Plasmon-Resonant Gold Nanostructures: Comparison
Please note this is a comparison between Version 2 by Peter Tang and Version 1 by Anisha Veeren.

Liposomes can sequester a variety of bioactive water-soluble ions, ligands and enzymes, and oligonucleotides. The bilayer that separates the liposome interior from the exterior solution provides a physical barrier to contents release and degradation. Tethering plasmon-resonant, hollow gold nanoshells to the liposomes, or growing gold nanoparticles directly on the liposome exterior, allows liposome contents to be released by nanosecond or shorter pulses of near-infrared light (NIR). Gold nanoshells or nanoparticles strongly adsorb NIR light; cells, tissues, and physiological media are transparent to NIR, allowing penetration depths of millimeters to centimeters. Nano to picosecond pulses of NIR light rapidly heat the gold nanoshells, inducing the formation of vapor nanobubbles, similar to cavitation bubbles.

  • : nanobubble
  • plasmon-resonant

1. Introduction

Following their discovery by Bangham in 1965 [1,2][1][2], liposomes have been one of the most thoroughly investigated nanocarriers for drug delivery [3,4,5,6,7,8,9,10,11,12,13][3][4][5][6][7][8][9][10][11][12][13]. Many promising drugs are discarded due to difficulties in maintaining a safe biodistribution in the concentration range necessary for efficacy [14]. Liposomes and other lipid-based drug carriers sequester toxic drugs within a lipid bilayer to alter drug biodistribution, thereby enhancing efficacy while minimizing damage to healthy tissue and organs [14]. Liposomes are capable of concentrating and stabilizing small molecule therapeutic compounds and larger oligonucleotides for delivery in vivo and in vitro [12,15,16,17,18][12][15][16][17][18]. Liposomes based on naturally occurring lipids are inherently biocompatible and hence can overcome many of the obstacles to cellular and tissue uptake. Small liposomes (<200 nm) are passively targeted to tumors and sites of inflammation by the enhanced permeation and retention (EPR) effect [18,19][18][19]. Liposomes can also be actively targeted by treating their surface with antibodies to ligands overexpressed on tumor cells [20,21,22,23,24,25,26][20][21][22][23][24][25][26]. Hydrophilic cargoes can be encapsulated within the aqueous liposome core, while hydrophobic molecules can be trapped within the hydrophobic bilayer interior [27].

2. Separating Cargo Sequestration from Rapid Release

Despite these advantages, it is difficult for a single bilayer to simultaneously prevent drug release and trigger rapid release at the preferred site of action [28]. It remains a challenge to initiate a biological or chemical response in cells or tissues with spatial and temporal control. Conventional liposomes often release small molecule drugs prematurely in the circulation due to degradation of the liposome bilayer or are taken up in the liver and spleen by interactions with the reticuloendothelial system (RES). “Stealth” coatings of polyethylene glycol and other polymer lipids can prolong the circulation time [29,30][29][30]. However, optimizing the liposome cargo retention in the circulation to maximize drug accumulation at the disease site may slow drug release from the liposome at the site of interest and reduce therapeutic activity. This is the case for clinical doxorubicin liposomes (Doxil) [31,32,33][31][32][33] and for clinically tested cisplatin liposomes [34]; drug release is so slow that the critical drug concentration is not easily achieved. As a result, multiple strategies for enhancing local drug release at the site of interest have included the incorporation of membrane destabilizing agents that respond to stimuli such as general hyperthermia [35[35][36][37],36,37], receptor targets [38[38][39][40],39,40], pH changes [41], or tumor-specific enzymes [42,43][42][43]. These strategies promote release but can compromise long-term liposome stability and decrease drug retention in the circulation. Specific ligands with high affinities to receptor overexpressed on disease cells can in turn lead to “binding-site barriers” where high concentrations of bound nanocarriers can prevent drug penetration into tissue [44].

3. Plasmon-Resonant Gold Nanostructures

To address the challenge of spatially and temporally controlled fast release, modified liposome nanocarriers have been developed in which the liposome functions primarily as the cargo container. Rapid spatial- and temporal-controlled contents release is addressed by tethered gold nanostructures triggered by an externally applied pulsed NIR laser. Such liposomes are also well-suited for in vitro, in vivo, and ex vivo use where spatial- and temporal-controlled delivery of cargo ranges from small molecules to complex proteins or oligonucleotides [45,46,47,48][45][46][47][48]. Plasmon-resonant hollow gold nanoshells (HGN) or simple gold nanoparticles [46,47,49,50][46][47][49][50] can be chemically tethered to conventional polyethylene glycol stabilized liposomes of any membrane composition to absorb and concentrate near-infrared light (NIR) [32,33,48,51,52,53][32][33][48][51][52][53]. A hollow shell structure of gold, silver, or alloys can strongly absorb NIR light at a characteristic localized surface plasmon resonance (LSPR) wavelength (Figure 1) that can be tuned by controlling the nanoparticle size and shell thickness [45]. As shown schematically in Figure 1, the absorption spectra of common physiological molecules such as water, oxygenated hemoglobin (HbO2), and hemoglobin (Hb) show a distinct minima in what has been called the “NIR window” [54]. As a result, NIR light can penetrate deeply within cell suspensions, tissue, and other physiological media. Unstructured solid silver or gold nanoparticles of 10–100 nm size absorb in the same wavelength range as hemoglobin (400–540 nm) and cannot be readily addressed in physiological fluids.
Figure 1. Schematic representation of optical adsorption of water (H2O), oxygenated hemoglobin (HbO2), and hemoglobin (Hb) showing the near-infrared window (NIR) from 650–900 nm over which physiological materials do not adsorb significant fractions of incident light, adapted from [54], published by Nature, 2001[54]. Hollow gold nanoshells (HGN) have a surface plasmon resonance determined by the HGN wall thickness and diameter that can be tuned to peak in the NIR Window. This causes HGN suspensions to appear dark blue to purple.
By manipulating the shape and organization of gold nanostructures, the surface plasmon-resonance can be moved to the NIR window. Light absorption from high power nanosecond or shorter pulses cause the gold nanostructures to undergo a rapid increase in temperature before any heat can be dissipated to the surrounding fluids [45,55][45][55]. Following this rapid heating of the nanoparticle, the hot nanoparticles begin to dissipate their thermal energy to the surrounding solution in microseconds. The high temperature gradients around the gold nanoparticles vaporize a minute amount of the surrounding water leading to the formation of unstable nanobubbles as the heat is dissipated (30 nm diameter particles thermally equilibrate with their surroundings in fractions of a microsecond). These nanobubbles rapidly expand and contract and give rise to microjets that have similar mechanical and thermal effects as ultrasound-induced cavitation (Figure 2) [51,56,57][51][56][57]. Nanobubble generation requires a threshold energy; any additional energy above threshold makes the nanobubbles grow larger. Once formed, the mechanical forces generated by the expansion and collapse of the nanobubbles can lyse endosomes, liposomes, or cells, providing rapid contents release [51[51][58][59][60][61],58,59,60,61], cell poration [62[62][63],63], or even cell death [62,63,64,65][62][63][64][65]. It is these mechanical cavitation forces that lead to endosome, liposome, and cell membrane rupture and contents release rather than the change in temperature. As a result, complex proteins and oligonucleotides, as well as small molecules, can be delivered without damage following NIR light triggering of the gold nanostructures [47,48,58,61,66][47][48][58][61][66].
Figure 2. Plasmon-resonant gold nanoparticles tethered to liposomes can be triggered by short pulses of NIR laser light (red lines) that heat the nanoparticles. At a threshold light fluence, heat dissipating from the nanoparticles boils a minute amount of water to form unstable vapor nanobubbles that rapidly expand and contract. As the nanobubbles collapse, liquid–vapor microjets form that can perforate cell and liposome membranes. This allows the liposome contents to be rapidly released and cells to be perforated.

References

  1. Bangham, A.D.; Standish, M.M.; Watkins, J.C. Diffusion of univalent ions across the lamellae of swollen phospholipids. J. Mol. Biol. 1965, 13, 238–252, IN26–IN27.
  2. Deamer, D.W. From “Banghasomes” to liposomes: A memoir of Alec Bangham, 1921–2010. FASEB J. 2010, 24, 1308–1310.
  3. Xing, H.; Hwang, K.; Lu, Y. Recent Developments of Liposomes as Nanocarriers for Theranostic Applications. Theranostics 2016, 6, 1336–1352.
  4. Pattni, B.S.; Chupin, V.V.; Torchilin, V.P. New Developments in Liposomal Drug Delivery. Chem. Rev. 2015, 115, 10938–10966.
  5. Filipczak, N.; Pan, J.Y.; Yalamarty, S.S.K.; Torchilin, V.P. Recent advancements in liposome technology. Adv. Drug Deliv. Rev. 2020, 156, 4–22.
  6. Kieler-Ferguson, H.M.; Chan, D.; Sockolosky, J.; Finney, L.; Maxey, E.; Vogt, S.; Szoka, F.C., Jr. Encapsulation, controlled release, and antitumor efficacy of cisplatin delivered in liposomes composed of sterol-modified phospholipids. Eur. J. Pharm. Sci. 2017, 103, 85–93.
  7. Szoka, F.; Papahadjopoulos, D. Comparative Properties and Methods of Preparation of Lipid Vesicles (Liposomes). Ann Rev Biophys. Bioeng. 1980, 9, 467–508.
  8. Anderson, L.J.E.; Hansen, E.; Lukianova-Hleb, E.Y.; Hafner, J.H.; Lapotko, D.O. Optically guided controlled release from liposomes with tunable plasmonic nanobubbles. J. Control. Release 2010, 144, 151–158.
  9. Discher, D.E.; Eisenberg, A. Polymer Vesicles. Science 2002, 297, 967–973.
  10. Kube, S.; Hersch, N.; Naumovska, E.; Gensch, T.; Hendriks, J.; Franzen, A.; Landvogt, L.; Siebrasse, J.-P.; Kubitscheck, U.; Hoffmann, B.; et al. Fusogenic Liposomes as Nanocarriers for the Delivery of Intracellular Proteins. Langmuir 2017, 33, 1051–1059.
  11. Mohamed, M.; Abu Lila, A.S.; Shimizu, T.; Alaaeldin, E.; Hussein, A.; Sarhan, H.A.; Szebeni, J.; Ishida, T. PEGylated liposomes: Immunological responses. Sci. Technol. Adv. Mater. 2019, 20, 710–724.
  12. Barenholz, Y. Doxil®—The first FDA-approved nano-drug: Lessons learned. J. Control. Release Off. J. Control. Release Soc. 2012, 160, 117–134.
  13. Has, C.; Sunthar, P. A comprehensive review on recent preparation techniques of liposomes. J. Liposome Res. 2020, 30, 336–365.
  14. Allen, T.M.; Cullis, P.R. Drug delivery systems: Entering the Mainstream. Science 2004, 303, 1818–1822.
  15. Schwendener, R.A. Liposomes in biology and medicine. In Bio-Applications of Nanoparticles; Chan, W.C.W., Ed.; Springer: Berlin, Germany, 2007; pp. 117–128.
  16. Allen, T.M.; Cullis, P.R. Liposomal drug delivery systems: From Concept to Clinical Applications. Adv. Drug Deliv. Rev. 2013, 65, 36–48.
  17. Wang, N.; Chen, M.N.; Wang, T. Liposomes used as a vaccine adjuvant-delivery system: From basics to clinical immunization. J. Control. Release 2019, 303, 130–150.
  18. Maruyama, K. Intracellular targeting delivery of liposomal drugs to solid tumors based on EPR effects. Adv. Drug Deliv. Rev. 2011, 63, 161–169.
  19. Torchilin, V. Tumor delivery of macromolecular drugs based on the EPR effect. Drug Deliv. Rev. 2011, 63, 131–135.
  20. Riaz, M.K.; Riaz, M.A.; Zhang, X.; Lin, C.; Wong, K.H.; Chen, X.; Zhang, G.; Lu, A.; Yang, Z. Surface functionalization and targeting strategies of liposomes in solid tumor therapy: A review. Int. J. Mol. Sci. 2018, 19, 195.
  21. Willis, M.; Forssen, E. Ligand-targeted liposomes. Adv. Drug Deliv. Rev. 1998, 29, 249–271.
  22. Maruyama, K. In Vivo Targeting by Liposomes. Biol. Pharm. Bull. 2000, 23, 791–799.
  23. Wang, M.; Thanou, M. Targeting nanoparticles to cancer. Pharmacol. Res. 2010, 62, 90–99.
  24. Deshpande, P.P.; Biswas, S.; Torchilin, V.P. Current trends in the use of liposomes for tumor targeting. Nanomedicine 2013, 8, 1509–1528.
  25. Noble, G.T.; Stefanick, J.F.; Ashley, J.D.; Kiziltepe, T.; Bilgicer, B. Ligand-targeted liposome design: Challenges and Fundamental Considerations. Trends Biotechnol. 2014, 32, 32–45.
  26. Agrawal, M.; Ajazuddin; Tripathi, D.K.; Saraf, S.; Saraf, S.; Antimisiaris, S.G.; Mourtas, S.; Hammarlund-Udenaes, M.; Alexander, A. Recent advancements in liposomes targeting strategies to cross blood-brain barrier (BBB) for the treatment of Alzheimer’s disease. J. Control. Release 2017, 260, 61–77.
  27. Tan, C.; Wang, J.; Sun, B. Biopolymer-liposome hybrid systems for controlled delivery of bioactive compounds: Recent advances. Biotechnol. Adv. 2021, 48, 107727.
  28. Allen, T.M.; Hansen, C.B.; Lopes de Menezes, D.E. Pharmacokinetics of long-circulating liposomes. Adv. Drug Deliv. Rev. 1995, 16, 267–284.
  29. Papahadjopolous, D.; Allen, T.M.; Gabizon, A.; Mayhew, E.; Matthay, K.; Huang, S.L.; Lee, K.D.; Woodle, M.C.; Lasic, D.D.; Redemann, C. Sterically stabilized liposomes: Improvements in Pharmacokinetics and Antitumour Therapeutic Efficacy. Proc. Nat. Acad. Sci. USA 1991, 88, 11460–11464.
  30. Gabizon, A.; Papahadjopoulos, D. Liposome formulations with prolonged circulation time in blood and enhanced uptake by tumors. Proc. Natl. Acad. Sci. USA 1988, 85, 6949–6953.
  31. Abraham, S.A.; Waterhouse, D.N.; Mayer, L.D.; Cullis, P.R.; Madden, T.D.; Bally, M.B. The Liposomal Formulation of Doxorubicin. Methods Enzymol. 2005, 391, 71–97.
  32. Forbes, N.; Shin, J.E.; Ogunyankin, M.; Zasadzinski, J.A. Inside-outside self-assembly of light-activated fast-release liposomes. Phys. Chem. Chem. Phys. 2015, 17, 15569–15578.
  33. Forbes, N.; Pallaoro, A.; Reich, N.O.; Zasadzinski, J.A. Rapid, reversible release from thermosensitive liposomes triggered by near infrared light. Part. Part. Syst. Charact. 2014, 31, 1158–1167.
  34. White, S.C.; Lorigan, P.; Margison, G.P.; Margison, J.M.; Martin, F.; Thatcher, N.; Anderson, H.; Ranson, M. Phase II study of SPI-77 (sterically stabilised liposomal cisplatin) in advanced non-small-cell lung cancer. Br. J. Cancer 2006, 95, 822–828.
  35. Ponce, A.M.; Vujaskovic, Z.; Yuan, F.; Needham, D.; Dewhirst, M.W. Hyperthermia mediated liposomal drug delivery. Int. J. Hyperth. 2006, 22, 205–213.
  36. Manzoor, A.A.; Lindner, L.H.; Landon, C.D.; Park, J.-Y.; Simnick, A.J.; Dreher, M.R.; Das, S.; Hanna, G.; Park, W.; Chilkoti, A.; et al. Overcoming Limitations in Nanoparticle Drug Delivery: Triggered, Intravascular Release to Improve Drug Penetration into Tumors. Cancer Res. 2012, 72, 5566–5575.
  37. Ta, T.; Porter, T.M. Thermosensitive liposomes for localized delivery and triggered release of chemotherapy. J. Control. Release 2013, 169, 112–125.
  38. Ruoslahti, E. Vascular zip codes in angiogenesis and metastasis. Biochem. Soc. Trans. 2004, 32, 397–402.
  39. Ruoslahti, E. Peptides as Targeting Elements and Tissue Penetration Devices for Nanoparticles. Adv. Mater. 2012, 24, 3747–3756.
  40. Noble, C.O.; Kirpotin, D.B.; Hayes, M.E.; Mamot, C.; Hong, K.; Park, J.W.; Benz, C.C.; Marks, J.D.; Drummond, D.C. Development of ligand-targeted liposomes for cancer therapy. Expert Opin. Ther. Targets 2004, 8, 335–353.
  41. Oishi, M.; Kataoka, K.; Nagasaki, Y. pH-responsive three-layered PEGylated polyplex micelle based on a lactosylated ABC triblock copolymer as a targetable and endosome disruptive nonviral gene vector. Bioconjugate Chem. 2006, 17, 677–688.
  42. Jørgensen, K.; Davidsen, J.; Mouritsen, O.G. Biophysical mechanisms of phospholipase A2 activation and their use in liposome-based drug delivery. FEBS Lett. 2002, 531, 23–27.
  43. Davidsen, J.; Jørgensen, K.; Andresen, T.L.; Mouritsen, O.G. Secreted phospholipase A2 as a new enzymatic trigger mechanism for localised liposomal drug release and absorption in diseased tissue. Biochim. Biophys. Acta (BBA)-Biomembr. 2003, 1609, 95–101.
  44. Barua, S.; Mitragotri, S. Challenges associated with penetration of nanoparticles across cell and tissue barriers: A review of current status and future prospects. Nano Today 2014, 9, 223–243.
  45. Ogunyankin, M.O.; Shin, J.E.; Lapotko, D.O.; Ferry, V.; Zasadzinski, J.A. Optimizing the NIR Fluence Threshold for Nanobubble Generation by Controlled Synthesis of 10–40 nm Hollow Gold Nanoshells. Adv. Funct. Mater. 2018, 28, 1705272.
  46. Li, X.; Che, Z.; Mazhar, K.; Price, T.J.; Qin, Z. Ultrafast Near-Infrared Light-Triggered Intracellular Uncaging to Probe Cell Signaling. Adv. Funct. Mater. 2017, 27, 1605778.
  47. Xiong, H.; Li, X.; Kang, P.; Perish, J.; Neuhaus, F.; PLoSki, J.E.; Kroener, S.; Ogunyankin, M.O.; Shin, J.E.; Zasadzinski, J.A.; et al. Near-Infrared Light Triggered-Release in Deep Brain Regions Using Ultra-photosensitive Nanovesicles. Angew. Chem. Int. Ed. 2020, 59, 8608–8615.
  48. Shin, J.E.; Ogunyankin, M.O.; Zasadzinski, J.A. Near Infrared-Triggered Liposome Cages for Rapid, Localized Small Molecule Delivery. Sci. Rep. 2020, 10, 1706–1711.
  49. Troutman, T.S.; Leung, S.J.; Romanowski, M. Light-Induced Content Release from Plasmon-Resonant Liposomes. Adv. Mater. 2009, 21, 2334–2338.
  50. Troutman, T.S.; Barton, J.K.; Romanowski, M. Biodegradable Plasmon Resonant Nanoshells. Adv. Mater. 2008, 20, 2604–2608.
  51. Wu, G.; Mikhailovsky, A.; Khant, H.A.; Fu, C.; Chiu, W.; Zasadzinski, J.A. Remotely Triggered Liposome Release by Near-Infrared Light Absorption via Hollow Gold Nanoshells. J. Am. Chem. Soc. 2008, 130, 8175–8177.
  52. Wu, G.H.; Mikhailovsky, A.; Khant, H.A.; Zasadzinski, J.A. Synthesis, characterization an optical response of gold nanoshells used to trigger release from liposomes. In Methods in Enzymology; Nejat, D., Ed.; Academic Press: Burlington, VT, USA, 2009; Volume 464, pp. 279–307.
  53. Forbes, N. Photothermally activated drug release from liposomes coupled to hollow gold nanoshells. SPIE Proc. Plasmon. Biol. Med. VIII 2011, 7911, 24–30.
  54. Weissleder, R. A clearer vision for in vivo imaging. Nat. Biotechnol. 2001, 19, 316–317.
  55. Prevo, B.G.; Esakoff, S.A.; Mikhailovsky, A.; Zasadzinski, J.A. Scalable Routes to Gold Nanoshells with Tunable Sizes and Response to Near-Infrared Pulsed-Laser Irradiation. Small 2008, 4, 1183–1195.
  56. Paliwal, S.; Mitragotri, S. Ultrasound-induced cavitation: Applications in Drug and Gene Delivery. Expert Opin. Drug Deliv. 2006, 3, 713–726.
  57. Pecha, R.; Gompf, B. Microimplosions: Cavitation Collapse and Shock Wave Emission on a Nanosecond Time Scale. Phys. Rev. Lett. 2000, 84, 1328–1330.
  58. Braun, G.B.; Pallaoro, A.; Wu, G.; Missirlis, D.; Zasadzinski, J.A.; Tirrell, M.; Reich, N.O. Laser-Activated Gene Silencing via Gold Nanoshell−siRNA Conjugates. ACS Nano 2009, 3, 2007–2015.
  59. Huang, X.; Hu, Q.; Braun, G.B.; Pallaoro, A.; Morales, D.P.; Zasadzinski, J.; Clegg, D.O.; Reich, N.O. Light-activated RNA interference in human embryonic stem cells. Biomaterials 2015, 63, 70–79.
  60. Huang, X.; Pallaoro, A.; Braun, G.B.; Morales, D.P.; Ogunyankin, M.O.; Zasadzinski, J.; Reich, N.O. Modular Plasmonic Nanocarriers for Efficient and Targeted Delivery of Cancer-Therapeutic siRNA. Nano Lett. 2014, 14, 2046–2051.
  61. Morales, D.P.; Braun, G.B.; Pallaoro, A.; Chen, R.; Huang, X.; Zasadzinski, J.A.; Reich, N.O. Targeted Intracellular Delivery of Proteins with Spatial and Temporal Control. Mol. Pharm. 2015, 12, 600–609.
  62. Lukianova-Hleb, E.Y.; Mutonga, M.B.G.; Lapotko, D.O. Cell-Specific Multifunctional Processing of Heterogeneous Cell Systems in a Single Laser Pulse Treatment. ACS Nano 2012, 6, 10973–10981.
  63. Lukianova-Hleb, E.Y.; Yvon, E.S.; Shpall, E.J.; Lapotko, D.O. All-in-one processing of heterogeneous human cell grafts for gene and cell therapy. Mol. Ther.-Methods Clin. Dev. 2016, 3, 16012.
  64. Lukianova-Hleb, E.Y.; Ren, X.; Sawant, R.R.; Wu, X.; Torchilin, V.P.; Lapotko, D.O. On-demand intracellular amplification of chemoradiation with cancer-specific plasmonic nanobubbles. Nat. Med. 2014, 20, 778–784.
  65. Lukianova-Hleb, E.Y.; Ren, X.Y.; Sawant, R.R.; Gillenwater, A.M.; Kupferman, M.; Hanna, E.Y.; Zasadzinski, J.A.; Wu, X.W.; Torchilin, V.P.; Lapotko, D.O. Plasmonic nanobubble theranostics for intra-operative and preventive treatment of head and neck squamous cell carcinoma. In Proceedings of the Conference on Photonic Therapeutics and Diagnostics X, San Francisco, CA, USA, 1–2 February 2014; Volume 8926.
  66. Morales, D.P.; Morgan, E.N.; McAdams, M.; Chron, A.B.; Shin, J.E.; Zasadzinski, J.A.; Reich, N.O. Light-Triggered Genome Editing: Cre Recombinase Mediated Gene Editing with Near-Infrared Light. Small 2018, 14, 1800543.
More
ScholarVision Creations