Skip to main content

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.

  • Protocol
  • Published:

Preparation and biomedical applications of programmable and multifunctional DNA nanoflowers

Abstract

We describe a comprehensive protocol for the preparation of multifunctional DNA nanostructures termed nanoflowers (NFs), which are self-assembled from long DNA building blocks generated via rolling-circle replication (RCR) of a designed template. NF assembly is driven by liquid crystallization and dense packaging of building blocks, which eliminates the need for conventional Watson-Crick base pairing. As a result of dense DNA packaging, NFs are resistant to nuclease degradation, denaturation or dissociation at extremely low concentrations. By manually changing the template sequence, many different functional moieties including aptamers, bioimaging agents and drug-loading sites could be easily integrated into NF particles, making NFs ideal candidates for a variety of applications in biomedicine. In this protocol, the preparation of multifunctional DNA NFs with highly tunable sizes is described for applications in cell targeting, intracellular imaging and drug delivery. Preparation and characterization of functional DNA NFs takes 5 d; the following biomedical applications take 10 d.

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

Access options

Buy this article

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

Figure 1: Schematic illustration of noncanonical self-assembly and biomedical applications of multifunctional DNA NFs.
Figure 2: Designing the RCR template and primer for DNA NF preparation.
Figure 3: SEM imaging of the morphology and size of NFs during the growth process with reaction time from 0.5 to 30 h.
Figure 4: Bioimaging of intracellular behavior of NFs and Dox delivered by NFs.

Similar content being viewed by others

References

  1. Rosi, N.L. et al. Oligonucleotide-modified gold nanoparticles for intracellular gene regulation. Science 312, 1027–1030 (2006).

    Article  CAS  Google Scholar 

  2. Wu, P. & Yan, X.-P. Doped quantum dots for chemo/biosensing and bioimaging. Chem. Soc. Rev. 42, 5489–5521 (2013).

    Article  CAS  Google Scholar 

  3. Fan, Q. et al. Transferring biomarker into molecular probe: melanin nanoparticle as a naturally active platform for multimodality imaging. J. Am. Chem. Soc. 136, 15185–15194 (2014).

    Article  CAS  Google Scholar 

  4. Yuan, Q. et al. Photon-manipulated drug release from a mesoporous nanocontainer controlled by azobenzene-modified nucleic acid. ACS Nano 6, 6337–6344 (2012).

    Article  CAS  Google Scholar 

  5. Zheng, J. et al. A spherical nucleic acid platform based on self-assembled DNA biopolymer for high-performance cancer therapy. ACS Nano 7, 6545–6554 (2013).

    Article  CAS  Google Scholar 

  6. Dunn, S.S. et al. Reductively responsive siRNA-conjugated hydrogel nanoparticles for gene silencing. J. Am. Chem. Soc. 134, 7423–7430 (2012).

    Article  CAS  Google Scholar 

  7. Kang, H. et al. Near-infrared light-responsive core–shell nanogels for targeted drug delivery. ACS Nano 5, 5094–5099 (2011).

    Article  CAS  Google Scholar 

  8. Yang, L. et al. Engineering polymeric aptamers for selective cytotoxicity. J. Am. Chem. Soc. 133, 13380–13386 (2011).

    Article  CAS  Google Scholar 

  9. Liu, X., Jin, X. & Ma, P.X. Nanofibrous hollow microspheres self-assembled from star-shaped polymers as injectable cell carriers for knee repair. Nat. Mater. 10, 398–406 (2011).

    Article  Google Scholar 

  10. Shim, M.S. & Xia, Y. A reactive oxygen species (ROS)-responsive polymer for safe, efficient, and targeted gene delivery in cancer cells. Angew. Chem. Int. Ed. 52, 6926–6929 (2013).

    Article  CAS  Google Scholar 

  11. Lovell, J.F. et al. Porphysome nanovesicles generated by porphyrin bilayers for use as multimodal biophotonic contrast agents. Nat. Mater. 10, 324–332 (2011).

    Article  CAS  Google Scholar 

  12. Wu, Y. et al. DNA aptamer–micelle as an efficient detection/delivery vehicle toward cancer cells. Proc. Natl. Acad. Sci. USA 107, 5–10 (2010).

    Article  CAS  Google Scholar 

  13. Chen, T. et al. DNA micelle flares for intracellular mRNA imaging and gene therapy. Angew. Chem. Int. Ed. 125, 2066–2070 (2013).

    Article  Google Scholar 

  14. Rothemund, P.W.K. Folding DNA to create nanoscale shapes and patterns. Nature 440, 297–302 (2006).

    Article  CAS  Google Scholar 

  15. Yin, P., Choi, H.M., Calvert, C.R. & Pierce, N.A. Programming biomolecular self-assembly pathways. Nature 451, 318–322 (2008).

    Article  CAS  Google Scholar 

  16. Bath, J. & Turberfield, A.J. DNA nanomachines. Nat. Nanotechnol. 2, 275–284 (2007).

    Article  CAS  Google Scholar 

  17. Jiang, Q. et al. DNA origami as a carrier for circumvention of drug resistance. J. Am. Chem. Soc. 134, 13396–13403 (2012).

    Article  CAS  Google Scholar 

  18. Lee, H. et al. Molecularly self-assembled nucleic acid nanoparticles for targeted in vivo siRNA delivery. Nat. Nanotechnol. 7, 389–393 (2012).

    Article  CAS  Google Scholar 

  19. Goodman, R.P. et al. Rapid chiral assembly of rigid DNA building blocks for molecular nanofabrication. Science 310, 1661–1665 (2005).

    Article  CAS  Google Scholar 

  20. Pei, H. et al. Reconfigurable three-dimensional DNA nanostructures for the construction of intracellular logic sensors. Angew. Chem. Int. Ed. 51, 9020–9024 (2012).

    Article  CAS  Google Scholar 

  21. Lee, J.B. et al. Multifunctional nanoarchitectures from DNA-based ABC monomers. Nat. Nanotechnol. 4, 430–436 (2009).

    Article  CAS  Google Scholar 

  22. Meng, H.-M. et al. DNA dendrimer: an efficient nanocarrier of functional nucleic acids for intracellular molecular sensing. ACS Nano 8, 6171–6181 (2014).

    Article  CAS  Google Scholar 

  23. Dirks, R.M. & Pierce, N.A. Triggered amplification by hybridization chain reaction. Proc. Natl. Acad. Sci. USA 101, 15275–15278 (2004).

    Article  CAS  Google Scholar 

  24. Choi, H.M. et al. Programmable in situ amplification for multiplexed imaging of mRNA expression. Nat. Biotechnol. 28, 1208–1212 (2010).

    Article  CAS  Google Scholar 

  25. Zhang, F., Nangreave, J., Liu, Y. & Yan, H. Structural DNA nanotechnology: state of the art and future perspective. J. Am. Chem. Soc. 136, 11198–11211 (2014).

    Article  CAS  Google Scholar 

  26. Zadeh, J.N. et al. NUPACK: analysis and design of nucleic acid systems. J. Comput. Chem. 32, 170–173 (2011).

    Article  CAS  Google Scholar 

  27. Douglas, S.M. et al. Rapid prototyping of 3D DNA-origami shapes with caDNAno. Nucleic Acids Res. 37, 5001–5006 (2009).

    Article  CAS  Google Scholar 

  28. Tamkovich, S.N. et al. Circulating DNA and DNase activity in human blood. Ann. NY Acad. Sci. 1075, 191–196 (2006).

    Article  CAS  Google Scholar 

  29. Hamblin, G.D. et al. Rolling circle amplification-templated DNA nanotubes show increased stability and cell penetration ability. J. Am. Chem. Soc. 134, 2888–2891 (2012).

    Article  CAS  Google Scholar 

  30. Shu, D. et al. Thermodynamically stable RNA three-way junction for constructing multifunctional nanoparticles for delivery of therapeutics. Nat. Nanotechnol. 6, 658–667 (2011).

    Article  CAS  Google Scholar 

  31. Zhu, G. et al. Noncanonical self-assembly of multifunctional DNA nanoflowers for biomedical applications. J. Am. Chem. Soc. 135, 16438–16445 (2013).

    Article  CAS  Google Scholar 

  32. Hu, R. et al. DNA nanoflowers for multiplexed cellular imaging and traceable targeted drug delivery. Angew. Chem. Int. Ed. 53, 5821–5826 (2014).

    Article  CAS  Google Scholar 

  33. Johne, R. et al. Rolling-circle amplification of viral DNA genomes using phi29 polymerase. Trends Microbiol. 17, 205–211 (2009).

    Article  CAS  Google Scholar 

  34. Shangguan, D. et al. Aptamers evolved from live cells as effective molecular probes for cancer study. Proc. Natl. Acad. Sci. USA 103, 11838–11843 (2006).

    Article  CAS  Google Scholar 

  35. Tuerk, C. & Gold, L. Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 249, 505–510 (1990).

    Article  CAS  Google Scholar 

  36. Xiong, X. et al. Nucleic acid aptamers for living cell analysis. Annu. Rev. Anal. Chem. 7, 405–426 (2014).

    Article  CAS  Google Scholar 

  37. Bagalkot, V., Farokhzad, O.C., Langer, R. & Jon, S. An aptamer–doxorubicin physical conjugate as a novel targeted drug-delivery platform. Angew. Chem. Int. Ed. 45, 8149–8152 (2006).

    Article  CAS  Google Scholar 

  38. Chaires, J.B., Herrera, J.E. & Waring, M.J. Preferential binding of daunomycin to 5′TACG and 5′TAGC sequences revealed by footprinting titration experiments. Biochemistry 29, 6145–6153 (1990).

    Article  CAS  Google Scholar 

  39. Frederick, C.A. et al. Structural comparison of anticancer drug-DNA complexes: adriamycin and daunomycin. Biochemistry 29, 2538–2549 (1990).

    Article  CAS  Google Scholar 

  40. Biffi, G., Tannahill, D., McCafferty, J. & Balasubramanian, S. Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 5, 182–186 (2013).

    Article  CAS  Google Scholar 

  41. Modi, S. et al. Two DNA nanomachines map pH changes along intersecting endocytic pathways inside the same cell. Nat. Nanotechnol. 8, 459–467 (2013).

    Article  CAS  Google Scholar 

  42. Wu, P., Hwang, K., Lan, T. & Lu, Y. A DNAzyme-gold nanoparticle probe for uranyl ion in living cells. J. Am. Chem. Soc. 135, 5254–5257 (2013).

    Article  CAS  Google Scholar 

  43. Sefah, K. et al. Development of DNA aptamers using cell-SELEX. Nat. Protoc. 5, 1169–1185 (2010).

    Article  CAS  Google Scholar 

  44. Dua, P. et al. Alkaline phosphatase ALPPL-2 is a novel pancreatic carcinoma-associated protein. Cancer Res. 73, 1934–1945 (2013).

    Article  CAS  Google Scholar 

  45. Fang, X. & Tan, W. Aptamers generated from cell-SELEX for molecular medicine: a chemical biology approach. Acc. Chem. Res. 43, 48–57 (2009).

    Article  Google Scholar 

  46. Shangguan, D. et al. Cell-specific aptamer probes for membrane protein elucidation in cancer cells. J. Proteome Res. 7, 2133–2139 (2008).

    Article  CAS  Google Scholar 

  47. Zhu, G. et al. Self-assembled aptamer-based drug carriers for bispecific cytotoxicity to cancer cells. Chem. Asian J. 7, 1630–1636 (2012).

    Article  CAS  Google Scholar 

  48. Zhu, G. et al. Self-assembled, aptamer-tethered DNA nanotrains for targeted transport of molecular drugs in cancer theranostics. Proc. Natl. Acad. Sci. USA 110, 7998–8003 (2013).

    Article  CAS  Google Scholar 

  49. Peer, D. et al. Nanocarriers as an emerging platform for cancer therapy. Nat. Nanotechnol. 2, 751–760 (2007).

    Article  CAS  Google Scholar 

  50. Petros, R.A. & DeSimone, J.M. Strategies in the design of nanoparticles for therapeutic applications. Nat. Rev. Drug Discov. 9, 615–627 (2010).

    Article  CAS  Google Scholar 

  51. Sun, W. et al. Cocoon-like self-degradable DNA nanoclew for anticancer drug delivery. J. Am. Chem. Soc. 136, 14722–14725 (2014).

    Article  CAS  Google Scholar 

  52. Wang, Y. et al. Aptamer/graphene oxide nanocomplex for in situ molecular probing in living cells. J. Am. Chem. Soc. 132, 9274–9276 (2010).

    Article  CAS  Google Scholar 

  53. Qian, R., Ding, L. & Ju, H. Switchable fluorescent imaging of intracellular telomerase activity using telomerase-responsive mesoporous silica nanoparticle. J. Am. Chem. Soc. 135, 13282–13285 (2013).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Key Scientific Program of China (2011CB911000), National Natural Science Foundation of China (NSFC) grants (21325520, 21327009, 21221003, J1210040, 21177036 and 21135001) and the China National Instrumentation Program 2011YQ03012412. It is also supported by the US National Institutes of Health (GM079359 and CA133086).

Author information

Authors and Affiliations

Authors

Contributions

G.Z., R.H., X.Z. and W.T. originated the method of preparing DNA NFs; Y.L., G.Z., R.H., X.Z., L.M., Q.L., L.Q., C.W. and W.T. were responsible for the content of the protocol. Y.L. and W.T. wrote the manuscript, and all authors contributed to the revision of the manuscript.

Corresponding authors

Correspondence to Xiaobing Zhang or Weihong Tan.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Predicted secondary structures of the linear RCR template and elongated concatamer RCR product.

(a) The linear RCR template is predicted to have a 3-way junction secondary structure. Sequence of template is shown in Table 1. (b) The elongated concatamer RCR product is predicted to have branched aptamer structures protruding and aligning on alternative sides and many dsDNA (for drug loading) on the backbone and stem. Structures were predicted using the Nupack software.

Supplementary Figure 2 Electrophoresis and confocal fluorescence microscopy characterization of NFs.

a) Agarose gel (2%) electrophoresis image indicating the elongation of DNA through RCR. Lane M: 20 bp DNA marker, lane 1: template, lane 2: primer, lane 3: template and primer after T4 DNA ligase treatment. The tailing results from the complicated structures formed through intra- and intermolecular base pairs. Lane 4: RCR product after phi29 DNA polymerase treatment. Polymerization of dNTP monomers and dense packaging-driven assembly results in huge molecular weight of RCR products; therefore, the corresponding band shows no migration. Hu, R. et al. DNA nanoflowers for multiplexed cellular imaging and traceable targeted drug delivery. Angew. Chem. Int. Ed. 2014. Volume 53. Pages 5821–5826. Copyright Wiley-VCH Verlag GmbH & Co. KGaA. Adapted with permission. b) Confocal fluorescence microscopy imaging of Cy5-dUTP-integrated NFs. The fluorescence signal (Red) of Cy5 (maximum excitation wavelength is 650 nm and maximum emission wavelength is 662 nm) is obvious and indicates the successful polymerization of Cy5-dUTP and the normal function of Cy5. NFs with a size of 2 μm (RCR for 20-30 h) are used for this imaging. Scale bar, 5 μm. Adapted with permission from J. Am. Chem. Soc. 2013, 135(44), 16438–16445. Copyright 2013 American Chemical Society.

Supplementary Figure 3 Characterizations of NFs.

a) Dynamic light scattering measurement of NFs. DLS data reveals the size distribution of NFs and the average radius is calculated to be about 150 nm. NFs with 15 h of RCR reaction time are used in this experiment. Hu, R. et al. DNA nanoflowers for multiplexed cellular imaging and traceable targeted drug delivery. Angew. Chem. Int. Ed. 2014. Volume 53. Pages 5821–5826. Copyright Wiley-VCH Verlag GmbH & Co. KGaA. Adapted with permission. b) TEM images of NFs displays ultrathin sheet sections (indicated by arrows). c) Enlarged partial view of the dashed box. These results provide clear evidence of the internal hierarchical structures and the dense DNA packaging in NFs. NFs with 10 h of RCR reaction time are used for this experiment. Adapted from ref. 31. d) AFM imaging of NFs. e) Cross-sectional plot indicated with the white line. AFM imaging displays that the NFs are monodisperse with a size of 100-200 nm. NFs with 10 h of RCR reaction time are used for this experiment. Adapted from ref. 31. f) Bright field. g) Polarized light imaging of (f). When exposed between crossed polarizers, NFs displayed disc-shaped optical textures as a result of the birefringence of liquid crystalline structure, which is an optical property of anisotropic materials. This is a direct demonstration of liquid crystalline structures of NFs. NFs with 10-15 h of RCR reaction time are used for this experiment. Panels b–g adapted with permission from J. Am. Chem. Soc. 2013, 135(44), 16438–16445. Copyright 2013 American Chemical Society.

Supplementary Figure 4 Exceptional stability of NFs.

(a−c) SEM images of NFs treated with DNase I (a, b, 5 U/mL) and human serum (c, 10% diluted) for 24 h at 37 °C. (d−f) SEM images displaying NFs heated at 170 °C for 0.5 h (d), treated with urea (5 M) for 0.5 h (e), and diluted 100 times (f). The maintenance of NF structural integrity indicates the high stability of NFs. Adapted with permission from J. Am. Chem. Soc. 2013, 135(44), 16438–16445. Copyright 2013 American Chemical Society.

Supplementary Figure 5 Flow cytometry results demonstrating the selective recognition of sgc8c-incorporated NFs.

(a) CEM cells (target cell line, suspension); dark green=cell only, light green=cells incubated with NFs. (b) Ramos cells (nontarget cell line, suspension); dark green=cell only, light green=cells incubated with NFs. (c) HeLa cells (target cell line, adherent); gray=cell only, red=cells incubated with NFs. Target cell lines show an obvious fluorescence signal shift compared with nontarget cell line, which demonstrates a high selectivity of NFs toward target cells. These results show that the conjugated aptamer preserves its binding affinity and specificity. NFs with a size of 200~300 nm (RCR for 10-15 h) are used for this experiment. Angew. Chem. Int. Ed. 2014. Volume 53. Pages 5821–5826. Copyright Wiley-VCH Verlag GmbH & Co. KGaA. Adapted with permission.

Supplementary Figure 6 Drug-loading into the preferable drug-associated sequences of NFs and the cytotoxicity of NFs toward target/nontarget cells.

(a) Drug-loading into the preferable drug-associated sequences of NFs. Left: NFs are incubated with Dox in Dulbecco’s PBS at room temperature. The whole mixture shows uniform red, which is the color of Dox solution. Right: After incubation for 24 hours, the mixture is centrifuged at 14,000 g for 5 min. The supernatant turns light upon losing part of the Dox, while the NF concentrate turns dark red upon Dox loading. Both observation with the naked eye and the following ultraviolet absorption measurement demonstrate the drug-loading ability of DNA NFs. NFs with a size of 200~300 nm (RCR for 10-15 h) are used for this experiment. (b) The correspondence between NF concentration and cell viability. Red line, cell viability of target CEM cells with increasing NF concentration; black line, cell viability of nonarget Ramos cells with increasing NF concentration. No significant cytotoxicity of NFs toward either target or nontarget cells is observed in panel a, which indicates the satisfactory biocompatibility of NFs. (c) SEM images of NFs used in this cytotoxicity experiment. Scale bar: 50 μm. Inset panel, high-resolution imaging of a single NF. Scale bar: 300 nm. NFs with a size of 200~300 nm (RCR for 10-15 h) are used for this experiment. Panels b and c adapted with permission from J. Am. Chem. Soc. 2013, 135(44), 16438–16445. Copyright 2013 American Chemical Society.

Supplementary Figure 7 MTS assay results show selective cytotoxicity of Dox delivered by NFs.

(a) Cell viability of target HeLa cells with NF-Dox complex or free Dox treatment. (b) Cell viability of target CEM cells with NF-Dox complex or free Dox treatment. (c) Cell viability of nontarget Ramos cells with NF-Dox complex or free Dox treatment. Red triangle, cytotoxicity of free Dox; black dots, cytotoxicity of NF-Dox complex. Free Dox shows nonselectivity toward both target (HeLa and CEM cells) and nontarget (Ramos cells) cell lines, while NF-Dox complex only shows cytotoxicity toward target cell lines (HeLa and CEM cells). In contrast to nonselective cytotoxicity of free Dox in both target cells and nontarget cells, the selective cytotoxicity of Dox delivered by NFs indicates their ability to perform targeted drug delivery. NFs with a size of 200~300 nm (RCR for 10-15 h) were used for this experiment. Adapted with permission from J. Am. Chem. Soc. 2013, 135(44), 16438–16445. Copyright 2013 American Chemical Society.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–7, Supplementary Method (PDF 980 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lv, Y., Hu, R., Zhu, G. et al. Preparation and biomedical applications of programmable and multifunctional DNA nanoflowers. Nat Protoc 10, 1508–1524 (2015). https://doi.org/10.1038/nprot.2015.078

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nprot.2015.078

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research