Yurena
Polo
a,
Jon
Luzuriaga
b,
Sergio
Gonzalez de Langarica
c,
Beatriz
Pardo-Rodríguez
b,
Daniel E.
Martínez-Tong
d,
Christos
Tapeinos
ef,
Irene
Manero-Roig
bg,
Edurne
Marin
c,
Jone
Muñoz-Ugartemendia
c,
Gianni
Ciofani
e,
Gaskon
Ibarretxe
b,
Fernando
Unda
b,
Jose-Ramon
Sarasua
c,
Jose Ramon
Pineda
*bh and
Aitor
Larrañaga
*c
aPolimerbio, Donostia-San Sebastian, Spain
bCell Signaling Lab, Department of Cell Biology and Histology, Faculty of Medicine and Nursing, University of the Basque Country (UPV/EHU), Leioa, Spain. E-mail: joseramon.pinedam@ehu.eus; Tel: +34 946 012 426
cGroup of Science and Engineering of Polymeric Biomaterials (ZIBIO Group), Department of Mining, Metallurgy Engineering and Materials Science, POLYMAT, University of the Basque Country (UPV/EHU), Bilbao, Spain. E-mail: aitor.larranagae@ehu.eus; Tel: +34 946 013 935
dPolymers and advanced materials: Physics, Chemistry and Technology, University of the Basque Country (UPV/EHU), Donostia-San Sebastian, Spain & Centro de Física de Materiales (UPV/EHU-CSIC), Donostia-San Sebastian, Spain
eSmart Bio-Interfaces, Istituto Italiano di Tecnologia, Pontedera, PI, Italy
fDivision of Pharmacy and Optometry, School of Health Sciences, Faculty of Biology, Medicine and Health, University of Manchester, Manchester, UK
gUniversité de Bordeaux IINS – UMR 5297, Bordeaux, France
hAchucarro Basque Center for Neuroscience Fundazioa, Leioa, Spain
First published on 8th February 2023
Stem cell-based therapies have shown promising results for the regeneration of the nervous system. However, the survival and integration of the stem cells in the neural circuitry is suboptimal and might compromise the therapeutic outcomes of this approach. The development of functional scaffolds capable of actively interacting with stem cells may overcome the current limitations of stem cell-based therapies. In this study, three-dimensional hydrogels based on graphene derivatives and cerium oxide (CeO2) nanoparticles are presented as prospective supports allowing neural stem cell adhesion, migration and differentiation. The morphological, mechanical and electrical properties of the resulting hydrogels can be finely tuned by controlling several parameters of the self-assembly of graphene oxide sheets, namely the amount of incorporated reducing agent (ascorbic acid) and CeO2 nanoparticles. The intrinsic properties of the hydrogels, as well as the presence of CeO2 nanoparticles, clearly influence the cell fate. Thus, stiffer adhesion substrates promote differentiation to glial cell lineages, while softer substrates enhance mature neuronal differentiation. Remarkably, CeO2 nanoparticle-containing hydrogels support the differentiation of neural stem cells to neuronal, astroglial and oligodendroglial lineage cells, promoting the in vitro generation of nerve tissue grafts that might be employed in neuroregenerative cell therapies.
Among the different approaches to regenerate the CNS, cell-based therapies, particularly those based on neural stem cells (NSCs), offer the most straightforward alternative to reestablish a functional neural network by producing therapeutic factors, promoting the self-restoration of the damaged tissue and ultimately replacing the lost neural cells.4,6 However, the integration of the cells into the host CNS remains challenging.7 In this regard, tissue engineering offers the possibility to combine NSCs with scaffolds to enhance cell integration on the damaged area.6 Within the recently coined field of materiobiology, scaffolds are considered multifunctional devices with the capability to finely tune biological functions.8 In the particular case of nervous system regeneration, materials based on graphene derivatives have attracted considerable attention9 thanks to the possibility to create moldable platforms (e.g., with tailored chemical, mechanical and electrical features) to promote the adhesion and differentiation of NSCs towards functional glial and neuronal lineages.10,11 Graphene consists of a single layer of carbon atoms organized in a honeycomb lattice monolayer that can be arranged either in two- (2D) or three-dimensional (3D) scaffolds, thus resembling the complex architecture of the extracellular matrix.12,13 Although most of the studies with NSCs have been performed in 2D graphene derivative-based scaffolds, recently, 3D scaffolds have been reported to better promote the proliferative ability of NSCs, while maintaining similar adhesion features.14 Moreover, the physical properties of these 3D scaffolds including stiffness or pore geometry can modulate the adhesion, proliferation and differentiation capabilities of the NSCs.13 Understanding how these features interact in the formation of complex neural networks that support mature neuronal and glial interplay will be vital to ensure the successful integration of the NSCs into the graphene derivative-based 3D scaffolds.
Three-dimensional composite scaffolds offer several benefits concerning single material scaffolds since they allow a simple modulation of their physicochemical, mechanical and electrical properties by simply modifying their composition while exploiting potential synergistic effects among their constituents.15 Herein, the combination of 3D scaffolds made of graphene-derivatives and cerium oxide (CeO2) nanoparticles will be explored as a platform for the regeneration of the CNS. CeO2 nanoparticles comprise a cubic fluorite arrangement that acts as redox reaction sites thanks to the oxygen deficiencies at the nanoscale order.16 This oxygen deprivation endows antioxidant features to the CeO2 nanoparticles that resemble the activity of antioxidant natural enzymes (i.e., superoxide dismutase (SOD) and catalase (CAT)), which is beneficial for the promotion of the angiogenesis and the restoration of the neural architecture.17,18 Accordingly, CeO2 nanoparticles have been reported to provide neuroprotective effects, as demonstrated on an adult spinal cord neuron model19 and in vivo on a pharmacologically induced brain oxidative stress model.20
In this study, we combine the potential of graphene-based materials to promote the adhesion, proliferation and differentiation of NSCs, with the additional antioxidant and neuroprotective effects associated with CeO2 nanoparticles as a prospective approach for the restoration of the injured CNS. To achieve this aim, we engineered and characterized 3D hydrogels based on the combination of graphene derivatives and CeO2 nanoparticles with tunable stiffness, porous geometry and electrical conductivity. We further studied the adhesion, integration and differentiation capabilities of the NSCs towards neuronal, astroglial, and oligodendroglial lineages at different time points. This allowed us to establish heterocellular cultures for in vitro studies that mimicked the in vivo CNS tissue architecture.
As the hydrogels have different diameters, 60000 cells were seeded on the GO
:
AsA 1
:
1 and 30
000 cells were seeded on the GO
:
AsA 1
:
4 and GO
:
AsA 1
:
4 + CeO2 0.25 hydrogels to keep a similar cell density. After 24 h, medium was changed to NeuroCult™ differentiation medium and cells were cultured as previously described.11
Primers | Sequence 5′–3′ | Lenght | Annealing | Amplicon (bp) |
---|---|---|---|---|
Gapdh (upstream) | GTATGACTCCACTCACGGCAA | 21 | 61.8 | 274 |
Gapdh (downstream) | CTTCCCATTCTCGGCCTTG | 19 | 60.2 | 274 |
Nestin (upstream) | CCCTGAAGTCGAGGAGCTG | 19 | 61.4 | 166 |
Nestin (downstream) | CTGCTGCACCTCTAAGCGA | 19 | 61.7 | 166 |
Gfap (upstream) | CTGGACTGCGTCATTTTCCC | 20 | 59.2 | 256 |
Gfap (downstream) | CGATGGAGCCTCAGGGATGA | 20 | 61.1 | 256 |
S100β (upstream) | TGGCTGCGGAAGTTGAGATT | 20 | 59.9 | 84 |
S100β (upstream) | ATGGCTCCCAGCAGCTAAAG | 20 | 60.1 | 84 |
Olig2 (upstream) | GTGGATGCTTATTACAGACC | 20 | 56.1 | 94 |
Olig2 (downstream) | ACCTTCCGAATGTGAATTAG | 20 | 58.1 | 94 |
Map2 (upstream) | GAAGAAACAGCTAATCTGCC | 20 | 58.1 | 423 |
Map2 (downstream) | CTCTTGCTTATTCCATCAGTG | 21 | 59.0 | 423 |
Dcx (upstream) | TCAGCATCTCCACCCAACCA | 20 | 61.1 | 94 |
Dcx (downstream) | TTGTGCTTTCCCGGTTGACA | 20 | 60.3 | 94 |
For the quantitative PCR, 4.5 μL of SsoAdvanced Universal SYBR® Green Supermix (1725271; BioRad, Hercules, California, USA) were mixed with 0.5 μL of primers (0.3125 M), 0.3 μL of cDNA (1.5 ng μL−1) and the necessary nuclease free water to reach 10 μL final volume reaction per well. Each primer was evaluated for optimal efficacy (>90%) and single product amplification using the melting curve approach. The 2−ΔΔCt technique was employed to determine the relative expression of each gene, with Gapdh serving as internal control.11,21 qPCR was carried out in triplicate using an ABI PRISM® 7000 (Thermo Fisher Scientific, Waltham, Massachusetts, USA). Data were examined using the CFX Manager™ program. For statistical analysis, 3 independent hydrogels were analyzed, and each of the samples measured in triplicate.
The absence or presence of oxygen functionalities will therefore have a direct impact on the final properties of the hydrogels.27 The different proportions of GO:
AsA allowed us to modulate both the reduction level and the arrangement of the GO sheets, which determined the pore size of the hydrogels (Fig. 1B). Scanning electron micrographs revealed the large porous structures formed by atomic wide walls of GO sheets. Increasing the amount of AsA diminished the repulsion forces between the GO sheets and enabled the formation of more compact structures with smaller pores. The GO
:
AsA 1
:
1 hydrogel showed the largest pores (7.3 ± 0.6 μm) with respect to the other two formulations (i.e., GO
:
AsA 1
:
4 showing pores of 3.9 ± 0.2 μm & 1
:
10 with pores of 4.3 ± 0.5 μm). In agreement with these results, X-ray diffraction demonstrated that the addition of an increasing amount of AsA triggered the reduction of GO sheets, resulting in the displacement of the diffraction peak to higher values (i.e., from 10° in the commercial GO to 24° in the GO
:
AsA 1
:
4 & 1
:
10) (Fig. 1C). According to Bragg's law, the interlayer distance of the commercial GO was 0.85 nm and decreased till 0.77 nm in the GO
:
AsA 1
:
1 sample. The chemical reduction of GO stimulated the self-assembly of the reduced GO sheets thanks to the reduction of hydroxyl, epoxy, carboxyl and carbonyl groups,28 that enabled the formation of new π–π binding sites between the GO sheets. The deletion of oxygenated functional groups also raised the hydrophobicity of the graphene sheets. The combined effect of these two events provoked a random overlapping of flexible graphene sheets, thus favouring the self-assembly of the 3D hydrogels. Remarkably, the addition of more AsA decreased the interlayer distance in the GO
:
AsA 1
:
4 and GO
:
AsA 1
:
10 hydrogels till 0.37 nm in both cases, but no differences were observed between these two formulations, suggesting that the reduction level was similar in both cases. These results were further corroborated by XPS (ESI Fig. 1†). The addition of AsA at a GO
:
AsA 1
:
1 ratio clearly reduced the area associated to oxygen functionalities with respect to the original GO. Oxygen functionalities were further reduced in the GO
:
AsA 1
:
4 ratio but no significant differences were observed when the GO
:
AsA ratio was increased to 1
:
10, suggesting that the maximum level of reduction had been reached. As observed, the reduced graphene oxide hydrogels still contain oxygen functionalities in their structure, which are absent in the structure of pure graphite.
The mechanical properties of the hydrogels were measured by rheology. The GO:
AsA 1
:
1 hydrogel showed the lowest shear modulus (G′) (22.8 ± 0.3 kPa), which in viscous materials is directly correlated with the elastic capabilities (Fig. 1D). The GO
:
AsA 1
:
4 and GO
:
AsA 1
:
10 presented increased shear modulus (178.4 ± 2.8 kPa vs. 186.4 ± 2.6 kPa), demonstrating the modulation of the mechanical properties in almost one order of magnitude by controlling the reduction level through simply modifying the amount of AsA added. As shown in other studies and by us, once the maximum reduction level is achieved, the mechanical properties of the hydrogels remain stable.29 Although the human brain presents a low shear modulus of around 1–2.5 kPa,30 other central nervous system areas like the spinal cord recorded shear modulus of 250–300 kPa
31 and the peripheral nervous system like ulnar and median nerves register a shear modulus around 10–20 kPa,32 which are similar to the mechanical properties presented by our hydrogels. Besides, it is reported the acquirement of well differentiated neural cultures of stem cells in vitro and in vivo with scaffolds that present even higher stiffness values,33 suggesting that all the hydrogels exhibited mechanical properties compatible with neural differentiation of stem cells.
GO is a poor electrical conductor due to the lack of percolating conduits between sp2 carbon atoms that act as electron carriers in graphene. The reduction process with AsA induces the deletion of oxygen functionalities and raises the amount of sp2 or π–π binding sites which consequently increases the conductivity of the material.34 In accordance with the data of other studies,35 the electrical conductivity of the hydrogels increased with the addition of AsA. The GO:
AsA 1
:
1 hydrogel exhibited the lowest electrical conductivity (0.6 S m−1), which increased in GO
:
AsA 1
:
4 (27 S m−1) and GO
:
AsA 1
:
10 (35 S m−1) (Fig. 1E). Accordingly, we were able to enhance the electrical conductivity thanks to the deletion of atomic-scale lattice defects of the GO sheets via the thermochemical reduction process.22 According to these results, other groups have reported electrical conductivities between 0.045 and 600 S m−1 on graphene-based hydrogels.36–38 Remarkably, as it has been previously reported, the impedance or the electrical conductivity are constant in graphene oxide as usually observed for highly conducting systems.39,40 The human brain has an electrical conductivity of around 0.33 S m−1, where grey matter exhibits a conductivity of around 0.47 ± 0.24 S m−1 and white matter around 0.22 ± 0.17 S m−1.41,42 However, materials showing electrical conductivity values between 0.08–1.3 S m−1 or even higher values are able to electrically stimulate neurons.43,44 Therefore, all our hydrogels showed conductivity values compatible with neural stimulation.
To confirm the presence of CeO2 nanoparticles in our samples, we performed transmission electron microscopy (TEM). TEM micrographs showed the GO sheets successfully decorated with CeO2 nanoparticles prior to their reduction with AsA (Fig. 2C). Moreover, the amount of CeO2 nanoparticles on the surface of the GO sheets increased with the addition of more CeO2 to the samples (data not shown). These results were further corroborated with Raman spectroscopy and energy-dispersive X-ray spectrometry (EDX) (Fig. 2D, E and ESI Fig. 2†). Raman spectroscopy demonstrated that all the samples that contained CeO2 nanoparticles showed a band at 453 cm−1 that can be ascribed to the first order scattering of CeO246 and increased in intensity as the amount of CeO2 nanoparticles raised. As expected, the GO
:
AsA 1
:
4 sample only exhibited the characteristic D and G bands (1350 and 1580 cm1 respectively) which are related to the disarranged sp2-hybridized carbon structure and expansion of the C–C bond in graphitic materials.11,47 In accordance with these results, EDX demonstrated the presence of Ce in the samples containing CeO2 nanoparticles and presented similar proportions for carbon (C) (73.22–79.74%) and oxygen (O) (19.75–25.19%) in all the samples.
SEM images demonstrated that, despite the addition of CeO2 nanoparticles, hydrogels preserved their highly porous structures formed by atomic wide walls of GO sheets (Fig. 2F). The estimated pore size was, however, slightly reduced after the incorporation of CeO2 nanoparticles. Accordingly, the estimated pore size were 3.9 ± 0.2 μm, 2.1 ± 0.1 μm, 2.6 ± 0.2 μm and 2.7 ± 0.2 μm for GO:
AsA 1
:
4, GO
:
AsA 1
:
4 + CeO2 0.25, GO
:
AsA 1
:
4 + CeO2 0.5 and GO
:
AsA 1
:
4 + CeO2 1, respectively. The XRD showed a peak at 24° in all the samples (Fig. 2G), demonstrating that the interlayer distance was not affected by the addition of CeO2 nanoparticles. Nevertheless, the intensity of the peak dropped down with the addition of CeO2 nanoparticles to the samples in a dose dependent manner, denoting an interlayer distortion of the GO sheets, mediated by the CeO2 nanoparticles. Remarkably, in all the samples there was a peak at 43° which corresponded to the plane of the graphene layer48 and indicated that the CeO2 nanoparticles were localized on the interlayer space without inducing any variation in the GO sheets.49
The mechanical behavior of the hydrogels decorated with CeO2 nanoparticles was again measured by rheology. The GO:
AsA 1
:
4 + CeO2 0.25 and GO
:
AsA 1
:
4 + CeO2 0.5 (187.7 ± 7.5 kPa and 164.6 ± 1.6, respectively) showed a shear modulus similar to the GO
:
AsA 1
:
4 hydrogel (178.4 ± 2.8 kPa). A gradual decrease on the mechanical properties was however observed in the GO
:
AsA 1
:
4 + CeO2 0.6 (71.6 ± 0.7 kPa), GO
:
AsA 1
:
4 + CeO2 0.75 (11.8 ± 0.1 kPa) and GO
:
AsA 1
:
4 + CeO2 1 (9.7 ± 0.1 kPa) hydrogels (Fig. 2H). The fact that the shear modulus decreased with the addition of more CeO2 nanoparticles to the samples indicated the possibility to tune the mechanical properties of the hydrogel by just modifying the CeO2 nanoparticle concentration. These results are in line with those obtained by other groups using dopant nanoparticle substances like platinum50 and can be explained due to the less organized structures associated to the incorporation of the CeO2 nanoparticles in the interlayer space of the GO sheets, as suggested by XRD.
We also measured the electrical properties of the hydrogels. GO:
AsA 1
:
4 + CeO2 0.25 and GO
:
AsA 1
:
4 + CeO2 0.5 (22 S m−1 and 17 S m−1) maintained almost the same electrical conductivity of the GO
:
AsA 1
:
4 (27 S m−1) (Fig. 2I). However, the addition of an increasing amount of CeO2 nanoparticles resulted in a decline of the electrical conductivity of the GO
:
AsA 1
:
4 + CeO2 0.6, GO
:
AsA 1
:
4 + CeO2 0.75 and GO
:
AsA 1
:
4 + CeO2 1 hydrogels (8.5 S m−1; 2.8 S m−1 and 1.6 S m−1 respectively), which may be attributed to the more disordered arrangement of the GO sheets due to the incorporation of the CeO2 nanoparticles in the interlayer space. Nevertheless, the electrical values obtained proved that all our hydrogels could potentially trigger the electrical excitation of the seeded neural cells.43,44
It is well known that cerium oxide can exhibit +3 and +4 states that support the formation of CeO2 and CeO2−x and provides antioxidant properties.51 CeO2 nanoparticles resemble the antioxidant enzymes superoxide dismutase (SOD) and catalase (CAT) and, hence, scavenge reactive oxygen species (ROS) and free radicals51,52 in physiological conditions.45,53,54 Here, we applied a concentration of 50 μM hydrogen peroxide (H2O2) and measured the antioxidant capabilities of the hydrogels containing increasing amounts of CeO2 nanoparticles. It was reported in the literature that this H2O2 concentration is able to mimic the pro-oxidative environment found in vivo which may cause a detrimental effect on important cellular structures, thus leading to oxidative distress.55 As expected, all the hydrogels containing CeO2 nanoparticles were able to reduce the hydrogen peroxide concentration in a dose dependent manner. In contrast, the GO:
AsA 1
:
4 sample had no antioxidant properties, demonstrating that the CeO2 nanoparticles were responsible of the decline on the hydrogen peroxide concentration (Fig. 2J). These results are in agreement with other studies were CeO2 nanoparticles have demonstrated to possess antioxidant properties in vitro and in vivo, creating a more favorable microenvironment for angiogenesis and nerve reconstruction, resulting accordingly in a neuroprotective effect.18,56
NSCs were seeded directly on the hydrogels and, as shown by SEM micrographs, were able to attach without a need of Fetal Bovine Serum (FBS) supplementation or extracellular matrix (ECM)-derived compound coating in 24 h (ESI Fig. 3†). They could even migrate inside the hydrogels after 7 days in culture (ESI Fig. 4†), demonstrating a good infiltration of the cells in the material. Indeed, NSC survival, adhesion and infiltration are crucial to facilitate the bench to clinic implementation of these hydrogels. Besides, the use of ECM-derived coatings like laminin have been correlated with cell proliferation of brain cancers such ependymoma or glioblastoma.58 Thus, by eluding its use, we also avoid the possibility of having degradation products that could represent a lethal risk in clinical approaches.59 In the same way, fetal serum supplementation in cell culture has been ascribed to be strongly immunogenic in both rodents and humans. Hence, by eluding its use in here, we also maximize the bench to clinic translation potential of our systems.
To assess the impact of the hydrogels on the establishment of a heterogeneous culture for future neural regeneration, after 7, 14 and 21 days in vitro (DIV), cells were fixed and immunostained for neural stem (Nestin), astroglial (GFAP and S100β), oligodendroglial (Olig2) or neuronal lineage markers (DCX and MAP2) to study the effect of each hydrogel on the differentiation fate of the NSCs. Additionally, RNA was extracted and quantitative retro transcriptase polymerase chain reaction (qPCR) was performed against Nestin, GFAP, S100β and MAP2 to better characterize the gene expression profile at messenger RNA (mRNA) level and corroborate the immunolabeling results.
NSC are remnant cells of the CNS development restricted in neurogenic niches and maintain the ability to differentiate towards neuronal and glial lineages.69 A balanced neuronal and glial cell differentiation is important for the long-term survival of the NSC-derived cell cultures due to the supporting function of the glial cells over the new-born neurons on the regulation of their oxidative and metabolic balance, and the neurophysiological processes of ion and neurotransmitter uptake and release, among others.11,70
We also characterized the NSC differentiation process towards astroglial lineages, by the loss of Nestin, the increase of the expression of glial fibrillary acidic protein (GFAP) and the appearance of the S100 calcium-binding protein β (S100β) marker during the astroglial differentiation (Fig. 3 and 4). The percentage of GFAP positive cells at DIV7 indicated that GO:
AsA 1
:
4 substrate promoted more astroglial differentiation (22.6 ± 8.2%, p < 0.05) than GO
:
AsA 1
:
1 (5.6 ± 3.4%, p < 0.05) or GO
:
AsA 1
:
4 + CeO2 0.25 (3.7 ± 0.2%, p < 0.05) hydrogels (Fig. 3B). At mRNA level, GO
:
AsA 1
:
4 also exhibited a higher expression of GFAP (9.7 ± 1.7, fold change, p < 0.05), in comparison with GO
:
AsA 1
:
1 (1.0 ± 0.2, fold change, p < 0.05) and GO
:
AsA 1
:
4 + CeO2 0.25 (0.7 ± 0.1, fold change, p < 0.05) (Fig. 3C). These results suggested that, as previously described, stiffer substrates like GO
:
AsA 1
:
4 promoted the differentiation towards glial lineages compared to softer ones.71,72
However, the presence of CeO2 nanoparticles could modulate the final fate of this astroglial phenotype into not only astrocytes, but also other types of glial cells or even the enhanced maintenance of non-differentiated stem cell phenotypes. In this sense, GO:
AsA 1
:
4 + CeO2 0.25 exhibited an augmented GFAP expression on mRNA level at DIV14 (66.3 ± 2.4, fold change, p < 0.05), in comparison with GO
:
AsA 1
:
1 (9.3 ± 2.8, fold change, p < 0.05) and GO
:
AsA 1
:
4 (12.7 ± 2.6, fold change, p < 0.05) (Fig. 3C). However, this expression was correlated with a similar increase on Nestin mRNA level, suggesting a possible expansion of the NSC population, rather than an astroglial cell differentiation. Indeed, the co-expression of GFAP and Nestin has been ascribed to the stem phenotype of neural cells and their proliferation.73,74
Interestingly, at DIV21 we found an abundant GFAP positive astroglial cell subpopulation with no Nestin expression in GO:
AsA 1
:
1 (36.5 ± 1.8%, p < 0.05) and GO
:
AsA 1
:
4 (31.1 ± 1.7%, p < 0.05) hydrogels, compared to GO
:
AsA 1
:
4 + CeO2 0.25 hydrogels where those GFAP + cells were relatively much less frequent (1.7 ± 0.1%, p < 0.05) (Fig. 3B). At mRNA level, GO
:
AsA 1
:
1 also exhibited a higher GFAP expression with respect to DIV7 (20.8 ± 0.9, fold change, p < 0.05), almost equaling the levels of the GO
:
AsA 1
:
4 (25.5 ± 0.8, fold change, p < 0.05). Remarkably, GO
:
AsA 1
:
4 + CeO2 0.25 exhibited a lower GFAP mRNA expression (11.2 ± 2.2, fold change, p < 0.05) further corroborating the finding of the immunofluorescence assays (Fig. 3C).
From the perspective of tissue engineering therapies, the presence of GFAP positive astrocytes may pose both advantages and disadvantages. On the one hand, mature astrocytes are an important supporting glial cell of the CNS, sustaining neuronal metabolism and function. On the other hand, astroglial cells have also been described to be involved in glial scar formation after CNS injury.75 Scar tissue has a dual component: first, the glial scar formed by glial precursors, reactive astrocytes and microglia found at the periphery of the lesion, and second, the fibrotic scar composed by phagocytic cells and fibroblasts at the lesion core.76 This biological process has been reported detrimental for an effective reinnervation of the damaged CNS.77 Taking into account that reactive astroglial phenotypes can be induced when these cells are exposed to a damaged CNS environment, the presence of high amounts of astrocytes (GFAP positive astroglial subpopulation) in the tissue engineered grafts may constitute a limitation for therapeutic purposes.76
During the maturation of the astrocytes, together with the expression of GFAP, cells acquire the expression of S100β.78,79 All the tested hydrogels presented a mature astroglial-like lineage population over time since the proportion of S100β positive cells were always higher than that of GFAP positive cells in all the conditions at protein level (Fig. 4A and B). qPCR results further corroborated the presence of S100β protein in all the conditions and demonstrated a higher expression of S100β mRNA on GO:
AsA 1
:
4 (3.9 ± 0.8 fold change, p < 0.05) at DIV21, compared to GO
:
AsA 1
:
1 (0.6 ± 0.0 fold change, p < 0.05) and GO
:
AsA 1
:
4 + CeO2 0.25 (1.1 ± 0.2 fold change, p < 0.05) (Fig. 4C), suggesting a possible establishment of a mature astroglial differentiation on the stiffer hydrogels with smaller pores and greater mechanical and electrical properties. It is noteworthy that S100β, apart from mature astroglial cells, is also a marker of oligodendroglial lineage cells.80 Oligodendrocyte precursors are known to express both S100β and Olig2 markers.80 Therefore, in view of this result, we also studied the expression of oligodendrocyte transcription factor Olig2 on the cell-seeded hydrogels. The expression of Olig2 in oligodendrocyte progenitors has been reported to be increased during the remyelination process of injured axons in multiple sclerosis (MS),81 and is characteristic of an in situ expanding oligodendrocyte population. At DIV7, the proportion of the oligodendroglial lineage population on the GO
:
AsA 1
:
1 (5.9 ± 0.4%, p < 0.05) was also lower compared to GO
:
AsA 1
:
4 (88.5 ± 3.0%, p < 0.05) and GO
:
AsA 1
:
4 + CeO2 0.25 (68.2 ± 0.2%, p < 0.05) (Fig. 4B). These results were further corroborated at mRNA level, where GO
:
AsA 1
:
4 at DIV7 exhibited a significantly higher Olig2 expression (6.8 ± 0.2, fold change, p < 0.05) compared to GO
:
AsA 1
:
1, demonstrating again that stiffer substrates promoted a faster differentiation towards glial cell lineages (Fig. 4C).
Interestingly, GO:
AsA 1
:
4 + CeO2 was able to support much better the oligodendroglial cell population for long culture periods until DIV21 (27.6 ± 0.2%, p < 0.05) compared to GO
:
AsA 1
:
1 (5.4 ± 0.1%, p < 0.05) and GO
:
AsA 1
:
4 (10.3 ± 0.5%, p < 0.05) (Fig. 4B), although at mRNA level all the hydrogels presented a decay on Olig2 expression (Fig. 4C). These results suggested that the presence of CeO2 nanoparticles might have helped on the establishment of a mature and healthy oligodendroglial lineage subpopulation expressing Olig2 protein within the hydrogel. Oligodendrocytes are known to be a particularly sensitive cell type to oxidative stress.82,83 Although the antioxidant effect of CeO2 0.25 using a concentrated source of exogenous H2O2 was limited, it might modulate the physiological levels of intracellular free radicals, which may explain the beneficial effect of the CeO2 nanoparticles addition on oligodendroglial survival in the scaffolds. Indeed, CeO2 nanoparticles have been reported to possess antioxidant and neuroprotective capabilities and even attenuate the inflammation and help on the recovery of demyelinating pathologies like MS,84 further suggesting their implication on oligodendrocyte function and survival. Our results also show that, even if no statistical difference was observed for the GO
:
AsA 1
:
4 + 0.25 CeO2 formulation on H2O2 reduction, the added CeO2 nanoparticles managed to decrease the accumulation of intracellular reactive oxygen species in NSCs (GO
:
AsA 1
:
1 100.0 ± 5.8, GO
:
AsA 1
:
4 88.7 ± 6.0 and GO
:
AsA 1
:
4 + CeO2 0.25 48.7 ± 2.2, p < 0.05) (ESI Fig. 5†), enhancing the survival and modulating the differentiation pattern of these cells. This may be ascribed to the reduction of other reactive oxygen species or via other mechanisms beyond the scope of this study.
Overall, our results indicated that stiffer substrates like GO:
AsA 1
:
4 promoted astroglial and oligodendrogial differentiation. Nevertheless, both GO
:
AsA 1
:
1 and GO
:
AsA 1
:
4 hydrogels were able to support a GFAP positive astrocyte-like subpopulation. Remarkably, the addition of CeO2 nanoparticles on the GO
:
AsA 1
:
4 + CeO2 hydrogel induced a much better maintenance of both Nestin + stem cell populations and also Olig2 + oligodendroglial lineage cell populations for longer culture periods until DIV21.
The functionality and successful engraftment of bioengineered nerve tissues implies a balanced generation of both mature and immature glial and neuronal cells. Hence, we also studied the expression of other mature neuronal markers like microtubule-associated protein 2 (MAP2), which is expressed in the dendrites of fully mature neurons.89 The percentage of MAP2 positive neurons in our hydrogels was much higher in GO:
AsA 1
:
4 + CeO2 0.25 (35.4 ± 0.2%, p < 0.05), than in GO
:
AsA 1
:
1 (0.04 ± 0.03% p < 0.05) and GO
:
AsA 1
:
4 (0.03 ± 0.02%, p < 0.05), which exhibited almost no mature neuronal generation at DIV7 (Fig. 5A and B). These results were further corroborated by qPCR, where GO
:
AsA 1
:
4 + CeO2 0.25 exhibited a higher expression for MAP2 (26.1 ± 7.5 fold change, p < 0.05) compared to GO
:
AsA 1
:
1 and GO
:
AsA 1
:
4 at DIV7 (Fig. 5C). On the contrary, no MAP2 positive cells were ever found at any time point on GO
:
AsA 1
:
4 hydrogel, but qPCR results showed an increase on MAP2 expression at DIV21 (2.3 ± 0.7 fold change p < 0.05 compared to DIV7), suggesting a possible delay of the neuronal differentiation on stiffer substrates with smaller pores and greater electrical conductivity due to the favored differentiation towards astroglial lineages.71,72 These results also corroborated the positive effect of the addition of CeO2 nanoparticles for a quicker neuronal differentiation over GO based 3D materials.
Interestingly, GO:
AsA 1
:
1 showed an increase on MAP2 positive cells (22.9 ± 0.4%, p < 0.05) at DIV14, and at DIV21 on mRNA content (9.4 ± 0.8 fold change, p < 0.05 compared to DIV7) (Fig. 5B and C). These results are in agreement with DCX results in GO
:
AsA 1
:
1 at DIV7, which exhibited an increase in DCX positive cells. It might be assumed that some of these progenitor cells would proceed through their differentiation process to eventually give rise to mature neurons at DIV14 and the higher expression of MAP2 at DIV21 and corroborates the findings of other studies in the literature, where softer substrates were reported to promote the differentiation towards neuronal lineages.71,72
However, the GO:
AsA 1
:
1 hydrogel was unable to support the survival of the MAP2 positive cells for longer periods (Fig. 5B). On the contrary, the GO
:
AsA 1
:
4 + CeO2 0.25 hydrogel managed to support a subpopulation of MAP2 positive mature neurons until DIV21. Our findings suggested that the addition of CeO2 nanoparticles to the hydrogels allowed them to support the terminal differentiation of NSCs towards fully mature neurons. It is noteworthy that here CeO2 nanoparticles were physically attached to the hydrogels, thus possibly preventing their cellular internalization and possible detrimental effects on the neuronal lineage differentiation of the seeded NSCs, as it has been suggested in other studies.90 Moreover, in accordance with our study, CeO2 nanoparticles have also been shown to protect cells against oxidative stress, improving neuronal function and delaying neuronal death after a traumatic brain injury both in vitro and in vivo.91
As previously stated, a glial and neuronal equilibrium is primordial for cell survival and functionality of nerve tissues.11 Herein, in the absence of CeO2 nanoparticles (GO:
AsA 1
:
1 and 1
:
4 hydrogels), the larger pores and lower stiffness and electrical conductivity of the GO
:
AsA 1
:
1 hydrogel enhanced the generation of neuronal lineage cells with respect to the smaller pores and greater mechanical and electrical properties of the GO
:
AsA 1
:
4, where there is a preferred differentiation towards glial lineages at shorter periods (Fig. 5D). But, remarkably, at DIV21, in both hydrogels, GO
:
AsA 1
:
1 and GO
:
AsA 1
:
4, astrocytes were the major persisting cell type, despite the detection of some populations of neuronal lineage cells at shorter time points, DIV7 and DIV14. This result comes in agreement with the fact that astrocytes are the metabolically most resistant (least demanding) cell type of the CNS, whose high endogenous antioxidant and glycolytic capacity endows them with a higher ability to survive in adverse conditions.83,92,93 It is very likely that most of the DCX+ and MAP2+ neuronal cells that were being generated in the GO
:
AsA 1
:
1 hydrogel at DIV7 and DIV14 eventually perished at DIV21, because of an insufficient antioxidant capacity to support the increased mitochondrial oxidative phosphorylation that comes along with mature neuronal differentiation.94 Interestingly, in the presence of CeO2 nanoparticles, the GO
:
AsA 1
:
4 + CeO2 0.25 hydrogel supported the co-generation of both MAP2 positive mature neuronal lineage cells together with Olig2 positive oligodendroglial lineage cells until DIV21. The results of this work clearly encourage the incorporation of neuroprotective and antioxidant systems like CeO2 nanoparticles trapped in tissue engineering scaffolds to boost survival of these two extremely necessary and highly vulnerable cell types of the CNS. The balanced generation of neurons, astrocytes and oligodendrocytes within the bioengineered construct is fundamental for an eventual success of CNS regeneration therapies. It should be emphasized that once a balanced oligodendrocytes, neurons and astrocytes population has been established within the graft, the close contact of the three different cell types may protect each other and improve xenocell survival prior to the integration into the host tissue.95,96 Indeed, several studies highlighted the necessary integrin mediated connexion between neuronal axons and oligodendrocytes for the survival of both neuronal and oligodendroglial cells in vitro.97,98 Moreover, oligodendrocytes have also been shown to be a glial cell subpopulation that play a key role on axonal regeneration,98,99 hence the importance of preserving both neurons and oligodendrocytes together in the same bioengineered construct. Here we present a 3D hydrogel based on graphene derivatives and cerium oxide nanoparticles as a promising therapeutic tool for neurodegenerative and demyelinating pathologies involving neuronal and/or oligodendroglial cell death.
Overall, our results showed that hydrogels based on graphene-derivatives supported both glial and neuronal lineage differentiation of NSCs in short-term cultures. Softer substrates like GO:
AsA 1
:
1 with larger pores and lower electrical conductivities promoted cell differentiation towards neuronal lineages, while stiffer substrates like GO
:
AsA 1
:
4 with smaller pores and greater electrical properties enhanced glial cell differentiation. However, for long-term cultures, graphene derivatives-based hydrogels alone were unable to sustain a balanced long-term survival of neurons and oligodendrocytes. In contrast, thanks to the antioxidant and neuroprotective capabilities of CeO2 nanoparticles embedded on GO
:
AsA 1
:
4 + CeO2 0.25, this hydrogel was able to support the generation of astroglial, oligodendroglial and neuronal cells until DIV21, providing a promising approach for CNS regeneration therapies.
This work has been funded by the Basque Government (GV/EJ) Department of Education (GIC21/131 IT1766-22, IT1751-22), Health Department (RIS3, 2021333012), Grants PID2019-106236GB-I00 and PID2019-104766RB-C21 funded by MCIN/AEI/10.13039/501100011033. Grant RYC-2013-13450 funded by MCIN/AEI/10.13039/501100011033 and by “ESF investing in your future” by the “European Union” and Achucarro Seed-Fund 003 (JRP), the University of the Basque Country (UPV/EHU) by GIU19/040, GIU 20/050, PPGA 20/22, COLAB19/03 and IKERTU-2020.0155. GV/EJ, Hazitek ZE-2019/00012-IMABI and ELKARTEK KK-2019/00093. Polimerbio and Y. P. have a Bikaintek PhD grant (20-AFW2-2018-00001). Part of this research was performed within the framework of the European Union's Horizon 2020 research and innovation program under the Marie Skłodowska-Curie Grant Agreement No. 793644 (BIONICS).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2nr06545b |
This journal is © The Royal Society of Chemistry 2023 |