Next Article in Journal
Neutralising Effects of Different Antibodies on Clostridioides difficile Toxins TcdA and TcdB in a Translational Approach
Previous Article in Journal
Enhanced CO2 Capture of Poly(amidoamine)-Modified Graphene Oxide Aerogels with the Addition of Carbon Nanotubes
Previous Article in Special Issue
Clinical Trials Targeting Secondary Damage after Traumatic Spinal Cord Injury
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Neurotrophic Factors as Regenerative Therapy for Neurodegenerative Diseases: Current Status, Challenges and Future Perspectives

1
Laboratory of Experimental Hematology, Vaccine and Infectious Disease Institute (Vaxinfectio), University of Antwerp, Universiteitsplein 1, B-2610 Antwerpen, Belgium
2
Department of Neurology, Antwerp University Hospital, B-2650 Edegem, Belgium
3
Neurology, Translational Neurosciences, Born Bunge Institute, Faculty of Medicine and Health Sciences, University of Antwerp, B-2610 Antwerpen, Belgium
4
Center for Cell Therapy and Regenerative Medicine (CCRG), Antwerp University Hospital, Drie Eikenstraat 655, B-2650 Edegem, Belgium
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3866; https://doi.org/10.3390/ijms24043866
Submission received: 21 September 2022 / Revised: 25 January 2023 / Accepted: 6 February 2023 / Published: 15 February 2023

Abstract

:
Neurodegenerative diseases, including Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), multiple sclerosis (MS), spinal cord injury (SCI), and amyotrophic lateral sclerosis (ALS), are characterized by acute or chronic progressive loss of one or several neuronal subtypes. However, despite their increasing prevalence, little progress has been made in successfully treating these diseases. Research has recently focused on neurotrophic factors (NTFs) as potential regenerative therapy for neurodegenerative diseases. Here, we discuss the current state of knowledge, challenges, and future perspectives of NTFs with a direct regenerative effect in chronic inflammatory and degenerative disorders. Various systems for delivery of NTFs, such as stem and immune cells, viral vectors, and biomaterials, have been applied to deliver exogenous NTFs to the central nervous system, with promising results. The challenges that currently need to be overcome include the amount of NTFs delivered, the invasiveness of the delivery route, the blood–brain barrier permeability, and the occurrence of side effects. Nevertheless, it is important to continue research and develop standards for clinical applications. In addition to the use of single NTFs, the complexity of chronic inflammatory and degenerative diseases may require combination therapies targeting multiple pathways or other possibilities using smaller molecules, such as NTF mimetics, for effective treatment.

1. Introduction

Neurodegenerative diseases of the central nervous system (CNS), such as multiple sclerosis (MS), Alzheimer′s disease (AD), Parkinson′s disease (PD), Huntington′s disease (HD), amyotrophic lateral sclerosis (ALS), and in acute cases, spinal cord injury (SCI), are still incurable and have high individual and societal costs [1,2,3]. PD and AD are the most common neurodegenerative diseases. As the world′s population ages, the prevalence of AD and PD is rapidly increasing. It is estimated that 50 million people worldwide suffer from neurodegenerative diseases, and this number will rise to 115 million by 2050 [4].
Unfortunately, currently available treatment options are inadequate to halt neurodegenerative processes [5,6]. Moreover, our understanding of the pathogenic processes and the consequent development of effective treatments is significantly complicated by the complexity of the mechanisms associated with neuronal loss and the conflicting physiological causes of these diseases. Furthermore, the difficulty in addressing widespread neuronal cell death, combined with the enormous limitations for the vast majority of drugs not to cross the blood–brain barrier (BBB), further complicates the treatment of these diseases [7,8].
From an evolutionary point of view, the nervous system would be able to protect itself from any injury [9]. In the early 20th century, pioneering work by Tello and Cajal demonstrated that the CNS has the ability to regenerate itself after injury [10,11,12]. In recent years, researchers have accumulated detailed in vitro and in vivo mechanistic evidence supporting the idea that an innate self-maintenance program is activated in the brain, not only during inflammatory and degenerative diseases, but also in healthy individuals [11,13,14]. These observations support the idea that chronic inflammatory and degenerative disorders of the brain can be the result of defective repair mechanisms, rather than uncontrollable pathogenic events [11,15,16,17]. We can, therefore, subscribe the idea that failure of molecular and cellular mechanisms sustaining the “brain-repair program”—which can be considered as an intrinsic part of the physiological activities of the brain—might be, at least partially, a cause of neurodegenerative diseases [11,18]. Therefore, research into the molecular and cellular events sustaining intrinsic brain-repair mechanisms and a better understanding of why they fail over time in chronic disorders might provide an attractive conceptual framework, in which new and efficacious therapies for neurodegenerative diseases can be developed.
Neurotrophic factors (NTFs) and their receptors play a crucial role in neural cell maturation and proliferation. NTFs regulate the development and survival of neurons, and they appear to be involved in the endogenous neuroprotection of different neurons. Several studies have reported that NTFs, particularly glial cell-derived neurotrophic factor (GDNF), ciliary neurotrophic factor (CNTF), brain-derived neurotrophic factor (BDNF), nerve growth factor (NGF), neurotrophin-3 (NT-3), and neurotrophin-4/5 (NT-4/5), act regeneratively in different animal models [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55] and patients [56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73] with neuroinflammatory and neurodegenerative diseases. Consistent with their known role in maintaining neuronal homeostasis, these NTFs, with regenerative properties, have been proposed as novel therapies for several neuroinflammatory and neurodegenerative diseases [74,75,76]. In this review, we provide an overview of the various and known NTFs described in the literature with their effects in the CNS. As well, we summarize the different approaches where NTFs have been administered via direct delivery or delivery through a vehicle, such as stem and immune cells, viral vectors, and biomaterials, into animal models or in patients suffering from a neurodegenerative disease.

2. Functions and Mechanisms of Neurotrophic Factors in Neurogenesis and Brain Repair

Glial cell-derived neurotrophic factor (GDNF) was originally isolated from the supernatant of a rat glioma cell line and found to have pronounced effects on the survival of midbrain dopaminergic neurons [77,78,79]. GDNF has further critical roles outside the nervous system in the regulation of kidney morphogenesis and spermatogenesis [80]. In the case of potential therapy for neurodegenerative diseases, GDNF has a relatively high specificity for dopaminergic neurons and, thus, has significant potential for the treatment of PD, which is mainly characterized by progressive depletion of dopaminergic cell populations in the midbrain [79]. Subsequently, GDNF was also shown to have trophic and protective effects on noradrenergic neurons in the locus coeruleus and on peripheral motor neurons, giving hope for its therapeutic potential in HD and ALS [24,25,35,51,81,82,83]. Translational research has focused mainly on the treatment of PD, where there has been reason for both celebration and caution [27,79,84,85,86,87,88,89,90,91,92,93]. A recent review by Manfredsson et al. [94] has highlighted that the therapeutic mechanism of action of GDNF is not fully well-defined, and that the degenerating brain of PD may be resistant to the neuroprotective potential of these proteins. The lack of clarity on the mechanism of action of GDNF may cause problems in appropriate model selection for preclinical therapeutic studies [94].
A second interesting NTF is the ciliary neurotrophic factor (CNTF), which is a member of the interleukin-6 family of cytokines. It has potent effects on the development and maintenance of the nervous system, as well as on cardiomyocytes, osteoblasts, immune cells, adipocytes, and skeletal muscle cells [95,96]. CNTF has been found to affect motor neuron survival in vitro, during development, after injury to motor neuron systems, and in genetic models of motor neuron degeneration [57], providing a rationale to develop CNTF as a treatment for ALS [56,57,62,66,67,68,69] and SCI [49,52], in which ventral motor neuron degeneration is extensive [57]. A drawback for CNTF administration is that it protects motor neurons from degenerative disease and injury, but also has some side effects, such as severe weight loss, hyperalgesia, coughing, muscle cramps, and pain [97]. Therefore, CNTF-related therapeutics will need to be designed to specifically target receptor mechanisms that protect motor neurons [98].
Apart from GDNF and CNTF, other known factors are the neurotrophins. This group consists of four members that share a common ancestral gene and have a similar structure. In the CNS, brain-derived neurotrophic factor (BDNF) is the major neurotrophin because of its abundant expression of tropomyosin receptor kinase B, also known as tyrosine receptor kinase B (TrkB) [99]. Studies of disease models of AD, in which BDNF was increased by using, for example, a lentivirus that expressed BDNF, showed that this factor is essential for multiple functions during adulthood, such as proper memory acquisition, memory retention, and cholinergic innervation [100,101]. BDNF is decreased within the brains, serum, and cerebro-spinal fluid (CSF) of patients with mild cognitive impairment and AD [101]. Also, low BDNF secretion in the serum of MS patients may be related to reduced neuroprotection [102,103]. As a result, low BDNF levels are expected to diminish the potential for remission in MS patients and induce the progressive phase of the disease [104]. To date, the potential beneficial effect of BDNF has been explored in several neurodegenerative and inflammatory diseases, such as animal models of AD, SCI, MS, and HD [21,54,99,105,106,107,108].
In the peripheral nervous system (PNS), nerve growth factor (NGF) is the dominant neurotrophin, which interacts on sympathetic and sensory neurons. In the CNS, NGF specifically provides trophic support to cholinergic neurons of the basal forebrain (BFCNs) that express TrkA (Figure 1), which would make it specifically interesting for AD [63,64,65,109,110]. NGF and its receptors, TrkA and p75, are known to play a bidirectional role between the immune and nervous systems. Recently, it has been extensively discussed that NGF plays a dual role in both anti- and pro-inflammatory response [111]. Moreover, cytokines, such as IL-1β, TNF-α, and IL-6, induce the overexpression of NGF [112].
Finally, neurotrophin-3 (NT-3) and neurotrophin-4/5 (NT-4/5) also have promising potential, albeit less studied than their counterparts. NT-3 is the third neurotrophic factor of the neurotrophin family, and, through activation of its tropomyosin-related kinase receptor C (TrkC) (Figure 1), it can modulate neuronal survival, support the differentiation of neurons, and stimulate the growth [113] and differentiation of new neurons and synapses [45]. Although this neurotrophin seems less popular, interesting in vivo studies have been done in various neurodegenerative diseases [39,45].
NT-4, also known as neurotrophin-5 (NT-5), is a neurotrophin that primarily signals via the TrkB receptor tyrosine kinase (Figure 1). The neurotrophins BDNF and NT-4 both bind to and activate TrkB receptors; however, they mediate different neuronal functions. The molecular mechanism of how TrkB activation by BDNF and NT-4 results in different outputs is not yet known. NT-4 is the least studied member of the neurotrophin family [53,114,115,116,117].
Unfortunately, the exact mechanism of NTFs is not yet fully understood. Nevertheless, research already reported the different NTF receptors and unravelled the pathways they activate to ensure the maintenance of cell growth, survival, development, and differentiation. BDNF, NGF, NT-3, and NT-4/5 bind to two families of receptors, namely, the tropomyosin kinase (Trk) receptors with high affinity, and with low affinity to the p75 receptor (Figure 1) [74,75,118]. Their actions are dependent on binding to the transmembrane receptor systems. Neurotrophins preferentially bind to specific receptors: NGF binds to TrkA, BDNF and NT-4 to TrkB, and NT-3 to TrkC [119]. However, there are a number of promiscuous interactions. All four neurotrophins can bind to the p75 receptor, and the association of p75 with Trk receptors can regulate the affinity of Trk receptors for each respective neurotrophin, allowing more control over ligand-receptor interactions within this system [120]. The Trk receptor binds with high affinity with NTFs to promote cell survival via phospholipase C-γ (PLC-γ), phosphoinositide-3-kinase (PI3K), and mitogen-activated protein kinase (MAPK) pathways that induce differentiation and survival via transcriptional events (Figure 1, green arrows). The MAPK pathway may be involved in ureteric branching during nephrogenesis and neurite outgrowth in the nervous system, but it also contributes to neuronal survival. The PI3K pathway is crucial for both neuronal survival and neurite outgrowth. The PLC-γ pathway regulates the intracellular level of Ca2+ ions by increasing the level of inositol (1,4,5) trisphosphate. Binding of NTFs to the low-affinity p75 receptor activates cell death via the JNK pathway. Activation of the JNK pathway similarly controls activation of several genes, some of which promote neuronal apoptosis. Neurotrophins are known to have a wide range of roles in the development and function of the nervous system. The characterisation of their receptors—the Trk receptor and p75 receptor—has significantly advanced research and enabled the characterisation of signalling pathways and the first steps to relate individual signalling pathways to specific developmental or functional roles of neurotrophins [119].
Today, we have learned to which receptors the various NTFs bind and what signalling pathways they activate. For instance, binding of CNTF to the CNTFRα receptor and two subunits, GP130 and leukaemia inhibitory factor (LIFRβ), activates the Janus kinase/signal transducer, an activator of transcription (JAK-STAT), MAPK, and PI3K pathways. The JAK-STAT pathway is associated with cell growth, survival, development, and differentiation (Figure 1, violet and blue arrows). Binding of GDNF to the GFRα receptor and tyrosine kinase RET receptor stimulates PLC-γ, MAPK, and PI3K (Figure 1, yellow arrows). RET activates various intracellular signalling cascades, which control cell survival, differentiation, proliferation, migration, chemotaxis, branching morphogenesis, neurite outgrowth, and synaptic plasticity [120]. Akt controls the activities of several proteins important for promoting cell survival, including substrates that directly regulate the caspase cascade, such as Bcl-2 agonist of cell death (BAD). Phosphorylated BAD prevents its pro-apoptotic activity (Figure 1, red inhibitory arrow). These different signaling pathways, which are activated by NTFs, work together to ensure normal neuronal function and to prevent neuronal cellular death (Figure 1).

3. Delivery of NTFs’ and Associated Challenges

3.1. Administration of NTF by Direct Infusion in the CNS

Various techniques have been used to get NTFs into the brain. The best known technique is direct intracerebroventricular (ICV) infusion. In particular, recombinant human (rh) GDNF and 125Iodine-labelled GDNF (125I-GDNF) have been shown to diffuse into the deep brain structures of rats [79,86], not only to significantly increase striatal and nigral dopamine (DA) levels, but also to increase hypothalamic DA levels, which could explain the decreased food and water consumption and body weight observed in in vivo experiments [87,88]. ICV injection of GDNF into 6-hydroxydopamine (6-OHDA)-treated rats, an animal model of PD, also appears to result in improved locomotor performance [87,88]. Furthermore, the ICV delivery route seems suitable for therapies that need to reach the BFCNs. Early and progressive degeneration of BFCNs contributes substantially to cognitive impairments of AD. Since BFCNs extend their axons through the hippocampus and neocortex, NGF administered in the lateral ventricle can act on the TrkA receptor to transmit trophic support signals to BFCNs. This approach has been shown to be particularly effective in preventing loss of BFCNs in rodents associated with injury and ageing [110,122,123]. However, the small volume of the rodent brain compared to the human brain raises important questions about the applicability of this technique in clinical studies. Therefore, ICV injections were also performed in non-human primates [90,91,92]. GDNF has been shown to produce significant improvements in motor activity in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-treated rhesus monkeys, a model of PD [89,90], and improvements in motor impairment and reductions in l-dopa-induced dyskinesia in marmosets [91]. In an autoradiographic study of the distribution of 125I-GDNF administered in the lateral ventricles of rhesus monkeys with a MPTP lesion, GDNF was not found to diffuse readily into the putamen. This finding contrasts with similar studies in rodents [79,86], suggesting that the success of ICV infusion in rodents might be a product of the smaller diffusion distance within their brain [87,88]. Moreover, the ICV delivery route was associated with serious side effects [110], such as hyperinnervation of cerebral blood vessels [123], hypophagia [110,122], Schwann cell hyperplasia with sprouting of sensory and sympathetic neurons [124], neuropathic pain [110], and dyskinesia [89,90,91], providing profound contra-indications for the applicability in clinical trials.
Because of these ICV-related side effects, the study by Tuszynski et al. [109] investigated whether intra-parenchymal infusion would be a well-tolerated way to administer NTFs to degenerating cholinergic neurons. In particular, intraparenchymal NGF infusion prevented degeneration of BFCNs, whereas glial responses were minimal in adult rats that underwent complete unilateral fornix transections, followed by intraparenchymal infusions of recombinant human NGF for a 2-week period. In addition, no apparent toxic effects of the infusions were observed, according to the researchers [109]. Other studies aimed to administer NTFs by a less invasive method. The group of Braschi et al. [21] tested whether intranasal (IN) administration of different concentrations of BDNF in AD11 transgenic mice, a model of AD, was able to rescue neuropathological and memory deficits. They found that IN administration of BDNF, but not with PBS, was adequate to completely rescue the performance of AD11 mice in both the object recognition test and the object context test. The strong improvement in memory performance in BDNF-treated mice was not accompanied by an improvement in AD-like pathology, amyloid-β (Aβ) load, tau hyperphosphorylation, and cholinergic deficiency [21]. Similarly, IN administration of NGF to Aβ peptide-expressing traumatic brain injury (TBI) rats, which are at risk of AD in later life, caused a marked reduction in Aβ42 deposits and restored motor and behavioural function [20]. Features such as non-invasive manipulations, rapid absorption rate, easy repetitive dosing, and reduction of non-target biodistribution make IN administration superior to the systemic and ICV routes of administration [19,20].
Finally, studies examined the effects of continuous intraputamenal administration of GDNF in both aged and MPTP-lesioned non-human primates [84,85,93]. Histological and biochemical analysis showed an increase in cell size and the number of dopaminergic neurons within the substantia nigra, as well as increased fibre density in the caudate nucleus, putamen, and globus pallidus. Primates with MPTP lesions showed improvements in the primate PD rating scale, whilst aged monkeys demonstrated improvement in general motor performance at high doses and increases in hand speed [84,85,93]. To assess the possible side effects of continuous administration of GDNF, a six-month toxicity study was conducted in rhesus monkeys. The results cast considerable doubt about the neuro-restorative potential of GDNF for the treatment of PD, given that they identified a number of pathological markers of toxicity, including reduced food intake and weight loss, meningeal thickening, and most concerning, multifocal cerebellar Purkinje cell loss [31]. Apart from the above-mentioned side effects, direct administration of NTFs into the brain also had some practical problems, such as invasiveness, BBB permeability [7,8,125], poor half-life, and rapid degradation [126]. This led to studies using cell therapy, where cells were modified to produce a specific protein.

3.2. Cells Modified to Express Neurotrophic Factors

During the last years, different cell types have been utilized to deliver NTFs to the injured sites. Mesenchymal stromal cells (MSCs) are described as adherent, fibroblast-like cells with prominent proliferation capacity [42,50,51,127]. Because of their low immunogenicity (low expression levels of major histocompatibility complex (MHC) class II), MSCs can survive after administration [128]. The existence of such capabilities makes MSCs a safe, tolerable, and efficient biological vector for the generation and delivery of therapeutic agents, such as NTFs, to the target sites [42,50,51]. Furthermore, different routes of administration were used to administer the modified MSCs, resulting in different outcomes. In a study by Suzuki et al. [51], human MSCs (hMSCs), derived from neonatal bone marrow aspirates which were modified to express GDNF, were administered intramuscularly as a "Trojan horse" to superoxide dismutase (SOD1)G93A rats, a rat model of familial ALS, to deliver GDNF to the terminals of motor neurons and to skeletal muscle. hMSC-GDNF survived in the muscle, secreted GDNF, and significantly increased the number of neuromuscular connections and motor neuron cell bodies in the spinal cord in the mid-stage of the disease. Moreover, hMSC-GDNF significantly slowed down disease progression [51]. In addition, several improvements have been reported when CNTF- [52], NT-3- [53], and BDNF-modified [54] MSCs were administered directly into the spinal cord of SCI rats, such as improvement in behavioural scores, motor function, axonal regeneration, and neuronal survival [52,53], and restoration of diaphragm muscle function [54]. Positive results with MSCs expressing NTFs were also observed after intravenous (iv) administration. A remarkable recovery of neuronal function was observed and demyelination was significantly reduced in EAE mice: the cumulative clinical scores were significantly decreased, and the disease onset was statistically delayed, after iv MSC-CNTF [55] and MSC-BDNF administration [105]. Moreover, BDNF-expressing MSCs can also reduce striatum atrophy and increase neurogenesis in HD mouse models [22]. In summary, MSCs represent a promising tool for cell therapy. There is currently much interest in the use of MSCs for the treatment of neurodegenerative diseases. There are several studies using the innate trophic support of MSCs or increased support by NTFs, such as the administration of BDNF, CNTF NTF-3, or GDNF to the CNS to support damaged neurons, using genetically engineered MSCs as delivery tools. Biosafety could be a potential difficulty in cell therapies when using genetically engineered MSCs. The random integration of vectors with genes for neurotrophic or other factors may pose the risk of insertional integration. However, homologous recombination and targeted gene transfer are advancing rapidly.
Neural stem cells (NSCs) are also used as a NTF vector, resulting in several positive effects. NSCs are characterised as multipotent and self-renewing cells with the capacity to differentiate into mature neurons and neuroglia cells [23,24,25,26]. In a rodent model of cervical SCI, it was shown that GDNF-expressing human induced pluripotent stem cell-derived NSCs (hiPSC-NSCs) showed greater differentiation into a neuronal phenotype than unmodified hiPSC-NSCs [27]. Furthermore, several improvements were seen with NSCs expressing GDNF in SOD1G93A ALS rats, when administered in the motor cortex [24] and in the spinal cord [25]. The results show improved survival, as well as enhanced proliferative and neuroprotective properties [24,25]. Moreover, human GDNF-expressing NSCs duly migrated to the disease site and integrated into the CNS after administration into the spinal cord of SOD1G93A ALS rats [25]. In addition, it has been shown that GDNF-expressing NSCs administration in the lateral ventricle promotes axonal regeneration and remyelination in chronic EAE rats [26].
A number of studies have indicated that immune cells are also useful as therapeutic biosystems to deliver various molecules into target areas [28,29]. Among the subsets of immune cells, macrophages are the most suitable target cells, as they are activated soon after the onset of the inflammatory response, can cross the BBB, and move to sites of neuronal degeneration [28,29]. In this regard, the monocyte-macrophage lineage could represent an efficient cellular system to deliver NTFs at the site of injury within the CNS. To support this hypothesis, Biju et al. used ex vivo transduced bone marrow-derived macrophages to deliver GDNF [28]. Axonal regeneration and retention of tyrosine hydroxylase (TH+) neurons were observed in both the striatum and substantia nigra regions [28]. Moreover, GDNF-expressing macrophages could successfully cross the BBB and deliver GDNF into the neuro-generated DA neurons after systemic administration [29].
Finally, other cells, such as fibroblasts, were also used as vectors to deliver NTFs. Specifically, fibroblasts modified to express BDNF were inoculated into SCI sites in rats, and these caused regenerative and sprouting responses at the sites of injury [106,107,108]. Similarly, genetically modified baby hamster kidney (BHK) cells and primary cells expressing NGF showed that they were able to rescue cholinergic function in damaged neurons in ageing models of both rodents and non-human primates [129,130,131]. More interestingly, the implanted cells maintained NGF secretion for at least 8 months in primate brains and did not cause the adverse side effects observed in studies with direct administration [132,133,134].
To date, research advances in cell-based therapies offer promising methods for treating neurodegenerative diseases. Although much work remains to be done, the increasing focus on preclinical studies and the recent translation of some of these therapies into clinical trials have paved the way for further progress. The use of modified cells expressing NTFs is likely to play a key role in future clinical strategies to treat neurodegenerative diseases by replacing dysfunctional neurons and providing neuroprotective functions. As mentioned earlier, a potential drawback that remains today is the biosafety.

3.3. Viral Delivery of Neurotrophic Factors

Viral vector-mediated gene delivery might be a more optimal approach instead of the techniques that have been previously described. Virus administration would permanently alter the cells’ ability to make its own NTF, requiring a single injection at the site of administration, rather than multiple injections [82,83,135,136], and eliminating the cumbersome cell preparation associated with the cell transfer technique [30,32,33,34,36,82,83,135,136,137].
Nakajima et al. [30] reported that injection of adenovirus (AV)-BDNF into bilateral sternomastoid muscles transferred vectors to the damaged sites, via retrograde transport using spinal accessory motor neurons, in SCI rats. The AV-BDNF was able to reach the spinal cord and reduce apoptotic signalling in neurons and oligodendrocytes [30]. Likewise, the application of retrograde AV-BDNF in bilateral sternomastoid muscles of chronically compressed SCI mice led to the recovery of oligodendrocyte progenitors and neurofilament expression via the axons of spinal accessory nerves [32]. However, there are some drawbacks using AV vectors, including immunogenicity, replicability, and the small insertion size of the vectors [30,32].
To date, adeno-associated virus (AAV)-mediated gene transfer of GDNF has been used and evaluated in a number of studies in rodents and primates, particularly for PD [136], HD [82,83], and SCI [33]. Eslamboli et al. [136] showed that unilateral intrastriatal injection of AAV-GDNF, resulting in the expression of high levels of GDNF in the striatum, induced a significant bilateral increase in tyrosine hydroxylase protein levels and DA turnover in a 6-OHDA lesion in marmosets. In addition, AAV-GDNF-treated rats scored better on a blinded semi-quantitative neurological scale compared to rats receiving the control AAV- Green Fluorescent Protein (GFP), which was supported by histological analyses [83]. Interestingly, Fouad et al. [33] reported that rats, with complete thoracic SCI, that received combined treatment, including self-complementary AAV-BDNF and NT-3 administration in the spinal cord, showed not only improved axonal regeneration, but also improved motor function of the hind limbs [33]. AAV vectors offer many of the same advantages as AV vectors, including a wide host-cell range and a relatively high transduction efficiency. In addition, AAV vectors do not express their own proteins and, therefore, would not elicit an immune response, making the technique even more attractive. However, the major drawback is the limited cloning capacity of the vector, which restricts its use in the gene delivery of large genes [33,82,83,136].
Next to AV- and AAV- mediated NTF delivery, viral delivery of GDNF by lentivirus (LV) reversed motor deficits and prevented nigrostriatal degeneration in MPTP-treated monkeys [137]. The delivery of LV expressing GDNF to AD mice models enhanced learning and memory function, while simultaneously improving the cognition capacity [34]. In addition, the group of Pereira de Almeida et al. [138,139] conducted two studies using tetracycline-regulated LV-mediated delivery of CNTF in a quinolinic acid (QA) rat model of HD. The 2001 study [138] showed that the extent of striatal damage was significantly reduced in the CNTF-treated rats, and the volume of the lesion was significantly reduced [138]. In 2002, they reported CNTF′s dose-dependent effects [139]. Remarkably, LV-based administration has numerous advantages, such as long-term transgene expression, low inflammation rate, and large-size gene insertion [35,36,140]. Despite these advantages, in some cases, oncogenic mutation may occur after integration of the LV gene into the host cell genome. This is cited as the main concern of safety in in vivo conditions.

3.4. Biomaterials to Deliver Neurotrophic Factors

Several of the above-mentioned strategies to deliver NTF to the site of injury in the spinal cord or brain, such as direct delivery, genetically engineered cells, and viral vectors, have a number of drawbacks, including viral vector spread beyond the target area, uncontrolled transgene expression, and immune rejection of transplanted cells. Therefore, there is a growing interest in using biomaterials as vehicles to deliver NTFs. Natural biomaterials are biocompatible, biodegradable, have remodelling advantages and a lower toxicity rate [141], while synthetic biomaterials have a more favourable mechanical and thermal resistance, no immune response capacity, and can be produced on large scales [37,38].
A recent study by Zhijiang et al. [141] used the natural biomaterial methylcellulose (MC), combined with hyaluronic acid (HAMC) hydrogel modified with the peptide KAFAK-LAARLYRKALARQLGVAA (KAFAK) and BDNF. They injected these into a lesion area of SCI rats and showed that locomotor function and axonal regeneration improved 8 weeks after SCI [141]. A similar study with NT-3 also showed that HAMC could release NT-3 for 28 days. The persistence of NT-3 in the target areas confirmed the regeneration and expansion of axons, without induction of the astroglial response, which can cause an inflammatory reaction [39]. Furthermore studies have used other natural bio-materials, such as bioactive scaffolds, to create a microenvironment conducive to endogenous regeneration of neuronal tissue in the SCI site. In particular, gelatin sponge scaffold, silk fibroin, chitosan, or a more developed multichannel nanofibrous gelatin scaffold have been used. These scaffolds were integrated into NT-3, with or without NSCs [44], adipose-derived stem cells [43], or MSCs [45,142]. The in vivo experiments have significantly improved neuronal differentiation, synaptic connection, and axonal remyelination, with reduced local inflammation at the SCI sites following bioactive scaffold implantation with NT-3. In addition the treatment has shown significant improvement in locomotor functionality [40,43,44,45,142].
Poly-lactide-co-glycolide (PLG) is one of the most frequently used synthetic biomaterials for drug delivery, due to its controlled and sustained release properties, low toxicity, and biocompatibility with tissue and cells [46,47]. PLG has been widely used as a material for spinal cord repair or peripheral nerve conduits [47]. Khalin et al. found that iv injection of poloxamer 188 (PX)-coated PLG nanoparticles with BDNF (PLG-BDNF) in TBI mice restored cognition and showed that this system is eligible to cross the BBB and deliver BDNF into the brain of the TBI model [38]. Furthermore, several studies with PLG-BDNF in animal models of SCI observed robust axon growth and remyelination 6 months after initial injury [39,47,48]. These positive findings of PLG-BDNF were not confirmed with CNTF. The latter would not be sufficient in vivo to promote oligodendrocyte remyelination in the glial-depleted environment of unilateral ethidium bromide lesions [49]. Similar to the PLG-BDNF results in SCI rats, poly N-isopropylacrylamide (PNIPAAm) with BDNF improved the axonal regeneration in SCI rats [37]. Finally, intrathecal infusion of N-terminal pegylated (PEG) BDNF (PEG-BDNF) was also used in an attempt to increase NTF release [143]. The authors showed that the PEG-BDNF was able to reach the spinal cord and that its expression was induced in that area. However, they could not observe an improved axonal response or recovery of motor function, which suggests that the amount of BDNF was insufficient [143].
As mentioned earlier, most NTFs have difficulties passing through the BBB and are, therefore, delivered directly into the brain in animal models and some clinical trials with patients using expensive and risky intracranial surgery [70,71,72]. The efficiency of delivery and the poor distribution of some NTFs in the brain are considered the main problems behind their modest effects in clinical trials. There is a great need for NTFs that can be administered systemically to avoid intracranial surgery. Nanoparticles (NPs) can be used to stabilise NTFs and facilitate their transport through the BBB [144]. For example, one study used plasmid DNA NPs encoding human GDNF (pGDNF) that were administered IN to a rat model of PD [145]. The amphetamine-induced rotational behaviour was reduced, and dopaminergic fibre density and cell counts in the lesioned substantia nigra and nerve terminal density in the lesioned striatum were significantly preserved in rats given IN pGDNF [145].

4. Clinical Trials with Neurotrophic Factors

In addition to studies in animal models, there were also studies in humans, in which NTFs were used for the purpose of regeneration. The first clinical trials with NTFs in ALS patients applied systemic administration of CNTF, while the protein did not readily cross the BBB and consequently did not reach a detectable concentration in the central parenchyma [56,57,66,67,68,69]. Side effects, including inflammation and cachexia, have been recorded after systemic administration, which were severe enough to terminate phase II/III clinical trials with CNTF in ALS patients [56,57,66,67,68,69] (Figure 2). This led to the NTFs being administered directly into the brain in subsequent clinical studies. In particular, GDNF was administered by monthly bolus injections into the cerebral ventricles of PD patients. No beneficial clinical effects were seen, whereas side effects, such as nausea, loss of appetite, tingling, Lhermitte sign, intermittent hallucinations, and depression, were reported. In addition, there was no evidence of the restoration of dopamine fibers in the striatum [70,71]. Bolus injection into the parenchyma exposed the patient to a higher risk of tissue trauma and denied the clinician the means to finetune and optimize dose delivery (Figure 2). The clinical phase I safety trial of Nikunj et al. delivered GDNF directly into the putamen of five patients with PD [72]. Afterwards, they continued to follow these patients for two years and concluded that direct intraputamenal GDNF infusion in patients with PD is safe, can be tolerated for two years, and leads to significant symptomatic improvement [73]. Interestingly, the same group performed another randomized, controlled, blinded clinical trial in order to confirm the initial clinical benefits. However, this trial did not confer the predetermined level of clinical benefit to patients with PD, despite increased (18)F-dopamine uptake [58].
The macro-encapsulation technique was a more sophisticated method. This technique was first conducted with CNTF in rats and non-human primate models of HD [59,60]. In brief, BHK cells engineered to synthesize and release large amounts of NTF, such as CNTF, have been introduced into a tube formed by a semipermeable membrane. The pores of this membrane are sized so that proteins can cross freely, whereas larger proteins (e.g., antibodies) and cells cannot. Due to the positive results of this technique, reduced side effects, and the ability of BHK-hCNTF to protect neurons from degeneration and restore neostriatal function in animal models [59,60], the group of Bachoud-Le Âvi et al. [61] and Aebischer et al. [62] used this macro-encapsulation technique in a phase I study in ALS and HD patients (Figure 2). In particular, a capsule was introduced into the lateral ventricle of six patients with HD [61] and ALS [62], using stereotactic neurosurgery. No signs of CNTF-induced toxicity were observed. According to the results, this phase I study demonstrated the safety, feasibility, and tolerability of this gene therapy procedure, but the heterogeneous cell survival indicates the need to improve a more uniform response. Furthermore, no clinical benefit was observed in any of the treated subjects, which could partly be due to the limited diffusion of CNTF through the ventricular wall to the adjacent putamen [61,62], similar to the limited diffusion of GDNF after ICV injection in non-human primates [89,90,91,92].
Finally, some clinical studies have used cells or viral vectors to bring the NTFs into the brain. Mark Tuszynski′s team [63] surgically implanted autologous fibroblasts, which were modified to secrete mature human NGF, into the basal forebrain of eight early stage AD patients. The mean Mini-Mental Status Examination (MMSE) scores showed an average decrease of 51% over a 22-month period, and an even greater decrease over 6 to 18 months. Moreover, there were cognitive improvements, and post-mortem analysis confirmed that there was NGF expression in the cell grafts and that cholinergic axons showed outgrowth. Overall, this study presented the first clinical evidence that NGF administration can provide therapeutic benefit, without side effects usually associated with NTF administration, such as nausea, loss of appetite, tingling, hallucinations, and depression [63]. Because AAV serotype 2 (AAV2)-NGF vectors represent a more convenient and less expensive method of gene delivery and resulted in long-term gene expression in non-human primate brains [64], Tuszynski et al. conducted a second phase 1 clinical trial on 10 patients with AD (Figure 2). Here, AAV2-NGF was injected in vivo into the basal forebrain region, genetically modifying cells of the brain itself, rather than employing grafts of autologous cells, as employed in the phase 1 ex vivo study [65]. This study showed that responses to NGF persist for up to 10 years after gene transfer. No adverse pathological effects were observed over a 7-year period, supporting the safety and rationale for the expanded clinical programs underway in AD, PD, and other neurological indications [65].

5. Challenges and Future Perspective of the Use of NTFs in Neurodegenerative Diseases

Neurodegenerative diseases that cause acute or chronic damage to neurons and glial cells represent a major socio-economic burden and loss of quality of life for millions of patients and their families worldwide [3]. With an ageing population, the number of patients will further increase [4], creating an urgent need for therapeutic strategies that can reverse or stop the degenerative process. NTFs, as discussed in this review, are important factors in both development and adulthood, and each is required by certain subsets of neurons for optimal function. From the results, GDNF would be of particular interest for PD, due to its high specificity for dopaminergic neurons [84,85,91,93,124]. In addition, CNTF seems important, especially for ALS [56,57,62,66,67,68,69] and SCI [49,52], due to its potent effects on motor neuron survival, after injury to motor neuron systems and in genetic models of motor neuron degeneration. NGF specifically provides trophic support to cholinergic neurons of the BFCNs that express TrkA, which would make it of particular interest for AD [63,64,65,109,110]. The potential beneficial effect of BDNF has been studied in several neurodegenerative and inflammatory diseases, including animal models of AD, SCI, MS and HD [21,54,99,105,106,107,108]. As well, neurotrophin-3 (NT-3) and neurotrophin-4/5 (NT-4/5) also have promising potential, however, they have been less studied than their counterparts. Decreased levels of one or more of these proteins may be responsible for at least some of the symptoms of AD, PD, ALS, HD, and MS [78,101,103,104,146,147]. Therefore, these factors have been investigated as a potential neuro-healing therapy in preclinical and/or clinical studies (Figure 2). In particular, NTFs can be delivered via direct infusion, cells modified to (over)express these factors, viral delivery, or biomaterials (Figure 2).
There are strong arguments showing that an increase of NTFs-delivery to degenerating neurons could be a powerful way to restore neuronal function, but the delivery of these NTFs into the brain seems challenging [148]. In particular, diseases of the CNS are known to be difficult to treat because of the presence of the BBB, which makes it virtually impossible for large proteins and complex connections to enter the brain from the blood [149,150,151]. The possibility that NTFs can cross the BBB is quite controversial [148]. For example, some authors state that it is not clear whether BDNF can easily pass the BBB [152], whereas others indicate that BDNF is able to do so [153]. Molinari et al. [153] published a recent paper on the possibility of using exogenous BDNF as a therapeutic approach in neurodegenerative diseases. His work showed, in in vitro experimental models, that a low BDNF dose can cross both the intestinal and BBB barrier [153]. An alternative way and more recent technique in the neuroscience to get large molecules across the BBB would be the use of low-frequency focused ultrasound combined with microbubbles. This non-invasive and reversible technique [154,155] can achieve a transient safe opening of the BBB [155,156]. Successful preclinical studies have already been performed with growth factors, antibodies, genes, viral vectors, and nanoparticles in rodent models of AD and PD [154,156,157]. Recent small clinical studies support the safety and feasibility of this strategy in patients [158]. Further research is needed to determine the safety when the MRI-guided BBB opening is used to improve the delivery of newly developed molecular therapies [156,157].
Furthermore, an upcoming way to improve BBB penetration after parenteral systemic administration is the use of chemical modification or antibody conjugation of native NTFs. Specifically, a covalent modification of NGF with the polyamine putrescine resulted in improved plasma pharmacokinetics and BBB permeability in rats, as compared with native NGF [159]. Moreover, a study by Wu and Pardridge [160] attached biotinylated polyethylene glycol-modified-BDNF to a monoclonal antibody against the transferrin receptor that was linked to streptavidin. This resulted in the ability of the chimeric molecule to bind to the transferrin receptor, which is abundant on brain endothelial cells, and subsequently to undergo receptor-mediated transcytosis through the BBB [160]. Although modification/conjugation strategies are promising for the CNS delivery of peripherally administered NTFs, a major challenge to the clinical implementation of such strategies is the anticipated difficulty in producing large quantities of pharmaceutical-grade preparations and in targeting the products to specific CNS areas [161].
Beyond the BBB permeability, it should be taken into account that, in general, transplanted cells manipulated to (over)express proteins may differentiate into undesirable cell types, with the possibility of tumour formation, risks of host rejection, and inflammation [162,163,164], limiting the widespread use of these manipulated cells, despite their advantages [164]. Viral vector-mediated delivery may already overcome some of the above-mentioned challenges. In particular, virus administration could permanently alter the cells′ ability to make their own NTFs, consequently requiring only a single injection and, thereby, decreasing the invasiveness of the treatment [165]. However, controlling the production of NTF proteins and terminating their expression warrants further research, since cytotoxic effects on host cells and inflammatory responses were seen after the development of self-inactivating viral vectors for in vivo applications [166].
The last delivery method discussed in this review is the application of biomaterials. In general, this method requires a less invasive manipulation with delivery of large amounts of NTFs to the damaged sites. When selecting the delivery method, a number of properties, such as degradability, safety, non-toxicity, and adaptability to release, must be taken into account [167]. Furthermore, biomaterials used for CNS regeneration should be injectable. It should be remembered that natural biomaterials can be immunogenic, but not toxic [167,168]. Synthetic components, on the other hand, do not cause inflammation, but may provoke cytotoxicity [167,169]. A recent technique, which is successfully developed for clinical use in neurodegenerative diseases, includes targeted nano-carriers for recombinant growth factors, therapeutic antibodies, enzymes, synthetic peptides, cell-penetrating peptide-drug conjugates, and RNAi sequences [170]. To enable challenging applications of nano-medicine and precision medicine in the treatment of neurodegenerative diseases, more in-depth research into bio-molecular delivery via nano-carriers for neuronal targeting and repair is needed. According to a recent review by Yu Wu et al., the successful use of macromolecular bio-therapeutics in clinical developments for neuronal regeneration will be aided by recent strategies to improve their bioavailability [170].
It is worth mentioning that many of the challenges discussed above may be overcome by small molecules that target the receptor for the NTF, instead of introducing the NTF itself. The development of small molecule mimetics, with an intrinsic neurotrophic activity and an improved pharmacokinetic profile, is a promising research area. This would allow for specific activation of only one type of receptor, such as TrkA or TrkB and not p75, or vice versa, potentially alleviating the side effects. Interestingly, it has recently been shown that neuro-inflammatory cytokines, such as TNF-α, downregulate both the mRNA and protein levels of TrkA, together with an increase of p75 mRNA expression [171]. This could shift NGF signalling from a neuroprotective to a neurotoxic role, showing that a specific binding of a certain receptor is interesting, especially during pathological (inflammatory) conditions [171]. The use of NTF therapy or NTF mimetics in combination with a TNF-α inhibitor could also be an interesting option. Because several synthetic TNF- α inhibitors induce serious adverse effects in various inflammatory diseases, patients and researchers have recently turned their attention to natural medicines, especially phytochemicals. Phytochemicals targeting TNF- α can significantly improve disease states with fewer side effects, according to the review by Subedi et al. [172]. Several experimental studies have also shown that the administration of bioactive molecules in low doses is effective to obtain pure biological effects with low risk of side effects [153,173].
The discovery and use of peptide mimetics [174] and small molecule ligands for the Trk receptors [175] have attracted considerable interest. Therefore, relatively stable peptide mimetics of NGF have, amongst others, been produced [176]. These analogues may be less immunogenic, more resistant to proteolytic degradation, and able to cross blood–tissue barriers, as compared with their parent molecules. These ligands may be more stable and less expensive to produce than recombinant proteins, and may eventually provide acceptable oral bio-availabilities unattainable with native NTFs. The use of a potent peptide BDNF mimetic that activates TrkB was shown to promote neuronal survival in embryonic sensory neurons of the dorsal root ganglion [177]. Small-molecule BDNF mimetics also have high potency and specificity against TrkB, and can promote neuronal survival, while also inducing differentiation and synaptic function in cultured hippocampal neurons [178]. When administered to mouse models of AD, HD, and PD, the small molecule could rescue cell death to the same extent as the full-length protein BDNF [178]. A number of clinical trials are also currently being conducted with NTF mimetics [175]. Results from these trials, especially in terms of side effects and efficacy, will broaden and improve NTF-based therapy for the treatment of neurodegenerative diseases with acute or chronic neuronal and glial damage.
Although NTF-based therapy has great potential, the greatest uncertainty is whether such an approach by itself is sufficient to halt and reverse the progression of neurodegenerative diseases. Due to the failures of monotherapy in the past, it may be interesting to use combination therapy, instead of the ′single magic bullet′ approach, to address the various disease-causing mechanisms simultaneously. In particular, a combination of several NTFs could be better than using a single NTF for neurodegenerative diseases. For example, studies have shown that BDNF and NT-3, when used in combination, are more effective than either factors alone in increasing the growth of host axons into transplanted spinal cord tissue following spinal cord hemisection in adult rats [33,179]. These synergistic effects may allow combinations of factors to be used at smaller doses than those required of any one factor used alone, diminishing adverse effects and potential for immunogenicity.
Moreover, combination therapy may be particularly useful in the treatment of CNS diseases in which there are multiple neuronal types affected, so that a NTF with maximal activity on a particular cell type can be administered together with another that acts on another cell type. For example, the capacity of NGF to stimulate cholinergic basal forebrain cells is also enhanced by BDNF, which can additionally potently stimulate dopaminergic cells in the midbrain [180]. We can, therefore, envision that a combined use of NTFs may work synergistically to restore neuronal function.
Besides NTFs, a number of other biological agents have emerged that show regenerative properties in neurodegenerative diseases, such as vascular endothelial growth factor (VEGF) [181,182,183,184,185,186,187,188], insulin-like growth factors (IGFs) [189,190,191,192,193,194,195,196,197,198,199,200,201,202,203], the cellular communication network (CCN) family [204,205,206,207,208], and erythropoietin (EPO) [209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234], with varying, but also promising results.

6. Conclusions

To date, several NTF distribution vectors and systems have been applied to deliver exogenous NTFs into the CNS, with variable results. In most cases, the translational capacity from bench to bedside was limited. The challenges that currently need to be overcome include the amount of NTFs released, BBB permeability if administered peripherally, the invasiveness of the delivery route, the half-life of the vehicle, and the occurrence of possible side effects. The combination of all these challenges is probably the reason why the application of NTFs has, so far, not been effective for the long-term regeneration of target tissues, especially in the brain. In addition, beyond the use of a single NTF, combination therapies, targeting multiple pathways or using smaller molecules, such as NTF mimetics, would be a more effective treatment option in neurodegenerative diseases. Nevertheless, it is important to continue research into the optimization of cellular-, viral vector-, and biomaterial systems to provide standards for clinical applications.

Author Contributions

Conceptualization, Y.E.O., I.W. and N.C.; methodology, Y.E.O., I.W. and N.C.; formal analysis, Y.E.O.; investigation, Y.E.O.; resources, I.W. and N.C.; data curation, Y.E.O.; writing—original draft preparation, Y.E.O., I.W. and N.C.; writing—review and editing, Y.E.O., B.W. and J.V.d.B.; visualization, Y.E.O., I.W. and N.C.; supervision, I.W. and N.C.; funding acquisition I.W. and N.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by BOF DOCPRO4 2020 (grant no. 42551) of the Special Research Fund (BOF) from the University of Antwerp, Belgium. Further support was provided through the Methusalem Funding Program from the University of Antwerp, and by the Belgian Charcot Foundation. In addition, this work received funding from the European Union’s Horizon 2020 research and innovation program under grant agreement 779316 (ReSToRe). Furthermore, Y.E.O holds a doctoral fellowship from the University of Antwerp (grant number: DOCPRO42551). J.V.d.B. is funded by the Belgian Charcot Foundation (grant number: FCS-2020-JVDB7). B.W. is supported by FWO-TBM (grant number T001121N), Start2Cure Foundation and the Queen Elisabeth Medical Foundation for Neurosciences. In addition, the funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

Y.E.O., J.V.d.B., N.C. and I.W. declare no conflict of interest. B.W. received honoraria for acting as a member of Scientific Advisory Boards for Almirall, Biogen, Celgene/BMS, Merck, Janssen, Novartis, Roche, Sandoz, Sanofi-Genzyme and speaker honoraria and travel support from Biogen, Celgene/BMS, Merck, Novartis, Roche, Sanofi-Genzyme; research and/or patient support grants from Biogen, Janssen, Merck, Sanofi-Genzyme., Roche. Honoraria and grants were paid to UZA/UZA Foundation.

References

  1. Metcalfe, S.M.; Bickerton, S.; Fahmy, T. Neurodegenerative Disease: A Perspective on Cell-Based Therapy in the New Era of Cell-Free Nano-Therapy. Curr. Pharm. Des. 2017, 23, 776–783. [Google Scholar] [CrossRef] [Green Version]
  2. Lindvall, O.; Kokaia, Z. Stem cells in human neurodegenerative disorders—Time for clinical translation? J. Clin. Investig. 2010, 120, 29–40. [Google Scholar] [CrossRef] [Green Version]
  3. Przedborski, S.; Vila, M.; Jackson-Lewis, V. Series Introduction: Neurodegeneration: What is it and where are we? J. Clin. Investig. 2003, 111, 3–10. [Google Scholar] [CrossRef] [Green Version]
  4. Havard Neuro Discovery Center. The Challenge of Neurodegenerative Diseases in an Aging Population. Trends Sci. 2017, 22, 6_92–6_93. [Google Scholar] [CrossRef] [Green Version]
  5. Volkman, R.; Offen, D. Concise Review: Mesenchymal Stem Cells in Neurodegenerative Diseases. Stem Cells 2017, 35, 1867–1880. [Google Scholar] [CrossRef] [Green Version]
  6. Frozza, R.L.; Lourenco, M.V.; de Felice, F.G. Challenges for Alzheimer’s disease therapy: Insights from novel mechanisms beyond memory defects. Front. Neurosci. 2018, 12, 37. [Google Scholar] [CrossRef]
  7. Ebrahimi, Z.; Talaei, S.; Aghamiri, S.; Goradel, N.H.; Jafarpour, A.; Negahdari, B. Overcoming the blood-brain barrier in neurodegenerative disorders and brain tumours. IET Nanobiotechnology 2020, 14, 441–448. [Google Scholar] [CrossRef]
  8. Zlokovic, B.V. The Blood-Brain Barrier in Health and Chronic Neurodegenerative Disorders. Neuron 2008, 57, 178–201. [Google Scholar] [CrossRef] [Green Version]
  9. Weil, Z.M.; Norman, G.J.; Devries, A.C.; Nelson, R.J. The Injured Nervous System: A Darwinian Perspective. Prog. Neurobiol. 2008, 86, 48–59. [Google Scholar] [CrossRef] [Green Version]
  10. y Cajal, S.R. Cajal’s Degeneration and Regeneration of the Nervous System; Oxford University Press: Oxford, UK, 1991. [Google Scholar] [CrossRef]
  11. Gianvito, M. How the brain repairs itself: New therapeutic strategies in inflammatory and degenerative CNS disorders. Lancet Neurol. 2004, 3, 372–378. [Google Scholar] [CrossRef]
  12. Fawcett, J.W. The Paper that Restarted Modern Central Nervous System Axon Regeneration Research. Trends Neurosci. 2018, 41, 239–242. [Google Scholar] [CrossRef]
  13. Monje, M.L.; Toda, H.; Palmer, T.D. Inflammatory Blockade Restores Adult Hippocampal Neurogenesis. Science 2003, 302, 1760–1765. [Google Scholar] [CrossRef]
  14. Kerschensteiner, M.; Raineteau, O.; Mettenleiter, T.C.; Schwab, M.E. The injured spinal cord spontaneously forms a new intraspinal circuit in adult rats. Nat. Neurosci. 2004, 7, 269–277. [Google Scholar]
  15. Jessberger, S. Neural repair in the adult brain. F1000 Res. 2016, 5, 169. [Google Scholar] [CrossRef] [Green Version]
  16. Maria, A.; Unit, T.; Raffaele, S. Adaptive functional changes in the cerebral cortex of patients with nondisabling multiple sclerosis correlate with the extent of brain structural damage. Ann. Neurol. 2002, 51, 330–339. [Google Scholar] [CrossRef]
  17. Mews, I.; Bergmann, M.; Bunkowski, S.; Gullotta, F.; Brück, W. Oligodendrocyte and axon pathology in clinically silent multiple sclerosis lesions. Mult. Scler. 1998, 4, 55–62. [Google Scholar] [CrossRef]
  18. Gemma, C. Neuroimmunomodulation and aging. Aging Dis. 2010, 1, 169–172. [Google Scholar] [CrossRef] [Green Version]
  19. Vaka, S.R.K.; Sammeta, S.M.; Day, L.B.; Murthy, S.N. Delivery of nerve growth factor to brain via intranasal administration and enhancement of brain uptake. J. Pharm. Sci. 2009, 98, 3640–3646. [Google Scholar] [CrossRef] [Green Version]
  20. Tian, L.; Guo, R.; Yue, X.; Lv, Q.; Ye, X.; Wang, Z.; Chen, Z.; Wu, B.; Xu, G.; Liu, X. Intranasal administration of nerve growth factor ameliorate β-amyloid deposition after traumatic brain injury in rats. Brain Res. 2012, 1440, 47–55. [Google Scholar] [CrossRef]
  21. Braschi, C.; Capsoni, S.; Narducci, R.; Poli, A. Intranasal delivery of BDNF rescues memory deficits in AD11 mice and reduces brain microgliosis. Aging Clin. Exp. Res. 2021, 33, 1223–1238. [Google Scholar] [CrossRef]
  22. Pollock, K.; Dahlenburg, H.; Nelson, H.; Fink, K.D.; Cary, W.; Hendrix, K.; Annett, G.; Torrest, A.; Deng, P.; Gutierrez, J.; et al. Human mesenchymal stem cells genetically engineered to overexpress brain-derived neurotrophic factor improve outcomes in huntington’s disease mouse models. Mol. Ther. 2016, 24, 965–977. [Google Scholar] [CrossRef]
  23. Mendes-Pinheiro, B.; Teixeira, F.G.; Anjo, S.I.; Manadas, B.; Behie, L.A.; Salgado, A.J. Secretome of Undifferentiated Neural Progenitor Cells Induces Histological and Motor Improvements in a Rat Model of Parkinson’s Disease. Stem Cells Transl. Med. 2018, 7, 829–838. [Google Scholar] [CrossRef] [Green Version]
  24. Thomsen, G.M.; Avalos, P.; Ma, A.A.; Alkaslasi, M.; Cho, N.; Wyss, L.; Vit, J.P.; Godoy, M.; Suezaki, P.; Shelest, O.; et al. Transplantation of Neural Progenitor Cells Expressing Glial Cell Line-Derived Neurotrophic Factor into the Motor Cortex as a Strategy to Treat Amyotrophic Lateral Sclerosis. Stem Cells 2018, 36, 1122–1131. [Google Scholar] [CrossRef] [Green Version]
  25. Suzuki, M.; McHugh, J.; Tork, C.; Shelley, B.; Klein, S.M.; Aebischer, P.; Svendsen, C.N. GDNF secreting human neural progenitor cells protect dying motor neurons, but not their projection muscule, in a rat model of familial ALS. PLoS ONE 2007, 2, e689. [Google Scholar] [CrossRef]
  26. Gao, X.; Deng, L.; Wang, Y.; Yin, L.; Yang, C.; Du, J.; Yuan, Q. GDNF Enhances Therapeutic Efficiency of Neural Stem Cells-Based Therapy in Chronic Experimental Allergic Encephalomyelitis in Rat. Stem Cells Int. 2016, 2016, 1431349. [Google Scholar] [CrossRef] [Green Version]
  27. Khazaei, M.; Ahuja, C.S.; Nakashima, H.; Nagoshi, N.; Li, L.; Wang, J.; Chio, J.; Badner, A.; Seligman, D.; Ichise, A.; et al. GDNF rescues the fate of neural progenitor grafts by attenuating Notch signals in the injured spinal cord in rodents. Sci. Transl. Med. 2020, 12, eaau3538. [Google Scholar] [CrossRef]
  28. Biju, K.; Zhou, Q.; Li, G.; Imam, S.Z.; Roberts, J.L.; Morgan, W.W.; Clark, R.A.; Li, S. Macrophage-mediated GDNF delivery protects against dopaminergic neurodegeneration: A therapeutic strategy for parkinson’s disease. Mol. Ther. 2010, 18, 1536–1544. [Google Scholar] [CrossRef]
  29. Zhao, Y.; Haney, M.J.; Gupta, R.; Bohnsack, J.P.; He, Z.; Kabanov, A.V.; Batrakova, E.V. GDNF-transfected macrophages produce potent neuroprotective effects in parkinson’s disease mouse model. PLoS ONE 2014, 9, e106867. [Google Scholar] [CrossRef] [Green Version]
  30. Nakajima, H.; Uchida, K.; Yayama, T.; Kobayashi, S.; Guerrero, A.R.; Furukawa, S.; Baba, H. Targeted retrograde gene delivery of brain-derived neurotrophic factor suppresses apoptosis of neurons and oligodendroglia after spinal cord injury in rats. Spine 2010, 35, 497–504. [Google Scholar] [CrossRef]
  31. Hovland, D.N.; Boyd, R.B.; Butt, M.T.; Engelhardt, J.A.; Moxness, M.S.; Ma, M.H.; Emery, M.G.; Ernst, N.B.; Reed, R.P.; Zeller, J.R.; et al. Six-month continuous intraputamenal infusion toxicity study of recombinant methionyl human glial cell line-derived neurotrophic factor (r-metHuGDNF) in rhesus monkeys. Toxicol. Pathol. 2007, 35, 676–692. [Google Scholar] [CrossRef]
  32. Uchida, K.; Nakajima, H.; Hirai, T.; Yayama, T.; Chen, K.; Guerrero, A.R.; Johnson, W.E.; Baba, H. The retrograde delivery of adenovirus vector carrying the gene for brain-derived neurotrophic factor protects neurons and oligodendrocytes from apoptosis in the chronically compressed spinal cord of twy/twy mice. Spine 2012, 37, 2125–2135. [Google Scholar] [CrossRef]
  33. Fouad, K.; Bennett, D.J.; Vavrek, R.; Blesch, A. Long-term viral brain-derived neurotrophic factor delivery promotes spasticity in rats with a cervical spinal cord hemisection. Front. Neurol. 2013, 4, 187. [Google Scholar] [CrossRef] [Green Version]
  34. Revilla, S.; Ursulet, S.; Álvarez-López, M.J.; Castro-Freire, M.; Perpiñá, U.; García-Mesa, Y.; Bortolozzi, A.; Giménez-Llort, L.; Kaliman, P.; Cristòfol, R.; et al. Lenti-GDNF Gene Therapy Protects Against Alzheimer’s Disease-Like Neuropathology in 3xTg-AD Mice and MC65 Cells. CNS Neurosci. Ther. 2014, 20, 961–972. [Google Scholar] [CrossRef] [Green Version]
  35. Popovic, N.; Maingay, M.; Kirik, D.; Brundin, P. Lentiviral gene delivery of GDNF into the striatum of R6/2 Huntington mice fails to attenuate behavioral and neuropathological changes. Exp. Neurol. 2005, 193, 65–74. [Google Scholar] [CrossRef]
  36. Humbel, M.; Ramosaj, M.; Zimmer, V.; Regio, S.; Aeby, L.; Moser, S.; Boizot, A.; Sipion, M.; Rey, M.; Déglon, N. Maximizing lentiviral vector gene transfer in the CNS. Gene Ther. 2021, 28, 75–88. [Google Scholar] [CrossRef]
  37. Conova, L.; Vernengo, J.; Jin, Y.; Himes, B.T.; Neuhuber, B.; Fischer, I.; Lowman, A. A pilot study of poly(N-isopropylacrylamide)-g-polyethylene glycol and poly(N-isopropylacrylamide)-g-methylcellulose branched copolymers as injectable scaffolds for local delivery of neurotrophins and cellular transplants into the injured spinal cord: Lab. J. Neurosurg. Spine 2011, 15, 594–604. [Google Scholar] [CrossRef] [Green Version]
  38. Khalin, I.; Alyautdin, R.; Wong, T.W.; Gnanou, J.; Kocherga, G.; Kreuter, J. Brain-derived neurotrophic factor delivered to the brain using poly (lactide-co-glycolide) nanoparticles improves neurological and cognitive outcome in mice with traumatic brain injury. Drug Deliv. 2016, 23, 3520–3528. [Google Scholar] [CrossRef] [Green Version]
  39. Donaghue, I.E.; Tator, C.H.; Shoichet, M.S. Sustained delivery of bioactive neurotrophin-3 to the injured spinal cord. Biomater. Sci. 2015, 3, 65–72. [Google Scholar] [CrossRef]
  40. Li, G.; Che, M.T.; Zeng, X.; Qiu, X.C.; Feng, B.; Lai, B.Q.; Shen, H.Y.; Ling, E.A.; Zeng, Y.S. Neurotrophin-3 released from implant of tissue-engineered fibroin scaffolds inhibits inflammation, enhances nerve fiber regeneration, and improves motor function in canine spinal cord injury. J. Biomed. Mater. Res.—Part A 2018, 106, 2158–2170. [Google Scholar] [CrossRef] [Green Version]
  41. McMurran, C.E.; Zhao, C.; Franklin, R.J.M. Toxin-based models to investigate demyelination and remyelination. Methods Mol. Biol. 2019, 1936, 377–396. [Google Scholar] [CrossRef]
  42. Moloney, T.C.; Rooney, G.E.; Barry, F.P.; Howard, L.; Dowd, E. Potential of rat bone marrow-derived mesenchymal stem cells as vehicles for delivery of neurotrophins to the Parkinsonian rat brain. Brain Res. 2010, 1359, 33–43. [Google Scholar] [CrossRef]
  43. Ji, W.C.; Li, M.; Jiang, W.T.; Ma, X.; Li, J. Protective effect of brain-derived neurotrophic factor and neurotrophin-3 overexpression by adipose-derived stem cells combined with silk fibroin/chitosan scaffold in spinal cord injury. Neurol. Res. 2020, 42, 361–371. [Google Scholar] [CrossRef]
  44. Sun, X.; Zhang, C.; Xu, J.; Zhai, H.; Liu, S.; Xu, Y.; Hu, Y.; Long, H.; Bai, Y.; Quan, D. Neurotrophin-3-Loaded Multichannel Nanofibrous Scaffolds Promoted Anti-Inflammation, Neuronal Differentiation, and Functional Recovery after Spinal Cord Injury. ACS Biomater. Sci. Eng. 2020, 6, 1228–1238. [Google Scholar] [CrossRef]
  45. Oudega, M.; Hao, P.; Shang, J.; Haggerty, A.E.; Wang, Z.; Sun, J.; Liebl, D.J.; Shi, Y.; Cheng, L.; Duan, H.; et al. Validation study of neurotrophin-3-releasing chitosan facilitation of neural tissue generation in the severely injured adult rat spinal cord. Exp. Neurol. 2019, 312, 51–62. [Google Scholar] [CrossRef]
  46. Makadia, H.K.; Siegel, S.J. Poly Lactic-co-Glycolic Acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef]
  47. Smith, D.R.; Dumont, C.M.; Ciciriello, A.J.; Guo, A.; Tatineni, R.; Munsell, M.K.; Cummings, B.J.; Anderson, A.J.; Shea, L.D. PLG Bridge Implantation in Chronic SCI Promotes Axonal Elongation and Myelination. ACS Biomater. Sci. Eng. 2019, 5, 6679–6690. [Google Scholar] [CrossRef] [Green Version]
  48. Tuinstra, H.M.; Aviles, M.O.; Shin, S.; Holland, S.J.; Zelivyanskaya, M.L.; Fast, A.G.; Ko, S.Y.; Margul, D.J.; Bartels, A.K.; Boehler, R.M.; et al. Multifunctional, multichannel bridges that deliver neurotrophin encoding lentivirus for regeneration following spinal cord injury. Biomaterials 2012, 33, 1618–1626. [Google Scholar] [CrossRef] [Green Version]
  49. Talbott, J.F.; Cao, Q.; Bertram, J.; Nkansah, M.; Benton, R.L.; Lavik, E.; Whittemore, S.R. CNTF promotes the survival and differentiation of adult spinal cord-derived oligodendrocyte precursor cells in vitro but fails to promote remyelination in vivo. Exp. Neurol. 2007, 204, 485–489. [Google Scholar] [CrossRef] [Green Version]
  50. Sun, S.; Zhang, Q.; Li, M.; Gao, P.; Huang, K.; Beejadhursing, R.; Jiang, W.; Lei, T.; Zhu, M.; Shu, K. GDNF Promotes Survival and Therapeutic Efficacy of Human Adipose-Derived Mesenchymal Stem Cells in a Mouse Model of Parkinson’s Disease. Cell Transplant. 2020, 29, 0963689720908512. [Google Scholar] [CrossRef] [Green Version]
  51. Suzuki, M.; McHugh, J.; Tork, C.; Shelley, B.; Hayes, A.; Bellantuono, I.; Aebischer, P.; Svendsen, C.N. Direct muscle delivery of GDNF with human mesenchymal stem cells improves motor neuron survival and function in a rat model of familial ALS. Mol. Ther. 2008, 16, 2002–2010. [Google Scholar] [CrossRef]
  52. Abbaszadeh, H.A.; Tiraihi, T.; Noori-Zadeh, A.; Delshad, A.R.; Sadeghizade, M.; Taheri, T. Human ciliary neurotrophic factor-overexpressing stable bone marrow stromal cells in the treatment of a rat model of traumatic spinal cord injury. Cytotherapy 2015, 17, 912–921. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, W.; Yan, Q.; Zeng, Y.; Zhang, X.; Xiong, Y.; Wang, J.; Chen, S. Implantation of adult bone marrow-derived mesenchymal stem cells transfected with the neurotrophin-3 gene and pretreated with retinoic acid in completely transected spinal cord. Brain Res. 2010, 1359, 256–271. [Google Scholar] [CrossRef]
  54. Gransee, H.M.; Zhan, W.Z.; Sieck, G.C.; Mantilla, C.B. Localized delivery of brain-derived neurotrophic factor-expressing mesenchymal stem cells enhances functional recovery following cervical spinal cord injury. J. Neurotrauma 2015, 32, 185–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lu, Z.; Hu, X.; Zhu, C.; Wang, D.; Zheng, X.; Liu, Q. Overexpression of CNTF in Mesenchymal Stem Cells reduces demyelination and induces clinical recovery in experimental autoimmune encephalomyelitis mice. J. Neuroimmunol. 2009, 206, 58–69. [Google Scholar] [CrossRef] [PubMed]
  56. Cedarbaum, J.M.; Chapman, C.; Charatan, M.; Stambler, N.; Andrews, L.; Zhan, C.; Radka, S.; Morrisey, D.; Lakings, D.; Brooks, B.R.; et al. A phase I study of recombinant human ciliary neurotrophic factor (rHCNTF) in patients with amyotrophic lateral sclerosis. Clin. Neuropharmacol. 1995, 18, 515–532. [Google Scholar] [CrossRef]
  57. Bongioanni, P.; Reali, C.; Sogos, V. Ciliary neurotrophic factor (CNTF) for amyotrophic lateral sclerosis or motor neuron disease. Cochrane Database Syst. Rev. 2004, 2004, CD004302. [Google Scholar] [CrossRef]
  58. Lang, A.E.; Gill, S.; Patel, N.K.; Lozano, A.; Nutt, J.G.; Penn, R.; Brooks, D.J.; Hotton, G.; Moro, E.; Heywood, P.; et al. Randomized controlled trial of intraputamenal glial cell line-derived neurotrophic factor infusion in Parkinson disease. Ann. Neurol. 2006, 59, 459–466. [Google Scholar] [CrossRef]
  59. Emerich, D.F.; Lindner, M.D.; Winn, S.R.; Chen, E.Y.; Frydel, B.R.; Kordower, J.H. Implants of encapsulated human CNTF-producing fibroblasts prevent behavioral deficits and striatal degeneration in a rodent model of Huntington’s disease. J. Neurosci. 1996, 16, 5168–5181. [Google Scholar] [CrossRef] [Green Version]
  60. Mittoux, V.; Joseph, J.M.; Conde, F.; Palfi, S.; Dautry, C.; Poyot, T.; Bloch, J.; Deglon, N.; Ouary, S.; Nimchinsky, E.A.; et al. Restoration of cognitive and motor functions by ciliary neurotrophic factor in a primate model of Huntington’s disease. Hum. Gene Ther. 2000, 11, 1177–1187. [Google Scholar] [CrossRef]
  61. Bachoud-Lévi, A.C.; Déglon, N.; Nguyen, J.P.; Bloch, J.; Bourdet, C.; Winkel, L.; Rémy, P.; Goddard, M.; Lefaucheur, J.P.; Brugières, P.; et al. Neuroprotective gene therapy for Huntington’s disease using a polymer encapsulated BHK cell line engineered to secrete human CNTF. Hum. Gene Ther. 2000, 11, 1723–1729. [Google Scholar] [CrossRef]
  62. Aebischer, P.; Schluep, M.; Déglon, N.; Joseph, J.M.; Hirt, L.; Heyd, B.; Goddard, M.; Hammang, J.P.; Zurn, A.D.; Kato, A.C.; et al. Intrathecal delivery of CNTF using encapsulated genetically modified xenogeneic cells in amyotrophic lateral sclerosis patients. Nat. Med. 1996, 2, 696–699. [Google Scholar] [CrossRef] [PubMed]
  63. Tuszynski, M.H.; Thal, L.; Pay, M.; Salmon, D.P.; U, H.S.; Bakay, R.; Patel, P.; Blesch, A.; Vahlsing, H.L.; Ho, G.; et al. A phase 1 clinical trial of nerve growth factor gene therapy for Alzheimer disease. Nat. Med. 2005, 11, 551–555. [Google Scholar] [CrossRef] [PubMed]
  64. Hadaczek, P.; Eberling, J.L.; Pivirotto, P.; Bringas, J.; Forsayeth, J.; Bankiewicz, K.S. Eight years of clinical improvement in MPTP-lesioned primates after gene therapy with AAV2-hAADC. Mol. Ther. 2010, 18, 1458–1461. [Google Scholar] [CrossRef]
  65. Tuszynski, M.H.; Yang, J.H.; Barba, D.; Hoi-Sang, U.; Bakay, R.A.E.; Pay, M.M.; Masliah, E.; Conner, J.M.; Kobalka, P.; Roy, S.; et al. Nerve growth factor gene therapy activation of neuronal responses in Alzheimer disease. JAMA Neurol. 2015, 72, 1139–1147. [Google Scholar] [CrossRef] [Green Version]
  66. Cedarbaum, J.M.; Chapman, C.; Charatan, M.T.; Stambler, N.; Andrews, L.; Zhan, C.; Radka, S.; Morrisey, D.; Lakings, D.; Brooks, B.R.; et al. The pharmacokinetics of subcutaneously administered recombinant human ciliary neurotrophic factor (rHCNTF) in patients with amyotrophic lateral sclerosis: Relation to parameters of the acute-phase response. Clin. Neuropharmacol. 1995, 18, 500–514. [Google Scholar] [CrossRef]
  67. Cedarbaum, J.M.; Brooks, B.R. A double-blind placebo-controlled clinical trial of subcutaneous recombinant human ciliary neurotrophic factor (rHCNTF) in amyotrophic lateral sclerosis. Neurology 1996, 46, 1244–1249. [Google Scholar] [CrossRef]
  68. Miller, R.G.; Petajan, J.H.; Bryan, W.W.; Armon, C.; Barohn, R.J.; Goodpasture, J.C.; Hoagland, R.J.; Parry, G.J.; Ross, M.A.; Stromatt, S.C. A placebo-controlled trial of recombinant human ciliary neurotrophic (rhCNTF) factor in amyotrophic lateral sclerosis. Ann. Neurol. 1996, 39, 256–260. [Google Scholar] [CrossRef] [PubMed]
  69. Miller, R.G.; Bryan, W.W.; Dietz, M.A.; Munsat, T.L.; Petajan, J.H.; Smith, S.A.; Goodpasture, J.C. Toxicity and tolerability of recombinant human ciliary neurotrophic factor in patients with amyotrophic lateral sclerosis. Neurology 1996, 47, 1329–1331. [Google Scholar] [CrossRef]
  70. Kordower, J.H.; Palfi, S.; Chen, E.Y.; Ma, S.Y.; Sendera, T.; Cochran, E.J.; Mufson, E.J.; Penn, R.; Goetz, C.G.; Comella, C.D. Clinicopathological findings following intraventricular glial-derived neurotrophic factor treatment in a patient with Parkinson’s disease. Ann. Neurol. 1999, 46, 419–424. [Google Scholar] [CrossRef]
  71. 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. Randomized, double-blind trial of glial cell line-derived neurotrophic factor (GDNF) in PD. Neurology 2003, 60, 69–73. [Google Scholar] [CrossRef]
  72. 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. [Google Scholar] [CrossRef] [PubMed]
  73. Patel, N.K.; Bunnage, M.; Plaha, P.; Svendsen, C.N.; Heywood, P.; Gill, S.S. Intraputamenal infusion of glial cell line-derived neurotrophic factor in PD: A two-year outcome study. Ann. Neurol. 2005, 57, 298–302. [Google Scholar] [CrossRef] [PubMed]
  74. Weissmiller, A.M.; Wu, C. Current advances in using neurotrophic factors to treat neurodegenerative disorders. Transl. Neurodegener. 2012, 1, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Oliveira, S.L.B.; Pillat, M.M.; Cheffer, A.; Lameu, C.; Schwindt, T.T.; Ulrich, H. Functions of neurotrophins and growth factors in neurogenesis and brain repair. Cytom. Part A 2013, 83A, 76–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Pietrucha-Dutczak, M.; Amadio, M.; Govoni, S.; Lewin-Kowalik, J.; Smedowski, A. The role of endogenous neuroprotective mechanisms in the prevention of retinal ganglion cells degeneration. Front. Neurosci. 2018, 12, 834. [Google Scholar] [CrossRef] [Green Version]
  77. MacMahon Copas, A.N.; McComish, S.F.; Fletcher, J.M.; Caldwell, M.A. The Pathogenesis of Parkinson’s Disease: A Complex Interplay Between Astrocytes, Microglia, and T Lymphocytes? Front. Neurol. 2021, 12, 771. [Google Scholar] [CrossRef]
  78. Lin, A.L.H.; Doherty, D.H.; Lile, J.D.; Bektesh, S.; Collins, F.; Lin, L.H. A Glial Cell LineDerived Neurotrophic GDNF: Factor for Midbrain Dopaminergic Neurons. Science 1993, 260, 1130–1132. [Google Scholar] [CrossRef]
  79. Lapchak, P.A.; Jiao, S.; Collins, F.; Miller, P.J. Glial cell line-derived neurotrophic factor: Distribution and pharmacology in the rat following a bolus intraventricular injection. Brain Res. 1997, 747, 92–102. [Google Scholar] [CrossRef]
  80. Sharma, M.; Braun, R.E. Cyclical expression of GDNF is required for spermatogonial stem cell homeostasis. Development 2018, 145, 10–11. [Google Scholar] [CrossRef] [Green Version]
  81. Zurn, A.D.; Winkel, L.; Menoud, A.; Djabali, K.; Aebischer, P. Combined effects of GDNF, BDNF, and CNTF on motoneuron differentiation in vitro. J. Neurosci. Res. 1996, 44, 133–141. [Google Scholar] [CrossRef]
  82. Kells, A.P.; Fong, D.M.; Dragunow, M.; During, M.J.; Young, D.; Connor, B. AAV-mediated gene delivery of BDNF or GDNF is neuroprotective in a model of Huntington disease. Mol. Ther. 2004, 9, 682–688. [Google Scholar] [CrossRef] [PubMed]
  83. McBride, J.L.; During, M.J.; Wuu, J.; Chen, E.Y.; Leurgans, S.E.; Kordower, J.H. Structural and functional neuroprotection in a rat model of Huntington’s disease by viral gene transfer of GDNF. Exp. Neurol. 2003, 181, 213–223. [Google Scholar] [CrossRef] [PubMed]
  84. Yi, A.; Markesbery, W.; Zhang, Z.; Grondin, R. Intraputamenal infusion of GDNF in aged rhesus monkeys_ Distribution and dopaminergic effects. J. Comp. Neurol. 2003, 461, 250–261. [Google Scholar]
  85. Maswood, N.; Grondin, R.; Zhang, Z.; Stanford, J.A.; Surgener, S.P.; Gash, D.M.; Gerhardt, G.A. Effects of chronic intraputamenal infusion of glial cell line-derived neurotrophic factor (GDNF) in aged Rhesus monkeys. Neurobiol. Aging 2002, 23, 881–889. [Google Scholar] [CrossRef] [PubMed]
  86. Martin, D.; Miller, G.; Fischer, N.; Dix, D.; Cullen, T.; Russell, D. Glial Cell Line-derived Neurotrophic Factor: The Lateral Cerebral Ventricle as a Site of Administration for Stimulation of the Substantia Nigra Dopamine System in Rats. Eur. J. Neurosci. 1996, 8, 1249–1255. [Google Scholar] [CrossRef]
  87. Bowenkamp, K.E.; Lapchak, P.A.; Hoffer, B.J.; Miller, P.J.; Bickford, P.C. Intracerebroventricular glial cell line-derived neurotrophic factor improves motor function and supports nigrostriatal dopamine neurons in bilaterally 6-hydroxydopamine lesioned rats. Exp. Neurol. 1997, 145, 104–117. [Google Scholar] [CrossRef]
  88. Lapchak, P.A.; Miller, P.J.; Collins, F.; Jiao, S. Glial cell line-derived neurotrophic factor attenuates behavioural deficits and regulates nigrostriatal dopaminergic and peptidergic markers in 6-hydroxydopamine-lesioned adult rats: Comparison of intraventricular and intranigral delivery. Neuroscience 1997, 78, 61–72. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, Z.; Miyoshi, Y.; Lapchak, P.A.; Collins, F.; Hilt, D.; Lebel, C.; Kryscio, R.; Gash, D.M. Dose response to intraventricular glial cell line-derived neurotrophic factor administration in Parkinsonian monkeys. J. Pharmacol. Exp. Ther. 1997, 282, 1396–1401. [Google Scholar]
  90. Greg, A.; Gerhardt, A.; Wayne, A.; Huett, P. GDNF improves dopamine function in the substantia nigra but not the putamen of unilateral MPTP-lesioned rhesus monkeys. Brain Res. 1999, 817, 163–171. [Google Scholar]
  91. Sergio, C.M.; Iravani, M. Glial cell line-derived neurotrophic factor concentration dependently improves disability and motor activity in MPTP-treated common marmosets. Eur. J. Pharmacol. 2001, 412, 45–50. [Google Scholar]
  92. Lapchak, P.A. Topographical distribution of w 125 I x -glial cell line-derived neurotrophic factor in unlesioned and MPTP-lesioned rhesus monkey brain following a bolus intraventricular injection. Brain Res. 1998, 789, 9–22. [Google Scholar] [CrossRef] [PubMed]
  93. Grondin, R.; Zhang, Z.; Yi, A.; Cass, W.A.; Maswood, N.; Andersen, A.H.; Elsberry, D.D.; Klein, M.C.; Gerhardt, G.A.; Gash, D.M. Chronic, controlled GDNF infusion promotes structural and functional recovery in advanced parkinsonian monkeys. Brain 2002, 125, 2191–2201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Manfredsson, F.P.; Polinski, N.K.; Subramanian, T.; Boulis, N.; Wakeman, D.R.; Mandel, R.J. The Future of GDNF in Parkinson’s Disease. Front. Aging Neurosci. 2020, 12, 593572. [Google Scholar] [CrossRef] [PubMed]
  95. Secombes, C.J.; Wang, T.; Bird, S. Vertebrate Cytokines and Their Evolution. In The Evolution of the Immune System; Academic Press: Cambridge, MA, USA, 2016; pp. 87–150. [Google Scholar] [CrossRef]
  96. Shanker, H.; Sharma, A. Neuropharmacology of Neuroprotection; Progress in Brain Research; Elsevier: Amsterdam, The Netherlands, 2022. [Google Scholar]
  97. Ramirez, B.U.; Retamal, L.; Vergara, C. Ciliary neurotrophic factor (CNTF) affects the excitable and contractile properties of innervated skeletal muscles. Biol. Res. 2003, 36, 303–312. [Google Scholar] [CrossRef] [Green Version]
  98. Lee, N.; Wanek, H.A.; MacLennan, A.J. Muscle ciliary neurotrophic factor receptor α helps maintain choline acetyltransferase levels in denervated motor neurons following peripheral nerve lesion. Exp. Neurol. 2019, 317, 202–205. [Google Scholar] [CrossRef]
  99. Miranda, M.; Morici, J.F.; Zanoni, M.B.; Bekinschtein, P. Brain-Derived Neurotrophic Factor: A Key Molecule for Memory in the Healthy and the Pathological Brain. Front. Cell. Neurosci. 2019, 13, 363. [Google Scholar] [CrossRef]
  100. Choi, S.H.; Bylykbashi, E.; Chatila, Z.K.; Lee, S.W.; Pulli, B.; Clemenson, G.D.; Kim, E.; Rompala, A.; Oram, M.K.; Asselin, C.; et al. Combined adult neurogenesis and BDNF mimic exercise effects on cognition in an Alzheimer’s mouse model. Science 2018, 361, eaan8821. [Google Scholar] [CrossRef] [Green Version]
  101. Blurton-Jones, M.; Kitazawa, M.; Martinez-Coria, H.; Castello, N.A.; Müller, F.J.; Loring, J.F.; Yamasaki, T.R.; Poon, W.W.; Green, K.N.; LaFerla, F.M. Neural stem cells improve cognition via BDNF in a transgenic model of Alzheimer disease. Proc. Natl. Acad. Sci. USA 2009, 106, 13594–13599. [Google Scholar] [CrossRef] [Green Version]
  102. Karimi, N.; Ashourizadeh, H.; Akbarzadeh, B.; Haghshomar, M.; Jouzdani, T.; Shobeiri, P.; Teixeira, A.L. Blood levels of brain-derived neurotrophic factor ( BDNF ) in people with multiple sclerosis ( MS ): A systematic review and meta-analysis. Mult. Scler. Relat. Disord. 2022, 65, 103984. [Google Scholar] [CrossRef]
  103. Azoulay, D.; Urshansky, N.; Karni, A. Low and dysregulated BDNF secretion from immune cells of MS patients is related to reduced neuroprotection. J. Neuroimmunol. 2008, 195, 186–193. [Google Scholar] [CrossRef]
  104. Knaepen, K.; Goekint, M.; Heyman, E.M.; Meeusen, R. Neuroplasticity-Exercise-Induced Response of Peripheral Brain-Derived Neurotrophic Factor A Systematic Review of Experimental Studies in Human Subjects. Sport. Med. 2010, 40, 765–801. [Google Scholar] [CrossRef] [PubMed]
  105. Makar, T.K.; Bever, C.T.; Singh, I.S.; Royal, W.; Sahu, S.N.; Sura, T.P.; Sultana, S.; Sura, K.T.; Patel, N.; Dhib-Jalbut, S.; et al. Brain-derived neurotrophic factor gene delivery in an animal model of multiple sclerosis using bone marrow stem cells as a vehicle. J. Neuroimmunol. 2009, 210, 40–51. [Google Scholar] [CrossRef] [PubMed]
  106. Tobias, C.A.; Shumsky, J.S.; Shibata, M.; Tuszynski, M.H.; Fischer, I.; Tessler, A.; Murray, M. Delayed grafting of BDNF and NT-3 producing fibroblasts into the injured spinal cord stimulates sprouting, partially rescues axotomized red nucleus neurons from loss and atrophy, and provides limited regeneration. Exp. Neurol. 2003, 184, 97–113. [Google Scholar] [CrossRef] [PubMed]
  107. Jin, Y.; Fischer, I.; Tessler, A.; Houle, J.D. Transplants of fibroblasts genetically modified to express BDNF promote axonal regeneration from supraspinal neurons following chronic spinal cord injury. Exp. Neurol. 2002, 177, 265–275. [Google Scholar] [CrossRef] [PubMed]
  108. McTigue, D.M.; Horner, P.J.; Stokes, B.T.; Gage, F.H. Neurotrophin-3 and brain-derived neurotrophic factor induce oligodendrocyte proliferation and myelination of regenerating axons in the contused adult rat spinal cord. J. Neurosci. 1998, 18, 5354–5365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Tuszynski, M.H. Intraparenchymal NGF infusions rescue degenerating cholinergic neurons. Cell Transplant. 2000, 9, 629–636. [Google Scholar] [CrossRef] [PubMed]
  110. Jönhagen, M.E.; Nordberg, A.; Amberla, K.; Bäckman, L.; Ebendal, T.; Meyerson, B.; Olson, L.; Seiger, Å.; Shigeta, M.; Theodorsson, E.; et al. Intracerebroventricular infusion of nerve growth factor in three patients with Alzheimer’s disease. Dement. Geriatr. Cogn. Disord. 1998, 9, 246–257. [Google Scholar] [CrossRef] [PubMed]
  111. Delivanoglou, N.; Boziki, M.; Theotokis, P.; Kesidou, E.; Touloumi, O.; Dafi, N.; Nousiopoulou, E.; Lagoudaki, R.; Grigoriadis, N.; Charalampopoulos, I.; et al. Spatio-temporal expression profile of NGF and the two-receptor system, TrkA and p75NTR, in experimental autoimmune encephalomyelitis. J. Neuroinflammation 2020, 17, 1–17. [Google Scholar] [CrossRef] [Green Version]
  112. Minnone, G.; De Benedetti, F.; Bracci-Laudiero, L. NGF and its receptors in the regulation of inflammatory response. Int. J. Mol. Sci. 2017, 18, 1028. [Google Scholar] [CrossRef] [Green Version]
  113. Zhou, L.; Baumgartner, B.J.; Hill-Felberg, S.J.; McGowen, L.R.; Shine, H.D. Neurotrophin-3 expressed in situ induces axonal plasticity in the adult injured spinal cord. J. Neurosci. 2003, 23, 1424–1431. [Google Scholar] [CrossRef] [Green Version]
  114. Haque, N.S.K.; Hlavin, M.L.; Fawcett, J.W.; Dunnett, S.B. The neurotrophin NT4/5, but not NT3, enhances the efficacy of nigral grafts in a rat model of Parkinson’s disease. Brain Res. 1996, 712, 45–52. [Google Scholar] [CrossRef] [PubMed]
  115. Akyol, O.; Sherchan, P.; Yilmaz, G.; Reis, C.; Ho, W.M.; Wang, Y.; Huang, L.; Solaroglu, I.; Zhang, J.H. Neurotrophin-3 provides neuroprotection via TrkC receptor dependent pErk5 activation in a rat surgical brain injury model. Exp. Neurol. 2018, 307, 82–89. [Google Scholar] [CrossRef]
  116. Yan, Z.; Shi, X.; Wang, H.; Si, C.; Liu, Q.; Du, Y. Neurotrophin-3 Promotes the Neuronal Differentiation of BMSCs and Improves Cognitive Function in a Rat Model of Alzheimer’ s Disease. Front. Cell. Neurosci. 2021, 15, 629356. [Google Scholar] [CrossRef] [PubMed]
  117. Donnelly, E.M.; Madigan, N.N.; Rooney, G.E.; Knight, A.; Chen, B.; Ball, B.; Kinnavane, L.; Garcia, Y.; Dockery, P.; Fraher, J.; et al. Lentiviral vector delivery of short hairpin RNA to NG2 and neurotrophin-3 promotes locomotor recovery in injured rat spinal cord. Cytotherapy 2012, 14, 1235–1244. [Google Scholar] [CrossRef] [PubMed]
  118. Kim, M.S.; Shutov, L.P.; Gnanasekaran, A.; Lin, Z.; Rysted, J.E.; Ulrich, J.D.; Usachev, Y.M. Nerve growth factor (NGF) regulates activity of nuclear factor of activated T-cells (NFAT) in neurons via the phosphatidylinositol 3-kinase (PI3K)-Akt-glycogen synthase kinase 3β (GSK3β) pathway. J. Biol. Chem. 2014, 289, 31349–31360. [Google Scholar] [CrossRef] [Green Version]
  119. Reichardt, L.F. Neurotrophin-regulated signalling pathways. Philos. Trans. R Soc. B Biol. Sci. 2006, 361, 1545–1564. [Google Scholar] [CrossRef] [Green Version]
  120. Mitre, M.; Mariga, A.; Chao, M.V. Neurotrophin signalling: Novel insights into mechanisms and pathophysiology. Clin. Sci. 2017, 131, 13–23. [Google Scholar] [CrossRef] [Green Version]
  121. Kashyap, M.P.; Roberts, C.; Waseem, M.; Tyagi, P. Drug Targets in Neurotrophin Signaling in the Central and Peripheral Nervous System. Mol. Neurobiol. 2018, 55, 6939–6955. [Google Scholar] [CrossRef]
  122. Williams, L.R. Hypophagia is induced by intracerebroventricular administration of nerve growth factor. Exp. Neurol. 1991, 113, 31–37. [Google Scholar] [CrossRef]
  123. Isaacson, L.G.; Saffran, B.N.; Crutcher, K.A. Intracerebral NGF infusion induces hyperinnervation of cerebral blood vessels. Neurobiol. Aging 1990, 11, 51–55. [Google Scholar] [CrossRef]
  124. Winkler, J.; Ramirez, G.A.; Kuhn, H.G.; Peterson, D.A.; Day-Lollini, P.A.; Stewart, G.R.; Tuszynski, M.H.; Gage, F.H.; Thal, L.J. Reversible schwann cell hyperplasia and sprouting of sensory and sympathetic neurites after intraventricular administration of nerve growth factor. Ann. Neurol. 1997, 41, 82–93. [Google Scholar] [CrossRef] [PubMed]
  125. Poduslo, J.F.; Curran, G.L. Permeability at the blood-brain and blood-nerve barriers of the neurotrophic factors: NGF, CNTF, NT-. Brain Res. Mol. Brain Res. 1996, 36, 280–286. [Google Scholar] [CrossRef] [PubMed]
  126. Blesch, A. Neurotrophic Factors in Neurodegeneration. Brain Pathol. 2006, 16, 295–303. [Google Scholar] [CrossRef] [PubMed]
  127. Sadan, O.; Shemesh, N.; Barzilay, R.; Dadon-Nahum, M.; Blumenfeld-Katzir, T.; Assaf, Y.; Yeshurun, M.; Djaldetti, R.; Cohen, Y.; Melamed, E.; et al. Mesenchymal stem cells induced to secrete neurotrophic factors attenuate quinolinic acid toxicity: A potential therapy for Huntington’s disease. Exp. Neurol. 2012, 234, 417–427. [Google Scholar] [CrossRef] [PubMed]
  128. Eggenhofer, E.; Luk, F.; Dahlke, M.H.; Hoogduijn, M.J. The life and fate of mesenchymal stem cells. Front. Immunol. 2014, 5, 148. [Google Scholar] [CrossRef] [Green Version]
  129. Tuszynski, M.H.; Roberts, J.; Senut, M.C.; Hs, U.; Gage, F.H. Gene therapy in the adult primate brain: Intraparenchymal grafts of cells genetically modified to produce nerve growth factor prevent cholinergic neuronal degeneration. Gene Ther. 1996, 3, 305–314. [Google Scholar]
  130. Emerich, D.F.; Winn, S.R.; Harper, J.; Hammang, J.P.; Baetge, E.E.; Kordower, J.H. Implants of polymer-encapsulated human NGF-secreting cells in the nonhuman primate: Rescue and sprouting of degenerating cholinergic basal forebrain neurons. J. Comp. Neurol. 1994, 349, 148–164. [Google Scholar] [CrossRef]
  131. Kordower, J.H.; Winn, S.R.; Liu, Y.T.; Mufson, E.J.; Sladek, J.R.; Hammang, J.P.; Baetge, E.E.; Emerich, D.F. The aged monkey basal forebrain: Rescue and sprouting of axotomized basal forebrain neurons after grafts of encapsulated cells secreting human nerve growth factor. Proc. Natl. Acad. Sci. USA 1994, 91, 10898–10902. [Google Scholar] [CrossRef] [Green Version]
  132. Smith, D.E.; Roberts, J.; Gage, F.H.; Tuszynski, M.H. Age-associated neuronal atrophy occurs in the primate brain and is reversible by growth factor gene therapy. Proc. Natl. Acad. Sci. USA 1999, 96, 10893–10898. [Google Scholar] [CrossRef] [Green Version]
  133. Conner, J.M.; Darracq, M.A.; Roberts, J.; Tuszynski, M.H. Nontropic actions of neurotrophins: Subcortical nerve growth factor gene delivery reverses age-related degeneration of primate cortical cholinergic innervation. Proc. Natl. Acad. Sci. USA 2001, 98, 1941–1946. [Google Scholar] [CrossRef] [Green Version]
  134. Chen, K.S.; Gage, F.H. Somatic gene transfer of NGF to the aged brain: Behavioral and morphological amelioration. J. Neurosci. 1995, 15, 2819–2825. [Google Scholar] [CrossRef] [PubMed]
  135. Bankiewicz, K.S.; Forsayeth, J.; Eberling, J.L.; Sanchez-Pernaute, R.; Pivirotto, P.; Bringas, J.; Herscovitch, P.; Carson, R.E.; Eckelman, W.; Reutter, B.; et al. Long-Term Clinical Improvement in MPTP-Lesioned Primates after Gene Therapy with AAV-hAADC. Mol. Ther. 2006, 14, 564–570. [Google Scholar] [CrossRef] [PubMed]
  136. Eslamboli, A.; Georgievska, B.; Ridley, R.M.; Baker, H.F.; Muzyczka, N.; Burger, C.; Mandel, R.J.; Annett, L.; Kirik, D. Continuous low-level glial cell line-derived neurotrophic factor delivery using recombinant adeno-associated viral vectors provides neuroprotection and induces behavioral recovery in a primate model of Parkinson’s disease. J. Neurosci. 2005, 25, 769–777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Kordower, J.H.; Emborg, M.E.; Bloch, J.; Ma, S.Y.; Chu, Y.; Leventhal, L.; McBride, J.; Chen, E.Y.; Palfi, S.; Roitberg, B.Z.; et al. Neurodegeneration prevented by lentiviral vector delivery of GDNF in primate models of Parkinson’s disease. Science 2000, 290, 767–773. [Google Scholar] [CrossRef]
  138. De Almeida, L.P.; Zala, D.; Aebischer, P.; Déglon, N. Neuroprotective effect of a CNTF-expressing lentiviral vector in the quinolinic acid rat model of Huntington’s disease. Neurobiol. Dis. 2001, 8, 433–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Régulier, E.; De Almeida, L.P.; Sommer, B.; Aebischer, P.; Déglon, N. Dose-dependent neuroprotective effect of ciliary neurotrophic factor delivered via tetracycline- regulated lentiviral vectors in the quinolinic acid rat model of Huntington’ s disease. Hum. Gene Ther. 2002, 13, 1981–1990. [Google Scholar] [CrossRef]
  140. Bensadoun, J.C.; Déglon, N.; Tseng, J.L.; Ridet, J.L.; Zurn, A.D.; Aebischer, P. Lentiviral vectors as a gene delivery system in the mouse midbrain: Cellular and behavioral improvements in a 6-OHDA model of Parkinson’s disease using GDNF. Exp. Neurol. 2000, 164, 15–24. [Google Scholar] [CrossRef]
  141. He, Z.; Zang, H.; Zhu, L.; Huang, K.; Yi, T.; Zhang, S.; Cheng, S. An anti-inflammatory peptide and brain-derived neurotrophic factor-modified hyaluronan-methylcellulose hydrogel promotes nerve regeneration in rats with spinal cord injury. Int. J. Nanomed. 2019, 14, 721–732. [Google Scholar] [CrossRef] [Green Version]
  142. Wu, G.H.; Shi, H.J.; Che, M.T.; Huang, M.Y.; Wei, Q.S.; Feng, B.; Ma, Y.H.; Wang, L.J.; Jiang, B.; Wang, Y.Q.; et al. Recovery of paralyzed limb motor function in canine with complete spinal cord injury following implantation of MSC-derived neural network tissue. Biomaterials 2018, 181, 15–34. [Google Scholar] [CrossRef]
  143. Ankeny, D.P.; McTigue, D.M.; Guan, Z.; Yan, Q.; Kinstler, O.; Stokes, B.T.; Jakeman, L.B. Pegylated brain-derived neurotrophic factor shows improved distribution into the spinal cord and stimulates locomotor activity and morphological changes after injury. Exp. Neurol. 2001, 170, 85–100. [Google Scholar] [CrossRef]
  144. Bondarenko, O.; Saarma, M. Neurotrophic Factors in Parkinson’s Disease: Clinical Trials, Open Challenges and Nanoparticle-Mediated Delivery to the Brain. Front. Cell Neurosci. 2021, 15, 682597. [Google Scholar] [CrossRef] [PubMed]
  145. Aly, A.E.E.; Harmon, B.T.; Padegimas, L.; Sesenoglu-Laird, O.; Cooper, M.J.; Waszczak, B.L. Intranasal Delivery of pGDNF DNA Nanoparticles Provides Neuroprotection in the Rat 6-Hydroxydopamine Model of Parkinson’s Disease. Mol. Neurobiol. 2019, 56, 688–701. [Google Scholar] [CrossRef] [PubMed]
  146. Park, J.B.; Lee, J.S.; Cho, B.P.; Rhee, K.J.; Baik, S.K.; Kim, J.; Kang, S.J.; Park, D.J.; Oh, J.E.; Shin, H.C.; et al. Adipose tissue-derived mesenchymal stem cells cultured at high cell density express brain-derived neurotrophic factor and exert neuroprotective effects in a 6-hydroxydopamine rat model of Parkinson’s disease. Genes Genom. 2015, 37, 213–221. [Google Scholar] [CrossRef]
  147. Williams, B.; Granholm, A.; Sambamurti, K. Age-Dependent Loss of Ngf Signaling in the Rat Basal Forebrain Is due to Disrupted Mapk Activation. Neurosci. Lett. 2007, 413, 110–114. [Google Scholar] [CrossRef]
  148. 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. [Google Scholar] [CrossRef] [PubMed]
  149. Xue, Y.-Q.; Ma, B.-F.; Zhao, L.-R. AAV9-mediated erythropoietin gene delivery into the brain protects nigral dopaminergic neurons in a rat model of Parkinson’s disease. Gene Ther. 2009, 17, 83–94. [Google Scholar] [CrossRef] [Green Version]
  150. Gunnarson, E.; Song, Y.; Kowalewski, J.M.; Brismar, H.; Brines, M.; Cerami, A. Erythropoietin modulation of astrocyte water permeability as a component of neuroprotection. Proc. Natl. Acad. Sci. USA 2009, 106, 1602–1607. [Google Scholar] [CrossRef] [Green Version]
  151. Tang, Z.; Sun, X.; Huo, G.; Xie, Y.; Shi, Q.; Chen, S.; Wang, X. Protective effects of erythropoietin on astrocytic swelling after oxygen-glucose deprivation and reoxygenation: Mediation through AQP4 expression and MAPK pathway. Neuropharmacology 2013, 67, 8–15. [Google Scholar] [CrossRef]
  152. Pardridge, W.M. Blood-brain barrier drug targeting enables neuroprotection in brain ischemia following delayed intravenous administration of neurotrophins. Adv. Exp. Med. Biol. 2002, 513, 397–430. [Google Scholar] [CrossRef]
  153. Molinari, C.; Morsanuto, V.; Ruga, S.; Notte, F.; Farghali, M.; Galla, R.; Uberti, F. The role of BDNF on aging-modulation markers. Brain Sci. 2020, 10, 285. [Google Scholar] [CrossRef]
  154. Ji, R.; Smith, M.; Niimi, Y.; Karakatsani, M.E.; Murillo, M.F.; Jackson-Lewis, V.; Przedborski, S.; Konofagou, E.E. Focused ultrasound enhanced intranasal delivery of brain derived neurotrophic factor produces neurorestorative effects in a Parkinson’s disease mouse model. Sci. Rep. 2019, 9, 19402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Wang, F.; Wu, H.; Hu, A.; Dong, L.; Lin, X.; Li, M.; Wang, Y.; Li, W.; Chang, L.; Chang, Y.; et al. Ultrasound combined with glial cell line-derived neurotrophic factor-loaded microbubbles for the targeted treatment of drug addiction. Front. Bioeng. Biotechnol. 2022, 10, 961728. [Google Scholar] [CrossRef] [PubMed]
  156. Fishman, P.S.; Fischell, J.M. Focused Ultrasound Mediated Opening of the Blood-Brain Barrier for Neurodegenerative Diseases. Front. Neurol. 2021, 12, 749047. [Google Scholar] [CrossRef] [PubMed]
  157. Karakatsani, M.E.; Blesa, J.; Konofagou, E.E. Blood–brain barrier opening with focused ultrasound in experimental models of Parkinson’s disease. Mov. Disord. 2019, 34, 1252–1261. [Google Scholar] [CrossRef]
  158. Idbaih, A.; Canney, M.; Belin, L.; Desseaux, C.; Vignot, A.; Bouchoux, G.; Asquier, N.; Law-Ye, B.; Leclercq, D.; Bissery, A.; et al. Safety and Feasibility of Repeated and Transient Blood–Brain Barrier Disruption by Pulsed Ultrasound in Patients with Recurrent Glioblastoma. Clin. Cancer Res. 2019, 25, 3793–3801. [Google Scholar] [CrossRef] [Green Version]
  159. Poduslo, J.F.; Curran, G.L.; Gill, J.S. Putrescine-Modified Nerve Growth Factor: Bioactivity, Plasma Pharmacokinetics, Blood-Brain/Nerve Barrier Permeability, and Nervous System Biodistribution. J. Neurochem. 2002, 71, 1998–2000. [Google Scholar] [CrossRef] [Green Version]
  160. Wu, D.; Pardridge, W.M. Neuroprotection with noninvasive neurotrophin delivery to the brain. Proc. Natl. Acad. Sci. USA 1999, 96, 254–259. [Google Scholar] [CrossRef] [Green Version]
  161. Saragovi, H.U.; Kalle, G. Development of pharmacological agents for targeting neurotrophins and their receptors. Trends Pharmacol. Sci. 2000, 21, 93–98. [Google Scholar] [CrossRef]
  162. Herberts, C.A. Risk factors in the development of stem cell therapy. J. Transl. Med. 2011, 9, 29. [Google Scholar] [CrossRef] [Green Version]
  163. Leventhal, A.; Chen, G.; Negro, A.; Boehm, M. The benefits and risks of stem cell technology. Oral Dis. 2013, 18, 217–222. [Google Scholar] [CrossRef]
  164. Master, Z.; Mcleod, M.; Mendez, I. Benefits, risks and ethical considerations in translation of stem cell research to clinical applications in Parkinson ’ s disease. J. Med. Ethics 2007, 33, 169–173. [Google Scholar] [CrossRef] [Green Version]
  165. Gray, S.J.; Woodard, K.T.; Samulski, R.J. Viral vectors and delivery strategies for CNS gene therapy. Ther. Deliv. 2015, 1, 517–534. [Google Scholar] [CrossRef] [Green Version]
  166. Nayak, S.; Herzog, R.W. Progress and Prospects: Immune Responses to Viral Vectors. Gene Ther. 2010, 17, 295–304. [Google Scholar] [CrossRef] [Green Version]
  167. Williams, D.F. Challenges With the Development of Biomaterials for Sustainable Tissue Engineering. Front. Bioeng. Biotechnol. 2019, 7, 127. [Google Scholar] [CrossRef] [PubMed]
  168. Gardner, A.B.; Lee, S.K.C.; Woods, E.C.; Acharya, A.P. Biomaterials-Based Modulation of the Immune System. Biomed. Res. Int. 2013, 2013, 732182. [Google Scholar] [CrossRef]
  169. Wang, M.O.; Etheridge, J.M.; Thompson, J.A.; Vorwald, C.E.; Dean, D. Evaluation of the In Vitro Cytotoxicity of Crosslinked Biomaterials. Biomacromolecules 2013, 14, 1321–1329. [Google Scholar] [CrossRef] [Green Version]
  170. Wu, Y.; Rakotoarisoa, M.; Angelov, B.; Deng, Y.; Angelova, A. Self-Assembled Nanoscale Materials for Neuronal Regeneration: A Focus on BDNF Protein and Nucleic Acid Biotherapeutic Delivery. Nanomaterials 2022, 12, 2267. [Google Scholar] [CrossRef] [PubMed]
  171. Guarnieri, G.; Sarchielli, E.; Comeglio, P.; Herrera-Puerta, E.; Piaceri, I.; Nacmias, B.; Benelli, M.; Kelsey, G.; Maggi, M.; Gallina, P.; et al. Tumor necrosis factor α influences phenotypic plasticity and promotes epigenetic changes in human basal forebrain cholinergic neuroblasts. Int. J. Mol. Sci. 2020, 21, 6128. [Google Scholar] [CrossRef]
  172. Subedi, L.; Lee, S.E.; Madiha, S.; Gaire, B.P.; Jin, M.; Yumnam, S.; Kim, S.Y. Phytochemicals against TNFα-mediated neuroinflammatory diseases. Int. J. Mol. Sci. 2020, 21, 764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Uberti, F.; Morsanuto, V.; Ghirlanda, S.; Ruga, S.; Clemente, N.; Boieri, C.; Boldorini, R.; Molinari, C. Highly Diluted Acetylcholine Promotes Wound Repair in an in Vivo Model. Adv. Wound Care 2018, 7, 121–133. [Google Scholar] [CrossRef] [Green Version]
  174. Adessi, C.; Soto, C. Converting a peptide into a drug: Strategies to improve stability and bioavailability. Curr. Med. Chem. 2002, 9, 963–978. [Google Scholar] [CrossRef]
  175. Price, R.D.; Milne, S.A.; Sharkey, J.; Matsuoka, N. Advances in small molecules promoting neurotrophic function. Pharmacol. Ther. 2007, 115, 292–306. [Google Scholar] [CrossRef] [PubMed]
  176. Lesauteur, L.; Ekiel, I.; Saragovi, H.U.; Gehring, K. Solution Structure and Internal Motion of a Bioactive Peptide Derived from Nerve Growth Factor. J. Biol. Chem. 1998, 273, 23652–23658. [Google Scholar] [CrossRef] [Green Version]
  177. Leary, P.D.O.; Hughes, R.A. Design of potent peptide mimetics of brain-derived neurotrophic factor. J. Biol. Chem. 2003, 278, 25738–25744. [Google Scholar] [CrossRef]
  178. Massa, S.M.; Yang, T.; Xie, Y.; Shi, J.; Bilgen, M.; Joyce, J.N.; Nehama, D.; Rajadas, J.; Longo, F.M. Small molecule BDNF mimetics activate TrkB signaling and prevent neuronal degeneration in rodents. J. Clin. Investig. 2010, 120, 8–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Bregman, B.S.; Mcatee, M.; Dai, H.N.; Kuhn, P.L. Neurotrophic Factors Increase Axonal Growth after Spinal Cord Injury and Transplantation in the Adult Rat. Exp. Neurol. 1997, 148, 475–494. [Google Scholar] [CrossRef]
  180. Knüsel, B.; Winslow, J.W.; Rosenthal, A.; Burton, L.E.; Seid, D.P.; Nikolics, K.; Hefti, F. Promotion of central cholinergic and dopaminergic neuron differentiation by brain-derived neurotrophic factor but not neurotrophin 3. Proc. Natl. Acad. Sci. USA 1991, 88, 961–965. [Google Scholar] [CrossRef] [Green Version]
  181. Storkebaum, E.; Carmeliet, P. VEGF: A critical player in neurodegeneration. J. Clin. Investig. 2004, 113, 14–18. [Google Scholar] [CrossRef]
  182. Oosthuyse, B.; Moons, L.; Storkebaum, E.; Beck, H.; Nuyens, D.; Brusselmans, K.; Van Dorpe, J.; Hellings, P.; Gorselink, M.; Heymans, S.; et al. Deletion of the hypoxia-response element in the vascular endothelial growth factor promoter causes motor neuron degeneration. Nat. Genet. 2001, 28, 131–138. [Google Scholar] [CrossRef]
  183. Facchiano, F.; Fernandez, E.; Mancarella, S.; Maira, G.; Miscusi, M.; D’Arcangelo, D.; Cimino-Reale, G.; Falchetti, M.L.; Capogrossi, M.C.; Pallini, R. Promotion of regeneration of corticospinal tract axons in rats with recombinant vascular endothelial growth factor alone and combined with adenovirus coding for this factor. J. Neurosurg. 2002, 97, 161–168. [Google Scholar] [CrossRef]
  184. Spuch, C.; Antequera, D.; Portero, A.; Orive, G.; Hernández, R.M.; Molina, J.A.; Bermejo-Pareja, F.; Pedraz, J.L.; Carro, E. The effect of encapsulated VEGF-secreting cells on brain amyloid load and behavioral impairment in a mouse model of Alzheimer’s disease. Biomaterials 2010, 31, 5608–5618. [Google Scholar] [CrossRef] [PubMed]
  185. Herrán, E.; Pérez-González, R.; Igartua, M.; Pedraz, J.L.; Carro, E.; Hernández, R.M. VEGF-releasing biodegradable nanospheres administered by craniotomy: A novel therapeutic approach in the APP/Ps1 mouse model of Alzheimer’s disease. J. Control. Release 2013, 170, 111–119. [Google Scholar] [CrossRef]
  186. Herran, E.; Perez- Gonzalez, R.; Igartua, M.; Pedraz, J.L.; Carro, E.; Hernandez, R.M. Enhanced Hippocampal Neurogenesis in APP/Ps1 Mouse Model of Alzheimer’s Disease After Implantation of VEGF-loaded PLGA Nanospheres. Curr. Alzheimer Res. 2015, 12, 932–940. [Google Scholar] [CrossRef] [PubMed]
  187. Xiong, N.; Zhang, Z.; Huang, J.; Chen, C.; Zhang, Z.; Jia, M.; Xiong, J.; Liu, X.; Wang, F.; Cao, X.; et al. VEGF-expressing human umbilical cord mesenchymal stem cells, an improved therapy strategy for Parkinsons disease. Gene Ther. 2011, 18, 394–402. [Google Scholar] [CrossRef] [PubMed]
  188. Yasuhara, T.; Shingo, T.; Muraoka, K.; Ji, Y.W.; Kameda, M.; Takeuchi, A.; Yano, A.; Nishio, S.; Matsui, T.; Miyoshi, Y.; et al. The differences between high and low-dose administration of VEGF to dopaminergic neurons of in vitro and in vivo Parkinson’s disease model. Brain Res. 2005, 1038, 1–10. [Google Scholar] [CrossRef]
  189. Yao, D.L.; Liu, X.; Hudson, L.D.; Webster, H.D. Insulin-like growth factor I treatment reduces demyelination and up- regulates gene expression of myelin-related proteins in experimental autoimmune encephalomyelitis. Proc. Natl. Acad. Sci. USA 1995, 92, 6190–6194. [Google Scholar] [CrossRef] [Green Version]
  190. Beck, K.D.; Powell-Braxtont, L.; Widmer, H.R.; Valverde, J.; Hefti, F. Igf1 gene disruption results in reduced brain size, CNS hypomyelination, and loss of hippocampal granule and striatal parvalbumin-containing neurons. Neuron 1995, 14, 717–730. [Google Scholar] [CrossRef] [Green Version]
  191. Borasio, G.D.; Robberecht, W.; Leigh, P.N.; Emile, J.; Guiloff, R.J.; Jerusalem, F.; Silani, V.; Vos, P.E.; Wokke, J.H.J.; Dobbins, T. A placebo-controlled trial of insulin-like growth factor-I in amyotrophic lateral sclerosis. Neurology 1998, 51, 583–586. [Google Scholar] [CrossRef]
  192. Beauverd, M.; Mitchell, J.D.; Wokke, J.H.; Borasio, G.D. Recombinant human insulin-like growth factor I (rhIGF-I) for the treatment of amyotrophic lateral sclerosis/motor neuron disease. Cochrane Database Syst. Rev. 2012, 11, CD002064. [Google Scholar] [CrossRef]
  193. Nagano, I.; Shiote, M.; Murakami, T.; Kamada, H.; Hamakawa, Y.; Matsubara, E.; Yokoyama, M.; Morita, K.; Shoji, M.; Abe, K. Beneficial effects of intrathecal IGF-1 administration in patients with amyotrophic lateral sclerosis. Neurol. Res. 2005, 27, 768–772. [Google Scholar] [CrossRef]
  194. Sorenson, E.J.; Windbank, A.J.; Mandrekar, J.N.; Bamlet, W.R.; Appel, S.H.; Armon, C.; Barkhaus, P.E.; Bosch, P.; Boylan, K.; David, W.S.; et al. Subcutaneous IGF-1 is not beneficial in 2-year ALS trial. Neurology 2008, 71, 1770–1775. [Google Scholar] [CrossRef] [PubMed]
  195. Vincent, A.M.; Mobley, B.C.; Hiller, A.; Feldman, E.L. IGF-I prevents glutamate-induced motor neuron programmed cell death. Neurobiol. Dis. 2004, 16, 407–416. [Google Scholar] [CrossRef] [PubMed]
  196. Apel, P.J.; Ma, J.; Callahan, M.; Northam, C.N.; Alton, T.B.; Sonntag, W.E.; Li, Z. Effect of locally delivered IGF-1 on nerve regeneration during aging: An experimental study in rats. Muscle Nerve 2010, 41, 335–341. [Google Scholar] [CrossRef] [Green Version]
  197. Cannella, B.; Pitt, D.; Capello, C.S.; Raine, C.S. Insulin-like growth factor-1 fails to enhance central nervous system myelin repair during autoimmune demyelination. Am. J. Pathol. 2000, 157, 933–943. [Google Scholar] [CrossRef] [PubMed]
  198. Ebert, A.D.; Beres, A.J.; Barber, A.E.; Svendsen, C.N. Human neural progenitor cells over-expressing IGF-1 protect dopamine neurons and restore function in a rat model of Parkinson’s disease. Exp. Neurol. 2008, 209, 213–223. [Google Scholar] [CrossRef]
  199. Lepore, A.C.; Haenggeli, C.; Gasmi, M.; Bishop, K.M.; Bartus, R.T.; Maragakis, N.J.; Rothstein, J.D. Intraparenchymal spinal cord delivery of adeno-associated virus IGF-1 is protective in the SOD1G93A model of ALS. Brain Res. 2007, 1185, 256–265. [Google Scholar] [CrossRef] [Green Version]
  200. Kaspar, B.K.; Lladó, J.; Sherkat, N.; Rothstein, J.D.; Gage, F.H. Retrograde viral delivery of IGF-1 prolongs survival in a mouse ALS model. Science 2003, 301, 839–842. [Google Scholar] [CrossRef] [Green Version]
  201. Dodge, J.C.; Haidet, A.M.; Yang, W.; Passini, M.A.; Hester, M.; Clarke, J.; Roskelley, E.M.; Treleaven, C.M.; Rizo, L.; Martin, H.; et al. Delivery of AAV-IGF-1 to the CNS extends survival in ALS mice through modification of aberrant glial cell activity. Mol. Ther. 2008, 16, 1056–1064. [Google Scholar] [CrossRef]
  202. Saenger, S.; Holtmann, B.; Nilges, M.R.; Schroeder, S.; Hoeflich, A.; Kletzl, H.; Spooren, W.; Ostrowitzki, S.; Hanania, T.; Sendtner, M.; et al. Functional improvement in mouse models of familial amyotrophic lateral sclerosis by PEGylated insulin-like growth factor i treatment depends on disease severity. Amyotroph. Lateral Scler. 2012, 13, 418–429. [Google Scholar] [CrossRef]
  203. Lai, E.C.; Felice, K.J.; Festoff, B.W.; Gawel, M.J.; Gelinas, D.F.; Kratz, R.; Murphy, M.F.; Natter, H.M.; Norris, F.H.; Rudnicki, S.A. Effect of recombinant human insulin-like growth factor-I on progression of ALS: A placebo-controlled study. Neurology 1997, 49, 1621–1630. [Google Scholar] [CrossRef]
  204. Gonzalez, D.; Rebolledo, D.L.; Correa, L.M.; Court, F.A.; Cerpa, W.; Lipson, K.E.; Van Zundert, B.; Brandan, E. The inhibition of CTGF/CCN2 activity improves muscle and locomotor function in a murine ALS model. Hum. Mol. Genet. 2018, 27, 2913–2926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Dombrowski, Y.; O’Hagan, T.; DIttmer, M.; Penalva, R.; Mayoral, S.R.; Bankhead, P.; Fleville, S.; Eleftheriadis, G.; Zhao, C.; Naughton, M.; et al. Regulatory T cells promote myelin regeneration in the central nervous system. Nat. Neurosci. 2017, 20, 674–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Dittmer, M.; Young, A.; O’Hagan, T.; Eleftheriadis, G.; Bankhead, P.; Dombrowski, Y.; Medina, R.J.; Fitzgerald, D.C. Characterization of a murine mixed neuron-glia model and cellular responses to regulatory T cell-derived factors. Mol. Brain 2018, 11, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. de la Vega Gallardo, N.; Penalva, R.; Dittmer, M.; Naughton, M.; Falconer, J.; Moffat, J.; de la Fuente, A.G.; Hombrebueno, J.R.; Lin, Z.; Perbal, B.; et al. Dynamic CCN3 expression in the murine CNS does not confer essential roles in myelination or remyelination. Proc. Natl. Acad. Sci. USA 2020, 117, 18018–18028. [Google Scholar] [CrossRef] [PubMed]
  208. Cheng, Z.; Zhang, Y.; Tian, Y.; Chen, Y. Cyr61 promotes Schwann cell proliferation and migration via αvβ3 integrin. BMC Mol. Cell Biol. 2021, 22, 21. [Google Scholar] [CrossRef] [PubMed]
  209. Armand-ugón, M.; Aso, E.; Moreno, J.; Riera-codina, M.; Sánchez, A. Memory Improvement in the AβPP/PS1 Mouse Model of Familial Alzheimer’s Disease Induced by Carbamylated-Erythropoietin is Accompanied by Modulation of Synaptic Genes. J. Alzheimer’s Dis. 2018, 45, 675–677. [Google Scholar] [CrossRef] [Green Version]
  210. Lee, S.T.; Chu, K.; Park, J.E.; Jung, K.H.; Jeon, D.; Lim, J.Y.; Lee, S.K.; Kim, M.; Roh, J.K. Erythropoietin improves memory function with reducing endothelial dysfunction and amyloid-beta burden in Alzheimer’s disease models. J. Neurochem. 2012, 120, 115–124. [Google Scholar] [CrossRef]
  211. Leist, M.; Ghezzi, P.; Grasso, G.; Bianchi, R.; Villa, P.; Fratelli, M.; Savino, C.; Bianchi, M.; Nielsen, J.; Gerwien, J.; et al. Derivatives of erythropoietin that are tissue protective but not erythropoietic. Science 2004, 305, 13–14. [Google Scholar] [CrossRef] [Green Version]
  212. Cetin, A.; Nas, K.; Büyükbayram, H.; Ceviz, A.; Olmez, G. The effects of systemically administered methylprednisolone and recombinant human erythropoietin after acute spinal cord compressive injury in rats. Eur. Spine J. 2006, 15, 1539–1544. [Google Scholar] [CrossRef]
  213. Simon, F.; Scheuerle, A.; Gröger, M.; Vcelar, B.; Mccook, O.; Möller, P.; Georgieff, M.; Calzia, E.; Radermacher, P.; Schelzig, H. Comparison of carbamylated erythropoietin-FC fusion protein and recombinant human erythropoietin during porcine aortic balloon occlusion-induced spinal cord ischemia / reperfusion injury. Intensive Care Med. 2011, 37, 1525–1533. [Google Scholar] [CrossRef]
  214. Rees, S.; Hale, N.; De Matteo, R.; Cardamone, L.; Tolcos, M.; Loeliger, M.; Mackintosh, A.; Shields, A.; Probyn, M.; Greenwood, D.; et al. Erythropoietin is neuroprotective in a preterm ovine model of endotoxin-induced brain injury. J. Neuropathol. Exp. Neurol. 2010, 69, 306–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Cerri, G.; Montagna, M.; Madaschi, L.; Merli, D.; Borroni, P.; Baldissera, F.; Gorio, A. Erythropoietin effect on sensorimotor recovery after contusive spinal cord injury: An electrophysiological study in rats. Neuroscience 2012, 219, 290–301. [Google Scholar] [CrossRef] [PubMed]
  216. Carelli, S.; Giallongo, T.; Viaggi, C.; Gombalova, Z.; Latorre, E. Grafted Neural Precursors Integrate Into Mouse Striatum, Differentiate and Promote Recovery of Function Through Release of Erythropoietin in MPTP-Treated Mice. ASN Neuro 2016, 8, 1759091416676147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Carelli, S. Counteracting neuroinflammation in experimental Parkinson’ s disease favors recovery of function: Effects of Er-NPCs administration. J. Neuroinflammation 2018, 15, 333. [Google Scholar] [CrossRef] [PubMed]
  218. Marfia, G.; Madaschi, L.; Marra, F.; Menarini, M.; Bottai, D.; Formenti, A.; Bellardita, C.; Maria, A.; Giulio, D.; Carelli, S.; et al. Adult neural precursors isolated from post mortem brain yield mostly neurons: An erythropoietin- dependent process. Neurobiol. Dis. 2022, 43, 86–98. [Google Scholar] [CrossRef]
  219. Carelli, S.; Giallongo, T.; Marfia, G.; Merli, D.; Ottobrini, L.; Degrassi, A.; Basso, M.D.; Maria, A.; Giulio, D.; Gorio, A. Exogenous adult postmortem neural precursors attenuate secondary degeneration and promote myelin sparing and functional recovery following experimental spinal cord injury. Cell Transplant. 2015, 24, 703–719. [Google Scholar] [CrossRef] [Green Version]
  220. Carelli, S.; Giallongo, T.; Gombalova, Z.; Merli, D.; Di Giulio, A.M.; Gorio, A. EPO-releasing neural precursor cells promote axonal regeneration and recovery of function in spinal cord traumatic injury. Restor. Neurol Neurosci. 2017, 35, 583–599. [Google Scholar] [CrossRef] [Green Version]
  221. Samy, D.M.; Ismail, C.A.; Nassra, R.A.; Zeitoun, T.M.; Nomair, A.M. Downstream modulation of extrinsic apoptotic pathway in streptozotocin-induced Alzheimer’s dementia in rats: Erythropoietin versus curcumin. Eur. J. Pharmacol. 2016, 770, 52–60. [Google Scholar] [CrossRef]
  222. Pedroso, I.; Bringas, M.L.; Aguiar, A.; Morales, L.; Alvarez, M.; Valdés, P.A. Use of Cuban recombinant human erythropoietin in Parkinson’ s disease treatment. Medicc Rev. 2012, 14, 11. [Google Scholar] [CrossRef]
  223. Jang, W.; Park, J.; Shin, K.J.; Kim, J.; Kim, J.S.; Youn, J. Safety and efficacy of recombinant human erythropoietin treatment of non-motor symptoms in Parkinson’ s disease. J. Neurol. Sci. 2013, 337, 47–54. [Google Scholar] [CrossRef]
  224. Lauria, G.; Campanella, A.; Filippini, G.; Martini, A.; Penza, P.; Maggi, L.; Antozzi, C.; Ciano, C.; Beretta, P.; Caldiroli, D.; et al. Erythropoietin in amyotrophic lateral sclerosis: A pilot, randomized, double-blind, placebo-controlled study of safety and tolerability. Amyotroph. Lateral Scler. 2009, 10, 410–415. [Google Scholar] [CrossRef] [PubMed]
  225. Lauria, G.; Bella, E.D.; Antonini, G.; Borghero, G.; Capasso, M.; Caponnetto, C.; Chiò, A.; Corbo, M.; Eleopra, R. Erythropoietin in amyotrophic lateral sclerosis: A multicentre, randomised, double blind, placebo controlled, phase III study. J. Neurol. Neurosurg. Psychiatry 2015, 86, 879–886. [Google Scholar] [CrossRef] [PubMed]
  226. Kim, H.Y.; Moon, C.; Kim, S.; Oh, W.; Oh, S.; Kim, J.; Kim, S.H. Recombinant Human Erythropoietin in Amyotrophic Lateral Sclerosis: A Pilot Study of Safety and Feasibility. J. Clin. Neurol. 2014, 10, 342–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Costa, D.D.; Beghi, E.; Carignano, P.; Pagliacci, C.; Faccioli, F.; Pupillo, E.; Messina, P.; Gorio, A.; Redaelli, T. Tolerability and efficacy of erythropoietin (EPO) treatment in traumatic spinal cord injury: A preliminary randomized comparative trial vs. methylprednisolone (MP). Neurol. Sci. 2015, 36, 1567–1574. [Google Scholar] [CrossRef]
  228. Cruz, Y.R.; Strehaiano, M.; Rodríguez Obaya, T.; Rodríguez, J.C.G.; Maurice, T. An intranasal formulation of erythropoietin (Neuro-EPO) prevents memory deficits and amyloid toxicity in the APPSwe transgenic mouse model of Alzheimer’s disease. J. Alzheimer’s Dis. 2016, 55, 231–248. [Google Scholar] [CrossRef]
  229. Reisi, P.; Arabpoor, Z.; Rashidi, B.; Alaei, H.; Salami, M.; Hamidi, G.; Shabrang, M.; Sharifi, M.; Dolatabadi, H.R. Erythropoietin improves neuronal proliferation in dentate gyrus of hippocampal formation in an animal model of Alzheimer′s disease. Adv. Biomed. Res. 2012, 1, 50. [Google Scholar] [CrossRef]
  230. Grignaschi, G.; Zennaro, E.; Tortarolo, M.; Calvaresi, N.; Bendotti, C. Erythropoietin does not preserve motor neurons in a mouse model of familial ALS. Amyotroph. Lateral Scler. 2007, 8, 31–35. [Google Scholar] [CrossRef]
  231. Grunfeld, J.F.; Barhum, Y.; Blondheim, N.; Rabey, J.; Melamed, E. Erythropoietin delays disease onset in an amyotrophic lateral sclerosis model. Exp. Cell Res. 2007, 204, 260–263. [Google Scholar] [CrossRef]
  232. Koh, S.; Kim, Y.; Kim, H.Y.; Cho, G.W.; Kim, K.S.; Kim, S.H. Recombinant human erythropoietin suppresses symptom onset and progression of G93A-SOD1 mouse model of ALS by preventing motor neuron death and inflammation. Eur. J. Neurosci. 2007, 25, 1923–1930. [Google Scholar] [CrossRef]
  233. Noh, M.Y.; Cho, K.A.; Kim, H.; Kim, S.; Kim, S.H. Erythropoietin modulates the immune- inflammatory response of a SOD1 (G93A) transgenic mouse model of amyotrophic lateral sclerosis (ALS). Neurosci. Lett. 2014, 574, 53–58. [Google Scholar] [CrossRef]
  234. Celik, M.; Gökmen, N.; Erbayraktar, S.; Akhisaroglu, M.; Konakç, S.; Ulukus, C.; Genc, S. Erythropoietin prevents motor neuron apoptosis and neurologic disability in experimental spinal cord ischemic injury. Proc. Natl. Acad. Sci. USA 2002, 99, 2258–2263. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The different NTF signalling pathways. The colours of the NTFs correlate with the colours of the arrows. Abbreviations used: Phospholipase C-γ (PLC-γ), phosphoinositide-3-kinase (PI3K), mitogen-activated protein kinase (MAPK), leukaemia inhibitory factor (LIFRβ), glycoprotein 130 (GP130), Janus kinase/signal transducer and activator of transcription (JAK), c-Jun N-terminal kinase (JNK), GDNF family receptor alpha-1 (GFRα1), Inositol trisphosphate (IP3), diacylglycerol (DAG), protein kinase C (PKC), Guanine nucleotide exchange factor (GEF) (Created with BioRender.com; adapted from Pietrucha-Dutczak et al. [76] and Kashyap et al. [121]).
Figure 1. The different NTF signalling pathways. The colours of the NTFs correlate with the colours of the arrows. Abbreviations used: Phospholipase C-γ (PLC-γ), phosphoinositide-3-kinase (PI3K), mitogen-activated protein kinase (MAPK), leukaemia inhibitory factor (LIFRβ), glycoprotein 130 (GP130), Janus kinase/signal transducer and activator of transcription (JAK), c-Jun N-terminal kinase (JNK), GDNF family receptor alpha-1 (GFRα1), Inositol trisphosphate (IP3), diacylglycerol (DAG), protein kinase C (PKC), Guanine nucleotide exchange factor (GEF) (Created with BioRender.com; adapted from Pietrucha-Dutczak et al. [76] and Kashyap et al. [121]).
Ijms 24 03866 g001
Figure 2. A summary of the current status of clinical trials applying neurotrophic factors. Several studies with NTFs for neurodegenerative diseases are still in the preclinical phase, whereas some of the clinical trials already initiated were terminated due to side effects or no clinical improvement. The different colours of the boxes correlate with the colours given to the NTF. Abbreviations used: glial cell-derived neurotrophic factor (GDNF), ciliary neurotrophic factor (CNTF), nerve growth factor (NGF), Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), amyotrophic lateral sclerosis (ALS), blood-brain barrier (BBB), Mini-Mental Status Examination (MMSE) scores, and N= enrolled patients (Created with BioRender.com) [59,60,61,63,64,65,66,67,69,70].
Figure 2. A summary of the current status of clinical trials applying neurotrophic factors. Several studies with NTFs for neurodegenerative diseases are still in the preclinical phase, whereas some of the clinical trials already initiated were terminated due to side effects or no clinical improvement. The different colours of the boxes correlate with the colours given to the NTF. Abbreviations used: glial cell-derived neurotrophic factor (GDNF), ciliary neurotrophic factor (CNTF), nerve growth factor (NGF), Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), amyotrophic lateral sclerosis (ALS), blood-brain barrier (BBB), Mini-Mental Status Examination (MMSE) scores, and N= enrolled patients (Created with BioRender.com) [59,60,61,63,64,65,66,67,69,70].
Ijms 24 03866 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

El Ouaamari, Y.; Van den Bos, J.; Willekens, B.; Cools, N.; Wens, I. Neurotrophic Factors as Regenerative Therapy for Neurodegenerative Diseases: Current Status, Challenges and Future Perspectives. Int. J. Mol. Sci. 2023, 24, 3866. https://doi.org/10.3390/ijms24043866

AMA Style

El Ouaamari Y, Van den Bos J, Willekens B, Cools N, Wens I. Neurotrophic Factors as Regenerative Therapy for Neurodegenerative Diseases: Current Status, Challenges and Future Perspectives. International Journal of Molecular Sciences. 2023; 24(4):3866. https://doi.org/10.3390/ijms24043866

Chicago/Turabian Style

El Ouaamari, Yousra, Jasper Van den Bos, Barbara Willekens, Nathalie Cools, and Inez Wens. 2023. "Neurotrophic Factors as Regenerative Therapy for Neurodegenerative Diseases: Current Status, Challenges and Future Perspectives" International Journal of Molecular Sciences 24, no. 4: 3866. https://doi.org/10.3390/ijms24043866

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop