Biomolecule Delivery in Neuroregeneration Strategies: Comparison
Please note this is a comparison between Version 2 by Jessie Wu and Version 1 by Angelina Angelova.

Neurodegenerative disorders have sophisticated etiology and represent a serious challenge for society. Among the various risk factors, oxidative stress and chronic neuroinflammation (which can be due to viral infection or other causes) are involved in the pathogenesis of Parkinson’s disease (PD), Alzheimer’s disease (AD), Huntington disease (HD), and amyotrophic lateral sclerosis (ALS). Enabling challenging applications of nanomedicine and precision medicine in the treatment of neurodegenerative disorders requires deeper investigations of nanocarrier-mediated biomolecular delivery for neuronal targeting and recovery. WResearchers place emphasis on nanoformulations for the delivery of brain-derived neurotrophic factor (BDNF) using different types of lipidic nanocarriers (liposomes, liquid crystalline or solid lipid nanoparticles), hydrogels, and scaffolds. 

  • BDNF
  • lipid nanoparticles
  • nanocarriers
  • growth factor delivery
  • nanomedicine
  • biotherapeutics
  • neuroregeneration

1. Neuroprotective Biomolecules and Nucleic Acids under Current Investigation

1.1. Neurotrophic Factor Protein-Based Therapies

Neurotrophic factors (NTFs) are a family of biomacromolecules (large peptides or small proteins) that support the growth, survival, and differentiation of developing and mature neurons by protecting them from injury and neurotoxins [34,35][1][2]. Nerve growth factor (NGF) was the first NTF discovered by Levi-Montalcini [36][3]. Subsequently, the neuroprotective functions of several other NTFs have been reported over the years [33,37,38,39,40,41,42,43,44,45,46][4][5][6][7][8][9][10][11][12][13][14]. They have been categorized into three main families: (i) the neurotrophin family, including NGF, BDNF, neurotrophin-3 (NT-3), and neurotrophin-4 (NT-4); (ii) the glial cell line-derived neurotrophic factor (GNDF) family, e.g., GDNF, neurturin (NRTN), artemin (ARTN), and persephin (PSPN); and (iii) the neuropoietic cytokines, e.g., ciliary neurotrophic factor (CNTF), leukemia inhibitory factor (LIF), and cardiotrophin (CT-1). Other proteins, such as fibroblast growth factor-1 and -2 (FGF-1 and FGF-2) and platelet-derived growth factor (PDGF), as well as polypeptides, including pituitary adenylate cyclase-activating peptide (PACAP), insulin-like growth factor 1 (IGF-1), human neuropeptide substance P, macrophage colony-stimulating factor (M-CSF), and granulocyte-macrophage colony-stimulating factor (GM-CSF), can also play a role as NTFs [47,48,49,50,51,52,53,54,55,56,57,58][15][16][17][18][19][20][21][22][23][24][25][26].
A novel family of unconventional NTFs, cerebral dopamine neurotrophic factor (CDNF) and mesencephalic astrocyte-derived neurotrophic factor (MANF), which are both structurally and mechanistically distinct from the other growth factors, have shown neurorestorative effects in animal models of PD [33][4]. These biotherapeutics localize to the lumen of the endoplasmic reticulum (ER) and likely modulate the unfolded protein response (UPR) pathway. Intermittent monthly bilateral intraputamenal infusions of CDNF have recently been tested in a randomized placebo-controlled phase I–II clinical trial in PD patients [33][4].
Studies of an AD rat model with amyloid-β-induced memory loss have demonstrated that granulocyte colony stimulating factor (GCSF), an endogenous neuronal hematopoietic factor protein, improves memory and neurobehavioral functions [39][7]. GCSF exerted neuroprotective activity associated with significant memory improvements, increased levels of antioxidant enzymes and total RNA expression in the brain, and reduced lipid peroxidation and acetylcholinesterase levels. In addition, GCSF induces neurogenesis, as evidenced by the increased number of progenitor CD34+ cells in the brain [39][7]. Clinical trials using GCSF for the treatment of AD and stroke have already been carried out [55,56,57][23][24][25]. The advantages of GCSF, as a good candidate for clinical trials in NDs, also include its capacity for crossing the BBB and its strong anti-apoptotic activity.
Several clinical trials have been conducted to examine the capacity of GDNF, NRTN and PGDF to rescue degenerating dopaminergic neurons in the substantia nigra and their axon terminals in the striatum [44,54][12][22]. GDNF has been studied as a candidate in clinical trials of PD considering its neurorestorative effects established in PD animal models [51,58][19][26]. The performed in vitro and in vivo studies with PD models have demonstrated the neuroprotective and neurorestorative effects of GDNF on midbrain dopaminergic neurons [51,52,53,54][19][20][21][22]. Unlike GCSF, the penetration of GDNF in the brain is strongly limited. Therefore, various strategies have been undertaken for GDNF delivery to the dopamine-depleted brain, e.g., implantation of microspheres, transfection by viral vectors, or ventricle and intraputaminal infusion of the protein [58,59,60][26][27][28]. The delivery of BDNF by nanoparticles and other biomimetic nanoscale assemblies will be presented in a separate section below.

1.2. siRNA-Based Therapy

Emerging strategies for the prevention or treatment of NDs are being developed based on selective silencing of mutant alleles. This approach aims to directly arrest the causative mutant genes for neurodegeneration [61][29]. RNA interference (RNAi) regulates the expression of genes by controlling the synthesis of proteins via a post-transcriptional gene-silencing mechanism. Long double-stranded RNA sequences are cleaved by the cytoplasmic enzyme Dicer into fragments (21–23 nucleotides long) called small interfering RNAs (siRNAs). siRNA is incorporated into a protein complex called the “RNA-induced silencing complex”, and then the sense strand of the siRNA is cleaved. The antisense strand guides the RNA-induced silencing complex to bind with a messenger RNA (mRNA), which is complementary to the antisense strand and degrades it. An important advantage of RNAi over small-molecule and protein therapeutics is that mutant alleles can be targeted with RNAi. In principle, any transcript that encodes a protein that causes or contributes to a disease can be targeted by RNAi [62][30]. Therefore, a major advantage of sequence-based targeting technologies is the ability to design precisely targeted biotherapeutics for almost any target sequence (coding or noncoding), regardless of the function of the gene product [63][31].
The therapeutic potential of RNAi in AD has been demonstrated through allele-specific gene silencing by short-hairpin RNA (shRNA) [62][30]. An anti-APPsw shRNA was delivered by the recombinant adeno-associated virus to the hippocampus of AD transgenic mice (APP/PS1) to selectively suppress mutant APP. No neuronal toxicity was detected in short- and long-term transduction experiments with the viral vector. Intravenously injected rabies virus glycoprotein (RVG)-targeted exosomes have specifically delivered siRNA to neural cells in the mouse brain. Strong mRNA (60%) and protein (62%) knockdown of BACE1 was achieved without noticeable immune stimulation. CBP-1 (acetyltransferase enzyme) has been inhibited by RNAi to evaluate the age-dependent mortality rate for 30 drugs used for protection of mammalian neurons. The genes of interest, which may be more specifically involved in the tau phosphorylation pathways in AD, are DYRK1A and AKAP13 [62][30].
Several obstacles remain for the clinical development of RNAi-based therapeutics [63][31]. The delivery issue represents a major challenge, as siRNA should be transferred to specific target sites, and the potential off-target effects should be taken into consideration as well. AD is a multifactor and genetically heterogeneous disorder. It cannot be treated by a single siRNA sequence. Therefore, new strategies should be envisioned to formulate the various RNAi components and successfully deliver them to the target sites.

2. Therapeutic Delivery Approaches for Neuroprotective Biomacromolecules

2.1. Invasive versus Noninvasive Administration of Carrier-Free Biomolecules

The major reason for the limited effect of therapeutic biomacromolecules (therapeutic peptides or proteins) in clinical trials has been attributed to the presence of the BBB [27,30][32][33]. Local delivery to the brain has been suggested via stereotactic cerebral injection or intracerebral infusion [32][34]. The problem of this approach is the difficulty in determining the most appropriate doses of each compound. For instance, intracerebral neurotrophic factor administration has shown no improvement of motor symptoms in PD (owing to the difficulty for the drug to cross the blood–brain barrier) and thus represents its limited efficacy in clinical trials [64][35]. Therefore, different approaches for biomolecule delivery are required to increase bioavailability [65,66,67,68][36][37][38][39].
A direct route to reach the brain without going through the BBB is the nasal-to-brain delivery route [69,70][40][41]. Intranasal drug administration avoids hepatic first-pass metabolism and has been considered a safe, noninvasive route [71,72,73][42][43][44]. In this method, the therapeutic drug, which is applied into the nasal cavity, can penetrate the central nervous system (CNS) via the olfactory and/or trigeminal nerves [73][44]. Different models have been used to evaluate nasal drug absorption both in vitro and in vivo [70,73,74,75][41][44][45][46]. Some biomolecules, such as CNTF, BDNF, and NT-4/5, have been successfully delivered to the hippocampus and cerebral cortex of rats. Quick absorption of BDNF has been observed due to the interaction of BDNF molecules (exposing cationic surface charges) and the nasal mucosa (negatively charged) [75][46].

2.2. Gene Delivery

Another strategy to alter local protein expression is based on gene delivery [76,77][47][48]. Several clinical trials have been performed to examine the capacity of neurotrophic factors to rescue degenerating neurons by viral vector-mediated gene delivery to the brain [76,77,78,79,80][47][48][49][50][51]. A cationic nanocarrier functionalized by dexamethasone and cell-penetrating peptides increased BDNF expression upon BDNF DNA delivery [77][48]. Many authors have demonstrated the tolerability of gene delivery to PD patients (e.g., intraputaminal injections of adeno-associated virus serotype 2-neurturin (CERE-120)) in a phase I open-label clinical test [78,79,80][49][50][51]. Although these gene therapy approaches have been shown to be safe, their efficacy in phase II clinical trials has been considered insufficient [80][51].

2.3. Carrier-Mediated Delivery Employing Different Nanoscale Materials

Recent research has focused on the development of neurotrophin delivery systems that can provide a safe and efficient neurotrophic supply over the long term [81,82,83,84,85,86,87,88,89][52][53][54][55][56][57][58][59][60]. It has been of special interest to combine such systems with implants, i.e., to explore implant-coupled drug delivery [81,82,83][52][53][54]. An encapsulated cell biodelivery (ECB) device has been demonstrated to be an efficient method to improve NGF levels in AD patients [89][60]. Other promising approaches have comprised electrode coating materials [82][53] as well as carrier systems such as hydrogels [83[54][55],84], microspheres [85][56], nanotubes [47][15], mesoporous silica supraparticles [86][57], or nanoparticles [87,88][58][59].

2.4. Nanoparticles modified by BDNF-derived peptides for drug delivery to neurons

The recognition mechanism of BDNF ligands has been used as a targeted strategy to the CNS [175][61]. Xu et al. demonstrated the internalization of PEG-PCL nanoparticles, whose surface was decorated by a BDNF-derived (IKRG) peptide, into neuronal cells [175][61]. The tetrapeptide (IKRG) amino acid sequence has been shown to mimic the function of BDNF in targeting TrkB receptors, which are abundant in neurons [175][61]. Enhanced uptake of peptide-modified PEG-PCL nanoparticles has been observed in TrkB-positive PC12 cells but not in TrkB-negative HeLa cells [175][61].

Dąbkowska et al. successfully delivered BDNF to neuronal SH-SY5Y cells via PEGylated poly(amidoamine) dendrimer (PAMAM) nanoparticles [176][62]. The BDNF-loaded nanoparticles were stabilized by electrostatic interactions (Figure 8). The studied BDNF-PAMAM-AF488-PEG nanoparticles have been characterized by slow release of the therapeutic agent and strong interaction with the cell membrane surface [176][62].

A nanofiber hydrogel has been formulated with a mixture of two peptides, one of which is a BDNF mimetic peptide [177][63]. The purpose has been to promote the promyelination of Schwann cells and the adhesion and proliferation of endothelial cells. RKKADP is a BDNF mimetic peptide that has self-assembled in water and has formed a hydrogel network [177][63]. Edelbrock et al. reported that BDNF mimetic peptide can activate BDNF-TrkB signaling as well as other downstream signaling cascades capable of promoting neuronal cell infiltration and functional maturation [178][64]. The regenerative efficacy, maturation of nerve fibers, and vascularization effect have also been confirmed in vivo.

 

References

  1. Subbiah, R.; Guldberg, R.E. Materials science and design principles of growth factor delivery systems in tissue engineering and regenerative medicine. Adv. Healthc. Mater. 2019, 8, e1801000.
  2. Ferenz, K.B.; Gast, R.E.; Rose, K.; Finger, I.E.; Hasche, A.; Krieglstein, J. Nerve growth factor and brain-derived neurotrophic factor but not granulocyte colony-stimulating factor, nimodipine and dizocilpine, require ATP for neuroprotective activity after oxygen–glucose deprivation of primary neurons. Brain Res. 2012, 1448, 20–26.
  3. Levi-Montalcini, R. Growth control of nerve cells by a protein factor and its antiserum: Discovery of this factor may provide new leads to understanding of some neurogenetic processes. Science 1964, 143, 105–110.
  4. Huttunen, H.J.; Saarma, M. CDNF protein therapy in Parkinson’s disease. Cell Transpl. 2019, 28, 349–366.
  5. Ivanova, L.; Karelson, M.; Dobchev, D.A. Identification of natural compounds against neurodegenerative diseases using in silico techniques. Molecules 2018, 23, 1847.
  6. Liu, R.; Hudalla, G.A. Using self-assembling peptides to integrate biomolecules into functional supramolecular biomaterials. Molecules 2019, 24, 1450.
  7. Prakash, A.; Medhi, B.; Chopra, K. Granulocyte colony stimulating factor (GCSF) improves memory and neurobehavior in an amyloid-β induced experimental model of Alzheimer’s disease. Pharm. Biochem Behav. 2013, 110, 46–57.
  8. Lucini, C.; D’Angelo, L.; Cacialli, P.; Palladino, A.; De Girolamo, P. BDNF, brain, and regeneration: Insights from zebrafish. Int. J. Mol. Sci. 2018, 19, 3155.
  9. Numakawa, T.; Odaka, H.; Adachi, N. Actions of brain-derived neurotrophin factor in the neurogenesis and neuronal function, and its involvement in the pathophysiology of brain diseases. Int. J. Mol. Sci. 2018, 19, 3650.
  10. Runeberg-Roos, P.; Piccinini, E.; Penttinen, A.M.; Mätlik, K.; Heikkinen, H.; Kuure, S.; Bespalov, M.M.; Peränen, J.; Garea-Rodríguez, E.; Fuchs, E.; et al. Developing therapeutically more efficient neurturin variants for treatment of Parkinson’s disease. Neurobiol. Dis. 2016, 96, 335–345.
  11. Fletcher, J.L.; Murray, S.S.; Xiao, J. Brain-derived neurotrophin factor in central nervous system nyelination: A New mechanism to promote myelin plasticity and repair. Int. J. Mol. Sci. 2018, 19, 4131.
  12. Sullivan, A.M.; Toulouse, A. Neurotrophic factors for the treatment of Parkinson’s disease. Cytokine Growth Factor Rev. 2011, 22, 157–165.
  13. Aron, L.; Klein, R. Repairing the parkinsonian brain with neurotrophic factors. Trends Neurosci. 2011, 34, 88–100.
  14. Paul, G.; Zachrisson, O.; Varrone, A.; Almqvist, P.; Jerling, M.; Lind, G.; Rehncrona, S.; Linderoth, B.; Bjartmarz, H.; Shafer, L.L.; et al. Safety and tolerability of intracerebroventricular PDGF-BB in Parkinson’s disease patients. J. Clin. Investig. 2015, 125, 1339–1346.
  15. Dharmadana, D.; Adamcik, J.; Ryan, T.M.; Appiah Danso, S.; Chong, C.J.H.; Conn, C.E.; Reynolds, N.P.; Mezzenga, R.; Valéry, C. Peptide substance P self-assembles into semi-flexible nanotubes that can be manipulated for nanotechnology. Nanoscale 2020, 12, 22680–22687.
  16. Schäbitz, W.R.; Krüger, C.; Pitzer, C.; Weber, D.; Laage, R.; Gassler, N.; Aronowski, J.; Mier, W.; Kirsch, F.; Dittgen, T.; et al. A neuroprotective function for the hematopoietic protein granulocyte-macrophage colony stimulating factor (GM-CSF). J. Cereb. Blood Flow Metab. 2008, 28, 29–43.
  17. Bianchi, V.E.; Locatelli, V.; Rizzi, L. Neurotrophic and neuroregenerative effects of GH/IGF1. Int. J. Mol. Sci. 2017, 18, 2441.
  18. Pradhan, K.; Das, G.; Gupta, V.; Mondal, P.; Barman, S.; Khan, J.; Ghosh, S. Discovery of neuroregenerative peptoid from amphibian neuropeptide that inhibits amyloid-β toxicity and crosses blood-brain barrier. ACS Chem. Neurosci. 2019, 10, 1355–1368.
  19. Chermenina, M.; Schouten, P.; Nevalainen, N.; Johansson, F.; Orädd, G.; Strömberg, I. GDNF is important for striatal organization and maintenance of dopamine neurons grown in the presence of the striatum. Neuroscience 2014, 270, 1–11.
  20. Grondin, R.; Littrell, O.M.; Zhang, Z.; Ai, Y.; Huettl, P.; Pomerleau, F.; Quintero, J.E.; Andersen, A.H.; Stenslik, M.J.; Bradley, L.H.; et al. GDNF revisited: A novel mammalian cell-derived variant form of GDNF increases dopamine turnover and improves brain biodistribution. Neuropharmacology 2019, 147, 28–36.
  21. Ibáñez, C.F.; Andressoo, J.O. Biology of GDNF and its receptors—Relevance for disorders of the central nervous system. Neurobiol. Dis. 2017, 97, 80–89.
  22. Smith, R.C.; O’Bryan, L.M.; Mitchell, P.J.; Leung, D.; Ghanem, M.; Wilson, J.M.; Hanson, J.C.; Sossick, S.; Cooper, J.; Huang, L.; et al. Increased brain bio-distribution and chemical stability and decreased immunogenicity of an engineered variant of GDNF. Exp. Neurol. 2015, 267, 165–176.
  23. Tsai, K.J.; Tsai, Y.C.; Shen, C.K. G-CSF rescues the memory impairment of animal models of Alzheimer’s disease. J. Exp. Med. 2007, 204, 1273–1280.
  24. Sanchez-Ramos, J.; Song, S.; Sava, V.; Catlow, B.; Lin, X.; Mori, T.; Cao, C.; Arendash, G.W. Granulocyte colony stimulating factor decreases brain amyloid burden and reverses cognitive impairment in Alzheimer’s mice. Neuroscience 2009, 163, 55–72.
  25. Schneider, A.; Krüger, C.; Steigleder, T.; Weber, D.; Pitzer, C.; Laage, R.; Aronowski, J.; Maurer, M.H.; Gassler, N.; Mier, W.; et al. The hematopoietic factor G-CSF is a neuronal ligand that counteracts programmed cell death and drives neurogenesis. J. Clin. Investig. 2005, 115, 2083–2098.
  26. Nutt, J.G.; Burchiel, K.J.; Comella, C.L.; Jankovic, J.; Lang, A.E.; Laws, E.R.; Lozano, A.M.; Penn, R.D.; Simpson, R.K.; Stacy, M.; et al. Implanted intracerebroventricular. Glial cell line-derived neurotrophic factor. Randomized, double-blind trial of glial cell line-derived neurotrophic factor (GDNF) in PD. Neurology 2003, 60, 69–73.
  27. Gill, S.S.; Patel, N.K.; Hotton, G.R.; O’Sullivan, K.; McCarter, R.; Bunnage, M.; Brooks, D.J.; Svendsen, C.N.; Heywood, P. Direct brain infusion of glial cell line-derived neurotrophic factor in Parkinson disease. Nat. Med. 2003, 9, 589–595.
  28. Whone, A.; Luz, M.; Boca, M.; Woolley, M.; Mooney, L.; Dharia, S.; Broadfoot, J.; Cronin, D.; Schroers, C.; Barua, N.U.; et al. Randomized trial of intermittent intraputamenal glial cell linederived neurotrophic factor in Parkinson’s disease. Brain 2019, 142, 512–525.
  29. Ramakrishna, S.; Muddashetty, R.S. Emerging role of microRNAs in dementia. J. Mol. Biol. 2019, 431, 1743–1762.
  30. Chen, S.; Ge, X.; Chen, Y.; Lv, N.; Liu, Z.; Yuan, W. Advances with RNA interference in Alzheimer’s disease research. Drug Des. Devel. Ther. 2013, 7, 117–125.
  31. Miller, V.M.; Gouvion, C.M.; Davidson, B.L.; Paulson, H.L. Targeting Alzheimer’s disease genes with RNA interference: An efficient strategy for silencing mutant alleles. Nucleic Acids Res. 2004, 32, 661–668.
  32. Asil, S.M.; Ahlawat, J.; Barroso, G.G.; Narayan, M. Nanomaterial based drug delivery systems for the treatment of neurodegenerative diseases. Biomater. Sci. 2020, 8, 4109–4128.
  33. Wong, K.H.; Riaz, M.K.; Xie, Y.; Zhang, X.; Liu, Q.; Chen, H.; Bian, Z.; Chen, X.; Lu, A.; Yang, Z. Review of current strategies for delivering Alzheimer’s disease drugs across the blood-brain barrier. Int. J. Mol. Sci. 2019, 20, 381.
  34. Angelova, A.; Angelov, B.; Drechsler, M.; Lesieur, S. Neurotrophin delivery using nanotechnology. Drug Discov. Today 2013, 18, 1263–1271.
  35. Hegarty, S.V.; Lee, D.J.; O’Keeffe, G.W.; Sullivan, A.M. Effects of intracerebral neurotrophic factor application on motor symptoms in Parkinson’s disease: A systematic review and meta-analysis. Parkinsonism Relat. Disord. 2017, 38, 19–25.
  36. Wen, Z.; Yan, Z.; He, R.; Pang, Z.; Guo, L.; Qian, Y.; Jiang, X.; Fang, L. Brain targeting and toxicity study of odorranalectin-conjugated nanoparticles following intranasal administration. Drug Deliv. 2011, 18, 555–561.
  37. Li, X.; Su, J.; Kamal, Z.; Guo, P.; Wu, X.; Lu, L.; Wu, H.; Qiu, M. Odorranalectin modified PEG-PLGA/PEG-PBLG curcumin-loaded nanoparticle for intranasal administration. Drug Dev. Ind. Pharm. 2020, 46, 899–909.
  38. Ahirrao, M.; Shrotriya, S. In vitro and in vivo evaluation of cubosomal in situ nasal gel containing resveratrol for brain targeting. Drug Dev. Ind. Pharm. 2017, 43, 1686–1693.
  39. Wu, H.; Li, J.; Zhang, Q.; Yan, X.; Guo, L.; Gao, X.; Qiu, M.; Jiang, X.; Lai, R.; Chen, H. A novel small Odorranalectin-bearing cubosomes: Preparation, brain delivery and pharmacodynamic study on amyloid-β₂₅₋₃₅-treated rats following intranasal administration. Eur. J. Pharm. Biopharm. 2012, 80, 368–378.
  40. Erdő, F.; Bors, L.A.; Farkas, D.; Bajza, Á.; Gizurarson, S. Evaluation of intranasal delivery route of drug administration for brain targeting. Brain Res. Bull. 2018, 143, 155–170.
  41. Samaridou, E.; Alonso, M.J. Nose-to-brain peptide delivery—The potential of nanotechnology. Bioorg. Med. Chem. 2018, 26, 2888–2905.
  42. Sajja, R.K.; Cudic, P.; Cucullo, L. In vitro characterization of odorranalectin for peptide-based drug delivery across the blood-brain barrier. BMC Neurosci. 2019, 20, 22.
  43. Fan, Y.; Chen, M.; Zhang, J.; Maincent, P.; Xia, X.; Wu, W. Updated progress of nanocarrier-based intranasal drug delivery systems for treatment of brain diseases. Crit. Rev. Ther. Drug Carr. Syst. 2018, 35, 433–467.
  44. Lochhead, J.J.; Thorne, R.G. Intranasal delivery of biologics to the central nervous system. Adv. Drug Deliv. Rev. 2012, 64, 614–628.
  45. Lochhead, J.J.; Wolak, D.J.; Pizzo, M.E.; Thorne, R.G. Rapid transport within cerebral perivascular spaces underlies widespread tracer distribution in the brain after intranasal administration. J. Cereb. Blood Flow Metab. 2015, 35, 371–381.
  46. Alcalá-Barraza, S.R.; Lee, M.S.; Hanson, L.R.; McDonald, A.A.; Frey, W.H.; McLoon, L.K. Intranasal delivery of neurotrophic factors BDNF, CNTF, EPO, and NT-4 to the CNS. J. Drug Target. 2010, 18, 179–190.
  47. Kirik, D.; Cederfjäll, E.; Halliday, G.; Petersén, Å. Gene therapy for Parkinson’s disease: Disease modification by GDNF family of ligands. Neurobiol. Dis. 2017, 97 Pt B, 179–188.
  48. Yoon, J.Y.; Yang, K.J.; Park, S.N.; Kim, D.K.; Kim, J.D. The effect of dexamethasone/cell-penetrating peptide nanoparticles on gene delivery for inner ear therapy. Int. J. Nanomed. 2016, 11, 6123–6134.
  49. Bartus, R.T.; Baumann, T.L.; Siffert, J.; Herzog, C.D.; Alterman, R.; Boulis, N.; Turner, D.A.; Stacy, M.; Lang, A.E.; Lozano, A.M.; et al. Safety/feasibility of targeting the substantia nigra with AAV2-neurturin in Parkinson patients. Neurology 2013, 80, 1698–1701.
  50. Marks, W.J., Jr.; Ostrem, J.L.; Verhagen, L.; Starr, P.A.; Larson, P.S.; Bakay, R.A.; Taylor, R.; Cahn-Weiner, D.A.; Stoessl, A.J.; Olanow, C.W.; et al. Safety and tolerability of intraputaminal delivery of CERE-120 (adeno-associated virus serotype 2-neurturin) to patients with idiopathic Parkinson’s disease: An open-label, phase I trial. Lancet Neurol. 2008, 7, 400–408.
  51. Marks, W.J., Jr.; Bartus, R.T.; Siffert, J.; Davis, C.S.; Lozano, A.; Boulis, N.; Vitek, J.; Stacy, M.; Turner, D.; Verhagen, L.; et al. Gene delivery of AAV2-neurturin for Parkinson’s disease: A double-blind, randomised, controlled trial. Lancet Neurol. 2010, 9, 1164–1172.
  52. Richardson, R.T.; Wise, A.K.; Thompson, B.C.; Flynn, B.O.; Atkinson, P.J.; Fretwell, N.J.; Fallon, J.B.; Wallace, G.G.; Shepherd, R.K.; Clark, G.M.; et al. Polypyrrole-coated electrodes for the delivery of charge and neurotrophins to cochlear neurons. Biomaterials 2009, 30, 2614–2624.
  53. Kikkawa, Y.S.; Nakagawa, T.; Ying, L.; Tabata, Y.; Tsubouchi, H.; Ido, A.; Ito, J. Growth factor-eluting cochlear implant electrode: Impact on residual auditory function, insertional trauma, and fibrosis. J. Transl. Med. 2014, 12, 280.
  54. Chikar, J.A.; Hendricks, J.L.; Richardson-Burns, S.M.; Raphael, Y.; Pfingst, B.E.; Martin, D.C. The use of a dual PEDOT and RGD-functionalized alginate hydrogel coating to provide sustained drug delivery and improved cochlear implant function. Biomaterials 2012, 33, 1982–1990.
  55. Endo, T.; Nakagawa, T.; Kita, T.; Iguchi, F.; Kim, T.S.; Tamura, T.; Iwai, K.; Tabata, Y.; Ito, J. Novel strategy for treatment of inner ears using a biodegradable gel. Laryngoscope 2005, 115, 2016–2020.
  56. Madduri, S.; Gander, B. Growth factor delivery systems and repair strategies for damaged peripheral nerves. J. Control. Release 2012, 161, 274–282.
  57. Wang, Y.; Wise, A.K.; Tan, J.; Maina, J.W.; Shepherd, R.K.; Caruso, F. Mesoporous silica supraparticles for sustained inner-ear drug delivery. Small 2014, 10, 4244–4248.
  58. Li, H.; Edin, F.; Hayashi, H.; Gudjonsson, O.; Danckwardt-Lillieström, N.; Engqvist, H.; Rask-Andersen, H.; Xia, W. Guided growth of auditory neurons: Bioactive particles towards gapless neural-electrode interface. Biomaterials 2017, 122, 1–9.
  59. Roy, S.; Glueckert, R.; Johnston, A.H.; Perrier, T.; Bitsche, M.; Newman, T.A.; Saulnier, P.; Schrott-Fischer, A. Strategies for drug delivery to the human inner ear by multifunctional nanoparticles. Nanomedicine 2012, 7, 55–63.
  60. Mitra, S.; Behbahani, H.; Eriksdotter, M. Innovative therapy for Alzheimer’s disease with focus on biodelivery of NGF. Front. Neurosci. 2019, 13, 38.
  61. Xu, J.; Chau, Y. Polymeric nanoparticles decorated with BDNF-derived peptide for neuron-targeted delivery of PTEN inhibitor. Eur. J. Pharm. Sci. 2018, 124, 37–45.
  62. Dąbkowska, M.; Łuczkowska, K.; Rogińska, D.; Sobuś, A.; Wasilewska, M.; Ulańczyk, Z.; Machaliński, B. Novel design of (PEG-ylated)PAMAM-based nanoparticles for sustained delivery of BDNF to neurotoxin-injured differentiated neuroblastoma cells. J. Nanobiotechnol. 2020, 18, 120.
  63. Lu, J.; Yan, X.; Sun, X.; Shen, X.; Yin, H.; Wang, C.; Liu, Y.; Lu, C.; Fu, H.; Yang, S.; et al. Synergistic effects of dual-presenting VEGF- and BDNF-mimetic peptide epitopes from self-assembling peptide hydrogels on peripheral nerve regeneration. Nanoscale 2019, 11, 19943–19958.
  64. Edelbrock, A.N.; Àlvarez, Z.; Simkin, D.; Fyrner, T.; Chin, S.M.; Sato, K.; Kiskinis, E.; Stupp, S.I. Supramolecular nanostructure activates TrkB receptor signaling of neuronal cells by mimicking brain-derived neurotrophic factor. Nano Lett. 2018, 18, 6237–6247.
More