Skip to main content
Intended for healthcare professionals
Free access
Research article
First published November 2006

Time Course of Post-Traumatic Mitochondrial Oxidative Damage and Dysfunction in a Mouse Model of Focal Traumatic Brain Injury: Implications for Neuroprotective Therapy

Abstract

In the present study, we investigate the hypothesis that mitochondrial oxidative damage and dysfunction precede the onset of neuronal loss after controlled cortical impact traumatic brain injury (TBI) in mice. Accordingly, we evaluated the time course of post-traumatic mitochondrial dysfunction in the injured cortex and hippocampus at 30 mins, 1, 3, 6, 12, 24, 48, and 72 h after severe TBI. A significant decrease in the coupling of the electron transport system with oxidative phosphorylation was observed as early as 30 mins after injury, followed by a recovery to baseline at 1 h after injury. A statistically significant (P < 0.0001) decline in the respiratory control ratio was noted at 3 h, which persisted at all subsequent time-points up to 72 h after injury in both cortical and hippocampal mitochondria. Structural damage seen in purified cortical mitochondria included severely swollen mitochondria, a disruption of the cristae and rupture of outer membranes, indicative of mitochondrial permeability transition. Consistent with this finding, cortical mitochondrial calcium-buffering capacity was severely compromised by 3h after injury, and accompanied by significant increases in mitochondrial protein oxidation and lipid peroxidation. A possible causative role for reactive nitrogen species was suggested by the rapid increase in cortical mitochondrial 3-nitrotyrosine levels shown as early as 30 mins after injury. These findings indicate that post-traumatic oxidative lipid and protein damage, mediated in part by peroxynitrite, occurs in mitochondria with concomitant ultrastructural damage and impairment of mitochondrial bioenergetics. The data also indicate that compounds which specifically scavenge peroxynitrite (ONOO) or ONOOderived radicals (e.g. ONOO + H+ → ONOOH → NO2 + OH) may be particularly effective for the treatment of TBI, although the therapeutic window for this neuroprotective approach might only be 3 h.

Introduction

Mitochondria are important in cellular bioenergetics and are sensitive to changes in the physiological state of cells. Accumulating evidence indicates that mitochondria play a pivotal role in both the necrotic and apoptotic processes in mammalian cells. Disruption of mitochondrial membrane potential (Δϕm) is considered to be an indicator of mitochondria damage and generally is defined as an early stage of apoptosis, preceding the efflux of macromolecules from the mitochondria (including cytochrome c, apoptosis-inducing factor, cellular inhibitor of apoptosis proteins (cIAPs), etc.) and caspase-9/caspase-3 cascade activation (Green and Reed, 1998; Spierings et al, 2005).
Mitochondria also appear to play a critical role in the secondary injury that occurs after traumatic brain injury (TBI) (Finkel, 2001; Hunot and Flavell, 2001), and mitochondrial dysfunction has been shown to be involved in excitatory amino acid (EAA)-induced neurotoxicity (White and Reynolds, 1996; Nicholls and Budd, 1998; Stout et al, 1998; Kroemer and Reed, 2000; Jiang et al, 2001; Brustovetsky et al, 2002). Mitochondrial dysfunction after TBI has been linked to impairment of brain mitochondrial electron transfer and energy transduction due to overloading of mitochondrion-associated calcium (Xiong et al, 1997), increased mitochondrial reactive oxygen species (ROS) production, oxidative damage, disruption of synaptic homeostasis (Azbill et al, 1997; Matsushita and Xiong, 1997; Sullivan et al, 1999a, b, c), and cell death (Robertson, 2004).
Most convincing with regard to the importance of mitochondrial failure in secondary brain injury, pharmacological agents that target mitochondria have been shown to be neuroprotective. Administration of cyclosporine A (CsA) after experimental TBI significantly reduces mitochondrial dysfunction (Sullivan et al, 1999c), cortical damage (Scheff and Sullivan, 1999; Sullivan et al, 2000b, c), and cytoskeleton and axonal dysfunction (Okonkwo et al, 1999; Okonkwo and Povlishock, 1999). Subsequently, Sullivan et al (2000a) have shown that dietary supplementation with creatine is also effective in ameliorating neuronal cell death by reducing mitochondrial ROS production and maintaining adenosine triphosphate levels after TBI. Although these prior studies clearly show the importance of mitochondrial dysfunction and failure in secondary brain injury, and the potential therapeutic value of pharmacological mitochondrial protection, the precise time course, characteristics and causes of post-traumatic mitochondrial dysfunction have not been fully defined. This is important to achieve to better understand the optimal mechanistic approaches and the therapeutic time window for being able to salvage mitochondrial function and achieve successful neuroprotection in the injured brain.
Thus, the current investigation was performed in the context of the mouse lateral controlled cortical impact (CCI)-TBI model (Smith et al, 1995; Raghupathi et al, 1998; Hannay et al, 1999; Sullivan et al, 1999a, c), and follows our recent definition of the time course of calpain-mediated cytoskeleton degradation and neurodegeneration in this paradigm (Hall et al, 2005). That study showed that cytoskeleton and neuronal degradation did not become apparent until 6 h after injury and do not peak until 24 to 48 h. In the present study, we report marked and sustained decreases in mitochondrial respiratory rates, respiratory control ratios (RCRs) and calcium-buffering capacity, indices of mitochondrial dysfunction, which become statistically significant by 3h and peaks by 12 h after TBI. The onset of metabolic dysfunction is coincident with mitochondrial oxidative damage, suggesting that oxygen radical-induced modification of membrane lipids and proteins might be pivotal factors in post-traumatic mitochondrial dysfunction. Based on its time course, a major player in this oxidative damage scenario appears to be the ROS peroxynitrite (ONOO).

Materials and methods

Materials

Mannitol, sucrose, bovine serum albumin (BSA), ethyleneglycol tetraacetate (EGTA), N−2-hydroxyethylpiperazine-N'−2-ethanesulfonic acid (HEPES) potassium salt, potassium phosphate monobasic anhydrous (KH2PO4), magnesium chloride (MgCl2), malate, pyruvate, adenosine 5'-diphosphate (ADP), oligomycin A, carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone (FCCP), and succinate were purchased from Sigma-Aldrich (St Louis, MO, USA). BCA protein assay kit was purchased from Pierce (Rockford, IL, USA).

Animals

The studies were performed using 350 young adult male CF-1 mice (Charles River, Portage, MI, USA) weighing 29 to 31 g. Ipsilateral cortical and ipsilateral hippocampal tissues from five mice were pooled as N1 and we used 3 N's (15 animals) for TBI and 1 N (5 animals) for sham at every time-point studied, as described in the present paper. All the protocols for performing surgery and TBI have been approved by the University of Kentucky Institutional Animal Care and Use Committee.

Mouse Model of Focal (Lateral Controlled Cortical Impact) Traumatic Brain Injury

Mice were initially anesthetized in a plexiglass chamber using 4.0% isoflurane and placed in a stereotaxic frame (David Kopf, Tujunga, CA, USA). During the injury procedure, anesthesia was maintained with 2.5% isoflurane delivered via a nose cone. The head was positioned in the horizontal plane with the nose bar set at zero. After a midline incision exposing the skull, a 4-mm craniotomy was made lateral to the sagittal suture and centered between the bregma and lambda. The skull cap at the craniotomy was carefully removed without damaging the underlying dura, and the exposed cortex was injured using a pneumatically controlled impactor device as described previously (Sullivan et al, 1999a, c). The impactor tip diameter was 3 mm and the impact velocity was set at 3.5 m/sec with a cortical compression of 1.0 mm. After injury, the craniotomy was closed by placement of a small disk of saline moistened Surgicil over the dura, followed by super gluing a 6 mm disk made of dental cement over the site. The animals were then placed in a Hova-Bator Incubator (model 1583; Randall Burkey Co, Boerne, TX, USA), which was set at 37°C until consciousness (i.e. return of righting reflex and mobility) was regained to prevent hypothermia. Injured mice were allowed to survive from 30 mins to 3 days depending on experimental group assignment. A sham group was included at all time-points which underwent anesthesia and a craniotomy, but no cortical impact or injury.

Isolation of Percoll-Purified Mitochondria

Brain mitochondria were extracted as described previously with some modifications (Sullivan et al, 2000a, 2003). Briefly, the mice were decapitated and the ipsilateral brain regions of interest (cortex and hippocampus) were quickly removed and pooled from five mice. The cortices and hippocampi were placed in a Potter-Elvejhem homogenizer containing five times the volume of ice-cold isolation buffer (five-fold buffer to tissue) with 1 mmol/L EGTA (215 mmol/L mannitol, 75 mmol/L sucrose, 0.1% BSA, 20 mmol/L HEPES, 1 mmol/L EGTA; pH adjusted to 7.2 with KOH), and the homogenates were subjected to differential centrifugation at 4°C. First, the homogenate was centrifuged twice at 1300g for 3 mins in an Eppendorf microcentrifuge at 4°C to remove cellular debris and nuclei. The pellet was discarded and the supernatant further centrifuged at 13,000g for 10 mins. The crude mitochondrial pellet obtained after differential centrifugation was then subjected to nitrogen decompression to release synaptic mitochondria, using a nitrogen cell disruption bomb, cooled to 4°C under a pressure of 1200 psi for 10 mins (Sullivan et al, 2003; Brown et al, 2004; Kristian et al, 2005). After nitrogen disruption the mitochondria were suspended in 3.5 mL of 15% percoll and further laid on two preformed layers consisting of 3.5 mL of 24% percoll and 3.5 mL of 40% percoll in 13 mL ultraclear tubes. The gradient was centrifuged in fixed angle rotor at 30,400g for 10 mins at 4°C. The fraction accumulated at the interphase of 40% and 24% percoll was carefully removed and diluted with isolation buffer without EGTA, then centrifuged at 12,300g for 10 mins at 4°C. The supernatants were carefully removed, the pellet resuspended in isolation buffer without EGTA and centrifuged at 13,000 × g for 10 mins at 4°C. The mitochondrial pellets at the bottom were then transferred to microcentrifuge tubes and topped off with isolation buffer without EGTA and centrifuged at 10,000g for 5 mins at 4°C to yield a tighter pellet. The final mitochondrial pellet was resuspended in isolation buffer without EGTA to yield a concentration of ~10mg/mL. The protein concentration was determined using the BCA protein assay kit measuring absorbance at 562 nm with a BioTek Synergy HT plate reader (Winooski, VT, USA).

Mitochondrial Respiration Studies

Mitochondrial respiratory rates were measured using a Clark-type electrode in a continuously stirred, sealed thermostatically controlled chamber (Oxytherm System, Hansatech Instruments Ltd) maintained at 37°Cas described previously (Sullivan et al, 2000a, 2003; Jiang et al, 2001; Sensi et al, 2003). In all, 25 to 40 μg of isolated mitochondrial protein was placed in the chamber containing 250 μL of KCl-based respiration buffer (125 mmol/L KCl, 2 mmol/L MgCl2, 2.5 mmol/L KH2PO4, 0.1% BSA, 20 mmol/L HEPES at pH 7.2) and allowed to equilibrate for 1 min. This was followed by the addition of complex-I substrates, 5 mmol/L pyruvate and 2.5 mmol/L malate, to monitor state II respiratory rate. Two boluses of 150 μmol/L ADP were added to the mitochondria to initiate state III respiratory rate for 2 mins, followed by the addition of 2 μmol/L oligomycin to monitor state IV respiration rate for an additional 2 mins. For the measurement of uncoupled respiratory rate, 2 μmol/L FCCP was added to the mitochondria in the chamber and oxygen consumption was monitored for another 2 mins. This was followed by the addition of 10 mmol/L succinate to monitor complex II-driven respiration. The RCR was calculated by dividing state III oxygen consumption (defined as the rate of respiration in the presence of ADP, second bolus addition) by the state IV oxygen consumption (rate obtained in the presence of oligomycin). Fresh mitochondria were prepared for each experiment and used within 4 h.

Mitochondrial Calcium-Buffering Studies

Mitochondrial calcium-buffering measurements were performed in a Shimadzu RF-5301 spectrofluorimeter maintained at 37°C. The extra-mitochondrial concentrations of Ca2+ were performed using the fluorescent Ca2+ -sensitive indicator Calcium Green 5 N (CaG5N) (Molecular Probes, Eugene, OR, USA) at a final concentration of 100 nmol/L with excitation and emission wavelengths 506 nm (slit 5 nm) and 532 nm (slit 10 nm), respectively. The mitochondrial transmembrane potential (Δϕm) was estimated using the fluorescence quenching of the cationic indicator, 150 nmol/L tetramethyl rhodamine ethyl ester (TMRE), which is accumulated and quenched inside energized mitochondria. The excitation wavelength was 550 nm (slit 5 nm) and the emission wavelength 575 nm (slit 10 nm) for TMRE. The experiment was initiated by the sequential additions of mitochondria (0.1 mg/2.0 mL protein) to a 2.0 mL cuvette containing KCl-based respiration buffer (125 mmol/L KCl, 2 mmol/L MgCl2, 2.5 mmol/L KH2PO4, 0.1% BSA, 20 mmol/L HEPES at pH 7.2), followed by 100 nmol/L CaG5N and 150nmol/L TMRE. The reaction was initiated by the addition of the oxidizable substrates, such as 5 mmol/L pyruvate and 2.5 mmol/L malate at 1 min, followed by the addition of 150 μmol/L ADP at 2 mins and 1 μmol/L oligomycin (final concentrations) at 3 mins, and the reaction was continued upto 5 mins. At 5 mins after the final addition, 32 mmol/L calcium, at a rate of 0.5 μL/min (equivalent to 80 nmol of calcium per mg of mitochondrial protein per min) was continuously infused into the cuvette using a syringe pump.

Measurement of Oxidative Damage Markers by Slot-Blot Analysis

A portion of isolated ipsilateral cortical and ipsilateral hippocampal mitochondria from sham (3 h), 30 mins, 3, and 12 h after injury was used for the quantitative measurements of 4-hydroxynonenal (4-HNE) as an index of lipid peroxidation, protein carbonyls, an index of protein oxidation, and 3-nitrotyrosine (3-NT) as an index of protein nitration, using the slot-blot technique. Briefly, approximately 2.5 μg of mitochondrial protein was transferred to a Protran (0.2 μm) nitrocellulose membrane (Schleicher & Schuell, Dassel, Germany) by a Minifold II vacuum slot blot apparatus (Schleicher & Schuell). Each slot was washed with 300 μL PBS and the membranes were dried. For the detection of protein tyrosine nitration, a rabbit polyclonal antityrosine antibody (Upstate Biotechnology, Milford, MA, USA) was used at a dilution of 1:2000 with overnight incubation at 4°C. For detection of HNE, a rabbit polyclonal anti-HNE antibody (Alpha Diagnostics International) was used at a dilution of 1:5000 with overnight incubation at 4°C. The levels of protein carbonyl were determined by using the OxyBlot Protein Oxidation Detection Kit (OxyBlot, Chemicon, Temecula, CA, USA) with rabbit polyclonal anti-dinitrophenyl hydrazine (DNP), with overnight incubation at 4°C. Positive bands detected with a peroxidase-coupled secondary IR-Dye 800CW goat anti-rabbit was used at a dilution of 1:500 (Vector Laboratories, CA, USA). All incubations and washing steps were performed according to the instructions given by the manufacturers. The intensity of the bands was visualized and quantitated using Li-Cor Odyssey Infrared Imager. On all blots, a blank without mitochondrial protein was employed to correct for nonspecific binding. The value of the blank was background subtracted from the values for all the other samples. In the case of 3-NT, we showed that preincubation with nitrated BSA blocked the reactivity of the antibody with the sample down to negligible levels.

Electron Microscopy

The purified mitochondrial pellets were fixed in 4% glutaraldehyde overnight at 4°C before being embedded for electron microscopy. The fixed mitochondrial pellets were then washed overnight at 4°C in 0.1 mol/L sodium cacodylate buffer, followed by 1 h secondary fixation at room temperature in 1% osmium tetroxide (in sodium cacodylate buffer). Then, the mitochondrial pellets were rinsed with distilled water and dehydrated in a gradient ethanol series up to 100% and twice in propylene oxide. The pellets were then placed in a 1:1 mixture of propylene oxide and Epon/Araldite resin and infiltrated overnight on a rotator. Next, 100% Epon/Araldite resin was added and rotated for 1 h at room temperature. Finally, fresh resin was prepared and degassed using a vacuum chamber. The mitochondrial pellets were added to the flat molds, filled with fresh resin, and baked overnight at 60°C. The 90 nm ultra-thin sections were cut using an RMC MT-7000 ultramicrotome mounted on 150 mesh copper grids and stained with uranyl acetate and lead citrate. A Zeiss 902 electron microscope was used to examine and photograph the samples.

Statistical Analysis

Results were expressed as the mean ± s.e.m. For single-variable comparisons, Student's t-test was used. For multiple-variable comparisons, data were analyzed by one-way analysis of variance (ANOVA) followed by post-hoc analysis using the Student-Neuman-Keul's (SNK) test.

Results

Total (synaptic and nonsynaptic population) ipsilateral cortical and hippocampal mitochondria were isolated using the ‘nitrogen bomb’ decompression method from sham (craniotomy, but no injury) and injured animals at 30 mins, 1, 3, 6, 12, 24, 48, and 72 h after injury. At all time-points, mitochondria isolated from either the cortices or hippocampi of sham animals were metabolically intact and wellcoupled, as shown by the RCR being > 6 (Figures 1 and 3). This indicates that in mitochondria isolated from sham animals at all time-points the electron transport system (ETS) was well coupled to oxidative phosphorylation and that any differences observed in the mitochondria from the injured animals were not due to preparation artifact.
Figure 1 Time course of mitochondrial respiratory dysfunction (A) and States III & IV respiratory rates (B) in mouse ipsilateral cortex after severe (1.0 mm) TBI. Mitochondrial oxygen consumption was measured using a Clark-type electrode in a continuously stirred, sealed chamber (Oxygraph System; Hansatech Instruments Ltd). Purified mitochondrial protein (25 to 35 μg) was suspended in respiration buffer (125 mmol/L KCl, 2 mmol/L MgCl2, 2.5 mmol/L KH2PO4, 0.1% BSA, 20 mmol/L HEPES, pH 7.2) in a final volume of 250 μL. Respiratory control ratio was calculated as the ratio of oxygen consumption in the presence of 5 mmol/L pyruvate and 2.5 mmol/L malate before (state IV) and after the second addition of l50 μmol/L ADP (state III). Data are presented as mean ± s.e.m. Statistical differences (one-way ANOVA and Student-Neuman-Keuls post hoc test): *P < 0.0001 versus sham; *P < 0.001 versus sham.

Time Course of Cortical and Hippocampal Mitochondrial Respiratory Dysfunction After Traumatic Brain Injury

Mitochondrial bioenergetics were assessed in animals after a severe TBI at 30 mins, 1, 3, 6, 12, 24, 48, and 72 h after injury. This revealed significant differences in the RCR that were due to changes in state III and IV respiratory rates. A significant decrease in RCR was observed in ipsilateral cortical mitochondria as early as 30 mins after injury, followed by a recovery to baseline at 1 h (Figure 1A). This significant decrease in RCR at 30 mins after injury was due to a significant increase in state IV respiration, indicative of increased proton conductance across the mitochondrial inner membrane (Figure 1B), which could in part be due to the onset of mitochondrial permeability transition (mPT). However, at 1 h, states III, IV and the RCR returned back to baseline. At 3 h, a secondary significant decrease (-70% compared with sham) in RCR was noted, which was followed by a partial recovery (Figure 1A). State IV respiration was increased by 233%. At 6h, state III respiration was seen to decrease significantly (-30%). This decrease in maximal respiratory capacity progressed down to −63% at 12 h. Although the RCR partially recovered toward the sham baseline, it remained significantly suppressed in comparison to that seen in mitochondria from the sham animals up to 72 h after injury. Similarly, state IV respiration remained significantly elevated. However, state III respiration appeared to recover completely. Figure 2 shows the contrast between the typical oxymetric respiratory traces from cortical mitochondria isolated from a sham versus 3 h postinjury brain.
Figure 2 A typical oxymetric trace showing alterations in cortical mitochondrial bioenergetics after a severe TBI. Mitochondria isolated 3 h after injury from the injured cortex of mice show a reduction in RCR (rate of respiration in the presence of ADP versus rate of respiration in the absence of ADP) and a loss of ETS capacity (rate of respiration in the presence of the uncoupler FCCP), compared with sham-operated animals.
Figure 3 Time course of mitochondrial respiratory dysfunction (A) and States III & IV respiratory rates (B) in mouse ipsilateral hippocampus after severe (1.0 mm) TBI. Mitochondrial protein (25 to 35 μg) was suspended in respiration buffer, in a final volume of 250 μL, and oxygen consumption assessed under various experimental conditions. Respiratory control ratio was calculated as the ratio of oxygen consumption in the presence of ADP (second addition of 150 μmol/L) or oligomycin (state IV), while using the oxidative substrates, pyruvate and malate. Data are presented as mean ± s.e.m. Statistical differences (one-way ANOVA and SNK post hoc test): *P < 0.0001 versus sham; *P < 0.001 versus sham.
Results similar to those seen in cortical mitochondria were obtained with mitochondria isolated from ipsilateral hippocampus with regard to changes in the RCR (Figure 3A) and state III and IV respiration rates (Figure 3B) as a function of time after injury. The one difference observed was a significant reduction in state III respiration at 72 h after injury in the hippocampal mitochondria. This indicates that there may be a tertiary phase of mitochondrial dysfunction between 24 and 72 h after injury. Although a similar decrease in state III respiration at this time-point was measured in the cortical mitochondria, it did not reach statistical significance.

Time Course of Cortical Mitochondrial Calcium Buffering Impairment After Traumatic Brain Injury

The ability of the cortical mitochondria to buffer Ca2+ before undergoing mPT was compared between sham and injured animals at 30 mins, 3 and 12 h (Figure 4) after injury. Calcium uptake was not significantly impaired at 30 mins after injury. However, by 3h, a > 50% decrease (P < 0.05) in calcium buffering was observed, which persisted at 12 h after injury. The inset in Figure 4 shows differences in mitochondrial calcium uptake (i.e. calcium-buffering capacity) using the Ca2+ -sensitive indicator, CaG5N, during a sustained calcium exposure, in injured versus noninjured mitochondria, indicating a reduced threshold for the onset of the mPT.
Figure 4 Post-traumatic time course of changes in cortical mitochondrial calcium buffering. Mitochondrial calcium buffering was assessed in a stirred cuvette mounted in a Shimadzu RF-5301 spectrofluorimeter maintained at 37°C. The extra-mitochondrial concentrations of Ca2+ were measured using the fluorescent Ca2+ -sensitive indicator CaG5N. Mitochondria (0.1mg/mL) were incubated with 5 mmol/L pyruvate and 2.5 mmol/L malate, 150 μmol/L ADP and 1 μmol/L oligomycin. At 5 mins after the final addition (oligomycin), 32 mmol/L calcium, at a rate of 0.5 μL/min, was infused continuously into the cuvette using a syringe pump. The inset is a typical trace that shows the differences in mitochondrial calcium uptake (i.e. calcium-buffering capacity), in injured (3 h after injury) versus noninjured (sham) mitochondria, indicating a reduced threshold for mpT. Bars are the mean ± s.e.m. Statistical differences (oneway ANOVA and SNK post hoc test): *P < 0.05 versus sham.

Time Course of Mitochondrial Oxidative Injury After Traumatic Brain Injury

Based on the time course of mitochondrial dysfunction, markers of oxidative damage were analyzed at the 30-min, 3-, and 12-h postinjury time-points for ipsilateral cortical (Figure 5A) and hippocampal (Figure 5B) mitochondria. In the cortex, an increase in lipid peroxidation products (HNE; P < 0.05) and protein nitration (3-NT; P < 0.05) was seen as early as 30 mins after injury. Protein carbonyl levels were significantly (P < 0.001) elevated at 3 h after injury compared with sham animals. At 12 h, only the HNE levels remained significantly increased, whereas the protein carbonyl and 3-NT had returned to within normal limits (i.e. in comparison to sham animals).
Figure 5 Time course of lipid (HNE) and protein (protein carbonyl) oxidative damage in percoll-purified mitochondria from the ipsilateral cortex (A) and ipsilateral hippocampus (B) after a severe (1.0 mm) TBI. Distribution of HNE, protein carbonyls, and 3-NT in cortical mitochondria (A), and hippocampal mitochondria (B), of mice at 30 mins, 3 and 12 h after injury and were compared with sham (3h) animals. Approximately 2.5 μg of percoll-purified mitochondrial protein isolated from cortical and hippocampal region was applied onto Protran nitrocellulose membrane (Schleicher & Schuell, Dassel, Germany), as described in the experimental section, and oxidative markers were quantitated using a Minifold II vacuum slot blot apparatus. Data are presented as mean ± s.e.m. Statistical differences (one-way ANOVA and SNK post hoc test): *P < 0.05 versus sham; **P < 0.001 versus sham; ***P < 0.0001 versus sham; @P < 0.005 versus sham.
Ipsilateral hippocampal mitochondria displayed a more delayed increase in oxidative damage, with HNE and protein carbonyl not being significantly increased until 3 h after injury. At 3 h, the HNE and 3-NT remained elevated along with the appearance of a significant (P < 0.0001) increase in protein oxidation-derived carbonyl levels. By 12 h, these had returned to normal sham levels, but a delayed increase in 3-NT (P < 0.005) was shown.

Time Course of Changes in Cortical Mitochondria After Traumatic Brain Injury

Several lines of evidence in the current study implicated the onset of the catastrophic mPT after TBI. To directly assess this question, the ultrastructural integrity of cortical mitochondria isolated from sham animals was qualitatively compared with those harvested from injured brains at 30 mins, 3 and 12 h after injury using transmission electron microscopy. The photomicrographs indicate that the cortical mitochondria isolated from sham animals displayed typical tight cristae and intact outer membranes (Figure 6A). In contrast, electron microscopy of the ipsilateral cortical mitochondria from TBI animals revealed obvious structural damage as early as 30 mins (Figure 6B), which appeared to get progressively worse at 3 h (Figure 6C) and 12 h (Figure 6D) after injury. The time-dependent damage consisted of mitochondrial swelling, disruption or loss of cristae, and breakage of the outer membranes indicative of osmotic swelling due to mPT. By 12 h, few, if any, normal-looking mitochondria were visible in any given sample of ipsilateral cortical mitochondria.
Figure 6 Electron microscopic pictures of percoll-purified mitochondria from ipsilateral cortex after severe postinjury reveal mitochondrial matrix condensation, cristae disruption and swelling. Percoll-purified mitochondrial pellets from sham (3 h), (A), 30 mins, (B), 3h (C) and 12 h (D) postinjury were fixed in 4% glutaraldehyde at 4°C before being embedded for electron microscopy. The scale bar represents 500 nm.

Discussion

This study describes the time course of structural and functional alterations in isolated cortical and hippocampal mitochondria during the first 72 h after a severe level of CCI-TBI in mice. Mitochondrial dysfunction was observed as early as 30 mins after injury, as evidenced by a decrease in RCR due to a significant increase in state IV respiration in both brain regions ipsilateral to the injury. The post-traumatic mitochondrial dysfunction occurred simultaneously with an increase in markers of mitochondrial oxidative damage, including lipid peroxidation (increased HNE), protein oxidation (increased protein carbonyl), and protein nitration (3-NT). This is very intriguing considering that oxidative stress is a known inducer of mPT susceptibility (Sullivan et al, 2004a, b, 2005). Oxidative stress is also known to significantly impair mitochondrial respiration specifically by altering complex-I, nicotinamide adenine dinucleotide (reduced form)-driven respiration (Sullivan et al, 2004a, c, 2005).

Role of Peroxynitrite in Traumatic Brain Injury Pathophysiology

Fifteen years ago, Beckman and coworkers introduced the theory that the principal ROS involved in producing tissue injury in a variety of neurological disorders is ONOO, which is formed by the combination of nitric oxide synthase (NOS) ONOO-generated NO radical and superoxide radical (Beckman, 1991). Since that time, the biochemistry of ONOO, which is often referred to as a reactive nitrogen species, has been clarified. ONOO-mediated oxidative damage is actually caused by ONOO decomposition products that possess potent free radical characteristics. The first involves the protonation of ONOO to form peroxynitrous acid (ONOOH), which can undergo homolytic decomposition to form the highly reactive nitrogen dioxide radical (NO2) and hydroxyl radical (OH). Probably more important physiologically, ONOO will react with carbon dioxide (CO2) to form nitrosoperoxocarbonate (ONOOCO2), which can decompose into NO2 and carbonate radical (CO3). Each of the ONOO-derived radicals (OH, NO2, and CO3) can initiate LP cellular damage by abstraction of an electron from a hydrogen atom bound to an allylic carbon in polyunsaturated fatty acids or cause protein carbonylation by reaction with susceptible amino acids (e.g. lysine, cysteine, arginine). Both of these oxidative damage markers were increased in cortical and hippocampal mitochondria during the first 3 h after injury. However, the coincidental increase in protein nitration (nitrotyrosine) in the isolated mitochondria as early as 30 mins after injury signals that a major portion of the ROS-induced damage is probably caused by ONOO.
The implication of ONOO in post-traumatic mitochondrial dysfunction and neurodegeneration is derived from five lines of evidence. First of all, all three NOS isoforms (endothelial, neuronal and inducible) are known to be upregulated during the first 24 h after TBI in rodents. Secondly, it has been shown that the acute treatment of injured mice or rats with NOS inhibitors can exert a neuroprotective effect and/or improve neurological recovery (Mesenge et al, 1996b, 1998a, b; Wallis et al, 1996; Wada et al, 1998a, b, 1999). Thirdly, biochemical footprints of ONOO-mediated damage have been documented in rodent TBI paradigms, including an increase in 3-NT levels (Mesenge et al, 1998a, b). The notion that these markers of ONOO -mediated damage are pathophysiologically important is supported by the finding that the NOS inhibitor L-arginine analog N-nitro-L-arginine methyl ester can lessen the accumulation of 3-NT in injured brains (Mesenge et al, 1998a, b) at the same doses that improve neurological recovery (Mesenge et al, 1996a).
Fourth, recent evidence indicates that the principal oxidative damage-producing ROS that is being formed by mitochondria is most likely ONOO. For instance, nitric oxide has been shown to be present in mitochondria (Lopez-Figueroa et al, 2000; Zanella et al, 2002), and a mitochondrial NOS isoform (mtNOS) has been isolated, characterized and shown to be similar to neuronal NOS (nNOS), except for post-translational acylation with myristic acid and C-terminal phosphorylation (Kanai et al, 2001; Elfering et al, 2002). Although physiological roles for mtNOS have been suggested, mtNOS and NO appear to be important in neuronal degeneration. For instance, NO production in mitochondria occurs during apoptosis (Bustamante et al, 2002), and mtNOS stimulation can cause cytochrome c release (Ghafourifar et al, 1999). Deregulation of mitochondrial NO generation and the aberrant production of its toxic metabolite ONOO appear to be involved in many, if not all, of the major acute and chronic neurodegenerative conditions (Heales et al, 1999). Exposure of mitochondria to Ca2+, which is known to cause them to become dysfunctional, leads to impairment of mitochondrial oxidative phosphorylation (Xiong et al, 1997) and ONOO generation, which in turn triggers mitochondrial Ca2+ release (i.e. limits their Ca2+ uptake or buffering capacity) (Bringold et al, 2000).
The fifth and possibly most compelling evidence supporting a role of ONOO in acute TBI comes from recent studies showing the beneficial effects of multiple compounds that possess the ability to directly scavenge ONOO or ONOO-derived radicals in TBI models. For instance, penicillamine, which can stoichiometrically react with ONOO (Althaus et al, 2000), is able to improve neurological recovery of mice subjected to moderately severe diffuse TBI (Hall et al, 1999). Similarly, the brainpenetrable pyrrolopyrimidine antioxidant U-101033E, which can react with ONOO (Rohn et al, 1998), also improves the neurological recovery of injured mice (Hall et al, 1995). Furthermore, the indolamine melatonin, which is also reactive with ONOO, has been reported to be neuroprotective in mice after diffuse TBI (Mesenge et al, 1998a, b) and rats after cortical contusion injury (Cirak et al, 1999; Sarrafzadeh et al, 2000). However, most promising as an anti-ONOO agent is the antioxidant tempol, which has been shown to catalytically scavenge ONOO derived NO2 and CO3 (Carroll et al, 2000; Bonini et al, 2002). Tempol has been reported to reduce post-traumatic brain edema and improve neurological recovery in a rat contusion injury model (Beit-Yannai et al, 1996; Zhang et al, 1998).

Dissection of the Time Course of Oxidative Damage and Mitochondrial Functional Impairment

The decrease in mitochondrial RCR, which first occurred at 30 mins, indicates that some mitochondrial uncoupling has taken place. The increase in state IV represents an increase in proton leak across the inner mitochondrial membrane. Although state III respiration also increased at 30 mins, its magnitude was less than the increase in state IV. Consequently, the RCR (state III/state IV) was decreased.
At 1 h after injury, mitochondrial function appeared to recover with the RCR, state III and IV respiration all returning to preinjury levels. The explanation for this early biphasic functional change could be twofold. The first possibility is that a fraction of the mitochondria in the injured cortical and hippocampal samples are undergoing mPT and thus irreversible function loss as early as 30 mins. Once they undergo mPT, they are beyond the point of salvage. The photomicrographs (Figure 6) showing that a portion of the mitochondria examined at 30 mins is severely swollen is consistent with the notion that those particular mitochondria are probably past the point of functional return. Consequently, the apparent recovery at 1 h may simply be that the initially damaged mitochondria seen at 30 mins are no longer part of the 1 h percoll isolation.
A second possibility is that an initial post-traumatic increase in intracellular and mitochondrial calcium leads to an early transient burst in ROS, which produces a reversible degree of oxidative damage and functional impairment. Indeed, a rapid increase in OH production in the injured mouse (Hall et al, 1993) and rat (Smith et al, 1994) brain has been shown after severe TBI. However, in both cases, the increase in OH ablated by 1 h. Some of the current findings support this explanation for the initial transient decrease in the RCR. Although an increase in mitochondrial HNE and 3-NT is seen in the 30-min cortical sample indicative of ONOO induced damage, the ability of the 30-min mitochondrial sample to buffer calcium is no different from that seen in sham, noninjured, brains. This suggests that most of the mitochondria in the 30-min postinjury sample may still be viable, although partially uncoupled and thus functionally impaired. The subsequent recovery in RCR, state III and state IV respiration may then signal that many mitochondria in the population have not actually undergone mPT, and are therefore able to recover from the initial wave of ROS production and oxidative damage.
At 3 h after injury, a secondary, more severe, wave of functional depression was seen, which was characterized by a 70% decrease in RCR (compared with sham) in both the cortical and the hippocampal mitochondria. As explained above concerning the initial 30-min decline in RCR, this secondary suppression in RCR is in part due to a significant increase in state IV respiration, which could imply that mPT was occurring in some of the isolated mitochondria, decreasing the overall RCR for the sample. Indeed, state IV was increased by 233% in the cortical mitochondria and 185% in the hippocampal sample. Maximal calcium-buffering capacity, examined in cortical mitochondria, was reduced by 53% at 3 h. This mitochondrial functional failure was paralleled by severe ultrastructural damage to the cortical mitochondria at 3 h, including swelling, dilated cristae and broken outer membranes, which are all indicative of the onset of mPT and subsequent osmotic swelling. Many fewer normal-looking mitochondria were observed in comparison to either the sham or the 30-min postinjury samples. Oxidative damage was observed in the mitochondria from both regions at 3h: increased protein carbonyl and 3-NT in the cortical sample and increased HNE and protein carbonyl in the hippocampal sample.
The increase in state IV respiration, indicative of an increased inner membrane proton conductance, was also coupled with a progressive decrease in state III respiration (oxidative phosphorylation) in the injured cortex and hippocampus at 6 and 12 h after injury. This decrease in state III respiration signals an impairment of electron transport chain and/or enzymes of the citric acid cycle. In addition, mitochondrial calcium uptake (i.e. calcium-buffering capacity) continued to be severely compromised at 12 h after injury in the injured cortex. Some evidence of mitochondrial oxidative damage persisted at the 12 h time-point (increase in HNE in cortical mitochondria, increase in 3-NT in hippocampal mitochondria). However, the overall time course of oxidative damage suggests that the peak was probably at 3 h. Ultrastructural examination of a cortical sample at 12 h revealed very few normal-looking mitochondria.
At 24 h after injury, mitochondrial oxidative phosphorylation (state III respiration) appeared to recover to sham levels in both brain regions. However, at the same time, proton conductance (state IV respiration) remained significantly increased, which resulted in a sustained reduction in RCR. The most logical explanation for this apparent recovery, in part based on the mitochondrial ultrastructural analysis, would be that after 12 h the majority of damaged mitochondria would not be present at 24 h after injury. The reason for this is that the Percoll isolation itself selects for relatively uninjured mitochondria based on density. Additionally, since the preparation contains mitochondria that are synaptic as well as nonsynaptic in origin, mitochondria from inflammatory cells that entered the CNS after TBI could also account for part of this apparent recovery (Lifshitz et al, 2003, 2004; Sullivan and Brown, 2005; Sullivan, 2005). Nevertheless, even at these later time-points a significant tertiary loss of mitochondrial bioenergetics was apparent and represented by a sustained decrease in RCR and increase in proton conductance during state IV respiration, indicating an ongoing disruption of mitochondrial functional integrity.
Ongoing studies in our laboratories are evaluating the relative efficacy of ONOO-directed pharmacological agents in preserving mitochondrial function against ONOO in both in vitro and in vivo TBI models. However, the current time-course analysis suggests that if ONOO is truly a major underlying cause of mitochondrial dysfunction, the therapeutic efficacy window may only be 3 h in length. Later treatment initiation might be able to protect against the late-phase damage and dysfunction that becomes manifest between 24 and 72 h, but would miss the initial major component that appears to occur during the first several hours. Ideally, an attempt to pharmacologically protect mitochondria from ONOO -induced oxidative damage would be initiated within 3 h and then maintained for at least 48 to 72 h to optimally reduce mitochondrial dysfunction throughout its full time course and to exert the best neuroprotective effect.

Acknowledgements

The authors thank Tonya Gibson Hurst, Kristen Day and Brian Thompson for expert technical assistance.

References

Althaus JS, Schmidt KR, Fountain ST, Tseng MT, Carroll RT, Galatsis P, Hall ED (2000) LC-MS/MS detection of peroxynitrite-derived 3-nitrotyrosine in rat microvessels. Free Radic Biol Med 29:1085–95.
Azbill RD, Mu X, Bruce-Keller AJ, Mattson MO, Springer JE (1997) Impaired mitochondrial function, oxidative stress and altered antioxidant enzyme activities following traumatic spinal cord injury. Brain Res 765:283–90.
Brown MR, Sullivan PG, Dorenbos KA, Modafferi EA, Geddes JW, Steward O (2004) Nitrogen disruption of synaptoneurosomes: an alternative method to isolate brain mitochondria. J Neurosci Meth 137:299–303.
Beckman JS (1991) The double-edged role of nitric oxide in brain function and superoxide-mediated injury. J Dev Physiol 15:53–9.
Beit-Yannai E, Zhang R, Trembovler V, Samuni A, Shohami E (1996) Cerebroprotective effect of stable nitroxide radicals in closed head injury in the rat. Brain Res 717:22–8.
Bonini MG, Mason RP, Augusto O (2002) The mechanism by which 4-hydroxy-2,2,6,6-tetramethylpiperidene-1-oxyl (tempol) diverts peroxynitrite decomposition from nitrating to nitrosating species. PG-506–11. Chem Res Toxicol 15:506–11.
Bringold U, Ghafourifar P, Richter C (2000) Peroxynitrite formed by mitochondrial NO synthase promotes mitochondrial Ca2+ release. Free Radic Biol Med 29:343–8.
Brustovetsky N, Brustovetsky T, Jemmerson R, Dubinsky JM (2002) Calcium–induced cytochrome c release from CNS mitochondria is associated with the permeability transition and rupture of the outer membrane. J Neurochem 80:207–18.
Bustamante J, Bersier G, Badin RA, Cymeryng C, Parodi A, Boveris A (2002) Sequential NO production by mitochondria and endoplasmic reticulum during induced apoptosis. Nitric Oxide 6:333–41.
Carroll RT, Galatsis P, Borosky S, Kopec KK, Kumar V, Althaus JS, Hall ED (2000) 4-hydroxy-2,2,6,6-tetramethylpiperidine-1-oxyl (tempol) inhibits peroxynitrite-mediated phenol nitration. Chem Res Toxicol 13:294–300.
Cirak B, Rousan N, Kocak A, Palaoglu S, Kilic K (1999) Melatonin as a free radical scavenger in experimental head trauma. Pediatr Neurosurg 31:298–301.
Elfering SL, Sarkela TM, Giulivi C (2002) Biochemistry of mitochondrial nitric oxide synthase. J Biol Chem 277:38079–86.
Finkel E (2001) The mitochondrion: is it central to apoptosis? Science 292:624–6.
Ghafourifar P, Schenk U, Klein SD, Richter C (1999) Mitochondrial nitric-oxide synthase stimulation causes cytochrome c release from isolated mitochondria. Evidence for intramitochondrial peroxynitrite formation. J Biol Chem 274:31185–8.
Green DR, Reed JC (1998) Mitochondria and apoptosis. Science 281:1309–12.
Hall ED, Andrus PK, Smith SI, Oostveen JA, Scherch HM, Lutzke BS, Raub TJ, Sawada GA, Palmer JR, Banitt LS, Tustin JM, Belonga KL, Ayer DE, Bundy GL (1995) Neuroprotective efficacy of microvascularly-localized versus brain-penetrating antioxidants. Acta Neurochir (Suppl) 66.
Hall ED, Andrus PK, Yonkers PA (1993) Brain hydroxyl radical generation in acute experimental head injury. J Neurochem 60:588–94.
Hall ED, Kupina NC, Althaus JS (1999) Peroxynitrite scavengers for the acute treatment of traumatic brain injury. Ann N Y Acad Sci 890:492–8.
Hall ED, Sullivan PG, Gibson TR, Pavel KM, Thompson BM, Scheff SW (2005) Spatial and temporal characteristics of neurodegeneration after controlled cortical impact in mice: more than a focal brain injury. J Neurotrauma 22:252–65.
Hannay HJ, Feldman Z, Phan P, Keyani A, Panwar N, Goodman JC, Robertson CS (1999) Validation of a controlled cortical impact model of head injury in mice. J Neurotrauma 16:1103–14.
Heales SJ, Bolanos JP, Stewart VC, Brookes PS, Land JM, Clark JB (1999) Nitric oxide, mitochondria and neurological disease. Biochim Biophys Acta 1410:215–28.
Hunot S, Flavell RA (2001) Apoptosis: death of a monopoly? Science 292:865–6.
Jiang D, Sullivan PG, Sensi SL, Steward O, Weiss JH (2001) Zn2+ induces permeability transition pore opening and release of pro-apoptotic peptides from neuronal mitochondria. J Biol Chem 276:47524–9.
Kanai AJ, Pearce LL, Clemens PR, Birder LA, VanBibber MM, Choi SY, De Groat WC, Peterson J (2001) Identification of a neuronal nitric oxide synthase in isolated cardiac mitochondria using electrochemical detection. Proc Natl Acad Sci USA 98:14126–31.
Kristian T, Hopkins IB, McKenna MC, Fiskum G (2005) Isolation of mitochondria with high respiratory control from primary cultures of neurons and astrocytes using nitrogen cavitation. J Neurosci Meth (in press).
Kroemer G, Reed JC (2000) Mitochondrial control of cell death. Nat Med 6:513–9.
Lifshitz J, Friberg H, Neumar RW, Raghupathi R, Welsh FA, Janmey P, Saatman KE, Wieloch T, Grady MS, McIntosh TK (2003) Structural and functional damage sustained by mitochondria after traumatic brain injury in the rat: evidence for differentially sensitive populations in the cortex and hippocampus. J Cereb Blood Flow Metab 23:219–31.
Lifshitz J, Sullivan PG, Hovda DA, Wieloch T, McIntosh TK (2004) Mitochondrial damage and dysfunction in traumatic brain injury. Mitochondrion 4:705–13.
Lopez-Figueroa MO, Caamano C, Morano MI, Ronn LC, Akil H, Watson SJ (2000) Direct evidence of nitric oxide presence within mitochondria. Biochem Biophys Res Commun 272:129–33.
Matsushita M, Xiong G (1997) Projections from the cervical enlargement to the cerebellar nuclei in the rat, studied by anterograde axonal tracing. J Comp Neurol 377:251–61.
Mesenge C, Charriaut-Marlangue C, Verrecchia C, Allix M, Boulu RR, Plotkine M (1998a) Reduction of tyrosine nitration after N(omega)-nitro-L-arginine-methylester treatment of mice with traumatic brain injury. Eur J Pharmacol 353:53–7.
Mesenge C, Margaill I, Verrecchia C, Allix M, Boulu RR, Plotkine M (1998b) Protective effect of melatonin in a model of traumatic brain injury in mice. J Pineal Res 25:41–6.
Mesenge C, Verrecchia C, Allix M, Boulu RR, Plotkine M (1996a) Reduction of the neurological deficit in mice with traumatic brain injury by nitric oxide synthase inhibitors. J Neurotrauma 13:209–14.
Mesenge C, Verrecchia C, Allix M, Boulu RR, Plotkine M (1996b) Reduction of the neurological deficit in mice with traumatic brain injury by nitric oxide synthase inhibitors. J Neurotrauma 13:11–6.
Nicholls DG, Budd SL (1998) Mitochondria and neuronal glutamate excitotoxicity. Biochim Biophys Acta 1366:97–112.
Okonkwo DO, Buki A, Siman R, Povlishock JT (1999) Cyclosporin A limits calcium-induced axonal damage following traumatic brain injury. NeuroReport 10:353–8.
Okonkwo DO, Povlishock JT (1999) An intrathecal bolus of cyclosporine A before injury preserves mitochondrial integrity and attenuates axonal disruption in traumatic brain injury. J Cereb Blood Flow Metab 19:443–51.
Raghupathi R, Fernandez SC, Murai H, Trusko SP, Scott RW, Nishioka WK, McIntosh TK (1998) BCL-2 overexpression attaenuates cortical cell loss after traumatic brain injury in transgenic mice. J Cereb Blood Flow Metab 18:1259–69.
Robertson CL (2004) Mitochondrial dysfunction contributes to cell death following traumatic brain injury in adult and immature animals. J Bioenerg Biomembr 36:363–8.
Rohn TT, Nelson LK, Waeg G, Quinn MT (1998) U-101033E (2,4-diaminopyrrolopyrimidine), a potent inhibitor of membrane lipid proxidation as assessed by the production of 4-hydroxynonenal, malondialdehyde, and 4-hydroxynonenal-protein adducts PG-1371–9. Biochem Pharmacol 56:1371–9.
Sarrafzadeh AS, Thomale UW, Kroppenstedt SN, Unterberg AW (2000) Neuroprotective effect of melatonin on cortical impact injury in the rat. Acta Neurochir (Wein) 142:1293–9.
Scheff SW, Sullivan PG (1999) Cyclosporin A significantly ameliorates cortical damage following experimental traumatic brain injury in rodents. J Neurotrauma 16:783–92.
Sensi SL, Ton-That D, Sullivan PG, Jonas EA, Gee KR, Kaczmarek LK, Weiss JH (2003) Modulation of mitochondrial function by endogenous Zn2+ pools. Proc Natl Acad Sci USA 100:6157–62.
Smith DH, Soares HD, Pierce JS, Perlman KG, Saatman KE, Meaney DF, Dixon CE, McIntosh TK (1995) A model of parasagittal controlled cortical impact in the mouse: cognitive and histopathologic effects. J Neurotrauma 12:169–78.
Smith SL, Andrus PK, Zhang JR, Hall ED (1994) Direct measurement of hydroxyl radicals, lipid peroxidation, and blood–brain barrier disruption following unilateral cortical impact head injury in the rat. J Neurotrauma 11:393–404.
Spierings D, McStay G, Saleh M, Bender C, Chipuk J, Maurer U, Green DR (2005) Connected to death: the (unexpurgated) mitochondrial pathway of apoptosis. Science 310:66–7.
Stout AK, Raphael HM, Kanterewicz BI, Klann E, Reynolds IJ (1998) Glutamate-induced neuron death requires mitochondrial calcium uptake. Nat Neurosci 1:366–73.
Sullivan PG (2005) Interventions with neuroprotective agents: novel targets and opportunities. Epilepsy Behav 7:S12–7.
Sullivan PG, Brown MR (2005) Mitochondrial aging and dysfunction in Alzheimer's disease. Prog Neuropsychopharmacol Biol Psychiatry 29:407–10.
Sullivan PG, Bruce-Keller AJ, Rabchevsky AG, Christakos S, Clair DK, Mattson MP, Scheff SW (1999a) Exacerbation of damage and altered NF-kappaB activation in mice lacking tumor necrosis factor receptors after traumatic brain injury. J Neurosci 19:6248–56.
Sullivan PG, Dube C, Dorenbos K, Steward O, Baram TZ (2003) Mitochondrial uncoupling protein-2 protects the immature brain from excitotoxic neuronal death. Ann Neurol 53:711–7.
Sullivan PG, Geiger JD, Mattson MP, Scheff SW (2000a) Dietary supplement creatine protects against traumatic brain injury. Ann Neurol 48:723–9.
Sullivan PG, Keller JN, Mattson MP, Scheff SW (1999b) Traumatic brain injury alters synaptic homeostasis: implications for impaired mitochondrial and transport function. J Neurotrauma 15:789–98.
Sullivan PG, Rabchevsky AG, Keller JN, Lovell M, Sodhi A, Hart RP, Scheff SW (2004a) Intrinsic differences in brain and spinal cord mitochondria: implication for therapeutic interventions. J Comp Neurol 474:524–34.
Sullivan PG, Rabchevsky AG, Waldmeier PC, Springer JE (2005) Mitochondrial permeability transition in CNS trauma: cause or effect of neuronal cell death? J Neurosci Res 79:231–9.
Sullivan PG, Springer JE, Hall ED, Schiff SW (2004b) Mitochondrial uncoupling as a therapeutic target following neuronal injury. J Bioenerg Biomembr 36:353–6.
Sullivan PG, Thompson M, Schiff SW (2000c) Continuous infusion of cyclosporin A postinjury significantly ameliorates cortical damage following traumatic brain injury. Exp Neurol 161:631–7.
Sullivan PG, Thompson M, Scheff SW (2000b) Continuous infusion of cyclosporin A significantly ameliorates cortical damage following traumatic brain injury. Exp Neurol 161:631–7.
Sullivan PG, Thompson MB, Scheff SW (1999 c) Cyclosporin A attenuates acute mitochondrial dysfunction following traumatic brain injury. Exp Neurol 160:226–30.
Wada K, Alonso OF, Busto R, Panetta J, Clemens JA, Ginsberg MD, Dietrich WD (1999) Early treatment with a novel inhibitor of lipid peroxidation (ly341122) improves histopathological outcome after moderate fluid percussion brain injury in rats. Neurosurgery 45:601–8.
Wada K, Chatzipanteli K, Busto R, Dietrich WD (1998a) Role of nitric oxide in traumatic brain injury in the rat. J Neurosurg 89:807–18.
Wada K, Chatzipanteli K, Kraydieh S, Busto R, Dietrich WD (1998b) Inducible nitric oxide synthase expression after traumatic brain injury and neuroprotection with aminoguanidine treatment in rats. Neurosugery 43:1427–36.
Wallis RA, Panizzon KL, Girard JM (1996) Traumatic neuroprotection with inhibitors of nitric oxide and ADP ribosylation. Brain Res 710:169–77.
White RJ, Reynolds IJ (1996) Mitochondrial depolarization in glutamate-stimulated neurons: an early signal specific to excitotoxin exposure. J Neurosci 16(18):5688–97.
Xiong Y, Gu Q, Peterson PL, Muizelaar JP, Lee CP (1997) Mitochondrial dysfunction and calcium perturbation induced by traumatic brain injury. J Neurotrauma 14:23–34.
Zanella B, Calonghi N, Pagnotta E, Masotti L, Guarnieri C (2002) Mitochondrial nitric oxide localization in H9c2 cells revealed by confocal microscopy. Biochem Biophys Res Commun 290:1010–4.
Zhang R, Shohami E, Beit-Yannai E, Bass R, Trembovler V, Samuni A (1998) Mechanism of brain protection by nitric oxide radicals in experimental model of closed-head injury. PG–332–40. Free Radic Biol Med 24.

Cite article

Cite article

Cite article

OR

Download to reference manager

If you have citation software installed, you can download article citation data to the citation manager of your choice

Share options

Share

Share this article

Share with email
EMAIL ARTICLE LINK
Share on social media

Share access to this article

Sharing links are not relevant where the article is open access and not available if you do not have a subscription.

For more information view the Sage Journals article sharing page.

Information, rights and permissions

Information

Published In

Article first published: November 2006
Issue published: November 2006

Keywords

  1. mitochondria
  2. mitochondrial permeability transition
  3. oxidative damage
  4. traumatic brain injury

Rights and permissions

© 2006 ISCBFM.
Request permissions for this article.
PubMed: 16538231

Authors

Affiliations

Indrapal N Singh1
Spinal Cord & Brain Injury Research Center and Department of Anatomy & Neurobiology, University of Kentucky Medical Center, Lexington, Kentucky, USA
Patrick G Sullivan1
Spinal Cord & Brain Injury Research Center and Department of Anatomy & Neurobiology, University of Kentucky Medical Center, Lexington, Kentucky, USA
Ying Deng
Spinal Cord & Brain Injury Research Center and Department of Anatomy & Neurobiology, University of Kentucky Medical Center, Lexington, Kentucky, USA
Lamin H Mbye
Spinal Cord & Brain Injury Research Center and Department of Anatomy & Neurobiology, University of Kentucky Medical Center, Lexington, Kentucky, USA
Edward D Hall
Spinal Cord & Brain Injury Research Center and Department of Anatomy & Neurobiology, University of Kentucky Medical Center, Lexington, Kentucky, USA

Notes

Correspondence: Dr ED Hall, Spinal Cord & Brain Injury Research Center, University of Kentucky Medical Center, B383 BBSRB, 741 S. Limestone Street, Lexington, KY 40536-0509, USA. E-mail: [email protected]
1
These authors contributed equally to this work.
This research was supported by funding from the NIH (NS 1R01 NS46566 and NS048191) and the Kentucky Spinal Cord & Head Injury Research Trust (KSCHIRT).

Metrics and citations

Metrics

Journals metrics

This article was published in Journal of Cerebral Blood Flow & Metabolism.

VIEW ALL JOURNAL METRICS

Article usage*

Total views and downloads: 2311

*Article usage tracking started in December 2016


Altmetric

See the impact this article is making through the number of times it’s been read, and the Altmetric Score.
Learn more about the Altmetric Scores



Articles citing this one

Receive email alerts when this article is cited

Web of Science: 257 view articles Opens in new tab

Crossref: 270

  1. Traumatic Brain Injury: A Comprehensive Review of Biomechanics and Mol...
    Go to citation Crossref Google Scholar
  2. Stretch-Induced Injury Affects Cortical Neuronal Networks in a Time- a...
    Go to citation Crossref Google Scholar
  3. Traumatic Brain Injury Recovery with Photobiomodulation: Cellular Mech...
    Go to citation Crossref Google Scholar
  4. The role of mitochondrial uncoupling in the regulation of mitostasis a...
    Go to citation Crossref Google Scholar
  5. Accelerometer-based head impact detection in soccer - Where are we?
    Go to citation Crossref Google Scholar
  6. Tracing the path of disruption: 13 C isoto...
    Go to citation Crossref Google Scholar
  7. Increased Expression of KNa1.2 Channel by MAPK Pathway Regulates Neuro...
    Go to citation Crossref Google Scholar
  8. Immunometabolic mechanisms of HIV-associated neurocognitive disorders ...
    Go to citation Crossref Google Scholar
  9. Nano‐Plumber Reshapes Glymphatic‐Lymphatic System to Sustain Microenvi...
    Go to citation Crossref Google Scholar
  10. A Systematic Review on Traumatic Brain Injury Pathophysiology and Role...
    Go to citation Crossref Google Scholar
  11. Cellular senescence in brain aging and cognitive decline
    Go to citation Crossref Google Scholar
  12. Colloidal therapeutics in the management of traumatic brain injury: Po...
    Go to citation Crossref Google Scholar
  13. Massive Solubility Changes in Neuronal Proteins upon Simulated Traumat...
    Go to citation Crossref Google Scholar
  14. Fundamental Neurochemistry Review: Microglial immunometabolism in trau...
    Go to citation Crossref Google Scholar
  15. Characterizing Sex Differences in Mitochondrial Dysfunction After Seve...
    Go to citation Crossref Google Scholar
  16. Human umbilical cord-derived mesenchymal stem cells-harvested mitochon...
    Go to citation Crossref Google Scholar
  17. Recent Advances in Biomaterials‐Based Therapies for Alleviation and Re...
    Go to citation Crossref Google Scholar
  18. Protecting the injured central nervous system: Do anesthesia or hypoth...
    Go to citation Crossref Google Scholar
  19. Pioglitazone restores mitochondrial function but does not spare cortic...
    Go to citation Crossref Google Scholar
  20. Mesenchymal Stromal Cells-Derived Exosome and the Roles in the Treatme...
    Go to citation Crossref Google Scholar
  21. Sex-based differences of antioxidant enzyme nanoparticle effects follo...
    Go to citation Crossref Google Scholar
  22. Mitochondria as a therapeutic: a potential new frontier in driving the...
    Go to citation Crossref Google Scholar
  23. FAM3A Ameliorates Brain Impairment Induced by Hypoxia–Ischemia in Neon...
    Go to citation Crossref Google Scholar
  24. Mechanism of Hypocalcemia Caused by Traumatic Brain Injury
    Go to citation Crossref Google Scholar
  25. Biomechanik und Pathophysiologie
    Go to citation Crossref Google Scholar
  26. Mesenchymal Stem Cell Transplantation: Neuroprotection and Nerve Regen...
    Go to citation Crossref Google Scholar
  27. Physical memory of astrocytes
    Go to citation Crossref Google Scholar
  28. Pathogenesis and management of traumatic brain injury (TBI): role of n...
    Go to citation Crossref Google Scholar
  29. Nandrolone Supplementation Promotes AMPK Activation and Divergent 18[F...
    Go to citation Crossref Google Scholar
  30. Programming axonal mitochondrial maintenance and bioenergetics in neur...
    Go to citation Crossref Google Scholar
  31. Mitochondrial function in spinal cord injury and regeneration
    Go to citation Crossref Google Scholar
  32. Traumatic Brain Injury Models in Zebrafish (Danio rerio)
    Go to citation Crossref Google Scholar
  33. A role of Na+, K+ -ATPase in spatial memory deficits and inflammatory/...
    Go to citation Crossref Google Scholar
  34. Traumatic brain injury and translational research: pharmacological and...
    Go to citation Crossref Google Scholar
  35. Pharmacological Therapies for Concussions
    Go to citation Crossref Google Scholar
  36. Physical Memory of Astrocytes
    Go to citation Crossref Google Scholar
  37. Diacylglycerol Lipase-β Knockout Mice Display a Sex-Dependent Attenuat...
    Go to citation Crossref Google Scholar
  38. Early onset senescence and cognitive impairment in a murine model of r...
    Go to citation Crossref Google Scholar
  39. Rescuing mitochondria in traumatic brain injury and intracerebral hemo...
    Go to citation Crossref Google Scholar
  40. Traumatic Brain Injury: Mechanisms of Glial Response
    Go to citation Crossref Google Scholar
  41. Integrative Neuroinformatics for Precision Prognostication and Persona...
    Go to citation Crossref Google Scholar
  42. Time-Course Evaluation of Brain Regional Mitochondrial Bioenergetics i...
    Go to citation Crossref Google Scholar
  43. Inhibiting Mitochondrial Cytochrome c Oxidase Downregulates Gene Trans...
    Go to citation Crossref Google Scholar
  44. Epigenetic Blockade of Hippocampal SOD2 Via DNMT3b-Mediated DNA Methyl...
    Go to citation Crossref Google Scholar
  45. Mitochondrial drug delivery systems
    Go to citation Crossref Google Scholar
  46. Temporal changes in inflammatory mitochondria-enriched microRNAs follo...
    Go to citation Crossref Google Scholar
  47. Effects of Pyruvate Administration on Mitochondrial Enzymes, Neurologi...
    Go to citation Crossref Google Scholar
  48. Treatment of Sports-Related Concussion
    Go to citation Crossref Google Scholar
  49. Early Cell Response to Mechanical Stimuli during TBI
    Go to citation Crossref Google Scholar
  50. The role of mitochondrial bioenergetics and oxidative stress in depres...
    Go to citation Crossref Google Scholar
  51. Protective effects of phenelzine administration on synaptic and non-sy...
    Go to citation Crossref Google Scholar
  52. Mitochondria focused neurotherapeutics for spinal cord injury
    Go to citation Crossref Google Scholar
  53. The Mitochondria-Associated ER Membranes Are Novel Subcellular Locatio...
    Go to citation Crossref Google Scholar
  54. Alternative substrate metabolism depends on cerebral metabolic state f...
    Go to citation Crossref Google Scholar
  55. Mitochondrial biogenesis as a therapeutic target for traumatic and neu...
    Go to citation Crossref Google Scholar
  56. Formoterol, a β2-adrenoreceptor agonist, induces mitochondrial biogene...
    Go to citation Crossref Google Scholar
  57. MicroRNA-146a Is a Wide-Reaching Neuroinflammatory Regulator and Poten...
    Go to citation Crossref Google Scholar
  58. Pharmacological inhibition of lipid peroxidative damage by the 21-amin...
    Go to citation Crossref Google Scholar
  59. Mitochondrial Dysfunction in Traumatic Brain Injury: Management Strate...
    Go to citation Crossref Google Scholar
  60. Oxygen Exposure During Cardiopulmonary Resuscitation Is Associated Wit...
    Go to citation Crossref Google Scholar
  61. The Changes of Brain Edema and Neurological Outcome, and the Probable ...
    Go to citation Crossref Google Scholar
  62. Bioenergetic restoration and neuroprotection after therapeutic targeti...
    Go to citation Crossref Google Scholar
  63. Applying univariate vs. multivariate statistics to investigate therape...
    Go to citation Crossref Google Scholar
  64. Antioxidant Therapies in Traumatic Brain Injury
    Go to citation Crossref Google Scholar
  65. Dihydrolipoic Acid–Gold Nanoclusters Regulate Microglial Polarization ...
    Go to citation Crossref Google Scholar
  66. Metformin Promotes Axon Regeneration after Spinal Cord Injury through ...
    Go to citation Crossref Google Scholar
  67. Increased Intracranial Pressure in Critically Ill Cancer Patients
    Go to citation Crossref Google Scholar
  68. Zelluläre Trauma-Biomechanik: Verletzungen des zentralen Nervensystems
    Go to citation Crossref Google Scholar
  69. Prenatal exposure to oxcarbazepine increases hippocampal apoptosis in ...
    Go to citation Crossref Google Scholar
  70. Effect of Chronic Continuous Normobaric Hypoxia on Functional State of...
    Go to citation Crossref Google Scholar
  71. Using complex II as an emerging therapeutic target for the treatment o...
    Go to citation Crossref Google Scholar
  72. Current progress of mitochondrial transplantation that promotes neuron...
    Go to citation Crossref Google Scholar
  73. Fractionated mitochondrial magnetic separation for isolation of synapt...
    Go to citation Crossref Google Scholar
  74. Traumatic Brain Injuries: Pathophysiology and Potential Therapeutic Ta...
    Go to citation Crossref Google Scholar
  75. Brain Trauma Disrupts Hepatic Lipid Metabolism: Blame It on Fructose?
    Go to citation Crossref Google Scholar
  76. Management of Head Trauma in the Neurocritical Care Unit
    Go to citation Crossref Google Scholar
  77. Intracranial pressure monitoring: Gold standard and recent innovations
    Go to citation Crossref Google Scholar
  78. Making sense of gut feelings in the traumatic brain injury pathogenesi...
    Go to citation Crossref Google Scholar
  79. Comprehensive Profile of Acute Mitochondrial Dysfunction in a Preclini...
    Go to citation Crossref Google Scholar
  80. Detection of brain specific cardiolipins in plasma after experimental ...
    Go to citation Crossref Google Scholar
  81. A1 rather than A2A adenosine receptor as a possible target of Guanosin...
    Go to citation Crossref Google Scholar
  82. Effects of Phenelzine Administration on Mitochondrial Function, Calciu...
    Go to citation Crossref Google Scholar
  83. Acute Mitochondrial Impairment Underlies Prolonged Cellular Dysfunctio...
    Go to citation Crossref Google Scholar
  84. Rebamipide Mitigates Impairments in Mitochondrial Function and Bioener...
    Go to citation Crossref Google Scholar
  85. Sex-dependent and chronic alterations in behavior and mitochondrial fu...
    Go to citation Crossref Google Scholar
  86. Fourier-filtering based size-encoded images for label-free tracking of...
    Go to citation Crossref Google Scholar
  87. Guanosine protects against Ca2+-induced mitochondrial dysfunction in r...
    Go to citation Crossref Google Scholar
  88. Aiming for the target: Mitochondrial drug delivery in traumatic brain ...
    Go to citation Crossref Google Scholar
  89. Cellular Injury Biomechanics of Central Nervous System Trauma
    Go to citation Crossref Google Scholar
  90. Increased Intracranial Pressure in Critically Ill Cancer Patients
    Go to citation Crossref Google Scholar
  91. Neuroprotective Effects of Cyclosporine in a Porcine Pre-Clinical Tria...
    Go to citation Crossref Google Scholar
  92. Disentangling oxidation/hydrolysis reactions of brain mitochondrial ca...
    Go to citation Crossref Google Scholar
  93. A Mild Traumatic Brain Injury in Mice Produces Lasting Deficits in Bra...
    Go to citation Crossref Google Scholar
  94. Activation of cyclin D1 affects mitochondrial mass following traumatic...
    Go to citation Crossref Google Scholar
  95. Mitochondrial uncoupling prodrug improves tissue sparing, cognitive ou...
    Go to citation Crossref Google Scholar
  96. Thiamine preserves mitochondrial function in a rat model of traumatic ...
    Go to citation Crossref Google Scholar
  97. Mitochondrial therapy promotes regeneration of injured hippocampal neu...
    Go to citation Crossref Google Scholar
  98. Effects of Female Sex Steroids Administration on Pathophysiologic Mech...
    Go to citation Crossref Google Scholar
  99. Synaptic Mitochondria are More Susceptible to Traumatic Brain Injury-i...
    Go to citation Crossref Google Scholar
  100. Hypertonic sodium lactate reverses brain oxygenation and metabolism dy...
    Go to citation Crossref Google Scholar
  101. Continuous Infusion of Phenelzine, Cyclosporine A, or Their Combinatio...
    Go to citation Crossref Google Scholar
  102. Discovery of Lipidome Alterations Following Traumatic Brain Injury via...
    Go to citation Crossref Google Scholar
  103. NaHS restores mitochondrial function and inhibits autophagy by activat...
    Go to citation Crossref Google Scholar
  104. The Effects of Memantine on Glutamic Receptor–Associated Nitrosative S...
    Go to citation Crossref Google Scholar
  105. Mitochondrial Damage in Traumatic CNS Injury
    Go to citation Crossref Google Scholar
  106. Acute spinal cord injury: A review of pathophysiology and potential of...
    Go to citation Crossref Google Scholar
  107. Mechanisms of Endogenous Neuroprotective Effects of Astrocytes in Brai...
    Go to citation Crossref Google Scholar
  108. Time courses of post-injury mitochondrial oxidative damage and respira...
    Go to citation Crossref Google Scholar
  109. Targeting mitochondrial dysfunction in CNS injury using Methylene Blue...
    Go to citation Crossref Google Scholar
  110. Traumatic brain injury, diabetic neuropathy and altered-psychiatric he...
    Go to citation Crossref Google Scholar
  111. Repeated Mild Traumatic Brain Injury: Potential Mechanisms of Damage
    Go to citation Crossref Google ScholarPub Med
  112. Repeated mild traumatic brain injury in female rats increases lipid pe...
    Go to citation Crossref Google Scholar
  113. Prospects for therapeutic mitochondrial transplantation
    Go to citation Crossref Google Scholar
  114. Direct Neuronal Reprogramming: Achievements, Hurdles, and New Roads to...
    Go to citation Crossref Google Scholar
  115. Tempol (4 hydroxy-tempo) inhibits anoxia-induced progression of mitoch...
    Go to citation Crossref Google Scholar
  116. Regulation of Mitochondrial Function and Glutamatergic System Are the ...
    Go to citation Crossref Google Scholar
  117. Phenelzine Protects Brain Mitochondrial Function In Vitr...
    Go to citation Crossref Google Scholar
  118. Synaptic Mitochondria Sustain More Damage than Non-Synaptic Mitochondr...
    Go to citation Crossref Google Scholar
  119. Mitochondria and microRNA crosstalk in traumatic brain injury
    Go to citation Crossref Google Scholar
  120. Comprehensive Profiling of Modulation of Nitric Oxide Levels and Mitoc...
    Go to citation Crossref Google Scholar
  121. Role of Melatonin in Reducing Amphetamine-Induced Degeneration in Subs...
    Go to citation Crossref Google ScholarPub Med
  122. Carbonyl Scavenging as an Antioxidant Neuroprotective Strategy for Acu...
    Go to citation Crossref Google Scholar
  123. Oxidative Damage in Neurodegeneration and Injury☆
    Go to citation Crossref Google Scholar
  124. MicroRNAs as diagnostic markers and therapeutic targets for traumatic ...
    Go to citation Crossref Google Scholar
  125. Neuroproteomic study of nitrated proteins in moderate traumatic brain ...
    Go to citation Crossref Google Scholar
  126. Imaging mass spectrometry reveals loss of polyunsaturated cardiolipins...
    Go to citation Crossref Google Scholar
  127. Ketogenic diet decreases oxidative stress and improves mitochondrial r...
    Go to citation Crossref Google ScholarPub Med
  128. Erythropoietin and Its Derivates Modulate Mitochondrial Dysfunction af...
    Go to citation Crossref Google Scholar
  129. Treatment Perspectives Based on Our Current Understanding of Concussio...
    Go to citation Crossref Google Scholar
  130. Neonatal anoxia leads to time dependent progression of mitochondrial l...
    Go to citation Crossref Google Scholar
  131. The Impact of Previous Physical Training on Redox Signaling after Trau...
    Go to citation Crossref Google Scholar
  132. Thioredoxin-Mimetic-Peptides Protect Cognitive Function after Mild Tra...
    Go to citation Crossref Google Scholar
  133. Mitochondrial specific therapeutic targets following brain injury
    Go to citation Crossref Google Scholar
  134. Brain injury and recovery
    Go to citation Crossref Google Scholar
  135. Epigenetic changes following traumatic brain injury and their implicat...
    Go to citation Crossref Google Scholar
  136. Dietary fructose aggravates the pathobiology of traumatic brain injury...
    Go to citation Crossref Google ScholarPub Med
  137. Traumatic Brain Injury Alters Methionine Metabolism: Implications for ...
    Go to citation Crossref Google Scholar
  138. Lipid peroxidation in brain or spinal cord mitochondria after injury
    Go to citation Crossref Google Scholar
  139. Altered Mitochondrial Dynamics and TBI Pathophysiology
    Go to citation Crossref Google Scholar
  140. Role of Glia in Memory Deficits Following Traumatic Brain Injury: Biom...
    Go to citation Crossref Google Scholar
  141. Traumatic Brain Injury Causes Aberrant Migration of Adult-Born Neurons...
    Go to citation Crossref Google Scholar
  142. Advanced and High-Throughput Method for Mitochondrial Bioenergetics Ev...
    Go to citation Crossref Google Scholar
  143. Potassium Aspartate Attenuates Brain Injury Induced by Controlled Cort...
    Go to citation Crossref Google Scholar
  144. Stimulating mitochondria to protect the brain following traumatic brai...
    Go to citation Crossref Google Scholar
  145. Therapeutic Targeting of Neuronal Mitochondria in Brain Injury
    Go to citation Crossref Google Scholar
  146. Persistently Altered Brain Mitochondrial Bioenergetics After Apparentl...
    Go to citation Crossref Google Scholar
  147. Mitochondrial bioenergetic alterations after focal traumatic brain inj...
    Go to citation Crossref Google Scholar
  148. Characterization of traumatic brain injury in human brains reveals dis...
    Go to citation Crossref Google Scholar
  149. Characterization of mitochondrial bioenergetics in neonatal anoxic mod...
    Go to citation Crossref Google Scholar
  150. Identification of ideal resuscitation pressure with concurrent traumat...
    Go to citation Crossref Google Scholar
  151. Preserved cardiac mitochondrial function and reduced ischaemia/reperfu...
    Go to citation Crossref Google Scholar
  152. Establishing a Clinically Relevant Large Animal Model Platform for ...
    Go to citation Crossref Google Scholar
  153. Evidence to support mitochondrial neuroprotection, in severe traumatic...
    Go to citation Crossref Google Scholar
  154. Changes in mitochondrial bioenergetics in the brain versus spinal cord...
    Go to citation Crossref Google Scholar
  155. Common biochemical defects linkage between post-traumatic stress disor...
    Go to citation Crossref Google Scholar
  156. Mitochondria-associated microRNAs in rat hippocampus following traumat...
    Go to citation Crossref Google Scholar
  157. Nrf2–ARE activator carnosic acid decreases mitochondrial dysfunction, ...
    Go to citation Crossref Google Scholar
  158. Traumatic Brain Injury and NADPH Oxidase: A Deep Relationship
    Go to citation Crossref Google Scholar
  159. A time course of NADPH-oxidase up-regulation and endothelial nitric ox...
    Go to citation Crossref Google Scholar
  160. The collective therapeutic potential of cerebral ketone metabolism in ...
    Go to citation Crossref Google Scholar
  161. Suppression of Oxidative Stress and 5-Lipoxygenase Activation by Edara...
    Go to citation Crossref Google Scholar
  162. N-acetylcysteine amide confers neuroprotection, improves bioenergetics...
    Go to citation Crossref Google Scholar
  163. Temporal and Spatial Dynamics of Nrf2-Antioxidant Response Elements Me...
    Go to citation Crossref Google Scholar
  164. Antioxidant gene therapy against neuronal cell death
    Go to citation Crossref Google Scholar
  165. Intracranial pressure after the BEST TRIP trial
    Go to citation Crossref Google Scholar
  166. Functional State of Myocardial Mitochondria in Ischemia Reperfusion of...
    Go to citation Crossref Google Scholar
  167. Neuroprotection by genipin against reactive oxygen and reactive nitrog...
    Go to citation Crossref Google Scholar
  168. Mitochondrial Polymorphisms Impact Outcomes after Severe Traumatic Bra...
    Go to citation Crossref Google Scholar
  169. Traumatic Brain Injury Pathophysiology and Treatments: Early, Intermed...
    Go to citation Crossref Google Scholar
  170. Current Understanding of Concussion: Treatment Perspectives
    Go to citation Crossref Google Scholar
  171. Beyond intracranial pressure: optimization of cerebral blood flow, oxy...
    Go to citation Crossref Google Scholar
  172. Cellular and temporal expression of NADPH oxidase (NOX) isotypes after...
    Go to citation Crossref Google Scholar
  173. PPAR Agonists as Therapeutics for CNS Trauma and Neurological Diseases
    Go to citation Crossref Google ScholarPub Med
  174. Intraparenchymal Microdialysis after Acute Spinal Cord Injury Reveals ...
    Go to citation Crossref Google Scholar
  175. Dose- and Time-Dependent Neuroprotective Effects of Pycnogenol ...
    Go to citation Crossref Google Scholar
  176. Impaired mitochondrial oxidative phosphorylation in the peroxisomal di...
    Go to citation Crossref Google Scholar
  177. The Role of Free Radicals in Traumatic Brain Injury
    Go to citation Crossref Google ScholarPub Med
  178. Voltage-Gated Calcium Channel Antagonists and Traumatic Brain Injury
    Go to citation Crossref Google Scholar
  179. Phenelzine Mitochondrial Functional Preservation and Neuroprotection a...
    Go to citation Crossref Google ScholarPub Med
  180. Administration of the Nrf2–ARE activators sulforaphane and carnosic ac...
    Go to citation Crossref Google Scholar
  181. Cellular Alterations in Human Traumatic Brain Injury: Changes in Mitoc...
    Go to citation Crossref Google Scholar
  182. Inhibition of Cathepsin S Produces Neuroprotective Effects after Traum...
    Go to citation Crossref Google Scholar
  183. The pathophysiology of traumatic brain injury at a glance
    Go to citation Crossref Google Scholar
  184. Relationship of nitric oxide synthase induction to peroxynitrite-media...
    Go to citation Crossref Google Scholar
  185. Blast‐induced neurotrauma leads to neurochemical changes and neuronal ...
    Go to citation Crossref Google Scholar
  186. Imaging in the diagnosis and prognosis of traumatic brain injury
    Go to citation Crossref Google Scholar
  187. Calpastatin reduces calpain and caspase activation in methamphetamine-...
    Go to citation Crossref Google Scholar
  188. Blood Glutamate Scavenging: Insight into Neuroprotection
    Go to citation Crossref Google Scholar
  189. Mapping of phospholipids by MALDI imaging (MALDI-MSI): realities and e...
    Go to citation Crossref Google Scholar
  190. Antioxidant therapies in traumatic brain and spinal cord injury
    Go to citation Crossref Google Scholar
  191. Cerebral Microdialysis Effects of Propofol versus Midazolam in Severe ...
    Go to citation Crossref Google Scholar
  192. Mitochondrial Injury after Mechanical Stretch of Cortical Neurons ...
    Go to citation Crossref Google Scholar
  193. Traumatic brain injury and trichloroethylene exposure interact and pro...
    Go to citation Crossref Google Scholar
  194. The mitochondrial permeability transition pore provides a key to the d...
    Go to citation Crossref Google Scholar
  195. Biomarkers of Traumatic Injury
    Go to citation Crossref Google Scholar
  196. Syndromics: A Bioinformatics Approach for Neurotrauma Research
    Go to citation Crossref Google Scholar
  197. The Melatonin MT1 Receptor Axis Modulates Mutant Huntingtin-Mediated T...
    Go to citation Crossref Google Scholar
  198. Agmatine-Promoted Angiogenesis, Neurogenesis, and Inhibition of Gliosi...
    Go to citation Crossref Google Scholar
  199. The involvement of Na+, K+-ATPase activity and free radical generation...
    Go to citation Crossref Google Scholar
  200. Post-Injury Administration of the Mitochondrial Permeability Transitio...
    Go to citation Crossref Google Scholar
  201. Melatonin attenuates brain contusion-induced oxidative insult, inactiv...
    Go to citation Crossref Google Scholar
  202. A selective role for ARMS/Kidins220 scaffold protein in spatial memory...
    Go to citation Crossref Google Scholar
  203. Cyclosporin A Preserves Mitochondrial Function after Traumatic Brain I...
    Go to citation Crossref Google Scholar
  204. Pharmacological inhibition of lipid peroxidation attenuates calpain-me...
    Go to citation Crossref Google Scholar
  205. Cerebral reactive oxygen species assessed by electron spin resonance s...
    Go to citation Crossref Google Scholar
  206. Conditional knockout of brain-derived neurotrophic factor in the hippo...
    Go to citation Crossref Google Scholar
  207. Temporal and Spatial Dynamics of Peroxynitrite-Induced Oxidative Damag...
    Go to citation Crossref Google Scholar
  208. Mild Traumatic Brain Injury Results in Extensive Neuronal Degeneration...
    Go to citation Crossref Google Scholar
  209. Therapeutic Window Analysis of the Neuroprotective Effects of Cyclospo...
    Go to citation Crossref Google Scholar
  210. Cerebral extracellular chemistry and outcome following traumatic brain...
    Go to citation Crossref Google Scholar
  211. Traumatic Brain Injury Pathophysiology/Models
    Go to citation Crossref Google Scholar
  212. Therapeutic Targets for Neuroprotection and/or Enhancement of Function...
    Go to citation Crossref Google Scholar
  213. Thioredoxin and glutaredoxin system proteins—immunolocalization in the...
    Go to citation Crossref Google Scholar
  214. Attenuation of Brain Nitrostative and Oxidative Damage by Brain Coolin...
    Go to citation Crossref Google Scholar
  215. The Role of Reactive Oxygen Species in the Pathogenesis of Traumatic B...
    Go to citation Crossref Google Scholar
  216. Mass‐spectrometry based oxidative lipidomics and lipid imaging: applic...
    Go to citation Crossref Google Scholar
  217. CLOCK and BMAL1 regulate MyoD and are nece...
    Go to citation Crossref Google Scholar
  218. Clinical characteristics and pathophysiological mechanisms of focal an...
    Go to citation Crossref Google Scholar
  219. Lipid Peroxidation-Derived Reactive Aldehydes Directly and Differentia...
    Go to citation Crossref Google Scholar
  220. Mitochondrial protection after traumatic brain injury by scavenging li...
    Go to citation Crossref Google Scholar
  221. Posttraumatic Seizures and Epileptogenesis
    Go to citation Crossref Google Scholar
  222. Age-related changes in mitochondrial respiration and oxidative damage ...
    Go to citation Crossref Google Scholar
  223. Neuroprotective effects of tempol on retinal ganglion cells in a parti...
    Go to citation Crossref Google Scholar
  224. Mitochondrial Damage in Traumatic CNS Injury
    Go to citation Crossref Google Scholar
  225. Metabolic Acetate Therapy for the Treatment of Traumatic Brain Injury
    Go to citation Crossref Google Scholar
  226. Molecular and Physiological Responses to Juvenile Traumatic Brain Inju...
    Go to citation Crossref Google Scholar
  227. α-Synuclein Levels Are Elevated in Cerebrospinal Fluid following Traum...
    Go to citation Crossref Google Scholar
  228. Biomarkers for the Diagnosis, Prognosis, and Evaluation of Treatment E...
    Go to citation Crossref Google Scholar
  229. Modulation of Multiple Pathways Involved in the Maintenance of Neurona...
    Go to citation Crossref Google Scholar
  230. Mitochondrial cytochrome c oxidase expression in the central nervous s...
    Go to citation Crossref Google Scholar
  231. New insights into mitochondrial structure during cell death
    Go to citation Crossref Google Scholar
  232. The optimal dosage and window of opportunity to maintain mitochondrial...
    Go to citation Crossref Google Scholar
  233. Conditional Knockout of Brain-Derived Neurotrophic Factor in the Hippo...
    Go to citation Crossref Google Scholar
  234. Early Mitochondrial Dysfunction after Cortical Contusion Injury
    Go to citation Crossref Google Scholar
  235. Temporal and Spatial Dynamics of Peroxynitrite-Induced Oxidative Damag...
    Go to citation Crossref Google Scholar
  236. Modern Approaches to Pediatric Brain Injury Therapy
    Go to citation Crossref Google Scholar
  237. Adaptation to oxidative challenge induced by chronic physical exercise...
    Go to citation Crossref Google Scholar
  238. Pharmacological evidence for a role of peroxynitrite in the pathophysi...
    Go to citation Crossref Google Scholar
  239. Emerging treatments for traumatic brain injury
    Go to citation Crossref Google Scholar
  240. Lactate and glucose as energy substrates and their role in traumatic b...
    Go to citation Crossref Google Scholar
  241. Comparative Neuroprotective Effects of Cyclosporin a and NIM811, a Non...
    Go to citation Crossref Google ScholarPub Med
  242. Excitotoxicity-Mediated Neurochemical Changes in Neurological Disorder...
    Go to citation Crossref Google Scholar
  243. Free Radicals and Neuroprotection in Traumatic Brain and Spinal Cord I...
    Go to citation Crossref Google Scholar
  244. Oxidative Damage in Neurodegeneration and Injury
    Go to citation Crossref Google Scholar
  245. Tempol protection of spinal cord mitochondria from peroxynitrite-induc...
    Go to citation Crossref Google Scholar
  246. The Nitrone Free Radical Scavenger NXY-059 Is Neuroprotective when Adm...
    Go to citation Crossref Google Scholar
  247. Traumatic Brain Injury Research Priorities: The Conemaugh Internationa...
    Go to citation Crossref Google Scholar
  248. Dosing and safety of cyclosporine in patients with severe brain injury
    Go to citation Crossref Google Scholar
  249. Induction of ketosis may improve mitochondrial function and decrease s...
    Go to citation Crossref Google Scholar
  250. Direct Isolation of Neural Stem Cells in the Adult Hippocampus after T...
    Go to citation Crossref Google Scholar
  251. EFFECT OF CYCLOPENTANONE PROSTAGLANDIN 15-DEOXY-Δ12,14PGJ2 ON EARLY FU...
    Go to citation Crossref Google Scholar
  252. Selective death of newborn neurons in hippocampal dentate gyrus follow...
    Go to citation Crossref Google Scholar
  253. Neuroprotective Effects of Tempol, a Catalytic Scavenger of Peroxynitr...
    Go to citation Crossref Google ScholarPub Med
  254. TEMPORAL WINDOW OF METABOLIC BRAIN VULNERABILITY TO CONCUSSION
    Go to citation Crossref Google Scholar
  255. TEMPORAL WINDOW OF METABOLIC BRAIN VULNERABILITY TO CONCUSSION
    Go to citation Crossref Google Scholar
  256. Multifaceted roles of sphingosine‐1‐phosphate: How does this bioactive...
    Go to citation Crossref Google Scholar
  257. A Time Course of Contusion-Induced Oxidative Stress and Synaptic Prote...
    Go to citation Crossref Google Scholar
  258. Glutathione production is regulated via distinct pathways in stressed ...
    Go to citation Crossref Google Scholar
  259. Attenuation of acute mitochondrial dysfunction after traumatic brain i...
    Go to citation Crossref Google Scholar
  260. Vasopressin serum levels and disorders of sodium and water balance in ...
    Go to citation Crossref Google Scholar
  261. Nitrotyrosine as an Oxidative Stress Marker: Evidence for Involvement ...
    Go to citation Crossref Google Scholar
  262. TEMPORAL WINDOW OF METABOLIC BRAIN VULNERABILITY TO CONCUSSIONS
    Go to citation Crossref Google Scholar
  263. Peroxynitrite‐mediated oxidative damage to brain mitochondria: Protect...
    Go to citation Crossref Google Scholar
  264. Selective early cardiolipin peroxidation after traumatic brain injury:...
    Go to citation Crossref Google Scholar
  265. Mitochondrial dysfunction early after traumatic brain injury in immatu...
    Go to citation Crossref Google Scholar
  266. Protective Effects of NIM811 in Transient Focal Cerebral Ischemia Sugg...
    Go to citation Crossref Google Scholar
  267. Brain neuroprotection by scavenging blood glutamate
    Go to citation Crossref Google Scholar
  268. Perihematomal Mitochondrial Dysfunction After Intracerebral Hemorrhage
    Go to citation Crossref Google Scholar
  269. Glutamate Receptors and Neurological Disorders
    Go to citation Crossref Google Scholar

Figures and tables

Figures & Media

Tables

View Options

View options

PDF/ePub

View PDF/ePub

Get access

Access options

If you have access to journal content via a personal subscription, university, library, employer or society, select from the options below:

ISCBFM members can access this journal content using society membership credentials.

ISCBFM members can access this journal content using society membership credentials.


Alternatively, view purchase options below:

Purchase 24 hour online access to view and download content.

Access journal content via a DeepDyve subscription or find out more about this option.