Skip to main content
  • Research article
  • Open access
  • Published:

Retroviral expression of human arginine decarboxylase reduces oxidative stress injury in mouse cortical astrocytes

Abstract

Background

In physiologic and pathologic conditions of the central nervous system (CNS), astrocytes are a double-edged sword. They not only support neuronal homeostasis but also contribute to increases in neuronal demise. A large body of experimental evidence has shown that impaired astrocytes play crucial roles in the pathologic process of cerebral ischemia; therefore, astrocytes may represent a breakthrough target for neuroprotective therapeutic strategies. Agmatine, an endogenous polyamine catalyzed from L-arginine by arginine decarboxylase (ADC), is a neuromodulator and it protects neurons/glia against various injuries.

Results

In this investigation, agmatine-producing mouse cortical astrocytes were developed through transduction of the human ADC gene. Cells were exposed to oxygen-glucose deprivation (OGD) and restored to a normoxic glucose-supplied condition. Intracellular levels of agmatine were measured by high performance liquid chromatography. Cell viability was evaluated by Hoechest/propidium iodide nuclear staining and lactate dehydrogenase assay. Expression of inducible nitric oxide synthase (iNOS) and matrix metalloproteinase s (MMPs) were assessed by a reverse transcription polymerase chain reaction, Western immunoblots, and immunofluorescence. We confirmed that ADC gene-expressed astrocytes produce a great amount of agmatine. These cells were highly resistant to not only OGD but also restoration, which mimicked ischemia-reperfusion injury in vivo. The neuroprotective effects of ADC seemed to be related to its ability to attenuate expression of iNOS and MMPs.

Conclusion

Our findings imply that astrocytes can be reinforced against oxidative stress by endogenous agmatine production through ADC gene transduction. The results of this study provide new insights that may lead to novel therapeutic approaches to reduce cerebral ischemic injuries.

Background

Astrocytes comprise the structural architecture of the brain and support neurons to maintain homeostasis [1, 2]. Physiologically, they stabilize the extracellular environment of neurons: they control ion/water distribution, regulate neurotransmitter recycling, release growth factors, and scavenge reactive oxygen species (ROS). Under pathologic conditions, when noxious stress exceeds limitations, astrocytes work in the opposite direction [1, 2]. Water accumulation induces brain edema, failure of neurotransmitter control, which causes glutamate excitotoxicity; and releases ROS/neurotoxic factors that contribute to further neuronal damage. In cerebral injuries, astrocytes are the main culprit of neurodegeneration and scar formation [2, 3]. Meanwhile, they also contribute to neurite outgrowth and neurogenesis [4]. Astrocytes have recently garnered attention as a breakthrough target in development of new therapeutic strategies for patients suffering from neuronal damage [2].

Agmatine, an endogenous polyamine derived from L-arginine by arginine decarboxylase (ADC), is naturally found in the mammalian central nervous system (CNS) and acts as a multifunctional neuromodulator [5–7]. It is packed into synaptic vesicles and released from synaptosomes by neuronal depolarization [8, 9]. Agmatine can stimulate α2-adrenergic and imidazoline receptors [10, 11], block the N-methyl-D-aspartic acid (NMDA) receptor and voltage-gated calcium channel [12, 13], and inhibit inducible/neuronal nitric oxide synthases (iNOS/nNOS) [14, 15]. Through these kinds of intracellular signaling, agmatine shows various antineurotoxic, anticonvulsant, antipsychotic, antidepressant, anxiolytic, and antinociceptive neuroprotective effects [7, 16]. Regarding ischemic insults, exogenous agmatine significantly reduces the infarct size and brain edema caused by ischemia-reperfusion injury in rodents [17–20]. Agmatine also rescues glial cells, as well as cortical neurons, from oxidative stress in vitro and in vivo[17, 21, 22].

In this study, a recombinant retroviral vector containing the human ADC (hADC) gene was introduced into mouse cortical astrocytes to maximize the astrocyte-protective effect of agmatine. Cells overexpressing hADC show increased production of agmatine and greater resistance against oxidative stress [23–25]. Here, we transduced primary cultured cortical astrocytes with hADC-expressing retroviral vector and assessed whether it induces synthesis of endogenous agmatine and attenuates oxidative stress injuries.

Methods

Cell culture

Primary cortical astrocyte cultures were prepared from postnatal 1–3 days old Crl:CD1 mouse pups (Samtako, Osan, Korea) [21, 26]. Our protocol was approved by the Institutional Animal Care and Use Committee of Yonsei University College of Medicine and all animals were treated in accordance with their guidelines. Animals were etherized and decapitated on ice-cold ethanol. Cortical hemispheres were dissected free of the skull and meninges, and then treated with 0.09% trypsin for 20 min at 37°C. The cells were mechanically dissociated into single cells and collected as a suspension. The cells were grown in minimum essential medium (MEM; Life Technologies, Grand Island, NY) containing 10% fetal bovine serum (FBS) and 10% equine serum (Thermo Fisher Scientific, Waltham, MA). The cultures were incubated at 37°C in a 5% CO2 incubator and the culture media was replaced every 3 days. One week after the first seeding, the other cells including microglia were removed by vigorous shaking. The highly-purified astrocytes (passage 3), at least 3 weeks after the first seeding, were used for experiments at about 80% confluence (Additional file 1).

Construction and infection of recombinant retroviral vector containing the hADC gene

A recombinant retroviral vector containing the hADC gene was constructed as described previously (Figure 1) [23–25]. Briefly, the full-length cDNA of hADC (GenBank accession no. AY325129) was amplified by a polymerase chain reaction (PCR) and ligated to the retroviral expression vector pLXSN (Clontech Laboratories, Mountain View, CA) that also contained a neomycin resistance gene. They were then amplified in E.coli DH5a and identified by restriction analysis. The hADC-expressing pLXSN vector was transfected into the retroviral packaging cell line PT67 using Lipofectamine 2000 (Sigma-Aldrich, St. Louis, MO). A stable clone selection was carried out by adding G418 (200 μg/mL; Sigma-Aldrich) to the culture medium of Dulbecco’s modified Eagle’s medium (DMEM, Life Technologies) supplemented with 10% FBS. High-titer clones were selected, and virus-containing supernatant was filtered through a 0.45-mm polysulfonic filter. Primary cultured mouse cortical astrocytes were infected with hADC pLXSN. The cells were incubated with the virus-containing media for 24 hrs and then maintained in normal culture medium for a week before being used for the experiments.

Figure 1
figure 1

Construction and infection of recombinant retroviral vector containing the human arginine decarboxylase (hADC) gene. (A) hADC pLXSN vector map: hADC pLXSN includes the Col E1 origin of replication and E.coli Ampr gene for propagation and antibiotic selection. The 5' viral LTR in this vector contains promoter/enhancer sequences that control expression of the gene of interest in the multiple cloning site. The SV40 early promoter (PSV40e) controls expression of the neomycin resistance gene (Neor), which allows antibiotic selection in eukaryotic cells. (B) Genetic confirmation by restriction analysis: Lane 1, size marker; Lane 2, restriction enzyme digestion using EcoRI and XhoI; Lane 3, control hADC pLXSN. (C) RNA levels of hADC were verified by RT-PCR in PT67 cells and mouse cortical astrocytes (Astro) after hADC pLXSN infection. (D) Expression of hADC was validated by immunocytochemistry in hADC pLXSN-infected astrocytes. Scale bar = 50 μm.

Oxygen-glucose deprivation and restoration

For oxygen-glucose deprivation (OGD), cultures were transferred to a closed anaerobic chamber (Forma Scientific Co., Seoul, Korea) in which the oxygen level (less than 0.1% O2) and temperature (37°C) were controlled automatically [21, 27]. After being washed three times with deoxygenated, glucose-free balanced salt solution (BSS0) at pH 7.4, cells were kept in the anaerobic chamber for 4 hours. Cultures were then restored with glucose at a concentration of 5.5 mM (BSS5.5) and recovered to normoxic conditions (37°C, 5% CO2) for up to 10 hours.

High performance liquid chromatography analysis

The concentration of agmatine was measured by high performance liquid chromatography (HPLC) [23, 24]. Harvested cells were homogenized in phosphate buffer containing 10% trichloroacetic acid and centrifuged at 12,000 rpm for 10 min. The collected supernatant was derivatized with o-phthalaldehyde and loaded onto the HPLC column (5 mm) connected with a fluorescence detector. Recovery of agmatine was calculated from an added external standard. The content of L-Arginine and putrescine was also measured using same method.

Hoechst and propidium iodide nuclear staining

Cell death was characterized by double nuclear staining with Hoechst 33258 and propidium iodide (PI) [23, 24]. Cells were stained with Hoechst 33258 dye (5 μg/mL; Sigma-Aldrich) for 30 min at 37°C and consecutively stained with PI (5 μg/mL; Sigma-Aldrich) just before observing it with a fluorescence microscope. Living cells showed Hoechst-positive blue fluorescence and nonviable cells had PI-positive red fluorescence.

Lactate dehydrogenase activity

Cell viability was quantified by measuring the amounts of lactate dehydrogenase (LDH) released from injured cells into the medium [27–30]. LDH released (% cytotoxicity) was calculated by dividing the value at the experimental time point by the maximum value × 100. Maximum LDH release values were obtained after freezing each culture at −70°C overnight and then rapidly thawing them, thus inducing nearly complete cell damage.

Reverse transcription polymerase chain reaction

Gene expression was assessed by reverse transcription PCR (RT-PCR) [23, 24]. Total RNA was extracted from cells with Trizol (Life Technologies) and quantified by measuring the absorbance at 260 nm. cDNA synthesis from mRNA and normalization of the samples were carried out by RT-PCR. PCR amplification of iNOS was performed at 94°C for 30 s, 57.5°C for 30 s, and 72°C for 30 s for 30 cycles. Amplification of MMP-2 was performed at 94°C for 30 s, 53°C for 30 s, and 72°C for 30 s for 35 cycles. Amplification for MMP-9 was performed at 94°C for 30 s, 59°C for 30 s, and 72°C for 30 s for 30 cycles. Specific primer sequences used are as follows: iNOS, sense 5'-TTT GAT GTG CTG CCT CTG GT-3' and antisense 5'-CAT TCT GCT TCT GGA AAC TAA GG-3'; MMP-2, sense 5'-GAG TTG GCA GTG CAA TAC CT-3' and antisense, 5'-GCC GTC CTT CTC AAA GTT GT-3'; MMP-9, sense 5'-TTA CCA GCG CCA GCC GAC TTT TG-3' and antisense 5'-CGT CGT CGT CGA AAT GGG CAT C-3'; Glyceraldehyde 3-phosphate dehydrogenase (GAPDH), sense 5'-ATG TCG TGG AGT CTA CTG GT-3' and antisense 5'-TGG CAT GGA CTG TGG TG-3'. PCR products were separated by electrophoresis in a 1.5% agarose gel stained with ethidium bromide.

Western immunoblots

Protein levels were determined by Western immunoblots [23, 24, 31]. Protein was extracted from whole cell lysates, and their concentration was calculated with the BCA protein assay (Thermo Fisher Scientific). Equal amounts of protein (40 μg) from each sample were resolved on a 10% sodium dodecyl sulfate-polyacrylamide gel. The separated proteins were then electrotransferred onto Immobilon-NC membranes (Millipore, Billerica, MA) and incubated overnight with primary antibodies against iNOS (1:1000 dilution), MMP-2 (1:2000 dilution), MMP-9 (1:2000 dilution), or β-actin (1:2000 dilution). The primary antibody against iNOS was purchased from Millipore (Billerica, MA) and the others were obtained from Santa Cruz Biotechnology (Dallas, TX). Immunoreactive bands were detected with horseradish peroxidase-conjugated secondary antibody and visualized using an enhanced chemiluminescent protein detection kit (Thermo Fisher Scientific).

Immunofluorescence staining

Protein expression was also confirmed by immunofluorescence [23, 31, 32]. Cells were fixed with 90% ethanol for 30 min on ice, and then treated with 1.6% hydrogen peroxide and 0.025% triton. They were reacted with specific antibodies (1:500 dilution) for iNOS (Millipore), MMP-2 (Santa Cruz Biotechnology), MMP-9 (Santa Cruz Biotechnology), and glial fibrillary acidic protein (GFAP; Abcam, Cambridge, MA). Subsequently, the cells were exposed to the corresponding fluorescent secondary antibodies (Life Technologies) and their nuclei were counterstained with 4’,6-diamidino-2-phenylindole (DAPI; Sigma-Aldrich). Finally, four random fields were imaged under a fluorescence microscope. Immunoreactive cells with fluorescence were manually counted at 200× magnification.

Statistical analysis

Data are presented at the mean of ± SD of at least three different experiments performed from separate cell preparations, with at least triplicate determination performed for each experiment. Kruskal-Wallis test and Mann–Whitney test were used in order to examine statistical significance (MedCalc Statistical Software version 12.7.8.; MedCalc Software bvba, Ostend, Belgium). P-values less than 0.05 were considered statistically significant.

Results

Transduction of mouse cortical astrocytes with recombinant retroviral vector containing the human arginine decarboxylase (hADC) gene

The vector map of hADC pLXSN is shown in Figure 1A. It includes the Col E1 origin of replication and E.coli Ampr gene for propagation and antibiotic selection. The 5' viral LTR contains promoter/enhancer sequences that control expression of the gene of interest in the multiple cloning site. The SV40 early promoter (PSV40e) controls expression of the neomycin resistance gene (Neor), which allows for antibiotic selection in eukaryotic cells. PSV40e was genetically confirmed by restriction analysis (Figure 1B). In astrocytes, as well as PT67 cells, infection of the hADC pLXSN vector effectively induced hADC transcription (Figure 1C). Expression of hADC was validated by immunocytochemistry in hADC pLXSN-infected astrocytes (Figure 1D).

Endogenous agmatine production from human arginine decarboxylase (hADC)-overexpressed astrocytes

Intracellular levels of agmatine were measured by HPLC (Figure 2). Endogenous agmatine concentration was markedly greater for astrocytes overexpressing hADC compared to that of control astrocytes (11.28-fold, p < 0.001). The amount of L-arginine and putrescine (precursor and metabolite of agmatine, respectively) showed similar tendency to that of agmatine. Meanwhile, OGD itself did not have a significant influence on the level of these amines.

Figure 2
figure 2

Endogenous agmatine measurement using high performance liquid chromatography (HPLC). The level of amines related to the agmatine pathway was determined in mouse cortical astrocytes infected with retrovirus containing the human arginine decarboxylase (hADC) gene. Cells underwent oxygen-glucose deprivation (OGD) for 4 hrs. Data are expressed as a mean ± SD of three different experiments performed from separate cell preparation, with triplicate determinations performed in each experiment. Asterisks indicate a p < 0.001 when the samples were compared to no treatment control.

Astrocyte-protective effect of human arginine decarboxylase (hADC) transduction against oxygen-glucose deprivation (OGD)

OGD damaged naive astrocytes and the majority of cells had PI-labeled nuclei (Figure 3B). However, astrocytes overexpressing hADC (hADC-astrocytes) were much less damaged by OGD, and the proportion of PI-positive cells were considerably lower (Figure 3D). hADC transduction itself did not have any significant effects on cell viability (Figure 3C), as determined by Hoechst 33258 and PI double nuclear staining.

Figure 3
figure 3

Astrocyte-protective effect of human arginine decarboxylase (hADC) transduction against oxygen-glucose deprivation (OGD). Cytotoxicity was assessed by Hoechst 33258 and propidium iodide (PI) nuclear staining; (A) No treatment control astrocytes, (B) astrocytes after 4 hrs OGD, (C) hADC-overexpressing astrocytes (hADC-astrocytes), (D) hADC-astrocytes after 4 hrs OGD. Scale bars = 100 μm. (E) Proportion of PI-labeled damaged cells with red fluorescence. Asterisks indicate a p < 0.05 when the samples were compared to no treatment control. Double asterisk indicate a p < 0.05 that that the two samples were significantly different which was confirmed by a post-hoc analysis.

Neuroprotective effect of human arginine decarboxylase (hADC) transduction on oxygen-glucose deprivation (OGD) and restoration injury

To mimic the ischemia-reperfusion injury in vivo, cells were restored to normoxic glucose-supplied conditions after 4 hrs of OGD (Figure 4). Cytotoxicity was measured by LDH leakage from the injured cells. Cytotoxicity gradually increased in a time-dependent manner for up to 10 hrs during restoration. Even in hADC-astrocytes, the restoration from OGD caused cytotoxicity in a time-dependent-manner; the extent of damage was much less than that in control astrocytes.

Figure 4
figure 4

Neuroprotective effect of human arginine decarboxylase (hADC) transduction against oxygen-glucose deprivation (OGD) and restoration injury. Astrocytes restored to normoxic glucose-supplied conditions after 4 hrs OGD. Cytotoxicity was assessed by the lactate dehydrogenase (LDH) assay. Data are expressed as mean of ± SD of the three different experiments performed from separate cell preparations, with triplicate determinations performed in each experiment. Asterisks indicate a p < 0.001.

Suppression of inducible nitric oxide synthase (iNOS) and matrix metalloproteinases (MMPs) in human arginine decarboxylase (hADC)-overexpressing astrocytes

Expression of iNOS and MMP-2/9 was assessed by RT-PCR (Figure 5A) and Western immunoblots (Figure 5B). OGD significantly increased levels of iNOS and MMP-2/9; however, hADC transduction attenuated the OGD effects. Introduction of the hADC gene itself did not change glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and β-actin levels, which served as internal controls for RT-PCR and Western immunoblots, respectively.

Figure 5
figure 5

Inducible nitric oxide synthase (iNOS) and matrix metalloproteinases (MMPs) expression in human arginine decarboxylase (hADC)-introduced astrocytes after oxygen-glucose deprivation (OGD). The mRNA levels were determined using reverse transcription polymerase chain reaction (RT-PCR) (A) and protein expression was evaluated using Western immunoblots (B). Glyceraldehyde 3-phosphate dehydrogenase (GAPDH) and β-actin were used as internal controls, respectively. Asterisks indicate a p < 0.05 when the samples were compared to no treatment control. Double asterisk indicate a p < 0.05 that two samples were significantly different which was confirmed by a post-hoc analysis.

Inhibited expression of inducible nitric oxide synthase (iNOS) and matrix metalloproteinases (MMPs) in human arginine decarboxylase (hADC)-transduced astrocytes

Naive astrocytes seldom expressed iNOS and MMP-2/9 (Figure 6A,D,G). After 4 hrs of OGD, control astrocytes exhibited expression of iNOS and MMP-2/9 (Figure 6B, E, H); however, hADC-astrocytes did not show immunoreactivity to either iNOS or MMP-2/9 (Figure 6C, F, I).

Figure 6
figure 6

Immunofluorescence staining of human arginine decarboxylase (hADC)-introduced astrocytes after oxygen-glucose deprivation (OGD). (A, D, G) No treatment control astrocytes; (B, E, H) astrocytes after 4 hrs OGD; (C, F, I) hADC-astrocytes after 4 hrs OGD. Cells reacted to inducible nitric oxide synthase (iNOS) (A-C), matrix metalloproteinase (MMP) 2 (D-F), and MMP 9 (G-I) showed green fluorescence. Cells were counter-stained with astrocyte marker glial fibrillary acidic protein (GFAP; red fluorescence) and nuclear marker 4',6-diamidino-2-phenylindole (DAPI; blue fluorescence). Scale bar = 50 μm.

Discussion

Since Reis and colleagues first discovered agmatine and ADC activity in the mammalian brain [5], agmatine has been widely studied. It is known to have various biological actions through multiple molecular targets, exhibiting affinity not only to α2-adrenoceptors and I1- and I2-imidazoline receptors but also to β-adrenoceptors and 5-HT3 serotonin, nicotinic cholinergic NMDA, D2-dopamine, κ-opioid, and adenosine A1 receptors [10–12]. Agmatine also inhibits iNOS/nNOS and monoamine oxidase [10, 14, 15]. As well, agmatine reduces heart rate and blood pressure [10], stimulates insulin and β-endorphin release, leading to increased cellular glucose uptake [33, 34], and accelerates glomerular ultrafiltration [35]. Agmatine protects systemic organs, especially against oxidative stress. Additionally, it shows cardioprotective and renoprotective effects on ischemic insult [36, 37], and it directly protects mitochondria from ROS [38, 39]. In the CNS, agmatine shows neuroprotective effects on various neuronal injuries: reduces the infarct size and edema after cerebral ischemia [17, 18, 40], attenuates brain damage and reactive gliosis caused by trauma [41, 42], and also decreases traumatic/ischemic spinal cord injuries [43, 44]. Agmatine also decreases neuropathic pain and convulsive events [45, 46], attenuates opioid dependence and alcohol withdrawal anxiety [47, 48], and it improves motor and cognitive functions in Parkinson's and Alzheimer's disease, respectively [49, 50]. Even in the eyes, agmatine lowers intraocular pressure and protects retinal ganglion cells [51, 52]. Taken together, agmatine seems to be a miracle cure rather than a specific drug.

In routine clinical settings, exogenous drug administration may be the most straightforward therapeutic strategy. However, exogenous drug administration is not always the best approach for real patients. Taken orally, agmatine is readily distributed throughout the body [53, 54] and can even cross the blood–brain barrier [55, 56]. Moreover, its reported half-life in the body is about 2 hrs [57]. The various functions of agmatine have pros and cons. In addition to neuronal effects, agmatine also has cardiovascular, endocrinal, renal, gastric, and growth effects [10]. Therefore, taken orally, agmatine may cause unwanted systemic responses. Thus, it is important to focus on the action of the drug at the site of injury and extend its duration of action. In this regard, in the ophthalmologic field, a topical agmatine ophthalmic solution has been formulated to preserve the optic nerves against glaucomatous damage [51]. Agmatine eye drops effectively lower intraocular pressure and protect retinal ganglion cells from chronic ocular pressure injuries. Along the same lines, investigators have attempted to develop a distinctive method for increasing endogenous agmatine production via a recombinant retroviral vector system containing the hADC gene. The hADC gene can be effectively delivered into mouse fibroblast NIH3T3 cells [23], primary mouse cortical neural stem cells [24], and human bone marrow mesenchymal stem cells [25]. Different types of transformed cells synthesize agmatine when they are exposed to hydrogen peroxide, resulting in resistance against oxidative stress. Irrespective of the cell type in an oxidative stressed environment, intracellular agmatine concentration is boosted by about 2 to 5-fold in hADC-overexpressing cells compared to no treatment control cells, as determined by HPLC [23–25]. While in normal culture conditions, agmatine synthesis is significantly increased in cortical neural stem cells [24] but not fibroblasts or mesenchymal stem cells [23, 25]. It is not yet clear why the activity of an introduced hADC gene differs according to cell type, especially in non-stressful conditions; nonetheless, it is assumed that agmatine controls the expression of the introduced hADC gene. If normal cells use little agmatine, the transformed cells are presumed not to utilize the introduced ADC without harmful stimuli. However, if normal cells actively use agmatine, the transformed cells are presumed to utilize the introduced ADC. When cells are attacked, all the cell types seem to use the introduced ADC.

In this investigation, we focused on cortical astrocytes, which have recently came to the forefront due to their suspected roles in many CNS injuries, including cerebral ischemia [2, 3]. Our results showed that intracellular levels of agmatine, as well as its precursor arginine and byproduct putrescine, are definitively increased (about 11-fold compared to the control, p < 0.001) in normal and oxidative stressed conditions upon hADC gene transduction, as measured by HPLC. Simultaneously, cell viability was determined by Hoechst /PI double nuclear staining and the LDH assay. hADC-overexpressing cells remained viable and healthy under oxidative stress conditions induced by OGD for 4 hrs. This astrocyte-rescuing property was gradually potentiated as restoration time proceeded for up to 10 hours. In addition, the neuroprotective effect of endogenous agmatine seemed to be related to iNOS intracellular signaling and the activity of MMPs, as assessed by RT-PCR, Western immunoblots, and immunofluorescence. Our present and previous findings are similar to those of other groups in regards to iNOS [15, 58–60] and MMPs [61, 62], respectively. As MMPs are upregulated by ischemic insult and degrade the basement membrane of brain microvessels [63], their suppression via ADC gene transduction may reduce ischemic injuries. Our retroviral system to deliver the hADC gene to target cells cannot be directly applied to non-dividing cells. Accordingly, another vector system using the Lentivirus is currently under development. It may be directly applied to non-dividing CNS neurons, specifically through in vivo transduction.

Conclusion

In summary, our findings imply that endogenous agmatine production through ADC gene transduction can protect astrocytes against oxidative stress. The results of this study provide new insights that may lead to novel therapeutic approaches to reduce oxidative stress-related diseases including cerebral ischemia.

Abbreviations

ADC:

Arginine decarboxylase

ANOVA:

One-way analysis of variance

BSS:

Balanced salt solution

CNS:

Central nervous system

DAPI:

4’,6-diamidino-2-phenylindole

DMEM:

Dulbecco’s modified Eagle’s medium

FBS:

Fetal bovine serum

GAPDH:

Glyceraldehyde 3-phosphate dehydrogenase

GFAP:

Glial fibrillary acidic protein

hADC:

Human arginine decarboxylase

HPLC:

High performance liquid chromatography

iNOS:

Inducible nitric oxide synthase

LDH:

Lactate dehydrogenase

MEM:

Minimum essential medium

MMP:

Matrix metalloproteinase

NMDA:

N-methyl-D-aspartic acid

nNOS:

Neuronal nitric oxide synthase

OGD:

Oxygen-glucose deprivation

PCR:

Polymerase chain reaction

PI:

Propidium iodide

ROS:

Reactive oxygen species

RT-PCR:

Reverse transcription polymerase chain reaction.

References

  1. Heneka MT, Rodríguez JJ, Verkhratsky A: Neuroglia in neurodegeneration. Brain Res Rev. 2010, 63: 189-211.

    Article  CAS  PubMed  Google Scholar 

  2. Barreto G, White RE, Ouyang Y, Xu L, Giffard RG: Astrocytes: targets for neuroprotection in stroke. Cent Nerv Syst Agents Med Chem. 2011, 11: 164-173.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  3. Nedergaard M, Dirnagl U: Role of glial cells in cerebral ischemia. Glia. 2005, 50: 281-286.

    Article  PubMed  Google Scholar 

  4. Chen Y, Swanson RA: Astrocytes and brain injury. J Cereb Blood Flow Metab. 2003, 23: 137-149.

    Article  PubMed  Google Scholar 

  5. Li G, Regunathan S, Barrow CJ, Eshraghi J, Cooper R, Reis DJ: Agmatine: an endogenous clonidine-displacing substance in the brain. Science. 1994, 263: 966-969.

    Article  CAS  PubMed  Google Scholar 

  6. Reis DJ, Regunathan S: Agmatine: an endogenous ligand at imidazoline receptors is a novel neurotransmitter. Ann N Y Acad Sci. 1999, 881: 65-80.

    Article  CAS  PubMed  Google Scholar 

  7. Halaris A, Plietz J: Agmatine: metabolic pathway and spectrum of activity in brain. CNS Drugs. 2007, 21: 885-900.

    Article  CAS  PubMed  Google Scholar 

  8. Reis DJ, Yang XC, Milner TA: Agmatine containing axon terminals in rat hippocampus form synapses on pyramidal cells. Neurosci Lett. 1998, 250: 185-188.

    Article  CAS  PubMed  Google Scholar 

  9. Goracke-Postle CJ, Nguyen HO, Stone LS, Fairbanks CA: Release of tritiated agmatine from spinal synaptosomes. Neuroreport. 2006, 17: 13-17.

    Article  PubMed  Google Scholar 

  10. Raasch W, Schäfer U, Chun J, Dominiak P: Biological significance of agmatine, an endogenous ligand at imidazoline binding sites. Br J Pharmacol. 2001, 133: 755-780.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Sugiura T, Kobuchi S, Tsutsui H, Takaoka M, Fujii T, Hayashi K, Matsumura Y: Preventive mechanisms of agmatine against ischemic acute kidney injury in rats. Eur J Pharmacol. 2009, 603: 108-113.

    Article  CAS  PubMed  Google Scholar 

  12. Yang XC, Reis DJ: Agmatine selectively blocks the N-methyl-D-aspartate subclass of glutamate receptor channels in rat hippocampal neurons. J Pharmacol Exp Ther. 1999, 288: 544-549.

    CAS  PubMed  Google Scholar 

  13. Wang G, Gorbatyuk OS, Dayanithi G, Ouyang W, Wang J, Milner TA, Regunathan S, Reis DJ: Evidence for endogenous agmatine in hypothalamo-neurohypophysial tract and its modulation on vasopressin release and Ca2+ channels. Brain Res. 2002, 932: 25-36.

    Article  CAS  PubMed  Google Scholar 

  14. Auguet M, Viossat I, Marin JG, Chabrier PE: Selective inhibition of inducible nitric oxide synthase by agmatine. Jpn J Pharmacol. 1995, 69: 285-287.

    Article  CAS  PubMed  Google Scholar 

  15. Demady DR, Jianmongkol S, Vuletich JL, Bender AT, Osawa Y: Agmatine enhances the NADPH oxidase activity of neuronal NO synthase and leads to oxidative inactivation of the enzyme. Mol Pharmacol. 2001, 59: 24-29.

    CAS  PubMed  Google Scholar 

  16. Uzbay TI: The pharmacological importance of agmatine in the brain. Neurosci Biobehav Rev. 2012, 36: 502-519.

    Article  CAS  PubMed  Google Scholar 

  17. Kim JH, Yenari MA, Giffard RG, Cho SW, Park KA, Lee JE: Agmatine reduces infarct area in a mouse model of transient focal cerebral ischemia and protects cultured neurons from ischemia-like injury. Exp Neurol. 2004, 189: 122-130.

    Article  CAS  PubMed  Google Scholar 

  18. Kim JH, Lee YW, Park KA, Lee WT, Lee JE: Agmatine attenuates brain edema through reducing the expression of aquaporin-1 after cerebral ischemia. J Cereb Blood Flow Metab. 2010, 30: 943-949.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Uranchimeg D, Kim JH, Kim JY, Lee WT, Park KA, Batbaatar G, Tundevrentsen S, Amgalanbaatar D, Lee JE: Recovered changes in the spleen by agmatine treatment after transient cerebral ischemia. Anat Cell Biol. 2010, 43: 44-53.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  20. Wang CC, Chio CC, Chang CH, Kuo JR, Chang CP: Beneficial effect of agmatine on brain apoptosis, astrogliosis, and edema after rat transient cerebral ischemia. BMC Pharmacol. 2010, 10: 11.

    Article  PubMed Central  PubMed  Google Scholar 

  21. Lee WT, Hong S, Yoon SH, Kim JH, Park KA, Seong GJ, Lee JE: Neuroprotective effects of agmatine on oxygen-glucose deprived primary-cultured astrocytes and nuclear translocation of nuclear factor-kappa B. Brain Res. 2009, 1281: 64-70.

    Article  CAS  PubMed  Google Scholar 

  22. Ahn SK, Hong S, Park YM, Lee WT, Park KA, Lee JE: Effects of agmatine on hypoxic microglia and activity of nitric oxide synthase. Brain Res. 2011, 1373: 48-54.

    Article  CAS  PubMed  Google Scholar 

  23. Moon SU, Kwon KH, Kim JH, Bokara KK, Park KA, Lee WT, Lee JE: Recombinant hexahistidine arginine decarboxylase (hisADC) induced endogenous agmatine synthesis during stress. Mol Cell Biochem. 2010, 345: 53-60.

    Article  CAS  PubMed  Google Scholar 

  24. Bokara KK, Kwon KH, Nho Y, Lee WT, Park KA, Lee JE: Retroviral expression of arginine decarboxylase attenuates oxidative burden in mouse cortical neural stem cells. Stem Cells Dev. 2011, 20: 527-537.

    Article  CAS  PubMed  Google Scholar 

  25. Seo SK, Yang W, Park YM, Lee WT, Park KA, Lee JE: Overexpression of human arginine decarboxylase rescues human mesenchymal stem cells against H2O2 toxicity through cell survival protein activation. J Korean Med Sci. 2013, 28: 366-373.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  26. Lee JE, Kim CY, Giaccia AJ, Giffard RG: The E6 and E7 genes of human papilloma virus-type 16 protect primary astrocyte cultures from injury. Brain Res. 1998, 795: 10-16.

    Article  CAS  PubMed  Google Scholar 

  27. Hong S, Lee JE, Kim CY, Seong GJ: Agmatine protects retinal ganglion cells from hypoxia-induced apoptosis in transformed rat retinal ganglion cell line. BMC Neurosci. 2007, 8: 81.

    Article  PubMed Central  PubMed  Google Scholar 

  28. Wroblewski F, Ladue JS: Lactic dehydrogenase activity in blood. Proc Soc Exp Biol Med. 1955, 90: 210-213.

    Article  CAS  PubMed  Google Scholar 

  29. Koh JY, Choi DW: Quantitative determination of glutamate mediated cortical neuronal injury in cell culture by lactate dehydrogenase efflux assay. J Neurosci Methods. 1987, 20: 83-90.

    Article  CAS  PubMed  Google Scholar 

  30. Hong S, Kim CY, Lee JE, Seong GJ: Agmatine protects cultured retinal ganglion cells from tumor necrosis factor-alpha-induced apoptosis. Life Sci. 2009, 84: 28-32.

    Article  CAS  PubMed  Google Scholar 

  31. Hong S, Iizuka Y, Kim CY, Seong GJ: Isolation of primary mouse retinal ganglion cells using immunopanning-magnetic separation. Mol Vis. 2012, 18: 2922-2930.

    PubMed Central  CAS  PubMed  Google Scholar 

  32. Bechara RI, Brown LA, Roman J, Joshi PC, Guidot DM: Transforming growth factor beta1 expression and activation is increased in the alcoholic rat lung. Am J Respir Crit Care Med. 2004, 170: 188-194.

    Article  PubMed  Google Scholar 

  33. Sener A, Lebrun P, Blachier F, Malaisse WJ: Stimulus-secretion coupling of arginine-induced insulin release. Insulinotropic action of agmatine. Biochem Pharmacol. 1989, 38: 327-330.

    Article  CAS  PubMed  Google Scholar 

  34. Chang CH, Wu HT, Cheng KC, Lin HJ, Cheng JT: Increase of beta-endorphin secretion by agmatine is induced by activation of imidazoline I(2A) receptors in adrenal gland of rats. Neurosci Lett. 2010, 468: 297-299.

    Article  CAS  PubMed  Google Scholar 

  35. Schwartz D, Peterson OW, Mendonca M, Satriano J, Lortie M, Blantz RC: Agmatine affects glomerular filtration via a nitric oxide synthase-dependent mechanism. Am J Physiol. 1997, 272: F597-F601.

    CAS  PubMed  Google Scholar 

  36. Greenberg S, George J, Wollman Y, Shapira I, Laniado S, Keren G: The effect of agmatine administration on ischemic-reperfused isolated rat heart. J Cardiovasc Pharmacol Ther. 2001, 6: 37-45.

    Article  CAS  PubMed  Google Scholar 

  37. Sugiura T, Tsutsui H, Takaoka M, Kobuchi S, Hayashi K, Fujii T, Matsumura Y: Protective effect of agmatine on ischemia/reperfusion-induced renal injury in rats. J Cardiovasc Pharmacol. 2008, 51: 223-230.

    Article  CAS  PubMed  Google Scholar 

  38. Arndt MA, Battaglia V, Parisi E, Lortie MJ, Isome M, Baskerville C, Pizzo DP, Ientile R, Colombatto S, Toninello A, Satriano J: The arginine metabolite agmatine protects mitochondrial function and confers resistance to cellular apoptosis. Am J Physiol Cell Physiol. 2009, 296: C1411-C1419.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  39. Battaglia V, Grancara S, Satriano J, Saccoccio S, Agostinelli E, Toninello A: Agmatine prevents the Ca(2+)-dependent induction of permeability transition in rat brain mitochondria. Amino Acids. 2010, 38: 431-437.

    Article  CAS  PubMed  Google Scholar 

  40. Huang YC, Tzeng WS, Wang CC, Cheng BC, Chang YK, Chen HH, Lin PC, Huang TY, Chuang TJ, Lin JW, Chang CP: Neuroprotective effect of agmatine in rats with transient cerebral ischemia using MR imaging and histopathologic evaluation. Magn Reson Imaging. 2013, 31: 1174-1181.

    Article  CAS  PubMed  Google Scholar 

  41. Sengul G, Takci E, Malcok UA, Akar A, Erdogan F, Kadioglu HH, Aydin IH: A preliminary histopathological study of the effect of agmatine on diffuse brain injury in rats. J Clin Neurosci. 2008, 15: 1125-1129.

    Article  CAS  PubMed  Google Scholar 

  42. Kuo JR, Lo CJ, Chang CP, Lin KC, Lin MT, Chio CC: Agmatine-promoted angiogenesis, neurogenesis, and inhibition of gliosis-reduced traumatic brain injury in rats. J Trauma. 2011, 71: E87-E93.

    Article  PubMed  Google Scholar 

  43. Gilad GM, Gilad VH: Accelerated functional recovery and neuroprotection by agmatine after spinal cord ischemia in rats. Neurosci Lett. 2000, 296: 97-100.

    Article  CAS  PubMed  Google Scholar 

  44. Yu CG, Marcillo AE, Fairbanks CA, Wilcox GL, Yezierski RP: Agmatine improves locomotor function and reduces tissue damage following spinal cord injury. Neuroreport. 2000, 11: 3203-3207.

    Article  CAS  PubMed  Google Scholar 

  45. Onal A, Delen Y, Ulker S, Soykan N: Agmatine attenuates neuropathic pain in rats: possible mediation of nitric oxide and noradrenergic activity in the brainstem and cerebellum. Life Sci. 2003, 73: 413-428.

    Article  CAS  PubMed  Google Scholar 

  46. Su RB, Wei XL, Zheng JQ, Liu Y, Lu XQ, Li J: Anticonvulsive effect of agmatine in mice. Pharmacol Biochem Behav. 2004, 77: 345-349.

    Article  CAS  PubMed  Google Scholar 

  47. Regunathan S: Agmatine: biological role and therapeutic potentials in morphine analgesia and dependence. AAPS J. 2006, 8: E479-E484.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Taksande BG, Kotagale NR, Patel MR, Shelkar GP, Ugale RR, Chopde CT: Agmatine, an endogenous imidazoline receptor ligand modulates ethanol anxiolysis and withdrawal anxiety in rats. Eur J Pharmacol. 2010, 637: 89-101.

    Article  CAS  PubMed  Google Scholar 

  49. Wiesinger H: Arginine metabolism and the synthesis of nitric oxide in the nervous system. Prog Neurobiol. 2001, 64: 365-391.

    Article  CAS  PubMed  Google Scholar 

  50. Condello S, Calabrò E, Caccamo D, Currò M, Ferlazzo N, Satriano J, Magazù S, Lentile R: Protective effects of agmatine in rotenone-induced damage of human SH-SY5Y neuroblastoma cells: fourier transform infrared spectroscopy analysis in a model of Parkinson's disease. Amino Acids. 2012, 42: 775-781.

    Article  CAS  PubMed  Google Scholar 

  51. Hong S, Kim CY, Lee WS, Shim J, Yeom HY, Seong GJ: Ocular hypotensive effects of topically administered agmatine in a chronic ocular hypertensive rat model. Exp Eye Res. 2010, 90: 97-103.

    Article  CAS  PubMed  Google Scholar 

  52. Hong S, Hara H, Shimazawa M, Hyakkoku K, Kim CY, Seong GJ: Retinal protective effects of topically administered agmatine on ischemic ocular injury caused by transient occlusion of the ophthalmic artery. Braz J Med Biol Res. 2012, 45: 212-215.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  53. Molderings GJ, Heinen A, Menzel S, Lubbecke F, Homann J, Gothert M: Gastrointestinal uptake of agmatine: distribution in tissues and organs and pathophysiologic relevance. Ann N Y Acad Sci. 2003, 1009: 44-51.

    Article  CAS  PubMed  Google Scholar 

  54. Haenisch B, von Kugelgen I, Bonisch H, Gothert M, Sauerbruch T, Schepke M, Marklein G, Hofling K, Schroder D, Molderings GJ: Regulatory mechanisms underlying agmatine homeostasis in humans. Am J Physiol Gastrointest Liver Physiol. 2008, 295: G1104-G1110.

    Article  CAS  PubMed  Google Scholar 

  55. Nguyen HO, Goracke-Postle CJ, Kaminski LL, Overland AC, Morgan AD, Fairbanks CA: Neuropharmacokinetic and dynamic studies of agmatine (decarboxylated arginine). Ann N Y Acad Sci. 2003, 1009: 82-105.

    Article  CAS  PubMed  Google Scholar 

  56. Piletz JE, May PJ, Wang G, Zhu H: Agmatine crosses the blood–brain barrier. Ann N Y Acad Sci. 2003, 1009: 64-74.

    Article  CAS  PubMed  Google Scholar 

  57. Huisman H, Wynveen P, Nichkova M, Kellermann G: Novel ELISAs for screening of the biogenic amines GABA, glycine, beta-phenylethylamine, agmatine, and taurine using one derivatization procedure of whole urine samples. Anal Chem. 2010, 82: 6526-6533.

    Article  CAS  PubMed  Google Scholar 

  58. Satriano J, Schwartz D, Ishizuka S, Lortie MJ, Thomson SC, Gabbai F, Kelly CJ, Blantz RC: Suppression of inducible nitric oxide generation by agmatine aldehyde: beneficial effects in sepsis. J Cell Physiol. 2001, 188: 313-320.

    Article  CAS  PubMed  Google Scholar 

  59. Regunathan S, Piletz JE: Regulation of inducible nitric oxide synthase and agmatine synthesis in macrophages and astrocytes. Ann N Y Acad Sci. 2003, 1009: 20-29.

    Article  CAS  PubMed  Google Scholar 

  60. Satriano J: Arginine pathways and the inflammatory response: interregulation of nitric oxide and polyamines: review article. Amino Acids. 2004, 26: 321-329.

    Article  CAS  PubMed  Google Scholar 

  61. Yang MZ, Mun CH, Choi YJ, Baik JH, Park KA, Lee WT, Lee JE: Agmatine inhibits matrix metalloproteinase-9 via endothelial nitric oxide synthase in cerebral endothelial cells. Neurol Res. 2007, 29: 749-754.

    Article  CAS  PubMed  Google Scholar 

  62. Jung HJ, Yang MZ, Kwon KH, Yenari MA, Choi YJ, Lee WT, Park KA, Lee JE: Endogenous agmatine inhibits cerebral vascular matrix metalloproteinases expression by regulating activating transcription factor 3 and endothelial nitric oxide synthesis. Curr Neurovasc Res. 2010, 7: 201-212.

    Article  CAS  PubMed  Google Scholar 

  63. Heo JH, Lucero J, Abumiya T, Koziol JA, Copeland BR, del Zoppo GJ: Matrix metalloproteinases increase very early during experimental focal cerebral ischemia. J Cereb Blood Flow Metab. 1999, 19: 624-633.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgment

This work was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (MEST) (No. 2011–0013288, 2012–0005440) and by the Brain Korea 21 Project for Medical Science, Yonsei University, Seoul, Republic of Korea.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jong Eun Lee.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

LJE, HS, and SMR designed the experiments. HS, SMR, and YK carried out the molecular studies and Lee JE supervised the whole process. HS and SMR wrote the bulk of the manuscript and LWT and PKA revised the manuscript. LJE has given final approval of the version to be published. All authors read approved the final manuscript.

Electronic supplementary material

12868_2014_3787_MOESM1_ESM.tiff

Additional file 1: Immunofluorescence staining of primary cultured astrocytes. Cells reacted to anti-glial fibrillary acidic protein (GFAP) (A, red fluorescence), and anti-CD11b (B, green fluorescence) antibodies. Nuclei were counterstained with Hoechst 33258 (C, blue fluorescence). Panel D is a merged image. Scale bar = 200 μm. (TIFF 559 KB)

Authors’ original submitted files for images

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 https://creativecommons.org/licenses/by/4.0/.

The Creative Commons Public Domain Dedication waiver (https://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

Hong, S., Son, M.R., Yun, K. et al. Retroviral expression of human arginine decarboxylase reduces oxidative stress injury in mouse cortical astrocytes. BMC Neurosci 15, 99 (2014). https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2202-15-99

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2202-15-99

Keywords