Skip to main content

Neural stem/progenitor cells from adult canine cervical spinal cord have the potential to differentiate into neural lineage cells

Abstract

Background

• Neural stem/progenitor cells (NSPCs) are multipotent self-renewing cells that can be isolated from the brain or spinal cord. As they need to be isolated from neural tissues, it is difficult to study human NSPCs. To facilitate NSPC research, we attempted to isolate NSPCs from dogs, as dogs share the environment and having many similar diseases with humans. We collected and established primary cultures of ependymal and subependymal cells from the central canal of the cervical spinal cord of adult dogs. To isolate pure NSPCs, we employed the monolayer culture and selective medium culture methods. We further tested the ability of the NSPCs to form neurospheres (using the suspension culture method) and evaluated their differentiation potential.

Results

• The cells had the ability to grow as cultures for up to 10 passages; the growth curves of the cells at the 3rd, 6th, and 9th passages showed similar patterns. The NSPCs were able to grow as neurospheres as well as monolayers, and immunostaining at the 3rd, 6th, and 9th passages showed that these cells expressed NSPC markers such as nestin and SOX2 (immunofluorescent staining). Monolayer cultures of NSPCs at the 3rd, 6th, and 9th passages were cultured for approximately 14 days using a differentiation medium and were observed to successfully differentiate into neural lineage and glial cells (astrocytes, neurons, and oligodendrocytes) at all the three passages tested.

Conclusion

• It is feasible to isolate and propagate (up to at least 10 passages) canine cervical spinal cord-derived NSPCs with the capacity to differentiate into neuronal and glial cells. To the best of our knowledge this is the first study to successfully isolate, propagate, and differentiate canine NSPCs derived from cervical spinal cord in the adult canine, and we believe that these cells will contribute to the field of spinal cord regeneration in veterinary and comparative medicine.

Peer Review reports

Background

Damage to the spinal cord can be devastating, as the neurons in the spinal cord do not regenerate easily, and the resulting loss of sensory and motor function can be permanent [1, 2]. In addition, such spinal cord injuries (SCIs) impose a high societal economic burden [3]. To date, there is no effective treatment for SCI, though this has been a popular topic of research [4, 5].

Stem cell transplantation in SCI has been suggested to have a beneficial effect on reconstructing neural circuits and remyelination [6,7,8]. Among stem cells, Neural stem/progenitor cells (NSPCs) are a potential source that has been demonstrated in in vivo models, resulting in post-mitotic astrocytes, oligodendrocytes, and neurons [9,10,11,12]. Furthermore, NSPCs have the ability to support neural protection, and neurite outgrowth; They also exert anti-inflammatory and antioxidant effects [13, 14]. These points highlight the potential of NPCs, however, it is difficult to study these cells in humans, because human-derived NSPCs must be harvested from neural tissues. Therefore, this study focused on dogs, which share similarities with human SCI with regard to genetic background, severity and location, and pathological processes [15,16,17].

Ependymal cells lining the central canal of the spinal cord have the ability to differentiate into neural cells in vitro, and are referred to as NSPCs [18, 19]. Although neural stem cells have been isolated from fetal canine spinal cord [20], there is currently no evidence that these cells can be isolated from adult canine spinal cord. In adult human spinal cord, there is a report that ependymal and subependymal cells around the central canal express neural precursor markers [21]. Here, we report on the isolation and culturing of NSPCs from the cervical spinal cord (ependymal and subependymal cells around the central canal) of adult dogs; we also identify the differentiation and self-renewal capacities of these cells for the first time.

Results

NSPC morphology

Following the collection of the cervical spinal cord from the dogs, a selective isolation of the central canal of the spinal cord was performed, followed by its culture in a suitable medium. The passage 0 cells were initially cultured as neurospheres and subsequently transitioned to a monolayer culture during the subculture. Rounded neurospheres were observed after three days of primary culture of the isolated NSPCs. The diameters of the neurospheres ranged from approximately 50–150 μm (Fig. 1). The neurospheres showed uniformity in size and shape. In monolayer culture, NSPCs adhered to the bottom of cell culture plates and exhibited intercellular connection by developing cell protrusions. Cells at passages 3, 6, and 9 showed similar cell morphology and intercellular connection (Fig. 2).

Fig. 1
figure 1

Neurospheres obtained by suspension culture of (A) cells at passage 3, (B) cells at passage 6, and (C) cells at passage 9. The scale bar represents 200 μm. (D) Scheme of experiment: neural stem/progenitor cells (NSPCs) were serially passaged in monolayer from passage 0 to 10. Cells that were cultured with monolayer were additionally cultured in neurospheres form only in 3rd, 6th, and 9th passage cells for microscopic and immunostaining observation. Growth curve and cell differentiation analysis were conducted in monolayer

Fig. 2
figure 2

Cervical spinal cord-derived neural stem/progenitor cells (NSPCs) in monolayer culture (A) and (B) cells at passage 3, (C) and (D) cells at passage 6, and (E) and (F) cells at passage 9. NSPCs exhibited intercellular connection by developing cell protrusions. Dotted boxes represent cell protrusions. The scale bar represents 200 μm

Cell growth characteristics and viability

The NSPCs were serially passaged to analyze their growth characteristics. At each passage, cell counts, and viability assessments were performed using trypan blue staining. The NSPCs maintained their ability to proliferate without any change in the proliferation rate for up to 10 passages (Fig. 3A). Cell viability measurements were performed during passaging and the viability was found to be > 97% at all passages tested (Fig. 3B). The growth curves of the cells at passages 3, 6, 9 are shown in Fig. 4. The NSPCs grew slowly from day 1 to day 3 of culture. From 3 days in vitro (DIV) onwards, the NSPCs showed a logarithmic growth pattern followed by a decrease in the proliferation rate at DIV 5–7 (Fig. 4).

Fig. 3
figure 3

Growth characteristics and viability of neural stem/progenitor cells (NSPCs) (A) NSPCs were serially passaged in monolayer for five days until 80–90% confluency at each passage. The population-doubling level (PDL) was measured based on the average of the total number of three plate (60 mm PLO/fibronectin-coated petri dishes) cells in each passage. (B) Cell viability was measured based on the average viability of three plate cells in each passage

Fig. 4
figure 4

Growth curve of cells at each passage. Each passage of cells (1 × 104) was seeded in each well of 24-well plate. The cell count was measured based on the average of the total number of three well cells. Daily cell counts were performed over seven successive days to construct the growth curve

Characterization of NSPCs by immunofluorescence

Immunostaining was conducted on pre- and post-differentiation cells from the 3rd, 6th, and 9th passages. To determine the cell characteristics in the pre-differentiation state, NSPCs cultured as monolayers or spheres were subjected to immunostaining. The cells were positive for the NSPC markers nestin and SOX2 (Fig. 5A, B). In the quantified data, nestin+/SOX2+ cells were more than 95% in all the three passages (Fig. 5C). To determine the differentiation potential of the cells, the cells were grown in differentiation medium 3. The differentiated cells expressed glial fibrillary acidic protein (GFAP) (an astrocyte marker), β3-tubulin (a neuronal marker), and myelin basic protein (MBP) (an oligodendrocyte marker) (Fig. 6A, B, C). These pre- and post-differentiation markers were expressed in cells at all the three passages. In the quantified data, GFAP+, β3-tubulin+, and MBP+ cells showed approximately 30–40%, 50–60%, and 30–40%, respectively, in all the three passages (Fig. 6D). These results showed a similar pattern in mRNA expression, with β3-tubulin expression being the highest, and GFAP and MBP expression at similar levels (Fig. 6E).

Fig. 5
figure 5

Immunocytochemistry of neural stem/progenitor cells (NSPCs) cultured as neurospheres or monolayers. (A) 3rd, 6th, and 9th passage cells cultured in a monolayer for 5 DIV. (B) 3rd, 6th, and 9th passage cells cultured as neurosphere for 5 DIV. Nuclei were stained with DAPI (blue) and NSPCs marker for SOX2 (green) and nestin (red). (C) Each well of the 24-well plate was immunostained with nestin and SOX2. The quantification was performed by calculating the average percentage of nestin+/SOX2+ cells out of the total cells in three wells. Quantification was carried out with cells cultured in monolayer

Fig. 6
figure 6

Immunocytochemistry of differentiated neural stem/progenitor cells (NSPCs). Medium 3 was used to differentiate cells 3rd, 6th, and 9th passage cells. (A) astrocyte marker glial fibrillary acidic protein (GFAP) (red), culture for 14 DIV. (B) immature neuron marker β-3 tubulin (red), culture for 14 DIV. (C) mature oligodendrocyte marker myelin basic protein (MBP) (green), culture for 21 DIV. Nuclei were stained with DAPI (blue). (D) Each well of the 24-well plate was immunostained with single marker and quantified. The quantification was performed by calculating the average percentage of GFAP+, β-3 tubulin+, and MBP+ cells out of the total cells in three wells. (E) Comparison of gene expression of differentiated cells (passage 3). The data were obtained from experiments biologically repeated two times, with each bar representing an average of the gene expression calculated with the formula 2−ΔΔCT; CT value was normalized to the housekeeping gene glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and mRNA expression is shown relative to the expression in the GFAP group. The error bar represents standard deviation. * denotes significance compared with the GFAP group at p < 0.05. ^ denotes significance compared with the MBP group at p < 0.05

Real-time PCR analysis of NSPCs

Real-time PCR analysis was carried out to assess the cell differentiation in both early and late-passages cells, as well as to investigate the impact of the differentiation medium on the cell differentiation. The incubation period for mRNA acquisition was undifferentiated cells for 5 days, differentiation into neurons and astrocytes for 14 days, and differentiation into oligodendrocytes for 21 days.

Real-time PCR was performed to compare gene expression of pre- and post-differentiation cells at early and late passages (Fig. 7). Differentiated cells at passage numbers 3, 6, and 9 exhibited significantly reduced expression of nestin. However, SOX2 expression was significantly decreased only at passage 9. Therefore, the differentiated cells showed reduced expression of pre-differentiation markers. The differentiation-related genes GFAP, MBP, and β3-tubulin were highly expressed in the differentiated cells at all passage numbers. GFAP and MBP expression was significantly high only in cells at passages 6 and 9, but β3-tubulin was significantly high in cells at all the three passages (passage numbers 3, 6, and 9).

Fig. 7
figure 7

Comparison of gene expression in cells before and after differentiation. Medium 3 was used to differentiate cells. (A) Heat map of mRNA expression (B) Relative mRNA expression of NSPCs and differentiation markers. The data were obtained from experiments biologically repeated two times, with each bar representing an average of the gene expression calculated with the formula 2−ΔΔCT; CT value was normalized to the housekeeping gene glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and mRNA expression is shown relative to the expression in the P3 group (passage 3 cells). The error bar represents standard deviation. * denotes significance compared with the P3 group at p < 0.05. ^ denotes significance compared with the P6 group at p < 0.05. # denotes significance compared with the P9 group at p < 0.05

mRNA is quantified to evaluate gene expression of cell differentiation using FBS and N-2 supplement which is commonly added to the differentiation medium. Differentiation-related gene expression was compared according to the differentiation medium by real-time PCR (Fig. 8). The gene expression of GFAP was the highest in the medium 2 (+ N2) group. In contrast, the medium 3 (+ FBS) and 4 (+ N2 + FBS) groups showed lower GFAP gene expression than the medium 1 (no FBS/N2) and 2 (+ N2) groups. It is considered that N-2 supplement promotes differentiation into astrocytes, but FBS inhibits differentiation into astrocytes. MBP expression was significantly lower in the medium 3 (+ FBS) and 4 (+ N2 + FBS) groups than in the medium 1 (no FBS/N2) and 2 (+ N2) groups. β3-tubulin gene expression was significantly higher in the group cultured in medium 3 (+ FBS) and 4 (+ N2 + FBS), and among them, the medium 4 (+ N2 + FBS) group showed the highest expression. From this, FBS and N-2 supplement are believed to promote differentiation into neurons rather than to oligodendrocytes.

Fig. 8
figure 8

Comparison of gene expression of cells according to differentiation medium. Passage 3 cells was used. (A) Heat map of mRNA expression (B) Relative mRNA expression of differentiation markers. The data were obtained from experiments biologically repeated two times, with each bar representing an average of the gene expression calculated with the formula 2−ΔΔCT; CT value was normalized to the housekeeping gene glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and mRNA expression is shown relative to the expression in the Medium 1 group. The error bar represents standard deviation. * denotes significance compared with the Medium 1 group at p < 0.05. ^ denotes significance compared with the Medium 2 group at p < 0.05. # denotes significance compared with the Medium 3 group at p < 0.05

Discussion

SCI can be irreversible because the CNS has considerably low regenerative capacity [1, 6, 22]. After incomplete SCI, propriospinal circuits can play a role in restoring neuronal function, but regeneration is limited due to the unfavorable environment for cell to survive [23,24,25,26]. Therefore, many studies have been conducted on the treatment of SCI by NSPC transplantation [17, 27,28,29]. However, these cells need to be obtained from CNS tissues, and NSPCs obtained from rodents has limitation of inherent species difference compared to human neural stem cells [30,31,32].Ultimately, for the treatment of SCI in humans, studies with experimental animals with disease courses similar to those in humans, such as dogs, are needed [15,16,17]. Accordingly, we performed this study on cervical spinal cord-derived NSPCs from adult dogs; adult dog-derived NSPCs have not yet been studied in the context of stem cell therapy for SCI treatment. Here, we demonstrated the feasibility of isolating cervical spinal cord-derived NSPCs in adult dogs; we successfully cultured these cells and characterized them for the first time.

Various cell culture methods have been reported for culturing NSPCs [33,34,35]. In this study, ependymal and subependymal cells of central canal were isolated and cells were cultured in petridishes. After cells were first confirmed to form neurospheres, then cultured with a monolayer during successive subcultures. This method provides a uniform growth environment for all cells, including nutrition and oxygenation, to create homogenous cell populations [35,36,37]. Cells at passage numbers 3, 6, and 9 were examined and showed homogenous cellular morphology; the ability to form neurospheres was also confirmed. The cell viability was more than 97% at all the passage numbers tested. Hence, monolayer culture was considered as an effective way to maintain the properties of NSPCs, as cell debris could be effectively removed through a washing process and uniform nutrition conditions could be provided to cells through an even distribution of the culture medium [36].

We used selective serum-free media with growth factors such as epidermal growth factor (EGF) and basic fibroblast growth factor (bFGF). The serum-free medium inhibits cell differentiation; among the added growth factors, EGF promotes cell survival, while bFGF promotes cell proliferation and exhibits cell protection effects [38,39,40]. Our results showed that the cells exhibited homogenous morphology and growth patterns at all the three passages examined, and the growth curves were also similar. In addition, both nestin and SOX2 were expressed at similar levels by cells at all the passages examined (passage numbers 3, 6, and 9). It has been reported that when primary cells are cultured in selective medium, cell types other than the target cell type do not survive in the culture environment and gradually disappear [35, 41]. Therefore, the selective medium method was chosen for obtaining pure preparations of canine-derived NSPCs; these cells were confirmed to be multipotent with self-renewal capabilities that could be expanded.

Multipotent NSPCs have the ability to differentiate into both neuronal and glial cells [42, 43]. To confirm whether the NSPCs isolated in this study possessed this capability, cell differentiation was induced using a medium without growth factors, as reported previously [44,45,46]. We observed the expression of neuronal, astrocyte, and oligodendrocyte markers in NSPCs after the induction of differentiation. In real-time PCR analysis, differentiated cells showed decreased expression of NSPC markers and increased expression of differentiation markers. However, while nestin expression was significantly reduced at all passages, SOX2 expression was significantly reduced only at passage 9. It is believed that cell differentiation is suppressed because the stemness associated with SOX2 is strongly expressed at early passages [47, 48]. In this context, GFAP and MBP, which are differentiation markers, did not show significantly high expression at passage 3. It is possible that this is due to the relatively lower level of differentiation, because SOX2 expression at the early passages was reported to be higher than that at late passages among differentiated cells [49, 50]. These results suggest that the NSPCs were multipotent cells and that the stemness associated with SOX2 was highly expressed in the early passages.

Differentiation is usually induced using a medium without growth factors, though supplements such as N-2 supplement and fetal bovine serum (FBS) are also commonly added [44,45,46]. The gene expression profiles of differentiated NSPCs in a medium supplemented with N-2 supplement and FBS were investigated to evaluate their ability to differentiate into specific cells. GFAP expression was highest in the cells cultured in medium 2 (+ N2). This is similar to the results reported in a previous study where the cells were differentiated into astrocytes using a medium supplemented with N-2 supplement [51]. FBS is also widely used for the differentiation of NSPCs into neurons [45, 46, 52]. Similar to previous studies, the cells grown in a medium with FBS showed high expression of β3-tubulin; however, the expression of MBP and GFAP was low. Interestingly, the cells cultured in medium supplemented with FBS and N-2 supplement showed the highest expression of β3-tubulin. Considering that N-2 is also used in neuronal differentiation [44, 46], it seems that FBS promotes differentiation into neurons and N-2 has an additional synergistic effect. Our results indicated that using a specific medium composition can promote differentiation of NSPCs into the desired cell type.

In this study, differentiation of NSPCs was confirmed by immunofluorescence staining, and differentiation-related genes according to the differentiation medium were confirmed through real-time PCR. Therefore, differentiation markers for each differentiation medium were not visually confirmed through immunofluorescence staining and only a single marker was used for astrocyte, oligodendrocyte and neural lineage cells in both the immunofluorescence and real-time PCR. In particular, in the case of MBP+ cells, extended cellular processes should be seen as mature oligodendrocytes, but considering that they are not, double stain with β3-tubulin is required because they can be seen as myelinating neural type cells. Therefore, in the case of oligodendrocytes, additional characterization is required due to their unusual appearance. In real-time PCR, GAPDH, which is widely used as a reference gene, was used, but additional reference genes were not evaluated. Hence, it is considered that further studies will be needed to compare various fluorescent markers of each neural lineage under various medium conditions and to validate the optimal reference gene panel for real time PCR in canine NSPCs differentiation.

Conclusion

In this paper we describe the isolation and characterization of NSPCs from the cervical spinal cord of adult canines. We show that adult canine-derived NSPCs can proliferate and differentiate into neural lineage and glial cells. Currently, various cell therapy studies are being conducted for the treatment of SCI, and studying canine spinal cord-derived NSPCs would be relevant in this regard. We believe that the adult canine-derived NSPCs characterized in this study can be used in spinal cord regeneration studies; this is expected to contribute not only to veterinary medicine but also to comparative medicine.

Methods

Tissue collection

The spinal cord was collected from euthanized dogs. Two male beagle dogs were used (weights 10.26 and 10.02 kg; age: 1 to 2 years), sourced from experimental animal supplier Saeronbio Inc. (Uiwang, Korea). This study was conducted according to the Animal Care and Use Guidelines (The Institute of Laboratory Animal Resources, Seoul National University). The study on tissue collection was approved by the Animal Care and Use Committee of Seoul National University (SNU-210609-1-1).

Isolation and culture of NSPCs

For NSPC isolation, the anesthesia protocols for euthanasia were as follows: alfaxalone (2 mg/kg IV; Alfaxan; Jurox), tramadol (4 mg/kg IV, Toranzin; Samsung Pharm. Ind. Co., Seoul, Korea), and acepromazine (10 µg/kg IV, Sedaject; Samu Median Co., Seoul, Korea). The maintenance of anesthesia was done with isoflurane (Ifran; Hana Pharm. Co., Seoul, Korea). Then, dorsal laminectomy was performed at cervical vertebrae 1–5 to expose the spinal cord. After injecting alfaxalone (2 mg/kg IV; Alfaxan; Jurox), 10 mL KCl solution (1 M) was injected to the cephalic vein for euthanasia. After confirming that euthanasia was performed through auscultation, spinal cord collection was carried out immediately, and it took about 1 min. Spinal cord tissue was transferred to 60-mm petridishes and washed with phosphate-buffered saline (PBS) to remove debris and blood vessels. Tissues around the central canal (ependymal and subependymal cells) of the spinal cord were collected under a dissecting microscope. These tissues are homogenized by using a microscissors and manually driven syringe and filtered through a 100-mm-pore nylon mesh. The filtered tissue homogenates were centrifuged at 220 × g for 5 min at 4 °C and then resuspended in culture medium. The culture medium consisted of Dulbecco’s Modified Eagle Medium/Nutrient Mixture F-12 (DMEM/F-12) with GlutaMAX™ (Gibco, Carlsbad, CA, USA), 2% B27 supplement (Gibco), 1% penicillin and streptomycin (PS; Gibco), 20 ng/mL EGF (R&D Systems, Minneapolis, MN, USA), and 20 ng/mL bFGF (R&D Systems).

First, the cells were cultured in a petri dish until the neurosphere was formed. Then, in the subculture, cells were cultured in monolayer. To establish monolayers, petridishes were coated with 15 µg/mL poly-L-ornithine solution (PLO; Sigma-Aldrich, St. Louis, MO, USA) and 1 µg/mL fibronectin solution (Sigma-Aldrich). While growing cells in monolayers, debris was removed through washing with PBS when changing medium. EGF and bFGF in the medium selectively proliferate NSPC while differentiated cells die. Cells were incubated at 37 °C with 20% O2 and 5% humidified CO2. The medium was replaced every 24 h until 80–90% confluence was attained (approximately five days). For subculture, the cells were disassociated with the StemPro™ Accutase™ Cell Dissociation Reagent (Gibco). This method was carried out up to the 10th passage.

In order to observe neurosphere formation in the 3rd, 6th, and 9th passage, cells from the previous passage that grew in the coating dish were collected and transferred to the uncoated petridishes and cultured. Cells were incubated at 37 °C with 20% O2 and 5% humidified CO2. The medium was replaced every 48 h by inspecting it under a microscope to ensure that it did not contain a neurosphere.

Differentiation

PLO/fibronectin-coated 12 mm cover slips were placed in 24-well plates. For differentiation induction, 2 × 104 cells were seeded in each well and allowed to grow. The differentiation into neurons and astrocytes was cultured for 14 days and the differentiation into oligodendrocytes was cultured for 21 days. Four types of differentiation media were used to evaluate the differentiation ability of the NSPCs, as follows: medium 1, DMEM/F-12 with GlutaMAX™ and 2% B27 supplement; medium 2: DMEM/F-12 with GlutaMAX, 2% B27 supplement, and 1% N-2 supplement (Gibco); medium 3: DMEM/F-12 with GlutaMAX™, 2% B27 supplement, and 1% FBS (Gibco); and medium 4: DMEM/F-12 with GlutaMAX, 2% B27 supplement, 1% FBS, and 1% N-2 supplement. In real-time PCR, medium 3, a basic differentiation medium that does not contain other supplements, was used.

Cell growth and viability

NSPCs were cultured in monolayer for five days until 80–90% confluency was attained, and passaged into three 60 mm PLO/fibronectin-coated petri dishes. At each passage, three plate cells were counted and viability assessments were performed using a Countess FL Automated Cell Counter (Thermo Fisher Scientific, Pittsburg, PA, USA) which is using trypan blue staining. The population doubling level (PDL), a cumulative counting technique, was applied using the following formula: PDL(n/n−1) = log (Nf/N0)/log 2, where n = passage number, Nf = final number of cells, and N0 = number of cells seeded at passage. To obtain the new PDL(n + 1), PDL(n/n − 1) was added to the previous PDL(n) [53].

For growth curve analysis, cells at the 3rd, 6th, and 9th passage were cultured in 24-well plates (coated 12-mm cover slips were inserted into each well). Cell counts were obtained daily for seven days using trypan blue to derive a weekly growth curve.

Immunofluorescence

Pre- and post-differentiation 3rd, 6th, and 9th passage cells were immunostained. The cells were fixed using 4% paraformaldehyde in PBS for 10 min at room temperature and then permeabilized with 0.1% Triton X-100 in PBS for 5 min. Blocking was performed using 10% FBS in PBS for 30 min. The following primary antibodies were used: nestin (1:400; sc-33,677; Santa Cruz Biotechnology, Dallas, TX, USA) and SOX2 (1:400; bs-23176R-A488; Bioss, Massachusetts, USA) (neural progenitor markers); GFAP (1:400; sc-33,673; Santa Cruz Biotechnology) (glial marker); MBP (1:400; bs-0380R-A488; Bioss) and β3-tubulin (1:400; sc-80,005; Santa Cruz Biotechnology) (neuronal markers). The cells were incubated with the primary antibodies overnight at 4 °C. After washing 3 times with PBS, Secondary antibodies (1:500; sc-533,656; Santa Cruz Biotechnology) were applied for 1 h at RT. The nuclei were stained with 4,6-diamidino-2-phenylindole (DAPI). The stained cells were then examined under a microscope (EVOS FL Imaging System, Stanwood, WA, USA). To exclude non-specific staining, differentiation markers were confirmed as not expressed in pre-differentiation cells which is cultured for 5 DIV (Supplement 1). Quantification was performed using (Image-J version 1.52a; National Institute of Health, Bethesda, MD, USA) by calculating the average percentage of fluorescence+ cells out of the total cells in three out of 24-well plates.

Real-time PCR

mRNA isolation was performed using QIAzole (Qiagen, Hilden, Germany) according to the manufacturer’s instructions. The purity and concentration of RNA were evaluated using an Epoch reader and Gen 5.2 reader (BioTek, Winooski, Vermont, VT, USA). A Prime Script II First-Strand cDNA Synthesis Kit (TaKaRa, Otsu, Japan) was used to synthesize cDNA. Subsequently, RT-PCR was performed using the ABI StepOnePlus Real-Time PCR system (Applied Biosystems, Foster City, CA, USA). The RT-PCR reaction mixture contained the following: primers (Table 1), cDNA, and SYBR Premix Ex Taq (TaKaRa). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH), a housekeeping gene, was used to normalize the mRNA expression data. Gene expression was compared using the 2−ΔΔCT method [54]. The data were obtained from experiments biologically repeated two times.

Table 1 Primer sequences used for real-time quantitative PCR

Statistical analysis

Quantitative data was analyzed using GraphPad Prism 8.0.1. and represented by mean ± standard deviation (SD). Two-way ANOVA analyses was performed with Tukey’s multiple comparison. Statistical significance was defined as p < 0.05.

Data Availability

All data generated or analysed during this study are included in this published article.

Abbreviations

NSPCs:

neural progenitor cells

SCI:

spinal cord injury

CNS:

central nervous system

GFAP:

glial fibrillary acidic protein

MBP:

myelin basic protein

DIV:

days in vitro

EGF:

epidermal growth factor

bFGF:

basic fibroblast growth factor

FBS:

fetal bovine serum

GAPDH:

glyceraldehyde-3-phosphate dehydrogenase

PDL:

population doubling level

PBS:

phosphate-buffered salin

DMEM/F-12:

Dulbecco’s Modified Eagle Medium/Nutrient Mixture F-12

PS:

penicillin and streptomycin

PLO:

poly-L-ornithine solution

DAPI:

4,6-diamidino-2-phenylindole

References

  1. Bradke F, Fawcett JW, Spira ME. Assembly of a new growth cone after axotomy: the precursor to axon regeneration. Nat Rev Neurosci. 2012;13:183–93.

    CAS  PubMed  Google Scholar 

  2. Nas K, Yazmalar L, Şah V, Aydın A, Öneş K. Rehabilitation of spinal cord injuries. World J Orthop. 2015;6:8–16.

    PubMed  PubMed Central  Google Scholar 

  3. Malekzadeh H, Golpayegani M, Ghodsi Z, Sadeghi-Naini M, Asgardoon M, Baigi V, et al. Direct cost of illness for spinal cord injury: a systematic review. Glob Spine J. 2022;12:1267–81.

    Google Scholar 

  4. Harvey LA. Physiotherapy rehabilitation for people with spinal cord injuries. J Physiother. 2016;62:4–11.

    PubMed  Google Scholar 

  5. Huang L, Fu C, Xiong F, He C, Wei Q. Stem cell therapy for spinal cord injury. Cell Transpl. 2021;30:963689721989266.

    Google Scholar 

  6. Yang B, Zhang F, Cheng F, Ying L, Wang C, Shi K, Wang J, Xia K, Gong Z, Huang X. Strategies and prospects of effective neural circuits reconstruction after spinal cord injury. Cell Death Dis. 2020;11:439.

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Ben-Hur T. Reconstructing neural circuits using transplanted neural stem cells in the injured spinal cord. J Clin Invest. 2010;120:3096–98.

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Hou Y, Liu X, Guo Y, Liu D, Guo P, Liu J. Strategies for effective neural circuit reconstruction after spinal cord injury: use of stem cells and biomaterials. World Neurosurg. 2022;161:82–9.

    PubMed  Google Scholar 

  9. Cizkova D, Kakinohana O, Kucharova K, Marsala S, Johe K, Hazel T, Hefferan MP, Marsala M. Functional recovery in rats with ischemic paraplegia after spinal grafting of human spinal stem cells. Neurosci. 2007;147:546–60.

    CAS  Google Scholar 

  10. Hefferan MP, Galik J, Kakinohana O, Sekerkova G, Santucci C, Marsala S, Navarro R, Hruska-Plochan M, Johe K, Feldman E. Human neural stem cell replacement therapy for amyotrophic lateral sclerosis by spinal transplantation. PLoS ONE. 2012;7:e42614.

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Kakinohana O, Juhasova J, Juhas S, Motlik J, Platoshyn O, Galik J, Hefferan M, Yuan S, Vidal J, Carson C. Survival and differentiation of human embryonic stem cell-derived neural precursors grafted spinally in spinal ischemia-injured rats or in naive immunosuppressed minipigs: a qualitative and quantitative study. Cell Transpl. 2012;21:2603–19.

    CAS  Google Scholar 

  12. Rosenzweig ES, Brock JH, Lu P, Kumamaru H, Salegio EA, Kadoya K, Weber JL, Liang JJ, Moseanko R, Hawbecker S. Restorative effects of human neural stem cell grafts on the primate spinal cord. Nat med. 2018;24:484–90.

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Martino G, Pluchino S. The therapeutic potential of neural stem cells. Nat Rev Neurosci. 2006;7:395–406.

    CAS  PubMed  Google Scholar 

  14. Mothe AJ, Tator CH. Review of transplantation of neural stem/progenitor cells for spinal cord injury. Int J Dev Neurosci. 2013;31:701–13.

    CAS  PubMed  Google Scholar 

  15. Moore SA, Granger N, Olby NJ, Spitzbarth I, Jeffery ND, Tipold A, et al. Targeting translational successes through CANSORT-SCI: using pet dogs to identify effective treatments for spinal cord injury. J Neurotrauma. 2017;34:2007–18.

    PubMed  PubMed Central  Google Scholar 

  16. Moore SA, Zidan N, Spitzbarth I, Nout-Lomas YS, Granger N, da Costa RC, Levine JM, Jeffery ND, Stein VM, Tipold A. Development of an International Canine spinal cord Injury observational registry: a collaborative data-sharing network to optimize translational studies of SCI. Spinal Cord. 2018;56:656–65.

    PubMed  PubMed Central  Google Scholar 

  17. Kirkness EF, Bafna V, Halpern AL, Levy S, Remington K, Rusch DB, Delcher AL, Pop M, Wang W, Fraser CM. The dog genome: survey sequencing and comparative analysis. Science. 2003;301:1898–903.

    PubMed  Google Scholar 

  18. Sabelström H, Stenudd M, Frisén J. Neural stem cells in the adult spinal cord. Exp Neurol. 2014;260:44–9.

    PubMed  Google Scholar 

  19. Marichal N, Reali C, Trujillo-Cenóz O, Russo RE. Spinal cord stem cells in their microenvironment: the ependyma as a stem cell niche. Adv Exp Med Biol. 2017;1041:55–79.

    CAS  PubMed  Google Scholar 

  20. Santos SIP, de Oliveira VC, Pieri NCG, Bressan FF, Ambrósio CE, Feitosa MLT. Isolation and characterization of neural stem cells from fetal canine spinal cord. Neurosci Lett. 2021;765:136293.

    CAS  PubMed  Google Scholar 

  21. Dromard C, Guillon H, Rigau V, et al. Adult human spinal cord harbors neural precursor cells that generate neurons and glial cells in vitro. J Neurosci Res. 2008;86:1916–26.

    CAS  PubMed  Google Scholar 

  22. Konomi T, Suda K, Ozaki M, Harmon SM, Komatsu M, Iimoto S, Tsuji O, Minami A, Takahata M, Iwasaki N. Predictive factors for irreversible motor paralysis following cervical spinal cord injury. Spinal Cord. 2021;59:554–62.

    PubMed  Google Scholar 

  23. Courtine G, Song B, Roy RR, Zhong H, Herrmann JE, Ao Y, et al. Recovery of supraspinal control of stepping via indirect propriospinal relay connections after spinal cord injury. Nat Med. 2008;14:69–74.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Flynn JR, Graham BA, Galea MP, Callister RJ. The role of propriospinal interneurons in recovery from spinal cord injury. Neuropharmacology. 2011;60:809–22.

    CAS  PubMed  Google Scholar 

  25. Fan B, Wei Z, Yao X, Shi G, Cheng X, Zhou X, Zhou H, Ning G, Kong X, Feng. Microenvironment imbalance of spinal cord injury. Cell Transpl. 2018;27:853–66.

    Google Scholar 

  26. Zhao Y, Xiao Z, Chen B, Dai J. The neuronal differentiation microenvironment is essential for spinal cord injury repair. Organogenesis. 2017;13:63–70.

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Yokota K, Kobayakawa K, Kubota K, Miyawaki A, Okano H, Ohkawa Y, Iwamoto Y, Okada S. Engrafted neural stem/progenitor cells promote functional recovery through synapse reorganization with spared host neurons after spinal cord injury. Stem cell Reports. 2015;5:264–77.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Medalha CC, Jin Y, Yamagami T, Haas C, Fischer I. Transplanting neural progenitors into a complete transection model of spinal cord injury. J Neurosci Res. 2014;92:607–18.

    CAS  PubMed  Google Scholar 

  29. Adler AF, Lee-Kubli C, Kumamaru H, Kadoya K, Tuszynski MH. Comprehensive monosynaptic rabies virus mapping of host connectivity with neural progenitor grafts after spinal cord injury. Stem cell rep. 2017;8:1525–33.

    Google Scholar 

  30. Ramos-Zúñiga R, González-Pérez O, Macías-Ornelas A, Capilla-González V, Quiñones-Hinojosa A. Ethical implications in the use of embryonic and adult neural stem cells. Stem Cells Int. 2012;2012:470949.

    PubMed  PubMed Central  Google Scholar 

  31. Mazzini L, Gelati M, Profico DC, Sgaravizzi G, Projetti Pensi M, Muzi G, et al. Human neural stem cell transplantation in ALS: initial results from a phase I trial. J Transl Med. 2015;13:17.

    PubMed  PubMed Central  Google Scholar 

  32. Jakel RJ, Schneider BL, Svendsen CN. Using human neural stem cells to model neurological disease. Nat Rev Genet. 2004;5:136–44.

    CAS  PubMed  Google Scholar 

  33. Mothe A, Tator CH. Isolation of neural stem/progenitor cells from the periventricular region of the adult rat and human spinal cord. J Vis Exp Jove. 2015;(99):e52732.

  34. Hugnot JP. Isolate and culture neural stem cells from the mouse adult spinal cord. In: Methods Mol Biol Springer. 2013;1059:53–63.

    Google Scholar 

  35. Rietze RL, Reynolds BA. Neural stem cell isolation and characterization. Meth Enzymol. 2006;419:3–23.

    CAS  Google Scholar 

  36. Yang S, Cao Z, Zhu J, Zhang Z, Zhao H, Zhao L, et al. In vitro monolayer culture of dispersed neural stem cells on the e-cadherin-based substrate with long-term stemness maintenance. ACS Omega. 2019;4:18136–46.

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Bez A, Corsini E, Curti D, Biggiogera M, Colombo A, Nicosia RF, et al. Neurosphere and neurosphere-forming cells: morphological and ultrastructural characterization. Brain Res. 2003;993:18–29.

    CAS  PubMed  Google Scholar 

  38. Janebodin K, Horst OV, Ieronimakis N, Balasundaram G, Reesukumal K, Pratumvinit B, et al. Isolation and characterization of neural crest-derived stem cells from dental pulp of neonatal mice. PLoS ONE. 2011;6:e27526.

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Sütterlin P, Williams EJ, Chambers D, Saraf K, von Schack D, Reisenberg M, et al. The molecular basis of the cooperation between EGF, FGF and eCB receptors in the regulation of neural stem cell function. Mol Cell Neurosci. 2013;52:20–30.

    PubMed  Google Scholar 

  40. Conti L, Cattaneo E. Neural stem cell systems: physiological players or in vitro entities? Nat Rev Neurosci. 2010;11:176–87.

    CAS  PubMed  Google Scholar 

  41. Ao Q, Wang AJ, Chen GQ, Wang SJ, Zuo HC, Zhang XF. Combined transplantation of neural stem cells and olfactory ensheathing cells for the repair of spinal cord injuries. Med Hypotheses. 2007;69:1234–7.

    CAS  PubMed  Google Scholar 

  42. Azari H, Rahman M, Sharififar S, Reynolds BA. Isolation and expansion of the adult mouse neural stem cells using the neurosphere assay. J Vis Exp Jove. 2010:e2393.

  43. Vescovi AL, Reynolds BA, Fraser DD, Weiss S. bFGF regulates the proliferative fate of unipotent (neuronal) and bipotent (neuronal/astroglial) EGF-generated CNS progenitor cells. Neuron. 1993;11:951–66.

    CAS  PubMed  Google Scholar 

  44. Xiong Z, Zhao S, Yan X. Nerve stem cell differentiation by a one-step cold atmospheric plasma treatment in vitro. J Vis Exp Jove. 2019;(143):e58663.

  45. Drew KL, McGee RC, Wells MS, Kelleher-Andersson JA. Growth and differentiation of adult hippocampal arctic ground squirrel neural stem cells. J Vis Exp Jove. 2011;(47):e2199.

  46. Mao X, Zhao S. Neuronal differentiation from mouse embryonic stem cells in vitro. J Vis Exp Jove. 2020;(160):e61190.

  47. Sun T, Wang XJ, Xie SS, Zhang DL, Wang XP, Li BQ, et al. A comparison of proliferative capacity and passaging potential between neural stem and progenitor cells in adherent and neurosphere cultures. Int J Dev Neurosci. 2011;29:723–31.

    PubMed  Google Scholar 

  48. Wang S, Guan S, Li W, Ge D, Xu J, Sun C, et al. 3D culture of neural stem cells within conductive PEDOT layer-assembled chitosan/gelatin scaffolds for neural tissue engineering. Mater Sci Eng C Mater Biol Appl. 2018;93:890–901.

    CAS  PubMed  Google Scholar 

  49. Cui CP, Zhang Y, Wang C, Yuan F, Li H, Yao Y, et al. Dynamic ubiquitylation of Sox2 regulates proteostasis and governs neural progenitor cell differentiation. Nat Commun. 2018;9:4648.

    PubMed  PubMed Central  Google Scholar 

  50. Graham V, Khudyakov J, Ellis P, Pevny L. SOX2 functions to maintain neural progenitor identity. Neuron. 2003;39:749–65.

    CAS  PubMed  Google Scholar 

  51. Perriot S, Canales M, Mathias A, Du Pasquier R. Differentiation of functional astrocytes from human-induced pluripotent stem cells in chemically defined media. Star Protoc. 2021;2:100902.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Liu Q, Lü L, Sun H, Zhang J, Ma W, Zhang T. Effect of serum on the differentiation of neural stem cells. Chin J Reparative Reconstr Surg. 2018;32:223–7.

    Google Scholar 

  53. Bischoff DS, Makhijani NS, Yamaguchi DT. Constitutive expression of human telomerase enhances the proliferation potential of human mesenchymal stem cells. BioResearch Open Access. 2012;1:273–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) method. Methods. 2001;25:402–8.

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all the members who contributed to this article.

Funding

This work was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (No. 2018R1D1A1B07047451).

Author information

Authors and Affiliations

Authors

Contributions

WKK was involved in study design and performed experiments, data analysis, and manuscript writing. YSS, JHL, WHK and BJK were involved in study design and revised the manuscript. All authors read and approved the manuscript.

Corresponding author

Correspondence to Byung-Jae Kang.

Ethics declarations

Ethics approval and consent to participate

The study is reported in accordance with ARRIVE guidelines. This study was conducted according to the Animal Care and Use Guidelines (The Institute of Laboratory Animal Resources, Seoul National University). The study was approved by the Animal Care and Use Committee of Seoul National University (SNU-210609-1-1).

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, W.K., Son, Y.S., Lim, JH. et al. Neural stem/progenitor cells from adult canine cervical spinal cord have the potential to differentiate into neural lineage cells. BMC Vet Res 19, 193 (2023). https://0-doi-org.brum.beds.ac.uk/10.1186/s12917-023-03757-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12917-023-03757-3

Keywords