Journal articles on the topic 'Organotypic spinal cord slices'

To see the other types of publications on this topic, follow the link: Organotypic spinal cord slices.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Organotypic spinal cord slices.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Biancotti, Juan C., Kendal A. Walker, Guihua Jiang, Julie Di Bernardo, Lonnie D. Shea, and Shaun M. Kunisaki. "Hydrogel and neural progenitor cell delivery supports organotypic fetal spinal cord development in an ex vivo model of prenatal spina bifida repair." Journal of Tissue Engineering 11 (January 2020): 204173142094383. http://dx.doi.org/10.1177/2041731420943833.

Full text
Abstract:
Studying how the fetal spinal cord regenerates in an ex vivo model of spina bifida repair may provide insights into the development of new tissue engineering treatment strategies to better optimize neurologic function in affected patients. Here, we developed hydrogel surgical patches designed for prenatal repair of myelomeningocele defects and demonstrated viability of both human and rat neural progenitor donor cells within this three-dimensional scaffold microenvironment. We then established an organotypic slice culture model using transverse lumbar spinal cord slices harvested from retinoic acid–exposed fetal rats to study the effect of fibrin hydrogel patches ex vivo. Based on histology, immunohistochemistry, gene expression, and enzyme-linked immunoabsorbent assays, these experiments demonstrate the biocompatibility of fibrin hydrogel patches on the fetal spinal cord and suggest this organotypic slice culture system as a useful platform for evaluating mechanisms of damage and repair in children with neural tube defects.
APA, Harvard, Vancouver, ISO, and other styles
2

Sypecka, Joanna, Sylwia Koniusz, Maria Kawalec, and Anna Sarnowska. "The Organotypic Longitudinal Spinal Cord Slice Culture for Stem Cell Study." Stem Cells International 2015 (2015): 1–10. http://dx.doi.org/10.1155/2015/471216.

Full text
Abstract:
The objective of this paper is to describe in detail the method of organotypic longitudinal spinal cord slice culture and the scientific basis for its potential utility. The technique is based on the interface method, which was described previously and thereafter was modified in our laboratory. The most important advantage of the presented model is the preservation of the intrinsic spinal cord fiber tract and the ventrodorsal polarity of the spinal cord. All the processes occurring during axonal growth, regeneration, synapse formation, and myelination could be visualized while being culturedin vitrofor up to 4-5 weeks after the slices had been isolated. Both pups and adult animals can undergo the same, equally efficient procedures when going by the protocol in question. The urgent need for an appropriatein vitromodel for spinal cord regeneration results from a greater number of clinical trials concerning regenerative medicine in the spinal cord injury and from still insufficient knowledge of the molecular mechanisms involved in the neuroreparative processes. The detailed method of organotypic longitudinal spinal cord slice culture is accompanied by examples of its application to studying biological processes to which both the CNS inhabiting and grafted cells are subjected.
APA, Harvard, Vancouver, ISO, and other styles
3

Haque, Azizul, Donald C. Shields, Arabinda Das, Abhay Varma, Russel J. Reiter, and Narendra L. Banik. "Melatonin receptor-mediated attenuation of excitotoxic cell death in cultured spinal cord slices." Melatonin Research 4, no. 2 (April 30, 2021): 336–47. http://dx.doi.org/10.32794/mr11250098.

Full text
Abstract:
Recent studies suggest ex vivo modeling of neuronal injury is a robust approach for the mechanistic study of neurodegeneration. Melatonin, an indolamine, is a versatile molecule with antioxidative, antiapoptotic, neuroprotective, and anti-inflammatory properties. While melatonin has been studied as a therapeutic agent for spinal cord injury (SCI) related neuronal cell loss, its actions in organotypic slice cultures approximating SCI effects are less well understood. The actions of melatonin were therefore examined following exposure of cultured rat spinal cord slices to glutamate excitotoxicity. Exposure to glutamate (500 μM) for 4 hours induced neuronal degeneration that was prevented by 0.5 μM melatonin (applied immediately or 4 hours following glutamate exposure). Decreased internucleosomal DNA fragmentation, Bax:Bcl-2 and calpain:calpastatin ratios, caspase 8, 9 and 3 activities in slice cultures were measured following melatonin treatment. Melatonin receptor (MTR1, MTR2) mRNA levels were increased in the melatonin treated spinal cord slices. To confirm melatonin receptor-mediated protection, slice cultures were treated with 10 or 25 μM luzindole (melatonin receptor antagonist) at 0 and 4 hours, respectively, after glutamate exposure. Luzindole significantly decreased the ability of melatonin to prevent cell death in the sliced culture model. These results suggest melatonin receptors may provide a pathway for therapeutic applications to prevent penumbral neuron loss following SCI.
APA, Harvard, Vancouver, ISO, and other styles
4

Shahar, A., S. Lustig, Y. Akov, Y. David, P. Schneider, and R. Levin. "Different pathogenicity of encephalitic togaviruses in organotypic cultures of spinal cord slices." Journal of Neuroscience Research 25, no. 3 (March 1990): 345–52. http://dx.doi.org/10.1002/jnr.490250311.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Ucar, Buket, Sedef Yusufogullari, and Christian Humpel. "Collagen hydrogels loaded with fibroblast growth factor-2 as a bridge to repair brain vessels in organotypic brain slices." Experimental Brain Research 238, no. 11 (August 29, 2020): 2521–29. http://dx.doi.org/10.1007/s00221-020-05907-7.

Full text
Abstract:
Abstract Vessel damage is a general pathological process in many neurodegenerative disorders, as well as spinal cord injury, stroke, or trauma. Biomaterials can present novel tools to repair and regenerate damaged vessels. The aim of the present study is to test collagen hydrogels loaded with different angiogenic factors to study vessel repair in organotypic brain slice cultures. In the experimental set up I, we made a cut on the organotypic brain slice and tested re-growth of laminin + vessels. In the experimental set up II, we cultured two half brain slices with a gap with a collagen hydrogel placed in between to study endothelial cell migration. In the experimental set up I, we showed that the number of vessels crossing the cut was tendencially increased with the addition of fibroblast growth factor-2 (FGF-2), vascular endothelial growth factor, or platelet-derived growth factor-BB compared to the control group. In the experimental set up II, we demonstrated that a collagen hydrogel loaded with FGF-2 resulted in a significantly increased number of migrated laminin + cells in the gap between the slices compared to the control hydrogel. Co-administration of several growth factors did not further potentiate the effects. Taken together, we show that organotypic brain slices are good models to study brain vessels and FGF-2 is a potent angiogenic factor for endothelial cell proliferation and migration. Our results provide evidence that the collagen hydrogels can be used as an extracellular matrix for the vascular endothelial cells.
APA, Harvard, Vancouver, ISO, and other styles
6

Liu, Jing-Jie, Xiao-Yan Ding, Li Xiang, Feng Zhao, and Sheng-Li Huang. "A novel method for oxygen glucose deprivation model in organotypic spinal cord slices." Brain Research Bulletin 135 (October 2017): 163–69. http://dx.doi.org/10.1016/j.brainresbull.2017.10.010.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Rybachuk, O. A., Yu A. Lazarenko, V. V. Krotov, and N. V. Voitenko. "Structural/Functional Characteristics of Organotypic Spinal Cord Slices under Conditions of Long-Lasting Culturing." Neurophysiology 49, no. 2 (April 2017): 162–64. http://dx.doi.org/10.1007/s11062-017-9647-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Phelps, P. E., R. P. Barber, and J. E. Vaughn. "Nonradial migration of interneurons can be experimentally altered in spinal cord slice cultures." Development 122, no. 7 (July 1, 1996): 2013–22. http://dx.doi.org/10.1242/dev.122.7.2013.

Full text
Abstract:
During development, many migrating neurons are thought to guide on radially oriented glia to reach their adult locations. However, members of the ‘U-shaped’ group of cholinergic interneurons in embryonic rat spinal cord appeared to migrate in a direction perpendicular to the orientation of radial glia. This ‘U-shaped’ group of cells was located around the ventral ventricular zone on embryonic day 16 and, during the next two days, the constituent cells dispersed into the dorsal horn or around the central canal. During this period, these cells could be identified with either ChAT immunocytochemistry or NADPH-diaphorase histochemistry and they appeared to be aligned along commissural axons, suggesting that such processes, rather than radial glia, might guide their migration. An organotypic spinal cord slice preparation was developed and utilized for three different experimental approaches to studying this migration. In the first experiments, slices of embryonic day 16 cervical spinal cord were cultured for one, two or three days, and a relatively histotypic dorsal migration of ‘U-derived’ cells could be inferred from these sequential cultures. A second set of experiments focused on the direct observation of dorsally directed migration in living spinal cord cultures. Embryonic day 16 slices were injected with a lipophilic fluorescent label near the dorsal boundary of the ‘U-shaped’ cell group and the dorsal movement of labeled cells was observed using confocal microscopy. These experiments confirmed the dorsal migratory pattern inferred from sequentially fixed specimens. A third experimental approach was to transect embryonic day 16 slice cultures microsurgically in order to disturb the migration of ‘U-derived’ cells. Depending upon the amount of ventral spinal cord removed, the source of cells was excised and/or their guidance pathway was perturbed. The number and position of ‘U-derived’ cells varied with the amount of ventral cord excised. If more than 400 microns was removed, no ‘U-derived’ diaphorase-labeled cells were present, whereas if only 200–300 microns was removed, the cultures contained such cells. However, in this instance, many of the ‘U-derived’ neurons did not move as far dorsally, nor did they display their characteristic dorsoventral orientation. When results from these three experiments are taken together, they provide strong evidence that nonradial neuronal migration occurs in developing spinal cord and that the ‘U-derived’ neurons utilize such a migration to move from their ventral generation sites to their dorsal adult locations.
APA, Harvard, Vancouver, ISO, and other styles
9

Ravikumar, Madhumitha, Seema Jain, Robert H. Miller, Jeffrey R. Capadona, and Stephen M. Selkirk. "An organotypic spinal cord slice culture model to quantify neurodegeneration." Journal of Neuroscience Methods 211, no. 2 (November 2012): 280–88. http://dx.doi.org/10.1016/j.jneumeth.2012.09.004.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Patar, Azim, Peter Dockery, Siobhan McMahon, and Linda Howard. "Ex Vivo Rat Transected Spinal Cord Slices as a Model to Assess Lentiviral Vector Delivery of Neurotrophin-3 and Short Hairpin RNA against NG2." Biology 9, no. 3 (March 15, 2020): 54. http://dx.doi.org/10.3390/biology9030054.

Full text
Abstract:
The failure of the spinal cord to regenerate can be attributed both to a lack of trophic support for regenerating axons and to upregulation of inhibitory factors such as chondroitin sulphate proteoglycans including NG2 following injury. Lentiviral vector-mediated gene therapy is a possible strategy for treating spinal cord injury (SCI). This study investigated the effect of lentiviral vectors expressing Neurotrophin-3 (NT-3) and short-hairpin RNA against NG2 (NG2 sh) to enhance neurite outgrowth in in vitro and ex vivo transection injury models. Conditioned medium from cells transduced with NT-3 or shNG2 lentiviruses caused a significant increase in neurite length of primary dorsal root ganglia neurons compared to the control group in vitro. In an ex vivo organotypic slice culture (OSC) transduction with Lenti-NT-3 promoted axonal growth. Transducing OSCs with a combination of Lenti-NT-3/NG2 sh lead to a further increase in axonal growth but only in injured slices and only within the region adjacent to the site of injury. These findings suggest that the combination of lentiviral NT-3 and NG2 sh reduced NG2 levels and provided a more favourable microenvironment for neuronal regeneration after SCI. This study also shows that OSCs may be a useful platform for studying glial scarring and potential SCI treatments.
APA, Harvard, Vancouver, ISO, and other styles
11

LIU, JINGJIE, XIAOYAN DING, LI XIANG, and SHENGLI HUANG. "Transplanted choroidal plexus epithelial cells can integrate with organotypic spinal cord slices into a new system." BIOCELL 46, no. 6 (2022): 1537–44. http://dx.doi.org/10.32604/biocell.2022.018441.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Czarnecki, Antonny, Vincent Magloire, and Jürg Streit. "Local oscillations of spiking activity in organotypic spinal cord slice cultures." European Journal of Neuroscience 27, no. 8 (April 2008): 2076–88. http://dx.doi.org/10.1111/j.1460-9568.2008.06171.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Liu, Jing-Jie, Ya-Juan Huang, Li Xiang, Feng Zhao, and Sheng-Li Huang. "A novel method of organotypic spinal cord slice culture in rats." NeuroReport 28, no. 16 (November 2017): 1097–102. http://dx.doi.org/10.1097/wnr.0000000000000892.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Kim, Hyuk Min, Hong Jun Lee, Man Young Lee, Seung U. Kim, and Byung Gon Kim. "Organotypic Spinal Cord Slice Culture to Study Neural Stem/Progenitor Cell Microenvironment in the Injured Spinal Cord." Experimental Neurobiology 19, no. 2 (September 30, 2010): 106–13. http://dx.doi.org/10.5607/en.2010.19.2.106.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Li, Bin, Xiao-Yun Liu, Zhe Li, Hui Bu, Meng-Meng Sun, Yan-Su Guo, and Chun-Yan Li. "Effect of ALS IgG on Motor Neurons in Organotypic Spinal Cord Cultures." Canadian Journal of Neurological Sciences / Journal Canadien des Sciences Neurologiques 35, no. 2 (May 2008): 220–25. http://dx.doi.org/10.1017/s0317167100008672.

Full text
Abstract:
Objective:Reports about the role of autoimmunity in amyotrophic lateral sclerosis (ALS) are inconsistent. The aim of this work was to investigate the effect of IgG from patients with ALS on motor neurons in a physiological-like surrounding.Methods:Using affinity chromatography, IgG from six ALS patients, four disease controls and five healthy subjects was purified. Organotypic spinal cord cultures, which conserve the structure of the spinal cord in a horizontal plane and are suitable for studies with long-term treatment, were used and IgG with different concentrations ranging from 0.05 mg/mL to 0.5 mg/mL was added to the culture medium. Ventral motor neuron survival was evaluated by morphology and SMI-32 immunohistochemistry staining. Lactate dehydrogenase (LDH) level in the culture medium was measured by colorimetry.Results:After cultures were treated with ALS IgG for three weeks, the number and morphology of motor neurons showed little change. In addition, there was no significant difference in lactate dehydrogenase release between cultures treated with medium alone, normal control IgG, disease control IgG or ALS IgG.Conclusions:The results indicate that IgG from these ALS patients was insufficient per se to induce motor neuron death in Organotypic slice cultures. However, this does not preclude the possibility that other changes may have occurred in the motor neurons. This work offered a new model to evaluate the role of IgG in the pathogenesis of ALS. Organotypic cultures contribute to study of the impact of IgG on motor neurons by mimicking physiological conditions.
APA, Harvard, Vancouver, ISO, and other styles
16

Pandamooz, Sareh, Mohammad Saied Salehi, Mohammad Nabiuni, and Leila Dargahi. "Valproic acid preserves motoneurons following contusion in organotypic spinal cord slice culture." Journal of Spinal Cord Medicine 40, no. 1 (August 31, 2016): 100–106. http://dx.doi.org/10.1080/10790268.2016.1213518.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

An, Sung Su, William A. Pennant, Yoon Ha, Jin Soo Oh, Hyo Jin Kim, So-Jung Gwak, Do Heum Yoon, and Keung Nyun Kim. "Hypoxia-induced expression of VEGF in the organotypic spinal cord slice culture." NeuroReport 22, no. 2 (January 2011): 55–60. http://dx.doi.org/10.1097/wnr.0b013e3283418b00.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Lee, Yu-Shang, Janie Baratta, Jen Yu, Vernon W. Lin, and Richard T. Robertson. "aFGF Promotes Axonal Growth in Rat Spinal Cord Organotypic Slice Co-Cultures." Journal of Neurotrauma 19, no. 3 (March 2002): 357–67. http://dx.doi.org/10.1089/089771502753594927.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Streit, J. "Regular oscillations of synaptic activity in spinal networks in vitro." Journal of Neurophysiology 70, no. 3 (September 1, 1993): 871–78. http://dx.doi.org/10.1152/jn.1993.70.3.871.

Full text
Abstract:
1. Spontaneous synaptic potentials were recorded in motoneurons grown in organotypic slice cultures of embryonic rat spinal cord. In 71 of 85 cells these potentials appeared without obvious temporal structure (random patterns); in the remaining 14 cells they appeared in bursts (rhythmic patterns). 2. Random activity patterns could be converted into rhythmic patterns by treating the cultures with strychnine, bicuculline, or both. The excitatory amino acid N-methyl-D-aspartate (NMDA) transiently increased the rate of spontaneous synaptic activity without inducing rhythmic patterns. The NMDA antagonist 7-chloro-kynurenate reduced the burst rate while leaving the burst length unchanged in rhythmic patterns. In random patterns it reduced the rate of spontaneous synaptic activity by 68%. 3. Histograms of interevent times of the random patterns were best fitted by the sum of two expontentials, suggesting that the random type of activity could not be described simply as a Poisson process but involved at least one additional mechanism. 4. Rhythmic patterns consisted of bursts of activity with a mean burst length of 2.2 s that were separated by interburst intervals with a mean length of 6.6 s. Within the bursts autocorrelograms revealed regular oscillations with a mean period of 226 ms in 6 of 11 experiments with rhythmic patterns. The period showed little variation between individual experiments (202-288 ms). In random patterns no oscillations were detected. 5. Within the spontaneous bursts the excitatory postsynaptic potentials (EPSPs) progressively declined in amplitude. A corresponding depression of EPSPs was observed when trains of electrical stimuli were applied at 5 Hz to the dorsal horns of the spinal cord slices.(ABSTRACT TRUNCATED AT 250 WORDS)
APA, Harvard, Vancouver, ISO, and other styles
20

Pandamooz, Sareh, Mohammad Saied Salehi, Mohammad Ismail Zibaii, Anahid Safari, Mohammad Nabiuni, Abolhassan Ahmadiani, and Leila Dargahi. "Modeling traumatic injury in organotypic spinal cord slice culture obtained from adult rat." Tissue and Cell 56 (February 2019): 90–97. http://dx.doi.org/10.1016/j.tice.2019.01.002.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Mazzone, Graciela L., and Andrea Nistri. "Electrochemical detection of endogenous glutamate release from rat spinal cord organotypic slices as a real-time method to monitor excitotoxicity." Journal of Neuroscience Methods 197, no. 1 (April 2011): 128–32. http://dx.doi.org/10.1016/j.jneumeth.2011.01.033.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Mazzone, Graciela L., and Andrea Nistri. "Effect of the PARP-1 Inhibitor PJ 34 on Excitotoxic Damage Evoked by Kainate on Rat Spinal Cord Organotypic Slices." Cellular and Molecular Neurobiology 31, no. 3 (December 29, 2010): 469–78. http://dx.doi.org/10.1007/s10571-010-9640-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Magloire, Vincent, and Jürg Streit. "Intrinsic activity and positive feedback in motor circuits in organotypic spinal cord slice cultures." European Journal of Neuroscience 30, no. 8 (October 2009): 1487–97. http://dx.doi.org/10.1111/j.1460-9568.2009.06978.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Calderó, J., N. Brunet, O. Tarabal, L. Piedrafita, M. Hereu, V. Ayala, and J. E. Esquerda. "Lithium prevents excitotoxic cell death of motoneurons in organotypic slice cultures of spinal cord." Neuroscience 165, no. 4 (February 2010): 1353–69. http://dx.doi.org/10.1016/j.neuroscience.2009.11.034.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Park, Hwan-Woo, Hyo-Jin Jeon, and Mi-Sook Chang. "Vascular endothelial growth factor enhances axonal outgrowth in organotypic spinal cord slices via vascular endothelial growth factor receptor 1 and 2." Tissue Engineering and Regenerative Medicine 13, no. 5 (October 2016): 601–9. http://dx.doi.org/10.1007/s13770-016-0051-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Elkhenany, Hoda, Pablo Bonilla, Esther Giraldo, Ana Alastrue Agudo, Michael J. Edel, María Jesus Vicent, Fernando Gisbert Roca, et al. "A Hyaluronic Acid Demilune Scaffold and Polypyrrole-Coated Fibers Carrying Embedded Human Neural Precursor Cells and Curcumin for Surface Capping of Spinal Cord Injuries." Biomedicines 9, no. 12 (December 16, 2021): 1928. http://dx.doi.org/10.3390/biomedicines9121928.

Full text
Abstract:
Tissue engineering, including cell transplantation and the application of biomaterials and bioactive molecules, represents a promising approach for regeneration following spinal cord injury (SCI). We designed a combinatorial tissue-engineered approach for the minimally invasive treatment of SCI—a hyaluronic acid (HA)-based scaffold containing polypyrrole-coated fibers (PPY) combined with the RAD16-I self-assembling peptide hydrogel (Corning® PuraMatrix™ peptide hydrogel (PM)), human induced neural progenitor cells (iNPCs), and a nanoconjugated form of curcumin (CURC). In vitro cultures demonstrated that PM preserves iNPC viability and the addition of CURC reduces apoptosis and enhances the outgrowth of Nestin-positive neurites from iNPCs, compared to non-embedded iNPCs. The treatment of spinal cord organotypic cultures also demonstrated that CURC enhances cell migration and prompts a neuron-like morphology of embedded iNPCs implanted over the tissue slices. Following sub-acute SCI by traumatic contusion in rats, the implantation of PM-embedded iNPCs and CURC with PPY fibers supported a significant increase in neuro-preservation (as measured by greater βIII-tubulin staining of neuronal fibers) and decrease in the injured area (as measured by the lack of GFAP staining). This combination therapy also restricted platelet-derived growth factor expression, indicating a reduction in fibrotic pericyte invasion. Overall, these findings support PM-embedded iNPCs with CURC placed within an HA demilune scaffold containing PPY fibers as a minimally invasive combination-based alternative to cell transplantation alone.
APA, Harvard, Vancouver, ISO, and other styles
27

Jeong, Dong-Kee, Cyrus E. Taghavi, Kyung-Jin Song, Kwang-Bok Lee, and Hyun-Wook Kang. "Organotypic Human Spinal Cord Slice Culture as an Alternative to Direct Transplantation of Human Bone Marrow Precursor Cells for Treating Spinal Cord Injury." World Neurosurgery 75, no. 3-4 (March 2011): 533–39. http://dx.doi.org/10.1016/j.wneu.2010.10.042.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Gerardo-Nava, Jose, Dorothee Hodde, Istvan Katona, Ahmet Bozkurt, Torsten Grehl, Harry W. M. Steinbusch, Joachim Weis, and Gary A. Brook. "Spinal cord organotypic slice cultures for the study of regenerating motor axon interactions with 3D scaffolds." Biomaterials 35, no. 14 (May 2014): 4288–96. http://dx.doi.org/10.1016/j.biomaterials.2014.02.007.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Min, Hyunjung, Li Xu, Roberta Parrott, Joanne Kurtzberg, and Anthony Filiano. "Abstract 4 Umbilical Cord-Derived Mesenchymal Stromal Cells Suppress Neuroinflammation and Promote Remyelination in the Spinal Cord." Stem Cells Translational Medicine 11, Supplement_1 (September 1, 2022): S6. http://dx.doi.org/10.1093/stcltm/szac057.004.

Full text
Abstract:
Abstract Introduction Mesenchymal stromal cells (MSCs) are one of the most widely tested cell therapies due to their ability to suppress inflammation. Demyelinating diseases of the central nervous system (CNS) are commonly accompanied by neuroinflammation, including infiltrating proinflammatory leukocytes and activation of resident macrophages (microglia and border-associated macrophages in the meninges). The precise mechanism of how MSCs suppress neuroinflammation is unclear; however, we recently described a novel pathway where MSCs physically interact with macrophages to reprogram the immune response through the transfer of a cytoplasmic organelle called a processing body (p-body). Objective We set out to study how human umbilical cord tissue-derived MSCs interact with CNS microglia and border macrophages to protect in a mouse model of spinal cord demyelination. Methods We manufactured several MSC lines under GMP conditions and tested their ability to interact with CNS macrophages and promote remyelination in multiple preclinical assays. To determine if p-bodies are critical for these steps, we depleted p-bodies using CRISPR/Cas9 targeting to DDX6. To test for effects on acute demyelination, we injected lysophosphatidylcholine (LPC) into the spinal cord or added LPC to organotypic cerebellar slice cultures. To test direct effects on microglia activation, we treated primary microglia or a microglia cell line with lipopolysaccharide. Results We determined that MSC promoted remyelination in the spinal cord and that this was dependent on p-bodies. MSC directly contacted CNS macrophages and inhibited their activation. Using primary microglia and a microglia cell line, we determined that MSCs suppress the release of proinflammation cytokines, IL-6, and TNF. Discussion We determined that MSCs suppress microglial cell activation and enhance remyelination of the spinal cord. Our study offers insight into a novel mechanism about how MSCs can alter the immune response via transferring p-bodies and enhance remyelination in CNS. Furthermore, our results may lead to a quick and simple potency assay to establish the efficacy of MSCs for suppressing neuroinflammation.
APA, Harvard, Vancouver, ISO, and other styles
30

Shichinohe, Hideo, Satoshi Kuroda, Sachiko Tsuji, Satoshi Yamaguchi, Shunsuke Yano, Jang-Bo Lee, Hiroyuki Kobayashi, Seiji Kikuchi, Kazutoshi Hida, and Yoshinobu Iwasaki. "Bone Marrow Stromal Cells Promote Neurite Extension in Organotypic Spinal Cord Slice: Significance for Cell Transplantation Therapy." Neurorehabilitation and Neural Repair 22, no. 5 (May 16, 2008): 447–57. http://dx.doi.org/10.1177/1545968308315596.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Amadio, Susanna, Chiara Parisi, Cinzia Montilli, Alberto Savio Carrubba, Savina Apolloni, and Cinzia Volonté. "P2Y12Receptor on the Verge of a Neuroinflammatory Breakdown." Mediators of Inflammation 2014 (2014): 1–15. http://dx.doi.org/10.1155/2014/975849.

Full text
Abstract:
In the CNS, neuroinflammation occurring during pathologies as amyotrophic lateral sclerosis (ALS) and multiple sclerosis (MS) is the consequence of an intricate interplay orchestrated by various cell phenotypes. Among the molecular cues having a role in this process, extracellular nucleotides are responsible for intercellular communication and propagation of inflammatory stimuli. This occurs by binding to several receptor subtypes, defined P2X/P2Y, which are widespread in different tissues and simultaneously localized on multiple cells. For instance, the metabotropic P2Y12subtype is found in the CNS on microglia, affecting activation and chemotaxis, on oligodendrocytes, possessing a hypothesized role in myelination, and on astrocytes. By comparative analysis, we have established here that P2Y12receptor immunolabelled by antibodies against C-terminus or second intracellular loop, is, respectively, distributed and modulated under neuroinflammatory conditions on ramified microglia or myelinated fibers, in primary organotypic cerebellar cultures, tissue slices from rat striatum and cerebellum, spinal cord sections from symptomatic/end stage SOD1-G93A ALS mice, and finally autoptic cortical tissue from progressive MS donors. We suggest that modulation of P2Y12expression might play a dual role as analytic marker of branched/surveillant microglia and demyelinating lesions, thus potentially acquiring a predictive value under neuroinflammatory conditions as those found in ALS and MS.
APA, Harvard, Vancouver, ISO, and other styles
32

Drexler, Berthold, Julia Grenz, Christian Grasshoff, and Bernd Antkowiak. "Allopregnanolone Enhances GABAergic Inhibition in Spinal Motor Networks." International Journal of Molecular Sciences 21, no. 19 (October 7, 2020): 7399. http://dx.doi.org/10.3390/ijms21197399.

Full text
Abstract:
The neurosteroid allopregnanolone (ALLO) causes unconsciousness by allosteric modulation of γ-aminobutyric acid type A (GABAA) receptors, but its actions on the spinal motor networks are unknown. We are therefore testing the hypothesis that ALLO attenuates the action potential firing of spinal interneurons and motoneurons predominantly via enhancing tonic, but not synaptic GABAergic inhibition. We used video microscopy to assess motoneuron-evoked muscle activity in organotypic slice cultures prepared from the spinal cord and muscle tissue. Furthermore, we monitored GABAA receptor-mediated currents by performing whole-cell voltage-clamp recordings. We found that ALLO (100 nM) reduced the action potential firing of spinal interneurons by 27% and that of α-motoneurons by 33%. The inhibitory effects of the combination of propofol (1 µM) and ALLO on motoneuron-induced muscle contractions were additive. Moreover, ALLO evoked a tonic, GABAA receptor-mediated current (amplitude: 41 pA), without increasing phasic GABAergic transmission. Since we previously showed that at a clinically relevant concentration of 1 µM propofol enhanced phasic, but not tonic GABAergic inhibition, we conclude that ALLO and propofol target distinct subpopulations of GABAA receptors. These findings provide first evidence that the combined application of ALLO and propofol may help to reduce intraoperative movements and undesired side effects that are frequently observed under total intravenous anesthesia.
APA, Harvard, Vancouver, ISO, and other styles
33

Cho, Jung-Sun, Hwan-Woo Park, Sang-Kyu Park, Sangho Roh, Soo-Kyung Kang, Ki-Suk Paik, and Mi-Sook Chang. "Transplantation of mesenchymal stem cells enhances axonal outgrowth and cell survival in an organotypic spinal cord slice culture." Neuroscience Letters 454, no. 1 (April 2009): 43–48. http://dx.doi.org/10.1016/j.neulet.2009.02.024.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Mazzone, G. L., M. Mladinic, and A. Nistri. "Excitotoxic cell death induces delayed proliferation of endogenous neuroprogenitor cells in organotypic slice cultures of the rat spinal cord." Cell Death & Disease 4, no. 10 (October 2013): e902-e902. http://dx.doi.org/10.1038/cddis.2013.431.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Biggs, James E., Paul A. Boakye, Naren Ganesan, Patrick L. Stemkowski, Aquilino Lantero, Klaus Ballanyi, and Peter A. Smith. "Analysis of the long-term actions of gabapentin and pregabalin in dorsal root ganglia and substantia gelatinosa." Journal of Neurophysiology 112, no. 10 (November 15, 2014): 2398–412. http://dx.doi.org/10.1152/jn.00168.2014.

Full text
Abstract:
The α2δ-ligands pregabalin (PGB) and gabapentin (GBP) are used to treat neuropathic pain. We used whole cell recording to study their long-term effects on substantia gelatinosa and dorsal root ganglion (DRG) neurons. Spinal cord slices were prepared from embryonic day 13 rat embryos and maintained in organotypic culture for >5 wk (neuronal age equivalent to young adult rats). Exposure of similarly aged DRG neurons (dissociated and cultured from postnatal day 19 rats) to GBP or PGB for 5–6 days attenuated high-voltage-activated calcium channel currents (HVA ICa). Strong effects were seen in medium-sized and in small isolectin B4-negative (IB4−) DRG neurons, whereas large neurons and small neurons that bound isolectin B4 (IB4+) were hardly affected. GBP (100 μM) or PGB (10 μM) were less effective than 20 μM Mn2+ in suppression of HVA ICa in small DRG neurons. By contrast, 5–6 days of exposure to these α2δ-ligands was more effective than 20 μM Mn2+ in reducing spontaneous excitatory postsynaptic currents at synapses in substantia gelatinosa. Spinal actions of gabapentinoids cannot therefore be ascribed to decreased expression of HVA Ca2+ channels in primary afferent nerve terminals. In substantia gelatinosa, 5–6 days of exposure to PGB was more effective in inhibiting excitatory synaptic drive to putative excitatory neurons than to putative inhibitory neurons. Although spontaneous inhibitory postsynaptic currents were also attenuated, the overall long-term effect of α2δ-ligands was to decrease network excitability as monitored by confocal Ca2+ imaging. We suggest that selective actions of α2δ-ligands on populations of DRG neurons may predict their selective attenuation of excitatory transmission onto excitatory vs. inhibitory neurons in substantia gelatinosa.
APA, Harvard, Vancouver, ISO, and other styles
36

Guertin, Pierre A., and Jørn Hounsgaard. "Conditional Intrinsic Voltage Oscillations in Mature Vertebrate Neurons Undergo Specific Changes in Culture." Journal of Neurophysiology 95, no. 3 (March 2006): 2024–27. http://dx.doi.org/10.1152/jn.00832.2005.

Full text
Abstract:
Although intrinsic neuronal properties in invertebrates are well known to undergo specific adaptive changes in culture, long-term adaptation of similar properties in mature vertebrate neurons remain poorly understood. To investigate this, we used an organotypic slice preparation from the spinal cord of adult turtles maintainable for several weeks in culture conditions. N-methyl-d-aspartate (NMDA)-induced-tetrodotoxin (TTX)-resistant voltage oscillations in motoneurons were ∼10 times faster in culture than in acute preparations. Oscillations in culture were abolished by NMDA receptor antagonists or by high extracellular Mg2+ concentrations. However, in contrast with results from motoneurons in the acute slice, NMDA-induced oscillations in culture did not depend on CaV1.3 channel activation as they still remained after nifedipine application. Other CaV1.3 channel-mediated properties such as metabotropic receptor-induced oscillations and plateau potentials failed to be induced in culture. This study shows that changes specifically affecting CaV1.3 channel contribution to intrinsic oscillatory property expression may occur in culture. The results contribute also to understanding further the potential for plasticity of mature vertebrate neurons.
APA, Harvard, Vancouver, ISO, and other styles
37

Gao, Po, Xiaowei Ding, Tahir Muhammad Khan, Weifang Rong, Heike Franke, and Peter Illes. "P2X7 receptor-sensitivity of astrocytes and neurons in the substantia gelatinosa of organotypic spinal cord slices of the mouse depends on the length of the culture period." Neuroscience 349 (May 2017): 195–207. http://dx.doi.org/10.1016/j.neuroscience.2017.02.030.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Tashiro, Jun, Seiji Kikuchi, Kazuyoshi Shinpo, Riichiro Kishimoto, Sachiko Tsuji, and Hidenao Sasaki. "Role of p53 in neurotoxicity induced by the endoplasmic reticulum stress agent tunicamycin in organotypic slice cultures of rat spinal cord." Journal of Neuroscience Research 85, no. 2 (February 1, 2007): 395–401. http://dx.doi.org/10.1002/jnr.21120.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Larkum, Matthew E., Thomas Launey, Alexander Dityatev, and Hans-R. Lüscher. "Integration of Excitatory Postsynaptic Potentials in Dendrites of Motoneurons of Rat Spinal Cord Slice Cultures." Journal of Neurophysiology 80, no. 2 (August 1, 1998): 924–35. http://dx.doi.org/10.1152/jn.1998.80.2.924.

Full text
Abstract:
Larkum, Matthew E., Thomas Launey, Alexander Dityatev, and Hans-R. Lüscher. Integration of excitatory postsynaptic potentials in dendrites of motoneurons of rat spinal cord slice cultures. J. Neurophysiol. 80: 924–935, 1998. We examined the attenuation and integration of spontaneous excitatory postsynaptic potentials (sEPSPs) in the dendrites of presumed motoneurons (MNs) of organotypic rat spinal cord cultures. Simultaneous whole cell recordings in current-clamp mode were made from either the soma and a dendrite or from two dendrites. Direct comparison of the two voltage recordings revealed that the membrane potentials at the two recording sites followed each other very closely except for the fast-rising phases of the EPSPs. The dendritic recording represented a low-pass filtered version of the somatic recording and vice versa. A computer-assisted method was developed to fit the sEPSPs with a generalized α-function for measuring their amplitudes and rise times (10–90%). The mean EPSP peak attenuation between the two recording electrodes was determined by a maximum likelihood analysis that extracted populations of similar amplitude ratios from the fitted events at each electrode. For each pair of recordings, the amplitude attenuation ratio for EPSP traveling from dendrite to soma was larger than that traveling from soma to dendrite. The linear relation between mean ln attenuation and distance between recording electrodes was used to map 1/ e attenuations into units of distance (μm). For EPSPs with typical time course traveling from the somatic to the dendritic recording electrode, the mean 1/ e attenuation corresponded to 714 μm; for EPSPs traveling in the opposite direction, the mean 1/ e attenuation corresponded to 263 μm. As predicted from cable analysis, fast EPSPs attenuated more in both the somatofugal and somatopetal direction than did slow EPSPs. For EPSPs with rise times shorter than ∼2.0 ms, the attenuation factor increased steeply. Compartmental computer modeling of the experiments with biocytin-filled and reconstructed MNs that used passive membrane properties revealed amplitude attenuation ratios of the EPSP traveling in both the somatofugal and somatopetal direction that were comparable to those observed in real experiments. The modeling of a barrage of sEPSPs further confirmed that the somato-dendritic compartments of a MN are virtually isopotential except for the fast-rising phase of EPSPs. Large, transient differences in membrane potential are locally confined to the site of EPSP generation. Comparing the modeling results with the experiments suggests that the observed attenuation ratios are adequately explained by passive membrane properties alone.
APA, Harvard, Vancouver, ISO, and other styles
40

Ulrich, D., R. Quadroni, and H. R. Luscher. "Electronic structure of motoneurons in spinal cord slice cultures: a comparison of compartmental and equivalent cylinder models." Journal of Neurophysiology 72, no. 2 (August 1, 1994): 861–71. http://dx.doi.org/10.1152/jn.1994.72.2.861.

Full text
Abstract:
1. Voltage-clamp, current-clamp, and morphological data were obtained from visually identified motoneurons in organotypic cocultures of rat embryonic spinal cord, dorsal root ganglia, and skeletal muscle. The cells were injected with Biocytin during whole-cell patch-clamp recordings and stained with horseradish peroxidase. 2. The somata and dendritic trees of the cells were reconstructed with a semiautomatic reconstruction system. The motoneurons had a common multipolar shape. An elliptic soma gave rise to 3-9 stem dendrites with a mean diameter of 2.5 +/- 0.9 (SD) micron terminating in 24 +/- 7 dendritic endings. The mean total dendritic path length was 3,306 +/- 1,075 microns. The mean total membrane surface area was 15,594 +/- 10,404 microns 2 with a dendritic to somatic membrane surface area ratio of 3.4 +/- 1.4 (n = 7 cells). 3. The ratio between the sum of the diameters of the two daughter branches and the diameter of the parental branch each raised to the 3/2 power at all branch points was 1.3 +/- 0.28 (n = 8 cells). The dendritic trees of the cells tapered continuously from the soma to the distal ends. The mean normalized dendritic trunk parameter of all cells was 0.62 +/- 0.22. 4. The motoneurons had a mean input resistance RN of 498 +/- 374 M delta, a mean membrane time constant (tau m) of 22 +/- 4.6 ms, and a mean dendritic dominance (rho) of 2.7 +/- 0.86 (n = 5 cells). The mean electronic length (L) calculated from tau m and the slowest voltage-clamp time constant (tau VC1) was 0.7 +/- 0.04 (n = 7 cells). 5. The specific membrane capacitance (Cm) estimated from the charge of the capacitive current during a voltage step and the total membrane surface area was 1.08 +/- 0.3 microF/cm2 (n = 6 cells). 6. Compartmental computer models were constructed of individual cells. Experimental and simulated voltage transients were matched with Cm = 1 microF/cm2, a uniform membrane resistivity (Rm) = tau m/Cm and a cytosolic resistivity (Ri) of 308 +/- 39 omega.cm (n = 3 cells). 7. The mean electrotonic length of the dendritic paths was 0.83 +/- 0.2 (n = 5 cells). The mean input resistance at the dendritic terminals (RT) was 1,413 +/- 260 M omega. Synaptic conductances were applied at all distal dendritic compartments of the model cells. The resulting synaptic currents were calculated at the input site and at the soma. The mean transient current attenuation ratio was 4.7 +/- 1.7 under idealized voltage-clamp conditions.(ABSTRACT TRUNCATED AT 400 WORDS)
APA, Harvard, Vancouver, ISO, and other styles
41

Cifra, Alessandra, Graciela L. Mazzone, Francesca Nani, Andrea Nistri, and Miranda Mladinic. "Postnatal developmental profile of neurons and glia in motor nuclei of the brainstem and spinal cord, and its comparison with organotypic slice cultures." Developmental Neurobiology 72, no. 8 (June 21, 2012): 1140–60. http://dx.doi.org/10.1002/dneu.20991.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Larkum, M. E., M. G. Rioult, and H. R. Luscher. "Propagation of action potentials in the dendrites of neurons from rat spinal cord slice cultures." Journal of Neurophysiology 75, no. 1 (January 1, 1996): 154–70. http://dx.doi.org/10.1152/jn.1996.75.1.154.

Full text
Abstract:
1. We examined the propagation of action potentials in the dendrites of ventrally located presumed motoneurons of organotypic rat spinal cord cultures. Simultaneous patch electrode recordings were made from the dendrites and somata of individual cells. In other experiments we visualized the membrane voltage over all the proximal dendrites simultaneously using a voltage-sensitive dye and an array of photodiodes. Calcium imaging was used to measure the dendritic rise in Ca2+ accompanying the propagating action potentials. 2. Spontaneous and evoked action potentials were recorded using high-resistance patch electrodes with separations of 30-423 microm between the somatic and dendritic electrodes. 3. Action potentials recorded in the dendrites varied considerably in amplitude but were larger than would be expected if the dendrites were to behave as passive cables (sometimes little or no decrement was seen for distances of > 100 microm). Because the amplitude of the action potentials in different dendrites was not a simple function of distance from the soma, we suggest that the conductance responsible for the boosting of the action potential amplitude varied in density from dendrite to dendrite and possibly along each dendrite. 4. The dendritic action potentials were usually smaller and broader and arrived later at the dendritic electrode than at the somatic electrode irrespective of whether stimulation occurred at the dendrite or soma or as a result of spontaneous synaptic activity. This is clear evidence that the action potential is initiated at or near the soma and spreads out into the dendrites. The conduction velocity of the propagating action potential was estimated to be 0.5 m/s. 5. The voltage time courses of previously recorded action potentials were generated at the soma using voltage clamp before and after applying 1 microM tetrodotoxin (TTX) over the soma and dendrites. TTX reduced the amplitude of the action potential at the dendritic electrode to a value in the range expected for dendrites that behave as passive cables. This indicates that the conductance responsible for the actively propagating action potentials is a Na+ conductance. 6. The amplitude of the dendritic action potential could also be initially reduced more than the somatic action potential using 1-10 mM QX-314 (an intracellular sodium channel blocker) in the dendritic electrode as the drug diffused from the dendritic electrode toward the soma. Furthermore, in some cases the action potential elicited by current injection into the dendrite had two components. The first component was blocked by QX-314 in the first few seconds of the diffusion of the blocker. 7. In some cells, an afterdepolarizing potential (ADP) was more prominent in the dendrite than in the soma. This ADP could be reversibly blocked by 1 mM Ni2+ or by perfusion of a nominally Ca2+-free solution over the soma and dendrites. This suggests that the back-propagating action potential caused an influx of Ca2+ predominantly in the dendrites. 8. With the use of a voltage-sensitive dye (di-8-ANEPPS) and an array of photodiodes, the action potential was tracked along all the proximal dendrites simultaneously. The results confirmed that the action potential propagated actively, in contrast to similarly measured hyperpolarizing pulses that spread passively. There were also indications that the action potential was not uniformly propagated in all the dendrites, suggesting the possibility that the distribution of Na+ channels over the dendritic membrane is not uniform. 9. Calcium imaging with the Ca2+ fluorescent indicator Fluo-3 showed a larger percentage change in fluorescence in the dendrites than in the soma. Both bursts and single action potentials elicited sharp rises in fluorescence in the proximal dendrites, suggesting that the back-propagating action potential causes a concomitant rise in intracellular calcium concentration...
APA, Harvard, Vancouver, ISO, and other styles
43

Cho, Jung-Sun, Hwan-Woo Park, Sang-Kyu Park, Sangho Roh, Soo-Kyung Kang, Ki-Suk Paik, and Mi-Sook Chang. "Corrigendum to “Transplantation of mesenchymal stem cells enhances axonal outgrowth and cell survival in an organotypic spinal cord slice culture” [Neurosci. Lett. 454 (1) (2009) 43–48]." Neuroscience Letters 460, no. 1 (August 2009): 102. http://dx.doi.org/10.1016/j.neulet.2009.05.026.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Lüscher, Hans-R., and Matthew E. Larkum. "Modeling Action Potential Initiation and Back-Propagation in Dendrites of Cultured Rat Motoneurons." Journal of Neurophysiology 80, no. 2 (August 1, 1998): 715–29. http://dx.doi.org/10.1152/jn.1998.80.2.715.

Full text
Abstract:
Lüscher, Hans-R. and Matthew E. Larkum. Modeling action potential initiation and back-propagation in dendrites of cultured rat motoneurons. J. Neurophysiol. 80: 715–729, 1998. Regardless of the site of current injection, action potentials usually originate at or near the soma and propagate decrementally back into the dendrites. This phenomenon has been observed in neocortical pyramidal cells as well as in cultured motoneurons. Here we show that action potentials in motoneurons can be initiated in the dendrite as well, resulting in a biphasic dendritic action potential. We present a model of spinal motoneurons that is consistent with observed physiological properties of spike initiation in the initial segment/axon hillock region and action potential back-propagation into the dendritic tree. It accurately reproduces the results presented by Larkum et al. on motoneurons in organotypic rat spinal cord slice cultures. A high Na+-channel density of ḡ Na = 700 mS/cm2 at the axon hillock/initial segment region was required to secure antidromic invasion of the somato-dendritic membrane, whereas for the orthodromic direction, a Na+-channel density of ḡ Na = 1,200 mS/cm2 was required. A “weakly” excitable ( ḡ Na = 3 mS/cm2) dendritic membrane most accurately describes the experimentally observed attenuation of the back-propagated action potential. Careful analysis of the threshold conditions for action potential initiation at the initial segment or the dendrites revealed that, despite the lower voltage threshold for spike initiation in the initial segment, an action potential can be initiated in the dendrite before the initial segment fires a spike. Spike initiation in the dendrite depends on the passive cable properties of the dendritic membrane, its Na+-channel density, and local structural properties, mainly the diameter of the dendrites. Action potentials are initiated more easily in distal than in proximal dendrites. Whether or not such a dendritic action potential invades the soma with a subsequent initiation of a second action potential in the initial segment depends on the actual current source-load relation between the action potential approaching the soma and the electrical load of the soma together with the attached dendrites.
APA, Harvard, Vancouver, ISO, and other styles
45

SUGAI, Fuminobu, Yoichi YAMAMOTO, and Saburo SAKODA. "Organotypic spinal cord culture using mice." Folia Pharmacologica Japonica 124, no. 1 (2004): 19–23. http://dx.doi.org/10.1254/fpj.124.19.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Glazova, Margarita V., Elena S. Pak, and Alexander K. Murashov. "Neurogenic potential of spinal cord organotypic culture." Neuroscience Letters 594 (May 2015): 60–65. http://dx.doi.org/10.1016/j.neulet.2015.03.041.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Luscher, H.-R., and J. Streit. "A Novel In Vitro Approach for Studying the Segmental Motor System." Physiology 7, no. 6 (December 1, 1992): 249–53. http://dx.doi.org/10.1152/physiologyonline.1992.7.6.249.

Full text
Abstract:
An organotypic culture system of rat spinal cord, dorsal root ganglia, and skeletal muscle is presented that develops and preserves many structural and functional properties of the in vivo spinal cord. This in vitro model enlarges the methodological repertoire of mammalian spinal cord physiology and is ideally suited for studying developmental aspects.
APA, Harvard, Vancouver, ISO, and other styles
48

Nakayama, Kiyomi, Hiroshi Nishimaru, and Norio Kudo. "Rhythmic Motor Activity in Thin Transverse Slice Preparations of the Fetal Rat Spinal Cord." Journal of Neurophysiology 92, no. 1 (July 2004): 648–52. http://dx.doi.org/10.1152/jn.01029.2003.

Full text
Abstract:
Networks generating locomotor-like rhythmic motor activity are formed during the last week of the fetal period in the rat spinal cord. We investigated the coordinated rhythmic motor activity induced in transverse slice preparations of the lumbar spinal cord taken from fetal rats as early as embryonic day (E) 16.5. In slices as thin as 100 μm, bath-application of 5-hydroxytryptamine (5-HT) induced rhythmic [Ca2+]i elevations in motoneurons labeled with Calcium Green-1 dextran. The rhythmic [Ca2+]i elevations were similar in frequency to that in the intact lumbar spinal cord, although there was no temporal correlation between the activity in the left and right sides of 100-μm slices. Such rhythmic [Ca2+]i elevations were observed in the slices taken from all lumbar segments. Moreover, the rhythmic activity was abolished by simultaneous blockade of glutamate, glycine, and GABAA receptors, indicating that synaptic transmission mediated by these receptors is important for the generation of the rhythm in these slices. Synchronous rhythmic activity between the left-right sides was found in slices thicker than 200 μm taken from any segmental level of the lumbar spinal cord. In these preparations, commissural neurons were activated synchronously with ipsilateral motoneurons. These results indicate that the neuronal networks sufficient to generate coordinated rhythmic activity are contained in one-half of a single lumbar segment at E16.5. Such spinal cord slices are a promising experimental model to investigate the neuronal mechanisms and the development of rhythm generation in the spinal cord.
APA, Harvard, Vancouver, ISO, and other styles
49

Asai, Tatsuya, Takashi Saka, Shuichi Terao, Hiroshi Ikeda, and Kazuyuki Murase. "Intrinsic optical signals in rat spinal cord slices." Neuroscience Research 31 (January 1998): S134. http://dx.doi.org/10.1016/s0168-0102(98)82027-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Weidenheim, K. M., Y. Kress, W. K. Rashbaum, and W. D. Lyman. "ANTIBODY-ASSOCIATED MYELINOPATHY IN HUMAN FETAL SPINAL CORD ORGANOTYPIC CULTURES." Journal of Neuropathology and Experimental Neurology 54, no. 3 (May 1995): 465. http://dx.doi.org/10.1097/00005072-199505000-00233.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography