Next Article in Journal
Recent Advances in Plant Early Signaling in Response to Herbivory
Next Article in Special Issue
Inorganic-Organic Hybrid Nanomaterials for Therapeutic and Diagnostic Imaging Applications
Previous Article in Journal
Micro Electromechanical Systems (MEMS) Based Microfluidic Devices for Biomedical Applications
Previous Article in Special Issue
Nanomedicine: Application Areas and Development Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Magnetic Nanoparticles to Gene Delivery

1
Research Team for Vascular Medicine, Tokyo Metropolitan Institute of Gerontology, 35-2 Sakae-cho, Itabashi-ku, Tokyo 173-0015, Japan
2
Laboratory for Medical Engineering, Division of Materials and Chemical Engineering, Yokohama National University, 79-1 Tokiwadai, Hodogaya-ku, Yokohama 240-8501, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2011, 12(6), 3705-3722; https://doi.org/10.3390/ijms12063705
Submission received: 6 May 2011 / Revised: 18 May 2011 / Accepted: 25 May 2011 / Published: 7 June 2011
(This article belongs to the Special Issue Bioactive Nanoparticles (special issue))

Abstract

:
Nanoparticle technology is being incorporated into many areas of molecular science and biomedicine. Because nanoparticles are small enough to enter almost all areas of the body, including the circulatory system and cells, they have been and continue to be exploited for basic biomedical research as well as clinical diagnostic and therapeutic applications. For example, nanoparticles hold great promise for enabling gene therapy to reach its full potential by facilitating targeted delivery of DNA into tissues and cells. Substantial progress has been made in binding DNA to nanoparticles and controlling the behavior of these complexes. In this article, we review research on binding DNAs to nanoparticles as well as our latest study on non-viral gene delivery using polyethylenimine-coated magnetic nanoparticles.

1. Introduction

Nanotechnology describes the creation and utilization of materials, devices, and systems through the control of nanometer-sized materials and their application to physics, chemistry, biology, engineering, materials science, medicine, and other endeavors. In particular, intensive efforts are in progress to develop nanomaterials for medical use as agents that can be targeted to specific organs, tissues, and cells. For example, magnetic nanoparticles (MNPs) are being used clinically as contrast agents for magnetic resonance imaging (MRI) (Table 1). MRI is a noninvasive technique that can provide real-time high-resolution soft tissue information [1,2]. MRI image quality can be further improved by utilizing contrast agents that alter proton relaxation rates [38]. MNP-based drug delivery systems (DDS) [911], and treatments of hyperthermia [1221], using MNPs have been studied for over a decade. Furthermore, researchers have reported that MNPs have been useful in hyperthermic treatment for various cancers in vivo [2231]. Nanotechnology-based anti-cancer agent DDS have already been approved, such as pegylated liposomal doxorubicin (DOXIL) for ovarian cancer [3237]. MNPs have been used effectively as transfection reagents for introducing nucleic acids (plasmids or siRNAs) [3853], or viruses (retrovirus, or adenovirus) [44,5456] into cells. Our own research is focused on MNP-mediated gene delivery systems (called as “Magnetofection”).

2. Gene Delivery

Gene delivery techniques efficiently introduce a gene of interest in order to express its encoded protein in a suitable host or host cell. Currently, there are three primary gene delivery systems that employ viral vectors (retroviruses and adenoviruses), nucleic acid electroporation, and nucleic acid transfection. These systems vary in efficacy (Table 2). Gene delivery by viral vectors can be highly efficient (80–90%) but may insert viral vector nucleic acid sequences into the host genome, potentially causing unwelcome effects, such as inappropriate expression of deleterious genes. Electroporation is also a highly efficient technique for introducing foreign genes into a host (50–70%); however, half of the recipient cells die due to the electrical stimulation. Transfection reagents do not efficiently deliver nucleic acids into cells (20–30%); however, cell viability is largely preserved and the method is safe enough for clinical use. Therefore, this method holds relatively more promise for medical applications, provided that its efficiency can be improved. MNPs are already in use by basic researchers to increase transfection efficiencies of cultured cells. Thus, MNP-nucleic acid complexes are added to cell culture media and then onto the cell surface by applying a magnetic force (Figure 1).
Oxide nanoparticles mixed with high magnetic moment compounds such as CoFe2O4, NiFe2O4, and MnFe2O4 exhibit superior performance compared to other magnetic materials [62,63]. However, these nanoparticles are highly toxic to cells, limiting their use for in vivo, and in vitro biomedical applications [6467]. However, iron oxides such as magnetite (Fe3O4) and maghemite (γ-Fe2O3), in particular, possess high magnetic moments, are relatively safe, and currently in clinical use as MRI contrast agents [5761]. These iron oxide based-magnetic materials are also suitable for biomedical applications. Fe3+ is widely dispersed in the human body so leaching of this metal ion from nanoparticles should not reach toxic concentrations [68,69]. As a result, maghemite is a popular choice for MNPs used biomedical applications. It is very important to modify the surface of MNPs so that they can be used for biomedical applications. Thus, MNPs are coated with compounds such as natural polymers (proteins and carbohydrates) [7075], synthetic organic polymers (polyethylene glycol), polyvinyl alcohol, poly-l-lactic acid) [72,7678], silica [79], and gold [80,81]. These surface coating agents prevent nanoparticle agglomeration, cytotoxicity, and add functionality. MNPs agglomerate readily in aqueous solutions around pH 7 [82], and it is difficult to control the properties and amounts of agglomerated MNPs. The greater toxicity of MNPs compared to those of microparticles can be attributed to their high surface to volume ratio [83]. Coating agents prevent the leaching of potentially toxic components from MNPs. In fact, the cytotoxicity of uncoated NiFeO4 MNPs is dramatically decreased by coating with cationic polymer, polyethylenimine (PEI) [8486]. PEI, a cationic polymer, is widely used for nucleic acid transfection [8789] and also serves as a nanoparticle dispersant [90]. PEI-coated MNPs enhance transfection efficiency [38,41,42,4446,48,49,51,54,55].

3. Cell Transplantation Therapy Using MNPs

Autologous cell transplantation has been widely used in the clinic for decades. Delivering therapeutic genes to patients using their own cells avoids using immunosuppressive drugs. We reasoned, therefore, that a non-viral gene delivery system using iron oxide-based MNPs could provide a powerful tool for next-generation therapies. Gene delivery using MNPs has been successful for delivering nucleic acids into living cells with high efficiency and low cytotoxicity [38,41,42,4446,48,49,51,54,55]. Currently, there are several methods for inducing cellular differentiation.
One of these methods, termed direct reprogramming, or direct conversion, has successfully yielded induced cardiomyocytes, induced neurons, reprogrammed pancreatic β cells, and induced pluripotent stem cells (iPSCs) [9195]. Direct reprogramming represents a more straightforward strategy to treat diseases involving loss of function by specific cell populations compared to approaches requiring an intermediate embryonic stem cell. Thus, patient-derived differentiated cells by gene transfer are suitable for autologous cell transplantation, potentially resulting in faster patient recoveries. The scheme is classified into ex vivo gene therapy. The steps involved in this technique are as follows: (1) Patient-derived cells (such as fibroblasts) are cultured in chemically defined media in vitro; (2) These cells are transfected by MNPs, and differentiated into functional cells; (3) Differentiated cells are isolated by fluorescence-activated cell sorting (FACS); (4) FACS-purified differentiated cells are transplanted into the patient’s target tissue (Figure 2).
Here we briefly describe the magnetofection [96], and our latest study concerning non-viral gene delivery using deacylated polyethylenimine coated MNPs.

4. Gene Delivery Using MNPs and Magnetic Force

The mechanism of magnetofection is similar to using transfection reagents (Lipofectamine 2000, FuGENE HD, and PEI). The only difference is that the plasmids form complexes with cationic polymer-coated MNPs (called as “Magnetoplex”) [42,48,9799] (Figure 3). Figure 3 shows the two difference techniques. The behavior of magnetoplex is readily controlled by magnetic force. Upon binding to the cell surface they are taken up by endocytosis [51,100,101]. Thus, the transfection efficiency was increased.
Many researchers have described magnetofection methods (Table 3). They modified the surface of iron oxide-based MNPs to increase transfection efficiency and reduce cytotoxicity. To achieve this, some investigators selected coating agents such as anionic surfactants (oleic acid, lauroyl sarcosinate) [42,50,102], a non-ionic water-soluble surfactant (Pluronic F-127) [42], fluorinated surfactant (lithium 3-[2-(perfluoroalkyl) ethylthio]propionate) [54], a polymer (polyethylene glycol, poly-l-lysine, poly(propyleneimine) dendrimers) [40,103,104], carbohydrates (Chitosan, Heparan sulfate) [41,47], silica particles (MCM48) [49], proteins (serum albumin, streptavidin) [40,55], hydroxyapatite [105], phospholipids [49,50], a cationic cell penetrating peptide (TAT peptide) [43], non-activated virus envelope (HVJ-E) [47], a transfection reagent (Lipofectamine 2000) [53], and viruses (adenovirus, retrovirus) [44,5456]. These coating agents are often used in conjunction with PEI. PEI is a well-known cationic gene carrier with high transfection efficiency. However, the high toxicity, depended on its molecular weight, has limited its use as a potential gene carrier. Thus, the PEI was modified to increase transfection efficiency, and decrease cytotoxicity [88,106]. To enhance transfection efficiency, most researchers used the PEI, or the modified PEI to coat the nanoparticle surface [38,41,42,4446,48,49,51,54,55,102,107]. PEI-coated MNPs are stable in water, bind nucleic acids, and control MNP behavior by magnetic force. In addition, linear PEI possesses low cytotoxicity compared with branched PEI in vivo and in vitro [108,109] The highest transfection efficiencies have been achieved using 25,000 molecular weight linear PEI [89]. However, PEI cytotoxicity due to its acyl groups has been described [88]. Therefore, our group focused on commercial deacylated PEI (Polyethylenimine “Max” (PEI “Max”), Polysciences Inc.) as an MNP (γ-Fe2O3, d = 70 nm, CIK NanoTek) coating agent.
Deacylated polyethylenimine (linear, 25,000 molecular weight) is built from the same polymer backbone as the popular linear polyethylenimine, and possesses high cationic reactivity. PEI “Max”-coated MNPs (PEI max-MNPs) are stable in deionized water, and positively charged. Thus, PEI max-MNPs electrostatically bind to plasmids. We attempted to introduce the green fluorescent protein (GFP) gene into a mouse embryonic carcinoma cell line, P19CL6 using PEI max-MNPs, and succeeded in establishing a highly efficient and low cytotoxic gene delivery system [107]. Furthermore, we applied this system to human fetal lung-derived fibroblasts (TIG-1 cells) using sixwell plates. Using MNPs, the transfected gene’s expression level increased 2-to 4-fold under optimum conditions (Figure 4, unpublished data). Furthermore, to assess whether the multiple plasmids were expressed in a single cell, we attempt to induce the expression of three fluorescent proteins GFP, cyan fluorescent protein (CFP), and yellow fluorescent protein (YFP). Most cells expressed these three proteins (Figure 5, unpublished data) indicating that gene delivery using MNPs could introduce and allow expression of multiple genes in a single cell.

5. Conclusions

The great promise of gene therapy for treating devastating, incurable diseases has yet to be realized. Less toxic and more efficient systems will be required, and robust research efforts in this regard are currently underway. Rapid advances have been made in adapting nanoparticle technology for basic biomedical and clinical research. Nanoparticles are already being used clinically to enhance MRI imaging, and drug delivery for cancer patients. Our own research has focused on gene delivery systems for autologous cell transplantation therapy, in which the patient’s own cells are transfected with the gene required to correct their condition. In particular, our laboratory and those of others have aimed to optimize magnetofection by developing better nanoparticle coating agents [38,4051,5355]. Nanoparticle size is another important parameter but there were few reports addressing this subject [111]. Since cells endocytose MNPs [51,100,101], MNP size has significant implications for transfection efficiency. PEI-MNPs forms magnetoplex, which increased its influence on the magnetic force. Furthermore, MNP size influences cytotoxicity [112], and more studies on this aspect of MNP technology will be crucial for enhancing transfection efficiencies.
The two research groups reported the important developments in the field of magnetofection. The first is the influence of the oscillating magnetic force on transfection [113,114]. The second is the use of MNP-heating, and -transfection [15,16]. The purpose of these studies have increased the efficiency of transfection, and/or induced a fever by oscillating MNPs for hyperthermia. The latter, a combination of MNP-heating and -transfection, was expected to research the efficacy of both hyperthermia and gene delivery. In the future, the studies of magnetofection using the oscillating MNPs could be developed as a novel methodology.
We found that PEI is an excellent cationic polymer for dispersing MNPs and that its water solubility, stability, and low toxicity contribute to enhancing transfection efficiency [95,115119]. Derivation of iPSCs with the use of non-viral gene delivery using PEI max MNPs should provide a powerful tool for treating diseases such as Alzheimer’s, Huntington’s, and Parkinson’s by autologous cell transplantation. Reprogramming cells requires the action of multiple transcription factors. Our studies demonstrate that MNP-mediated transfection efficiently introduces at least three genes in a single cell. This indicates the feasibility of our system for one-step reprogramming.

References

  1. Chouly, C; Pouliquen, D; Lucet, I; Jeune, JJ; Jallet, P. Development of superparamagnetic nanoparticles for MRI: Effect of particle size, charge and surface nature on biodistribution. J. Microencapsul 1996, 13, 245–255. [Google Scholar]
  2. Schlorf, T; Meincke, M; Kossel, E; Gluer, CC; Jansen, O; Mentlein, R. Biological properties of iron oxide nanoparticles for cellular and molecular magnetic resonance imaging. Int. J. Mol. Sci 2010, 12, 12–23. [Google Scholar]
  3. Yoo, B; Pagel, MD. An overview of responsive MRI contrast agents for molecular imaging. Front. Biosci 2008, 13, 1733–1752. [Google Scholar]
  4. Sun, C; Fang, C; Stephen, Z; Veiseh, O; Hansen, S; Lee, D; Ellenbogen, RG; Olson, J; Zhang, M. Tumor-targeted drug delivery and MRI contrast enhancement by chlorotoxin-conjugated iron oxide nanoparticles. Nanomedicine (Lond. UK) 2008, 3, 495–505. [Google Scholar]
  5. Medarova, Z; Rashkovetsky, L; Pantazopoulos, P; Moore, A. Multiparametric monitoring of tumor response to chemotherapy by noninvasive imaging. Cancer Res 2009, 69, 1182–1189. [Google Scholar]
  6. Wu, YL; Ye, Q; Sato, K; Foley, LM; Hitchens, TK; Ho, C. Noninvasive evaluation of cardiac allograft rejection by cellular and functional cardiac magnetic resonance. JACC Cardiovasc. Imaging 2009, 2, 731–741. [Google Scholar]
  7. Wu, YL; Ye, Q; Foley, LM; Hitchens, TK; Sato, K; Williams, JB; Ho, C. In situ labeling of immune cells with iron oxide particles: An approach to detect organ rejection by cellular MRI. Proc. Natl. Acad. Sci. USA 2006, 103, 1852–1857. [Google Scholar]
  8. Chen, CL; Zhang, H; Ye, Q; Hsieh, WY; Hitchens, TK; Shen, HH; Liu, L; Wu, YJ; Foley, LM; Wang, SJ; et al. A New Nano-sized Iron Oxide Particle with High Sensitivity for Cellular Magnetic Resonance Imaging. Mol Imaging Biol 2010. [Google Scholar] [CrossRef]
  9. Matsumura, Y; Maeda, H. A new concept for macromolecular therapeutics in cancer chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs. Cancer Res 1986, 46, 6387–6392. [Google Scholar]
  10. Maeda, H; Matsumura, Y. Tumoritropic and lymphotropic principles of macromolecular drugs. Crit. Rev. Ther. Drug Carrier Syst 1989, 6, 193–210. [Google Scholar]
  11. Oh, KT; Baik, HJ; Lee, AH; Oh, YT; Youn, YS; Lee, ES. The reversal of drug-resistance in tumors using a drug-carrying nanoparticular system. Int. J. Mol. Sci 2009, 10, 3776–3792. [Google Scholar]
  12. Jordan, A; Scholz, R; Wust, P; Fahling, H; Roland, F. Magnetic fluid hyperthermia (MFH): Cancer treatment with AC magnetic field induced excitation of biocompatible superparamagnetic nanoparticles. J. Magn. Magn. Mater 1999, 201, 413–419. [Google Scholar]
  13. Mornet, S; Vasseur, S; Grasset, F; Veverka, P; Goglio, G; Demourgues, A; Portier, J; Pollert, E; Duguet, E. Magnetic nanoparticle design for medical applications. Prog. Solid State Chem 2006, 34, 237–247. [Google Scholar]
  14. Kim, DH; Kim, KN; Kim, KM; Lee, YK. Targeting to carcinoma cells with chitosan- and starch-coated magnetic nanoparticles for magnetic hyperthermia. J. Biomed. Mater. Res. , Part A 2009, 88, 1–11. [Google Scholar]
  15. Ito, A; Shinkai, M; Honda, H; Kobayashi, T. Heat-inducible TNF-alpha gene therapy combined with hyperthermia using magnetic nanoparticles as a novel tumor-targeted therapy. Cancer Gene Ther 2001, 8, 649–654. [Google Scholar]
  16. Tang, QS; Zhang, DS; Cong, XM; Wan, ML; Jin, LQ. Using thermal energy produced by irradiation of Mn-Zn ferrite magnetic nanoparticles (MZF-NPs) for heat-inducible gene expression. Biomaterials 2008, 29, 2673–2679. [Google Scholar]
  17. Salloum, M; Ma, RH; Weeks, D; Zhu, L. Controlling nanoparticle delivery in magnetic nanoparticle hyperthermia for cancer treatment: Experimental study in agarose gel. Int. J. Hyperthermia 2008, 24, 337–345. [Google Scholar]
  18. Wust, P; Gneveckow, U; Johannsen, M; Bohmer, D; Henkel, T; Kahmann, F; Sehouli, J; Felix, R; Ricke, J; Jordan, A. Magnetic nanoparticles for interstitial thermotherapy—feasibility, tolerance and achieved temperatures. Int. J. Hyperth 2006, 22, 673–685. [Google Scholar]
  19. Ito, A; Honda, H; Kobayashi, T. Cancer immunotherapy based on intracellular hyperthermia using magnetite nanoparticles: A novel concept of “heat-controlled necrosis” with heat shock protein expression. Cancer Immunol. Immunother 2006, 55, 320–328. [Google Scholar]
  20. Tanaka, K; Ito, A; Kobayashi, T; Kawamura, T; Shimada, S; Matsumoto, K; Saida, T; Honda, H. Intratumoral injection of immature dendritic cells enhances antitumor effect of hyperthermia using magnetic nanoparticles. Int. J. Cancer 2005, 116, 624–633. [Google Scholar]
  21. Ito, A; Tanaka, K; Honda, H; Abe, S; Yamaguchi, H; Kobayashi, T. Complete regression of mouse mammary carcinoma with a size greater than 15 mm by frequent repeated hyperthermia using magnetite nanoparticles. J. Biosci. Bioeng 2003, 96, 364–369. [Google Scholar]
  22. Muggia, FM. Doxorubicin-polymer conjugates: Further demonstration of the concept of enhanced permeability and retention. Clin. Cancer Res 1999, 5, 7–8. [Google Scholar]
  23. Gabizon, A; Chemla, M; Tzemach, D; Horowitz, AT; Goren, D. Liposome longevity and stability in circulation: Effects on the in vivo delivery to tumors and therapeutic efficacy of encapsulated anthracyclines. J. Drug Target 1996, 3, 391–398. [Google Scholar]
  24. Sakakibara, T; Chen, FA; Kida, H; Kunieda, K; Cuenca, RE; Martin, FJ; Bankert, RB. Doxorubicin encapsulated in sterically stabilized liposomes is superior to free drug or drug-containing conventional liposomes at suppressing growth and metastases of human lung tumor xenografts. Cancer Res 1996, 56, 3743–3746. [Google Scholar]
  25. Harrington, KJ; Mohammadtaghi, S; Uster, PS; Glass, D; Peters, AM; Vile, RG; Stewart, JS. Effective targeting of solid tumors in patients with locally advanced cancers by radiolabeled pegylated liposomes. Clin. Cancer Res 2001, 7, 243–254. [Google Scholar]
  26. Noe, LL; Becker, RV, III; Gradishar, WJ; Gore, M; Trotter, JP. The cost effectiveness of tamoxifen in the prevention of breast cancer. Am J Manag Care 1999, 5(Suppl 6), S389–406. [Google Scholar]
  27. Ibrahim, NK; Desai, N; Legha, S; Soon-Shiong, P; Theriault, RL; Rivera, E; Esmaeli, B; Ring, SE; Bedikian, A; Hortobagyi, GN; et al. Phase I and pharmacokinetic study of ABI- 007, a Cremophor-free, protein-stabilized, nanoparticle formulation of paclitaxel. Clin. Cancer Res 2002, 8, 1038–1044. [Google Scholar]
  28. Ibrahim, NK; Samuels, B; Page, R; Doval, D; Patel, KM; Rao, SC; Nair, MK; Bhar, P; Desai, N; Hortobagyi, GN. Multicenter phase II trial of ABI-007, an albumin-bound paclitaxel, in women with metastatic breast cancer. J. Clin. Oncol 2005, 23, 6019–6026. [Google Scholar]
  29. Pinder, MC; Ibrahim, NK. Nanoparticle albumin-bound paclitaxel for treatment of metastatic breast cancer. Drugs Today 2006, 42, 599–604. [Google Scholar]
  30. Hamaguchi, T; Kato, K; Yasui, H; Morizane, C; Ikeda, M; Ueno, H; Muro, K; Yamada, Y; Okusaka, T; Shirao, K; et al. A phase I and pharmacokinetic study of NK105, a paclitaxel-incorporating micellar nanoparticle formulation. Br. J. Cancer 2007, 97, 170–176. [Google Scholar]
  31. Hamaguchi, T; Matsumura, Y; Suzuki, M; Shimizu, K; Goda, R; Nakamura, I; Nakatomi, I; Yokoyama, M; Kataoka, K; Kakizoe, T. NK105, a paclitaxel-incorporating micellar nanoparticle formulation, can extend in vivo antitumour activity and reduce the neurotoxicity of paclitaxel. Br. J. Cancer 2005, 92, 1240–1246. [Google Scholar]
  32. Muggia, FM; Hainsworth, JD; Jeffers, S; Miller, P; Groshen, S; Tan, M; Roman, L; Uziely, B; Muderspach, L; Garcia, A; et al. Phase II study of liposomal doxorubicin in refractory ovarian cancer: Antitumor activity and toxicity modification by liposomal encapsulation. J. Clin. Oncol 1997, 15, 987–993. [Google Scholar]
  33. Kikumori, T; Kobayashi, T; Sawaki, M; Imai, T. Anti-cancer effect of hyperthermia on breast cancer by magnetite nanoparticle-loaded anti-HER2 immunoliposomes. Breast Cancer Res. Treat 2009, 113, 435–441. [Google Scholar]
  34. Johannsen, M; Thiesen, B; Wust, P; Jordan, A. Magnetic nanoparticle hyperthermia for prostate cancer. Int. J. Hyperth 2010, 26, 790–795. [Google Scholar]
  35. Rao, W; Deng, ZS; Liu, J. A review of hyperthermia combined with radiotherapy/chemotherapy on malignant tumors. Crit. Rev. Biomed. Eng 2010, 38, 101–116. [Google Scholar]
  36. Yallapu, MM; Othman, SF; Curtis, ET; Gupta, BK; Jaggi, M; Chauhan, SC. Multi-functional magnetic nanoparticles for magnetic resonance imaging and cancer therapy. Biomaterials 2011, 32, 1890–1905. [Google Scholar]
  37. Chen, B; Wu, W; Wang, X. Magnetic iron oxide nanoparticles for tumor-targeted therapy. Curr. Cancer Drug Targets 2011, 11, 184–189. [Google Scholar]
  38. Zhang, H; Lee, MY; Hogg, MG; Dordick, JS; Sharfstein, ST. Gene delivery in three-dimensional cell cultures by superparamagnetic nanoparticles. ACS Nano 2010, 4, 4733–4743. [Google Scholar]
  39. Pickard, MR; Barraud, P; Chari, DM. The transfection of multipotent neural precursor/stem cell transplant populations with magnetic nanoparticles. Biomaterials 2011, 32, 2274–2284. [Google Scholar]
  40. Lee, JH; Lee, K; Moon, SH; Lee, Y; Park, TG; Cheon, J. All-in-one target-cell-specific magnetic nanoparticles for simultaneous molecular imaging and siRNA delivery. Angew. Chem., Int. Ed. Engl 2009, 48, 4174–4179. [Google Scholar]
  41. Kievit, FM; Veiseh, O; Bhattarai, N; Fang, C; Gunn, JW; Lee, D; Ellenbogen, RG; Olson, JM; Zhang, M. PEI-PEG-Chitosan Copolymer Coated Iron Oxide Nanoparticles for Safe Gene Delivery: Synthesis, complexation, and transfection. Adv. Funct. Mater 2009, 19, 2244–2251. [Google Scholar]
  42. Mykhaylyk, O; Antequera, YS; Vlaskou, D; Plank, C. Generation of magnetic nonviral gene transfer agents and magnetofection in vitro. Nat. Protoc 2007, 2, 2391–2411. [Google Scholar]
  43. Song, HP; Yang, JY; Lo, SL; Wang, Y; Fan, WM; Tang, XS; Xue, JM; Wang, S. Gene transfer using self-assembled ternary complexes of cationic magnetic nanoparticles, plasmid DNA and cell-penetrating Tat peptide. Biomaterials 2010, 31, 769–778. [Google Scholar]
  44. Scherer, F; Anton, M; Schillinger, U; Henke, J; Bergemann, C; Kruger, A; Gansbacher, B; Plank, C. Magnetofection: Enhancing and targeting gene delivery by magnetic force in vitro and in vivo. Gene Ther 2002, 9, 102–109. [Google Scholar]
  45. Shi, Y; Zhou, L; Wang, R; Pang, Y; Xiao, W; Li, H; Su, Y; Wang, X; Zhu, B; Zhu, X; Yan, D; Gu, H. In situ preparation of magnetic nonviral gene vectors and magnetofection in vitro. Nanotechnology 2010, 21, 115103. [Google Scholar]
  46. Ang, D; Nguyen, QV; Kayal, S; Preiser, PR; Rawat, RS; Ramanujan, RV. Insights into the mechanism of magnetic particle assisted gene delivery. Acta Biomater 2011, 7, 1319–1326. [Google Scholar]
  47. Morishita, N; Nakagami, H; Morishita, R; Takeda, S; Mishima, F; Terazono, B; Nishijima, S; Kaneda, Y; Tanaka, N. Magnetic nanoparticles with surface modification enhanced gene delivery of HVJ-E vector. Biochem. Biophys. Res. Commun 2005, 334, 1121–1126. [Google Scholar]
  48. Namgung, R; Singha, K; Yu, MK; Jon, S; Kim, YS; Ahn, Y; Park, IK; Kim, WJ. Hybrid superparamagnetic iron oxide nanoparticle-branched polyethylenimine magnetoplexes for gene transfection of vascular endothelial cells. Biomaterials 2010, 31, 4204–4213. [Google Scholar]
  49. Yiu, HH; McBain, SC; Lethbridge, ZA; Lees, MR; Dobson, J. Preparation and characterization of polyethylenimine-coated Fe3O4-MCM-48 nanocomposite particles as a novel agent for magnet-assisted transfection. J. Biomed. Mater. Res. , Part A 2010, 92, 386–392. [Google Scholar]
  50. Namiki, Y; Namiki, T; Yoshida, H; Ishii, Y; Tsubota, A; Koido, S; Nariai, K; Mitsunaga, M; Yanagisawa, S; Kashiwagi, H; et al. A novel magnetic crystal-lipid nanostructure for magnetically guided in vivo gene delivery. Nat. Nanotechnol 2009, 4, 598–606. [Google Scholar]
  51. Arsianti, M; Lim, M; Marquis, CP; Amal, R. Polyethylenimine based magnetic iron-oxide vector: The effect of vector component assembly on cellular entry mechanism, intracellular localization, and cellular viability. Biomacromolecules 2010, 11, 2521–3251. [Google Scholar]
  52. Kim, TS; Lee, SH; Gang, GT; Lee, YS; Kim, SU; Koo, DB; Shin, MY; Park, CK; Lee, DS. Exogenous DNA uptake of boar spermatozoa by a magnetic nanoparticle vector system. Reprod. Domest. Anim 2009, 45, e201–e206. [Google Scholar]
  53. Yang, SY; Sun, JS; Liu, CH; Tsuang, YH; Chen, LT; Hong, CY; Yang, HC; Horng, HE. Ex vivo magnetofection with magnetic nanoparticles: A novel platform for nonviral tissue engineering. Artif. Organs 2008, 32, 195–204. [Google Scholar]
  54. Tresilwised, N; Pithayanukul, P; Mykhaylyk, O; Holm, PS; Holzmuller, R; Anton, M; Thalhammer, S; Adiguzel, D; Doblinger, M; Plank, C. Boosting oncolytic adenovirus potency with magnetic nanoparticles and magnetic force. Mol. Pharmaceutics 2010, 7, 1069–1089. [Google Scholar]
  55. Hashimoto, M; Hisano, Y. Directional gene-transfer into the brain by an adenoviral vector tagged with magnetic nanoparticles. J. Neurosci. Methods 2011, 194, 316–320. [Google Scholar]
  56. Mah, C; Fraites, TJ, Jr; Zolotukhin, I; Song, S; Flotte, TR; Dobson, J; Batich, C; Byrne, BJ. Improved method of recombinant AAV2 delivery for systemic targeted gene therapy. Mol. Ther 2002, 6, 106–112. [Google Scholar]
  57. Basti, H; Ben Tahar, L; Smiri, LS; Herbst, F; Vaulay, MJ; Chau, F; Ammar, S; Benderbous, S. Catechol derivatives-coated Fe3O4 and gamma-Fe2O3 nanoparticles as potential MRI contrast agents. J. Colloid Interface Sci 2010, 341, 248–254. [Google Scholar]
  58. Gamarra, LF; Amaro, E, Jr; Alves, S; Soga, D; Pontuschka, WM; Mamani, JB; Carneiro, SM; Brito, GE; Figueiredo Neto, AM. Characterization of the biocompatible magnetic colloid on the basis of Fe3O4 nanoparticles coated with dextran used as contrast agent in magnetic resonance imaging. J. Nanosci. Nanotechnol 2010, 10, 4145–4153. [Google Scholar]
  59. Jung, CW; Jacobs, P. Physical and chemical properties of superparamagnetic iron oxide MR contrast agents: Ferumoxides, ferumoxtran, ferumoxsil. Magn. Reson. Imaging 1995, 13, 661–674. [Google Scholar]
  60. Martina, MS; Fortin, JP; Menager, C; Clement, O; Barratt, G; Grabielle-Madelmont, C; Gazeau, F; Cabuil, V; Lesieur, S. Generation of superparamagnetic liposomes revealed as highly efficient MRI contrast agents for in vivo imaging. J. Am. Chem. Soc 2005, 127, 10676–10685. [Google Scholar]
  61. Widder, DJ; Greif, WL; Widder, KJ; Edelman, RR; Brady, TJ. Magnetite albumin microspheres: A new MR contrast material. Am. J. Roentgenol 1987, 148, 399–404. [Google Scholar]
  62. Sun, X; Gutierrez, A; Yacaman, MJ; Dong, X; Jin, S. Investigations on magnetic properties and structure for carbon encapsulated nanoparticles of Fe, Co, Ni. Mater. Sci. Eng. A 2000, 286, 157–160. [Google Scholar]
  63. Tomitaka, A; Kobayashi, H; Yamada, T; Jeun, M; Bae, S; Takemura, Y. Magnetization and self-heating temperature of NiFe2O4 nanoparticles measured by applying ac magnetic field. J. Phys.: Conf. Ser 2010, 200, 122010. [Google Scholar]
  64. Cho, WS; Duffin, R; Poland, CA; Duschl, A; Oostingh, GJ; Macnee, W; Bradley, M; Megson, IL; Donaldson, K. Differential pro-inflammatory effects of metal oxide nanoparticles and their soluble ions in vitro and in vivo; zinc and copper nanoparticles, but not their ions, recruit eosinophils to the lungs. Nanotoxicology 2011, in press. [Google Scholar]
  65. George, S; Xia, T; Rallo, R; Zhao, Y; Ji, Z; Lin, S; Wang, X; Zhang, H; France, B; Schoenfeld, D; et al. Use of a high-throughput screening approach coupled with in vivo zebrafish embryo screening to develop hazard ranking for engineered nanomaterials. ACS Nano 2011, 5, 1805–1817. [Google Scholar]
  66. Giri, J; Pradhan, P; Somani, V; Chelawat, H; Chhatre, S; Banerjee, R; Bahadur, D. Synthesis and characterizations of water-based ferrofluids of substituted ferrites [Fe1-xBxFe2O4, B = Mn, Co (x = 0−1)] for biomedical applications. J. Magn. Magn. Mater 2008, 320, 724–730. [Google Scholar]
  67. Karlsson, HL; Cronholm, P; Gustafsson, J; Moller, L. Copper oxide nanoparticles are highly toxic: A comparison between metal oxide nanoparticles and carbon nanotubes. Chem. Res. Toxicol 2008, 21, 1726–1732. [Google Scholar]
  68. McBain, SC; Yiu, HH; Dobson, J. Magnetic nanoparticles for gene and drug delivery. Int. J. Nanomed 2008, 3, 169–180. [Google Scholar]
  69. Buyukhatipoglu, K; Clyne, AM. Superparamagnetic iron oxide nanoparticles change endothelial cell morphology and mechanics via reactive oxygen species formation. J. Biomed. Mater. Res. , Part A 2011, 96, 186–195. [Google Scholar]
  70. Schroder, U; Segren, S; Gemmefors, C; Hedlund, G; Jansson, B; Sjogren, HO; Borrebaeck, CA. Magnetic carbohydrate nanoparticles for affinity cell separation. J. Immunol. Methods 1986, 93, 45–53. [Google Scholar]
  71. Berry, CC; Wells, S; Charles, S; Curtis, AS. Dextran and albumin derivatised iron oxide nanoparticles: Influence on fibroblasts in vitro. Biomaterials 2003, 24, 4551–7455. [Google Scholar]
  72. Nitin, N; LaConte, LE; Zurkiya, O; Hu, X; Bao, G. Functionalization and peptide-based delivery of magnetic nanoparticles as an intracellular MRI contrast agent. J. Biol. Inorg. Chem 2004, 9, 706–712. [Google Scholar]
  73. Ito, A; Ino, K; Kobayashi, T; Honda, H. The effect of RGD peptide-conjugated magnetite cationic liposomes on cell growth and cell sheet harvesting. Biomaterials 2005, 26, 6185–6193. [Google Scholar]
  74. de la Fuente, JM; Penades, S. Glyconanoparticles: Types, synthesis and applications in glycoscience, biomedicine and material science. Biochim. Biophys. Acta 2006, 1760, 636–651. [Google Scholar]
  75. McDonald, MA; Watkin, KL. Investigations into the physicochemical properties of dextran small particulate gadolinium oxide nanoparticles. Acad. Radiol 2006, 13, 421–427. [Google Scholar]
  76. Mertz, CJ; Kaminski, MD; Xie, Y; Finck, MR; Guy, S; Rosengart, AJ. In vitro studies of functionalized magnetic nanospheres for selective removal of a simulant biotoxin. J. Magn. Magn. Mater 2005, 293, 572–577. [Google Scholar]
  77. Mikhaylova, M; Jo, Y; Kim, D; Bobrysheva, N; Andersson, Y; Eriksson, T; Osmolowsky, M; Semenov, V; Muhammed, M. The Effect of Biocompatible Coating Layers on Magnetic Properties of Superparamagnetic Iron Oxide Nanoparticles. Hyperfine Interact 2004, 156–157, 257–263. [Google Scholar]
  78. Qiu, XP; Winnik, F. Preparation and characterization of PVA coated magnetic nanoparticles. Chin. J. Polym. Sci 2000, 18, 535–539. [Google Scholar]
  79. Yiu, HHP; Wright, PA; Botting, NP. Enzyme immobilisation using SBA-15 mesoporous molecular sieves with functionalised surfaces. J. Mol. Catal. B: Enzym 2001, 15, 81–92. [Google Scholar]
  80. Ameur, S; Martelet, C; Jaffrezic-Renault, N; Chovelon, J-M. Sensitive immunodetection through impedance measurements onto gold functionalized electrodes. Appl. Biochem. Biotechnol 2000, 89, 161–170. [Google Scholar]
  81. Arsianti, M; Lim, M; Lou, SN; Goon, IY; Marquis, CP; Amal, R. Bi-functional gold-coated magnetite composites with improved biocompatibility. J. Colloid Interface Sci 2011, 354, 536–545. [Google Scholar]
  82. Williams, D; Gold, K; Holoman, T; Ehrman, S; Wilson, O. Surface modification of magnetic nanoparticles using gum arabic. J. Nanopart. Res 2006, 8, 749–753. [Google Scholar]
  83. Klabunde, KJ; Stark, J; Koper, O; Mohs, C; Park, DG; Decker, S; Jiang, Y; Lagadic, I; Zhang, D. Nanocrystals as stoichiometric reagents with unique surface chemistry. J. Phys. Chem 1996, 100, 12142–12153. [Google Scholar]
  84. Zhang, H; Xia, T; Meng, H; Xue, M; George, S; Ji, Z; Wang, X; Liu, R; Wang, M; France, B; et al. Differential expression of syndecan-1 mediates cationic nanoparticle toxicity in undifferentiated versus differentiated normal human bronchial epithelial cells. ACS Nano 2011, 5, 2756–2769. [Google Scholar]
  85. Sunoqrot, S; Bae, JW; Jin, SE; Ryan, MP; Liu, Y; Hong, S. Kinetically controlled cellular interactions of polymer-polymer and polymer-liposome nanohybrid systems. Bioconjugate Chem 2011, 22, 466–474. [Google Scholar]
  86. Schweiger, C; Pietzonka, C; Heverhagen, J; Kissel, T. Novel magnetic iron oxide nanoparticles coated with poly(ethylene imine)-g-poly(ethylene glycol) for potential biomedical application: Synthesis, stability, cytotoxicity and MR imaging. Int. J. Pharm 2011, 408, 130–137. [Google Scholar]
  87. Boussif, O; Lezoualc’h, F; Zanta, MA; Mergny, MD; Scherman, D; Demeneix, B; Behr, JP. A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: Polyethylenimine. Proc. Natl. Acad. Sci. USA 1995, 92, 7297–7301. [Google Scholar]
  88. Thomas, M; Lu, JJ; Ge, Q; Zhang, C; Chen, J; Klibanov, AM. Full deacylation of polyethylenimine dramatically boosts its gene delivery efficiency and specificity to mouse lung. Proc. Natl. Acad. Sci. USA 2005, 102, 5679–5684. [Google Scholar]
  89. Abdallah, B; Hassan, A; Benoist, C; Goula, D; Behr, JP; Demeneix, BA. A powerful nonviral vector for in vivo gene transfer into the adult mammalian brain: Polyethylenimine. Hum. Gene Ther 1996, 7, 1947–1954. [Google Scholar]
  90. Zuo, KH; Jiang, DL; Zhang, JX; Lin, QL. Forming nanometer TiO2 sheets by nonaqueous tape casting. Ceram. Int 2007, 33, 477–481. [Google Scholar]
  91. Ieda, M; Fu, JD; Delgado-Olguin, P; Vedantham, V; Hayashi, Y; Bruneau, BG; Srivastava, D. Direct reprogramming of fibroblasts into functional cardiomyocytes by defined factors. Cell 2010, 142, 375–386. [Google Scholar]
  92. Takeuchi, JK; Bruneau, BG. Directed transdifferentiation of mouse mesoderm to heart tissue by defined factors. Nature 2009, 459, 708–711. [Google Scholar]
  93. Vierbuchen, T; Ostermeier, A; Pang, ZP; Kokubu, Y; Sudhof, TC; Wernig, M. Direct conversion of fibroblasts to functional neurons by defined factors. Nature 2010, 463, 1035–1041. [Google Scholar]
  94. Zhou, Q; Brown, J; Kanarek, A; Rajagopal, J; Melton, DA. In vivo reprogramming of adult pancreatic exocrine cells to beta-cells. Nature 2008, 455, 627–632. [Google Scholar]
  95. Takahashi, K; Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar]
  96. Laurent, N; Sapet, CD; Le Gourrierec, L; Bertosio, E; Zelphati, O. Nucleic acid delivery using magnetic nanoparticles: The Magnetofection™ technology. Ther. Deliv 2011, 2, 471–482. [Google Scholar]
  97. Sonawane, ND; Szoka, FC, Jr; Verkman, AS. Chloride accumulation swelling in endosomes enhances DNA transfer by polyamine-DNA polyplexes. J. Biol. Chem 2003, 278, 44826–44831. [Google Scholar]
  98. Brunner, S; Sauer, T; Carotta, S; Cotten, M; Saltik, M; Wagner, E. Cell cycle dependence of gene transfer by lipoplex, polyplex and recombinant adenovirus. Gene Ther 2000, 7, 401–407. [Google Scholar]
  99. Nishiyama, N; Kataoka, K. Current state, achievements, and future prospects of polymeric micelles as nanocarriers for drug and gene delivery. Pharmacol. Ther 2006, 112, 630–648. [Google Scholar]
  100. Yu, JH; Quan, JS; Huang, J; Nah, JW; Cho, CS. Degradable poly(amino ester) based on poly(ethylene glycol) dimethacrylate and polyethylenimine as a gene carrier: Molecular weight of PEI affects transfection efficiency. J. Mater. Sci.: Mater. Med 2009, 20, 2501–2510. [Google Scholar]
  101. Veiseh, O; Kievit, FM; Gunn, JW; Ratner, BD; Zhang, M. A ligand-mediated nanovector for targeted gene delivery and transfection in cancer cells. Biomaterials 2009, 30, 649–657. [Google Scholar]
  102. Pan, X; Guan, J; Yoo, JW; Epstein, AJ; Lee, LJ; Lee, RJ. Cationic lipid-coated magnetic nanoparticles associated with transferrin for gene delivery. Int. J. Pharm 2008, 358, 263–270. [Google Scholar]
  103. Li, Z; Xiang, J; Zhang, W; Fan, S; Wu, M; Li, X; Li, G. Nanoparticle delivery of anti-metastatic NM23-H1 gene improves chemotherapy in a mouse tumor model. Cancer Gene Ther 2009, 16, 423–429. [Google Scholar]
  104. Gonzalez, B; Ruiz-Hernandez, E; Feito, MJ; Lopez de Laorden, C; Arcos, D; Ramirez-Santillan, C; Matesanz, C; Portoles, MT; Vallet-Regi, M. Covalently bonded dendrimer-maghemite nanosystems: Nonviral vectors for in vitro gene magnetofection. J. Mater. Chem 2011, 21, 4598–4604. [Google Scholar]
  105. Wu, H-C; Wang, T-W; Bohn, MC; Lin, F-H; Spector, M. Novel magnetic hydroxyapatite nanoparticles as non-viral vectors for the glial cell line-derived neurotrophic factor Gene. Adv. Funct. Mater 2010, 20, 67–77. [Google Scholar]
  106. Zhao, D; Gong, T; Zhu, D; Zhang, Z; Sun, X. Comprehensive comparison of two new biodegradable gene carriers. Int J Pharm 2011, in press. [Google Scholar]
  107. Kami, D; Takeda, S; Makino, H; Toyoda, M; Itakura, Y; Gojo, S; Kyo, S; Umezawa, A; Watanabe, M. Efficient transfection method using deacylated polyethylenimine-coated magnetic nanoparticles. J Artif Organs 2011, in press. [Google Scholar]
  108. Moghimi, SM; Symonds, P; Murray, JC; Hunter, AC; Debska, G; Szewczyk, A. A twostage poly(ethylenimine)-mediated cytotoxicity: Implications for gene transfer/therapy. Mol. Ther 2005, 11, 990–995. [Google Scholar]
  109. Zou, SM; Erbacher, P; Remy, JS; Behr, JP. Systemic linear polyethylenimine (L-PEI)-mediated gene delivery in the mouse. J. Gene Med 2000, 2, 128–134. [Google Scholar]
  110. Biswas, S; Gordon, LE; Clark, GJ; Nantz, MH. Click assembly of magnetic nanovectors for gene delivery. Biomaterials 2011, 32, 2683–2688. [Google Scholar]
  111. Ahmed, M; Deng, Z; Narain, R. Study of transfection efficiencies of cationic glyconanoparticles of different sizes in human cell line. ACS Appl. Mater. Interfaces 2009, 1, 1980–1987. [Google Scholar]
  112. Frohlich, E; Kueznik, T; Samberger, C; Roblegg, E; Wrighton, C; Pieber, TR. Sizedependent effects of nanoparticles on the activity of cytochrome P450 isoenzymes. Toxicol. Appl. Pharmacol 2010, 242, 326–332. [Google Scholar]
  113. McBain, SC; Griesenbach, U; Xenariou, S; Keramane, A; Batich, CD; Alton, EWFW; Dobson, J. Magnetic nanoparticles as Gene delivery agents: Enhanced transfection in the presence of oscillating magnet arrays. Nanotechnology 2008, 19, 405102. [Google Scholar]
  114. Kamau, SW; Hassa, PO; Steitz, B; Petri-Fink, A; Hofmann, H; Hofmann-Amtenbrink, M; von Rechenberg, B; Hottiger, MO. Enhancement of the efficiency of non-viral gene delivery by application of pulsed magnetic field. Nucleic Acids Res 2006, 34, e40. [Google Scholar]
  115. Dimos, JT; Rodolfa, KT; Niakan, KK; Weisenthal, LM; Mitsumoto, H; Chung, W; Croft, GF; Saphier, G; Leibel, R; Goland, R; et al. Induced pluripotent stem cells generated from patients with ALS can be differentiated into motor neurons. Science 2008, 321, 1218–1221. [Google Scholar]
  116. Ebert, AD; Yu, J; Rose, FF, Jr; Mattis, VB; Lorson, CL; Thomson, JA; Svendsen, CN. Induced pluripotent stem cells from a spinal muscular atrophy patient. Nature 2009, 457, 277–280. [Google Scholar]
  117. Okita, K; Ichisaka, T; Yamanaka, S. Generation of germline-competent induced pluripotent stem cells. Nature 2007, 448, 313–317. [Google Scholar]
  118. Okita, K; Nakagawa, M; Hyenjong, H; Ichisaka, T; Yamanaka, S. Generation of mouse induced pluripotent stem cells without viral vectors. Science 2008, 322, 949–953. [Google Scholar]
  119. Takahashi, K; Tanabe, K; Ohnuki, M; Narita, M; Ichisaka, T; Tomoda, K; Yamanaka, S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 2007, 131, 861–782. [Google Scholar]
Figure 1. MNP gene delivery system (Magnetofection). Plasmids are bound to MNPs, which then move from the media to the cell surface by applying a magnetic force.
Figure 1. MNP gene delivery system (Magnetofection). Plasmids are bound to MNPs, which then move from the media to the cell surface by applying a magnetic force.
Ijms 12 03705f1
Figure 2. Strategy for cell transplantation therapy. A patient’s cells are cultured in chemically defined media. MNP-transfected cells by the introduced gene are isolated by FACS. FACS-purified differentiated cells are transplanted into the patient.
Figure 2. Strategy for cell transplantation therapy. A patient’s cells are cultured in chemically defined media. MNP-transfected cells by the introduced gene are isolated by FACS. FACS-purified differentiated cells are transplanted into the patient.
Ijms 12 03705f2
Figure 3. Gene delivery systems using a transfection reagent (cationic polymer) and MNPs: (A) Gene delivery system using transfection reagent. The polyplex moves randomly in culture medium; (B) Magnetofection system. The magnetoplex only moves to the cell surface.
Figure 3. Gene delivery systems using a transfection reagent (cationic polymer) and MNPs: (A) Gene delivery system using transfection reagent. The polyplex moves randomly in culture medium; (B) Magnetofection system. The magnetoplex only moves to the cell surface.
Ijms 12 03705f3
Figure 4. Optimum conditions for PEI max-MNPs magnetofection. To optimize conditions, we varied volume (A) and time on the magnetic plate (B). These results were evaluated by quantitative real-time RT-PCR. The relative expression level (GFP/GAPDH) in the human fetal lung-derived fibroblasts (TIG-1 cells) treated with PEI max alone (A), and in the absence of magnetic force (0 h) (B) was defined as 1. Optimal transfection conditions were established when TIG-1 cells were treated with 0.8 μg PEI max-MNPs and 2.0 μg pCAG-GFP for 8 h on the magnetic plate in either a six-well plate or a 35 mm dish. The asterisk (*) indicates a significant difference (P < 0.05).
Figure 4. Optimum conditions for PEI max-MNPs magnetofection. To optimize conditions, we varied volume (A) and time on the magnetic plate (B). These results were evaluated by quantitative real-time RT-PCR. The relative expression level (GFP/GAPDH) in the human fetal lung-derived fibroblasts (TIG-1 cells) treated with PEI max alone (A), and in the absence of magnetic force (0 h) (B) was defined as 1. Optimal transfection conditions were established when TIG-1 cells were treated with 0.8 μg PEI max-MNPs and 2.0 μg pCAG-GFP for 8 h on the magnetic plate in either a six-well plate or a 35 mm dish. The asterisk (*) indicates a significant difference (P < 0.05).
Ijms 12 03705f4
Figure 5. Transfection of TIG-1 cells with multiple genes using PEI max-MNPs. TIG-1 cells were simultaneously transfected with GFP, CFP, and YFP expression vector plasmids. TIG-1 cells were treated with 0.8 μg of PEI max-MNPs and 0.7 μg each of pCAG-GFP (GFP, provided by Dr. Nishino), pPhi-Yellow-N (YFP, Evrogen), and pAmCyan1-C1 (CFP, Clonetech) for 8 h on the magnetic plate in a six-well plate or a 35 mm dish. White bar indicates 200 μm.
Figure 5. Transfection of TIG-1 cells with multiple genes using PEI max-MNPs. TIG-1 cells were simultaneously transfected with GFP, CFP, and YFP expression vector plasmids. TIG-1 cells were treated with 0.8 μg of PEI max-MNPs and 0.7 μg each of pCAG-GFP (GFP, provided by Dr. Nishino), pPhi-Yellow-N (YFP, Evrogen), and pAmCyan1-C1 (CFP, Clonetech) for 8 h on the magnetic plate in a six-well plate or a 35 mm dish. White bar indicates 200 μm.
Ijms 12 03705f5
Table 1. Biomedical Applications of Magnetic Nanoparticles (MNPs).
Table 1. Biomedical Applications of Magnetic Nanoparticles (MNPs).
PurposeReferences
MRIDiagnosis[18,5761]
DDSAnti-cancer therapy, Enzyme therapy[911,2231]
HyperthermiaAnti-cancer therapy[1218,3337]
Gene DeliveryAnti-cancer therapy, Cell transplantation therapy[3855]
Table 2. Gene delivery systems.
Table 2. Gene delivery systems.
Expression TypeEfficiency (%)Cell Viability (%)Safety
Virus*Stable, or Transient80–90%80–90%Low
ElectroporationTransient50–70%40–50%High
TF reagent **Transient20–30%80–90%High
*Virus including adenovirus (transient), retrovirus (stable), and lentivirus (stable);
**TF reagent, transfection reagents including PEI (Polysciences Inc.), FuGENE HD (Promega), and Lipofectamine 2000 (Invitrogen);
All values are ours (unpublished experiments).
Table 3. Summary of magnetofection literature.
Table 3. Summary of magnetofection literature.
AuthorYearVectorMagnetic NanoparticlesModifying AgentTargeting Cell, or TissueTF EfficiencyCell Viability (% of Control)Reference
Kami D2011PlasmidIron oxide (γ-Fe2O3)PEI max (MW: 25 k)P19CL6* 82%100%[107]

Pickard MR2011PlasmidNeuroMag-Neural precursor cell* 30%70%[39]

Hashimoto M2011Adenovirus, BiotinSPIONPEI, StreptoavidinHeLa** 4-fold-[55]
Adenovirus, BiotinSPIONPEI, StreptoavidinNIH3T3** 10-fold-
Adenovirus, BiotinSPIONPEI, StreptoavidinMouse embryonic brain--

Biswas S2011PlasmidIron oxide (Fe3O4)Aminooxy, Oxime etherMCF-7** 1425-fold89%[110]

B González2011PlasmidSPIONPoly(propyleneimine) dendrimersSaos-2 osteoblasts* 12%75%[104]

Zhang H2010PlasmidSPIONBranch PEI (MW: 25 k)NIT3T3* 64%100%[38]
siRNASPIONBranch PEI (MW: 25 k)NIT3T3* 77%100%

Song HP2010PlasmidPolyMagTat peptideU251* 60%80%[43]
PlasmidPolyMagTat peptideRat spinal cord** 2-fold-

Arsianti M2010PlasmidIron oxideBranch PEI (MW: 25 k)BHK-21-60–90%[51]

Shi Y2010PlasmidMagnetiteHyperbranch PEI (MW: 10 k)COS-7** 13-fold-[45]

Ang D2010PlasmidMagnetiteBranch PEI (MW: 25 k)COS-7** 6-fold70%[46]

Tresilwised N2010AdenovirusIron oxide (Fe2O3, Fe3O4)Branch PEI (MW: 25 k), Zonyl FSA fluorosurfactantEPP85-181RDB** 10-fold-[54]

Namgung R2010PlasmidSPIONPEG, Branch PEI (MW: 25 k)HUVEC** 12-fold80%[48]

Yiu HH2010PlasmidIron oxide (Fe3O4)PEI (MW: 25 k), MCM48 (Silica particle)NCI-H292** 4-fold-[49]

HC Wu2010PlasmidMagnetiteHydroxyapatiteRat marrow stromal cells* 60–70%100%[105]

Namiki Y2009PlasmidMagnetiteOleic acid, PhospholipidHSC45** 8-fold-[50]
siRNAMagnetiteOleic acid, PhospholipidTissue sample from gastric cancer--

Kim TS2009PlasmidPolyMag-Boar spermatozoa--[52]

Kievit FM2009PlasmidSPIONPEI (MW: 25 k)C6* 90%10%[41]
PlasmidSPIONPEI (MW: 25 k), ChitosanC6* 45%100%
PlasmidPolyMag-C6* 32%66%

Lee JH2009siRNAMnMEIOSerum albumin, PEG-RGDMDA-MB-435-GFP* 30%-[40]

Li Z2009PlasmidIron oxidePoly-l-lysineLung tissue*** 60%-[103]

Yang SY2008PlasmidIron oxide (Fe3O4)Lipofectamine 2000He99--[53]
PlasmidIron oxide (Fe3O4)DOTAP:DOPEHe99--

Pan X2008PlasmidMagnetiteOleic acid, Branch PEI (MW: 25 k), TransferrinKB** 300-fold92%[102]

Mykhaylyk O2007PlasmidIron oxide (Fe2O3, Fe3O4)Branch PEI (MW: 25 k)H441* 49%-[42]
PlasmidIron oxide (Fe2O3, Fe3O4)Pluronic F-127H441* 37%-
PlasmidIron oxide (Fe2O3, Fe3O4)Lauroyl sarcosinateH441--
PlasmidIron oxide (Fe2O3, Fe3O4)Branch PEI (MW: 25 k), Lauroyl sarcosinateH441--

Morishita N2005PlasmidIron oxide (γFe2O3)HVJ-E, protamine sulfateBHK-21** 4-fold-[47]
PlasmidIron oxide (γ-Fe2O3)HVJ-E, heparin sulfateLiver, BALB/c mice (8 weeks age)** 3-fold-

Scherer F2002PlasmidSPIONPEI (MW: 800 k)NIH3T3** 5-fold-[44]
AdenovirusSPIONPEI (MW: 800 k)K562** 100-fold-
RetrovirusSPIONPEI (MW: 800 k)NIH3T3* 20%-

Mah C2002AdenovirusAvidinylated magnetiteBiotunylated heparan sulfateC12S* 75%-[56]
AdenovirusAvidinylated magnetiteBiotunylated heparan sulfateAdult 129/SvJ mice--
*indicates % of fluorescent positive cells analyzed by flow cytometric analysis.
**indicates analysis by luciferase activity assay compared with control. Transfection efficiency was indicated optimal transfection condition.
***indicates transfection without magnetic force.
PEI: Polyethylenimine; PEI max: Deacaylated PEI; MNP: Magnetic nanoparticle; SPION: Superparamagnetic iron oxide nanoparticle; MW: Molecular weight; TF: transfection; PolyMag: Commercial Magnetofection reagent), NeuroMag (Commercial Magnetofection reagent); HVJ-E: hemagglutinating virus of Japan-envelope; DOTAP: 1,2-dioleoyl- 3-trimethylammonium-propane; DOPE: 1,2-dioleoyl-3-sn- phosphatidyl-ethanolamine; Tat peptide: cationic cell penetrating peptide; MeMEIO: Manganese-doped magnetism-engineered iron oxide; PEG: polyethylene glycol, Zonyl FSA fluorosurfactant: Lithium 3-[2-(perfluoroalkyl)ethylthio]propionate).

Share and Cite

MDPI and ACS Style

Kami, D.; Takeda, S.; Itakura, Y.; Gojo, S.; Watanabe, M.; Toyoda, M. Application of Magnetic Nanoparticles to Gene Delivery. Int. J. Mol. Sci. 2011, 12, 3705-3722. https://doi.org/10.3390/ijms12063705

AMA Style

Kami D, Takeda S, Itakura Y, Gojo S, Watanabe M, Toyoda M. Application of Magnetic Nanoparticles to Gene Delivery. International Journal of Molecular Sciences. 2011; 12(6):3705-3722. https://doi.org/10.3390/ijms12063705

Chicago/Turabian Style

Kami, Daisuke, Shogo Takeda, Yoko Itakura, Satoshi Gojo, Masatoshi Watanabe, and Masashi Toyoda. 2011. "Application of Magnetic Nanoparticles to Gene Delivery" International Journal of Molecular Sciences 12, no. 6: 3705-3722. https://doi.org/10.3390/ijms12063705

Article Metrics

Back to TopTop