Gene delivery to adult neural stem cells

Share Embed


Descrição do Produto

Experimental Cell Research 279, 34 –39 (2002) doi:10.1006/excr.2002.5569

Gene Delivery to Adult Neural Stem Cells Anna Falk,* Niklas Holmstro¨m,* ,1 Marie Carle´n,* ,1 Robert Cassidy,* Cecilia Lundberg,† and Jonas Frise´n* ,2 *Department of Cell and Molecular Biology, Medical Nobel Institute, Karolinska Institute, SE-171 77 Stockholm, Sweden; and †Wallenberg Neuroscience Center, Division of Neurobiology, Lund University, SE-221 84 Lund, Sweden

have unveiled single genes that drive neural stem cells to specific fates [3], suggesting that neurons for transplantation could be generated by expressing certain genes in neural stem cells. A conceptually different way to replace lost neurons would be to stimulate endogenous stem cells to produce more neurons of a desired type, and understanding the regulation of this process may elucidate ways to influence this process by expressing desired genes in stem cells. Another situation in which it may be very attractive to express genes in neural stem cells is for gene therapy in the central nervous system. There are a number of neurological diseases caused by defined genetic defects, such as lysosomal storage diseases, and stem cells may be ideal targets for gene therapy. Much effort has been focused on developing methods to transduce stem cells, not least in the hematopoietic system [4, 5], but this has proved difficult. We therefore sought to evaluate the efficacy of different gene delivery methods, both viral and nonviral, for introducing genes into adult neural stem cells in vitro and in vivo.

Neural stem cells may present an ideal route for gene therapy as well as offer new possibilities for the replacement of neurons lost to injury or disease. However, it has proved difficult to express ectopic genes in stem cells. We report methods to introduce genes into adult neural stem cells using viral and nonviral vectors in vitro and in vivo. Adenoviral and VSV-Gpseudotyped retroviral vectors are more efficient than plasmid transfection or VSV-G lentiviral transduction in vitro. We further show that adult neural stem cells can be directed to a neuronal fate by ectopic expression of neurogenin 2 in vitro. Plasmids can be delivered in vivo when complexed with linear polyethyleneimine, and gene expression can be targeted specifically to neural stem or progenitor cells by the use of specific promoters. These techniques may be utilized both to study the function of various genes in the differentiation of neural stem cells to specific cell fates and, ultimately, for gene therapy or to generate specific differentiated progeny for cell transplantation. © 2002 Elsevier Science (USA)

INTRODUCTION

MATERIALS AND METHODS

The discovery of neurogenesis from stem cells in the adult brain has raised hope for the development of new regenerative therapies for neurological diseases [1]. Transplantation of neurons from aborted human fetuses to patients with Parkinson’s disease has demonstrated that cell therapy can be an effective treatment for neurodegenerative disease [2]. However, the supply of human fetal tissue is highly limited and its use ethically controversial, making adult neural stem cells an attractive alternative source of neurons for transplantation. Although neural stem cells can generate therapeutically relevant types of neurons, there is currently no way to efficiently derive a homogenous population of a specific neuronal type from stem cells. Insights into the molecular regulation of nervous system development

Neural stem cell cultures. The lateral walls of the lateral ventricles of adult male C57/bl mice (Charles River) were dissected and enzymatically dissociated in 0.5 ␮g/␮l trypsin (Sigma) and 0.7 ␮g/␮l hyaluronidase (Sigma) at 37°C for 30 min. The cells were then processed as described [6] and cultured in DMEM/F12 supplemented with B27 (Life Technologies), 20 ng/ml EGF (BD Bioscience), 100 U/ml penicillin, and 100 ␮g/ml streptomycin (Life Technologies). In vitro plasmid transfection. We tested four different cationic lipids, each at six different lipid/DNA concentrations (Table 1). Lipid was incubated with culture medium without supplements for 30 min at room temperature before pEGFP-N1 plasmid DNA (Clontech) was added to the lipid solution and incubated for 15 min at room temperature. A total of 10,000 cells were added to each well in a final volume of 300 ␮l/well in a 24-well plate. The cells were from primary neurospheres which had been passaged the day before. After incubation for 4 h at 37°C, 0.6 ml prewarmed culture medium with 1.5⫻ B27, penicillin, and streptomycin was added to each well. To induce differentiation, neurospheres were transferred to polyornithinecoated glass coverslips 24 h after transfection. After an additional 24 h of culture, the cells were fixed in 4% formaldehyde for 10 min at room temperature and stained with 1.5 ␮g/ml DAPI (Sigma). Three independent experiments were conducted and the average percentage of cells expressing eGFP was calculated.

1

These authors contributed equally to this work. To whom correspondence and reprint requests should be addressed. Fax: ⫹46 8 324927. E-mail: [email protected]. 2

0014-4827/02 $35.00 © 2002 Elsevier Science (USA) All rights reserved.

34

35

GENE DELIVERY TO ADULT NEURAL STEM CELLS

TABLE 1 Transfection Protocol Amount of lipid ⫹ DNA (␮l ⫹ ␮g)

LipofectAMINE Cellfectin Lipofectin DMRIE-C

1

2

3

4

5

6

0.16 ⫹ 0.05 0.23 ⫹ 0.05 0.16 ⫹ 0.05 0.08 ⫹ 0.05

0.31 ⫹ 0.1 0.47 ⫹ 0.1 0.31 ⫹ 0.1 0.16 ⫹ 0.1

0.625 ⫹ 0.2 0.94 ⫹ 0.2 0.625 ⫹ 0.2 0.31 ⫹ 0.2

1.25 ⫹ 0.4 1.87 ⫹ 0.4 1.25 ⫹ 0.4 0.625 ⫹ 0.4

2.5 ⫹ 0.8 3.75 ⫹ 0.8 2.5 ⫹ 0.8 1.25 ⫹ 0.8

5 ⫹ 1.6 7.5 ⫹ 1.6 5 ⫹ 1.6 2.5 ⫹ 1.6

In this initial screen we found that DMRIE-C was most efficient. After further optimization we established the following protocol. In each well of a 24-well plate, 125 ␮l culture medium without supplements was mixed with 1.5 ␮l DMRIE-C (Life Technologies). Plasmid DNA (0.8 ␮g pEGFP-N1) was diluted in 125 ␮l culture medium without supplements and added to the DMRIE-C mixture. The plate was incubated at room temperature for 45 min and secondary neurospheres were then transfected as above. Adenovirus. GFP (Clontech) was cloned into the pQBI-AdCMV5 vector (Q-BIOgene) and the Adeno-quest kit (Q-BIOgene) was used to produce adenovirus serotype 5 expressing GFP under the control of the CMV promoter. Virus was purified through two discontinuous CsCl gradients, and CsCl was then removed with a NAP-5 desalting column (Pharmacia). Secondary neurospheres passaged the previous day were infected with 161 PFU per cell and induced to differentiate 20 h after infection on polyornithine-coated glass coverslips. The cells were fixed 24 h later in 4% formaldehyde for 10 min at room temperature, stained, and counted as above. Retrovirus. Vesicular stomatitis virus G-protein (VSV-G)pseudotyped retroviral vectors were produced using the packaging cell line 293 GPG as described [7]. The cells were transfected with 25 ␮g of plasmid DNA, and the transfection was ended 8 h later by adding 10 ml of fresh medium. The medium was collected from the cells, and fresh medium added, every day for 8 days. Viral particles were centrifuged at 44,000g for 90 min and the pellet was resuspended in 200 ␮l of DMEM/F12. The titer was calculated by infection of COS-4 cells, and the titer was determined to be 1.5 ⫻ 10 8 transducing units (TU)/ml. Secondary neurospheres were infected at a multiplicity of infection (MOI) of 3.5 and the cells were cultured another 20 h before differentiation was induced. Two days after transduction, the cells were fixed, stained, mounted, and counted as described above. Some cells were analyzed by immunocytochemistry with a mouse anti-␤III-tubulin antibody (Tuj1 0.7 ␮g/ml, Babco) and either DAPI (1.5 ␮g/ml, Sigma) or mouse anti-glial fibrillary acidic protein (GFAP) antibody (0.4 ␮g/ml, Sigma). The primary antibodies were detected with Cy3 goat anti-mouse antibodies (2.8 ␮g/ml, Jackson) or AMCA goat anti-rabbit antibodies (75 ␮g/ml, Jackson). Three independent experiments were conducted and the average percentage of cells expressing both eGFP and ␤-tubulin was calculated. Lentivirus. VSV-G-pseudotyped lentiviral vectors were produced by cotransfecting 293T cells with pHR⬘CMV-GFP-W (vector plasmid), pCMV⌬R8.91 (packaging plasmid), and pMD.G (envelope plasmid) as described [8, 9]. The medium was collected 2 and 3 days after transfection and the particles were collected by ultracentrifugation at 26,000 rpm for 90 min. The pellet was resuspended in 200 ␮l of medium. The viral titer was calculated by infection of 293T cells and was determined to be 1.2 ⫻ 10 8 TU/ml. Secondary neurospheres were infected at a MOI of 20 and the rest of the experiment was as for retrovirus. In vivo plasmid transfection. Endotoxin-free DNA was prepared using affinity columns (EndoFree plasmid mega kit, Qiagen). DNA– polyethyleneimine (PEI) complexes were prepared according to instructions by the manufacturer. In short, plasmid DNA (10 ␮g) was

diluted in a sterile solution of 5% glucose to a final volume of 48.2 ␮l and complexed with 1.8 ␮l linear PEI (22 kDa, MBI Fermentas) in a ratio of 6 PEI nitrogens per phosphate. Five microliters of the PEI/ plasmid mix was stereotaxically injected into the left lateral ventricle of adult mice (0.5 mm posterior and 0.7 mm lateral to bregma and 2 mm below the dura mater). To establish the average number of EGFP-expressing cells we analyzed 20 consecutive cryostat sections (50 ␮m), spanning the lateral ventricles. To limit gene expression to neural stem and progenitor cells, we used the plasmid pNES1852tk/ lacZ, encoding the lacZ gene under control of the enhancer elements in the second intron of the human nestin gene [10]. The ependymal layer of the lateral ventricles was labeled by a stereotaxic injection, into the left lateral ventricle, of 2 ␮l red fluorescent RetroBeads (Lumafluor) diluted 1:100 in PBS. Four days later 2 ␮g plasmid DNA (pNes1852tk/lacZ) complexed with linear PEI was injected into the contralateral lateral ventricle. Four days after the PEI/plasmid injection, animals were perfused with 4% paraformaldehyde. Cryostat sections (50 ␮m) were incubated with rabbit anti-␤-galactosidase antibody (0.6 ␮g/ml, Cappel) overnight at 4°C and detected with Alexa 488 goat anti-rabbit antibody (1 ␮g/ml, Molecular Probes). Sections were examined with a Zeiss laser scanning confocal microscope.

RESULTS AND DISCUSSION

To establish a nonviral gene expression protocol, we transfected adult mouse primary neural stem cells with a plasmid encoding enhanced green fluorescent protein (eGFP) under the control of the CMV promoter. A panel of different cationic lipids was analyzed with respect to efficiency and toxicity at different DNA/lipid concentrations (Table 1). Stem cells were cultured as clonal aggregates, often referred to as neurospheres [11, 12], which were dissociated the day before transfection to increase the exposure of individual cells. The cells were plated on an adhesive substrate 24 h after the transfection, and the number of eGFP-expressing cells was quantified 24 h later. The efficiency and toxicity varied greatly among the lipids (Fig 1A). DMRIE-C was most efficient, resulting in 11% of cells expressing eGFP under the initial conditions, and 15% after optimizing this protocol, with very limited cell death (Fig. 1). Thus, adult primary neural stem cells can be transfected with a relatively high efficiency (Figs. 1B and 1C). We next evaluated the utility of different viral vectors in vitro (Fig. 1). Retroviruses produced from the packaging cell lines BOSC 23 or Eco Pack gave no or

36

FALK ET AL.

FIG. 1. Gene delivery to adult neural stem cells in vitro. Four different cationic lipids were tested for their efficiency in transfecting adult neural stem cells in vitro (A). Increasing lipid/DNA concentrations were analyzed for each lipid (see Table 1). The highest lipid/DNA concentrations induced cell death, indicated by a cross. Transfection with DMRIE-C proved most efficient and was optimized further and compared to infection with VSV-G-pseudotyped lentivirus (lenti), VSV-G-pseudotyped retrovirus (retro), and adenovirus (adeno) for efficiency in transducing adult neural stem cells in vitro (B). Bars represent the mean transfection efficiency ⫾ SEM from three independent experiments. (C–F) Representative images of expression of eGFP in transfected or virally transduced cells. Nuclei are stained with DAPI and appear blue. Bar, 40 ␮m.

very limited transduction. Retroviruses bind to specific cell surface receptors to infect cells, whereas VSV-Gpseudotyped viruses interact with a phopholipid component in the plasma membrane giving an extremely broad host range [13]. VSV-G psdeudotyped retrovirus gave transduction efficiencies approaching 60% in neural stem cells (Figs. 1B and 1E). VSV-G-pseudotyped lentivirus was also examined for its ability to infect neural stem cells, but was markedly less efficient than VSV-G-psdeudotyped retrovirus (⬍10%, Figs. 1B and 1D). The low efficiency observed with the VSV-Gpseudotyped lentivirus in our culture system is consis-

tent with other observations of lower transduction of mouse than rat cells in vitro with this virus [14]. Adenovirus was most efficient in our analysis (68%, Figs. 1B and 1F), consistent with prominent expression of CVADR, the receptor for adenovirus in neural stem cells [6]. However, we have occasionally seen altered differentiation of adenovirus infected neural stem cells (data not shown, Ref. [15]), which we did not see with any of the other vectors. To test whether these gene delivery techniques can be used to direct the differentiation of adult neural stem cells to a certain fate, we engineered a VSV-G-

GENE DELIVERY TO ADULT NEURAL STEM CELLS

37

FIG. 2. Neurogenin 2-induced neurogenesis from adult neural stem cells in vitro. Adult neural stem cells were transduced with VSV-G-pseudotyped retrovirus encoding either IRES2-eGFP (A) or neurogenin 2-IRES2-eGFP (B and C). (A) Approximately 1% of cells transduced with eGFP alone differentiate to neurons, as indicated by ␤III-tubulin immunoreactivity (red). (B) When neural stem cell cultures are transduced with neurogenin 2 retrovirus, 90% of transduced cells differentiate to a neuronal fate. (C) Only rarely do neurogenin 2-expressing cells take on an astrocytic identity, as indicated by GFAP immunoreactivity (shown in blue). Nuclei in (A) and (B) are labeled with DAPI and appear blue. Virally transduced cells are identified by eGFP expression (shown in green). Bar in A and B, 40 ␮m; bar in C, 20 ␮m. FIG. 3. Gene delivery to adult neural stem cells in vivo. Linear PEI was used to deliver plasmid DNA (pEGFP-N1) to cells lining the lateral ventricles, via intraventricular injection. (A) Quantification of the number of eGFP-positive cells based on the location in the lateral ventricle (LV) wall: ventral (V), lateral (L), dorsolateral (DL), dorsal (D), and medial (M). Bars show the mean ⫾ SEM from three animals. (B) The dorsolateral horn (indicated by box in A) and the ventral wall of the injected lateral ventricle had the highest density of eGFP-positive cells. Nuclei are labeled with propidium iodide and appear red. (C, D) To specifically target gene expression to neural stem/progenitor cells, plasmid DNA (pNES1852tk/lacZ), encoding lacZ driven by the regulatory elements of the nestin gene was complexed with PEI and injected into the lateral ventricle. The ependymal layer was labeled with rhodamine beads injected into the contralateral ventricle 4 days prior to transfection. LacZ-expressing cells (green) were detected in both the ependymal layer (E) as well as in the subventricular zone (SVZ). Bar in B, 20 ␮m; bar in C and D, 10 ␮m.

pseudotyped retrovirus driving the expression of the neural determination basic helix–loop– helix protein neurogenin 2 [16] followed by an internal ribosomal entry site and eGFP. Following induction of differentiation, 90% of neurogenin 2-infected cells were immunoreactive for the neuron specific protein ␤III-tubulin

(Fig. 2B), compared to 1% for cells transduced with the control retrovirus (Fig. 2A). Of the neurogenin 2-transduced cells, very few expressed the astrocyte marker GFAP after differentiation (Fig. 2C), in comparison to cells receiving no virus or the control virus, where the GFAP-positive cells were in majority (data not shown).

38

FALK ET AL.

Thus, VSV-G-pseudotyped retroviral particles can be used in vitro to reproducibly transduce adult neural stem cells and efficiently induce differentiation to a neuronal fate. Viral vectors have also been used to transduce cells lining the ventricular system of the brain in vivo. Neural stem cells reside in the ependymal layer, which lines the ventricular system, and in the adjacent subventricular zone [6, 17]. Adenovirus very efficiently infect ependymal cells [18], which express high levels of CVADR, a cell surface protein that acts as a receptor for adenovirus [6]. However, deeper cell layers cannot be reached by intraventricular injection [18]. Moreover, adenoviral infection evokes an immune response that may hamper the analysis of the effect of specific genes as well as be a serious problem for therapeutic use [19]. Both retrovirus and lentivirus are extremely ineffective in infecting ventricular wall cells in the adult mouse brain [6] (and data not shown). Although most efforts have been directed to engineering viral vectors for in vivo gene delivery, the application of nonviral techniques can provide several potential benefits. For example, larger DNA fragments can be accommodated in nonviral, plasmid-based, gene delivery systems than in most viral systems. In this study we used the polycation PEI, which can efficiently condense DNA. PEI/DNA complexes taken into cellular endosomal compartments are protected from degradation and are readily released into the cytoplasm [20]. PEI/DNA complexes have been shown to transfect both neurons and glia following injection in the mouse brain [21, 22]. Disruption of the ependymal layer is sometimes seen after PEI/DNA injections, which may indicate some toxicity. To establish a nonviral gene expression protocol for in vivo gene delivery to adult neural stem cells we used linear PEI to condense plasmid DNA (CMVeGFP) and injected the PEI/plasmid complexes into the lateral ventricle of adult mice. The ipsilateral ventricle consistently showed more eGFP-positive cells than the contralateral ventricle (Fig. 3A). The dorsolateral horn and the ventral part of the ipsilateral ventricle showed the highest density of eGFP-expressing cells (Figs. 3A and 3B). Furthermore, PEI/DNA complexes transfected both ependymal and subependymal cells. Thus, in contrast to the suggestion in a recent study [23], we do not find that the ability of PEI complexes to transfect cells following intraventricular injection is limited to a single discrete cell population. A characteristic feature of central nervous system stem and progenitor cells is expression of the intermediate filament protein nestin, and the second intron of the human and rat nestin gene functions as an enhancer directing gene expression to neural stem and progenitor cells [10, 24]. To limit gene expression to neural stem cells we used a plasmid encoding the lacZ gene under control of the enhancer elements in the

second intron of the human nestin gene [10]. To visualize the distribution of the transfected cells, the ependymal layer was first labeled by intraventricular injection of fluorescent beads [17], 4 days prior to the PEI/plasmid injection. As expected, lacZ-positive cells were detected in both the ependymal layer and the subventricular zone (Figs. 3C and 3D) at a much lower frequency than observed with the ubiquitous CMV promoter. The transient nature of plasmid expression limits its use and in situations where prolonged expression of a gene is desired stable integration may be required. However, there are several applications for which transient expression is sufficient and even desired. For example, genes directing cellular differentiation are, in general, transiently expressed during development, and a limited period of expression may be advantageous in situations where one may want to study the role of such genes in stem cells. We thank Carina Lothian and Urban Lendahl for the gift of pNes1852tk/lacZ and Richard Mulligan for providing the 293 GPG cells, Diogo Castro for providing adenovirus, and Elisabeth Andersson for the plasmid construct used to prepare the neurogenin 2 retrovirus. This project was supported by the Swedish Foundation for Strategic Research, the Swedish Medical Research Council, Project A.L.S., EU QLG3-2000-01343, the Karolinska Institute, and the Swedish Cancer Foundation.

REFERENCES 1.

Gage, F. H. (1998). Cell therapy. Nature 392(Suppl.), 18 –24.

2.

Bjo¨ rklund, A., and Lindvall, O. (2000). Cell replacement therapies for central nervous system disorders. Nature Neurosci. 3, 537–544.

3.

Tanabe, Y., William, C., and Jessell, T. M. (1998). Specification of motor neuron identity by the MNR2 homeodomain protein. Cell 95, 67– 80.

4.

Heim, D. A., and Dunbar, C. E. (2000). Hematopoietic stem cell gene therapy: Towards clinically significant gene transfer efficiency. Immunol. Rev. 178, 29 –38.

5.

Klug, C. A., Chesier, S., and Weissman, I. L. (2000). Inactivation of a GFP retrovirus occurs at multiple levels in long-term repopulating stem cells and their differentiated progeny. Blood 96, 894 –901.

6.

Johansson, C. B., Momma, S., Clarke, D. L., Risling, M., Lendahl, U., and Frise´ n, J. (1999). Identification of a neural stem cell in the adult mammalian central nervous system. Cell 96, 25–34.

7.

Ory, D. S., Neugeboren, B. A., and Mulligan, R. C. (1996). A stable human-derived packaging cell line for production of high titer retrovirus/vesicular stomatitis virus G pseudotypes. Proc. Natl. Acad. Sci. USA 93, 11400 –11406.

8.

Zufferey, R., Nagy, D., Mandel, R. J., Naldini, L., and Trono, D. (1997). Multiply attenuated lentiviral vectors achieves efficient gene delivery in vivo. Nature Biotechnol. 15, 871– 875.

9.

Zufferey, R., Donello, J. E., Trono, D., and Hope, T. J. (1999). Woodchuck hepatitis virus posttranscriptional regulatory element enhances expression of transgenes delivered by retroviral vectors. J. Virol. 73, 2886 –2892.

GENE DELIVERY TO ADULT NEURAL STEM CELLS 18.

10.

Lothian, C., and Lendahl, U. (1997). An evolutionarily conserved region in the second intron of the human nestin gene directs gene expression to CNS progenitor cells and to early neural crest cells. Eur. J. Neurosci. 9, 452– 462.

11.

Reynolds, B. A., and Weiss, S. (1992). Generation of neurons and astrocytes from isolated cells of the adult mammalian nervous system. Science 255, 1707–1710.

12.

Reynolds, B. A., and Weiss, S. (1996). Clonal and population analyses demonstrate that an EGF-responsive mammalian embryonic precursor is a stem cell. Dev. Biol. 175, 1–13.

13.

Yee, J. K., Friedmann, T., and Burns, J. C. (1994). Generation of high-titer pseudotyped retroviral vectors with very broad host range. Methods Cell Biol. 43, 99 –112.

21.

14.

Englund, U., Ericson, C., Rosenblad, C., Mandel, R. J., Trono, D., Wictorin, K., and Lundberg, C. (2000). The use of a recombinant lentiviral vector for ex vivo gene transfer into the rat CNS. NeuroReport 11, 3973.

22.

15.

Hughes, S. M., Moussavi-Harami, F., Sauter, S. L., and Davidsson, B. L. (2002). Viral-mediated gene transfer to mouse primary neural progenitor cells. Mol. Ther. 5, 16 –24.

16.

Anderson, D. J. (1999). Lineages and transcription factors in the specification of vertebrate primary sensory neurons. Curr. Opin. Neurobiol. 9, 517–524.

17.

Doetsch, F., Caille, I., Lim, D. A., Garcia-Verdugo, J. M., and Alvarez-Buylla, A. (1999). Subventricular zone astrocytes are neural stem cells in the adult mammalian brain. Cell 97, 703– 716.

Received May 1, 2002

19.

20.

23.

24.

39

Davidson, B. L., and Bohn, M. C. (1997). Recombinant adenovirus: A gene transfer vector for study and treatment of CNS disease. Exp. Neurol. 144, 125–130. Wood, M. J. A., Charlton, H. M., Wood, K. J., Kajiwara, K., and Byrnes, A. P. (1996). Immune responses to adenovirus vectors in the nervous system. Trends Neurosci. 19, 497–501. Boussif, O., Lezoualch, F., Zanta, M. A., Mergny, M. D., Scherman, D., Demeneix, B., and Behr, J. P. (1995). A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: Polyethylenimine. Proc. Natl. Acad. Sci. USA 92, 7297– 7301. Abdallah, B., Hassan, A., Benoist, C., Goula, D., Behr, J. P., and Demeneix, B. A. (1996). A powerful nonviral vector for in vivo gene transfer into the adult mammalian brain: Polyethylenimine. Hum. Gene Ther. 20, 1947–1954. Lemkine, G. F., Goula, D., Becker, N., Paleari, L., Levi, G., and Demeneix, B. A. (1999). Optimisation of polyethyleniminebased gene delivery to mouse brain. J. Drug Target 7, 305–312. Lemkine, G. F., Mantero, S., Migne´ , C., Raji, A., Goula, D., Normandie, P., Levi, G., and Demeneix, B. A. (2002). Preferential transfection of adult mouse neural stem cells and their immediate progeny in vivo with polyethylenimine. Mol. Cell. Neurosci. 19, 165–174. Zimmerman, L., Lendahl, U., Cunningham, M., McKay, R., Parr, B., Gavin, B., Mann, J., Vassileva, G., and McMahon, A. (1994). Independent regulatory elements in the nestin gene direct transgene expression to neural stem cells or muscle precursors. Neuron 12, 11–24.

Lihat lebih banyak...

Comentários

Copyright © 2017 DADOSPDF Inc.