complexity and heterogeneity of the native tissue microenvironment. These hybrid scaffolds can provide a range of physical, chemical, and mechanical cues that influence cell behavior, including cell adhesion, migration, proliferation, and differentiation. Additionally, the combination of scaffolds can enhance the functionality of the 3D cell culture systems by incorporating specific features, such as the controlled release of growth factors or the inclusion of microvascular networks. Utilizing diverse scaffold types in 3D cell culture offers an innovative and promising approach for advancing our understanding of tissue biology, disease mechanisms, and developing more effective therapies. Bassi et al. [ 98 ] addressed the limitations of conventional therapies for osteosarcoma, a type of bone cancer, by introducing two innovative approaches in tumor engineering that aim to improve therapy outcomes. The study utilized hydroxyapatite-based scaffolds that mimic the in vivo TME, specifically emphasizing the CSC niche. Two types of scaffolds were employed: a biomimetic hybrid composite scaffold obtained through biomineralization, involving the direct nucleation of magnesium-doped hydroxyapatite (MgHA) on self-assembling collagen fibers (MgHA/Coll), and porous hydroxyapatite scaffolds (HA) produced by a direct foaming process. These scaffolds provided a framework for the subsequent investigation of the biological performance of human osteosarcoma cell lines (MG63 and SAOS-2) and enriched CSCs within these complex 3D cell culture models. Immunofluorescence and other characterization techniques were employed to evaluate the response of the osteosarcoma cell lines and CSCs to the biomimetic scaffolds. The results demonstrated the successful formation of sarcospheres, which are stable spheroids enriched with CSCs, with a minimum diameter of 50 µm. Comparing the advanced 3D cell culture models with conventional 2D culture systems, the study revealed the former’s superiority in mimicking the osteosarcoma stem cell niche and enhancing the predictivity of preclinical studies. The findings underscore the significance of the TME and emphasize the potential of combining CSCs with biomimetic scaffolds as a promising approach to developing novel therapeutic strategies for osteosarcoma. Further efforts can be focused on developing more sophisticated 3D models that accurately replicate the heterogeneity of the osteosarcoma microenvironment, incorporating patient-derived cells and elements such as immune cells and vasculature. Additionally, the advanced 3D cell culture models can serve as valuable tools for drug screening and personalized medicine approaches, further contributing to advancing osteosarcoma research and treatment strategies.

A unique cell culture technique known as “sequential culture” was used to establish a biomimetic bone microenvironment that facilitated the EMT of metastasized prostate cancer cells [ 141 ]. The approach involved incorporating bioactive factors from the osteogenic induction of human mesenchymal stem cells (MSCs) within porous 3D scaffolds, specifically polymer–clay composite (PCN) scaffolds, by incorporating hydroxyapatite (HAP) clay into PCL. The researchers also modified sodium clay Montmorillonite (Na-MMT) clay using 5-amino valeric acid to create HAPclay through in situ hydroxyapatite biomineralization into the intercalated nano clay. They performed RNA extraction and quantitative real-time polymerase chain reaction (qRT-PCR) analysis to investigate gene expression changes. Additionally, they conducted a comparative analysis of bone metastasis between the low and high metastatic cell lines, providing insights into their differential responses to the bone microenvironment. It was shown that both, the highly metastatic prostate cancer cell line PC-3 and the non-metastatic cell line MDAPCa2b, underwent MET transition when exposed to the biomimetic bone microenvironment in the 3D scaffold model. However, notable differences were observed in their morphological characteristics and cell–cell adhesion, suggesting distinct responses to the microenvironment. Additionally, quantitative variations in gene expression were observed between tumors generated using the two cell lines in the bone microenvironment. These findings are essential for developing targeted therapeutic strategies against prostate cancer bone metastasis. Bai et al. [ 142 ] conducted a study in which they incorporated graphene oxide (GO) onto a copolymer of polyacrylic acid-g-polylactic acid (PAA-g-PLLA) to create a stimuli-responsive scaffold. This scaffold, combined with PCL and gambogic acid (GA), exhibited a selective response towards tumors and demonstrated a significant accumulation of GO/GA in vitro breast tumor cells (MCF-7 cells) under acidic conditions (pH 6.8), while showing minimal impact on normal cells (MCF-10A cells) at physiological pH (pH 7.4). The study further revealed that the synergistic use of pH-responsive photo-thermal conversion was more effective in inhibiting tumor growth than independent treatments. In vivo experiments showed remarkable tumor suppression (99% reduction within 21 days) through tumor tissue disintegration, degeneration, and overall tumor suppression when treated with GO-GA scaffolds combined with photo-thermal therapy, in comparison to control groups or those treated with either GO-GA scaffolds or near-infrared (NIR) irradiation alone.

Microfluidics provide a versatile platform for 3D cell culture, offering both scaffold-based and scaffold-free approaches. Researchers can tailor the platform to suit the specific requirements of their experiments, whether involving cell-laden scaffolds or the aggregation of cells to form spheroids or organoids. The microfluidic setup allows for precise control over the microenvironment, including the flow of nutrients and oxygen, as well as the ability to introduce gradients of specific molecules. Lee et al. [ 143 ] utilized soft lithography to fabricate a 7-channel microchannel plate using poly-dimethylsiloxane (PDMS). Within separate channels, PANC-1 pancreatic cancer cells and pancreatic stellate cells (PSCs) were cultured within a collagen I matrix. The study observed the formation of 3D tumor spheroids by PANC-1 cells within five days. Intriguingly, the presence of co-cultured PSCs resulted in an increased number of spheroids, suggesting a potential influence of PSCs on tumor growth. In the co-culture setup, PSCs exhibited heightened expression of α-smooth muscle actin (α-SMA), a marker associated with fibroblast activation, as well as various EMT-related markers, including vimentin, transforming growth factor-beta (TGF-β), TIMP1, and IL-8. These findings indicated that PSCs may induce an EMT-like phenotype in PANC-1 cells, potentially promoting tumor invasiveness, chemoresistance, and metastasis. Upon treating the co-culture with gemcitabine, the survival of the spheroids did not exhibit significant changes. However, when combined with paclitaxel, the tumor spheroids demonstrated a notable inhibitory effect on growth. The model revealed a complex interplay between PANC-1 cells and PSCs within the TME. Nonetheless, the combination of gemcitabine and paclitaxel showed promise to overcome resistance and inhibit tumor growth. The implications of these findings are significant for understanding the complex interplay between tumor cells and the surrounding stromal cells within the TME. Tumor-stroma interactions play a critical role in cancer progression and therapy response. Using microfluidic-based 3D co-culture models allows researchers to better recapitulate the in vivo conditions, providing a more accurate representation of tumor behavior and therapeutic responses.

Likewise, Chen et al. [ 144 ] developed a microchannel plate-based co-culture model to recreate the in vivo TME by combining Hepa1-6 tumor spheroids with JS-1 stellate cells (liver cancer)—the novel model aimed to mimic key aspects of EMT and chemoresistance observed in tumors. The integration of these cell types in 3D concave microwells allowed for the formation of 3D tumor spheroids in 3 days. The experimental setup was optimized to ensure optimal culture proliferation conditions and appropriate interactions between Hepa1-6 and JS-1 cells. Co-cultured JS-1 cells displayed noticeable changes in cellular morphology, including an increase in the expression of α-SMA. In contrast, the co-cultured Hepa1-6 spheroids exhibited higher expression levels of TGF-β1 than those cultured alone. These findings suggested that JS-1 stellate cells induced an EMT-like phenotype in the Hepa1-6 cells, potentially contributing to increased invasiveness and resistance to chemotherapy. Jeong et al. [ 145 ] conducted a similar study involving the formation of 3D spheroids composed of human colorectal carcinoma cells (HT-29) using a microfluidic chip. They reported a notable enhancement in HT-29 growth when co-cultured with fibroblasts (see Fig.  5 ). This enhancement was demonstrated by a 1.5-fold increase in the percentage change in spheroid diameter over 5 days. Furthermore, after 6 days of culture, the co-cultured spheroids exhibited reduced expression of Ki-67, a marker associated with proliferation, while showing increased fibronectin expression. These findings indicated altered cellular behavior compared to the spheroid monocultures. The presence of fibroblasts in the co-culture environment also led to their activation, as evidenced by an upregulation in the expression of α-smooth muscle actin (α-SMA) and an increase in migratory activity. This reciprocal interaction between the spheroids and fibroblasts within a microfluidic chip established a dynamic relationship. Additionally, when exposed to paclitaxel, the co-culture displayed a survival advantage over 2D monoculture, suggesting the potential role of fibroblasts in conferring drug resistance. Integrating the 3D tumor spheres and CAFs within a collagen matrix-incorporated microfluidic chip provided a valuable tool for studying the TME and evaluating drug screening and efficacy. This approach allowed for the replication of essential interactions between tumor cells and stromal components, which are known to influence cancer progression and therapeutic response. By utilizing the proposed microfluidic chip-based model, researchers can delve into the intricate dynamics of the TME and explore novel therapeutic approaches. The ability to control and better mimic the in vivo conditions within the chip provides a valuable platform for investigating drug responses and evaluating the effectiveness of anticancer treatments. Further exploration and refinement of this model could lead to significant advancements in our understanding of tumor biology and the development of targeted therapies for improved patient outcomes. Table 8 summarizes some studies using microfluidic-based systems to develop 3D cell cultures.

figure 5

illustration of the microfluidic chip used in 3D co-culture of human colorectal cancer cells (HT-29) and normal colorectal fibroblasts (CCD-18Co) in a collagen matrix. The chip comprised 4 units, each featuring 7 channels for cell loading or media fill. Cancer and fibroblast cells were loaded into channels 4 and 2 in the co-culture, while channels 1 and 3 were designated for media fill. A cell loading channel’s detailed structure and dimensions are illustrated at the bottom left. Figure adapted from [ 145 ]

Challenges and future prospectives

While 3D cell culture offers many advantages over traditional 2D culture, it also presents some unique challenges that must be addressed to realize its potential for advancing research fully. One significant challenge is maintaining a stable and reproducible culture system. 3D cell culture systems often require specialized equipment, such as bioreactors and microfluidic devices, which can be expensive and difficult to use. These systems can be more challenging to reproduce compared to 2D systems due to the increased complexity and high heterogeneity of the culture environment, as cells are often embedded in matrices or scaffolds, making it difficult to control factors such as temperature, pH, and the presence of growth factors and/or other signaling molecules [ 149 ]. In addition, there is often a high degree of variability between different batches of cells and between experiments, making it difficult to draw statistically supported conclusions. Considering 3D cell cultures, adhering to Good Manufacturing Practices (GMP) principles is essential for translating these advanced models from research to clinical and commercial applications. However, several challenges and considerations arise when implementing GMP standards, including standardization of culture conditions, scalability, quality control, raw materials and biologics sourcing, regulatory compliance, data integrity, and documentation. GMP-compliant manufacturing processes require high reproducibility and control over critical parameters such as cell sourcing, culture media, culture supplements, and environmental conditions [ 150 , 151 ]. As mentioned above, achieving this consistency can be challenging, given the inherent biological variability of primary cells and the sensitivity of 3D cultures to slight changes in culture conditions. Furthermore, meeting regulatory requirements is a paramount challenge in translating 3D spheroid cultures to clinical applications. Regulatory bodies, such as the Food and Drug Administration (FDA) in the United States and the European Medicines Agency (EMA) in Europe, have specific guidelines for the use of cell-based therapies and products [ 152 ]. GMP compliance is necessary to navigate these regulatory pathways and obtain approval for clinical trials and commercialization.

Moreover, oxygen accessibility is a critical consideration in 3D cell culture methods, and its heterogeneity within these environments poses a significant challenge in replicating physiological conditions and obtaining accurate experimental results. Cells located in the interior of 3D structures, such as spheroids, often encounter limited oxygen availability due to microenvironmental factors (i.e., tumor spheroids naturally develop hypoxic regions due to irregular vascularization in tumors) and diffusion barriers (e.g., densely packed cells, ECM, scaffolding matrices) [ 153 ]. As cells proliferate and form 3D structures, the demand for oxygen increases due to the larger volume that oxygen must traverse. Oxygen diffusion from the surrounding culture medium becomes progressively hindered as the distance from the culture surface to the interior of the 3D structure increases. This results in an oxygen gradient, where cells near the periphery have sufficient oxygen, but those in the core encounter oxygen deficiency, leading to hypoxia. Hypoxic core cells often exhibit altered gene expression, reduced proliferation, and changes in metabolic pathways as they enter a dormant state and cease cycling when deprived of oxygen and nutrients. This reduced activity renders them relatively resistant to cytostatic drugs that predominantly target actively dividing cells, leading to increased drug resistance, as is often observed in solid tumors [ 154 , 155 ]. Confocal microscopy can be used to visualize dormant cells by labeling them with a nucleoside analog, allowing for their quantification and distinction from actively proliferating cells. This analog gets diluted in actively dividing cells. Still, it remains retained in quiescent, non-dividing cancer cells, thus providing a valuable tool for distinguishing them from the surrounding actively proliferating cells [ 156 ]. Leveraging this characteristic of 3D spheroids, they offer potential avenues for developing novel therapeutics targeting cancer cells resistant to cytostatic anticancer drugs. Wenzel et al. [ 157 ] cultivated T47D breast cancer cells in 3D cultures and used confocal imaging to differentiate cells within the inner core from those in the surrounding outer core. Cells in the inner core, experiencing limited access to oxygen and nutrients, exhibited reduced metabolic activity compared to their counterparts in the outer core. Through screening small molecule libraries against these 3D cultures, the authors identified nine compounds that selectively targeted and killed the inner core cancer cells while sparing the more actively proliferating outer cells. The identified drugs primarily affected the respiratory chain pathway, aligning with the altered metabolic activity of oxygen-deprived cells transitioning from aerobic to anaerobic metabolism. Hence, compounds selectively targeting dormant cancer cells significantly improved the effectiveness of commonly employed cytostatic anticancer drugs. Alternatively, the use of microfluidic devices that enable the creation of controlled oxygen gradients within cultures, the incorporation of oxygen-permeable materials, and the addition of oxygen-releasing compounds to provide a more uniform distribution of oxygen in vitro. However, it is important to acknowledge that these strategies may not fully replicate the complexity of oxygen gradients in real tissues [ 158 ]. Boyce et al. [ 159 ] presented the design and characterization of a modular device that capitalized on the gas-permeable properties of silicone to create oxygen gradients within cell-containing regions. The microfabricated device was constructed by stacking laser-cut acrylic and silicone rubber sheets, where the silicone not only facilitated oxygen gradient formation but also served as a barrier, separating the flowing gases from the cell culture medium to prevent evaporation or bubble formation during extended incubation periods. The acrylic components provided structural stability, ensuring a sterile culture environment. Using oxygen-sensing films, gradients with varying ranges and steepness in the microdevice can be achieved by adjusting the composition of gases flowing through the silicone elements. Furthermore, a cell-based reporter assay illustrated that cellular responses to hypoxia were directly proportional to the oxygen tension established within the system, proving efficacy.

Another practical challenge in 3D cultures arises from the intricacy of extracting cells from biomaterial-based 3D constructs. Typically, the construction of degradable hydrogel scaffolds involves integrating breakable crosslinks and/or cleavable components into the polymer structure or incorporating naturally biodegradable ECM constituents such as hyaluronic acid, laminin, fibronectin, and collagen [ 160 ]. Yet, traditional dissociation techniques prove to be notably inefficient and are influenced by the inherent structural complexities of the culture system. Enzymatic degradation, for example by collagenase, is a widely employed method for retrieving cells from 3D cell culture collagen-based scaffolds. The enzyme is selected to match the specific collagen type in the scaffold. During incubation, collagenase enzymatically cleaves the collagen fibers, releasing cells that were embedded or adhered to these fibers. Once the collagen has been broken down, the cells are collected as a suspension in the culture medium [ 161 ]. Cell viability and functionality assessments are typically performed to maintain the cells’ health and functionality. While using enzymatic degradation for 3D cell culture scaffolds is common, it remains an intricate approach associated with several limitations. It is important not to underestimate the impact of collagenase or other enzymes on cell viability and functionality. Careful optimization of digestion time and enzyme concentration is essential to balance efficient scaffold degradation and preserving cell quality [ 162 ]. Additionally, potential changes in cell phenotype during digestion are a significant concern, necessitating diligent monitoring of digestion parameters. In complex 3D scaffolds, particularly those with intricate structures, enzymatic digestion may be less effective, prompting researchers to explore alternative retrieval methods or adapt the digestion process. Ethical considerations also come into play, especially when working with human or animal-derived cells, raising concerns about using enzymes like collagenase. Adherence to ethical guidelines and institutional regulations is crucial for maintaining responsible and ethical research practices.

Hence, extensive research efforts have been directed toward developing improved techniques for cell retrieval from scaffold-based 3D cell cultures without compromising the cells’ integrity. For instance, Kyykallio et al. [ 163 ] developed an innovative pipeline for extracting extracellular vesicles (EVs) from 3D cancer spheroids using nanofibrillar cellulose (NFC) scaffolds as a cell culture matrix. This pipeline encompassed two distinct approaches: a batch method optimized for maximal EV yield at the conclusion of the culture period, and a harvesting method designed to facilitate time-dependent EV collection, allowing integration of EV profiling with spheroid development. Both approaches provided convenient setup, quick execution, and reliably produced a significant number of electric vehicles (EVs). Compared to scaffold-free 3D spheroid cultures on ultra-low affinity plates, the NFC-based approach demonstrated similar EV production per cell, offering scalability, preserved cell phenotype and integrity, and greater operational simplicity, ultimately leading to higher EV yields. Another approach is based on cell-mediated degradation of hydrogel scaffolds, where living cells actively break down the hydrogel structure [ 164 ]. This degradation mechanism is particularly relevant in tissue engineering and regenerative medicine. When cells are encapsulated within a hydrogel scaffold, they can secrete enzymes and other molecules that interact with its components, leading to its gradual breakdown. As cells proliferate and remodel their microenvironment, they may alter the scaffold’s properties and eventually facilitate its degradation. This dynamic process allows for the controlled release of cells, growth factors, and other bioactive substances within the hydrogel, making it a valuable technique for drug delivery applications.

While synthetic degradable polymer scaffolds are significant for developing 3D cell culture models, a concern regarding their in vitro and in vivo biocompatibility pertains to the presence of potentially toxic elements and chemicals utilized during the polymerization of synthetic hydrogels or the crosslinking of natural polymer hydrogel precursors, especially when the reaction conversion is less than 100%. These substances release unreacted monomers, stabilizers, initiators, organic solvents, and emulsifiers. These are integral to the hydrogel preparation process but may pose harm if they seep into the seeded cells or tissues [ 165 , 166 ]. For instance, widely employed free radical photo-initiators (e.g., Irgacure) have been observed to diminish cell viability, even at minimal concentrations [ 167 , 168 ]. Consequently, hydrogel scaffolds intended for embedding cells in 3D cultures typically require purification (e.g., by dialysis or solvent washing) to eliminate any residual hazardous chemicals before seeding. However, in certain scenarios, the purification of hydrogel scaffolds is more challenging or unfeasible, particularly when dealing with hydrogels generated through in situ gelation. In such cases, cells are introduced to the reactants necessary for hydrogel synthesis while still in a pre-polymer solution. As a result, when employing in situ gelation techniques, utmost caution must be exercised to ensure that all components are non-toxic and safe.

Furthermore, another challenge associated with 3D cell culture is the difficulty characterizing the cellular response to drugs and other therapeutic agents. In 2D cell culture, cells are typically analyzed using a range of standard assays that are well-established and easy to interpret. However, in 3D cell culture, there is often a lack of such standardized assays and protocols. Fang and Eglen [ 169 ] highlighted that the cultures’ complex morphology, functionality, and architecture hampered the application of some well-developed biochemical assays to 3D systems. Cells tend to aggregate into dense and/or large clusters over time, even in macroporous scaffolds, causing diffusional limitations when carrying out in situ characterization assays. Limitations arise due to the impeded diffusion and confinement of gases, nutrients, waste, and reagents within the system, compounded by challenges when quantifying and normalizing data between different biomimetic cultures [ 170 , 171 , 172 ]. For instance, Totti et al. [ 173 ] demonstrated that assessing a culture of pancreatic cancer cells in macroporous polyurethane foam-type scaffolds with the 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium (MTS) assay showed minimal differences between various scaffold conditions (e.g., ECM coatings on the scaffolds). However, sectioning, immunostaining, and imaging revealed clearer cell proliferation, morphology, and growth distinctions between the conditions. Likewise, the 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) assay failed in capturing the differences in pancreatic cells’ viability cultured in polyurethane scaffolds after drug and irradiation screening, which were realized using advanced microscopy and imaging [ 174 ]. Hence, it is crucial for researchers to carefully consider the appropriate analytical approach that aligns with their study objectives before commencing the analysis of any 3D cultures. Also, they must be aware that some of the classical gold-standard approaches used in 2D cultures may not be directly applicable in 3D settings, as Hamdi et al. [ 175 ] showed that it is unfeasible to extract cells from spheroids for colony formation assays, which are used for developing post-treatment survival curves. Consequently, the researchers suggested in situ characterization readouts, which are novel and/or different from the existing 2D culture protocols.

Using stem cells and differentiated markers is crucial for characterizing and monitoring the cellular composition and differentiation status within 3D spheroids. These markers can help researchers achieve specific goals and outcomes, such as assessing the differentiation potential of stem cells, tracking the progression of differentiation, and studying the dynamics of cell populations in the spheroids [ 176 , 177 ]. However, using such markers in 3D spheroid cultures presents certain challenges that need to be addressed for accurate and meaningful results. One primary challenge is the heterogeneity of stem cells within spheroids. Spheroids often comprise a mixture of stem cells and differentiated cells, so the stem cell markers may not exclusively identify and isolate the stem cell population, leading to difficulty in studying the specific behavior of stem cells within the spheroid. Another challenge is the variability in the expression of stem cell markers. These markers’ expression can fluctuate spatially and temporally within the spheroid, making it complex to track and interpret changes in marker expression over time. Additionally, in larger spheroids, stem cell markers may not effectively penetrate the core of the spheroid, limiting the ability to assess the stem cell population in the inner regions [ 176 , 177 ]. Researchers can employ several strategies to overcome these challenges and effectively use stem cell markers in 3D spheroid cultures [ 178 , 179 ]. An alternative method involves combining stem cells and other cellular markers to better understand the cellular composition within the spheroid. This multi-marker approach can help mitigate the issues related to marker heterogeneity. Moreover, live imaging techniques, such as confocal microscopy, can provide real-time insights into the dynamics of marker expression within spheroids. Controlling the size of spheroids is another strategy to enhance marker penetration and access to the innermost cells. Utilizing microfluidic techniques allows for the accurate regulation of spheroid size, ensuring effective penetration of markers throughout all regions of the spheroid [ 178 , 179 ]. Additionally, single-cell analysis methods, such as single-cell RNA sequencing and proteomic analysis, enable the characterization of individual cells within spheroids. This approach can identify unique gene or protein expression patterns and shed light on the behavior of stem cell populations. Another valuable strategy is creating spheroids with genetically encoded stem cell reporters, which produce fluorescent or luminescent signals in stem cells, making them more visible and trackable. Lastly, mimicking the stem cell niche or microenvironment within 3D culture conditions can help maintain stemness and marker expression in spheroids [ 179 ].

Although imaging provides valuable information about cell distribution and binding, quantitative measurements using image analysis in 3D cultures are often lacking because they require cell count consistency across samples [ 180 ]. The challenge lies in the inability to visualize the whole-cell population, leading to difficulties obtaining accurate and reliable data from the entire culture. This is due to the hampered diffusion of fluorescent markers, primarily due to their large size, governed by the inherent heterogeneity of 3D cultures. One potential solution is to measure cell number from imageable cross-sections; however, Sirenko et al. [ 181 ] noted that light interferences and dye diffusion limitations resulted in unreliable results, as the number of cells counted substantially differed from the number of cells seeded. In addition, technical limitations such as prohibitive costs and limited scalability must also be considered [ 149 ]. Implementing 3D culture systems may incur higher costs compared to 2D culture systems, attributed to the requirement for specialized equipment, materials, and expertise [ 182 , 183 ]. Similarly, scaling up 3D culture systems for industrial or clinical applications can be challenging due to the increased complexity of the culture environment and the need for specialized equipment [ 184 ]. This can limit the potential for the widespread adoption of 3D culture techniques in these settings.

Significant strides have been made in creating dynamic scaffolds that can respond to or guide resident cells [ 185 ]. For example, thermoresponsive hydrogels like poly-N-isopropylacrylamide (pNIPAm) have been proven effective for cell population harvesting [ 186 , 187 ]. Moreover, the fusion of microscale technologies for cell culture with adaptable hydrogel designs has facilitated various investigations. These include investigating cell migration within microfluidic hydrogels and establishing high-throughput screening platforms to explore interactions between cells and materials [ 188 ]. Notably, the mechanobiology field is intrigued by various mechanically dynamic hydrogels that can either stiffen, soften, or reversibly transition between these states to examine cellular responses. These dynamic substrates offer a means to scrutinize how mechanical cues influence cell behavior, similar to the study of soluble factors over decades [ 189 ]. Techniques for introducing heterogeneity and multiple cell types within 3D constructs are also advancing. This includes innovative methods where hydrogels serve as bio-inks to print cells, either layer-by-layer from a 2D base or directly within a 3D space enclosed by another hydrogel. As these platforms progress, they are expected to become more widely accessible [ 190 , 191 ]. In the interim, it remains crucial to maintain an open and collaborative dialogue between cell biologists, materials scientists, and engineers. This collaborative effort will ensure that the next generation of scaffold-based 3D cell culturing systems is well-equipped to address the significant challenges posed by the increasing biological and technical complexities.

To conclude, scaffold-based 3D cell culture has emerged as a valuable tool in cancer research, providing a more physiologically relevant environment for studying tumor behavior, drug responses, and interactions between cancer cells and the surrounding microenvironment. Various scaffold materials, including polymers, decellularized tissue, hydrogels, and hybrids with microfluidics, have been explored to create complex and biomimetic 3D models. Polymer-based scaffolds offer tunable mechanical properties and are relatively easy to fabricate, making them versatile for 3D cell culture. The choice of polymers can influence cell behavior, proliferation, and migration, allowing researchers to study cancer progression and metastasis in a more realistic context. Additionally, incorporating bioactive molecules into polymer scaffolds can enable the controlled release of drugs and growth factors, facilitating drug screening and targeted therapy development. Furthermore, hydrogels offer high biocompatibility and can be functionalized with bioactive signals to direct cell behavior and tissue formation. In cancer research, hydrogels provide a platform to investigate the effect of mechanical cues on tumor growth, immune cell infiltration, and angiogenesis. Additionally, the ease of incorporating multiple cell types within hydrogels enables the study of tumor-stroma interactions. Likewise, decellularized tissue scaffolds retain native ECM composition, topography, and mechanical properties, closely mimicking the natural tumor microenvironment. As a result, cancer cells cultured in decellularized tissue scaffolds can exhibit more accurate tumor behaviors, including invasion and angiogenesis. Moreover, these scaffolds can be derived from patient-specific tissues, enabling personalized medicine approaches and improving the predictability of drug responses. Lastly, hybrid scaffolds that integrate microfluidic channels offer unique advantages for cancer research. By combining 3D cell culture with microfluidics, researchers can study tumor angiogenesis, metastasis, and drug penetration in a more physiologically relevant manner. Furthermore, microfluidics can facilitate high-throughput screening of anticancer drugs, enabling rapid and cost-effective testing of potential therapies.

Data availability

Not applicable.

Siegel RL, Miller KD, Fuchs HE, Jemal A. Cancer Statistics, 2021. CA Cancer J Clin. 2021;71(1):7–33.

Article   PubMed   Google Scholar  

Grønning T. History of cancer. In: Colditz GA, editor. The sage encyclopedia of cancer and society. 2nd ed. CA: SAGE Publications, Inc.; 2015. p. 549–54.

Google Scholar  

Guimarães* I dos S, Daltoé* RD, Herlinger AL, Madeira KP, Ladislau T, Valadão IC, Junior PCML, Fernandes Teixeira S, Amorim GM, Santos DZ dos, Demuth KR, Rangel LBA. Conventional Cancer Treatment. Cancer Treatment—Conventional and Innovative Approaches. 2013; Available from: https://www.intechopen.com/chapters/42057 .

De Vita VT. The evolution of therapeutic research in cancer. N Engl J Med. 2010;298(16):907–10.

Loessner D, Holzapfel BM, Clements JA. Engineered microenvironments provide new insights into ovarian and prostate cancer progression and drug responses. Adv Drug Deliv Rev. 2014;15(79):193–213.

Article   Google Scholar  

Xu X, Farach-Carson MC, Jia X. Three-dimensional in vitro tumor models for cancer research and drug evaluation. Biotechnol Adv. 2014;32(7):1256–68.

Article   PubMed   PubMed Central   CAS   Google Scholar  

Pickup MW, Mouw JK, Weaver VM. The extracellular matrix modulates the hallmarks of cancer. EMBO Rep. 2014;15(12):1243–53.

Brassart-Pasco S, Brézillon S, Brassart B, Ramont L, Oudart JB, Monboisse JC. Tumor microenvironment: extracellular matrix alterations influence tumor progression. Front Oncol. 2020;10:397.

Article   PubMed   PubMed Central   Google Scholar  

Nazemi M, Rainero E. Cross-talk between the tumor microenvironment, extracellular matrix, and cell metabolism in cancer. Front Oncol. 2020;10.

Jackson HW, Defamie V, Waterhouse P, Khokha R. TIMPs: versatile extracellular regulators in cancer. Nat Rev Cancer. 2017;17(1):38–53.

Article   PubMed   CAS   Google Scholar  

Li J, Xu R. Obesity-associated ECM remodeling in cancer progression. Cancers (Basel). 2022;14(22):5684.

Popovic A, Tartare-Deckert S. Role of extracellular matrix architecture and signaling in melanoma therapeutic resistance. Front Oncol. 2022;12.

Gordon-Weeks A, Yuzhalin AE. Cancer extracellular matrix proteins regulate tumour immunity. Cancers (Basel). 2020;12(11):1–25.

Romero-López M, Trinh AL, Sobrino A, Hatch MMS, Keating MT, Fimbres C, Lewis DE, Gershon PD, Botvinick EL, Digman M, Lowengrub JS, Hughes CCW. Recapitulating the human tumor microenvironment: colon tumor-derived extracellular matrix promotes angiogenesis and tumor cell growth. Biomaterials. 2017;116:118–29.

Brown Y, Hua S, Tanwar PS. Extracellular matrix in high-grade serous ovarian cancer: advances in understanding of carcinogenesis and cancer biology. Matrix Biol. 2023;118:16–46.

Frantz C, Stewart KM, Weaver VM. The extracellular matrix at a glance. J Cell Sci. 2010;123(Pt 24):4195–200.

Langhans SA. Three-dimensional in vitro cell culture models in drug discovery and drug repositioning. Front Pharmacol. 2018;9(JAN):6.

Rozario T, DeSimone DW. The extracellular matrix in development and morphogenesis: a dynamic view. Dev Biol. 2010;341(1):126–40.

Luca AC, Mersch S, Deenen R, Schmidt S, Messner I, Schäfer KL, Baldus SE, Huckenbeck W, Piekorz RP, Knoefel WT, Krieg A, Stoecklein NH. Impact of the 3D microenvironment on phenotype, gene expression, and EGFR inhibition of colorectal cancer cell lines. PLoS ONE. 2013;8(3):e59689.

Li L, Zhao Q, Kong W. Extracellular matrix remodeling and cardiac fibrosis. Matrix Biol. 2018;1(68–69):490–506.

Kiss DL, Windus LCE, Avery VM. Chemokine receptor expression on integrin-mediated stellate projections of prostate cancer cells in 3D culture. Cytokine. 2013;64(1):122–30.

Urbanczyk M, Layland SL, Schenke-Layland K. The role of extracellular matrix in biomechanics and its impact on bioengineering of cells and 3D tissues. Matrix Biol. 2020;1(85–86):1–14.

Harunaga JS, Yamada KM. Cell-matrix adhesions in 3D. Matrix Biol. 2011;30(7–8):363–8.

Schmitz A, Fischer SC, Mattheyer C, Pampaloni F, Stelzer EHK. Multiscale image analysis reveals structural heterogeneity of the cell microenvironment in homotypic spheroids. Sci Rep. 2017;7(1):1–13.

Imamura Y, Mukohara T, Shimono Y, Funakoshi Y, Chayahara N, Toyoda M, Kiyota N, Takao S, Kono S, Nakatsura T, Minami H. Comparison of 2D- and 3D-culture models as drug-testing platforms in breast cancer. Oncol Rep. 2015;33(4):1837–43.

Loessner D, Stok KS, Lutolf MP, Hutmacher DW, Clements JA, Rizzi SC. Bioengineered 3D platform to explore cell-ECM interactions and drug resistance of epithelial ovarian cancer cells. Biomaterials. 2010;31(32):8494–506.

Zhou CH, Yang SF, Li PQ. Human lung cancer cell line SPC-A1 contains cells with characteristics of cancer stem cells. Neoplasma. 2012;59(6):685–92.

Gamerith G, Rainer J, Huber JM, Hackl H, Trajanoski Z, Koeck S, Lorenz E, Kern J, Kofler R, Kelm JM, Zwierzina H, Amann A, Gamerith G, Rainer J, Huber JM, Hackl H, Trajanoski Z, Koeck S, Lorenz E, Kern J, Kofler R, Kelm JM, Zwierzina H, Amann A. 3D-cultivation of NSCLC cell lines induce gene expression alterations of key cancer-associated pathways and mimic in-vivo conditions. Oncotarget. 2017;8(68):112647–61.

Jiang R, Huang J, Sun X, Chu X, Wang F, Zhou J, Fan Q, Pang L. Construction of in vitro 3-D model for lung cancer-cell metastasis study. BMC Cancer. 2022;22(1):1–9. https://doi.org/10.1186/s12885-022-09546-9 .

Article   CAS   Google Scholar  

Chong YK, Toh TB, Zaiden N, Poonepalli A, Leong SH, Ong CEL, Yu Y, Tan PB, See SJ, Ng WH, Ng I, Hande MP, Kon OL, Ang BT, Tang C. Cryopreservation of neurospheres derived from human glioblastoma multiforme. Stem Cells. 2009;27(1):29.

Panchalingam K, Paramchuk W, Hothi P, Shah N, Hood L, Foltz G, Behie LA, Panchalingam K, Paramchuk W, Hothi P, Shah N, Hood L, Foltz G, Behie LA. Large-scale production of human glioblastoma-derived cancer stem cell tissue in suspension bioreactors to facilitate the development of novel oncolytic therapeutics. Cancer Stem Cells—The Cutting Edge. 2011;475–502.

Sutherland RM, Inch WR, McCredie JA, Kruuv J. A multi-component radiation survival curve using an in vitro tumour model. Int J Radiat Biol Relat Stud Phys Chem Med. 1970;18(5):491–5.

Youn BS, Sen A, Behie LA, Girgis-Gabardo A, Hassell JA. Scale-up of breast cancer stem cell aggregate cultures to suspension bioreactors. Biotechnol Prog. 2006;22(3):801–10. https://doi.org/10.1021/bp050430z .

Kuo CT, Chiang CL, Chang CH, Liu HK, Huang GS, Huang RYJ, Lee H, Huang CS, Wo AM. Modeling of cancer metastasis and drug resistance via biomimetic nano-cilia and microfluidics. Biomaterials. 2014;35(5):1562–71.

Ware MJ, Keshishian V, Law JJ, Ho JC, Favela CA, Rees P, Smith B, Mohammad S, Hwang RF, Rajapakshe K, Coarfa C, Huang S, Edwards DP, Corr SJ, Godin B, Curley SA. Generation of an in vitro 3D PDAC stroma rich spheroid model. Biomaterials. 2016;108:129–42.

Rice AJ, Cortes E, Lachowski D, Cheung BCH, Karim SA, Morton JP, Del Río Hernández A. Matrix stiffness induces epithelial-mesenchymal transition and promotes chemoresistance in pancreatic cancer cells. Oncogenesis. 2017;6(7):e352.

Kuo CT, Chiang CL, Huang RYJ, Lee H, Wo AM. Configurable 2D and 3D spheroid tissue cultures on bioengineered surfaces with acquisition of epithelial–mesenchymal transition characteristics. NPG Asia Mater. 2012;4(9):e27.

Michy T, Massias T, Bernard C, Vanwonterghem L, Henry M, Guidetti M, Royal G, Coll JL, Texier I, Josserand V, Hurbin A. Verteporfin-loaded lipid nanoparticles improve ovarian cancer photodynamic therapy in vitro and in vivo. Cancers. 2019;11(11):1760.

Hirst J, Pathak HB, Hyter S, Pessetto ZY, Ly T, Graw S, Koestler DC, Krieg AJ, Roby KF, Godwin AK. Licofelone enhances the efficacy of paclitaxel in ovarian cancer by reversing drug resistance and tumor stem-like properties. Cancer Res. 2018;78(15):4370–85. https://doi.org/10.1158/0008-5472.CAN-17-3993 .

Goodwin TJ, Milburn Jessup J, Wolf DA. Morphologic differentiation of colon carcinoma cell lines HT-29 and HT-29KM in rotating-wall vessels. In Vitro Cell Dev Biol. 1992;28A(1):47–60.

Dainiak MB, Savina IN, Musolino I, Kumar A, Mattiasson B, Galaev IY. Biomimetic macroporous hydrogel scaffolds in a high-throughput screening format for cell-based assays. Biotechnol Prog. 2008;24(6):1373–83.

Chandrasekaran S, Deng H, Fang Y. PTEN deletion potentiates invasion of colorectal cancer spheroidal cells through 3D Matrigel. Integr Biol (Camb). 2015;7(3):324–34.

Poornima K, Francis AP, Hoda M, Eladl MA, Subramanian S, Veeraraghavan VP, El-Sherbiny M, Asseri SM, Hussamuldin ABA, Surapaneni KM, Mony U, Rajagopalan R. Implications of three-dimensional cell culture in cancer therapeutic research. Front Oncol. 2022;12.

Gunti S, Hoke ATK, Vu KP, London NR. Organoid and spheroid tumor models: techniques and applications. Cancers (Basel). 2021;13(4):1–18.

Lovitt CJ, Shelper TB, Avery VM. Advanced cell culture techniques for cancer drug discovery. Biology. 2014;3(2):345–67.

Zeilinger K, Freyer N, Damm G, Seehofer D, Knöspel F. Cell sources for in vitro human liver cell culture models. Exp Biol Med (Maywood). 2016;241(15):1684–98.

Eglen RM, Reisine T. Human iPS cell-derived patient tissues and 3D cell culture part 2: spheroids, organoids, and disease modeling. SLAS Technol. 2019;24(1):18–27.

Su J, Zhang L, Zhang W, Choi DS, Wen J, Jiang B, Chang CC, Zhou X. Targeting the biophysical properties of the myeloma initiating cell niches: a pharmaceutical synergism analysis using multi-scale agent-based modeling. PLoS ONE. 2014;9(1):e85059.

Valyi-Nagy K, Kormos B, Ali M, Shukla D, Valyi-Nagy T. Stem cell marker CD271 is expressed by vasculogenic mimicry-forming uveal melanoma cells in three-dimensional cultures. Mol Vis. 2012;18:588.

PubMed   PubMed Central   CAS   Google Scholar  

Lombardo Y, Filipovicá A, Molyneux G, Periyasamy M, Giamas G, Hu Y, Trivedi PS, Wang J, Yaguë E, Michel L, Coombes RC. Nicastrin regulates breast cancer stem cell properties and tumor growth in vitro and in vivo. Proc Natl Acad Sci U S A. 2012;109(41):16558–63.

Sgodda M, Dai Z, Zweigerdt R, Sharma AD, Ott M, Cantz T. A scalable approach for the generation of human pluripotent stem cell-derived hepatic organoids with sensitive hepatotoxicity features. Stem Cells Dev. 2017;26(20):1490–504.

Meier F, Freyer N, Brzeszczynska J, Knöspel F, Armstrong L, Lako M, Greuel S, Damm G, Ludwig-Schwellinger E, Deschl U, Ross JA, Beilmann M, Zeilinger K. Hepatic differentiation of human iPSCs in different 3D models: a comparative study. Int J Mol Med. 2017;40(6):1759.

Weng KC, Kurokawa YK, Hajek BS, Paladin JA, Shirure VS, George SC. Human induced pluripotent stem-cardiac-endothelial-tumor-on-a-chip to assess anticancer efficacy and cardiotoxicity. Tissue Eng Part C Methods. 2020;26(1):44–55. https://doi.org/10.1089/ten.tec.2019.0248 .

Liu S, Fang C, Zhong C, Li J, Xiao Q. Recent advances in pluripotent stem cell-derived cardiac organoids and heart-on-chip applications for studying anti-cancer drug-induced cardiotoxicity. Cell Biol Toxicol. 2023;39:2527.

Han R, Sun Q, Wu J, Zheng P, Zhao G. Sodium butyrate upregulates miR-203 expression to exert anti-proliferation effect on colorectal cancer cells. Cell Physiol Biochem. 2016;39(5):1919–29.

Fontoura JC, Viezzer C, dos Santos FG, Ligabue RA, Weinlich R, Puga RD, Antonow D, Severino P, Bonorino C. Comparison of 2D and 3D cell culture models for cell growth, gene expression and drug resistance. Mater Sci Eng, C. 2020;1(107): 110264.

Ingram M, Techy GB, Saroufeem R, Yazan O, Narayan KS, Goodwin TJ, Spaulding GF. Three-dimensional growth patterns of various human tumor cell lines in simulated microgravity of a NASA bioreactor. In Vitro Cell Dev Biol Anim. 1997;33(6):459–66.

Lv D, Yu SC, Ping YF, Wu H, Zhao X, Zhang H, Cui Y, Chen B, Zhang X, Dai J, Bian XW, Yao XH. A three-dimensional collagen scaffold cell culture system for screening anti-glioma therapeutics. Oncotarget. 2016;7(35):56904.

Spill F, Reynolds DS, Kamm RD, Zaman MH. Impact of the physical microenvironment on tumor progression and metastasis. Curr Opin Biotechnol. 2016;40:41–8.

Ravi M, Paramesh V, Kaviya SR, Anuradha E, Paul Solomon FD. 3D cell culture systems: advantages and applications. J Cell Physiol. 2015;230(1):16–26. https://doi.org/10.1002/jcp.24683 .

Campuzano S, Pelling AE. Scaffolds for 3D cell culture and cellular agriculture applications derived from non-animal sources. Front Sustain Food Syst. 2019;17(3):38.

Hippler M, Lemma ED, Bertels S, Blasco E, Barner-Kowollik C, Wegener M, Bastmeyer M. 3D scaffolds to study basic cell biology. Adv Mater. 2019. https://doi.org/10.1002/adma.201808110 .

Cavo M, Delle Cave D, D’Amone E, Gigli G, Lonardo E, del Mercato LL. A synergic approach to enhance long-term culture and manipulation of MiaPaCa-2 pancreatic cancer spheroids. Sci Rep. 2020;10(1).

Ware MJ, Colbert K, Keshishian V, Ho J, Corr SJ, Curley SA, Godin B. Generation of homogenous three-dimensional pancreatic cancer cell spheroids using an improved hanging drop technique. Tissue Eng Part C Methods. 2016;22(4):312–21.

Foty R. A simple hanging drop cell culture protocol for generation of 3D spheroids. J Vis Exp. 2011;51(51).

Timmins NE, Nielsen LK. Generation of multicellular tumor spheroids by the hanging-drop method. Methods Mol Med. 2007;140:141–51.

Zhao L, Xiu J, Liu Y, Zhang T, Pan W, Zheng X, Zhang X. A 3D printed hanging drop dripper for tumor spheroids analysis without recovery. Sci Rep. 2019;9(1):1–14.

Gao W, Wu D, Wang Y, Wang Z, Zou C, Dai Y, Ng CF, Teoh JYC, Chan FL. Development of a novel and economical agar-based non-adherent three-dimensional culture method for enrichment of cancer stem-like cells. Stem Cell Res Ther. 2018;9(1):243. https://doi.org/10.1186/s13287-018-0987-x .

Madoux F, Tanner A, Vessels M, Willetts L, Hou S, Scampavia L, Spicer TP. A 1536-well 3D viability assay to assess the cytotoxic effect of drugs on spheroids. SLAS Discov. 2017;22(5):516–24.

Khawar IA, Park JK, Jung ES, Lee MA, Chang S, Kuh HJ. Three dimensional mixed-cell spheroids mimic stroma-mediated chemoresistance and invasive migration in hepatocellular carcinoma. Neoplasia. 2018;20(8):800–12.

Song Y, Kim JS, Choi EK, Kim J, Kim KM, Seo HR, Song Y, Kim JS, Kyung Choi E, Kim J, Mo Kim K, Ran Seo H. TGF-β–independent CTGF induction regulates cell adhesion mediated drug resistance by increasing collagen I in HCC. Oncotarget. 2017;8(13):21650–62.

Ma L, Li Y, Wu Y, Aazmi A, Zhang B, Zhou H, Yang H. The construction of in vitro tumor models based on 3D bioprinting. Biodes Manuf. 2020;3(3):227–36.

Xie M, Gao Q, Fu J, Chen Z, He Y. Bioprinting of novel 3D tumor array chip for drug screening. Biodes Manuf. 2020;3(3):175–88.

Datta P, Dey M, Ataie Z, Unutmaz D, Ozbolat IT. 3D bioprinting for reconstituting the cancer microenvironment. NPJ Precis Oncol. 2020;4(1).

LaBarge W, Morales A, Pretorius D, Kahn-Krell AM, Kannappan R, Zhang J. Scaffold-free bioprinter utilizing layer-by-layer printing of cellular spheroids. Micromachines (Basel). 2019;10(9):570.

Khoshnood N, Zamanian A. A comprehensive review on scaffold-free bioinks for bioprinting. Bioprinting. 2020;1(19): e00088.

Norotte C, Marga FS, Niklason LE, Forgacs G. Scaffold-free vascular tissue engineering using bioprinting. Biomaterials. 2009;30(30):5910–7.

Anil-Inevi M, Delikoyun K, Mese G, Tekin HC, Ozcivici E. Magnetic levitation assisted biofabrication, culture, and manipulation of 3D cellular structures using a ring magnet based setup. Biotechnol Bioeng. 2021;118(12):4771–85.

Anil-Inevi M, Yaman S, Yildiz AA, Mese G, Yalcin-Ozuysal O, Tekin HC, Ozcivici E. Biofabrication of in situ self assembled 3D cell cultures in a weightlessness environment generated using magnetic levitation. Sci Rep. 2018;8(1).

Türker E, Demirçak N, Arslan-Yildiz A. Correction: scaffold-free three-dimensional cell culturing using magnetic levitation. Biomater Sci. 2018;6(7):1996.

Haisler WL, Timm DM, Gage JA, Tseng H, Killian TC, Souza GR. Three-dimensional cell culturing by magnetic levitation. Nat Protoc. 2013;8(10):1940–9.

Jaganathan H, Gage J, Leonard F, Srinivasan S, Souza GR, Dave B, Godin B. Three-dimensional in vitro co-culture model of breast tumor using magnetic levitation. Sci Rep. 2014;4.

Caleffi JT, Aal MCE, de Gallindo H, Caxali GH, Crulhas BP, Ribeiro AO, Souza GR, Delella FK. Magnetic 3D cell culture: state of the art and current advances. Life Sci. 2021;286:120028.

Rogers M, Sobolik T, Schaffer DK, Samson PC, Johnson AC, Owens P, Codreanu SG, Sherrod SD, McLean JA, Wikswo JP, Richmond A. Engineered microfluidic bioreactor for examining the three-dimensional breast tumor microenvironment. Biomicrofluidics. 2018;12(3):34102.

Xu Z, Gao Y, Hao Y, Li E, Wang Y, Zhang J, Wang W, Gao Z, Wang Q. Application of a microfluidic chip-based 3D co-culture to test drug sensitivity for individualized treatment of lung cancer. Biomaterials. 2013;34(16):4109–17.

Ruedinger F, Lavrentieva A, Blume C, Pepelanova I, Scheper T. Hydrogels for 3D mammalian cell culture: a starting guide for laboratory practice. Appl Microbiol Biotechnol. 2015;99(2):623–36.

Zhang N, Milleret V, Thompson-Steckel G, Huang NP, Vörös J, Simona BR, Ehrbar M. Soft hydrogels featuring in-depth surface density gradients for the simple establishment of 3D tissue models for screening applications. SLAS Discovery. 2017;22(5):635–44.

Rijal G, Bathula C, Li W. Application of synthetic polymeric scaffolds in breast cancer 3D tissue cultures and animal tumor models. Int J Biomater. 2017;2017:9.

Unnikrishnan K, Thomas LV, Ram Kumar RM. Advancement of scaffold-based 3D cellular models in cancer tissue engineering: an update. Front Oncol. 2021;11.

van Tienderen GS, Conboy J, Muntz I, Willemse J, Tieleman J, Monfils K, Schurink IJ, Demmers JAA, Doukas M, Koenderink GH, van der Laan LJW, Verstegen MMA. Tumor decellularization reveals proteomic and mechanical characteristics of the extracellular matrix of primary liver cancer. Biomater Adv. 2023;1(146): 213289.

Li X, Valadez AV, Zuo P, Nie Z. Microfluidic 3D cell culture: potential application for tissue-based bioassays. Bioanalysis. 2012;4(12):1509.

Zhong Q, Ding H, Gao B, He Z, Gu Z. Advances of microfluidics in biomedical engineering. Adv Mater Technol. 2019;4(6).

Santos SC, Custódio CA, Mano JF. Human protein-based porous scaffolds as platforms for xeno-free 3D cell culture. Adv Healthc Mater. 2022;11(12).

Afewerki S, Sheikhi A, Kannan S, Ahadian S, Khademhosseini A. Gelatin-polysaccharide composite scaffolds for 3D cell culture and tissue engineering: Towards natural therapeutics. Bioeng Transl Med. 2018;4(1):96–115.

Chen L, Xiao Z, Meng Y, Zhao Y, Han J, Su G, Chen B, Dai J. The enhancement of cancer stem cell properties of MCF-7 cells in 3D collagen scaffolds for modeling of cancer and anti-cancer drugs. Biomaterials. 2012;33(5):1437–44.

McGrath J, Panzica L, Ransom R, Withers HG, Gelman IH. Identification of genes regulating breast cancer dormancy in 3D bone endosteal niche cultures. Mol Cancer Res. 2019;17(4):860–9.

Arya N, Sardana V, Saxena M, Rangarajan A, Katti DS. Recapitulating tumour microenvironment in chitosan-gelatin three-dimensional scaffolds: an improved in vitro tumour model. J R Soc Interface. 2012;9(77):3288–302.

Bassi G, Panseri S, Dozio SM, Sandri M, Campodoni E, Dapporto M, Sprio S, Tampieri A, Montesi M. Scaffold-based 3D cellular models mimicking the heterogeneity of osteosarcoma stem cell niche. Sci Rep. 2020;10(1):22294.

Kievit FM, Wang K, Erickson AE, Lan Levengood SK, Ellenbogen RG, Zhang M. Modeling the tumor microenvironment using chitosan-alginate scaffolds to control the stem-like state of glioblastoma cells. Biomater Sci. 2016;4(4):610.

Li W, Hu X, Wang S, Xing Y, Wang H, Nie Y, Liu T, Song K. Multiple comparisons of three different sources of biomaterials in the application of tumor tissue engineering in vitro and in vivo. Int J Biol Macromol. 2019;130:166–76.

Palomeras S, Rabionet M, Ferrer I, Sarrats A, Garcia-Romeu ML, Puig T, Ciurana J. Breast cancer stem cell culture and enrichment using Poly(ε-Caprolactone) scaffolds. Molecules. 2016;21(4):537.

Pradhan S, Clary JM, Seliktar D, Lipke EA. A three-dimensional spheroidal cancer model based on PEG-fibrinogen hydrogel microspheres. Biomaterials. 2017;1(115):141–54.

Godugu C, Patel AR, Desai U, Andey T, Sams A, Singh M. AlgiMatrix TM based 3D cell culture system as an in-vitro tumor model for anticancer studies. PLoS ONE. 2013;8(1): e53708. https://doi.org/10.1371/journal.pone.0053708 .

Andersen T, Auk-Emblem P, Dornish M. 3D cell culture in alginate hydrogels. Microarrays. 2015;4(2):133–61.

Yang Z, Zhao X. A 3D model of ovarian cancer cell lines on peptide nanofiber scaffold to explore the cell–scaffold interaction and chemotherapeutic resistance of anticancer drugs. Int J Nanomed. 2011;6:310.

Song H, Cai GH, Liang J, Ao DS, Wang H, Yang ZH. Three-dimensional culture and clinical drug responses of a highly metastatic human ovarian cancer HO-8910PM cells in nanofibrous microenvironments of three hydrogel biomaterials. J Nanobiotechnol. 2020;18(1):90.

Suo A, Xu W, Wang Y, Sun T, Ji L, Qian J. Dual-degradable and injectable hyaluronic acid hydrogel mimicking extracellular matrix for 3D culture of breast cancer MCF-7 cells. Carbohydr Polym. 2019;211:336–48.

Wang J, Xu W, Qian J, Wang Y, Hou G, Suo A. Photo-crosslinked hyaluronic acid hydrogel as a biomimic extracellular matrix to recapitulate in vivo features of breast cancer cells. Colloids Surf B Biointerfaces. 2022;209:112159.

Kim MJ, Chi BH, Yoo JJ, Ju YM, Whang YM, Chang IH. Structure establishment of three-dimensional (3D) cell culture printing model for bladder cancer. PLoS ONE. 2019;14(10):e0223689.

Loessner D, Meinert C, Kaemmerer E, Martine LC, Yue K, Levett PA, Klein TJ, Melchels FPW, Khademhosseini A, Hutmacher DW. Functionalization, preparation and use of cell-laden gelatin methacryloyl-based hydrogels as modular tissue culture platforms. Nat Protoc. 2016;11(4):727–46.

Arya AD, Hallur PM, Karkisaval AG, Gudipati A, Rajendiran S, Dhavale V, Ramachandran B, Jayaprakash A, Gundiah N, Chaubey A. Gelatin methacrylate hydrogels as biomimetic three-dimensional matrixes for modeling breast cancer invasion and chemoresponse in vitro. ACS Appl Mater Interfaces. 2016;8(34):22005–17.

Curtis KJ, Schiavi J, Mc Garrigle MJ, Kumar V, McNamara LM, Niebur GL. Mechanical stimuli and matrix properties modulate cancer spheroid growth in three-dimensional gelatin culture. J R Soc Interface. 2020;17(173):20200568.

Cavo M, Fato M, Peñuela L, Beltrame F, Raiteri R, Scaglione S. Microenvironment complexity and matrix stiffness regulate breast cancer cell activity in a 3D in vitro model. Sci Rep. 2016;6(1):1–13.

VandenHeuvel SN, Farris HA, Noltensmeyer DA, Roy S, Donehoo DA, Kopetz S, Haricharan S, Walsh AJ, Raghavan S. Decellularized organ biomatrices facilitate quantifiable in vitro 3D cancer metastasis models. Soft Matter. 2022;18(31):5791–806.

Landberg G, Fitzpatrick P, Isakson P, Jonasson E, Karlsson J, Larsson E, Svanström A, Rafnsdottir S, Persson E, Gustafsson A, Andersson D, Rosendahl J, Petronis S, Ranji P, Gregersson P, Magnusson Y, Håkansson J, Ståhlberg A. Patient-derived scaffolds uncover breast cancer promoting properties of the microenvironment. Biomaterials. 2020;1(235): 119705.

Zhang X, Chen X, Hong H, Hu R, Liu J, Liu C. Decellularized extracellular matrix scaffolds: recent trends and emerging strategies in tissue engineering. Bioact Mater. 2021;10:15–31.

D’angelo E, Natarajan D, Sensi F, Ajayi O, Fassan M, Mammano E, Pilati P, Pavan P, Bresolin S, Preziosi M, Miquel R, Zen Y, Chokshi S, Menon K, Heaton N, Spolverato G, Piccoli M, Williams R, Urbani L, Agostini M. Patient-derived scaffolds of colorectal cancer metastases as an organotypic 3D model of the liver metastatic microenvironment. Cancers (Basel). 2020;12(2):364.

Varinelli L, Guaglio M, Brich S, Zanutto S, Belfiore A, Zanardi F, Iannelli F, Oldani A, Costa E, Chighizola M, Lorenc E, Minardi SP, Fortuzzi S, Filugelli M, Garzone G, Pisati F, Vecchi M, Pruneri G, Kusamura S, Baratti D, Cattaneo L, Parazzoli D, Podestà A, Milione M, Deraco M, Pierotti MA, Gariboldi M. Decellularized extracellular matrix as scaffold for cancer organoid cultures of colorectal peritoneal metastases. J Mol Cell Biol. 2023;14(11).

Tian X, Werner ME, Roche KC, Hanson AD, Foote HP, Yu SK, Warner SB, Copp JA, Lara H, Wauthier EL, Caster JM, Herring LE, Zhang L, Tepper JE, Hsu DS, Zhang T, Reid LM, Wang AZ. Organ-specific metastases obtained by culturing colorectal cancer cells on tissue-specific decellularized scaffolds. Nat Biomed Eng. 2018;2(6):443–52.

Li H, Dai W, Xia X, Wang R, Zhao J, Han L, Mo S, Xiang W, Du L, Zhu G, Xie J, Yu J, Liu N, Huang M, Zhu J, Cai G. Modeling tumor development and metastasis using paired organoids derived from patients with colorectal cancer liver metastases. J Hematol Oncol. 2020;13(1):1–6. https://doi.org/10.1186/s13045-020-00957-4 .

Parkinson GT, Salerno S, Ranji P, Håkansson J, Bogestål Y, Wettergren Y, Ståhlberg A, Bexe Lindskog E, Landberg G. Patient-derived scaffolds as a model of colorectal cancer. Cancer Med. 2021;10(3):867–82.

Jian M, Ren L, Ren L, He G, He G, Lin Q, Lin Q, Tang W, Tang W, Chen Y, Chen J, Chen J, Liu T, Ji M, Ji M, Wei Y, Wei Y, Chang W, Chang W, Xu J, Xu J. A novel patient-derived organoids-based xenografts model for preclinical drug response testing in patients with colorectal liver metastases. J Transl Med. 2020;18(1):1–17. https://doi.org/10.1186/s12967-020-02407-8 .

Pinto ML, Rios E, Silva AC, Neves SC, Caires HR, Pinto AT, Durães C, Carvalho FA, Cardoso AP, Santos NC, Barrias CC, Nascimento DS, Pinto-do-Ó P, Barbosa MA, Carneiro F, Oliveira MJ. Decellularized human colorectal cancer matrices polarize macrophages towards an anti-inflammatory phenotype promoting cancer cell invasion via CCL18. Biomaterials. 2017;124:211–24.

Helal-Neto E, Brandão-Costa RM, Saldanha-Gama R, Ribeiro-Pereira C, Midlej V, Benchimol M, Morandi V, Barja-Fidalgo C. Priming endothelial cells with a melanoma-derived extracellular matrix triggers the activation of αvβ3/VEGFR2 axis. J Cell Physiol. 2016;231(11):2464–73.

Fontana F, Raimondi M, Marzagalli M, Sommariva M, Limonta P, Gagliano N. Epithelial-to-mesenchymal transition markers and CD44 isoforms are differently expressed in 2D and 3D cell cultures of prostate cancer cells. Cells. 2019;8(2):143.

Li GN, Livi LL, Gourd CM, Deweerd ES, Hoffman-Kim D. Genomic and morphological changes of neuroblastoma cells in response to three-dimensional matrices. Tissue Eng. 2007;13(5):1035–47.

Sarwar M, Sykes PH, Chitcholtan K, Evans JJ. Collagen I dysregulation is pivotal for ovarian cancer progression. Tissue Cell. 2022;74.

Tsai S, McOlash L, Palen K, Johnson B, Duris C, Yang Q, Dwinell MB, Hunt B, Evans DB, Gershan J, James MA. Development of primary human pancreatic cancer organoids, matched stromal and immune cells and 3D tumor microenvironment models. BMC Cancer. 2018;18(1):1–13. https://doi.org/10.1186/s12885-018-4238-4 .

Sha H, Zou Z, Xin K, Bian X, Cai X, Lu W, Chen J, Chen G, Huang L, Blair AM, Cao P, Liu B. Tumor-penetrating peptide fused EGFR single-domain antibody enhances cancer drug penetration into 3D multicellular spheroids and facilitates effective gastric cancer therapy. J Control Release. 2015;28(200):188–200.

Monteiro MV, Gaspar VM, Ferreira LP, Mano JF. Hydrogel 3D in vitro tumor models for screening cell aggregation mediated drug response. Biomater Sci. 2020;8(7):1855–64.

Kundu B, Saha P, Datta K, Kundu SC. A silk fibroin based hepatocarcinoma model and the assessment of the drug response in hyaluronan-binding protein 1 overexpressed HepG2 cells. Biomaterials. 2013;34(37):9462–74.

Simon KA, Mosadegh B, Minn KT, Lockett MR, Mohammady MR, Boucher DM, Hall AB, Hillier SM, Udagawa T, Eustace BK, Whitesides GM. Metabolic response of lung cancer cells to radiation in a paper-based 3D cell culture system. Biomaterials. 2016;1(95):47–59.

Loessner D, Rizzi SC, Stok KS, Fuehrmann T, Hollier B, Magdolen V, Hutmacher DW, Clements JA. A bioengineered 3D ovarian cancer model for the assessment of peptidase–mediated enhancement of spheroid growth and intraperitoneal spread. Biomaterials. 2013;34(30):7389–400.

Stratmann AT, Fecher D, Wangorsch G, Göttlich C, Walles T, Walles H, Dandekar T, Dandekar G, Nietzer SL. Establishment of a human 3D lung cancer model based on a biological tissue matrix combined with a Boolean in silico model. Mol Oncol. 2014;8(2):351–65. https://doi.org/10.1016/j.molonc.2013.11.009 .

Dunne LW, Huang Z, Meng W, Fan X, Zhang N, Zhang Q, An Z. Human decellularized adipose tissue scaffold as a model for breast cancer cell growth and drug treatments. Biomaterials. 2014;35(18):4940–9.

Çetin EA. Investigation of the viability of different cancer cells on decellularized adipose tissue. Celal Bayar Univ J Sci. 2023;19(2):113–9.

Sensi F, D’angelo E, Piccoli M, Pavan P, Mastrotto F, Caliceti P, Biccari A, Corallo D, Urbani L, Fassan M, Spolverato G, Riello P, Pucciarelli S, Agostini M. Recellularized colorectal cancer patient-derived scaffolds as in vitro pre-clinical 3D model for drug screening. Cancers (Basel). 2020;12(3).

Mazza G, Telese A, Al-Akkad W, Frenguelli L, Levi A, Marrali M, Longato L, Thanapirom K, Vilia MG, Lombardi B, Crowley C, Crawford M, Karsdal MA, Leeming DJ, Marrone G, Bottcher K, Robinson B, Del Rio HA, Tamburrino D, Spoletini G, Malago M, Hall AR, Godovac-Zimmermann J, Luong TV, De Coppi P, Pinzani M, Rombouts K. Cirrhotic human liver extracellular matrix 3D scaffolds promote smad-dependent TGF-β1 epithelial mesenchymal transition. Cells. 2019;9(1):83.

Zhao L, Huang L, Yu S, Zheng J, Wang H, Zhang Y. Decellularized tongue tissue as an in vitro model for studying tongue cancer and tongue regeneration. Acta Biomater. 2017;58:122–35.

Blanco-Fernandez B, Rey-Vinolas S, Baǧcl G, Rubi-Sans G, Otero J, Navajas D, Perez-Amodio S, Engel E. Bioprinting decellularized breast tissue for the development of three-dimensional breast cancer models. ACS Appl Mater Interfaces. 2022;14(26):29467–82. https://doi.org/10.1021/acsami.2c00920 .

Molla MDS, Katti DR, Katti KS. An in vitro model of prostate cancer bone metastasis for highly metastatic and non-metastatic prostate cancer using nanoclay bone-mimetic scaffolds. Springer Link. 2019;4(21):1207–13. https://doi.org/10.1557/adv.2018.682 .

Bai G, Yuan P, Cai B, Qiu X, Jin R, Liu S, Li Y, Chen X. Stimuli-responsive scaffold for breast cancer treatment combining accurate photothermal therapy and adipose tissue regeneration. Adv Funct Mater. 2019. https://doi.org/10.1002/adfm.201904401 .

Lee JH, Kim SK, Khawar IA, Jeong SY, Chung S, Kuh HJ. Microfluidic co-culture of pancreatic tumor spheroids with stellate cells as a novel 3D model for investigation of stroma-mediated cell motility and drug resistance. J Exp Clin Cancer Res. 2018;37(1):1–12. https://doi.org/10.1186/s13046-017-0654-6 .

Chen Y, Sun W, Kang L, Wang Y, Zhang M, Zhang H, Hu P. Microfluidic co-culture of liver tumor spheroids with stellate cells for the investigation of drug resistance and intercellular interactions. Analyst. 2019;144(14):4233–40.

Jeong SY, Lee JH, Shin Y, Chung S, Kuh HJ. Co-culture of tumor spheroids and fibroblasts in a collagen matrix-incorporated microfluidic chip mimics reciprocal activation in solid tumor microenvironment. PLoS ONE. 2016;11(7):e0159013.

Shin CS, Kwak B, Han B, Park K. Development of an in vitro 3D tumor model to study therapeutic efficiency of an anticancer drug. Mol Pharm. 2013;10(6):2167–75. https://doi.org/10.1021/mp300595a .

Hübner J, Raschke M, Rütschle I, Gräßle S, Hasenberg T, Schirrmann K, Lorenz A, Schnurre S, Lauster R, Maschmeyer I, Steger-Hartmann T, Marx U. Simultaneous evaluation of anti-EGFR-induced tumour and adverse skin effects in a microfluidic human 3D co-culture model. Sci Rep. 2018;8(1):1–12.

Dorayappan KDP, Gardner ML, Hisey CL, Zingarelli RA, Smith BQ, Lightfoot MDS, Gogna R, Flannery MM, Hays J, Hansford DJ, Freitas MA, Yu L, Cohn DE, Selvendiran K. A microfluidic chip enables isolation of exosomes and establishment of their protein profiles and associated signaling pathways in ovarian cancer. Cancer Res. 2019;79(13):3503–13.

Booij TH, Price LS, Danen EHJ. 3D cell-based assays for drug screens: challenges in imaging, image analysis, and high-content analysis. SLAS Discovery. 2019;24(6):615–27.

Eskes C, Boström AC, Bowe G, Coecke S, Hartung T, Hendriks G, Pamies D, Piton A, Rovida C. Good cell culture practices & in vitro toxicology. Toxicol In Vitro. 2017;1(45):272–7.

Coecke S, Balls M, Bowe G, Davis J, Gstraunthaler G, Hartung T, Hay R, Merten OW, Price A, Schechtman L, Stacey G, Stokes W. Guidance on good cell culture practice: a report of the Second ECVAM Task Force on good cell culture practice. Altern Lab Anim. 2005;33(3):261–87.

Kasendra M, Troutt M, Broda T, Bacon WC, Wang TC, Niland JC, Helmrath MA. The engineered gut: use of stem cells and tissue engineering to study physiological mechanisms and disease processes: intestinal organoids: roadmap to the clinic. Am J Physiol Gastrointest Liver Physiol. 2021;321(1):G1.

Al-Ani A, Toms D, Kondro D, Thundathil J, Yu Y, Ungrin M. Oxygenation in cell culture: critical parameters for reproducibility are routinely not reported. PLoS ONE. 2018;13(10):e0204269.

Blatchley MR, Hall F, Ntekoumes D, Cho H, Kailash V, Vazquez-Duhalt R, Gerecht S. Discretizing three-dimensional oxygen gradients to modulate and investigate cellular processes. Adv Sci. 2021;8(14):2100190. https://doi.org/10.1002/advs.202100190 .

Martinez NJ, Titus SA, Wagner AK, Simeonov A. High-throughput fluorescence imaging approaches for drug discovery using in vitro and in vivo three-dimensional models. Expert Opin Drug Discov. 2015;10(12):1347–61.

Robertson FM, Ogasawara MA, Ye Z, Chu K, Pickei R, Debeb BG, Woodward WA, Hittelman WN, Cristofanilli M, Barsky SH. Imaging and analysis of 3D tumor spheroids enriched for a cancer stem cell phenotype. J Biomol Screen. 2010;15(7):820–9.

Wenzel C, Riefke B, Gründemann S, Krebs A, Christian S, Prinz F, Osterland M, Golfier S, Räse S, Ansari N, Esner M, Bickle M, Pampaloni F, Mattheyer C, Stelzer EH, Parczyk K, Prechtl S, Steigemann P. 3D high-content screening for the identification of compounds that target cells in dormant tumor spheroid regions. Exp Cell Res. 2014;323(1):131–43.

Palacio-Castañeda V, Velthuijs N, Le Gac S, Verdurmen WPR. Oxygen control: the often overlooked but essential piece to create better in vitro systems. Lab Chip. 2022;22(6):1068–92.

Boyce MW, Simke WC, Kenney RM, Lockett MR. Generating linear oxygen gradients across 3D cell cultures with block-layered oxygen controlled chips (BLOCCs). Anal Methods. 2020;12(1):18.

Li Y, Yang HY, Lee DS. Biodegradable and injectable hydrogels in biomedical applications. Biomacromol. 2022;23(3):609–18. https://doi.org/10.1021/acs.biomac.1c01552 .

Osaki T, Sivathanu V, Kamm RD. Engineered 3D vascular and neuronal networks in a microfluidic platform. Sci Rep. 2018;8(1):1–13.

Virumbrales-Muñoz M, Ayuso JM, Lacueva A, Randelovic T, Livingston MK, Beebe DJ, Oliván S, Pereboom D, Doblare M, Fernández L, Ochoa I. Enabling cell recovery from 3D cell culture microfluidic devices for tumour microenvironment biomarker profiling. Sci Rep. 2019;9(1):1–14.

Kyykallio H, Faria AVS, Hartmann R, Capra J, Rilla K, Siljander PRM. A quick pipeline for the isolation of 3D cell culture-derived extracellular vesicles. J Extracell Vesicles. 2022;11(10).

Hubbell JA. Materials as morphogenetic guides in tissue engineering. Curr Opin Biotechnol. 2003;14(5):551–8.

Bartnikowski M, Bartnikowski NJ, Woodruff MA, Schrobback K, Klein TJ. Protective effects of reactive functional groups on chondrocytes in photocrosslinkable hydrogel systems. Acta Biomater. 2015;1(27):66–76.

El-Sherbiny IM, Yacoub MH. Hydrogel scaffolds for tissue engineering: progress and challenges. Glob Cardiol Sci Pract. 2013;2013(3):316.

PubMed   PubMed Central   Google Scholar  

Bryant SJ, Nuttelman CR, Anseth KS. Cytocompatibility of UV and visible light photoinitiating systems on cultured NIH/3T3 fibroblasts in vitro. J Biomater Sci Polym Ed. 2000;11(5):439–57.

Liu S, Borjigin T, Schmitt M, Morlet-Savary F, Xiao P, Lalevée J. High-performance photoinitiating systems for LED-induced photopolymerization. Polymers (Basel). 2023;15(2):342.

Fang Y, Eglen RM. Three-dimensional cell cultures in drug discovery and development. SLAS Discovery. 2017;22(5):456–72. https://doi.org/10.1177/1087057117696795 .

Sirenko O, Mitlo T, Hesley J, Luke S, Owens W, Cromwell EF. High-content assays for characterizing the viability and morphology of 3d cancer spheroid cultures. Assay Drug Dev Technol. 2015;13(7):402–14. https://doi.org/10.1089/adt.2015.655 .

Zustiak SP, Boukari H, Leach JB. Solute diffusion and interactions in cross-linked poly(ethylene glycol) hydrogels studied by Fluorescence Correlation Spectroscopy. Soft Matter. 2010;6(15):3609–18.

Shanbhag S, Woo Lee J, Kotov N. Diffusion in three-dimensionally ordered scaffolds with inverted colloidal crystal geometry. Biomaterials. 2005;26(27):5581–5.

Totti S, Allenby MC, Dos Santos SB, Mantalaris A, Velliou EG. A 3D bioinspired highly porous polymeric scaffolding system for in vitro simulation of pancreatic ductal adenocarcinoma. RSC Adv. 2018;8(37):20928–40.

Gupta P, Totti S, Pérez-Mancera PA, Dyke E, Nisbet A, Schettino G, Webb R, Velliou EG. Chemoradiotherapy screening in a novel biomimetic polymer based pancreatic cancer model. RSC Adv. 2019;9(71):41649–63.

Hamdi DH, Barbieri S, Chevalier F, Groetz JE, Legendre F, Demoor M, Galera P, Lefaix JL, Saintigny Y. In vitro engineering of human 3D chondrosarcoma: a preclinical model relevant for investigations of radiation quality impact. BMC Cancer. 2015;15(1):1–14. https://doi.org/10.1186/s12885-015-1590-5 .

Okere B, Alviano F, Costa R, Quaglino D, Ricci F, Dominici M, Paolucci P, Bonsi L, Iughetti L. In vitro differentiation of human amniotic epithelial cells into insulinproducing 3D spheroids. Int J Immunopathol Pharmacol. 2015;28(3):390–402.

Paris F, Marrazzo P, Pizzuti V, Marchionni C, Rossi M, Michelotti M, Petrovic B, Ciani E, Simonazzi G, Pession A, Bonsi L, Alviano F. Characterization of perinatal stem cell spheroids for the development of cell therapy strategy. Bioengineering. 2023;10(2):189.

Kim W, Park E, Yoo HS, Park J, Jung YM, Park JH. Recent advances in monitoring stem cell status and differentiation using nano-biosensing technologies. Nanomaterials. 2022;12(17):2934.

Burdick JA, Vunjak-Novakovic G. Engineered microenvironments for controlled stem cell differentiation. Tissue Eng Part A. 2009;15(2):205–19.

De León SE, Pupovac A, McArthur SL. Three-Dimensional (3D) cell culture monitoring: opportunities and challenges for impedance spectroscopy. Biotechnol Bioeng. 2020;117(4):1230–40. https://doi.org/10.1002/bit.27270 .

Sirenko O, Hancock MK, Hesley J, Hong D, Cohen A, Gentry J, Carlson CB, Mann DA. Phenotypic characterization of toxic compound effects on liver spheroids derived from ipsc using confocal imaging and three-dimensional image analysis. Assay Drug Dev Technol. 2016;14(7):381–94. https://doi.org/10.1089/adt.2016.729 .

Jensen C, Teng Y. Is it time to start transitioning from 2D to 3D cell culture? Front Mol Biosci. 2020;7:33.

Ajjarapu SM, Tiwari A, Kumar S. Applications and utility of three-dimensional in vitro cell culture for therapeutics. Future Pharmacol. 2023;3(1):213–28.

Basu A, Dydowiczová A, Trosko JE, Bláha L, Babica P. Ready to go 3D? A semi-automated protocol for microwell spheroid arrays to increase scalability and throughput of 3D cell culture testing. Toxicol Mech Methods. 2020;30(8):590–604. https://doi.org/10.1080/15376516.2020.1800881 .

Burdick JA, Murphy WL. Moving from static to dynamic complexity in hydrogel design. Nat Commun. 2012;3(1):1–8.

Nash ME, Healy D, Carroll WM, Elvira C, Rochev YA. Cell and cell sheet recovery from pNIPAm coatings; motivation and history to present day approaches. J Mater Chem. 2012;22(37):19376–89.

Abuwatfa WH, Awad NS, Pitt WG, Husseini GA. Thermosensitive polymers and thermo-responsive liposomal drug delivery systems. Polymers. 2022;14(5):925.

Young EWK, Beebe DJ. Fundamentals of microfluidic cell culture in controlled microenvironments. Chem Soc Rev. 2010;39(3):1036–48.

Guvendiren M, Burdick JA. Stiffening hydrogels to probe short- and long-term cellular responses to dynamic mechanics. Nat Commun. 2012;3.

Highley CB, Rodell CB, Burdick JA. Direct 3D printing of shear-thinning hydrogels into self-healing hydrogels. Adv Mater. 2015;27(34):5075–9.

Malda J, Visser J, Melchels FP, Jüngst T, Hennink WE, Dhert WJA, Groll J, Hutmacher DW. 25th anniversary article: engineering hydrogels for biofabrication. Adv Mater. 2013;25(36):5011–28.

Download references

Acknowledgements

The authors would like to acknowledge the financial support of the American University of Sharjah Faculty Research Grants, the Al-Jalila Foundation [AJF 2015555], the Al Qasimi Foundation, the Patient’s Friends Committee-Sharjah, the Biosciences and Bioengineering Research Institute [BBRI18-CEN-11], GCC Co-Fund Program [IRF17-003], the Takamul Program [P OC-00028-18], the Technology Innovation Pioneer (T IP) Healthcare Awards, Sheikh Hamdan Award for Medical Sciences [MRG-57-2019-2020], and the Dana Gas Endowed Chair for Chemical Engineering. We also would like to acknowledge student funding from the Material Science and Engineering Ph.D. program at AUS.

Author information

Authors and affiliations.

Materials Science and Engineering Ph.D. Program, College of Arts and Sciences, American University of Sharjah, P.O. Box. 26666, Sharjah, United Arab Emirates

Waad H. Abuwatfa & Ghaleb A. Husseini

Department of Chemical and Biological Engineering, College of Engineering, American University of Sharjah, P.O. Box 26666, Sharjah, United Arab Emirates

Department of Chemical Engineering, Brigham Young University, Provo, UT, 84602, USA

William G. Pitt

You can also search for this author in PubMed   Google Scholar

Contributions

WHA drafted the manuscript. GAH and WGP reviewed and edited the manuscript. All authors read and approved the final version.

Corresponding author

Correspondence to Ghaleb A. Husseini .

Ethics declarations

Competing interests.

The authors declare no competing interest.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Abuwatfa, W.H., Pitt, W.G. & Husseini, G.A. Scaffold-based 3D cell culture models in cancer research. J Biomed Sci 31 , 7 (2024). https://doi.org/10.1186/s12929-024-00994-y

Download citation

Received : 03 August 2023

Accepted : 04 January 2024

Published : 14 January 2024

DOI : https://doi.org/10.1186/s12929-024-00994-y

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Three-dimensional (3D) cell culture
  • Decellularized tissues
  • Microfluidics
  • Extracellular matrix (ECM)

Journal of Biomedical Science

ISSN: 1423-0127

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

cancer cell research studies

U.S. flag

An official website of the United States government

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.

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.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Cancer Cell Lines Are Useful Model Systems for Medical Research

Affiliations.

  • 1 IRCCS SDN, 80143 Naples, Italy. [email protected].
  • 2 IRCCS SDN, 80143 Naples, Italy.
  • PMID: 31374935
  • PMCID: PMC6721418
  • DOI: 10.3390/cancers11081098

Cell lines are in vitro model systems that are widely used in different fields of medical research, especially basic cancer research and drug discovery. Their usefulness is primarily linked to their ability to provide an indefinite source of biological material for experimental purposes. Under the right conditions and with appropriate controls, authenticated cancer cell lines retain most of the genetic properties of the cancer of origin. During the last few years, comparing genomic data of most cancer cell lines has corroborated this statement and those that were observed studying the tumoral tissue equivalents included in the The Cancer Genome Atlas (TCGA) database. We are at the disposal of comprehensive open access cell line datasets describing their molecular and cellular alterations at an unprecedented level of accuracy. This aspect, in association with the possibility of setting up accurate culture conditions that mimic the in vivo microenvironment (e.g., three-dimensional (3D) coculture), has strengthened the importance of cancer cell lines for continuing to sustain medical research fields. However, it is important to consider that the appropriate use of cell lines needs to follow established guidelines for guaranteed data reproducibility and quality, and to prevent the occurrence of detrimental events (i.e., those that are linked to cross-contamination and mycoplasma contamination).

Keywords: biological samples; cell lines; leukemia; medical research; solid cancer.

PubMed Disclaimer

Conflict of interest statement

The authors declare no conflict of interest.

List of major breakthroughs in…

List of major breakthroughs in the historical progress of cancer cell lines.

Similar articles

  • Continuous hematopoietic cell lines as model systems for leukemia-lymphoma research. Drexler HG, Matsuo AY, MacLeod RA. Drexler HG, et al. Leuk Res. 2000 Nov;24(11):881-911. doi: 10.1016/s0145-2126(00)00070-9. Leuk Res. 2000. PMID: 11086173 Review.
  • Translational Metabolomics of Head Injury: Exploring Dysfunctional Cerebral Metabolism with Ex Vivo NMR Spectroscopy-Based Metabolite Quantification. Wolahan SM, Hirt D, Glenn TC. Wolahan SM, et al. In: Kobeissy FH, editor. Brain Neurotrauma: Molecular, Neuropsychological, and Rehabilitation Aspects. Boca Raton (FL): CRC Press/Taylor & Francis; 2015. Chapter 25. In: Kobeissy FH, editor. Brain Neurotrauma: Molecular, Neuropsychological, and Rehabilitation Aspects. Boca Raton (FL): CRC Press/Taylor & Francis; 2015. Chapter 25. PMID: 26269925 Free Books & Documents. Review.
  • False and mycoplasma-contaminated leukemia-lymphoma cell lines: time for a reappraisal. Drexler HG, Dirks WG, MacLeod RA, Uphoff CC. Drexler HG, et al. Int J Cancer. 2017 Mar 1;140(5):1209-1214. doi: 10.1002/ijc.30530. Int J Cancer. 2017. PMID: 27870004
  • Contaminated and misidentified cell lines commonly use in cancer research. Cheung ST, Chan SL, Lo KW. Cheung ST, et al. Mol Carcinog. 2020 Jun;59(6):573-574. doi: 10.1002/mc.23189. Epub 2020 Mar 20. Mol Carcinog. 2020. PMID: 32196757
  • Polyurethane foam scaffold as in vitro model for breast cancer bone metastasis. Angeloni V, Contessi N, De Marco C, Bertoldi S, Tanzi MC, Daidone MG, Farè S. Angeloni V, et al. Acta Biomater. 2017 Nov;63:306-316. doi: 10.1016/j.actbio.2017.09.017. Epub 2017 Sep 18. Acta Biomater. 2017. PMID: 28927931
  • Characterisation of Canine and Feline Breast Tumours, Their Metastases, and Corresponding Primary Cell Lines Using LA-REIMS and DESI-MS Imaging. Molnár A, Horkovics-Kováts GS, Kucsma N, Szegő Z, Tauber B, Egri A, Szkupien Z, Deák BA, McKenzie JS, Thuróczy J, Schäffer R, Schlosser G, Szakács G, Balog J. Molnár A, et al. Int J Mol Sci. 2024 Jul 15;25(14):7752. doi: 10.3390/ijms25147752. Int J Mol Sci. 2024. PMID: 39062995 Free PMC article.
  • Failure to progress: breast and prostate cancer cell lines in developing targeted therapies. James C, Whitehead A, Plummer JT, Thompson R, Badal S. James C, et al. Cancer Metastasis Rev. 2024 Jul 26. doi: 10.1007/s10555-024-10202-w. Online ahead of print. Cancer Metastasis Rev. 2024. PMID: 39060878 Review.
  • cancercelllines.org-a novel resource for genomic variants in cancer cell lines. Paloots R, Baudis M. Paloots R, et al. Database (Oxford). 2024 Apr 30;2024:baae030. doi: 10.1093/database/baae030. Database (Oxford). 2024. PMID: 38687868 Free PMC article.
  • Organoids as preclinical models of human disease: progress and applications. Chen B, Du C, Wang M, Guo J, Liu X. Chen B, et al. Med Rev (2021). 2024 Mar 14;4(2):129-153. doi: 10.1515/mr-2023-0047. eCollection 2024 Apr. Med Rev (2021). 2024. PMID: 38680680 Free PMC article. Review.
  • Induced pluripotent stem cells (iPSCs): molecular mechanisms of induction and applications. Cerneckis J, Cai H, Shi Y. Cerneckis J, et al. Signal Transduct Target Ther. 2024 Apr 26;9(1):112. doi: 10.1038/s41392-024-01809-0. Signal Transduct Target Ther. 2024. PMID: 38670977 Free PMC article. Review.
  • Masters J.R.W. Human cancer cell lines: Fact and fantasy. Nat. Rev. Mol. Cell Biol. 2000;1:233–236. doi: 10.1038/35043102. - DOI - PubMed
  • Neve R.M., Chin K., Fridlyand J., Yeh J., Baehner F.L., Fevr T., Clark L., Bayani N., Coppe J.P., Tong F., et al. A collection of breast cancer cell lines for the study of functionally distinct cancer subtypes. Cancer Cell. 2006;10:515–527. doi: 10.1016/j.ccr.2006.10.008. - DOI - PMC - PubMed
  • Drexler H.G., Matsuo Y., MacLeod R.A.F. Continuous hematopoietic cell lines as model systems for leukemia-lymphoma research. Leuk. Res. 2000;24:881–911. doi: 10.1016/S0145-2126(00)00070-9. - DOI - PubMed
  • Asfari M., Janjic D., Meda P., Li G.D., Halban P.A., Wollheim C.B. Establishment Of 2-mercaptoethanol-dependent differentiated insulin-secreting cell-lines. Endocrinology. 1992;130:167–178. doi: 10.1210/endo.130.1.1370150. - DOI - PubMed
  • Sramkoski R.M., Pretlow T.G., Giaconia J.M., Pretlow T.P., Schwartz S., Sy M.S., Marengo S.R., Rhim J.S., Zhang D.S., Jacobberger J.W. A new human prostate carcinoma cell line, 22R upsilon 1. In Vitro Cell. Dev. Biol. Animal. 1999;35:403–409. doi: 10.1007/s11626-999-0115-4. - DOI - PubMed

Publication types

  • Search in MeSH

Related information

Grants and funding.

  • Progetti di Ricerca Corrente/Ministero della Salute
  • 5 per Mille/IRCCS SDN

LinkOut - more resources

Full text sources.

  • Europe PubMed Central
  • PubMed Central

full text provider logo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Information

  • Author Services
  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials

Journal Description

Cancer cell research, journal menu.

  • Cancer Cell Research Home
  • Call for papers
  • Author Guideline
  • Editorial Board
  • About Journal

Journal Browser

  • 01   (2014)
  • 02   (2014)
  • 03   (2014)
  • 04   (2014)
  • 05   (2015)
  • 06   (2015)
  • 07   (2015)
  • 08   (2015)
  • 09   (2016)

Highly Accessed Articles

Latest books.

Ready to start planning your care?  Call us at 800-525-2225 to make an appointment.

MSK Research Highlights, July 31, 2024

Wednesday, July 31, 2024

Detail shot of a scientist purifying proteins

New research from Memorial Sloan Kettering Cancer Center (MSK) showed giving chemotherapy shortly after a stem cell or bone marrow transplant from a less than perfectly matched donor greatly reduces the chances that the patient will develop graft-versus-host disease (GVHD); and sheds new light on cell state changes in prostate cancer.

Improving transplant success in patients without fully matched blood stem cell donors

Stem cell and bone marrow transplants (BMTs) provide the possibility of a cure for certain types of blood cancer, including many cases of  leukemia and  lymphoma . But for those transplants that require blood stem cells from a donor — known as  allogeneic transplants — it can be hard to find an exact match. This is especially true for patients with certain genetic ancestries, especially African, Latin American, and Asian. Their tissues, or HLA types, are more diverse.

In a new retrospective study, MSK  bone marrow transplant specialist and cellular therapist Brian Shaffer, MD ,  and a multicenter team of investigators demonstrated how to improve transplants in patients who are not able to find fully matched donors. The study was overseen and funded by the  National Marrow Donor Program (NMDP). The research confirmed that for patients receiving only partially matched cells, giving chemotherapy shortly after the transplant greatly reduces the chances that the patient will develop  graft-versus-host disease (GVHD) , a potentially fatal side effect of BMTs that occurs when the donor’s immune system attacks the recipient’s healthy tissues.

The analysis included data from more than 10,000 BMT recipients, nearly half of whom received cyclophosphamide chemotherapy after their transplant. (The rest received anti-rejection drugs called calcineurin inhibitors, or CNIs.) Patients who received chemotherapy after their transplant had better outcomes than those who received CNIs. In fact, outcomes for patients with partially matched donors who received chemotherapy were almost as good as for those who had fully matched donors. An  ongoing clinical trial at MSK is now looking at whether lower doses of chemotherapy are as effective as higher doses at preventing GVHD and whether they reduce the side effects of chemotherapy.

Read more in the  Journal of Clinical Oncology .

Single cell analysis sheds light on cell state changes in prostate cancer

Cancer can resist treatment by  changing its identity , a transformation known as lineage plasticity. A notable example is when prostate adenocarcinoma, the most common type, transforms into neuroendocrine prostate cancer (NEPC), a more aggressive form. It is important to understand how this plasticity affects the expression of cell surface antigens – especially because promising new drugs are being tested that  target specific surface antigens . A team led by MSK  physician-scientist Charles Sawyers, MD , and  Michael Haffner, MD, PhD , of the Fred Hutchinson Cancer Center used immunohistochemistry and single-cell analysis of treatment-resistant prostate cancer and found a surprisingly high variety of cell states. This heterogeneity is not captured by conventional histology-based methods. They further showed that the cell states could be identified by gene regulatory networks that could improve diagnostic precision and help in the selection of patients who will respond to therapies aimed at specific cell surface antigens. Read more in the  Proceedings of the National Academies of Sciences .

National Cancer Institute - Cancer.gov

Dietary glutamine may be linked to B-cell lymphomas in abdominal lymph nodes

B cells

B cells migrating outside of the center of the B cell follicle in mice without GNA13 . Image provided by Jagan Muppidi, M.D., Ph.D.

Researchers have discovered that dietary glutamine, an amino acid found in many foods, may be linked to the development of certain kinds of B-cell lymphomas in abdominal structures called mesenteric lymph nodes in mouse models. The study appeared July 18, 2024, in Nature Immunology .

B-cell lymphoma is a common form of cancer of the lymphatic system. It can originate from B cells at different stages of activation. These diverse origins may contribute to how patients respond to therapies.

“One of the field’s bigger efforts over the last 20 years has been trying to define what targeted therapies might be effective in different forms of this common lymphoma,” said Jagan R. Muppidi, M.D., Ph.D. , Stadtman Investigator in the Lymphoid Malignancies Branch , and the lead investigator of the study.

The investigators focused on germinal center B cells. These cells are responsible for making antibodies in germinal centers, which are clusters of immune cells that form in lymph nodes. Patients with cancers that are derived from germinal center B cells commonly have loss of function mutations in a gene called GNA13 , which encodes the protein G-alpha-13. This protein typically inhibits migration of germinal center B cells.

The researchers used mouse models that lacked the G-alpha-13 protein in B cells and found that these mice tended to develop lymphoma from germinal centers within mesenteric lymph nodes in the abdomen. Additionally, they found gene expression studies showed the absence of G-alpha-13   led to increased activity of a protein called mTOR and increased expression of the protein MYC, both of which are involved in cell growth and proliferation, specifically in germinal center B cells in mesenteric lymph nodes.

“Part of the novelty of our study is that it provides evidence for different anatomic locations being linked to different biology in lymphoma,” said Muppidi. “At present, there are no comprehensive databases that link mutations in lymphoma to the sites of disease, so it isn’t clear whether people with tumors that have GNA13 mutations are more likely to develop disease in the mesenteric lymph nodes as the mice did.”

The researchers next explored possible mechanisms that could explain why the mice developed cancer in this particular location. Within the gut, the lymphatic system transports nutrients such as fats, proteins and vitamins from the intestines to the mesenteric lymph node. Since the G-alpha-13 protein typically confines germinal center B cells to the center of the B cell follicle, researchers hypothesized that these B cells in mice without G-alpha-13 could be migrating outside of the center of the cells and be exposed to substances in the lymphatic system that originated in the gut.

The researchers did discover that this was the case, and specifically found that diet-derived glutamine supported increased levels of MYC and mTOR activity in G-alpha-13-deficient germinal center B cells, leading to increased proliferation.

“We are interested in more completely defining the molecular pathway by which dietary glutamine and potentially other nutrients support the expression of MYC  in G-alpha-13-deficient cells,” Muppidi said. These findings are only applicable in mouse models today, but if extended into future human studies, this research could eventually identify therapeutic targets for G-alpha13-deficient lymphoma. Next, the researchers hope to learn whether alteration of dietary glutamine or other dietary modifications can suppress the progression of existing lymphoma in animal models. 

Related News

Acute myelocytic leukemia

Clinical trial researching CAR T-cell therapy for acute myeloid leukemia

Patient with doctor

Clinical trial researching immunotherapy for cervical cancer

metastatic tumor CT

New immunotherapy approach shows potential in some people with metastatic solid tumors, NIH researchers say

U.S. flag

An official website of the United States government

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.

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.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • SAGE Open Med

New approaches and procedures for cancer treatment: Current perspectives

Dejene tolossa debela.

1 Center for Innovative Drug Development and Therapeutic Trials for Africa (CDT-Africa), College of Health Sciences, Addis Ababa University, Addis Ababa, Ethiopia

Seke GY Muzazu

2 Enteric Diseases and Vaccines Research Unit, Centre for Infectious Disease Research in Zambia (CIDRZ), Lusaka, Zambia

Kidist Digamo Heraro

3 Wachemo University, Hossana, Ethiopia

Maureen Tayamika Ndalama

Betelhiem woldemedhin mesele.

4 Kotebe Metropolitan University, Addis Ababa, Ethiopia

Dagimawi Chilot Haile

5 University of Gondar, Gondar, Ethiopia

Sophia Khalayi Kitui

Tsegahun manyazewal.

Cancer is a global health problem responsible for one in six deaths worldwide. Treating cancer has been a highly complex process. Conventional treatment approaches, such as surgery, chemotherapy, and radiotherapy, have been in use, while significant advances are being made in recent times, including stem cell therapy, targeted therapy, ablation therapy, nanoparticles, natural antioxidants, radionics, chemodynamic therapy, sonodynamic therapy, and ferroptosis-based therapy. Current methods in oncology focus on the development of safe and efficient cancer nanomedicines. Stem cell therapy has brought promising efficacy in regenerating and repairing diseased or damaged tissues by targeting both primary and metastatic cancer foci, and nanoparticles brought new diagnostic and therapeutic options. Targeted therapy possessed breakthrough potential inhibiting the growth and spread of specific cancer cells, causing less damage to healthy cells. Ablation therapy has emerged as a minimally invasive procedure that burns or freezes cancers without the need for open surgery. Natural antioxidants demonstrated potential tracking down free radicals and neutralizing their harmful effects thereby treating or preventing cancer. Several new technologies are currently under research in clinical trials, and some of them have already been approved. This review presented an update on recent advances and breakthroughs in cancer therapies.

Introduction

Cancer is a global health problem responsible for one in six deaths worldwide. In 2020, there were an estimated 19.3 million new cancer cases and about 10 million cancer deaths globally. Cancer is a very complicated sequence of disease conditions progressing gradually with a generalized loss of growth control. 1 – 3 There were only a few options of cancer treatment for patients for many decades which include surgery, radiation therapy, and chemotherapy as single treatments or in combination. 4 , 5 But recently, many pathways involved in cancer therapy progression and how they can be targeted has improved dramatically, with combinatorial strategies, involving multiple targeted therapies or “traditional” chemotherapeutics, such as the taxanes and platinum compounds, being found to have a synergistic effect. 6 New approaches, such as drugs, biological molecules, and immune-mediated therapies, are being used for treatment even if the excepted therapy level has not reached that resists the mortality rate and decreases the prolonged survival time for metastatic cancer.

The creation of a new revolution in neoplastic cancer or targeting drugs depends on the pathways and characteristics of different tumor entities. 7 Chemotherapy is considered the most effective and widely used modality in treating cancers as used alone or in combination with radiotherapy. Genotoxicity is how chemotherapy drugs target the tumor cells mainly producing reactive oxygen species that largely destroy tumor cells. 8 Hormonal treatments are also widely used for cancer malignancies and considered as cytostatic because it restricts tumor development by limiting the hormonal growth factors acting through the direction of hypothalamic–pituitary–gonadal axis (HPGA), hormone receptor blockage, and limiting of adrenal steroid synthesis. 9

In this narrative review, a general overview of the most advanced and novel cancer therapies was provided. In addition, also new strategies currently under investigation at the research stage that should overwhelm the drawbacks of standard therapies; different strategies to cancer diagnosis and therapy; and their current status in the clinical context, underlining their impact as innovative anti-cancer approaches.

Cancer treatment modalities

We can see cancer treatment modalities by dividing them into conventional (traditional) and advanced or novel or modern categories. In this era worldwide, over half of all ongoing medical treatment trials are focusing on cancer treatments. 7 Entities, such as the type of cancer, its site, and severity, guide to select treatment options and its progress. The most widely used traditional treatment methods are surgery, chemotherapy, and radiotherapy, while modern modalities include hormone therapy, anti-angiogenic, stem cell therapies, immunotherapy, and dendritic cell-based immunotherapy. 10

Conventional cancer therapies

The most recommended conventional cancer treatment strategies include surgical resection of the tumors followed by radiotherapy with x-rays and/or chemotherapy. 11 Of these modalities, surgery is most effective at an early stage of disease progression. Radiation therapy can damage healthy cells, organs, and tissues. Although chemotherapy has reduced morbidity and mortality, virtually all chemotherapeutic agents damage healthy cells, especially rapidly dividing and growing cells. 12 Drug resistance, a major problem with chemotherapy, is a phenomenon wherein cancer cells that initially were suppressed by an anti-cancer drug develop resistance to the drug. This is caused primarily by reduced drug uptake and increased drug efflux. 13 Limitations of conventional chemotherapeutic modality, such as dosage selection difficulty, lack of specificity, rapid drug metabolism, and mainly harmful side effects. 14

Advanced and innovative cancer therapies

Among the obstacles of cancer, drug resistance and its delivery systems are the most problem in cancer cure and decreasing signs and symptoms; but currently, there are many approved treatment approaches and drugs. The efficiency of conventional cancer is reduced due to tumor pathology and architectural abnormality of tumor tissue blood vessels. 15 The following are the advanced and innovative cancer therapy types with their benefits and challenges.

Stem cells therapy

Stem cells are undifferentiated cells present in the bone marrow (BM) with an ability to differentiate into any type of body cell. Stem cell therapeutic strategy is also one of the treatment options for cancer which are considered to be safe and effective. Application of stem cell is yet in the experimental clinical trial; for example, their use in the regeneration of other damaged tissue is being explored. Mesenchymal stem cells (MSCs) are currently being used in trials that are delivered from the BM, fat tissues, and connective tissues. 16

Pluripotent stem cells

Embryonic stem cells (ESCs) isolated from the uniform inner mass cells of the embryo possess the flexibility to administer rise to any or all kinds of cells except those within the placenta. In 2006, the invention of Yamanaka factors to induce pluripotent stem cells (iPSCs) from physical cells in a culture marked a breakthrough in cell biology. 17 Avoiding ethical issues from embryo destruction, iPSCs and ESCs have the same characteristics. Hematopoietic embryonic stem cells (hESCs) and iPSCs are currently used for the induction of effector T cells and natural killer (NK) cells, 18 and anti-tumor vaccine preparation. 19

Adult stem cells

Adult stem cells (ASCs) groups often used in tumor therapy include hematopoietic stem cells (HSCs), MSCs, and neural stem cells (NSCs). HSCs, located in BM, can form all mature blood cells in the body. Currently, only approved by the Food and Drug Administration (FDA) is the infusion of HSCs derived from cord blood to treat multiple myeloma and leukemia. 20 MSCs are found in many tissues and organs, playing important roles in tissue repair and regeneration into cells, such as osteocytes, adipocytes, and chondrocytes. MSCs have special biological characteristics and are used as complimentary with other approaches in treating tumors. 21 NSCs can self-renew and generate new neurons and glial cells and are used for treating both primary and metastatic breast and other tumors. 22

Cancer stem cells

Cancer stem cells (CSCs) are generated in normal stem cells or precursor/progenitor cells by the epigenetic mutations process. Their role in tumor treatment includes cancer growth, metastasis, and recurrence, so that it could give promise in the treatment of solid tumors. 23 Stem cells have several action mechanisms in treating the tumor. The homing process is one mechanism which is a rapid migration of HSCs into defined stem cell niches in BM after that the transplants undergo the engraftment process before giving rise to specialized blood cells. This mechanism is dependent on the active interaction between stem cell CXCR4 receptors and requires their interaction with endothelial cells through LFA-1, VLA-4/5, CD44, and the secretion of matrix degradable enzyme MMP-2/9. 22 The second mechanism is the tumor-tropic effect in which the migration of MSCs toward tumor microenvironment (TM) after attraction by CXCL16, SDF-1, CCL-25, and IL-6 secreted by tumor cells and differentiation of MSCs within the tumor cells which contributes to tumor stromal development. 24 Stem cells also act by paracrine factor secretion, including extracellular vesicles (EVs) and soluble materials, 25 and their differentiation capacity, such as transplanted HSCs, can give rise to all blood cell types. 26

Generally, cancer treatment using stem cell therapy by various strategies, including transplantation of HSC, 27 MSC infusion, 28 therapeutic carriers, 29 generation of immune effector cells, 30 and vaccine production. 31 The stem cell cancer therapy approach confronted the following side effects: (1) tumorigenesis, (2) adverse events in allogeneic HSC transplantation, (3) drug toxicity and drug resistance, (4) increased immune responses and autoimmunity, and (5) viral infection. 22 Despite several successes, there are challenges, such as therapeutic dose control, low cell targeting, and retention in tumor sites, that should be investigated and overcome in the future. In addition, existing results from stem cell technologies are highly encouraging for tumor treatment but it still needs further efforts to improve the safety and efficacy before they could enter clinical trials. Table 1 summarized the licensed list of stem cell therapies.

Licensed stem cell therapies.

S. no.Stem cell therapiesExamplesAuthorityIndication
01Pluripotent stem cellsiPSC (sipuleucel-T)FDAProstate cancer
02Adult stem cellsMSC-INFβFDAOvarian tumor
03Cancer stem cellsVenetoclaxFDAAML

AML: acute myelogenous leukemia; FDA: The US Food and Drug Administration; iPSC: induced pluripotent stem cell; MSC-INβ: mesenchymal stem cells with interferon beta.

Targeted drug therapy

Targeted cancer therapies are drugs or other substances which are sometimes interchangeably used as “molecularly targeted drugs,” “molecularly targeted therapies,” and “precision medicines.” Those drugs’ mechanism of action is by interfering with growth molecules which leads to blocking the growth and spreading of cancer. 34 Tumor initiation and progression are determined by the TM of an atypical tumor which comprises endothelial cells, pericytes, smooth muscle cells, fibroblasts, various inflammatory cells, dendritic cells, and CSCs. There are various signaling mechanisms and pathways that TM-forming cells dynamically interact with the cancerous cells which are suitable for sustaining a reasonably high cellular proliferation. So, it is the area of research interest using TM conditions to mediate effective targeting measures for cancer therapy. 35

Selectively treating cancer cells with conventional chemotherapy is difficult since it is similar to normal cells. So those problems are intervened by cellular mechanisms, such as cell cycle arrest, apoptosis induction, proliferation prevention, and interfering with metabolic reprogramming by targeted drug therapy agents. 36 Modifying TM and targeting TM for drug delivery for effective treatment are two strategies that can be used for the treatment of cancer. 37 Targeted therapy drugs do work in different ways from standard chemotherapy drugs treatment like attacking cancer cells while doing less damage to normal cells which is a programming that sets them apart from normal, healthy cells. 38

Using targeted therapy markedly increased the survival rate for some diseases, for example, from 17% to 24% in patients with advanced pancreatic cancer, the addition of erlotinib to standard chemotherapy. Imatinib has had a dramatic effect on chronic myeloid leukemia, and rituximab, sunitinib, and trastuzumab have revolutionized the treatment of renal cell carcinoma and breast cancer, respectively. 39

We can classify the targeted cell agents based on the mechanism of their work or their target site. Some enzymes serve as signals for cancer cells to grow. Some targeted therapies inhibit enzymes that are signals for cancer cells to grow. These drugs are called enzyme inhibitors. Blocking these cell signals can inhibit cancer from getting bigger and spreading. 40

Some targeted therapies are called apoptosis-inducing drugs because they are aimed right at the parts of the cell that control whether cells live or die. The examples are serine/threonine kinase, protein kinase B (PKB/Akt), which promotes cell survival, and inhibitors of this protein are in the preclinical phase. 41

These agents stop the tumors from making new blood vessels which helps cut off the tumors’ blood supply so that tumors cannot grow. In addition, they arrest tumor growth that involves by curtailing blood supply to the tumor by inhibiting angiogenic factors, such as vascular endothelial growth factor (VEGF) or its receptors. The study showed the survival of patients with advanced colorectal carcinoma extended by months after the use of Avastin (bevacizumab) in combination with 5-fluorouracil-based chemotherapy. 42

Types of target agents

Monoclonal antibodies.

Antibody drugs are man-made versions of immune system proteins administered intravenously to attack certain targets on cancer cells. They contain a more proportion of human components than murine components. 43 Their attack mechanisms of action are recruiting host immune functions to attack the target cell, binding to ligands or receptors thereby interrupting essential cancer cell processes, and carrying a lethal payload, such as radioisotope or toxin, to the target cell. 44 Gemtuzumab is an example of a CD-33-specific monoclonal antibody currently used for AML treatment by conjugating with calicheamicin. 45 In addition, ibritumomab tiuxetan is an anti-CD20, a 90Y metal isotope-based is developed in clinical therapy. 46 Delivery of active therapeutics, prodrug activation enzymes, and chemotherapy toxins are also another use of target agents of monoclonal antibodies. 47

Small molecule inhibitors

These are smaller protein in size (⩽500 Da) than monoclonal antibodies, so that they can simply translocate through plasma membranes and can be taken orally. Their main function is interrupting cellular processes by interfering with the intracellular signaling of tyrosine kinases which leads to the inhibition of tyrosine kinase signaling and initiates a molecular cascade that can lead to the inhibition of cell growth, proliferation, migration, and angiogenesis in malignant tissues. 48 Examples of small molecule inhibitors are gefitinib and erlotinib which inhibit epidermal growth factor receptor (EGFR) kinase and EGFR in non-small cell lung cancer (NSCLC) patients, respectively. There are also lapatinib and sorafenib which act on the inhibition of EGFR/Erb-B2 Receptor Tyrosine Kinase 2 (ERBB2) for ERBB2-positive breast cancer and VEGFR kinase, in renal cancer. 49

Ablation cancer therapy

Ablation is a treatment technique that destroys tumors without removing them mostly indicated for small-size tumors of less than 3 cm and the surgical option is contraindicated. Ablation is also used with embolization for larger tumors. However, this technique might not be indicated for treating tumors near major blood vessels, the diaphragm, or major bile ducts due to destroying some of the normal tissue around the tumor. 50

Thermal ablation

This technique uses extreme hyperthermia or hypothermia to destruct tumor tissue concentrating on a focal zone in and around the tumor. Similar to surgery, thermal ablation removes the tumor and a 5–10 mm thick margin of seemingly normal tissue but the tissue is killed in situ and then absorbed by the body later. The procedure is similar to surgery using an open, laparoscopic, or endoscopic approach but is commonly applied using a percutaneous or non-invasive approach. The type of tumor, site, physician’s choice, and health status determine the approach. 51

Radiofrequency ablation (RFA), microwave ablation, high-intensity focused ultrasound, and cryoablation are currently being used in the clinical setting. Cryoablation uses a hypothermic modality to induce tissue damage by a freeze-thaw process against others. All these treatments operate on the principle of hyperthermia except cryoablation. Of all the ablation techniques, cryoablation demonstrated the highest potential to elicit a post-ablative immunogenic response. 52

Recent studies showed additional to tissue disruption RFA and cryoablation can modulate the immune system that they were applied as therapy on TM and in the systematic circulation. Evidence has shown that ablation procedures affect carcinogenesis due to its local inflammatory response leading to an immunogenic gene signature. 53

The advantage of this procedure over surgery is that it provides a minimal (e.g. percutaneously or laparoscopically) or non-invasive approach to cancer therapy and gets attention as an alternative to standard surgical therapies. 54

Cryoablation

Cryoablation is one of the ablation techniques which ablates the extensive tissue by freezing to lethal temperatures followed by liquid formation, causing extensive tissue. Benign and malignant primary tumors are mostly treated by this therapy. 55 James Arnott reported that the freezing temperatures can impair cancer cell viability after he attempted the usage of cold temperatures by salt and ice solutions for the generation of local numbness before surgical operations in the nineteenth century. He suggested cryoablation as an attractive therapeutic option and increased a patient’s survival. 56

Cryoablation techniques are based on the principle of the Joule–Thomson effect which was studied in the 1930s by many researchers and concluded using liquid CO 2 under high pressure, liquid air, and liquid oxygen to achieve the cooling effect and the subsequent formation of ice crystals so employed to treat lesions, warts, and keratosis. However, after 1950, Allington replaced liquid N 2 for the treatment of various skin lesion disorders. 47

RFA therapy

RFA is a minimally invasive procedure and an image-guided technique using hyperthermic (high-frequency electrical currents) conditions to destroy cancer cells. Imaging techniques, such as ultrasound, computed tomography (CT), or magnetic resonance imaging (MRI), guide needle electrodes into a tumor cell. Generally, RFA is the most effective approach for treating small-size tumors of less than 3 cm in diameter. RFA can be used in combination with other conventional cancer treatment options. 57 After starting the use of deployable devices or multiple-electrode systems, RFA can treat medium tumors (up to 5 cm diameter). 58

Gene therapy

Gene therapy is the insertion of a normal copy of a defective gene in the genome to cure a specific disorder. The first application dates back to 1990 when a retroviral vector was exploited to deliver the adenosine deaminase (ADA) gene to T cells in patients with severe combined immunodeficiency (SCID). Approximately, about 2900 gene therapy clinical trials are currently ongoing, two-third of which are related to cancer. Strategies, such as expression of proapoptotic and chemosensitizing genes, expression of wild-type tumor suppressor genes, expression of genes able to solicit specific anti-tumor immune responses, and targeted silencing of oncogenes, are under evaluation for cancer gene therapy. 47

Thymidine kinase (TK) gene delivery is effective for the administration of prodrug ganciclovir to activate its expression and induce specific cytotoxicity. 59 The p53 tumor suppressor gene which is vectors carrying has been assessed for the clinical purpose very recently. ONYX-015 has been tested in NSCLC patients and gave a high response rate when given alone or combined with chemotherapy. 60 Gendicine, a recombinant adenovirus carrying wild-type p53-induced complete disease regression in head and neck squamous cell cancer had similar success when combined with radiotherapy. 61

Some challenges that have been faced with gene therapy are the selection of the right conditions and the choice of the best delivery mechanism. Identified drawbacks of this therapy are genome integration, limited efficacy in specific subsets of patients, and high chances of being neutralized by the immune system. Basic research and medical translation used RNA interference (RNAi) as an efficient technology that able to produce targeted gene silencing. 62 RNA-induced silencing complex (RISC) mediates the targeted gene silencing process by cleaving the messenger RNA (mRNA) and interference with protein synthesis. 63 A siRNAs can be designed to block desired targets, involving cell proliferation and metastatic invasion; hence, precise molecular mechanisms are a triggering factor for tumor formation. This method relies on siRNA-mediated gene silencing of anti-apoptotic proteins, transcription factors (i.e. c-myc gene), 64 , 65 or cancer mutated genes (i.e. K-RAS). 66

Advantages of siRNA-based drugs are safety, high efficacy, specificity, few side effects, and low costs of production. 67 However, occasionally, they can induce off-target effects or elicit innate immune responses, followed by specific inflammation. 68 Delivery methods currently under study are chemical modification (insertion of a phosphorothioate at 3’ end, introduction of a 2’ O-methyl group, and modification by 2,4-dinitrophenol) and lipid encapsulation, or conjugation with organic molecules (polymers, peptides, lipids, antibodies, small molecules) efficiently target to spontaneously cross cell membranes of naked siRNAs. 69 Interaction of cationic liposomes with negatively charged nucleic acids facilitates easy transfection by simple electrostatic interactions. 70 They can be constituted by 1,2-dioleoyl-3-trimethylammonium propane (DOTAP) and N-[1-(2,3-dioleoyloxy) propyl]-N, Ntrimethylammonium methyl sulfate (DOTMA). 71 Currently, a Phase I clinical trial is recruiting patients for evaluating the safety of Eph receptor A2 (EphA2) targeting 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) encapsulated siRNA (siRNA-EphA2- DOPC) in patients with advanced and recurrent cancer. 72 siRNAs can be concentrated in cationic polymers, such as chitosan, cyclodextrin, and polyethyleneimine (PEI). 73 CALAA-01 is one of the cyclodextrin polymers conjugated with human transferrin is being entered a Phase I clinical trial. PEI has been used as an anti-cancer by forming small cationic nanoparticles and loading with human epidermal growth factor receptor 2 (HER-2 receptor)-specific siRNA. 74 Phase II clinical trial has been started to evaluate Local Drug EluteR (siG12D LODER) directed to mutated Kirsten rat sarcoma (K-RAS) oncogene for the treatment of advanced pancreatic cancer. Conjugating to peptides, antibodies, and aptamers improves stability during circulation and enhances cellular uptake of siRNAs. 75 The introduction of nanocarriers has largely improved siRNAs stability, pharmacokinetics and biodistribution properties, and targeting specificity. Polyallylamine phosphate nanocarriers have been developed to release siRNAs in the cytoplasm after disassembly at low endosomal pH. 76

Dose correction and variabilities between individuals and different stages of disease are challenging issues on clinical translation of the siRNA-based approach. In the future, the needed research is on setting up the best-personalized therapy and toward controlled release to reach only specific targets on treating the tumor. Table 2 summarizes the gene therapy drugs based on their mechanism of action and induction.

Summary of gene therapy approaches.

S. no.Gene therapyMechanism of actionCategory
01Oncolytic virotherapy (OV)Directly lyses tumor cells and introduces wild-type tumor suppressor genes into cellsNaturally occurring or genetically modified virusesTumor immunotherapy
02GendicineInduces the expression of p53, restores its activity, and destroys the tumor cellsNon-replicative adenoviral vectorNeck and head squamous cell carcinoma
03Oncorine (rAd5-H101)Causes oncolysisReplicative, oncolytic recombinant ad5Refractory nasopharyngeal cancer
04ImlygicCauses apoptosis of tumor cellGenetically modified oncolytic HSV-1Non-resectable metastatic melanoma
05Rexin-GInhibits cell cycle in the G1 phaseReplication-incompetent retroviral vectorMetastatic cancers
06KymriahInitiates the anti-tumor effect through CD3 domainCAR T cell-based geneRelapsed B-cell acute lymphoblastic leukemia
07ZalmoxisEnhances immune reconstitutionAllogeneic hematopoietic stem cell transplantation (allo-HSCT)Hematopoietic malignancies

Natural antioxidants

Day to day, the anatomy undergoes many exogenous insults, such as ultraviolet (UV) rays, pollution, and tobacco smoke, that end in the assembly of reactive species, particularly oxidants and free radicals, liable for the onset of many diseases, together with cancer. These molecules can even be made as a consequence of clinical administration of medication; however, they are additionally naturally created within our cells and tissues by mitochondria and peroxisomes, and from macrophages metabolism, throughout traditional physiological aerobic processes. 47

Oxidative stress and radical oxygen species can significantly change the regulation of transcription factors by damaging the DNA and other bio-macromolecule. 77

Vitamins, polyphenols, and plant-derived bioactive compounds are natural antioxidants used as preventive and therapeutic drugs against these molecules that damage the body due to their anti-inflammatory and antioxidant properties. 78 Studies added to cancer therapy after appreciating their anti-proliferative and proapoptotic properties. Compounds, such as vitamins, alkaloids, flavonoids, carotenoids, curcumin, berberine, quercetin, and others, are examples of natural antioxidants screened in vitro and in vivo. 79

Limited bioavailability and/ or toxicity is one of the challenges of natural drugs while their translation into clinical practice. 47 Curcumin has cytotoxic effects in different kinds of tumors, such as the brain, lung, leukemia, pancreatic, and hepatocellular carcinoma, 80 while sparing normal cells at effective therapeutic doses. The curcumin’s biological properties, treatment duration and efficient therapeutic doses are under study. 80 This day, about 27 clinical trials are done, while 40 are under study on curcumin.

Berberine is an alkaloid compound that has been studied to be effective against different cancers as a chemopreventive agent, modulating many signaling pathways. Different nanotechnological strategies have been developed to facilitate its delivery across cell membranes due to their poorly soluble in water. 81 Six clinical trials are under study and two have been completed.

Quercetin is another natural plant origin agent that is effective alone and also in combination with chemotherapeutic agents in treating many cancers, such as lung, prostate, liver, colon, and breast cancers. 82 Quercetin’s mechanism of action is by binding to cellular receptors and the interference of several signaling pathways. 83 Currently, six clinical trials are under study and seven studies have been completed.

Current clinical trials

In recent years, analysis of cancer medication has taken outstanding steps toward more practical, precise, and fewer invasive cancer treatments in the research of clinical trials ( Figure 1 ).

An external file that holds a picture, illustration, etc.
Object name is 10.1177_20503121211034366-fig1.jpg

Current status of clinical trials of innovative and novel strategies of cancer therapy.

Currently, the most frequent entries focusing on cancer therapies in the database of clinical trials ( www.clinicalTrials.gov ) include the terminologies stem cell, targeted therapy, immunotherapy, and gene therapy because they are very promising and effective. Table 3 summarizes the potential advantages and disadvantages of the new treatment approaches.

Comparison of advantageous and disadvantageous of new cancer therapies.

S. no.Treatment approachAdvantagesDisadvantages
01Stem cell therapySafe and effective
Can be combined with other strategies
Decreases tumor volumes and extend survival
Treatment not durable
Potential tumorigenesis
02Targeted therapyHigh specificity
Reduced adverse reactions
Long-term side effects in question
03Ablation therapyPrecise treatment
Possibility to perform along with MRI imaging (magnetic hyperthermia)
Efficiency mainly to localized areas
Low penetration power
Needs skilled operator
04Gene therapyExpression of proapoptotic and chemosensitizing genes
Expression of wild-type tumor suppressor genes
Expression of genes able to solicit specific anti-tumor immune responses
Targeted silencing of oncogenes and safety (RNAi)
Genome integration
Limited efficacy in specific subsets of patients
High chances to be neutralized by the immune system
Off-target effects and inflammation (RNAi)
Need for ad hoc delivery systems (RNAi)
Setup of doses and suitable conditions for controlled release (RNAi)
05Natural antioxidantsEasily available in large quantities
The exploitation of their intrinsic properties
Limited bioavailability
Possible toxicity

MRI: magnetic resonance imaging.

Table 4 summarizes the approaches to advanced cancer therapies and their respective delivery systems with examples.

Advanced therapy approaches and delivery systems.

S. no.Types of therapyDelivery systemExample
01Stem cellNanoparticles Hyaluronic acid (HA)
Polyvinyl alcohol
02Immune therapy Nanoparticles
Scaffolds
Hydrogels
Antigen-TLR agonist fusion vaccines
Porous 3D scaffolds
Anti-PD-1 mAbs
03Gene therapy Viral gene delivery
Non-viral gene delivery
Polysaccharides
Polyethylemine (PEI)
Lipid
Naked DNA
04Natural antioxidantsNano delivery systems Solid nanocrystals
Nanoemulsion
Nanoliposomes

Current methods in oncology focus on the development of safe and efficient cancer nanomedicines. Targeted medical care helped rising the biodistribution of recent or already tested chemotherapeutical agents around the specific tissue to be treated; different methods, such as sequence medical care, siRNAs delivery, therapy, and inhibitor molecules, supply new potentialities to cancer patients. Gene therapy acts by direct in situ insertion of exogenous genes into benign tumors. Noticeably, stem cells can be used as regenerative medicine, therapeutic carriers, drug targeting, and generation of immune cells because of having unique biological actions on other cells. 22 On the opposite hand, thermal ablation and magnetic hyperthermia are promising alternatives to the growth surgical process. Finally, radionics and pathomics approaches facilitate the management of huge knowledge sets from cancer patients to enhance prognosis and outcomes. Much progress has been made, but many others are likely to come soon, producing more and more ad hoc personalized therapies. Further development and refinement of drug delivery systems are essential for improving therapeutic outcomes.

Acknowledgments

The authors thank the Center for Innovative Drug Development and Therapeutic Trials for Africa (CDT-Africa), Addis Ababa University, for the support rendered.

Author contributions: D.T.D. is a major contributor in writing the manuscript. All others reviewed and approved the manuscript.

Declaration of conflicting interests: The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Funding: The author(s) received no financial support for the research, authorship, and/or publication of this article.

An external file that holds a picture, illustration, etc.
Object name is 10.1177_20503121211034366-img1.jpg

  • Open access
  • Published: 29 July 2024

Pericytes recruited by CCL28 promote vascular normalization after anti-angiogenesis therapy through RA/RXRA/ANGPT1 pathway in lung adenocarcinoma

  • Ying Chen 1 , 2 , 3 ,
  • Zhiyong Zhang 1 , 2 , 3 ,
  • Fan Pan 3 , 6 ,
  • Pengfei Li 1 , 2 , 3 ,
  • Weiping Yao 3 , 6 ,
  • Yuxi Chen 1 , 2 , 3 ,
  • Lei Xiong 5 ,
  • Tingting Wang 1 , 2 , 3 ,
  • Yan Li 4 &
  • Guichun Huang 6 , 7  

Journal of Experimental & Clinical Cancer Research volume  43 , Article number:  210 ( 2024 ) Cite this article

157 Accesses

Metrics details

It has been proposed that anti-angiogenesis therapy could induce tumor "vascular normalization" and further enhance the efficacy of chemotherapy, radiotherapy, target therapy, and immunotherapy for nearly twenty years. However, the detailed molecular mechanism of this phenomenon is still obscure.

Overexpression and knockout of CCL28 in human lung adenocarcinoma cell line A549 and murine lung adenocarcinoma cell line LLC, respectively, were utilized to establish mouse models. Single-cell sequencing was performed to analyze the proportion of different cell clusters and metabolic changes in the tumor microenvironment (TME). Immunofluorescence and multiplex immunohistochemistry were conducted in murine tumor tissues and clinical biopsy samples to assess the percentage of pericytes coverage. Primary pericytes were isolated from lung adenocarcinoma tumor tissues using magnetic-activated cell sorting (MACS). These pericytes were then treated with recombinant human CCL28 protein, followed by transwell migration assays and RNA sequencing analysis. Changes in the secretome and metabolome were examined, and verification of retinoic acid metabolism alterations in pericytes was conducted using quantitative real-time PCR, western blotting, and LC–MS technology. Chromatin immunoprecipitation followed by quantitative PCR (ChIP-qPCR) was employed to validate the transcriptional regulatory ability and affinity of RXRα to specific sites at the ANGPT1 promoter.

Our study showed that after undergoing anti-angiogenesis treatment, the tumor exhibited a state of ischemia and hypoxia, leading to an upregulation in the expression of CCL28 in hypoxic lung adenocarcinoma cells by the hypoxia-sensitive transcription factor CEBPB. Increased CCL28 could promote tumor vascular normalization through recruiting and metabolic reprogramming pericytes in the tumor microenvironment. Mechanistically, CCL28 modified the retinoic acid (RA) metabolism and increased ANGPT1 expression via RXRα in pericytes, thereby enhancing the stability of endothelial cells.

We reported the details of the molecular mechanisms of "vascular normalization" after anti-angiogenesis therapy for the first time. Our work might provide a prospective molecular marker for guiding the clinical arrangement of combination therapy between anti-angiogenesis treatment and other therapies.

Introduction

Lung cancer stands as the foremost cause of mortality among patients with malignant tumors, with lung adenocarcinoma being a predominant pathological subtype [ 1 , 2 ]. Angiogenesis is a crucial hallmark of lung adenocarcinoma, and significant strides have been made in applying anti-angiogenesis therapy for its treatment [ 3 ]. Extensive research data indicates that anti-angiogenic tumor therapy exhibits the potential to augment the effectiveness of diverse therapeutic modalities, including chemotherapy, radiotherapy, tyrosine kinase inhibitors (TKIs), and immunotherapy. Nevertheless, a subset of patients fails to derive benefits from the combined approach of anti-angiogenic therapy and other anti-tumor treatments. Unraveling the synergistic mechanisms underlying anti-angiogenic therapy holds substantial clinical implications for optimizing combination therapies.

Since the 2000s, there has been a fundamental shift in the understanding of anti-angiogenic therapy for tumors, transitioning from the initial concept of solely inhibiting angiogenesis to inducing vascular normalization in tumors [ 4 ]. Presently, there is a consensus that optimizing the effectiveness of additional anti-tumor treatments through anti-angiogenic therapy is imperative. This notion is central to anti-angiogenic drugs' capability to induce transient vascular normalization in tumor blood vessels [ 5 ]. Preliminary investigations undertaken by our research group propose that anti-angiogenic drugs, employing diverse mechanisms of action, can prompt vascular normalization in lung adenocarcinoma [ 6 , 7 , 8 , 9 ]. The balance between pro-angiogenesis and anti-angiogenesis factors was postulated as the key molecular base of vascular normalization [ 10 ]. However, uncertainties persist regarding the timing and the limited duration of vascular normalization post-treatment. Addressing this issue is critical, as the temporal aspects of vascular normalization and its short-lived nature need clarification [ 7 , 8 , 9 , 11 , 12 ]. The pressing clinical challenge involves establishing protocols to monitor and regulate vascular normalization in lung adenocarcinoma, offering valuable guidance for developing combined anti-angiogenic therapies with other anti-tumor treatment strategies. A comprehensive understanding of the molecular mechanisms underpinning anti-angiogenic therapy-induced vascular normalization in lung adenocarcinoma forms the foundational basis for resolving this intricate clinical dilemma.

Pericytes, one of the structural cells found in capillary vessels, envelop the basal membrane of endothelial cells and interact with vascular endothelial cells to ensure the stability and functionality of microvessels [ 13 , 14 ]. Commonly used markers for identifying pericytes include platelet-derived growth factor receptor β (PDGFRβ), chondroitin sulfate proteoglycan 4 (CSPG4, also known as NG2), and alpha-smooth muscle actin (α-SMA). Several reports suggest that platelet-derived growth factor B (PDGFB) can promote vascular normalization in colon cancer and malignant melanoma by recruiting pericytes [ 15 ]. However, PDGFB does not perform this function in pericytes in lung adenocarcinoma [ 16 ]. The mechanism of pericyte recruitment in lung adenocarcinoma following anti-angiogenesis therapy remains unclear.

Human CC motif chemokine ligands 28 (CCL28) can recruit various cell types, including lymphatic endothelial cells [ 13 ], T cells, and plasma cells, through its receptors CCR3 or CCR10. Apart from its chemotactic properties, CCL28 has been implicated in promoting tumor development in ovarian cancer [ 17 ], gastric cancer [ 18 ], liver cancer, and lung adenocarcinoma [ 19 ]. Under hypoxic conditions, tumor cells mobilize various cell types using different chemokines to induce angiogenesis in response to nutritional and oxygen requirements. Notably, hypoxia induces the expression of CCL28 in ovarian cancer, which, in turn, enhances angiogenesis by recruiting regulatory T cells (Tregs) [ 17 ]. In contrast, CCL28 has been identified as a negative regulator of tumor growth and bone invasion in oral squamous cell carcinoma [ 20 ]. These diverse findings suggest that CCL28 may play distinct roles in various tumors. While one study discovered that CCL28 increased the proliferation, migration, and secretion of IL-6 and HGF in oral fibroblasts [ 21 ], the functional role of CCL28 in pericytes in lung adenocarcinoma has not been elucidated. Furthermore, previous studies have reported that CCL28 was involved in angiogenesis in various diseases, including tumors, skin wound healing [ 22 ], and rheumatoid arthritis [ 23 ]. However, little is known about the mechanism of CCL28 in tumor vascular normalization.

Our previous report highlighted that CCL28 can moderately enhance the angiogenesis in lung adenocarcinoma [ 19 ]. Interestingly, we found a normalized vasculature in CCL28 highly expressed tumors. The current study delves into the previously undisclosed connection between pericytes and CCL28 in lung adenocarcinoma. Both in vivo and in vitro experiments collectively illustrate that CCL28 derived from tumor cells is pivotal in advancing vascular normalization by mobilizing and reprogramming pericytes.

Materials and methods

Cancer cell lines.

The human lung adenocarcinoma cell line (A549, SPC-A1, and H1975) and mouse Lewis lung cancer cell line (LLC) were purchased from the Shanghai Institute of Biochemistry and Cell Biology (SIBCB) and maintained in our lab. Lung cancer cells were cultured in RPMI-1640, or DMEM medium (Gibco, LifeTech, USA) supplemented with 10% fetal bovine serum (FBS), 1% antibiotics (100U/ml penicillin and 100 μg/ml streptomycin) at 37℃ in a humidified 5% CO 2 atmosphere.

Isolation and identification of pericytes

Pericytes were isolated and cultured as we previously described [ 24 ]. Fresh lung cancer samples were collected from patients in Jinling Hospital (Nanjing, China). Tissue samples were cut into small blocks of approximately 1 ~ 2 mm in diameter, digested with trypsin and 0.5% collagenase, and filtered through the cell strainer to obtain single-cell suspension. PDGFRβ + cells were isolated by PE-PDGFRβ antibody and anti-PE magnetic beads (Miltenyi Biotec,130–123-772). The separated cells were cultured in F12K medium (Gibco, LifeTech, USA) containing 10% fetal bovine serum (Gibco, Life Tech, USA) with 100 U/mL penicillin and 100 μg/mL streptomycin. After two or three passages, pericytes were identified by morphology and immunofluorescence staining for α-Smooth Muscle Actin (α-SMA, CST, #19,245), platelet-derived growth factor receptor β (PDGFRβ, Abcam, ab69506) and chondroitin sulfate proteoglycan 4 (NG2, Abcam, ab129051). Patients with incomplete data were excluded to evaluate the clinical effect. Written informed consent was obtained from all subjects before collecting the samples. All the methods followed the institutional guidelines and were approved by the Ethical Review Committee of Jinling Hospital, Nanjing, China(2022DZGZR-QH-005).

Hypoxic culture model and RNA sequence assay

As previously reported, the hypoxic cell culture model was established with hypoxic chambers in our lab [ 19 ]. Briefly, lung adenocarcinoma cells were cultured under two different oxygen concentrations, 1% and 20%, respectively. The model was testified by the expression changes of HIF-1α and its regulated genes, such as GLUT1 and VEGFA (Fig. 1 C). Pericytes were cultured under different stimulation. After culturing for 24 h, the cells were collected, and the total amount of RNA or protein was extracted for western blot or quantitative PCR. RNA sequence assay was applied to detect the gene expression differences of pericytes cultured with or without the stimulation of CCL28 (MCE, HY-P7250) under hypoxia.

figure 1

CCL28 expression is upregulated after anti-angiogenesis therapy by hypoxia-sensitive transcription factor CEBPB in lung adenocarcinoma

RNA extraction and quantitative PCR

Total RNA was extracted from different cell lines using Trizol reagent (Invitrogen, USA). Subsequently, reverse transcription and quantitative PCR were performed using SYBR Green with an ABI StepOne Plus System (Applied Biosystems, Life Tech, USA). Relative gene expression was calculated by the ΔΔCt method based on glyceraldehyde-3-phosphate dehydrogenase (GAPDH) or β-actin levels. All reactions were run in triplicate.

Luciferase reporter assay

Wild-type (pGL3 wt, 5’-TGATTATGCAATGG-3') and mutant (pGL3 mut, 5’-ACTAATACGTTACC-3') promoters of the CCL28 gene were constructed into pGL3 firefly luciferase reporter plasmid vector purchased from Nanjing Realgene Bio-Technology Company (Nanjing China). The pRL-TK vector expressing renilla luciferase was used as an internal control. The reporter plasmid was co-transfected with pRL-TK vector and CEBPB expressing pcDNA3.1 vector or control vector into A549 cell. Dual-Luciferase Reporter Assay System (Promega, USA) was used to detect the luciferase activity after 48-h incubation.

Chromatin Immunoprecipitation (ChIP)

Briefly, pericytes treated with or without recombinant human CCL28 or CCR3 (R&D Systems, MAB155-100) neutralizing antibodies were collected and fixed by adding a cross-linking agent, formaldehyde, to stabilize the interactions between chromatin proteins and DNA. Subsequently, chromatin immunoprecipitation was conducted to enrich DNA fragments bound by chromatin proteins. After cross-link reversal and DNA purification, real-time quantitative PCR (qPCR) technology is employed to measure enriched DNA, determining the relative abundance at specific gene loci.

Liquid Chromatography-Mass Spectrometry (LC–MS)

Liquid chromatography-mass spectrometry was applied to measure retinoic acid (RA, HY-14649) in pericytes. Briefly, after treatment with or without CCL28, pericytes were subjected to enzymatic digestion, washed three times with PBS, flash-frozen in liquid nitrogen for one minute, and then stored at -80 °C. When performing the detection, begin by thawing the sample at 4 °C. Next, add 0.5 mL of methanol solution, followed by 10 min of sonication and 30 min of shaking for extraction. Subsequently, place the centrifuge tube in a low-temperature centrifuge and centrifuge for 10 min at 4 °C and 12,000 rpm. Collect 800μL of the supernatant, evaporate it, and then add 100μL of methanol solution. Finally, retain 80 μL of the supernatant for subsequent liquid chromatography-mass spectrometry analysis. A standard curve is generated to calculate the sample's content using a retinoic acid standard (yuanye Bio-Technology).

Single-cell RNA sequencing

Lewis lung cancer cells overexpressing CCL28 were inoculated in C57BL/6 mice subcutaneously (1 × 10 6 cells per mouse). When the tumor grew to about 100 mm 3 , fresh tumor samples were collected and dissociated into single cells. The tissue was washed with PBS, then chopped and incubated in an enzyme digestion solution for 40 min at 37 °C. After digestion, the solution was filtered through a 40 μm filter and centrifuged. Red blood cells were removed using a red blood cell lysis solution, followed by cell counting. The cell suspension was filtered using FACS tubes, centrifuged again, and washed twice. Finally, after microscopic examination and cell counting, the single-cell suspension was used for scRNA sequencing.

The scRNA sequencing analysis was carried out according to a single cell analysis workflow with BD Rhapsody™ Systems (BD Biosciences, USA), and RNA sequencing and data analysis were completed on the platform of NovaSeq 6000 system (Illumina, USA). The entired scRNA sequencing workflow was provided in the Supplementary document 1.

Animal model

Six-week-old female BALB/c nude mice and C57BL/6 female mice were used for tumor model establishment. LLC (1 × 10 6 per mouse) with or without overexpressing CCL28 were subcutaneously injected in C57BL/6 mice ( n  = 6, each group). Mice inoculated with CCL28 knock-out LLC cells and wild-type LLC were randomly divided into two groups, respectively, and treated with or without RA by gavage daily (n = 6, each group). Mice were observed, and the tumors' length and width were measured daily. When tumors grew to a specific size, subcutaneous tumors were collected and photographed. The A549 (CCL28 overexpression or CCL28 knock-out) tumor model followed the same procedure. All animal experiments were carried out following the institutional guidelines and approved by the Ethical Review Committee of Comparative Medicine, Jinling Hospital, Nanjing, China (2022DZGKDWLS-0091).

Western blot and ELISA

Proteins were extracted from the cultured cells by lysis buffer, separated by SDS-PAGE, and transferred to PVDF membranes (Millipore, USA). The filters were blocked in Tris-buffered saline containing 0.2% Tween plus 5% non-fat milk and incubated with primary antibodies overnight at 4 °C. Secondary antibodies were used for visualization through chemiluminescence (ECL, Amersham Pharmacia Biotech, UK). Primary antibodies against RXRα (CST,1:1000), RARα (CST,1:1000), RDH13 (Proteintech,1:1000), DHRS11 (Proteintech,1:1000), CCR10 (Abmart,1:1000), β-actin (Servicebio, 1:1000) and neutralizing antibody against CCR3 (R&D Systems, USA) were applied in the present study. Anti-rabbit IgG, HRP-linked Antibody (CST, 1:1000) and anti-mouse IgG, HRP-linked Antibody (CST, 1:1000) were utilized as the secondary antibodies. CCL28 in serum was detected by an ELISA kit (Abcam, USA) according to the manufacturer's instructions.

Immunofluorescence and multiplex Immunohistochemistry (mIHC)

Tumor samples were collected and fixed in 10% formalin before processing and paraffin embedding. Immunofluorescence was conducted on 5 µm sections. For culture cell staining, lung adenocarcinoma-associated pericytes were washed with 1 × PBS and fixed with acetone for 10 min on ice. Tumor microvascular endothelial cells and cancer pericytes were stained by CD31 antibody (Abcam, ab281583) and rat anti-human/mouse monoclonal NG2 antibody (Abcam, ab129051). At the same time, tumor cells were stained by mouse anti-human pan-cytokeratin (Pan-ck) monoclonal antibody (Abcam, ab215838). Goat anti-mouse IgG antibody (labeled with Alex Fluor 488), goat anti-rabbit IgG antibody (labeled with Alex Fluor 555 or Alex Fluor 488), and goat anti-rat IgG antibody (labeled with Alex Fluor 555 or Alex Fluor 488) (Abcam) were applied as secondary antibody in immunofluorescence staining analysis.

For multiplex immunohistochemistry, the slides were incubated with the primary antibodies (anti-CD31, anti-NG2, anti-Pan-CK antibody, and anti-CCL28 (Proteintech, 18,214–1-AP), respectively) and horseradish peroxidase-conjugated secondary antibody, and tyramine signal amplification (TSA) was performed following the pre-optimized antibody concentration and the order of staining. Antibody stripping and antigen retrieval were performed after each round of TSA. DAPI (Sigma-Aldrich, USA) was used for nuclei staining. A whole slide scan of the multiplex tissue sections produced multispectral fluorescent images visualized in SlideViewer software. A specialized pathologist chose representative regions of interest (ROI), and multiple fields of view were acquired at 20 × power for further analysis. The mean fluorescence intensity of CCL28 in each ROI was calculated by the ImageJ program ( https://imagej.nih.gov/ij/ ).

CRISPR-cas9 knock out

Briefly, three designed sgRNAs were synthesized and co-transfected with Cas9 nuclease/sgRNA in A549 lung adenocarcinoma cells. After 72 h, the cells were seeded in a 96-well plate for single-cell cloning by limiting dilution analysis (LDA). After 7 to 15 days, about ten single-cell clones were selected for PCR and ELISA. Mutant sites were further confirmed by DNA sequencing.

Knockdown of RDH13 and DHRS11

For RDH13 and DHRS11 gene silence, pericytes were infected with negative control siRNAs, RDH13-targeting, or DHRS11-targeting siRNAs (genepharma). Briefly, pericytes were seeded in 24-well plates, and the cells were transfected with siRNAs using Lipofectamine RNAiMAX (Invitrogen) according to the manufacturer’s instructions.

Bioinformatic analysis

Transcription factor binding sites prediction in promotor of CCL28 gene was conducted in PROMO 3.0, virtual laboratory TFSEARCH (ver.1.3) and JASPAR ( https://alggen.lsi.upc.edu/rerecer/menu_recerca.html , https://diyhpl.us/~bryan/irc/protocol-online/protocol-cache/TFSEARCH.html , https://jaspar.elixir.no/ ). RNA sequencing dates of lung adenocarcinoma tumor samples or lung adenocarcinoma cell lines were extracted on the cBioPortal platform ( https://www.cbioportal.org/ ). The relationship between expression levels of CEBPB and CCL28, RXRα, and ANNGPT1 were calculated online, as well as the predictive value of CEBPB for the survival of lung adenocarcinoma patients.

Statistical analyses

Data were presented as mean ± SEM. Student's unpaired two-tailed tests were used for comparisons between two groups. Pearson or Spearman correlation was applied to analyze the relationship between the expression scores of two genes. Statistical analyses were performed on GraphPad Prism 8.0. Multiple comparisons were analyzed by one-way ANOVA using the LSD test. Statistical significance was confirmed when p  < 0.05.

CCL28 expression is up-regulated after anti-angiogenesis therapy by hypoxia-sensitive transcription factor CEBPB in lung adenocarcinoma

The GEO database ( https://www.ncbi.nlm.nih.gov/gds/ ) was retrieved to elucidate the impact of anti-angiogenesis therapy on the gene expression of cancer patients. Molecular expression profile data (GSE61676) of lung adenocarcinoma patients undergoing anti-angiogenesis treatment was reanalyzed. Twenty-three stage IV lung adenocarcinoma patients, treated with a combination of anti-VEGF monoclonal antibody (Bevacizumab) and TKI, were selected for analysis. The changes in mRNA profile (Affymetrix Human Exon 1.0 ST Arrays) were examined before and 24 h after Bevacizumab treatment. The findings revealed a significant increase in the hypoxia index (GLUT1, one of the primary target genes regulated by hypoxia-inducing factor HIF-1) and up-regulation of CCL28 expression in lung adenocarcinoma patients 24 h after VEGF monoclonal antibody treatment compared to baseline levels (Fig. 1 A, 1B). After treatment with bevacizumab, CCL28 expression showed a log2 fold change of 1.258 compared to baseline, with a corresponding p-value of 0.0152, as depicted in Fig. 1 A. After anti-angiogenesis therapy, tumor cells often experience a state of ischemia and hypoxia. To simulate this hypoxic state of tumor cells, nine cell lines of 4 different tumor types were culture in hypoxia and normoxia conditions, including two hepatoma cell lines (HEPG2 and SMMC), three lung adenocarcinoma cell lines (A549, SPC-A1, and H1975), two breast carcinoma cell lines (MCF7 and MDA-MB-231), and two colorectal cancer cell lines (HCT116 and SW480). The expression of GLUT1 and VEGFA, which was driven by the hypoxia-induced factor (HIF), was detected by qRT-PCR. VEGFA and GLUT1 were up-regulated in all nine cell lines in the hypoxic chamber, indicating that the hypoxic microenvironment was established (Fig. 1 C). We previously reported that high expression of CCL28 was induced in hypoxic lung adenocarcinoma cell lines, A549 and SPC-A1. To further confirm our study, we examined the expression of CCL28 under hypoxia conditions in the nine cell lines. CCL28 was up-regulated in all three lung adenocarcinoma cell lines, whereas others were not (Fig. 1 D). Then, we utilized PROMO 3.0 to predict the transcription factors binding sites to the CCL28 promoter region and intersected them with previously reported hypoxia-related transcription factors, revealing six transcription factors involved (Fig. 1 E). CEBPB and Sp1 can potentially regulate CCL28, while only CEBPB was highly expressed in hypoxic tumor cells (as listed in Supplementary Table 1). We then used a public integrative database (cBioPortal) to analyze the correlation of the expression of CEBPB with that of CCL28 in 44 lung adenocarcinoma cell lines (as listed in Supplementary Table 2). As expected, CEBPB expression strongly correlated with CCL28 (Fig. 1 F). The CEBPB binding motif was shown in Fig. 1 G. Thus, we speculated that hypoxia-induced high expression of CCL28 was mediated by CEBPB. To verify our prediction, we generated luciferase reporter plasmids harboring either wild-type (WT) or mutated (MUT) CEBPB binding sequencing within the promotor element of the CCL28 gene. Luciferase reporter analysis indicated that CEBPB could directly regulate the expression of the CCL28 gene in A549 cells (Fig. 1 H). In addition, disease-free survival and overall survival of lung adenocarcinoma patients with high CEBPB expression were significantly decreased ( p  = 0.0065 and p  = 0.032, respectively) (Fig. 1 I).

Tumor-derived CCL28 promotes vascular normalization and pericyte recruitment in the tumor microenvironment

To investigate the effects of CCL28 on tumor growth and vascular normalization, we established CCL28 overexpression and knock-out lung adenocarcinoma cells by lentivirus vectors and Cas9 nuclease/sgRNA (Supplementary Fig. 1), respectively. As we previously reported, in vivo studies indicated that CCL28 could promote tumor growth in A549 human lung adenocarcinoma (Fig. 2 A). Microvessels and pericytes were subsequently assessed by immunofluorescence staining using antibodies against CD31 (red) and NG2 (green) in tumor tissue in different groups. The percentage of pericyte coverage and microvessel density (MVD) in A549-CCL28 tumor tissue significantly increased compared to A549-NC, and this effect was reversed after CCL28 was knocked out (Fig. 2 B). These results suggested that CCL28 could promote vascular normalization and mobilize pericytes.

figure 2

Tumor-derived CCL28 recruits pericytes to promote vascular normalization in the tumor microenvironment

Further, we isolated primary pericytes from lung cancer tissues to explore how CCL28 recruits pericytes in vitro. The characteristics of pericytes were identified by immunofluorescence staining α-SMA, PDGFRβ, and NG2 (Fig. 2 C). Cell viability remained unaffected by different concentrations of recombinant CCL28 in pericytes and A549 (Supplementary Fig. 2A,2B), but enhanced cell migration of A549 (Supplementary Fig. 2C). The transwell migration assay was performed to confirm the effect of CCL28 on the migration of pericytes by using different concentrations of human recombinant CCL28 protein (Fig. 2 D, left panel). The number of cells migrated to the lower chamber was calculated (Fig. 2 D, right panel), and we found that the optimal concentration of CCL28 for enhancing the migration efficiency of pericytes is 250 ng/ml. Thus, we selected a concentration of 250 ng/ml to conduct subsequent migration experiments. To further verify the recruitment effect of CCL28 on pericytes, cells were inoculated around matrigel mixed with or without CCL28 recombinant protein. The graphs visually illustrate cell migration distance and quantity (Fig. 2 E, left panel). CCL28 significantly increased the migration distance of pericytes compared to the control group (Fig. 2 E, right panel). Subsequently, we assessed the expression of two CCL28 receptors, CCR3 and CCR10, in pericytes with or without CCL28. Western blot experiments revealed that the expression of CCR3 was significantly higher than that of CCR10 in pericytes, suggesting that CCR3 likely plays a predominant role in mediating CCL28 signal transduction within pericytes (Fig. 2 F). Blocking CCR3 with neutralizing antibodies diminished the chemotactic effect of CCL28 on pericytes (Fig. 2 G). Further, we validated the relationship between CCL28 and pericytes coverage in biopsy tissue using multiplex immunofluorescence. CCL28 was mainly expressed in lung adenocarcinoma cells due to the fluorescence overlapping between CCL28 and PAN-CK (Fig. 2 H). We found a significant positive correlation between CCL28 expression levels and the percent of pericytes coverage (Fig. 2 I). The above findings indicate that CCL28 could recruit pericytes through the receptor CCR3, thereby promoting vascular normalization. Because the action of CCL28 leads to vascular normalization, making the vascular network healthier and more regular, it may also increase the density and functionality of tumor blood vessels, providing more nutrients and oxygen to promote tumor growth.

Tumor-derived CCL28 promotes expression of angiopoietin-1 via CCR3 in pericytes

Pericytes interact with vascular endothelial cells to promote the maturation of neovasculature [ 25 ]. However, the molecular mechanisms underlying pericytes-induced vascular normalization in lung adenocarcinoma after anti-angiogenesis therapy remain unclear. To clarify the exact role and potential mechanism of CCL28 in vascular normalization, we mimicked the hypoxic microenvironment in lung adenocarcinoma, cultured vascular pericytes, and conducted secretome analysis after CCL28 treatment. We detected augmented production of angiopoietin-1 in pericytes stimulated by CCL28 (Fig. 3 A). Compared to the control group, the expression of ANGPT1 changed by 5.415-fold under CCL28 stimulation, with a p-value of 0.033. RNA-seq results were confirmed by quantitative PCR in different concentrations of recombinant CCL28 protein under hypoxia conditions (Fig. 3 A right panel). Angiopoietin-1, a protein typically associated with angiogenesis and vascular normalization, maintains vascular stability and integrity by interacting with endothelial cells. Consistent with prior studies [ 24 , 26 ], western blot experiments showed that recombinant human angiopoietin-1 could up-regulate endothelial nitric oxide synthase (eNOS) expression in a dose-dependent manner in human umbilical vein endothelial cell (HUVEC) (Fig. 3 B). Likewise, recombinant human angiopoietin-1promoted cell survival by activating the PI3K-AKT signaling pathway in a Tie2-dependent manner in endothelial cells (Fig. 3 C). To further explore the regulatory mechanism of ANGPT1 expression, we identified a significant increase in the transcription factor RXRa and RARα after CCL28 stimulation in transcriptome data, verified by qPCR (Fig. 3 D and E ). However, protein level detected by western blot, CCL28 up-regulated RXRα and CCR3 but not RARα (Fig. 3 F). Also, the effect of CCL28 on RXRα up-regulation was reversed by adding a CCR3 neutralizing antibody (Fig. 3 G). Next, we confirmed the relation between transcription factor RXRα and expression of ANGPT1. Correlation analysis showed that RXRα expression positively correlates with ANGPT1 expression (Fig. 3 H). The binding motif of the RARα and RXRα complex (Fig. 3 I) and the transcription factor binding sites in the promotor area of the CCL28 gene (Fig. 3 J) were shown. CHIP-qPCR was performed to investigate the binding of transcription factor RXRα to the promoter region of the target gene ANGPT1 . Compared to the NC group, CCL28 enhanced the degree of enrichment, and this effect could be dampened after the CCR3 neutralization, suggesting a direct interaction between RXRα and the ANGPT1 promoter. The electrophoresis image displayed the specificity of the amplified DNA fragments (Fig. 3 K).

figure 3

Tumor-derived CCL28 promotes the expression of angiopoietin-1 via CCR3 in pericytes

Furthermore, the mRNA level of ANGPT1 was increased by the addition of CCL28 but attenuated by the blockade of CCR3 (Fig. 3 L). RXRα is a subtype of nuclear receptor closely associated with vitamin A metabolism and retinoic acid signaling. These results led us to speculate whether retinoic acid metabolism was also involved in regulating vascular normalization by CCL28. Thus, we measured the retinoic acid content in pericytes using LC–MS technology. Interestingly, the retinoic acid production was notably increased (almost 2.5-fold change) compared to the control group in pericytes stimulated by CCL28 (Fig. 3 M). Additionally, the exogenous addition of retinoic acid in pericytes could promote the expression of ANGPT1 in a dose-dependent manner (Fig. 3 N). These results indicate that CCL28 can activate the nuclear transcription factor RXRα through the CCR3 receptor, promoting the expression of ANGPT1 in pericytes. In interaction with endothelial cells, pericytes-deriore, the mRNA level of ANGPT1 was increased by the addition of CCL28 but attenuated by the blockade ofon.

CCL28 activates retinoic acid signaling in pericytes through CCR3

We further analyzed the sequencing data to understand better the mechanism responsible for retinoic acid accumulation in pericytes after CCL28 stimulation. Enrichment analysis of gene pathways showed that RA signaling pathways were activated after CCL28 stimulation (Fig. 4 A). Two key enzymes involved in retinoic acid metabolism have changed, including RDH13 and DHRS11 (Fig.4B). RDH13 plays a role in converting retinol to retinaldehyde, a critical rate-limiting enzyme step in the retinoid metabolic pathway. However, DHRS11 reduces retinoic acid metabolites to related forms of retinol. In this way, the balance between RDH13 and DHRS11 can affect the level of retinoic acid in cells. The metabolic balance model diagram of retinoic acid shows the transformation process of the three substances (Fig. 4 C). In this metabolic process, RDH13 was up-regulated, but DHRS11 was decreased in pericytes after treatment of CCL28 (Fig. 4 B). These results were confirmed by qPCR (Fig. 4 D) and western blot (Fig. 4 E). The correlation analysis exhibited a positive relationship between RDH13 and CCL28 in lung adenocarcinoma (Fig. 4 F). Expression changes of RDH13 and DHRS11 in pericytes stimulated by CCL28 could be reversed by blocking CCR3 (Fig. 4 G). However, treating pericytes directly with retinoic acid did not result in changes in key molecules involved in RA metabolism (Supplementary Fig. 3 A). To further investigate whether the expression of two key enzymes in the retinoic acid synthesis process is correlated with the expression of ANGPT1, we knocked down the expression of RDH13 and DHRS11 using siRNA and then examined the expression of ANGPT1. After knockdown of RDH13 (Fig. 4 H), RXRα and ANGPT1 were significantly reduced (Fig. 4 I, J). However, DHRS11 did not exhibit this effect (Supplementary Fig. 3B, 3C). To further elucidate the relationship between CCL28 and its downstream molecules, immunofluorescence staining of key molecules—RDH13, DHRS11, and ANGPT1—was performed on tumor biopsy tissues (Fig. 4 K, left panel), followed by correlation analysis. In these samples, we observed a positive correlation between CCL28 expression and the levels of ANGPT1 and RDH13, while a negative correlation was found with DHRS11 (Fig. 4 K, right panel).All these results illuminated that CCL28 promoted the retinoic acid synthesis process by disturbing the balance between RDHs and DHRS, but the changes are temporary and dynamic.

figure 4

Retinoic acid signaling is activated by CCL28 in pericytes through CCR3 A , Volcano plot of changes in metabolic pathways after CCL28 stimulation. B Volcano plot of the enrichment of gene expression after CCL28 stimulation. C Diagram of the metabolic conversion process in the retinoic acid metabolic signaling pathway. D and E Expression of RDH13 and DHRS11 detected by qPCR and western blot with or without exogenous supplement of CCL28. F Correlation of expression of CCL28 with RDH13 in lung adenocarcinoma. G The protein level of DHRS11 and RDH13 stimulated with or without CCL28 and CCR3 neutralizing antibody in pericytes (left) and gray value was calculated(right). H Knockdown efficiency of RDH13 was confirmed by qPCR. I and J Relative expression of RXRα and ANGPT1 after knockdown of RDH13 with or without stimulation of CCL28. K Representative immunofluorescence images of PAN-CK, NG2, CCL28 with DHRS11 or RDH13 or Angiopoietin-1 on biopsy tissues from lung cancer patients (left panel). Scale bar = 100 μm. The correlation between the expression of CCL28 and the levels of DHRS11, RDH13, and angiopoietin-1 (right panel). Data with error bars are shown as mean ± SEM. Each symbol represents data from a replicate. Each panel is a representative experiment of at least three independent biological replicates. *, **, *** represent p  < 0.05, p  < 0.01 and p  < 0.001, respectively. Abbreviation: MFI, Mean fluorescence intensity

Both CCL28 and retinoic acid could promote vascular normalization in vivo

Further, we investigated the role of CCL28 in tumor growth and vascular normalization in immunocompetent mice. We overexpressed CCL28 in LLC cells, which was validated by ELISA. Compared to the control group, CCL28 overexpression showed a significant increase in CCL28 concentration in both the cell supernatants and serum in mice implanted with tumors (Supplementary Fig. 4A, 4B). As expected, in vivo studies indicated that CCL28 could promote tumor growth in Lewis lung adenocarcinoma (LLC) (Fig. 5 A). Compared to the control group, NG2 + pericytes accumulation and angiogenesis enhanced in the LLC-CCL28 group (Fig. 5 B).

figure 5

To more clearly explore the function of CCL28 on stromal cells and immune cells in the tumor microenvironment, we performed single-cell sequencing analysis on tumor tissues of LLC-NC and LLC-CCL28. Single-cell sequencing analysis provided a panoramic study of the tumor microenvironment, and the cell population in the tumor microenvironment underwent significant alterations. In total, 6380 cells were obtained after quality filtering in the 2 conditions. Specifically, there were 3829 cells in the CCL28 overexpression group, and 2551 cells in the control group. The t-SNE diagrams showed that these cells were divided into 19 cell clusters with a total of 10 cell types (Fig. 5 C and Supplementary Fig. 4D, 4E). Of all the cell types, three types showed the most significant changes in cell proportions: fibroblasts, pericytes, and macrophages, with fold changes of 2.39, 1.95, and 1.45, respectively. The percentage of pericytes was 1.09% in LLC-NC and rose to the rate of 2.14% in LLC-CCL28 (Fig. 5 D). In the CCL28 overexpressing group, we identified 82 pericytes, whereas the control group had 28 pericytes. This significant difference suggested that CCL28 may play a pivotal role in influencing pericyte populations within the tumor microenvironment. Moreover, a previous study has reported that CCL28 could recruit Tregs and promote angiogenesis in ovarian carcinoma [ 17 ]. However, the percentage of T cells was reduced in the CCL28 up-regulated LLC, indicating that CCL28 might play different roles in different cancer types. Moreover, the number of DC cells remains unchanged, but their proportion decreased (Fig. 5 D).

Subsequently, we focused on the analysis of pericytes. Cluster 14, which was CSPG4 + and ACTA2 + , was recognized as pericytes. The marker genes t-SNE plot for Cluster 14 was presented in Supplementary Fig. 5. Interestingly, metabolic pathway enrichment analysis revealed that the retinol metabolism was activated in 13 clusters, including DC, macrophages, fibroblasts, and pericytes, suggesting that the synthesis and metabolism of retinoic acid play essential physiological roles in tumor microenvironment (Fig. 5 E). In the cellular communication network, pericytes interact closely with various cell types. The heat map indicated that pericytes exhibited the strongest interaction with themselves, followed by cluster 3 and cluster 6, representing tumor and endothelial cells. Cluster 3 was characterized as Ki-67-positive and Top2a-positive, and cluster 6 was positive for Pecam (Fig. 5 F). Our results in Fig. 3 M also suggested CCL28 can promote retinoic acid accumulation in pericytes in vitro. We further investigated the effects of retinoic acid by daily gavage on tumor growth and vascular normalization in vivo. Knocking out CCL28 (LLC-CCL28-KO) or supplementing retinoic acid (LLC-NC + RA) could suppress tumor growth in LLC models (Fig. 5 G and Supplementary Fig. 4C). However, the suppression effect is more pronounced after knocking out CCL28 than the control group (LLC-NC), and CCL28 knockout with supplementing RA could further synergistically inhibit tumor growth (Fig. 5 G). Multiplex immunofluorescence was used to detect the normalization of tumor blood vessels. We found that compared to the control group, knocking out CCL28 resulted in a significant reduction in the proportion of NG2 + cells and decreased coverage of pericytes. As expected, supplementation of retinoic acid was able to promote a certain degree of restoration of pericytes coverage both in wild-type and CCL28 knockout tumors (Fig. 5 H). Then, we quantified the vascular density in different groups and found that, compared to the control group, CCL28 knockout resulted in significantly decreased vascular density, while retinoic acid had only a slight effect (Fig. 5 H). Additionally, we evaluated the effect of retinoic acid on tumor cell viability using a CCK8 assay (Supplementary Fig. 4F). The results indicated that retinoic acid did not significantly affect tumor cell viability at pharmacological concentrations (1–2 μg/ml, equivalent to 3.33–6.66 μM). However, cytotoxic effects were observed at higher concentrations, which exceed pharmacologically relevant levels in the body necessary to support essential cellular functions. To confirm the influence on tumor hypoxia, we immunofluorescently labeled mouse tumor tissues using CA9 (Carbonic Anhydrase 9), which is considered an indicator of hypoxia. Our findings indicated that compared to the control group, tumor hypoxia significantly increased in tumors with CCL28 knockout. However, retinoic acid does not appear to have a significant effect on hypoxia (Supplementary Fig. 4G). These results indicated that RA inhibited tumor growth and enhanced vascular normalization, rather than angiogenesis.

The above data indicate that CCL28-regulated retinoic acid plays a crucial role in vascular normalization, suggesting its potential as a therapeutic agent in anti-angiogenic tumor treatment. To investigate whether the combination of retinoic acid and bevacizumab has a synergistic effect, we established an A549 mouse model and administered intravenous injections of bevacizumab, coupled with oral gavage of retinoic acid. We observed that retinoic acid and bevacizumab could attenuate tumor growth in mice, respectively. Moreover, simultaneous administration of retinoic acid and bevacizumab had a synergistic effect on tumor growth (Fig. 5 I). These data indicate that retinoic acid can be used in combination with bevacizumab for tumor treatment, providing a new direction for clinical combination therapy.

CCL28 is involved in bevacizumab-mediated vascular normalization

Moreover, we investigated the synergistic effects of CCL28 knock-out and VEGF blocking on the tumor growth and vascular normalization of lung adenocarcinoma. Subcutaneously implanted lung adenocarcinoma cells with CCL28 knock-out grew much slower than wild-type tumor cells, while combination of VEGF blocker could stop the growth of the tumors (Fig. 6 A). In addition, significant promotion of vascular normalization is observed after bevacizumab treatment. However, this effect disappeared after knocking out CCL28, regardless of whether bevacizumab was added (Fig. 6 B). Knocking out CCL28 could inhibit tumor angiogenesis, and the effect was more pronounced when bevacizumab was combined (Fig. 6 B). These results indicated that CCL28 could participate in bevacizumab-mediated vascular normalization, and CCL28 might be a potential target for anti-angiogenesis therapy in lung adenocarcinoma.

figure 6

Further, we evaluated the expression levels of CCL28 within the tumor tissue (Fig. 6 C, upper panel). The mean fluorescence intensity of CCL28 was significantly upregulated when bevacizumab was used compared to the control group (Fig. 6 C, lower panel), consistent with the results we found in the clinical samples. Also, the expression level of CCL28 in lung adenocarcinoma correlates with therapeutic efficacy. In lung adenocarcinoma patients, based on their treatment outcomes with bevacizumab, they were categorized into two groups: poor responders and good responders. PAN-CK, NG2, CD31, and CCL28 were stained in tumor tissues to analyze their expression levels with the therapeutic efficacy of bevacizumab treatment. Patients who responded well to bevacizumab showed significantly increased CCL28 expression, and high expression of CCL28 in tumor cells,, was associated with enhanced vascular maturity and favorable treatment outcomes (Fig. 6 D, 6E). The conclusion was evidenced by a notable decrease in malignant pericardial and pleural effusion and a significant reduction in metastatic lymph nodes after bevacizumab-based treatment (Fig. 6 F). Then, we analyzed the expression levels of DHRS11 and RDH13 in two groups of patients. Consistently, the therapeutic efficacy of bevacizumab appeared to be positively correlated with RDH13 expression and negatively correlated with DHRS11 expression (Fig. 6 G, H). These results suggest that the favorable response to anti-angiogenic therapy in patients may be attributed to the activation of CCL28, thereby deepening our understanding of treatment response variations.

Anti-angiogenesis therapy stands as a pivotal approach for metastatic lung adenocarcinoma [ 3 , 27 ], targeting the well-established VEGF/VEGFR pathway with numerous drugs developed over the last four decades [ 28 , 29 ]. However, its clinical impact has proven more complicated than initially anticipated. Recognizing the vasculature normalization effects of anti-angiogenic drugs has led to their integration into combination regimens with chemotherapy, radiotherapy, immunotherapy, and targeted therapy [ 4 ]. For example, bevacizumab, through the modulation of angiogenesis and improvement of the tumor microenvironment, lowers tumor vascular density and edema, thereby enhancing the efficiency of oxygen supply and making other treatment modalities more effective. Consequently, anti-angiogenesis drugs are employed as modulators for tumor vasculature and the microenvironment within combination regimens.

Anti-angiogenic therapy inhibits abnormal blood vessel formation, fostering vascular normalization through mechanisms like reducing vessel density, remodeling the extracellular matrix, regulating inflammation, balancing growth factors, and modulating the immune system. The tipped balance between pro-angiogenesis and anti-angiogenesis factors contributes to maintaining tumor vascular growth [ 30 ]. Anti-angiogenic therapy disrupts this tipped balance by inhibiting VEGF, enhancing the relative action of angiopoietin-1(ANGPT1), which aids in regulating and inducing vascular normalization. This rebalancing contributes to vascular normalization, ensuring newly formed blood vessels exhibit a more organized structure and function more normally. The action of angiopoietin-1 helps consolidate and stabilize the newly formed vessels, making them a more effective transportation system.

However, it is also widely accepted that agents targeting VEGF/VEGFR may ultimately elevate hypoxia levels within tumors [ 4 ]. This hypoxic microenvironment triggers alternative pro-angiogenesis molecular pathways in tumor or stromal cells within the tumor microenvironment, including FGF-2, HGF, DLL4/Notch, CCL28, and others [ 19 , 21 , 22 , 31 ]. Furthermore, hypoxia prompts a cascade of biological responses, leading tumor cells to adapt their metabolic pathways to low-oxygen conditions. The dysregulated metabolism observed in rapidly proliferating tumor cells is a hallmark of malignancy [ 32 ], contributing to the activation of various metabolism-related genes, including several hypoxia-related transcription factors like hypoxia-inducible factor (HIF), nuclear factor kappa-B, CREB, AP-1, p53, Sp1/3, Egr-1, and CEBPB [ 33 ].

In this study, we identified CCL28 as another crucial molecular target for vascular normalization in lung adenocarcinoma. It was observed that CCL28 could be up-regulated under hypoxic conditions. However, the molecular mechanism underlying hypoxia-induced transcriptional activation of the CCL28 gene remains unclear. Our investigation revealed that the transcription factors CEBEPB could regulate the expression of CCL28. CCL28, belonging to the subfamily of small cytokine CC chemokines, binds to chemokine receptors CCR3 and CCR10 [ 19 ]. It has been reported that CCL28 exhibits chemotactic activity for various immune cells and plays a role in the physiology of extracutaneous epithelial tissues [ 34 ]. Several studies have delved into the functions of CCL28 in the tumor microenvironment [ 17 , 18 , 20 ]. Therefore, we hypothesized that CCL28 might play a pivotal role in modulating the tumor microenvironment in lung adenocarcinoma. Here, we noted an augmentation in pericyte accumulation within the tumor microenvironment associated with the overexpression of CCL28. However, the mechanisms of pericyte recruitment in lung adenocarcinoma remain unknown. Thus, we examined the expression of CCR3/CCR10 on various tumor stromal cells, revealing widespread expression of CCR3 on endothelial cells, cancer-associated fibroblasts, and pericytes in lung adenocarcinoma. Two chemotactic experiments substantiated the effects of CCL28 on recruiting pericytes through the receptor CCR3.

Abnormal differentiation of stromal cells is an essential characteristic of malignant tumors and is correlated with abnormal metabolism. Many tumor stromal cells were differentiated from anti-tumor type to pro-tumor type, such as tumor-associated macrophage, cancer-associated fibroblasts, tumor-associated neutrophils, etc. [ 35 , 36 , 37 ]. Reprogramming the metabolism and modulating tumor stromal cell differentiation is a promising cancer treatment strategy. Interestingly, the present study found that retinoic acid metabolism was activated in a series of stromal cells in lung cancer.

Importantly, chemotactic factors play a crucial role in cell migration, influencing not only the cell's movement [ 38 ] but also closely interacting with cellular metabolism. They regulate energy production, metabolic pathways, and antioxidant responses, ensuring that cells have sufficient energy and resources during migration. Chemotactic factors influence cellular metabolism by regulating intracellular signaling pathways, such as PI3K/AKT, MAPK, AMPK, and mTOR [ 39 , 40 ]. The activation or inhibition of these signaling pathways can modulate the activity of intracellular metabolic enzymes, affecting processes such as glucose metabolism, lipid synthesis, and amino acid utilization. In the present study, we found chemokine CCL28 could influence the cellular levels of retinoic acid by modulating its synthesis.

Retinoic acid (RA), also known as vitamin A acid, and its related analogs participate in regulating the gene networks involved in cell growth, differentiation, homeostasis, and apoptosis. RA is an active metabolite of retinol. Retinol (Vitamin A) is a fat-soluble essential micronutrient that plays a crucial role in embryonic development, organ formation, immune system function, and vision [ 26 ]. Retinol can be metabolized into retinal by the retinol dehydrogenases (RDHs), and retinal can also be metabolized into retinol by the dehydrogenase/reductase SDR family (DHRS) [ 41 ]. Retinal can be further irreversibly metabolized to RA by the retinaldehyde dehydrogenases (ALDHs). The synthesis of RA depends on the balance between RDHs and DHRS. The present study found that CCL28 could disturb the balance between RDH13 and DHRS11. After being treated with CCL28, RDH13 in pericytes was significantly upregulated in a dose-dependent manner. However, the effect curve of CCL28 on DHRS11 expression in pericytes is not linear and might be bell-shaped, like the effects of VEGFA on endothelial cells. In addition, there might be an alternative pathway to maintain the balance of DHRS11 and RDH13 in pericytes. RA binds to RARα, promoting the formation of a heterodimer with retinoid X receptor alpha (RARα/RXRα) in the cell nucleus. This heterodimer binds to retinoic acid response elements (RAREs) in the promoter region of target genes [ 42 ], including ANGPT1 .

The angiopoietin family, including ANGPT1, ANGPT2, ANGPT3, and ANGPT4, is crucial in vascular development and normalization. They regulate endothelial cells' survival, proliferation, and migration by interacting with Tie receptors, thus vital to vascular normalization. ANGPT1 can activate the TIE2 receptor on endothelial cells, maintaining endothelial cell stability, enhancing tight connections between endothelial cells, and reducing microvascular permeability through a series of signaling pathways [ 26 ]. These signaling pathways include tyrosine kinase-related protein DOKR (also known as DOK2), endothelial nitric oxide synthase (eNOS), SH2 domain-containing phosphatase (SHP2), growth factor receptor-binding protein 2 (GRB2), and PI3K-Akt [ 24 ]. Various regulatory mechanisms influence the expression of the ANGPT1 . These include the hypoxia-inducible factor-1α (HIF-1α) signaling pathway under low oxygen conditions, interaction with vascular endothelial growth factor (VEGF), involvement of anti-inflammatory factors and growth factors, cell–cell interactions, as well as the nuclear factor-κB (NF-κB) signaling pathway and hormonal regulation. In the present study, we identified another regulatory mechanism controlling the expression of ANGPT1.

Vitamin A and its metabolites, particularly retinoic acid, exert regulatory effects on the vascular system through various pathways [ 43 ]. These include maintaining endothelial cell function, regulating angiogenesis, inhibiting vascular smooth muscle cell proliferation, suppressing inflammatory responses, and promoting cell differentiation. This comprehensive regulatory process contributes to maintaining blood vessels' stable structure and function, playing a crucial role in embryonic development, tissue repair, angiogenesis, and vascular health during inflammation and diseases. Interestingly, vitamin A and its metabolites have been used as a differentiation modulator of malignant cells for cancer treatment. The present study proposed a new strategy to modulate the differentiation of cancer stromal by vitamin A and its metabolites.

Conclusions

In conclusion, we elucidated that a specific chemokine CCL28 induced after anti-angiogenesis therapy can alter tumor stromal cell metabolism and reshape the tumor microenvironment. In summary, we identified a mechanism through which CCL28 promotes vascular normalization via a CCR3-pericytes-RA-RXRα-ANGPT1-dependent pathway (Fig. 7 ). With an in-depth exploration of the interplay between chemokine and ANGPT1, our work may provide valuable insights into the regulatory network of vascular normalization, offering new avenues for modulating tumor microenvironment and overcoming resistance to anti-angiogenesis therapy in lung adenocarcinoma treatment.

figure 7

A schematic diagram of tumor microenvironment modulation effects of CCL28

Availability of data and materials

The data that support the findings of this study are available from the corresponding author upon reasonable request.Competing interests.

The authors declare that they have no competing interests.

Abbreviations

Platelet-derived growth factor receptor β

Chondroitin sulfate proteoglycan 4

Human CC motif chemokine ligands 28

Human umbilical vein endothelial cell

Dehydrogenase/reductase 11

Retinol dehydrogenase 13

Bevacizumab

Zhang Y, Vaccarella S, Morgan E, Li M, Etxeberria J, Chokunonga E, et al. Global variations in lung cancer incidence by histological subtype in 2020: a population-based study. Lancet Oncol. 2023;24:1206–18.

Article   PubMed   Google Scholar  

Xia C, Dong X, Li H, Cao M, Sun D, He S, et al. Cancer statistics in China and United States, 2022: profiles, trends, and determinants. Chin Med J (Engl). 2022;135:584–90.

Reck M, Popat S, Grohé C, Corral J, Novello S, Gottfried M, et al. Anti-angiogenic agents for NSCLC following first-line immunotherapy: rationale, recent updates, and future perspectives. Lung Cancer. 2023;179:107173.

Article   PubMed   CAS   Google Scholar  

Augustin HG, Koh GY. Antiangiogenesis: vessel regression, vessel normalization, or both? Cancer Res. 2022;82:15–7.

Zhang N, Yin R, Zhou P, Liu X, Fan P, Qian L, et al. DLL1 orchestrates CD8(+) T cells to induce long-term vascular normalization and tumor regression. Proc Natl Acad Sci U S A. 2021;118(22):e2020057118.

Huang G, Chen L. Recombinant human endostatin improves anti-tumor efficacy of paclitaxel by normalizing tumor vasculature in Lewis lung carcinoma. J Cancer Res Clin Oncol. 2010;136:1201–11.

Shi S, Chen L, Huang G. Antiangiogenic therapy improves the antitumor effect of adoptive cell immunotherapy by normalizing tumor vasculature. Med Oncol. 2013;30:698.

Shi S, Tao L, Song H, Chen L, Huang G. Synergistic antitumor effect of combining metronomic chemotherapy with adoptive cell immunotherapy in nude mice. APMIS. 2014;122:380–91.

Tao L, Huang G, Shi S, Chen L. Bevacizumab improves the antitumor efficacy of adoptive cytokine-induced killer cells therapy in non-small cell lung cancer models. Med Oncol. 2014;31:777.

Jain RK. Normalization of tumor vasculature: an emerging concept in antiangiogenic therapy. Science. 2005;307:58–62.

Shi S, Wang R, Chen Y, Song H, Chen L, Huang G. Combining antiangiogenic therapy with adoptive cell immunotherapy exerts better antitumor effects in non-small cell lung cancer models. PLoS One. 2013;8:e65757.

Article   PubMed   PubMed Central   CAS   Google Scholar  

Ramjiawan RR, Griffioen AW, Duda DG. Anti-angiogenesis for cancer revisited: is there a role for combinations with immunotherapy? Angiogenesis. 2017;20:185–204.

Article   PubMed   PubMed Central   Google Scholar  

Hartmann DA, Berthiaume AA, Grant RI, Harrill SA, Koski T, Tieu T, et al. Brain capillary pericytes exert a substantial but slow influence on blood flow. Nat Neurosci. 2021;24:633–45.

Figueiredo AM, Villacampa P, Dieguez-Hurtado R, Jose Lozano J, Kobialka P, Cortazar AR, et al. Phosphoinositide 3-kinase-regulated pericyte maturation governs vascular remodeling. Circulation. 2020;142:688–704.

Zhang Y, Cedervall J, Hamidi A, Herre M, Viitaniemi K, D’Amico G, et al. Platelet-specific PDGFB ablation impairs tumor vessel integrity and promotes metastasis. Cancer Res. 2020;80:3345–58.

Hosaka K, Yang Y, Seki T, Nakamura M, Andersson P, Rouhi P, et al. Tumour PDGF-BB expression levels determine dual effects of anti-PDGF drugs on vascular remodelling and metastasis. Nat Commun. 2013;4:2129.

Facciabene A, Peng X, Hagemann IS, Balint K, Barchetti A, Wang LP, et al. Tumour hypoxia promotes tolerance and angiogenesis via CCL28 and T(reg) cells. Nature. 2011;475:226–30.

Ji L, Qian W, Gui L, Ji Z, Yin P, Lin GN, et al. Blockade of beta-catenin-induced CCL28 suppresses gastric cancer progression via inhibition of treg cell infiltration. Cancer Res. 2020;80:2004–16.

Huang G, Tao L, Shen S, Chen L. Hypoxia induced CCL28 promotes angiogenesis in lung adenocarcinoma by targeting CCR3 on endothelial cells. Sci Rep. 2016;6:27152.

Park J, Zhang X, Lee SK, Song N-Y, Son SH, Kim KR, et al. CCL28-induced RARβ expression inhibits oral squamous cell carcinoma bone invasion. J Clin Investig. 2019;129:5381–99.

Incio J, Ligibel JA, McManus DT, Suboj P, Jung K, Kawaguchi K, et al. Obesity promotes resistance to anti-VEGF therapy in breast cancer by up-regulating IL-6 and potentially FGF-2. Sci Transl Med. 2018;10:10.

Article   Google Scholar  

Cascone T, Xu L, Lin HY, Liu W, Tran HT, Liu Y, et al. The HGF/c-MET pathway is a driver and biomarker of VEGFR-inhibitor resistance and vascular remodeling in non-small cell lung cancer. Clin Cancer Res. 2017;23:5489–501.

Chen Z, Kim SJ, Essani AB, Volin MV, Vila OM, Swedler W, et al. Characterising the expression and function of CCL28 and its corresponding receptor, CCR10, in RA pathogenesis. Ann Rheum Dis. 2015;74:1898–906.

Huang H, Bhat A, Woodnutt G, Lappe R. Targeting the ANGPT-TIE2 pathway in malignancy. Nat Rev Cancer. 2010;10:575–85.

Teichert M, Milde L, Holm A, Stanicek L, Gengenbacher N, Savant S, et al. Pericyte-expressed Tie2 controls angiogenesis and vessel maturation. Nat Commun. 2017;8:16106.

Khan KA, Wu FT, Cruz-Munoz W, Kerbel RS. Ang2 inhibitors and Tie2 activators: potential therapeutics in perioperative treatment of early stage cancer. EMBO Mol Med. 2021;13:e08253.

Wang M, Herbst RS, Boshoff C. Toward personalized treatment approaches for non-small-cell lung cancer. Nat Med. 2021;27:1345–56.

Ellis LM, Hicklin DJ. VEGF-targeted therapy: mechanisms of anti-tumour activity. Nat Rev Cancer. 2008;8:579–91.

Geindreau M, Ghiringhelli F, Bruchard M. Vascular endothelial growth factor, a key modulator of the anti-tumor immune response. Int J Mol Sci. 2021;22:22.

Yang T, Xiao H, Liu X, Wang Z, Zhang Q, Wei N, et al. Vascular normalization: a new window opened for cancer therapies. Front Oncol. 2021;11:719836.

Li J-L, Sainson RCA, Oon CE, Turley H, Leek R, Sheldon H, et al. DLL4-Notch signaling mediates tumor resistance to anti-VEGF therapy in vivo. Can Res. 2011;71:6073–83.

Article   CAS   Google Scholar  

Hanahan D. Hallmarks of cancer: new dimensions. Cancer Discov. 2022;12:31–46.

Cummins EP, Taylor CT. Hypoxia-responsive transcription factors. Pflugers Arch. 2005;450:363–71.

Mohan T, Deng L, Wang B-Z. CCL28 chemokine: an anchoring point bridging innate and adaptive immunity. Int Immunopharmacol. 2017;51:165–70.

Zhang X, Ji L, Li MO. Control of tumor-associated macrophage responses by nutrient acquisition and metabolism. Immunity. 2023;56:14–31.

Zhang H, Yue X, Chen Z, Liu C, Wu W, Zhang N, et al. Define cancer-associated fibroblasts (CAFs) in the tumor microenvironment: new opportunities in cancer immunotherapy and advances in clinical trials. Mol Cancer. 2023;22:159.

Que H, Fu Q, Lan T, Tian X, Wei X. Tumor-associated neutrophils and neutrophil-targeted cancer therapies. Biochim Biophys Acta Rev Cancer. 2022;1877:188762.

Stoll G, Pol J, Soumelis V, Zitvogel L, Kroemer G. Impact of chemotactic factors and receptors on the cancer immune infiltrate: a bioinformatics study revealing homogeneity and heterogeneity among patient cohorts. Oncoimmunology. 2018;7:e1484980.

Xu H, Lin S, Zhou Z, Li D, Zhang X, Yu M, et al. New genetic and epigenetic insights into the chemokine system: the latest discoveries aiding progression toward precision medicine. Cell Mol Immunol. 2023;20:739–76.

Lazennec G, Rajarathnam K, Richmond A. CXCR2 chemokine receptor - a master regulator in cancer and physiology. Trends Mol Med. 2024;30:37–55.

Adams MK, Belyaeva OV, Wu L, Kedishvili NY. The retinaldehyde reductase activity of DHRS3 is reciprocally activated by retinol dehydrogenase 10 to control retinoid homeostasis. J Biol Chem. 2014;289:14868–80.

Ghyselinck NB, Duester G. Retinoic acid signaling pathways. Development. 2019;146:146.

DiKun KM, Gudas LJ. Vitamin A and retinoid signaling in the kidneys. Pharmacol Ther. 2023;248:108481.

Download references

Acknowledgements

Not applicable.

This work was supported by the National Natural Science Foundation of China (Grant No. 82273326), Jiangsu Provincial Youth Medical Key Talents Project (Grant No. QNRC2016887), the Bethune Charitable Foundation (Grant No. KY202301-17) and Postgraduate Research & Practice Innovation Program of Jiangsu Province (Grant No. KYCX22_0176).

Author information

Authors and affiliations.

The State Key Laboratory of Pharmaceutical Biotechnology, Chemistry and Biomedicine Innovation Center (ChemBIC), Division of Immunology, Medical School, Nanjing University, Nanjing, 210093, China

Ying Chen, Zhiyong Zhang, Pengfei Li, Yuxi Chen & Tingting Wang

Jiangsu Key Laboratory of Molecular Medicine, Division of Immunology, Medical School, Nanjing University, Nanjing, 210093, China

Medical Schoolof, Nanjing University, Nanjing, Jiangsu, 210093, China

Ying Chen, Zhiyong Zhang, Fan Pan, Pengfei Li, Weiping Yao, Yuxi Chen & Tingting Wang

Department of Respiratory Critical Care Medicine, Affiliated Hospital of Medical School, Nanjing Drum Tower Hospital, Nanjing University, Nanjing, Jiangsu, 210008, China

Department of Cardio-Thoracic Surgery, Affiliated Hospital of Medical School, Jinling Hospital, Nanjing University, Nanjing, Jiangsu, 210008, China

Department of Medical Oncology, Affiliated Hospital of Medical School, Jinling Hospital, Nanjing University, Nanjing, Jiangsu, 210008, China

Fan Pan, Weiping Yao & Guichun Huang

Department of Oncology, Medical School, Zhongda Hospital, Southeast University, Nanjing, 210009, China

Guichun Huang

You can also search for this author in PubMed   Google Scholar

Contributions

TTW, YL and GCH: Provided funding support, conceptualization, methodological guidance, and manuscript revision for this study.. YC: conceived of the study, performed the experiments and wrote the paper. WPY: participated in animal experiments. ZYZ and FP: performed the statistical analysis. PFL, YXC and LX: Participated in experiment design and sample collection. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Tingting Wang , Yan Li or Guichun Huang .

Ethics declarations

Ethics approval and consent to participate.

All animal experiments were carried out following the institutional guidelines and approved by the Ethical Review Committee of Comparative Medicine, Jinling Hospital, Nanjing, China. Written informed consent was obtained from all patients before collecting the samples. All the methods were carried out following the institutional guidelines and approved by the Ethical Review Committee of Jinling Hospital, Nanjing, China.

Consent for publication

All authors have consented for publication.

Competing interests

Additional information, publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Supplementary material 1., supplementary material 2., supplementary material 3., supplementary material 4., rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Chen, Y., Zhang, Z., Pan, F. et al. Pericytes recruited by CCL28 promote vascular normalization after anti-angiogenesis therapy through RA/RXRA/ANGPT1 pathway in lung adenocarcinoma. J Exp Clin Cancer Res 43 , 210 (2024). https://doi.org/10.1186/s13046-024-03135-3

Download citation

Received : 21 March 2024

Accepted : 22 July 2024

Published : 29 July 2024

DOI : https://doi.org/10.1186/s13046-024-03135-3

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Anti-angiogenesis therapy
  • Vascular normalization
  • Lung adenocarcinoma
  • Retinoic acid

Journal of Experimental & Clinical Cancer Research

ISSN: 1756-9966

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

cancer cell research studies

Advertisement

Iron Chelation Therapy Elicits Innate Immune Control of Metastatic Ovarian Cancer

ORCID logo

  • Article contents
  • Figures & tables
  • Supplementary Data
  • Peer Review
  • Get Permissions
  • Cite Icon Cite
  • Search Site
  • Accepted Manuscript July 29 2024

Tito A. Sandoval , Camilla Salvagno , Chang-Suk Chae , Deepika Awasthi , Paolo Giovanelli , Matias Marin Falco , Sung-Min Hwang , Eli Teran-Cabanillas , Lasse Suominen , Takahiro Yamazaki , Hui-Hsuan Kuo , Jenna E. Moyer , M. Laura Martin , Jyothi Manohar , Kihwan Kim , Maria A. Sierra , Yusibeska Ramos , Chen Tan , Alexander Emmanuelli , Minkyung Song , Diana K. Morales , Dmitriy Zamarin , Melissa K. Frey , Evelyn Cantillo , Eloise Chapman-Davis , Kevin Holcomb , Christopher E. Mason , Lorenzo Galluzzi , Zhen Ni Zhou , Anna Vaharautio , Suzanne M. Cloonan , Juan R. Cubillos-Ruiz; Iron Chelation Therapy Elicits Innate Immune Control of Metastatic Ovarian Cancer. Cancer Discov 2024; https://doi.org/10.1158/2159-8290.CD-23-1451

Download citation file:

  • Ris (Zotero)
  • Reference Manager

Iron accumulation in tumors contributes to disease progression and chemoresistance. While targeting this process can influence various hallmarks of cancer, the immunomodulatory effects of iron chelation in the tumor microenvironment are unknown. Here, we report that treatment with deferiprone, an FDA-approved iron chelator, unleashes innate immune responses that restrain ovarian cancer. Deferiprone reprogrammed ovarian cancer cells towards an immunostimulatory state characterized by production of type I interferon (IFN) and overexpression of molecules that activate natural killer (NK) cells. Mechanistically, these effects were driven by innate sensing of mitochondrial DNA in the cytosol and concomitant activation of nuclear DNA damage responses triggered upon iron chelation. Deferiprone synergized with chemotherapy and prolonged the survival of mice with ovarian cancer by bolstering type I IFN responses that drove NK cell-dependent control of metastatic disease. Hence, iron chelation may represent an alternative immunotherapeutic strategy for malignancies that are refractory to current T cell-centric modalities.

Article PDF first page preview

Client account, citing articles via, email alerts.

  • Online First
  • Online ISSN 2159-8290
  • Print ISSN 2159-8274

AACR Journals

  • Blood Cancer Discovery
  • Cancer Discovery
  • Cancer Epidemiology, Biomarkers & Prevention
  • Cancer Immunology Research
  • Cancer Prevention Research
  • Cancer Research
  • Cancer Research Communications
  • Clinical Cancer Research
  • Molecular Cancer Research
  • Molecular Cancer Therapeutics
  • Information on Advertising & Reprints
  • Information for Institutions/Librarians

cancer cell research studies

  • Privacy Policy
  • Copyright © 2023 by the American Association for Cancer Research.

This Feature Is Available To Subscribers Only

Sign In or Create an Account

Cancer Research

A light pink dendritic cell communicating with a blue T cell.

Driving discovery. Powering progress.

Nci research areas.

  • Cancer Biology
  • Causes of Cancer
  • See All Research Areas

Cancer Research & Infrastructure at NCI

  • NCI's Clinical Trials Programs
  • NCI-Designated Cancer Centers
  • Frederick National Laboratory for Cancer Research
  • Intramural Research
  • Extramural Research
  • Cancer Research Workforce

Research Resources & NCI Initiatives

  • Resources for Researchers
  • The Cancer Moonshot℠
  • Childhood Cancer Data Initiative
  • Equity and Inclusion Program
  • See All Key Initiatives

Research Advances

cancer cell research studies

The 2022 report includes data on new cancer cases and deaths plus a detailed examination of trends in pancreatic cancer and its subtypes.

National Cancer Plan

NCI joins the cancer community in advancing the goals of the National Cancer Plan as part of its research programs.

Abstract illustration depicting artificial intelligence in medical sciences.

Advances in technology and access to large volumes of data have led to promising applications of AI in cancer research and care.

People standing against a wall.

Participate in Cancer Research

Explore the many ways you can participate in cancer research, including treatment and non-treatment studies.

cancer cell research studies

Inside Cancer Careers: The Future of AI and Cancer Research

NIH, National Cancer Institute, Division of Cancer Treatment and Diagnosis (DCTD)

Please enable JavaScript or use a Browser that supports JavaScript

  • Advanced Search
  • Browse by Analyte
  • Browse by Pre-analytical Factor
  • Search SOPs
  • SOP Compendiums
  • Suggest a New Paper
  • Submit an SOP
  • Curator Home
  • Back Office

The impact of impaired tissue fixation in resected non-small-cell lung cancer on protein deterioration and DNA degradation.

Author(s): Butter R, Halfwerk H, Radonic T, Lissenberg-Witte B, Thunnissen E

Publication: Lung Cancer , 2023, Vol. 178 , Page 108-115

PubMed ID: 36812759 PubMed Review Paper? No

Purpose of Paper

The purpose of this paper was to assess the effects of inadequate fixation on tissue quality (determined macroscopically and microscopically), immunohistochemical staining of 10 antigens, and DNA degradation using formalin-fixed, paraffin-embedded (FFPE) non-small cell lung cancer (NSCLC) tissue.

Conclusion of Paper

Based on either macroscopic or microscopic review a similar number of fixed tissue specimens were classified as adequately fixed (8 versus 6 specimens), inadequately fixed (2 specimens for each, and having both adequately and inadequately fixed regions (15 versus 17 specimens). 

Mean IHC scores of eight antigens (programmed death-ligand 1, PD-L1; cytokeratin CAM5.2, CK7; mesenchymal-epithelial transition factor, c-MET; cytokeratin KER-MNF116; Napsin A; p40; thyroid transcription factor 1, TTF) were compared between adequately and inadequately fixed areas of the 25 tumors collected for the study; because specimens were excluded from analysis if immunopositive staining was absent in both adequately and inadequately fixed tissue, a different number of specimens were analyzed for each antigen. Mean H-scores for both KER-MNF116 (256 versus 151, respectively; p<0.001) and p40 (293 versus 248, respectively; p=0.028) were significantly higher in regions that were adequately fixed than those that were inadequately fixed. When mean H-scores were only examined in the subset of tumors that displayed both adequately and inadequately fixed regions, KER-MNF116 alone differed significantly, with a higher mean H-score in the adequately fixed region compared to the inadequately fixed region (249 versus 131; p=0.003). Necrotic zones within a tumor had significantly lower mean H-scores than adequately fixed regions for CK7 (80 versus 186; p=0.044), KER-MNF116 (63 versus 256; p<0.001), and p40 (0 versus 293; p=0.002). 

Evidence of DNA fragmentation (the absence of 300 and 400 bp amplicons) was prominent in DNA isolated from both adequately and inadequately fixed regions, and the concentration of PCR amplicons did not differ between the two region types for any of the amplicon lengths investigated.  Interactions between DNA degradation and cold ischemia time and fixation time were observed, and tumor specimens that had a cold ischemia time >16 h had a lower concentration of 300 (0 versus 0.13 ng/µl; p=0.004) and 400 bp (0 versus 0.0059 ng/µl; p=0.046) PCR amplicons than those with a delay of <6 h, and tumor specimens that were fixed in formalin for >24 h had a lower concentration of 300 (0.02 versus 0.14 ng/µl; p=0.024) and 400 bp (0 versus 0.0067 ng/µl; p=0.045) PCR amplicons than those that were fixed for <24 h. A potential influence of cold ischemia time and fixation time on immunohistochemical staining was not investigated.  

Study Purpose

The purpose of this study was to assess the effects of under- and over-fixation of formalin-fixed, paraffin-embedded (FFPE) non-small cell lung cancer (NSCLC) tissue on tissue quality (determined macroscopically and microscopically), immunohistochemical staining of 10 antigens, and DNA degradation.  Surgically resected lung tissue from 25 patients diagnosed with NSLC(9 with non-mucinous adenocarcinoma, 9 with squamous cell carcinoma, 2 with pleomorphic adenoma, 2 with carcinoid tumors, 1 with mucinous adenocarcinoma, and two with large cell carcinoma) were partially sliced along the axial plane every 1.5 cm, bandage mesh was placed between slices, and the tissue specimen was fixed in 10% neutral buffered formalin for 12-24 h. Cold ischemia time (time between resection and fixation) ranged between 0.5 to 21.5 h (mean 3.4 h).  Fixed specimens were sectioned again (0.5 cm thick sections) and examined macroscopically for adequate (specimen was a white/gray color throughout) or inadequate (specimen contained pink/red areas) fixation.  Specimens then underwent additional tissue processing and paraffin embedding (no additional details provided).  FFPE tissue sections were stained with hematoxylin and eosin for microscopic evaluation and by immunohistochemistry (IHC) for the following antigens: ALK, PD-L1, GAM2.5, CK7, c-Met, KER-MNF116, NapsinA, p40, ROS1, TTF1. For tissue slides stained by immunohistochemistry, areas containing the maximum and minimum staining intensity, necrotic zones, and evidence of adequate and inadequate fixation (detachment of the basement membrane) fixation were estimated.  To evaluate DNA degradation, adequately fixed and inadequately fixed regions were macroscopically dissected from 10 µm thick FFPE sections and DNA was isolated with the QIAamp DNA FFPE Tissue Kit, then assessed by multiplex PCR (100, 200, 300, 400 bp amplicons; gene not specified). The concentration of PCR amplicons was determined by electrophoresis with an Agilent Bioanalyzer. 

Summary of Findings:

Based on macroscopic evaluation of the fixed tissue segment before paraffin-embedding, 8 of the 25 specimens were adequately fixed (tissue was white or grey), 2 of the specimens were inadequately fixed (tissue was pink or red), and 15 specimens contained regions of both. Microscopic review of H&E-stained slides generated similar results, as 6 specimens were adequately fixed, 2 were inadequately fixed (detachment of the basement membrane), and 17 specimens contained regions of both.  Mean IHC scores of eight antigens were compared between adequately and inadequately fixed areas of the 25 tumors collected for the study; because specimens were excluded from analysis if immunopositive staining absent in both adequately and inadequately fixed tissue, a different number of specimens were analyzed for each antigen. Mean H-scores for both KER-MNF116 (256 versus 151, respectively; p<0.001) and p40 (293 versus 248, respectively; p=0.028) were significantly higher in regions that were adequately fixed and those that were inadequately fixed. Necrotic zones within a tumor had significantly lower mean H-scores than adequately fixed regions for CK7 (80 versus 186; p=0.044), KER-MNF116 (63 versus 256; p<0.001), and p40 (0 versus 293; p=0.002).  When mean H-scores were only examined in the subset of tumors that displayed both adequately and inadequately fixed regions, KER-MNF116 alone differed significantly, with a higher mean H-score in the adequately fixed region compared to the inadequately fixed region (249 versus 131; p=0.003). The minimum and maximum H-scores within a tumor differed significantly for all eight antigens evaluated (PD-L1, CAM5.2, CK7, c-MET, KER-MNF116, Napsin A, p40, TTF1; p≤0.008 for all).  Evidence of DNA fragmentation (the absence of 300 and 400 bp amplicons) was prominent in both adequately and inadequately fixed regions, and the concentration of PCR amplicons did not differ between the two region types for any of the amplicon lengths investigated.  Interactions between DNA degradation and cold ischemia time and fixation time were observed. Tumor specimens that had a cold ischemia time >16 h had a lower concentration of 300 (0 versus 0.13 ng/µl; p=0.004) and 400 bp (0 versus 0.0059 ng/µl; p=0.046) PCR amplicons than those with a delay of <6 h. Tumor specimens that were fixed in formalin for >24 h had a lower concentration of 300 (0.02 versus 0.14 ng/µl; p=0.024) and 400 bp (0 versus 0.0067 ng/µl; p=0.045) PCR amplicons than those that were fixed for <24 h. A potential influence of cold ischemia time and fixation time on immunohistochemical staining was not investigated.  

Please enable JavaScript or use a Browser that supports JavaScript for more on this study.

Biospecimens

  • Tissue - Lung

Preservative Types

  • Neoplastic - Benign
  • Neoplastic - Carcinoma
AnalyteTechnology Platform
Morphology Macroscopic observation
DNA PCR
Protein Immunohistochemistry
Morphology H-and-E microscopy

Pre-analytical Factors:

ClassificationPre-analytical FactorValue(s)
<24 h
>24 h
<6 h
>16 h
ALK
PD-L1
GAM2.5
CK7
c-Met
KER-MNF116
NapsinA
p40
ROS1
TTF1
100 bp
200 bp
300 bp
400 bp

You Recently Viewed  

News and announcements.

  • New Articles on the GTEx Project are Now FREELY Available!
  • Just Published!
  • April 24, 2024: Biobanking for Precision Medicine Seminar
  • U.S. Department of Health and Human Services
  • National Institutes of Health
  • National Cancer Institute

IMPDH2 suppression impedes cell proliferation by instigating cell cycle arrest and stimulates apoptosis in pediatric hepatoblastoma

  • Open access
  • Published: 01 August 2024
  • Volume 150 , article number  377 , ( 2024 )

Cite this article

You have full access to this open access article

cancer cell research studies

  • Linman Li 1 , 3   na1 ,
  • Yichi Wu 1 , 3   na1 ,
  • Hong-ting Huang 1   na1 ,
  • June-kong Yong 1 ,
  • Zicheng Lv 1 , 2 ,
  • Yi Zhou 3   na1 ,
  • Xuelin Xiang 3 ,
  • Jie Zhao 1 ,
  • Zhifeng Xi 1 , 3 ,
  • Hao Feng 1 , 2 , 3 &
  • Qiang Xia 1 , 3  

Hepatoblastoma (HB) is the most common pediatric liver tumor, presenting significant therapeutic challenges due to its high rates of recurrence and metastasis. While Inosine Monophosphate Dehydrogenase 2(IMPDH2) has been associated with cancer progression, its specific role and clinical implications in HB have not been fully elucidated.

This study utilized Quantitative Real-Time Polymerase Chain Reaction (qRT-PCR) and Tissue Microarray (TMA) for validation. Following this, IMPDH2 was suppressed, and a series of in vitro assays were conducted. Flow cytometry was employed to assess apoptosis and cell cycle arrest. Additionally, the study explored the synergistic therapeutic effects of mycophenolate mofetil (MMF) and doxorubicin (DOX) on HB cell lines.

The study identified a marked overexpression of IMPDH2 in HB tissues, which was strongly correlated with reduced Overall Survival (OS) and Event-Free Survival (EFS). IMPDH2 upregulation was also found to be associated with key clinical-pathological features, including pre-chemotherapy alpha-fetoprotein (AFP) levels, presence of preoperative metastasis, and the pre-treatment extent of tumor (PRETEXT) staging system. Knockdown of IMPDH2 significantly inhibited HB cell proliferation and tumorigenicity, inducing cell cycle arrest at the G0/G1 phase. Notably, the combination of MMF, identified as a specific IMPDH2 inhibitor, with DOX, substantially enhanced the therapeutic response.

The overexpression of IMPDH2 was closely linked to adverse outcomes in HB patients and appeared to accelerate cell cycle progression. These findings suggest that IMPDH2 may serve as a valuable prognostic indicator and a potential therapeutic target for HB.

The present study unveiled a significant overexpression of inosine monophosphate dehydrogenase 2 (IMPDH2) in hepatoblastoma (HB) tissues, particularly in association with metastasis and recurrence of the disease. The pronounced upregulation of IMPDH2 was found to be intimately correlated with adverse outcomes in HB patients. This overexpression appears to accelerate the progression of the cell cycle, suggesting that IMPDH2 may serve as a promising candidate for both a prognostic marker and a therapeutic target in the context of HB.

Avoid common mistakes on your manuscript.

Introduction

Hepatoblastoma (HB) is the predominant pediatric liver cancer, typically affecting children aged below three years (Czauderna et al. 2014 ). Despite its rarity, the prevalence of HB has climbed annually in recent decades, constituting approximately 80% of pediatric liver malignancies (Bell et al. 2017 ; Linabery and Ross 2008 ). Clinical presentations often include abdominal distension and the detection of an abdominal mass through imaging. Elevated levels of AFP, a fetal hepatocyte-derived protein, are commonly observed and considered as a crucial tumor indicator for monitoring disease progression and response to comprehensive treatment. The rising incidence is speculated to be linked to improved survival rates among prematurely born or with extreme low weight, a population susceptible to HB (Heck et al. 2013 ; Fine Licht et al. 2012 ).

Primary therapeutic approaches for HB involve complete surgical resection and chemotherapy, with cisplatin-based regimens proving effective for unresectable tumors and contributing to increased survival rates (Davies et al. 2004 ). Advances in surgical techniques and postoperative chemotherapy have enhanced overall outcomes, yielding 5-year survival rates averaging 82% (Tulla et al. 2015 ). However, approximately 30% of patients, even with surgery and neoadjuvant chemotherapy, exhibit suboptimal outcomes (Aronson et al. 2019 ; Meyers et al. 2017 ). In cases of advanced HB, liver transplantation stands as the sole effective treatment option (Trobaugh-Lotrario et al. 2016 ).

Mechanistically, ever-increasing evidence implicates that HB originates from less differentiated cells, with disorders of developmental and self-regeneration pathways playing a pivotal role in hepato-carcinogenesis. The Wnt/beta-catenin pathway, in relation to stem cells, is critically important, influencing the activation and expansion of these cells during embryonic development and liver regeneration, thus fostering liver homeostasis (Monga 2015 ; Russell and Monga 2018 ). HB often exhibits resistance to chemotherapy at high levels (Marin et al. 2018 ). Unfortunately, the mechanisms underlying this resistance remain poorly understood.

IMPDH is recognized as a pivotal rate-controlling enzyme within the purine de novo synthesis pathway (Shu and Nair 2008 ). This enzyme comprises two isotypes: IMPDH1 and IMPDH2 (Natsumeda et al. 1990 ). with IMPDH1 exhibiting relatively stable expression in both normal and tumor cells. In contrast, IMPDH2 manifests significantly heightened expression levels across various malignant tumors (Collart et al. 1992 ). Accumulating evidence from previous investigations has underscored the substantial elevation of IMPDH2 expression, which correlates closely with tumor progression and a poor prognosis in a diverse array of malignancies (Zhou et al. 2014 ). Notably, overexpression of IMPDH2 has been identified in specific patient subgroups that demonstrate a diminished response to chemotherapy, thereby establishing it as an independent prognostic indicator for chemotherapeutic response and event-free survival (EFS) (Fellenberg et al. 2007 ). Recent research has unveiled markedly upregulated IMPDH2 expression in prostate, kidney, bladder, and liver cancers, suggesting its potential utility as a biomarker for these diseases(Yang et al. 2014 ; Zou et al. 2015 ; Jia et al. 2022 ). However, the precise role of IMPDH2 in the molecular mechanisms underlying hepatoblastoma (HB) remains insufficiently characterized within the current literature.

Cyclin-Dependent Kinase Inhibitor 1 A (CDKN1A), commonly referred to as p21, is a critical regulator of cell cycle progression, particularly during the G1 phase (Xiong et al. 1993 ). Functionally, p21 plays a multifaceted role in orchestrating cell cycle arrest, facilitating DNA repair, and regulating apoptosis (Mauro et al. 2012 ; Gartel and Tyner 2002 ). By inhibiting the activity of cyclin-dependent kinases (CDKs), p21 effectively prevents the transition of cells from the G1 to the S phase, thereby significantly contributing to the regulation of cell division and preventing aberrant cell proliferation. Furthermore, in response to DNA damage, p21 can induce cell cycle arrest, providing sufficient time for DNA repair to occur. Under certain conditions, p21 also participates in the modulation of programmed cell death, or apoptosis.

Mycophenolate mofetil (MMF), approved by the FDA in 1995 for the prevention of transplant graft failure (Lipsky 1996 ), is a widely utilized immunosuppressive agent in the post-organ transplantation setting. Acting as a prodrug, MMF is rapidly metabolized into mycophenolic acid (MPA), which targets IMPDH, thereby reducing the synthesis of guanine nucleotides (Ransom 1995 ; Suthanthiran and Strom 1997 ). In the therapeutic management of hepatoblastoma (HB), particularly in high-risk patients, a combination regimen of cisplatin and doxorubicin (DOX) is frequently employed (Trobaugh-Lotrario and Katzenstein 2012 ). However, both of these chemotherapy agents are associated with specific side effects. Cisplatin, for example, is known to induce severe ototoxicity. Dox, an anthracycline, exerts its therapeutic effects through a variety of mechanisms, including DNA intercalation, interference with type II topoisomerases, generation of free radicals, and inhibition of DNA and RNA synthesis (Zsíros et al. 2010 ). Cardiomyopathy stands as a recognized side effect of Dox therapy (Zsiros et al. 2013 ). Unfortunately, specific medications targeting DOX-induced cardiotoxicity are currently unavailable.

In the present study, our analysis of public databases revealed a significant overexpression of IMPDH2 in hepatoblastoma (HB) tissues, particularly in cases associated with metastasis and recurrence of the disease. Subsequently, we conducted an immunohistochemical assessment of IMPDH2 protein expression using tissue microarray (TMA) analysis. This was followed by an investigation into the correlations between IMPDH2 protein expression levels and various clinicopathological variables, as well as the overall survival (OS) and event-free survival (EFS) of the patients. Furthermore, we experimentally knocked down IMPDH2 in HB cell lines to determine its effects on cellular functions and its relationship with the cell cycle. Ultimately, our research led to the discovery of a synergistic interaction between mycophenolate mofetil (MMF), an inhibitor of the IMPDH2 target, and the chemotherapeutic agent doxorubicin (DOX).

Materials and methods

Patients and specimens.

Tissue samples were obtained from a cohort of 129 patients diagnosed with hepatoblastoma (HB) at Renji Hospital, Shanghai, over a decade-long period from May 2013 to August 2023. For the study, we constructed tissue microarrays (TMAs) that included a total of 129 paraffin-embedded HB tumor specimens, 21 paraffin-embedded specimens of adjacent normal liver tissue from HB patients, and 10 paraffin-embedded specimens of normal liver tissue from controls. Furthermore, we randomly selected eight pairs of tumor and corresponding normal tissues for analysis via reverse transcription quantitative polymerase chain reaction (RT-qPCR). Surgical interventions for all patients included in the study were performed at Renji Hospital. The study protocol was approved by the Ethics Committee of Renji Hospital, and we obtained written informed consent from the guardians of all participating patients.

Cell culture

The HB cell lines (Huh6, HepG2) and embryonic kidney cell lines (HEK293) were nurtured in Dulbecco’s modified Eagle’s medium, complemented with 10% fetal bovine serum. Incubation was sustained in a humidified environment at 37 °C, with 5% carbon dioxide. Cell passages were restricted to < 6 months before experimentation.

Dataset acquisition and process

Three microarray datasets (GSE151347, GSE131329, GSE133039) sourced from the GEO database were utilized for investigating IMPDH2 expression in HB cancer tissues and normal tissues. Differential expression profiles of genes were generated through GEO2R analysis. Subsequent analysis and visualization of Gene Ontology (GO) results were conducted employing R Studio software (version 4.2.3). The HSA Synergy Score was assessed utilizing the online tool SynergyFinder. ( https://tangsoftwarelab.shinyapps.io/synergyfinder/_w_f98c38d7/#!/dashboard ).

TMA staining and immunohistochemistry (IHC) assay

TMA sections, each 4 μm thick and prepared using a Leica RM2245 microtome, underwent heat treatment at 60 °C for 30 min, followed by three cycles of immersion in xylene for 10 min per cycle and subsequent immersion in absolute ethanol for 10 min per cycle. After rinsing with running water, experiments were conducted. IHC was performed to investigate IMPDH2 expression in the TMA, utilizing anti-IMPDH2 as the primary antibody. A semi-quantitative scoring system was devised, considering both the proportion of cells displaying positive staining and the intensity of staining. The positive staining proportions were graded as follows: 0 for absence, 1 for 0–25%, 2 for 25–50%, 3 for 50–75%, and 4 for 75–100%. Staining intensity was rated on a scale of 0 for absence, 1 for weak, 2 for moderate, and 3 for strong. The combined score was calculated by multiplying these two sub-scores, allowing for the classification of samples into categories reflecting negative expression (0), low expression (1–3), moderate expression (4–6), and high expression (9–12). Two blinded pathologists independently assessed the IHC staining results without access to clinicopathologic data.

Western blot

Twenty micrograms of protein samples were applied onto 10% sodium dodecyl sulfate-polyacrylamide gel electrophoresis. After separation, these samples were transferred from the gel to nitrocellulose membranes (Millipore, MA, USA), which underwent blocking using 5% bovine serum albumin for 1 h. Subsequently, the membranes were subjected to an overnight incubation at 4 °C with anti-β-actin, IMPDH2, or specific antibodies as indicated. Following washing with PBST, secondary antibody incubation was conducted for 1 h, and subsequent visualization was achieved by exposure of the membranes to photographic film. Table S2 contains detailed information about the utilized antibodies.

Quantitative real-time PCR (RT-qPCR)

Real-time PCR analyses were conducted in triplicate, utilizing 25 ng of cDNA per reaction and 10 µM forward and reverse primers, employing 2× ChamQ SYBR Color qPCR Master Mix (Vazyme Biotech) on a Bio-Rad CFX 96 system (Bio-Rad, California, USA). Details of primer sequences are available in Tables S3. The PCR cycling protocol initiated with a 30-second denaturation at 95 °C, succeeded by 40 cycles consisting of 10 s at 95 °C for denaturation and 30 s for annealing/elongation. Normalization of target gene expression to GAPDH was performed, and the relative expression of target genes was determined employing the 2 −ΔΔCT method.

Cell proliferation and colony formation assays

Cell proliferation was evaluated utilizing the CCK-8 Kit (Beyotime, China). Assessment of cell colony-forming ability post-IMPDH2 silencing was conducted via colony formation assay. Cells, transfected with shIMPDH2 or shNC controls, were seeded into 12-well plates and incubated for a duration of 2 weeks. Subsequently, they were rinsed twice with PBS, followed by fixation using 4% formaldehyde for 15 min, staining with crystal violet for an additional 15 min, and subsequent photographic documentation.

HepG2 and Huh6 cell lines were maintained in a serum-depleted medium within the upper transwell chamber (Corning, USA) containing an 8.0 μm pore polycarbonate membrane. The lower chamber was supplemented with medium containing 20% FBS. After a 24-hour incubation period, cells that traversed the membrane and adhered to its undersurface were fixed and stained using crystal violet.

Cell transfection

Lentiviral vectors containing IMPDH2 shRNA plasmids (shIM#1, shIM#2, Supplement) and an empty vector (shNC) were procured from Shanghai Jiaotong University School of Medicine. To develop stable IMPDH2 knockdown HB cell lines, lentiviral particles were produced through co-transfection of the shIMPDH2 plasmid with a mix of packaging plasmids (pMDL, VSVG, pRSV-Rev at a ratio of 5:3:2) using PEI MAX transfection reagent (Polysciences, 24765-100) in HEK-293T cells. The viral particles were collected 48–72 h post-transfection. Subsequently, HepG2 and Huh6 cells were exposed to the lentivirus containing IMPDH2 shRNA or control lentivirus for 48 h and subsequently subjected to puromycin screening to select for cells exhibiting stable knockdown.

Flow Cytometry

Cell cycle analysis was performed using the Cell Cycle Staining Kit, while apoptosis assessment utilized the Annexin V-FITC/PI apoptosis Kit (MULTI SCIENCES, Shanghai, China) as per the manufacturer’s protocols. BD FACSCanto II (USA) was employed for cell collection and analysis. ModFit LT 3.2 software was utilized to compute cell cycle distribution and apoptosis ratio, and Flowjo 10.8.1 was used for the visualization of apoptosis-related images.

Tumorigenesis assays in nude mice

All animal experiments adhered to the Institutional Animal Care and Use Committee guidelines at Shanghai Jiaotong University. Four-week-old male BALB/c nude mice were randomly allocated to specified groups. Subcutaneous tumor models were established using HepG2 and Huh6 cells. The study comprised three groups: two experimental groups with IMPDH2 knockdown (HepG2 shIM#1, HepG2 shIM#2) and one control group (HepG2 shNC), each consisting of three nude mice. Each mouse received subcutaneous injections of HepG2 cells (5 × 10 6 in 50% Matrigel™, BD Biosciences). The huh6 groups were similar to HepG2 groups, differing only in the number of injected cells (1.0 × 10 7 in 25% Matrigel™). Tumor volume was assessed every 2 days using calipers and calculated using the formula (width 2 x length)/2. Finally, humane euthanasia using carbon dioxide was performed, and tumors were harvested and prepared for further analysis.

Statistical analysis

All experiments were conducted in triplicate, and GraphPad Prism software was utilized for data analysis. Results are expressed as mean ± standard error of mean (SEM). Unpaired and paired two-tailed student’s t-tests were employed as the analytical method to assess differences between the two groups. Kaplan–Meier methodology was applied for OS and EFS calculations, and between-group disparities were evaluated using the log-rank test. Clinicopathological characteristics in HB were subjected to analysis using χ2 tests conducted via SPSS software. A P value less than 0.05 was considered statistically significant (* P  < 0.05, ** P  < 0.01, *** P  < 0.001).

IMPDH2 is highly expressed in HB tissues and is associated with metastasis and recurrence

To explore the differential expression of IMPDH2 in hepatoblastoma (HB) relative to healthy tissue, an extensive analysis was undertaken, leveraging datasets from the Gene Expression Omnibus (GEO). The datasets included GSE151347, GSE131329 (comparing tumor to non-tumor tissue), GSE133039 (analyzing recurrence versus non-recurrence), and again GSE151347 (examining metastasis versus non-metastasis). Our results consistently revealed a significant overexpression of IMPDH2 in HB tissue, which was particularly pronounced in cases associated with metastasis and recurrence in HB patients ( Fig.  1 A ) . To substantiate the validity of these public data, we conducted a preliminary assessment of IMPDH2 expression levels using reverse transcription quantitative polymerase chain reaction (RT-qPCR) in a series of HB tissues and their matched healthy liver tissue counterparts at the transcriptome level. These analyses confirmed a marked upregulation of IMPDH2 in HB patients ( Fig.  1 B ) . With the goal of achieving a comprehensive understanding of IMPDH2 expression at the protein level, we constructed a tissue microarray (TMA) encompassing a considerable number of HB cases. Examination of the TMA with an IMPDH2 antibody led to the clear stratification of cases into negative, low, moderate, and high expression groups based on IMPDH2 levels ( Fig.  1 C-D ) . It is noteworthy that the vast majority, 90.2%, showed positive IMPDH2 expression ( Fig.  1 E ) . Furthermore, stark differences were observed between the high/medium and negative/low IMPDH2 expression groups with respect to pre-chemotherapy alpha-fetoprotein (AFP) levels, pre-operative metastasis, and the PRETEXT staging system ( Table  1 ), along with a notable correlation with Ki67 expression. In addition, HB patients with negative/low IMPDH2 expression exhibited significantly better overall survival (OS) and event-free survival (EFS) compared to those with high/medium IMPDH2 expression ( Fig.  1 F-G ) . In aggregate, these observations strongly suggest that the pronounced overexpression of IMPDH2 adversely affects the prognosis of HB patients. Moreover, IMPDH2 holds promise as a potential prognostic and risk prediction biomarker.

figure 1

IMPDH2 is highly expressed in HB tissues and is associated with metastasis and recurrence. A. Left, Venn diagrams displayed DEGs retrieved from the GEO database (GSE151347 T vs. N, GSE131329 T vs. N); the Right, illustrated Venn diagrams of DEGs from the GEO database (GSE133039 R vs. T, GSE151347 M + vs. M-); The middle section depicted the collective genes overlapping across all datasets. P  < 0.001, log2FC > 1. B. The RT-qPCR analysis unveiled the mRNA expression levels of IMPDH2 in HB tissues, encompassing eight pairs of tumors and their corresponding paired normal tissues. C: The schematic diagram illustrating the collection, classification, and analysis of tumor samples. D: Representative images and proportion of HB TMA with different staining intensities. E: The proportion of HB TMA with different staining intensities. F. Patient EFS between IMPDH2 −/low and IMPDH2 medium/high group. G. Patient OS between IMPDH2 −/low and IMPDH2 medium/high group

Knockdown of IMPDH2 suppresses HB cell malignancy in vitro

To elucidate the biological role of inosine monophosphate dehydrogenase 2 (IMPDH2) in hepatoblastoma (HB) cells, we experimentally downregulated IMPDH2 expression in HB cell lines ( Fig.  2 A ) . Subsequently, the proliferation rate of these IMPDH2-knockdown cell lines was evaluated using the CCK-8 assay. Our results demonstrated a significant reduction in the proliferative capacity of the HB cell lines upon IMPDH2 suppression ( Fig.  2 B-C ) . In parallel, the clonogenic potential of the HB cell lines was markedly diminished following IMPDH2 knockdown ( Fig.  2 D ) . Furthermore, transwell migration assays indicated a pronounced inhibition of the migratory capacity of the HB cell lines post-IMPDH2 knockdown ( Fig.  2 E ) . Importantly, previous studies have reported a significant increase in apoptosis following IMPDH2 inhibition in cell lines derived from triple-negative breast cancer and diffuse large B-cell lymphoma(Wang et al. 2021 ; Gao et al. 2023 ). Prompted by these findings, we investigated the impact of IMPDH2 knockdown on apoptosis using flow cytometry. Our analysis revealed a substantial increase in apoptotic cell death following IMPDH2 knockdown ( Fig.  2 F ) . Collectively, these observations suggest that IMPDH2 plays a critical role in promoting the malignant progression of HB cells while concurrently suppressing apoptotic processes within these cell lines.

figure 2

Knockdown of IMPDH2 suppresses HB cell malignancy in vitro and in vivo. A. Western blot assays validating the efficiencies of IMPDH2 knockdown in HB cell lines. B-C. The CCK-8 assay was employed to evaluate proliferative potential in HB cell lines post transfection with IMPDH2 shRNA (shIM#1 and shIM#2) and control shRNA. ( n  = 3, t-test). D. Colony formation assays were utilized to assess the proliferative capacity of HB cell lines following transfection with IMPDH2 shRNA and control shRNA. ( n  = 3, t-test). E. Transwell assays were employed to evaluate the metastatic potential of HB cell lines subsequent to transfection with IMPDH2 shRNA and control shRNA. ( n  = 3, t-test). F. A Flow cytometry analysis was used to assess the apoptosis rate of HB cell lines post transfection with IMPDH2 shRNA and control shRNA. ( n  = 3, t-test). G. The in vivo tumorigenic potential of HepG2, Huh6 cells transfected with IMPDH2 shRNA and control shRNA, respectively. B-C. Assessment of tumor volumes and weights of transfected HepG2 and Huh6 cells was conducted every 2 days, respectively

Knockdown of IMPDH2 inhibits tumor proliferation in vivo

To further investigate the impact of IMPDH2 on hepatoblastoma (HB) proliferation in vivo, HepG2 and Huh6 cells were transfected with stable short hairpin negative control (shNC) and short hairpin IMPDH2 (shIMPDH2) constructs, specifically shIM#1 and shIM#2. These cells were then inoculated into nude mice. After a period of 19 days following the injection of the HepG2 cell line, xenograft tumors formed at the injection site in all mice. Notably, the group treated with shIMPDH2 demonstrated significantly reduced tumor growth rates when compared to the shNC group. A comparable trend was observed in the results from the Huh6 cell line ( Fig.  2 G ) . These findings corroborate that the suppression of IMPDH2 markedly hinders tumor formation and growth.

IMPDH2 regulates the progression of HB through the cell cycle signaling pathway

To delineate the molecular mechanisms underlying the inhibitory effects observed upon IMPDH2 knockdown, we performed RNA sequencing on hepatoblastoma (HB) cell lines with IMPDH2 knockdown. The analysis of differentially expressed genes (DEGs) revealed a significant upregulation of genes linked to cell cycle regulation. Gene Ontology (GO) term enrichment analysis underscored the role of these genes in processes such as the regulation of the mitotic cell cycle, cell cycle phase control, and the transition through the mitotic cell cycle phase ( Fig.  3 A ). This prompted us to focus on the influence of IMPDH2 on the cell cycle in HB cells. The impact of IMPDH2 knockdown on the cell cycle was evaluated using flow cytometry. We observed a pronounced accumulation of cells in the G0/G1 phase in the IMPDH2 knockdown group compared to the IMPDH2 short hairpin negative control (shNC) group, accompanied by a significant decrease in the G2/M phase cell population ( Fig.  3 B ) . The increase in apoptosis observed following IMPDH2 knockdown ( Fig.  2 F ) supports our inference that IMPDH2 knockdown results in G0/G1 cell cycle arrest, which subsequently initiates apoptosis. It is noteworthy that the cell cycle inhibitory protein p21 was significantly upregulated in the IMPDH2 knockdown group relative to the IMPDH2 shNC group. In contrast, genes associated with the G1/S transition, such as CDK6 and cyclin D1, showed marked downregulation. These results were further validated by reverse transcription quantitative polymerase chain reaction (RT-qPCR) analysis ( Fig.  3 C-D ) and confirmed through Western blot analysis ( Fig.  3 E). In summary, our results elucidate that IMPDH2 knockdown induces G0/G1 phase cell cycle arrest and triggers apoptosis in HB cell lines. This is achieved through the upregulation of the cell cycle inhibitory protein p21 and the concurrent downregulation of cell cycle transition genes CDK6 and cyclin D1.

figure 3

IMPDH2 regulates the progression of HB through the Cell Cycle Signaling Pathway. Combination of Doxorubicin with MMF remarkably reduced tumor proliferation and promoted apoptosis in vitro. A. Results of Gene Ontology analysis in HB cell lines. B. Analysis of cell cycle distribution in HB cell lines using flow cytometry following IMPDH2 knockdown: graphical representation of cell cycle distribution and statistical analysis. C-D. Quantification of IMPDH2, P21, CDK6, and cyclin D1 mRNA expression levels in HB cell lines transfected with IMPDH2 shRNA and control shRNA, assessed via real-time PCR. E. Validation of changes in cell cycle-related proteins following knockdown of IMPDH2 using Western blot assays. F. Cellular viability of HB cell lines post-administration of DOX and MMF was evaluated utilizing the CCK-8

Combination of doxorubicin with mmf remarkably reduced tumor proliferation and promoted apoptosis in vitro

Cisplatin, a chemotherapy agent extensively utilized in the treatment of hepatoblastoma (HB), exerts its therapeutic effect by disrupting DNA repair mechanisms, thereby inhibiting cancer cell growth. The efficacy of cisplatin is significantly enhanced when administered in combination with doxorubicin (DOX). Mycophenolate mofetil (MMF) functions as an inhibitor of purine synthesis by targeting inosine monophosphate dehydrogenase (IMPDH) activity, which in turn suppresses the de novo biosynthesis of purine nucleotides. Our hypothesis was centered on the potential synergistic effect of combining DOX’s inhibition of DNA synthesis with MMF’s blockade of de novo DNA synthesis, thereby amplifying their therapeutic impact on tumors. To evaluate this hypothesis, HB cell lines were treated with DOX, MMF, and their combination. The resulting dose-response relationships for each drug were depicted (Fig.  3 F). Subsequently, the computation of the HSA Synergy Score revealed the striking superiority of the DOX-MMF combination treatment (40 nM DOX and 1 µM MMF) over the individual drugs (Fig.  3 G-H). These results indicate that the addition of a specific dosage of MMF can significantly enhance the tumor-suppressive effects of DOX. Furthermore, our examination of the impact of combination therapy on cell apoptosis clearly demonstrated a pronounced pro-apoptotic effect of the combined treatment when compared to monotherapy (Fig. 5I-K). Consequently, we propose that optimizing the concentration of MMF in conjunction with DOX administration could lead to improved treatment outcomes. This strategy could potentially enable the use of reduced DOX dosages, thereby minimizing the side effects associated with MMF.

As the most prevalent class of primary liver malignancies, hepatoblastoma (HB) exhibits a pronounced propensity for metastasis, particularly to the lungs, and often displays inherent resistance to chemotherapy. The poor prognosis and overall survival rates observed in HB patients are largely attributable to recurrent episodes and the dissemination of metastases. Consequently, our study was designed to explore the fundamental mechanisms and potential biomarkers underlying HB.

Within this investigation, elevated IMPDH2 expression was observed among HB patients, which correlated with unfavorable prognostic indicators. Silencing IMPDH2 expression resulted in a significant deceleration of in vitro cellular proliferation and induced cell cycle arrest at the G0/G1 phase in HB cells. Additionally, IMPDH2 suppression led to an increase in P21 expression, disrupting the functionality and binding capacities of CDK6 and Cyclin D1, thereby promoting cellular apoptosis and inhibiting cellular proliferation. These observations position IMPDH2 as a significant contributor to HB progression and suggest its potential as a promising prognostic marker for this disease.

In our study, we collected postoperative samples from HB patients at Renji Hospital spanning the past decade. These samples encompassed tumor tissues as well as paired adjacent normal tissues, with comprehensive clinical data gathered from the corresponding patients. Utilizing these samples, we established the most extensive hepatoblastoma tissue microarray (TMA) to date. Subsequent analysis of publicly accessible databases concerning HB revealed a notable increase in IMPDH2 expression at the transcriptional level, significantly correlating with instances of metastasis and recurrence.

Further evaluation of IMPDH2 protein expression using the TMA platform confirmed its elevated levels in HB. Integration of clinical parameters revealed a significant association, with patients exhibiting heightened IMPDH2 expression tending to present at older ages (Table  1 ), validating prior research findings (Haeberle et al. 2020 ). Moreover, approximately 20% of pediatric HB cases present with metastatic disease at diagnosis, predominantly affecting the lungs and significantly impacting patient prognosis(Hu et al. 2021 ; Angelico et al. 2019 ). Our investigation identified a strong correlation between HB metastasis and increased IMPDH2 expression. Elevated AFP levels serve as a crucial diagnostic criterion for HB, with the majority of patients demonstrating abnormally high levels. Clinical conditions often parallel AFP levels, and existing literature suggests that both low and extremely high AFP levels in pediatric HB patients correspond to poorer prognoses (Meyers et al. 2017 ; Schweinitz et al. 1997 ). Our findings indicate that among patients with heightened IMPDH2 expression, there is a notably higher prevalence of individuals with AFP levels exceeding 10,000 ng/mL, aligning with previous literature findings in patients exhibiting high IMPDH2 expression. Remarkably, both overall survival (OS) and event-free survival (EFS) demonstrated a significant decline in the cohort characterized by increased IMPDH2 expression, mirroring outcomes observed in various solid malignancies(Duan et al. 2018 ; He et al. 2018 ; Shireman et al. 2021 ). These outcomes underscore the robust reliability of our TMA platform in identifying novel biomarkers. We propose that our platform holds promise for uncovering additional pertinent biomarkers in future research endeavors.

In addition to the cell line sequencing outcomes following IMPDH2 knockdown, which affirmed the correlation between IMPDH2 and the cell cycle, our analysis of GSE131329 yielded substantial findings. The Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway examination (Supplemental Fig.  1 ) identified the cell cycle and Wnt signaling pathway as the foremost pathways of significance. The Wnt/beta-catenin pathway, a pivotal developmental pathway associated with progenitor/stem cells, plays a critical role in activating and expanding these cells during embryogenesis and liver regeneration, thereby facilitating hepatic homeostasis(Monga 2015 ; Russell and Monga 2018 ). Notably, HB demonstrates the highest incidence of beta-catenin mutations among human cancers, closely associated with aberrant Wnt/beta-catenin signaling, reaching rates of up to 90%. This not only validates the reliability of the dataset but also substantiates our investigation. This is evident in our observation that the inhibition of IMPDH2 expression in HB cells resulted in the arrest of the cell cycle at the G0/G1 phase.

Purine nucleotides hold a pivotal role in governing cell cycle dynamics, where IMPDH2 actively participates in their synthesis. These nucleotides are indispensable for DNA and RNA construction, fundamental processes driving cell growth and division. Consequently, IMPDH2 indirectly modulates cell cycle progression by influencing the synthesis of purine nucleotides. Our findings substantiate an expanding body of literature indicating IMPDH2 involvement across diverse cancer types. Notably, heightened IMPDH2 expression in colon cancer cells has been linked to methotrexate resistance(Peñuelas et al. 2005 ). In glioblastoma, IMPDH2 amplifies rRNA and tRNA synthesis, fostering accelerated cell proliferation(Kofuji et al. 2019 ). Furthermore, elevated IMPDH2 levels have demonstrated suppression of cancer cell apoptosis through the regulation of multifaceted pathways, including the PI3K/AKT/mTOR and PI3K/c-Myc/AFF4 pathways (Gao et al. 2023 ; Ni et al. 2023 ). In this current investigation, we elucidate a novel mechanism whereby IMPDH2 augments the proliferative and tumorigenic potential of HB cells by modulating the cell cycle pathway.

Surprisingly, our investigation unveiled mycophenolate mofetil (MMF), an FDA-approved medication widely used for post-transplant rejection, as a specific inhibitor of IMPDH2(Lipsky 1996 ). Its combination with doxorubicin (DOX) demonstrated a notably synergistic effect. Therefore, we postulate that, for post-transplant HB patients, utilizing MMF, as opposed to other immunosuppressants, may offer a dual benefit of anti-rejection and tumor suppression. This hypothesis warrants further substantiation through clinical investigations, forming the cornerstone of our subsequent research endeavors.

In summary, our identification of IMPDH2 as a novel prognostic marker and therapeutic target in HB, particularly in cases of metastasis or recurrence, suggests that targeting this factor holds potential to enhance patient prognosis and overall quality of life.

Data availability

No datasets were generated or analysed during the current study.

Angelico R, Grimaldi C, Gazia C, Saffioti MC, Manzia TM, Castellano A, Spada M (2019) How Do Synchronous Lung Metastases Influence the Surgical Management of Children with Hepatoblastoma? An Update and Systematic Review of the Literature Cancers (Basel), 11(11)

Aronson DC, Weeda VB, Maibach R, Czauderna P, Dall’Igna P, de Ville de J, Goyet S, Branchereau G, Perilongo P, Brock J, Zsiros M, Semeraro C, Chardot B, Wildhaber B, Morland, Brugières L (2019) Microscopically positive resection margin after hepatoblastoma resection: what is the impact on prognosis? A Childhood Liver Tumours Strategy Group (SIOPEL) report. Eur J Cancer 106:126–132

Article   PubMed   Google Scholar  

Bell D, Ranganathan S, Tao J, Monga SP (2017) Novel advances in understanding of Molecular Pathogenesis of Hepatoblastoma: a Wnt/β-Catenin perspective. Gene Expr 17(2):141–154

Article   CAS   PubMed   Google Scholar  

Collart FR, Chubb CB, Mirkin BL, Huberman E (1992) Increased inosine-5’-phosphate dehydrogenase gene expression in solid tumor tissues and tumor cell lines. Cancer Res 52(20):5826–5828

CAS   PubMed   Google Scholar  

Czauderna P, Lopez-Terrada D, Hiyama E, Häberle B, Malogolowkin MH, Meyers RL (2014) Hepatoblastoma state of the art: pathology, genetics, risk stratification, and chemotherapy. Curr Opin Pediatr 26(1):19–28

Davies JQ, de la Hall PM, Kaschula RO, Sinclair-Smith CC, Hartley P, Rode H, Millar AJ (2004) Hepatoblastoma–evolution of management and outcome and significance of histology of the resected tumor. A 31-year experience with 40 cases. J Pediatr Surg 39(9):1321–1327

de Fine Licht S, Schmidt LS, Rod NH, Schmiegelow K, Lähteenmäki PM, Kogner P, Träger C, Stokland T, Schüz J (2012) Hepatoblastoma in the nordic countries. Int J Cancer 131(4):E555–E561

Duan S, Huang W, Liu X, Liu X, Chen N, Xu Q, Hu Y, Song W, Zhou J (2018) IMPDH2 promotes colorectal cancer progression through activation of the PI3K/AKT/mTOR and PI3K/AKT/FOXO1 signaling pathways. J Exp Clin Cancer Res 37(1):304

Article   CAS   PubMed   PubMed Central   Google Scholar  

Fellenberg J, Bernd L, Delling G, Witte D, Zahlten-Hinguranage A (2007) Prognostic significance of drug-regulated genes in high-grade osteosarcoma. Mod Pathol 20(10):1085–1094

Gao G, Xue Q, He J, Wu M, Jiang Y, Li Q, Zhang Y, Shi W (2023) Single-cell RNA sequencing in double-hit lymphoma: IMPDH2 induces the progression of lymphoma by activating the PI3K/AKT/mTOR signaling pathway. Int Immunopharmacol, 125(Pt A): p. 111125

Gartel AL, Tyner AL (2002) The role of the cyclin-dependent kinase inhibitor p21 in apoptosis. Mol Cancer Ther 1(8):639–649

Haeberle B, Rangaswami A, Krailo M, Czauderna P, Hiyama E, Maibach R, Lopez-Terrada D, Aronson DC, Alaggio R, Ansari M, Malogolowkin MH, Perilongo G, O’Neill AF, Trobaugh-Lotrario AD, Watanabe K, Schmid I, von Schweinitz D, Ranganathan S, Yoshimura K, Hishiki T, Tanaka Y, Piao J, Feng Y, Rinaldi E, Saraceno D, Derosa M, Meyers RL (2020) The importance of age as prognostic factor for the outcome of patients with hepatoblastoma: analysis from the children’s hepatic tumors international collaboration (CHIC) database. Pediatr Blood Cancer 67(8):e28350

He Y, Zheng Z, Xu Y, Weng H, Gao Y, Qin K, Rong J, Chen C, Yun M, Zhang J, Ye S (2018) Over-expression of IMPDH2 is associated with tumor progression and poor prognosis in hepatocellular carcinoma. Am J Cancer Res 8(8):1604–1614

CAS   PubMed   PubMed Central   Google Scholar  

Heck JE, Meyers TJ, Lombardi C, Park AS, Cockburn M, Reynolds P, Ritz B (2013) Case-control study of birth characteristics and the risk of hepatoblastoma. Cancer Epidemiol 37(4):390–395

Article   PubMed   PubMed Central   Google Scholar  

Hu H, Zhang W, Wang Y, Zhang Y, Yi Y, Gao Y, Chen L, Huang D (2021) Prognostic analysis for children with hepatoblastoma with lung metastasis: a single-center analysis of 98 cases. Asia Pac J Clin Oncol 17(5):e191–e200

Jia X, Liu Y, Cheng Y, Wang Y, Kang H, Ma Z, Chen K (2022) Inosine monophosphate dehydrogenase type1 sustains tumor growth in hepatocellular carcinoma. J Clin Lab Anal 36(5):e24416

Kofuji S, Hirayama A, Eberhardt AO, Kawaguchi R, Sugiura Y, Sampetrean O, Ikeda Y, Warren M, Sakamoto N, Kitahara S, Yoshino H, Yamashita D, Sumita K, Wolfe K, Lange L, Ikeda S, Shimada H, Minami N, Malhotra A, Morioka S, Ban Y, Asano M, Flanary VL, Ramkissoon A, Chow LML, Kiyokawa J, Mashimo T, Lucey G, Mareninov S, Ozawa T, Onishi N, Okumura K, Terakawa J, Daikoku T, Wise-Draper T, Majd N, Kofuji K, Sasaki M, Mori M, Kanemura Y, Smith EP, Anastasiou D, Wakimoto H, Holland EC, Yong WH, Horbinski C, Nakano I, DeBerardinis RJ, Bachoo RM, Mischel PS, Yasui W, Suematsu M, Saya H, Soga T, Grummt I, Bierhoff H, Sasaki AT (2019) IMP dehydrogenase-2 drives aberrant nucleolar activity and promotes tumorigenesis in glioblastoma. Nat Cell Biol 21(8):1003–1014

Linabery AM, Ross JA (2008) Trends in childhood cancer incidence in the U.S. (1992–2004). Cancer 112(2):416–432

Lipsky JJ (1996) Mycophenolate mofetil. Lancet 348(9038):1357–1359

Marin JJG, Briz O, Herraez E, Lozano E, Asensio M, Di Giacomo S, Romero MR, Osorio-Padilla LM, Santos-Llamas AI, Serrano MA, Armengol C, Efferth T, Macias RIR (2018) Molecular bases of the poor response of liver cancer to chemotherapy. Clin Res Hepatol Gastroenterol 42(3):182–192

Mauro M, Rego MA, Boisvert RA, Esashi F, Cavallo F, Jasin M, Howlett NG (2012) p21 promotes error-free replication-coupled DNA double-strand break repair Nucleic Acids Res, 40(17): pp. 8348-60

Meyers RL, Maibach R, Hiyama E, Häberle B, Krailo M, Rangaswami A, Aronson DC, Malogolowkin MH, Perilongo G, von Schweinitz D, Ansari M, Lopez-Terrada D, Tanaka Y, Alaggio R, Leuschner I, Hishiki T, Schmid I, Watanabe K, Yoshimura K, Feng Y, Rinaldi E, Saraceno D, Derosa M, Czauderna P (2017) Risk-stratified staging in paediatric hepatoblastoma: a unified analysis from the children’s hepatic tumors international collaboration. Lancet Oncol 18(1):122–131

Monga SP (2015) β-Catenin signaling and roles in Liver Homeostasis, Injury, and Tumorigenesis. Gastroenterology 148(7):1294–1310

Natsumeda Y, Ohno S, Kawasaki H, Konno Y, Weber G, Suzuki K (1990) Two distinct cDNAs for human IMP dehydrogenase. J Biol Chem 265(9):5292–5295

Ni C, Liu W, Zheng K, Guo S, Song B, Jing W, Li G, Li B, Ni C, Shi K, Jin G, Yu G (2023) PI3K/ c-Myc/AFF4 axis promotes pancreatic tumorigenesis through fueling nucleotide metabolism. Int J Biol Sci 19(6):1968–1982

Peñuelas S, Noé V, Ciudad CJ (2005) Modulation of IMPDH2, survivin, topoisomerase I and vimentin increases sensitivity to methotrexate in HT29 human colon cancer cells. Febs j 272(3):696–710

Ransom JT (1995) Mechanism of action of mycophenolate mofetil. Ther Drug Monit 17(6):681–684

Russell JO, Monga SP (2018) Wnt/β-Catenin signaling in Liver Development, Homeostasis, and Pathobiology. Annu Rev Pathol 13:351–378

Shireman JM, Atashi F, Lee G, Ali ES, Saathoff MR, Park CH, Savchuk S, Baisiwala S, Miska J, Lesniak MS, James CD, Stupp R, Kumthekar P, Horbinski CM, Ben-Sahra I, Ahmed AU (2021) De novo Purine Biosynth is Major Driv Chemoresistance Glioblastoma Brain 144(4):1230–1246

Google Scholar  

Shu Q, Nair V (2008) Inosine monophosphate dehydrogenase (IMPDH) as a target in drug discovery. Med Res Rev 28(2):219–232

Suthanthiran M, Strom TB (1997) Immunoregulatory drugs: mechanistic basis for use in organ transplantation. Pediatr Nephrol 11(5):651–657

Trobaugh-Lotrario AD, Katzenstein HM (2012) Chemotherapeutic approaches for newly diagnosed hepatoblastoma: past, present, and future strategies. Pediatr Blood Cancer 59(5):809–812

Trobaugh-Lotrario AD, Meyers RL, Tiao GM, Feusner JH (2016) Pediatric liver transplantation for hepatoblastoma. Transl Gastroenterol Hepatol 1:44

Tulla M, Berthold F, Graf N, Rutkowski S, von Schweinitz D, Spix C, Kaatsch P (2015) Incidence, trends, and survival of children with embryonal tumors. Pediatrics 136(3):e623–e632

von Schweinitz D, Hecker H, Schmidt-von-Arndt G, Harms D (1997) Prognostic factors and staging systems in childhood hepatoblastoma. Int J Cancer 74(6):593–599

Article   Google Scholar  

Wang W, Wu Y, Chen S, Liu X, He J, Wang S, Lu W, Tang Y, Huang J (2021) Shikonin is a novel and selective IMPDH2 inhibitor that target triple-negative breast cancer. Phytother Res 35(1):463–476

Xiong Y, Hannon GJ, Zhang H, Casso D, Kobayashi R, Beach D (1993) 21 is a universal inhibitor of cyclin kinases. Nature 366(6456):701–704

Yang LY, Chang RM, Lau WY, Ou DP, Wu W, Zeng ZJ (2014) Mesohepatectomy for centrally located large hepatocellular carcinoma: indications, techniques, and outcomes. Surgery 156(5):1177–1187

Zhou L, Xia D, Zhu J, Chen Y, Chen G, Mo R, Zeng Y, Dai Q, He H, Liang Y, Jiang F, Zhong W (2014) Enhanced expression of IMPDH2 promotes metastasis and advanced tumor progression in patients with prostate cancer. Clin Transl Oncol 16(10):906–913

Zou J, Han Z, Zhou L, Cai C, Luo H, Huang Y, Liang Y, He H, Jiang F, Wang C, Zhong W (2015) Elevated expression of IMPDH2 is associated with progression of kidney and bladder cancer. Med Oncol 32(1):373

Zsiros J, Brugieres L, Brock P, Roebuck D, Maibach R, Zimmermann A, Childs M, Pariente D, Laithier V, Otte JB, Branchereau S, Aronson D, Rangaswami A, Ronghe M, Casanova M, Sullivan M, Morland B, Czauderna P, Perilongo G (2013) Dose-dense cisplatin-based chemotherapy and surgery for children with high-risk hepatoblastoma (SIOPEL-4): a prospective, single-arm, feasibility study Lancet Oncol, 14(9): pp. 834 – 42

Zsíros J, Maibach R, Shafford E, Brugieres L, Brock P, Czauderna P, Roebuck D, Childs M, Zimmermann A, Laithier V, Otte JB, de Camargo B, MacKinlay G, Scopinaro M, Aronson D, Plaschkes J, Perilongo G (2010) Successful treatment of childhood high-risk hepatoblastoma with dose-intensive multiagent chemotherapy and surgery: final results of the SIOPEL-3HR study. J Clin Oncol 28(15):2584–2590

Download references

Acknowledgements

This study was funded by Shanghai Science and Technology Development Foundation (Outstanding academic leader) (HF, 23XD1423100, National Natural Science Foundation, China (XQ, 82241221). G-H. The HSA Synergy Score was computed using the online analytical tool SynergyFinder. I-K. Apoptosis in HB cell lines under single (40 nM DOX; 1µM MMF) and combination drug treatments was analyzed using flow cytometry.

Author information

Linman Li, Yichi Wu, Hong-ting Huang and Yi Zhou contributed equally to this work.

Authors and Affiliations

Department of Liver Surgery, Renji Hospital (Punan Branch), School of Medicine, Shanghai Jiao Tong University, Shanghai, 200127, China

Linman Li, Yichi Wu, Hong-ting Huang, June-kong Yong, Zicheng Lv, Jie Zhao, Zhifeng Xi, Hao Feng & Qiang Xia

Clinical Research Unit, Renji Hospital, School of Medicine, Shanghai Jiao Tong University, Shanghai, 200127, China

Zicheng Lv & Hao Feng

Shanghai Engineering Research Centre of Transplantation and Immunology, Shanghai, 200127, China

Linman Li, Yichi Wu, Yi Zhou, Xuelin Xiang, Zhifeng Xi, Hao Feng & Qiang Xia

You can also search for this author in PubMed   Google Scholar

Contributions

LLM, FH, and XQ participated in the conception and design of this study. HHT, LLM, and FH was the project manager and coordinated patient recruitment. LLM, ZZJ, YJ, HHT, TH were involved in the acquisition, analysis, or interpretation of data. LLM, WYC, LZ, HHT and ZZJ drafted the manuscript. All the authors contributed to the critical review and final approval of the manuscript. FH, XQ, and LZC accessed and verified the underlying study data. All authors were responsible for the decision to submit the manuscript.

Corresponding authors

Correspondence to Hao Feng or Qiang Xia .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Li, L., Wu, Y., Huang, Ht. et al. IMPDH2 suppression impedes cell proliferation by instigating cell cycle arrest and stimulates apoptosis in pediatric hepatoblastoma. J Cancer Res Clin Oncol 150 , 377 (2024). https://doi.org/10.1007/s00432-024-05858-4

Download citation

Received : 28 March 2024

Accepted : 17 June 2024

Published : 01 August 2024

DOI : https://doi.org/10.1007/s00432-024-05858-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Liver transplantation
  • Hepatocellular carcinoma
  • Neoadjuvant therapy
  • Recurrence-free survival
  • Immunotherapy
  • Hepatoblastoma
  • Pediatric cancer
  • Find a journal
  • Publish with us
  • Track your research

COMMENTS

  1. Cancer Treatment Research

    Find research articles on cancer treatment, including news stories, clinical trials, blog posts, and descriptions of active studies. ... A new NCI-supported study showed that altering cancer cell metabolism by feeding mice a diet very low in the nutrient methionine improved the ability of chemotherapy and radiation therapy to shrink tumors.

  2. Clinical Trials Information

    You can donate your medical data and health samples >. See how donating your existing medical records and health samples can save lives. Explains clinical trials, including what they are, why they are important, things to think about when deciding to take part, and questions to ask your doctor.

  3. Cancer Cell Lines Are Useful Model Systems for Medical Research

    1. Introduction. Cancer cell lines are valuable in vitro model systems that are widely used in cancer research and drug discovery [ 1 ]. Their use is primarily linked to their peculiar capability to provide an indefinite source of biological material for experimental purposes [ 2 ]. The establishment of a new cell line is a very complex process ...

  4. Cancer Biology, Epidemiology, and Treatment in the 21st Century

    The Biology of Cancer. Cancer is a disease that begins with genetic and epigenetic alterations occurring in specific cells, some of which can spread and migrate to other tissues. 4 Although the biological processes affected in carcinogenesis and the evolution of neoplasms are many and widely different, we will focus on 4 aspects that are particularly relevant in tumor biology: genomic and ...

  5. The changing landscape of cancer cell therapies: clinical trials and

    The number of new cell therapy trials has decreased in the past two years. Our analysis shows that the number of new trials peaked in 2021 and has since decreased by 3.9% in 2022 (led by a 10.5% ...

  6. DCB

    Credit: National Cancer Institute. Research in cancer cell biology seeks to define the biological basis underlying the differences between normal cells and cancerous cells. This includes studies of the fundamental mechanisms that drive pre-cancer states, oncogenic transformation, and that support tumor growth and behavior.

  7. Cell Press: Cancer Cell

    About. Cancer Cell. Cancer Cell provides a high-profile forum to promote major advances in cancer research and oncology. The primary criterion for considering manuscripts is whether the studies provide major advances into answering important questions relevant to naturally occurring cancers. More.

  8. Five decades of advances in cancer research: Cell

    Paradigm-shifting work then taught us that our own cells harbor causative oncogenes. Our understanding of cancer—its origins, evolution, and vulnerabilities—exploded from there. Along the way, research outcomes enabled earlier and more accurate cancer diagnosis and drug development, saving millions of lives. On this 50 th anniversary of ...

  9. Cell cycle control in cancer

    Cancer is a group of diseases in which cells divide continuously and excessively. Cell division is tightly regulated by multiple evolutionarily conserved cell cycle control mechanisms, to ensure ...

  10. The Clinical Relevance of Cancer Cell Lines

    Cross-Contamination of Cancer Cell Lines. Nelson-Rees and colleagues were perhaps the first researchers, back in the 1970s, to question experimentally the relevance of these in vitro cancer models ().Their cytogenetic studies revealed that many cancer cell lines used to model various types of cancers were in fact derived from the HeLa cervical cancer line and not from other cancers (6,8).

  11. The next big questions in cancer research: Cell

    This is one of the big questions to address in cancer, and thanks to critical advances in recent years, we are much closer to achieving this goal. Pan-cancer studies, incorporating single-cell analyses and spatial transcriptomics, have revealed intricate cellular ecosystems connecting different cell types in the TME and defined spatial niches ...

  12. Scaffold-based 3D cell culture models in cancer research

    Three-dimensional (3D) cell cultures have emerged as valuable tools in cancer research, offering significant advantages over traditional two-dimensional (2D) cell culture systems. In 3D cell cultures, cancer cells are grown in an environment that more closely mimics the 3D architecture and complexity of in vivo tumors. This approach has revolutionized cancer research by providing a more ...

  13. Cancer Cell Lines for Drug Discovery and Development

    In the early 1990s, the National Cancer Institute (NCI; Bethesda, MD) introduced a "disease-oriented" drug screening approach using a panel of 60 human cancer cell lines derived from nine different types of cancer (brain, colon, leukemia, lung, melanoma, ovarian, renal, breast, and prostate; ref. 1).This approach was designed to facilitate high-throughput screening of large numbers of ...

  14. Cancer Cell Lines Are Useful Model Systems for Medical Research

    Abstract. Cell lines are in vitro model systems that are widely used in different fields of medical research, especially basic cancer research and drug discovery. Their usefulness is primarily linked to their ability to provide an indefinite source of biological material for experimental purposes. Under the right conditions and with appropriate ...

  15. Cancer Cell Research

    Cancer Cell Research (ISSN 2161-2609) is an international, peer-reviewed open access journal on oncology. It publishes reviews, regular research papers and short communications. ... Cancer Cell Research publishes highly significant studies in a broad range of cancer research. Molecular and cellular mechanisms of fundamental cancer biology, such ...

  16. Cancer Cell Research

    Cancer Cell Research. Cancer Cell Research is an open access journal dedicated to publishing findings resulting from conferences, workshops, and similar events, in all areas of chemistry. The conference organizers and proceedings editors are responsible for managing the peer-review process and selecting papers for conference proceedings.

  17. Researchers devise novel solution to preventing relapse after CAR T

    In experiments in patient-derived laboratory cancer cell lines and other studies, CAR-Enhancer treatment succeeded in eradicating all tumor cells, clearing the way for clinical trials of this ...

  18. Research Areas: Cancer Biology

    For example, studies of cell signaling pathways in normal cells and cancer cells have contributed greatly to our knowledge about the disease, revealing molecular alterations that are shared among different types of cancer and pointing to possible treatment strategies. Decades of basic research in cancer biology have created a broad base of ...

  19. MSK Research Highlights, July 31, 2024

    New research from Memorial Sloan Kettering Cancer Center (MSK) showed giving chemotherapy shortly after a stem cell or bone marrow transplant from a less than perfectly matched donor greatly reduces the chances that the patient will develop graft-versus-host disease (GVHD); and sheds new light on cell state changes in prostate cancer.

  20. Dietary glutamine may be linked to B-cell lymphomas in abdominal lymph

    Researchers have discovered that dietary glutamine, an amino acid found in many foods, may be linked to the development of certain kinds of B-cell lymphomas in abdominal structures called mesenteric lymph nodes in mouse models. The study appeared July 18, 2024, in Nature Immunology. B-cell lymphoma is a common form of cancer of the lymphatic ...

  21. Advances in diffuse large B-cell lymphoma treatment

    Study Provides Insights about Chemorefractory Diffuse Large B-Cell Lymphoma Patients who relapse early after first-line chemotherapy are often, but not always, refractory to second-line chemotherapy July 30, 2024 / Cancer / Research

  22. New approaches and procedures for cancer treatment: Current

    Cancer is a global health problem responsible for one in six deaths worldwide. Treating cancer has been a highly complex process. Conventional treatment approaches, such as surgery, chemotherapy, and radiotherapy, have been in use, while significant advances are being made in recent times, including stem cell therapy, targeted therapy, ablation therapy, nanoparticles, natural antioxidants ...

  23. Recent study reveals key immune cells as critical factors in lung

    Non-small cell lung cancer accounts for ~85% of lung tumors and is a leading cause of death in adults. ... Apart from any fair dealing for the purpose of private study or research, no part may be ...

  24. Pericytes recruited by CCL28 promote vascular normalization after anti

    Lung cancer stands as the foremost cause of mortality among patients with malignant tumors, with lung adenocarcinoma being a predominant pathological subtype [1, 2].Angiogenesis is a crucial hallmark of lung adenocarcinoma, and significant strides have been made in applying anti-angiogenesis therapy for its treatment [].Extensive research data indicates that anti-angiogenic tumor therapy ...

  25. Advancing Cancer Research and Medicine with Single-Cell Genomics

    Single-cell sequencing (SCS) has impacted on many areas of cancer research and improved our understanding of intratumor heterogeneity, the tumor microenvironment, metastasis, and therapeutic resistance. The development and refinement of SCS technologies has led to massive reductions in costs, increased cell throughput, and improved reproducibility, paving the way for clinical applications.

  26. Pre-existing skin-resident CD8 and γδ T cell circuits mediate immune

    Abstract. Merkel cell carcinoma (MCC) is an aggressive neuroendocrine skin cancer with a ~50% response rate to immune checkpoint blockade (ICB) therapy. To identify predictive biomarkers, we integrated bulk and single-cell RNA-seq with spatial transcriptomics from a cohort of 186 samples from 116 patients, including bulk RNA-seq from 14 matched pairs pre- and post-ICB. In non-responders ...

  27. Iron Chelation Therapy Elicits Innate Immune Control of Metastatic

    Abstract. Iron accumulation in tumors contributes to disease progression and chemoresistance. While targeting this process can influence various hallmarks of cancer, the immunomodulatory effects of iron chelation in the tumor microenvironment are unknown. Here, we report that treatment with deferiprone, an FDA-approved iron chelator, unleashes innate immune responses that restrain ovarian ...

  28. Cancer Research

    The 2022 report includes data on new cancer cases and deaths plus a detailed examination of trends in pancreatic cancer and its subtypes. NCI joins the cancer community in advancing the goals of the National Cancer Plan as part of its research programs. NCI-Designated Cancer Centers deliver cutting-edge cancer treatments and conduct state-of ...

  29. The impact of impaired tissue fixation in resected non-small-cell lung

    Studies. Study Purpose. The purpose of this study was to assess the effects of under- and over-fixation of formalin-fixed, paraffin-embedded (FFPE) non-small cell lung cancer (NSCLC) tissue on tissue quality (determined macroscopically and microscopically), immunohistochemical staining of 10 antigens, and DNA degradation.

  30. IMPDH2 suppression impedes cell proliferation by instigating ...

    Background Hepatoblastoma (HB) is the most common pediatric liver tumor, presenting significant therapeutic challenges due to its high rates of recurrence and metastasis. While Inosine Monophosphate Dehydrogenase 2(IMPDH2) has been associated with cancer progression, its specific role and clinical implications in HB have not been fully elucidated. Methods This study utilized Quantitative Real ...