Skip to main content

Expression profiling of stemness markers in testicular germline stem cells from neonatal and adult Swiss albino mice during their transdifferentiation in vitro

Abstract

Background

Spermatogonial stem cells (SSCs) were considered to be stem cells with limited potencies due to their existence in adult organisms. However, the production of spermatogonial stem cell colonies with broader differentiation capabilities in primary germ cell cultures from mice of select genetic backgrounds (C57BL6/Tg14, ddY, FVB and 129/Ola) indicated that SSCs from these strains were pluripotent.

Methods

We established primary cultures of SSCs from neonatal and adult Swiss 3T3 Albino mice. Stemness of SSC colonies were evaluated by performing real-time PCR and immunofluorescence analysis for a panel of chosen stemness markers. Differentiation potentials of SSCs were examined by attempting the generation of embryoid bodies and evaluating the expression of ectodermal, mesodermal and endodermal markers using immunofluorescence and real-time PCR analysis.

Results

Spermatogonial stem cells from neonatal and mature mice testes colonised in vitro and formed compact spermatogonial stem cell colonies in culture. The presence of stem cell markers ALPL, ITGA6 and CD9 indicated stemness in these colonies. The differentiation potential of these SSC colonies was demonstrated by their transformation into embryoid bodies upon withdrawal of growth factors from the culture medium. SSC colonies and embryoid bodies formed were evaluated using immunofluorescence and real-time PCR analysis. Embryoid body like structures derived from both neonatal and adult mouse testis were quite similar in terms of the expression of germ layer markers.

Conclusion

These results strongly suggest that SSC-derived EB-like structures could be used for further differentiation into cells of interest in cell-based therapeutics.

Introduction

Embryonic Stem Cells (ESCs) are pluripotent cells having the potential to differentiate into any type of cell in the body and can undergo differentiation in vitro by means of genetic modification or adding exogenous factors into the culture medium but may form teratomas in the accepting host [1]. Though ESCs offer wider potential for therapeutic applications in regenerative medicine, its use is hindered by the ethical concerns around the destruction of embryos for this purpose [2]. On the other hand, the use of adult stems cells in cell-based therapeutics is limited by their narrow range of potency, despite their potential for autologous stem cell donation, which may help to avoid issues of immune rejection [3]. Though pluripotency can be induced in somatic cells [4], both ESCs and iPSCs express several abnormalities during reprogramming and in prolonged culture. These include acquired abnormal karyotype [5,6,7], copy number variations (CNV) leading to mutations [8, 9], elimination of residual pluripotent cells [10] etc. iPSC induced teratoma was more aggressive than those induced by ES cells [11] and the molecular signature of iPSCs can be influenced by the cell type of origin [12]. iPSCs pose side effects related to transplantation in the accepting host, like formation of tumors due to the residual mass of pluripotent cells and hence their applications are limited currently. Recent research has established an optimized tool that will allow specific and selective removal of iPSCs by using LVCAGs iPSCs (transgenic) [1, 13].

Mammalian testis has a small population of Germ line stem cells (GSCs), which are descendants of primordial germ cells (PGCs) and have the ability to both self-renew and generate daughter cells that begin spermatogenesis [14]. Specification of PGCs consists of three main steps: repressing somatic programming, regaining the potential of pluripotency and epigenetic reprogramming in the entire genome. Prdm14 is a transcriptional regulator which influences pluripotency and epigenetic reprogramming and is specifically expressed in the PSCs and germlines [15, 16]. GSCs constitute about 0.03% of the germ cells in the testis [17]. The genomic integrity of these cells appears to be higher because evolutionary selective forces act only on mutations of the germline genome and not on those in somatic cells [18]. Moreover, GSCs do not invoke the ethical concerns associated with the use of embryos for deriving embryonic stem cells (ESCs) for research [19]. Considering the advantages of GSCs over somatic stem cells, attempts have been made to derive pluripotent cells from GSCs. In-vitro proliferation of spermatogonial stem cells from testis of a new born transgenic mouse line C57BL6/Tg14 bred into DBA/2 background in growth factor supplemented media was reported in 2003 [20]. Two types of colonies showing typical characteristics of GS cell colonies and ES cell colonies were developed from neonatal ddY and DBA mice testis cultures [21]. SSCs from adult mice (C57BL/6, FVB and 129/Ola) testis responded to culture conditions and acquired embryonic stem cell properties and produced multipotent adult germline stem cells (maGSCs). Those maGSCs could generate three germ layers in vitro, produced teratomas in immunodeficient mice and could participate in development when injected into blastocysts [22]. Subsequently, putative SSCs and/or their progenitors were shown to reprogram to pluripotency when removed from their stem cell niche and when appropriate growth factors and reagents in embryonic stem cell medium are added [23]. In mice, PGC like cells (PGCLCs) were derived from ESCs and iPSCs by means of in-vitro culture through aggregates of Epiblast-like cells or EPiLc [16, 24]. Recent reports reveal the derivation of induced PGC-like cells (iPGCLCs) from mouse iPSCs, which could re-establish spermatogenesis following transplantation into the testis of infertile W/Wv mice [25]. As the derivation of SSC colonies depended on the genetic background [20], the differentiation capabilities of GSCs have been examined in selected mouse strains only. In this study, we have attempted to grow germline stem cells from testis of neonatal and adult Swiss albino mice and to evaluate their differentiation capability in vitro.

Materials and methods

Animals

Healthy male mice (Mus musculus, Swiss strain) bred in the institute animal facility, housed at temperature 27 ± 1° C and humidity-controlled conditions under 14 h light: 10 h dark and provided with food and water ad libitum, were used for the study. Neonatal (Postnatal day 12, P12), adolescent (postnatal day 30, P30) and adult (postnatal day 90, P90) male mice were used in this study. Use of laboratory animals for experiments was duly approved by the Institutional Animal Ethics Committee of Rajiv Gandhi Centre for Biotechnology, Thiruvananthapuram, vide Approval No. IAEC/66/PRK/2008.

Reagents

Primary antibodies (GFRA1, ITGA6, ALPL, CD9, NES, TBXT and PDX-1), goat anti-rabbit IgG-FITC, goat anti-rabbit IgG- Alexa Fluor 488, Rabbit anti-goat IgG- Alexa Fluor 488, (Santa Cruz Biotechnology, CA, USA); CD9 antibody (BioLegend, San Diego, CA, USA); Reverse Transcriptase PCR primers (Sigma Genosys, Bangalore, India); Taq DNA polymerase, 100 bp & 1 kb ladder (New England Biolabs, MA, USA); DAPI, TRI reagent, Ethidium Bromide, agarose, glycine, Trizma Base and paraformaldehyde were purchased from Sigma- Aldrich, MO, USA.

Preparation of single cell suspension from mouse testis

Neonatal and adult male mice were sacrificed by cervical dislocation and the testes were dissected out. Testis was then washed twice in sterile PBS and the tunica albuginea removed. A sequential enzymatic digestion was used to obtain single cell suspension of mice testis [26, 27]. The seminiferous tubules were digested with 1 mg/mL collagenase I and 5 µg/ mL DNase in Dulbecco’s Modified Eagle’s Medium (DMEM)/ F-12 supplemented with 1% Minimum Essential Medium, Non-Essential Amino Acids (MEM-NEAA) and 1% Antibiotic- Antimycotic solution for 20 min at 37° C. The tubes were kept at room temperature for 5 min to allow the seminiferous tubules to settle. The supernatant containing interstitial and peritubular cells was removed. The tubular fragments were subjected to a second digestion step with 1 mg/ mL collagenase I and 5 µg/ mL DNase and dispase solution (1mL/ 100 mg testis tissue) for 30 min at 32° C. The digestion mixture was gently agitated every 5 min during digestion to aid dissociation of the tubules. The digestion was stopped by adding DMEM /F12 FBS, mixed thoroughly and centrifuged at 270 × g for 5 min at room temperature. The supernatant was discarded and the pellet resuspended in DMEM/ F12 FBS. Filtration through a 40 μm cell strainer (BD Falcon, NJ, USA) removed the debri from the solution. The cells were observed under a microscope to ensure the absence of cell clumps and that the cells were intact.

Primary cell culture of testicular cells

Primary cultures of germ cells were established from neonatal and adult mice testicular cells [23, 26, 28]. The single cell suspension of testicular cells prepared as mentioned above was resuspended in SF medium supplemented with FBS (SF) [21, 22]. The SF medium composed of StemPro34 SFM base (Invitrogen, CA, USA) supplemented with StemPro34 nutrient supplement, non-growth factor components [5 mg/ mL Bovine serum albumin (Calbiochem, USA); 6 mg/ mL d-(+) Glucose, 10 µg/ mL d-Biotin, 25 µg/ mL Insulin, 30 µg/ mL Pyruvic acid sodium salt, 0.06% dl-Lactic acid (60% solution), 100 µM Ascorbic acid, 30 µM Sodium selenite, 60 µM Putrescine, 100 µg/ mL Bovine Apo-transferrin, 60 ng/ mL Progesterone, 30 ng/ mL β-Estradiol 17-cypionate, 10 µM 2-Mercaptoethanol (Sigma- Aldrich, MO, USA); 2 mM L-Glutamine, 1X Antibiotic–Antimycotic, 1X MEM vitamins, 1X Nonessential amino acids (Invitrogen, CA, USA); 1% Fetal bovine serum (Hyclone, Logan, USA)] and growth factor components [103 U/ mL LIF, 10 ng/ mL Recombinant human basic FGF (Sigma Aldrich, MO, USA), 20 ng/ mL Recombinant mouse EGF and 15 ng/ mL Recombinant rat GDNF (R&D Systems, Minneapolis, MN)]. The resuspended cells were plated on a gelatin coated 12 well plate at a cell density of 2 × 105 cells/mL and incubated for 16–24 h at 37° C in a humidified incubator with 5% CO2. The somatic cells of the testes attached to the gelatin plate and the germ cells did not. The floating germ cells were harvested using a P 1000 pipette and were transferred into a new 12 well plate after centrifugation at 270 × g for 5 min at RT. The germ cells were replenished with fresh SF medium every 3 days and the cultures were examined daily under a stereomicroscope for the appearance of SSC colonies The SSC colonies which started appearing by day 5 in culture were harvested using a micropipette and were placed in new 12 well plates containing fresh SF medium. Two such colonies were processed for immunofluorescence analysis of each of the stemness markers which was screened. The remaining colonies were transferred to new 12 well plates containing fresh SF medium [21] deprived of growth factors LIF and GDNF to induce embryoid body (EB) formation as reported earlier [29,30,31]. These cultures were maintained weeks in same medium for 2–3 [21] (Fig. 1).

Fig. 1
figure 1

Graphical representation of the experiment design for generating ES-like colonies and embryoid like body development from spermatogonial stem cells harvested from the mouse testes (Created with BioRender.com)

Immunofluorescence analysis of SSC colonies and embryoid bodies

SSC colonies and embryoid bodies were grown on poly-L-Lysine coated coverslips in a 12 well plate in appropriate medium at 37° C in a humidified incubator with 5% CO2 for 24 h. They were fixed with 4% paraformaldehyde in PBS, permeabilized in 0.1% Triton X-100 for 10 min and blocked for 30 min in 1% BSA. The SSC colonies were probed with antibodies against GFRA1, ITGA6, ALPL and CD9 (at a dilution of 1:200) for overnight at 4° C. Similarly, the embryoid body like structures were probed with antibodies against NES, TBXT (BRACHYURY) and PDX-1 (at a dilution of 1:200). They were further probed with fluorochrome conjugated goat anti rabbit (GFRA1, ITGA6, CD9) or rabbit anti-goat secondary (ALPL, NES, TBXT and PDX-1) antibodies (at a dilution of 1: 500) for 1 h. The dispersed cells from SSC colonies were counterstained with DAPI for 5 min at a final concentration of 0.50 µg/ mL to visualize the nuclei. The coverslips were washed in PBS, dried and mounted on glass slides using 60% glycerol. The cells were imaged on a Leica TCS-SP2 Confocal Microscope equipped with AOBS system (Leica, Mannheim, Germany).

RNA preparation

Total RNA was prepared from SSC colonies/embryoid bodies derived from primary cultures of neonatal (one week old), adult (three months old) mice, using TRI reagent according to the manufacturer’s instructions. The SSC colonies/embryoid bodies were homogenized in 1 ml of TRI reagent (5 pulses at 1626 × g for 30 s each with 30 s intervals between pulses) using a PT-100 homogenizer probe (Kinematica AG, Luzernerstrasse, Lucerne). 200 µL of chloroform was added to the homogenate, incubated for 15 min and centrifuged at 12,000 × g for 15 min at 4° C. The upper transparent layer was transferred into a new eppendorf tube and RNA was precipitated using 0.5 mL isopropanol. The RNA pellet was washed in 70% ethyl alcohol, air dried and suspended in 35 µL sterile DEPC water.

RT-PCR analysis

Single cell suspensions from the testes of P12, P30 and P90 mice were prepared as mentioned earlier. The cells were grown in DMEM/F12 medium supplemented with 10% FBS, 1% antibiotic–antimycotic and 1% non-essential amino acids (NEAA) in gelatin coated plates for 24 h. The floating germ cells were harvested on the following day using a P 1000 pipette, centrifuged at 270 × g for 5 min at RT and were transferred into a new low attachment 12 well plate. Further the cells were grown in DMEM/F12 medium supplemented with 10% FBS, 1% antibiotic–antimycotic and 1% NEAA without any growth factor. It was maintained and observed under microscope for 5–6 days. RNA was isolated from the SSC colonies derived from the testes from P12, P30 and P90 mice using Trizol reagent as explained above. 1 µg of RNA was converted to cDNA using Verso cDNA synthesis kit (Thermo scientific). Postnatal day 8 (P8) mouse testes cDNA was used as a positive control for this experiment. RT-PCR was performed for pluripotency factors such as c-Myc, Klf4, Lin28, Nanog, Oct4 and Sox2 [4, 32,33,34]. The PCR products were resolved on 1.5% agarose gel. The expression level of beta-actin was used as an internal control. The primer pairs used for this experiment are listed in Table 1.

Table 1 Sequences of primers used in the real time and RT-PCR experiment

Real-time PCR analysis

Real-time PCR analyses were done for floating germ cell population (FGCP), spermatogonial stem cell clusters (SSCs) and SSCs derived embryoid bodies (EB) from the immature and adult testis. 1 µg RNA was reverse transcribed using SuperScript™ VILO™ cDNA synthesis kit (Invitrogen, Carlsbad, CA, USA) following manufacturer’s instructions. Primers used in the real time PCR experiment are listed in Table 1. PCR for pluripotency and germ cell markers and differentiation markers was done using a gene specific forward and reverse primer of each in a total of 5 µl reaction with SYBR green master mix (Applied biosystems, WA, UK) in ABI 7900 HT Sequence Detection System (Applied Biosystems, Netherlands) under standard qPCR temperature conditions. The following cycling parameters were used: 50º C for 10 min, 90ο C for 10 min and 95ο C for 10 min. This was followed by 40 cycles at 95ο C for 10 s and a combined annealing/extension temperature of 60º C for 2 min. Expression level of beta actin was used as internal control. Two biological replicates, each with three technical replicates and with appropriate nontemplate controls (NTCs) were analysed. Fold changes were calculated using the comparative delta delta Ct (ΔΔCt) method for relative quantitation. Statistical analysis of these genes between the floating germ cell population versus SSCs, and SSCs of mature and immature testes versus SSCs derived embryoid bodies was done using two-tailed Student’s t-test on averaged Δ∆Ct values of these genes.

Results

Primary culture of mouse testicular cells

Primary culture was established from germ cells from neonatal (one week old) and adult (three months old) mouse testes. Cell suspension prepared from the seminiferous tubules of neonatal mouse was plated on gelatin coated plates on the first day. The floating germ cells were replated on day 2 (Fig. 2, A). The testicular somatic cells remained attached to the plate and proliferated rapidly forming a monolayer, while the floating germ cells formed clusters and later transformed into small colonies by day 5 (Fig. 2, C). Similarly, cell suspension from the seminiferous tubules of adult mouse (Fig. 2, B) also generated small clusters on day 4 which transformed into compact SSC colonies by day 5 (Fig. 2, D) respectively. The colonies formed in the cultures of SSCs derived from neonatal testes were larger in size when compared to its counterparts from the adult testes. These SSC colonies were manually transferred onto a new 12 well plate and replenished with fresh SF media on every 7th day and were successfully maintained up to two months.

Fig. 2
figure 2

Formation of SSC colonies in primary cultures of neonatal (one week old) and adult (3 month old) mice testicular cells. Cell suspension prepared from the seminiferous tubules of neonatal (A) and adult (B) mice were followed up for 4 weeks. SSC colonies were formed on day 5 in cultures of germ cells from neonatal (C) and adult (D). Bar = 20 μm

Evaluation for stemness in SSC colonies from neonatal testis

The stemness of SSC colonies grown in primary culture for 24 days was analyzed by evaluating the expression of a subset of known pluripotency markers using immunofluorescence microscopy. Immunolocalization studies revealed that SSC colonies were positive for GFRA1 (Fig. 3A), ITGA6 (Fig. 3B), ALPL (Fig. 3C) and CD9 (Fig. 3D). The corresponding phase contrast images are shown in Fig. 3, E-H).

Fig. 3
figure 3

Confocal microscopy of SSC colonies derived from neonatal mice testis in day 24 of culture, showing the expression of GFRA1 (A), ITGA6 (B) and ALPL (C) and CD9 (D). The corresponding phase contrast images (E-H) of the SSC colonies are also shown. Bar = 20 μm. Real time PCR analysis of the expression of stem cell markers and a germ cell marker in SSC colonies on day 24 of culture (I). Bar = 20 μm

The pluripotency/ stemness and germline status of the SSC colonies generated were assessed by evaluating the expression levels of markers such as Alpl, Gfra1, Itga6, Ifitm3, Nanog, Pou5f1 and Ddx4 in the floating germ cells and the SSC colonies. The stemness markers Alpl, Itga6 and Ifitm3 were upregulated in SSCs colonies, while the key germline marker Ddx4 and pluripotency markers such as Nanog and Pou5f1 were downregulated significantly (Fig. 3, I). However, the mRNA levels of the SSC marker Gfra1 was significantly downregulated, which was in contrast with the observation made from our immunocytochemical studies.

The expression levels of key pluripotency markers such as c-Myc, Klf4, Lin28, Nanog, Oct4 and Sox2 were analysed in SSC-derived ES-like colonies derived from P12, P30 and P90 mouse testes grown in DMEMF12 medium without any added growth factors. The colonies derived from the SSCs from P12 and P30 testes showed abundance of c-Myc, Klf4, Lin28, while Nanog, Oct4, and Sox2 were sparsely expressed. However, ES like colonies derived from SSCs of adult testes showed a reduction in the expression level of c-Myc, Klf4, Lin28 when compared with the other two groups, whereas Nanog, Oct4, and Sox2 were not detected (Fig. 4).

Fig. 4
figure 4

Expression of c-Myc, Klf4, Nanog, Lin28, Oct4, and Sox-2 in ES-like cells derived from neonatal (P12), adolescent (P30) and adult (P90) mouse testicular cells after 5–6 days culture in growth factors free DMEM-F12 medium. Postnatal day 8 mouse whole testes (T) used as a positive control. Actb was used as an internal control. L-100 bp DNA ladder

SSCs differentiated into embryoid bodies

Spermatogonial stem cell colonies produced from neonatal mice were maintained in medium containing growth factors LIF and GDNF. The maintenance of these SSC colonies in medium devoid of these growth factors induced differentiation in them. Transformation of SSC colonies (Fig. 5A) into embryoid bodies was observed within 5–7 days after removal of LIF and GDNF in SSCs derived from the neonatal mouse testes (Fig. 5B). The SSC derived embryoid bodies were evaluated for the expression of all the three germ layer markers using immunolocalisation studies. The embryoid bodies expressed NESTIN (Fig. 6A), BRACHYURY (Fig. 6B) and PDX-1 (Fig. 6C) indicating the presence of ectoderm, mesoderm and endoderm layers in them. A negative control, which was probed only with the secondary antibody, is also shown (Fig. 6D). The corresponding DIC images are shown in Fig. 6, E-H. Both the SSC clusters and the embryoid bodies were analyzed by real time PCR for the expression of germ layer markers. The relative quantities of the express of the three germ layer markers (Nes for ectoderm, Des for mesoderm and Afp for endoderm) are shown in Fig. 6, I. We observed 33-fold increase in the expression of Nes, 10-fold increase of Des and 13-fold increase of Afp in the EB-like colonies when compared to SSC colonies (Fig. 6, I).

Fig. 5
figure 5

Embryoid body (EB) formation on withdrawal of growth factors. Spermatogonial Stem Cells (SSCs) colonies from neonatal testis (A), were cultured for 10 days after removal of growth factors LIF and GDNF (B). Bar = 20 μm

Fig. 6
figure 6

Expression of NESTIN (A and E), BRACHYURY (B and F), and PDX-1 (C and G) embryoid bodies (EBs) derived from neonatal mice testicular cells 10 days after removal of LIF and GDNF, (D and H) are secondary antibody control. Bar = 20 μm. Real-time PCR analysis of germ layer markers Nes, Des, Afp in SSC-derived EBs 10 days after removal of LIF and GDNF (I). Actb was used for internal normalization and the expression levels of these genes in EBs were further normalized with their corresponding expression levels in SSCs

SSC colonies and EB-like structures from SSCs from adult testis

SSCs isolated from adult mouse testis produced relatively loose grape-like clusters. Further the dispersed SSCs colony cells stained positive for GFRA1, ITGA6, ALPL and CD9 (Fig. 7, A-D). These cells counter-stained with DAPI are presented in Fig. 7, E-H. The withdrawal of LIF and GDNF from the culture medium induced the formation of compact bodies, which stained positive for NES, BRACHYURY and PDX1 (Fig. 8, A-C). A negative control was also run with the exclusion of primary antibody from the incubation (Fig. 8, D). The corresponding DIC images are shown in Fig. 8, E-H.

Fig. 7
figure 7

Expression of GFRA1 (A), ITGA6 (B), ALPL (C) and CD9 (D) in dispersed cells of SSCs derived from adult mice testicular cells. The corresponding DAPI-stained images are shown in E-H. Bar = 20 μm

Fig. 8
figure 8

Expression of NESTIN (A and E), TBXT (BRACHYURY) (B and F), and PDX-1 (C and G) embryoid bodies (EBs) derived from adult mice testicular cells 10 days after removal of LIF and GDNF, (D and H) are secondary antibody control. Bar = 20 μm. Real time PCR analysis of germ layer markers Nes, Des, Afp in SSC-derived EBs 10 days after removal of LIF and GDNF (I). Actb was used for internal normalization and the expression levels of these genes in EBs were further normalized with their corresponding expression levels in SSCs

Realtime PCR analysis was performed to evaluate the expression levels of the three primary germ layer markers Nes, Des and Afp in these EB-like structures A significant increase in the relative expression levels of all the three markers in EB-like structures, when compared with that from SSC colonies was observed. Thus, the fold changes in the expression levels of Nes, Des and Afp in EB-like structures were 4, 86 and 30 respectively (Fig. 8, I).

Discussion

The long-held notion that spermatogonial stem cells (SSCs) of the testes are unipotent was unravelled by the production of ES-like cells from neonatal mouse testis, which formed teratomas and could also participate in normal embryonic development [21]. Studies on the developmental fate of a single spermatogonial stem cell revealed that conversion of SSC into a pluripotent cell type is accompanied by loss of spermatogenic potential [2]. In line with this observation, this study also demonstrates a significant reduction in the expression of a key germ cell marker Ddx4 in the SSC-derived ES-like colonies derived from immature testes when cultured without any feeder cells. This result indicates a possible eraser of germ cell imprints in SSCs due to loss of their niche. Hickford et al. had reported that Ddx4 was crucial for germ cell proliferation and differentiation and its expression was highly regulated by epigenetic modification specifically dimethylation of arginine motifs and CpG islands at its promoter region [35].In addition, studies had reported the upregulation of Ddx4 expression in germ cells once they enter the gonads, whereas it’s expression was not detected in pluripotent ES, EG or inner cell mass (ICM) cells [36]. However, germ cell colonies could be initiated from ICR or C57BL/6 × DBA/2 F1 (BDF1) mouse strains but not from C57BL/6 or 129/Sv genetic backgrounds. Hence, the derivation of germ cell colonies in vitro depended upon the niche and the genetic background of the mouse strain under study [20]. Later, Guan et al. [22] showed that SSCs from adult mice (C57BL/6, FVB and 129/Ola) testis responded to culture conditions and acquired embryonic stem cell like properties and produced multipotent adult germline stem cells (maGSCs) which could differentiate into three germ layers in vitro. Further, it produced teratomas in immunodeficient mice and could participate in development when injected into blastocysts [22]. Our study demonstrates for the first time the successful generation of ES-like cells from the neonatal and adult testes SSCs from a common mouse strain like Swiss Albino. Also, these SSC-derived ES-like cells could produce EBs comprising of the three primordial germ layers, thereby establishing their pluripotent like state in vitro once again.

The formation of ES-like colonies in germ cell cultures from human testis has been reported [37, 38]. However, the time required for stem cell colony formation was lesser in the germ cell cultures of murine testis compared to those reported for human testicular germ cell cultures reported in these studies. Compact SSC colonies were formed in both neonatal and adult mice testicular cell culture maintained with growth factors (Fig. 2). These SSC colonies remained in pluripotent like state for months when grown in growth factor supplemented medium. Though ES-like clusters formed in SSC cultures from neonatal and adult mouse testes when grown in feeder free and growth factor free media, the efficiency was less and the colonies were unable to maintain their pluripotent nature for longer periods. Plethora of earlier studies had well documented the importance of growth factors in SSC maintenance [39, 40]. These SSC derived ES-like colonies expressed ALPL, GFRA1, ITGA6 and CD9 indicating their stemness [41, 42].

The expression of stemness markers such as Alpl, Itga6 and Ifitm3 in the SSC derived ES like colonies (Fig. 3, I) indicates their pluripotent status. Moreover, the ES like clusters formed in SSC cultures in growth factor or feeder free media were also positive for c-Myc, Klf4, Lin24, Oct-4, Nanog and Sox2, indicating their inherent pluripotency state. However, the expression of these markers was reduced in ES-like colonies derived from SSCs from adult testes when compared with those from immature (Fig. 4), though it did not affect their differentiation potential. Reduced expression of these pluripotency factors in SSC colonies derived from adult testes might be due to their exposure to the differentiation priming cues produced by the testicular niche and the differentiation primed epigenetic modifications. Recently, Sojoudi et al. reported that mouse SSCs exhibited heterogeneity in colonies formed during in vitro cultures, as they differed in their appearances and molecular marker expression. However, the reason behind this heterogeneity is not understood. CTCF, SMAD3 and SOX2 were important transcription factors predicted to transform the SSCs committed to spermatogenesis into pluripotent form in p53−/− mice in long-term cultures in the absence of exogenous growth factors. SMAD3 is a prerequisite for reprogramming of the SSCs into pluripotent state in long term cultures. Since it was found that p53 knockout affected SMAD3 induction in SSCs, there is a connection between p53 gene expression and pluripotency associated factors. Thus, this study provides a new insight into SSC reprogramming mechanism and tumorogenesis of GSCs [43]. The expression of pluripotency markers SSEA-4, OCT-4 and SOX 2 was detected in human spermatogonial stem cell derived ES-like cells [39].

The differentiation potentials of the stem cell colonies derived from the mouse testis were studied by withdrawing growth factors LIF and GDNF from the culture medium. Differentiation could be induced in ES cells following LIF withdrawal leading to the formation of embryoid bodies [40]. The withdrawal of LIF and GDNF from the culture medium was based on reports showing that the addition of LIF to SSC cultures is superfluous and that GDNF is required for SSC maintenance [41]. GDNF produced by Sertoli cells and endothelial cells in the testis regulates the expression of genes implicated in the maintenance of self-renewing state and/or prevention of differentiation of SSCs (Aundiff) which include Nanos2, Etv5, Lhx1, T(Brachyury), Bcl6b, Id4 and Cxcr4 [44,45,46,47,48,49] The loss or withdrawal of GDNF might downregulate the expression of self-renewal genes under its signalling or regulation. Therefore, the down regulation of self-renewal associated genes in SSCs cultured in growth factor deprived condition appears to lead to their differentiation. In testis, SSC differentiation is mediated by retinoic acid (RA) signalling and canonical WNT signalling. In the absence of such differentiation factors in the culture medium, the formation of EB like bodies from SSC colonies was rather spontaneous and their differentiation induction appears to be primarily due to the loss of in vitro niche required for spermatogenic differentiation. Further, such EB like bodies developed from SSC colonies derived from neonatal and adult mouse testes upon removal of growth factors expressed ectodermal (Nes), mesodermal (BRACHYURY (TBXT) and Des) and endodermal (Afp and PDX-1) cell linage markers (Figs. 5 and 7) suggesting natural transdifferentiation of SSC colonies considering the upregulation of Ifitm3 and downregulation of Pou5f1, Nanog and Gfra1 in cultured SSC colonies (Fig. 3I). The use of NES [42], TBXT [50], DES [43], AFP [50] and PDX-1 [44] as germ layer markers has been reported. Trans-differentiation is a form of lineage reprogramming or cellular process which involves the direct conversion of one type of differentiated cell into another distinct cell type from its original lineage without going through an intermediate pluripotent state, reflecting a high level of cellular plasticity [51, 52]. This transformation involves the direct conversion of one specialized cell type into another, often from one tissue or germ layer to another through specific molecular changes by alterations in gene expression, epigenetic modifications and signalling pathway activation that drives the cells toward the new fate [53,54,55].

During natural transdifferentiation, first the cell dedifferentiates and later, the innate developmental programme is activated allowing the cell to differentiate into a new lineage [56]. Newt lens regeneration perfectly illustrates the naturally occurring transdifferentiation process during which, pigmented epithelial cells (PECs) dedifferentiates, proliferates by re-entering into cell cycle and finally differentiates into crystallin expressing mature lens cell [57, 58]. In the present study, Ifitm3 which is expressed in migrating primordial germ cells (PGCs) and is implicated to have roles in germ cell development [59] was found to be upregulated during in vitro SSC culture, suggesting possible dedifferentiation of SSC to an intermediate primitive state and those cells which in turn transdifferentiates into all three germ lineages. Further, involvement of BMP and upregulation of Wnt signalling during transdifferentiation from cornea to lens in Xenopus laevis was reported [60]. BMPs and WNT mediated signalling are implicated in early and late germ cell development like PGC induction, proliferation, migration and gametogenesis [61, 62]. Therefore, the embryoid body like body formation from SSC colonies might also have utilized BMP and WNT mediated dedifferentiation and transdifferentiation processes. A comparative analysis of small RNA signatures from SSCs, Sertoli cells, developing germ cells, ESCs and MSCs using high throughput sequencing revealed that the miRNA signature in mouse SSCs were similar to those of ESCs [45]. Thus, we could establish ES-like clusters from germ cells isolated from the testis of neonatal and adult Swiss albino mice which expressed pluripotency genes and upon induction of differentiation formed embryoid bodies similar to those derived from ES cells.

Conclusion

GSCs colonized in primary germ cell culture from Swiss albino mice testicular cells and its stemness was confirmed by demonstrating the expression of pluripotency markers. Majority of the studies on GSCs till date employed GS cells established from transgenic mouse lines. This study reports for the first-time isolation of GSCs from Swiss albino mice testicular cells which could form colonies, maintain stemness and form EBs upon induction of differentiation. Further, both neonatal and adult mouse-derived GSCs have comparable ability to make SSC colonies and EB-like bodies and have similar gene expression. The germline stem cell derived embryoid bodies were comparable to those formed by embryonic stem cells in vitro. The authenticity and differentiation potential of the germline stem cell derived embryoid bodies was confirmed by localizing germ layer markers to these embryoid bodies. Transplantation assays and lineage specific differentiation of these testis derived stem cells are underway.

Data availability

Not applicable.

Abbreviations

ALP:

Alkaline Phosphatase

BSA:

bovine serum albumin

CNV:

copy number variation

DEPC:

Diethyl pyrocarbonate

DMEM:

Dulbecco’s Modified Eagle’s Medium

DNA:

Deoxy ribonucleic acid

EB:

Embryoid body

ES:

Embryonic Stem Cell

DAPI:

4’,6-diamidino-2-phenylindole

FBS:

fetal bovine serum;FGF:fibroblast growth factor

FGCP:

floating germ cell population

GDNF:

glial derived neurotrophic factor

GSC:

germline stem cell

iPSC:

induced Pluripotent Stem cell

LIF:

leukemia inhibiting factor

MEM:

minimum essential medium

M-MuLV:

Moloney Murine Leukemia Virus

NEAA:

non-essential amino acid

NTP:

nucleotide triphosphate

OCT:

Octamer binding protein

PBS:

Phosphate buffered saline

PCR:

polymerase chain reaction

PDX:

Pancreatic and duodenal homeobox 1

PGC:

primordial germ cell

PI:

propidium iodide

RNA:

ribonucleic acid

RT:

room temperature

SSC:

Spermatogonial Stem Cell

References

  1. Mousaei Ghasroldasht M, Seok J, Park HS, Liakath Ali FB, Al-Hendy A. Stem cell therapy: from idea to clinical practice. Int J Mol Sci. 2022;23(5).

  2. Kanatsu-Shinohara M, Lee J, Inoue K, Ogonuki N, Miki H, Toyokuni S, et al. Pluripotency of a single spermatogonial stem cell in mice. Biol Reprod. 2008;78(4):681–7.

    Article  CAS  PubMed  Google Scholar 

  3. Durnaoglu S, Genc S, Genc K. Patient-specific pluripotent stem cells in neurological diseases. Stem Cells Int. 2011;2011:212487.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006;126(4):663–76.

    Article  CAS  PubMed  Google Scholar 

  5. Draper JS, Smith K, Gokhale P, Moore HD, Maltby E, Johnson J, et al. Recurrent gain of chromosomes 17q and 12 in cultured human embryonic stem cells. Nat Biotechnol. 2004;22(1):53–4.

    Article  CAS  PubMed  Google Scholar 

  6. Baker DE, Harrison NJ, Maltby E, Smith K, Moore HD, Shaw PJ, et al. Adaptation to culture of human embryonic stem cells and oncogenesis in vivo. Nat Biotechnol. 2007;25(2):207–15.

    Article  CAS  PubMed  Google Scholar 

  7. Mayshar Y, Ben-David U, Lavon N, Biancotti JC, Yakir B, Clark AT, et al. Identification and classification of chromosomal aberrations in human induced pluripotent stem cells. Cell Stem Cell. 2010;7(4):521–31.

    Article  CAS  PubMed  Google Scholar 

  8. Laurent LC, Ulitsky I, Slavin I, Tran H, Schork A, Morey R, et al. Dynamic changes in the copy number of pluripotency and cell proliferation genes in human ESCs and iPSCs during reprogramming and time in culture. Cell Stem Cell. 2011;8(1):106–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Hussein SM, Batada NN, Vuoristo S, Ching RW, Autio R, Närvä E, et al. Copy number variation and selection during reprogramming to pluripotency. Nature. 2011;471(7336):58–62.

    Article  CAS  PubMed  Google Scholar 

  10. Miura K, Okada Y, Aoi T, Okada A, Takahashi K, Okita K, et al. Variation in the safety of induced pluripotent stem cell lines. Nat Biotechnol. 2009;27(8):743–5.

    Article  CAS  PubMed  Google Scholar 

  11. Gutierrez-Aranda I, Ramos-Mejia V, Bueno C, Munoz-Lopez M, Real PJ, Mácia A, et al. Human induced pluripotent stem cells develop teratoma more efficiently and faster than human embryonic stem cells regardless the site of injection. Stem Cells. 2010;28(9):1568–70.

    Article  PubMed  Google Scholar 

  12. Bar-Nur O, Russ HA, Efrat S, Benvenisty N. Epigenetic memory and preferential lineage-specific differentiation in induced pluripotent stem cells derived from human pancreatic islet beta cells. Cell Stem Cell. 2011;9(1):17–23.

    Article  CAS  PubMed  Google Scholar 

  13. Dahlke J, Schott JW, Vollmer Barbosa P, Klatt D, Selich A, Lachmann N et al. Efficient genetic safety switches for future application of iPSC-Derived cell transplants. J Personalized Med. 2021;11(6).

  14. Seydoux G, Braun RE. Pathway to totipotency: lessons from germ cells. Cell. 2006;127(5):891–904.

    Article  CAS  PubMed  Google Scholar 

  15. Kurimoto K, Yamaji M, Seki Y, Saitou M. Specification of the germ cell lineage in mice: a process orchestrated by the PR-domain proteins, Blimp1 and Prdm14. Cell Cycle (Georgetown Tex). 2008;7(22):3514–8.

    Article  CAS  PubMed  Google Scholar 

  16. Yao C, Yao R, Luo H, Shuai L. Germline specification from pluripotent stem cells. Stem Cell Res Ther. 2022;13(1):74.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Tegelenbosch RA, de Rooij DG. A quantitative study of spermatogonial multiplication and stem cell renewal in the C3H/101 F1 hybrid mouse. Mutat Res. 1993;290(2):193–200.

    Article  CAS  PubMed  Google Scholar 

  18. Puri MC, Nagy A. Concise review: embryonic stem cells versus induced pluripotent stem cells: the game is on. Stem Cells. 2012;30(1):10–4.

    Article  CAS  PubMed  Google Scholar 

  19. McLaren A. Ethical and social considerations of stem cell research. Nature. 2001;414(6859):129–31.

    Article  CAS  PubMed  Google Scholar 

  20. Kanatsu-Shinohara M, Ogonuki N, Inoue K, Miki H, Ogura A, Toyokuni S, et al. Long-term proliferation in culture and germline transmission of mouse male germline stem cells. Biol Reprod. 2003;69(2):612–6.

    Article  CAS  PubMed  Google Scholar 

  21. Kanatsu-Shinohara M, Inoue K, Lee J, Yoshimoto M, Ogonuki N, Miki H, et al. Generation of pluripotent stem cells from neonatal mouse testis. Cell. 2004;119(7):1001–12.

    Article  CAS  PubMed  Google Scholar 

  22. Guan K, Nayernia K, Maier LS, Wagner S, Dressel R, Lee JH, et al. Pluripotency of spermatogonial stem cells from adult mouse testis. Nature. 2006;440(7088):1199–203.

    Article  CAS  PubMed  Google Scholar 

  23. Golestaneh N, Kokkinaki M, Pant D, Jiang J, DeStefano D, Fernandez-Bueno C, et al. Pluripotent stem cells derived from adult human testes. Stem Cells Dev. 2009;18(8):1115–26.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Zvetkova I, Apedaile A, Ramsahoye B, Mermoud JE, Crompton LA, John R, et al. Global hypomethylation of the genome in XX embryonic stem cells. Nat Genet. 2005;37(11):1274–9.

    Article  CAS  PubMed  Google Scholar 

  25. Li Y, Wang X, Feng X, Liao S, Zhang D, Cui X, et al. Generation of male germ cells from mouse induced pluripotent stem cells in vitro. Stem cell Res. 2014;12(2):517–30.

    Article  CAS  PubMed  Google Scholar 

  26. Falciatori I, Lillard-Wetherell K, Wu Z, Hamra FK, Garbers DL. Deriving mouse spermatogonial stem cell lines. (Clifton NJ). 2008;450:181–92. Methods in molecular biology.

    CAS  Google Scholar 

  27. Gassei K, Ehmcke J, Schlatt S. Efficient enrichment of undifferentiated GFR alpha 1 + spermatogonia from immature rat testis by magnetic activated cell sorting. Cell Tissue Res. 2009;337(1):177–83.

    Article  CAS  PubMed  Google Scholar 

  28. Kanatsu-Shinohara M, Shinohara T. Germline modification using mouse spermatogonial stem cells. Methods Enzymol. 2010;477:17–36.

    Article  CAS  PubMed  Google Scholar 

  29. Kurosawa H. Methods for inducing embryoid body formation: in vitro differentiation system of embryonic stem cells. J Biosci Bioeng. 2007;103(5):389–98.

    Article  CAS  PubMed  Google Scholar 

  30. Thomas C, Doetschman HE. Margot Katz, Werner Schmidt and Rolf Kemler. The in vitro development of blastocyst-derived embryonic stem cell lines: formation of visceral yolk sac, blood islands and myocardium. J Experimental Morphology. 1985;87:27–45.

    Google Scholar 

  31. Desbaillets I, Ziegler U, Groscurth P, Gassmann M. Embryoid bodies: an in Vitro Model of Mouse Embryogenesis. Exp Physiol. 2004;85(6):645–51.

    Article  Google Scholar 

  32. Liu X, Huang J, Chen T, Wang Y, Xin S, Li J, et al. Yamanaka factors critically regulate the developmental signaling network in mouse embryonic stem cells. Cell Res. 2008;18(12):1177–89.

    Article  CAS  PubMed  Google Scholar 

  33. Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, Tian S, et al. Induced pluripotent stem cell lines derived from human somatic cells. Science. 2007;318(5858):1917–20.

    Article  CAS  PubMed  Google Scholar 

  34. Thomson M, Liu SJ, Zou LN, Smith Z, Meissner A, Ramanathan S. Pluripotency factors in embryonic stem cells regulate differentiation into germ layers. Cell. 2011;145(6):875–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Hickford DE, Frankenberg S, Pask AJ, Shaw G, Renfree MB. DDX4 (VASA) is conserved in germ cell development in marsupials and monotremes. Biol Reprod. 2011;85(4):733–43.

    Article  CAS  PubMed  Google Scholar 

  36. Bowles J, Koopman P. Sex determination in mammalian germ cells: extrinsic versus intrinsic factors. Reproduction. 2010;139(6):943–58.

    Article  CAS  PubMed  Google Scholar 

  37. Mizrak SC, Chikhovskaya JV, Sadri-Ardekani H, van Daalen S, Korver CM, Hovingh SE, et al. Embryonic stem cell-like cells derived from adult human testis. Hum Reprod (Oxford England). 2010;25(1):158–67.

    Article  CAS  Google Scholar 

  38. Sadri-Ardekani H, Mizrak SC, van Daalen SK, Korver CM, Roepers-Gajadien HL, Koruji M, et al. Propagation of human spermatogonial stem cells in vitro. JAMA. 2009;302(19):2127–34.

    Article  CAS  PubMed  Google Scholar 

  39. Jeong D, McLean DJ, Griswold MD. Long-term culture and transplantation of murine testicular germ cells. J Androl. 2003;24(5):661–9.

    Article  PubMed  Google Scholar 

  40. Wang P, Suo LJ, Wang YF, Shang H, Li GX, Hu JH, et al. Effects of GDNF and LIF on mouse spermatogonial stem cells proliferation in vitro. Cytotechnology. 2014;66(2):309–16.

    Article  CAS  PubMed  Google Scholar 

  41. Kanatsu-Shinohara M, Toyokuni S, Shinohara T. CD9 is a surface marker on mouse and rat male germline stem cells. Biol Reprod. 2004;70(1):70–5.

    Article  CAS  PubMed  Google Scholar 

  42. Oka M, Tagoku K, Russell TL, Nakano Y, Hamazaki T, Meyer EM, et al. CD9 is associated with leukemia inhibitory factor-mediated maintenance of embryonic stem cells. Mol Biol Cell. 2002;13(4):1274–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Liu S, Wei R, Liu H, Liu R, Li P, Zhang X, et al. Analysis of chromatin accessibility in p53 deficient spermatogonial stem cells for high frequency transformation into pluripotent state. Cell Prolif. 2022;55(3):e13195.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Chen C, Ouyang W, Grigura V, Zhou Q, Carnes K, Lim H, et al. ERM is required for transcriptional control of the spermatogonial stem cell niche. Nature. 2005;436(7053):1030–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Oatley JM, Avarbock MR, Telaranta AI, Fearon DT, Brinster RL. Identifying genes important for spermatogonial stem cell self-renewal and survival. Proc Natl Acad Sci USA. 2006;103(25):9524–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Oatley JM, Avarbock MR, Brinster RL. Glial cell line-derived neurotrophic factor regulation of genes essential for self-renewal of mouse spermatogonial stem cells is dependent on src family kinase signaling. J Biol Chem. 2007;282(35):25842–51.

    Article  CAS  PubMed  Google Scholar 

  47. Oatley MJ, Kaucher AV, Racicot KE, Oatley JM. Inhibitor of DNA binding 4 is expressed selectively by single spermatogonia in the male germline and regulates the self-renewal of spermatogonial stem cells in mice. Biol Reprod. 2011;85(2):347–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Sada A, Hasegawa K, Pin PH, Saga Y. NANOS2 acts downstream of glial cell line-derived neurotrophic factor signaling to suppress differentiation of spermatogonial stem cells. Stem Cells. 2012;30(2):280–91.

    Article  CAS  PubMed  Google Scholar 

  49. Yang QE, Kim D, Kaucher A, Oatley MJ, Oatley JM. CXCL12-CXCR4 signaling is required for the maintenance of mouse spermatogonial stem cells. J Cell Sci. 2013;126(Pt 4):1009–20.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Sojoudi K, Azizi H, Skutella T. A Fundamental Research in In Vitro Spermatogonial Stem Cell Culturing: what are clump cells? Cell Reprogram. 2023;25(2):65–72.

    Article  CAS  PubMed  Google Scholar 

  51. Clayton RM. The molecular basis for competence, determination and transdifferentiation: a hypothesis. Adv Exp Med Biol. 1982;158:23–38.

    Article  CAS  PubMed  Google Scholar 

  52. Selman K, Kafatos FC. Transdifferentiation in the labial gland of silk moths: is DNA required for cellular metamorphosis? Cell Differ. 1974;3(2):81–94.

    Article  CAS  PubMed  Google Scholar 

  53. Li WC, Yu WY, Quinlan JM, Burke ZD, Tosh D. The molecular basis of transdifferentiation. J Cell Mol Med. 2005;9(3):569–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Eguizabal C, Montserrat N, Veiga A, Izpisua Belmonte JC. Dedifferentiation, transdifferentiation, and reprogramming: future directions in regenerative medicine. Semin Reprod Med. 2013;31(1):82–94.

    Article  PubMed  Google Scholar 

  55. Cieślar-Pobuda A, Knoflach V, Ringh MV, Stark J, Likus W, Siemianowicz K, et al. Transdifferentiation and reprogramming: overview of the processes, their similarities and differences. Biochim et Biophys acta Mol cell Res. 2017;1864(7):1359–69.

    Article  Google Scholar 

  56. Jopling C, Boue S, Izpisua Belmonte JC. Dedifferentiation, transdifferentiation and reprogramming: three routes to regeneration. Nat Rev Mol Cell Biol. 2011;12(2):79–89.

    Article  CAS  PubMed  Google Scholar 

  57. Tsonis PA, Madhavan M, Tancous EE, Del Rio-Tsonis K. A newt’s eye view of lens regeneration. Int J Dev Biol. 2004;48(8–9):975–80.

    Article  PubMed  Google Scholar 

  58. Maki N, Martinson J, Nishimura O, Tarui H, Meller J, Tsonis PA, et al. Expression profiles during dedifferentiation in newt lens regeneration revealed by expressed sequence tags. Mol Vis. 2010;16:72–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Tanaka SS, Yamaguchi YL, Tsoi B, Lickert H, Tam PP. IFITM/Mil/fragilis family proteins IFITM1 and IFITM3 play distinct roles in mouse primordial germ cell homing and repulsion. Dev Cell. 2005;9(6):745–56.

    Article  CAS  PubMed  Google Scholar 

  60. Day RC, Beck CW. Transdifferentiation from cornea to lens in Xenopus laevis depends on BMP signalling and involves upregulation of wnt signalling. BMC Dev Biol. 2011;11:54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Lochab AK, Extavour CG. Bone morphogenetic protein (BMP) signaling in animal reproductive system development and function. Dev Biol. 2017;427(2):258–69.

    Article  CAS  PubMed  Google Scholar 

  62. Cantú AV, Laird DJ. Primordial germ cell migration and the wnt signaling pathway. Anim Reprod. 2017;14(1):89–101.

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by Grants BT/PR11207/MED/31/58/2008 from Department of Biotechnology and 37 (1618)/ 13/ EMR –II from CSIR to Kumar PG. Indu S was supported by Grant No: (I .no. 10 − 2(5)0.2005(ii)-E.U.II, May10th 2006) from University Grants Commission, New Delhi, India. Jiji V, Bindu Asokan and Gopikrishnan K at the Confocal Imaging Core Facility provided technical assistance in imaging.

Funding

This work was supported by Grants BT/PR11207/MED/31/58/2008 from Department of Biotechnology and 37 (1618)/ 13/ EMR –II from CSIR to Kumar PG. Indu S was supported by Grant No: (I .no. 10 − 2 [5]0.2005(ii)-E.U.II, May10th 2006) from University Grants Commission, New Delhi, India. The funding body played no role in the design of the study and collection, analysis, and interpretation of data and in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

PGK conceived the theme, designed the experiments, analyzed and interpreted data and edited the manuscript; SI, designed the experiments, performed the experiments, analyzed and interpreted the data and drafted the manuscript; AND, JS and MS performed the experiments, acquired the data and drafted the manuscript; SRK and AB contributed in drafting and finalizing this manuscript. All authors approved this manuscript.

Corresponding author

Correspondence to Pradeep G. Kumar.

Ethics declarations

Author information

PGK was a senior scientist heading the Molecular Reproduction Lab-1 at Rajiv Gandhi Centre for Biotechnology until his superannuation on 31 July 2022 and is currently an Emeritus Scientist of Indian Council of Medical Research, New Delhi at Department of Biotechnology, University of Kerala, Thiruvananthapuram. He is a Life member of professional societies including Indian Society for the Study of Reproduction and Fertility (ISSRF), Society of Biological Chemists India (SBCI), Indian Society of Cell Biology (ISCB) and Society for Reproductive Biology and Comparative Endocrinology (SRBCE). SI is a post-doctoral fellow; AND was a PhD student; MS is a Senior Research Fellow; JS and SRK are doctoral students and AB was a short-term research intern.

Ethics approval and consent to participate

Title of the approved project - Identification and functional evaluation of factors regulating testicular germ line stem cell division and differentiation Name of the institutional approval committee - Institutional Animal Ethics Committee (IAEC) of Rajiv Gandhi Centre for Biotechnology, Thiruvananthapuram Approval number - IAEC/66/PRK/2008 Date of Approval – 12 June 2008 Consent to participate – Not applicable as there is no human subject involved in this study.

Consent for publication

All the authors have provided their consents for the publication of this article.

Competing interests

The authors declare no competing intetests.

Additional information

Publisher’s Note

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

Electronic supplementary material

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

Indu, S., Devi, A.N., Sahadevan, M. et al. Expression profiling of stemness markers in testicular germline stem cells from neonatal and adult Swiss albino mice during their transdifferentiation in vitro. Stem Cell Res Ther 15, 93 (2024). https://doi.org/10.1186/s13287-024-03701-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13287-024-03701-8

Keywords