Next Article in Journal
Current Challenges and Future Directions in Recombinant AAV-Mediated Gene Therapy of Duchenne Muscular Dystrophy
Previous Article in Journal / Special Issue
Phorbol Ester Modulation of Ca2+ Channels Mediates Nociceptive Transmission in Dorsal Horn Neurones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Voltage-Gated Calcium Channel Antagonists and Traumatic Brain Injury

1
Department of Neurological Surgery, One Shields Avenue, University of California, Davis, CA 95616, USA
2
NSF Center for Biophotonics Science and Technology, Suite 2700 Stockton Blvd, Suite 1400, Sacramento, CA, 95817, USA
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2013, 6(7), 788-812; https://doi.org/10.3390/ph6070788
Submission received: 29 March 2013 / Revised: 6 June 2013 / Accepted: 6 June 2013 / Published: 26 June 2013
(This article belongs to the Special Issue Calcium Antagonists)

Abstract

:
Traumatic brain injury (TBI) is a leading cause of death and disability in the United States. Despite more than 30 years of research, no pharmacological agents have been identified that improve neurological function following TBI. However, several lines of research described in this review provide support for further development of voltage gated calcium channel (VGCC) antagonists as potential therapeutic agents. Following TBI, neurons and astrocytes experience a rapid and sometimes enduring increase in intracellular calcium ([Ca2+]i). These fluxes in [Ca2+]i drive not only apoptotic and necrotic cell death, but also can lead to long-term cell dysfunction in surviving cells. In a limited number of in vitro experiments, both L-type and N-type VGCC antagonists successfully reduced calcium loads as well as neuronal and astrocytic cell death following mechanical injury. In rodent models of TBI, administration of VGCC antagonists reduced cell death and improved cognitive function. It is clear that there is a critical need to find effective therapeutics and rational drug delivery strategies for the management and treatment of TBI, and we believe that further investigation of VGCC antagonists should be pursued before ruling out the possibility of successful translation to the clinic.

1. Introduction

1.1. Significance of Traumatic Brain Injury

It is estimated that well over 5.3 million people live in the United States with deficits related to traumatic brain injury (TBI) [1], with over 1.7 new TBI cases annually. A recent meta-analysis examining the prevalence of TBI in the general adult population found that that approximately 12% of the general adult population has a history of TBI with loss of consciousness (16.7% for males and 8.5% for females) [2]. Many patients suffering severe [3,4,5] as well as mild or moderate TBI [5,6,7] are unable to return to work, maintain steady employment, and they struggle with daily tasks due to persistent cognitive deficits. In fact, in a recent study by the TBI Model Systems National Data Centre only 38% of patients followed were employed 2-years following their TBI [8]. Approximately $48.3 billion dollars is spent on TBI patient care in the US each year, with over $31.7 billion (65%) for those that survive [9]. The annual cost for new cases of acute TBI care and rehabilitation is estimated to be approximately $10 billion in the U.S. alone. To date, no effective pharmacotherapy has been shown to improve outcome following TBI, in spite of intensive research in this area [10]. Therefore, the need to develop effective drugs for TBI continues to be pressing. In this paper we review the role of voltage gated calcium channels (VGCC) in the pathophysiology of TBI, and evidence that antagonists of VGCC can be neuroprotective in animal models of TBI and may have potential for clinical use.

1.2. Regulation of Intracellular Calcium

Calcium ions (Ca2+) are major regulators of vital cellular functions. Through their interactions with specific calcium binding proteins, including calmodulin, parvalbumin, calbindin and calretinin, calcium ions are involved in the regulation of secretory functions (e.g., neurotransmitters, hormones), enzyme activity, intracellular transport, contractile processes, glycolysis, respiration, mitosis, membrane potential and intracellular communication [11,12,13,14,15]. These functions are critical for cell survival and disruption of calcium regulation can be catastrophic for the cell. It is therefore not surprising that interference with Ca2+ homeostasis contributes to cell injury and death in a number of pathological conditions, including traumatic brain injury [16]. In fact, calcium has been called the “final common pathway” for toxic cell death [17].
Intracellular cytosolic calcium levels are maintained at low resting levels between 50–100 nM via the combined activity of VGCC’s, receptor operated channels (ROC), store operated channels (SOC), calcium-ATPase transporters in the plasma membrane (PMCA) and smooth endoplasmic reticulum (SERCA), as well as the Na+/Ca2+ exchanger in endoplasmic reticulum, mitochondria and plasma membrane. Some inorganic calcium is also bound to bicarbonate, phosphate and phosphatides. Intracellular stores contained within the endoplasmic reticulum and mitochondria also influence cytosolic calcium levels. For example, receptor-generated inositol triphosphate (IP3) can release calcium from stores within the endoplasmic reticulum [18,19] and ryanodine receptor activation on the endoplasmic reticulum and mitochondria represent an additional important source of intracellular Ca2+ [20].

2. Voltage Gated Calcium Channels

2.1. Voltage Gated Calcium Channel Structure

Voltage-gated calcium channels are heteromultimers formed by an α1 subunit and three auxiliary subunits α2-δ, β, and γ [21] (Figure 1). The α1 subunit is the largest (190–250 kDa) and incorporates the conduction pore, voltage sensor and gating apparatus. It is also the major site of channel regulation by second messengers, drugs, and toxins. The α1 subunit is comprised of four homologous domains composed of six transmembrane helical segments (S1–S6) that determine key channel characteristics [21]. For example, the S4 segment functions as the voltage sensor and the pore loop between segments S5 and S6 determines ion selectivity and conductance. Ten distinct α1 subunits have been described, associated with six different classes of VGCCs [22,23,24,25,26]. In addition there are also several ancillary proteins (β, α2δ, and γ) associated with the α1 subunit that form a multimeric complex (see Figure 1). These ancillary subunits modify the biophysical properties of the VGCC, second messenger modulation and intracellular transport [27,28]. Finally, there is a growing literature related to the structure and function of splice variants of the L- [29,30], N- [31], P/Q- [31], R- [32] and T-type [33] VGCCs. In fact there is evidence that alternate splice variants of both the L- [34,35] and N-type [36] VGCC can influence the affinity and/or sensitivity of channels to specific antagonists.
Figure 1. Subunit structure of voltage gated calcium channels (VGCC). The α1 is the pore-forming subunit which contains voltage-sensing machinery and the binding sites of channel blockers. α1 subunit contains 4 homologous domains (I–IV), each containing 6 transmembrane helices (S1–S6). The α2δ and β subunits enhance expression and modulate the voltage dependence and gating kinetics of VGCCs.
Figure 1. Subunit structure of voltage gated calcium channels (VGCC). The α1 is the pore-forming subunit which contains voltage-sensing machinery and the binding sites of channel blockers. α1 subunit contains 4 homologous domains (I–IV), each containing 6 transmembrane helices (S1–S6). The α2δ and β subunits enhance expression and modulate the voltage dependence and gating kinetics of VGCCs.
Pharmaceuticals 06 00788 g001

2.2. Voltage Gated Calcium Channel Subtypes

Classification of the six subtypes of VGCCs was originally based on their biophysical and pharmacological properties (Table 1). They are typically referred to as L-, N-, P-, Q-, R- and T-type channels [22,23,25,37]. However, a newer gene-based nomenclature is in use [38]. Under the new nomenclature L-type channels are designed Cav1.1-1.4, P/Q as Cav 2.1, N as Cav 2.2, R as Cav 2.3, and T-type channels as Cav3.1-3.3. It addition there is evidence of addition R-type Ca2+ current that is separate from the Cav 2.3 channel [39].
Table 1. Summary of the properties and common antagonists of voltage-gated calcium channel blockers.
Table 1. Summary of the properties and common antagonists of voltage-gated calcium channel blockers.
Channel TypeL
Cav1.1-1.4
N
Cav2.2
P
Cav2.1
Q
Cav2.1
R
Cav2.3
T
Cav3.1-3.3
Ref.
Conductance (pS) 2511 to 209 to 2015 to 1615 to 208[40]
Selectivity (Ca2+>Ba2+)2:12:12:1ND1.3:11:1[40]
Activation Potential (mV) −10 to −50−20−50−50−25 to −40−70[40]
Inactivation Kinetics (msec)150–2,000 100–200500–1,000500–1,00050–10010–70[40]
Calcium Blockers (IC50)
ω-conotoxin MVIIA None78 nM–1 µMNoneNoneNoneNone[40,41,42]
ω-conotoxin GVIA None28 nM–2 µM NoneNoneNoneNone[40,41,42]
ω-Agatoxin AgaIVA NoneNone15 nM 50 nM–1 µM50 nmNone[41,43]
ω-conotoxin MVIIC None18 nM18 nM50 nM–1 µMNoneNone[41,44]
ω-Agatoxin AgaIIIA 1 nm1 nmIC50 N/AIC50 N/ANoneNone[45]
SNX-482NoneNone30–750 nm30–750 nm 15–30 nM None[41,46,47]
Nimodipine0.135–2.6 µMNoneNoneNoneNone5–11 µM[48,49,50,51]
Nifedipine 100 nMNoneNoneNoneNone39 µM[50]
Efonidipine10 µMNoneNoneNoneNone1.3–13 µM[51,52,53]
Amplodipine3–5 µMNoneNoneNoneNone4–13 µM[51,54,55]
Nicardipine9–26 µMNone32–97 µM32–97 µMNone5–13 µM[55]
Verapamil0.6–1 µMNoneNoneNoneNone20–30 µM[49,56,57]
Diltiazem3–33 µMNoneNoneNoneNone30 µM[49,57]
Mibefradil1.7–21 µMNone208 µM208 µMNone0.5–11 µM[51,58,59]
For simplicity we will retain the older nomenclature in this review (i.e., L, N, P, Q, R & T). As shown in the table above, the various channels differ in their conductance rates, voltage activation threshold, and rate of inactivation. A complete review of calcium channel biophysics is beyond the scope of this review. However, L-type channels are high-threshold, large-conductance, slowly inactivating calcium channels [35]. The L-type channel is well understood pharmacologically and defined by its sensitivity to blockade by dihydropyridines such as nimodipine. The N-, P, Q and R-type voltage gated calcium channels are intermediate, having medium conductances, intermediate inactivation kinetics, and medium to high activation thresholds [21]. These calcium channels are selectively blocked by various snail and spider toxins as summarized above. The T-type channel is a low-threshold, low-conductance, rapidly inactivating calcium channel [60]. T-type channels can be blocked by Mibefradil with a 10–30 times higher potency than L-type channels [61,62]. All six types of VGCC’s are found in brain, with the L, N-, P- and Q- and R-type channels involved in synaptic transmission [21]. The N, P, Q and R-type channels are also critically involved in regulating calcium-dependent neurotransmitter release from presynaptic terminals, while the L- and T-type channels appear to make little contribution to this process [24]. The low-voltage activated T-type channel appears to be primarily involved in the generation of rhythmic burst firing of neurons, and is thought to contribute to network synchrony and epilepsy [63].

2.3. Voltage Gated Calcium Channel Distribution in the Nervous System

The distributions of VGCCs in mammalian brain have been examined by receptor autoradiography [64,65] and immunocytochemistry [66,67,68]. L-type VGCC’s are widely distributed in muscle, endocrine cells and brain, but are less abundant in neuronal tissues compared to other VGCC’s [23]. N-type VGCC’s show a wide distribution in rodent brain, with highest densities found in cerebral cortex, dendritic zones of the hippocampus, amygdala, septal nuclei, medial geniculate, superior colliculus, molecular layers of cerebellar cortex, n. solitary tract and spinal cord (layers 1–3) [23,64,66,67,69]. P/Q-type calcium channels, although not as widely distributed as N-type channels, are located on cerebellar granule cells, interneurons and Purkinje cells, and in the hippocampus on pyramidal and granule cells [25,66]. While less well described, there is evidence of R-type VGCC in the cortex [70], thalamus [71], and hippocampus [72,73]. T-type VGCC’s have been demonstrated in cerebellum, thalamus, olfactory bulb and hippocampus [74]. At the cellular level N and P/Q-type VGCC’s are found both pre- and post-synaptically, and interact directly with presynaptic core proteins syntaxin and SNAP-25, providing a molecular basis for Ca+2 influx into nerve terminals and transmitter release [75,76]. Astrocytes have L- and possibly T-type channels, while, to date, VGCCs have not been observed on mature oligodendrocytes [77].

2.4. Characterizing Voltage Gated Calcium Channels Based on Pharmacology

Dihydropyridine antagonists (e.g., nimodipine, nifedipine, verapamil, amlodipine) robustly block L-type channels [78]. While originally it was thought that the remaining VGCC were dihydropyridine insensitive, it was demonstrated in Xenopus oocytes that several of these blockers (including amlodipine) also antagonized N-, P- and Q-type channels at higher concentrations [55]. It was also observed that the dihydropyridine efonidipine inhibits not only the L- but also the T-type VGCC [79]. Finally, it has recently been suggested that, at the concentrations used, the effect of dihydropyridines on the R-type channel have been occluded. Specifically, at a 500 nm concentration, isradipine was able to partially antagonize the R-type current in cardiomyocytes [80]. In conclusion, while the ability of dihydropyridines to antagonize L-type channels is undisputed, these compounds appear to have effects across the spectrum of VGCC. It is clear, therefore, that more research is warranted into the antagonism of each of the VGCC subtypes by dihydropyridines. The N-, P- and Q-type channels can be blocked by specific ω-conotoxins and ω-agatoxins [24,25,81]. The ω-conotoxins or ω-conopeptides are basic, water soluble, 24–29 amino acid peptides isolated from the venom of fish-hunting marine snails belonging to the genus Conus [81]. The best-characterized are the ω-conotoxins GVIA, TVIA (SNX185) and MVIIA (i.e., SNX111) that block the N-type calcium channel [24,82]. A large number of highly selective analogs (e.g., SNX111 and SNX185) have been synthesized and carry the designation “SNX” in the literature. The ω-agatoxin AgaIVA is isolated from the venom of funnel-web spiders and is a potent blocker of P- (Kd1-3 nM) and Q-type (Kd90 nM) VGCC’s. The P- and Q-type VGCC’s are differentiated pharmacologically by their sensitivity, but not selectivity to AgaIVA [24]. The N-, P- and Q-type channels are all blocked by the ω-conotoxin MVIIC [24]. The R-type VGCC’s are resistant to ω-conotoxins and ω-agatoxins, but have been reported to be blocked by a selective inhibitor SNX-482 although the selectivity of SNX-482 for T-type VGCCs has recently been challenged [83,84,85]. T-type blockers include several dihydropyridine antagonists including efonidipine, felodipine and nitrendipine [51,61,86] as well as nimodipine [51]. Mibefradil has also been identified as a potent T-type channel blocker [87]. There is also evidence that specific heavy metals have the ability to block VGCC including lead for L-type [88], nickel [89] and zinc [90] for T-type, and zinc and copper for R-type channels [91]. To date, antagonism of either L-Type (α1S and α1D) or N-type (α1B) channels has been evaluated as a strategy for improving outcome following TBI, but similar studies have not yet been carried out with blockers of P/Q, T or R-type VGCCs.

3. Experimental Evidence that Blockade of VGCC’s Can be Neuroprotective

3.1. Pathological Calcium Accumulation Following TBI

The pathophysiology of TBI has typically been separated into primary and secondary injuries. Primary injury occurs at the time of impact as a result of mechanical tissue deformation resulting in contusions, lacerations, shearing of axonal connections and hemorrhage. Primary injury also initiates a cascade of secondary injury mechanisms, including a large influx of calcium into damaged cells, which can trigger further cell death and lead to substantially increased morbidity [17,92,93,94,95,96].
There is ample experimental evidence that intracellular calcium overload occurs after brain injury, and is a key early step in the activation of secondary injury mechanisms (Figure 2) [17,92,93,94,95,96]. Several in vitro studies have documented a large increase in [Ca2+]i that occurs following traumatic injury to neurons and astrocytes in culture [97,98,99,100,101,102,103,104,105,106,107]. Calcium influx from injury in neurons can result from activation of the N-methyl-D-Aspartate (NMDA) receptor [105,108,109,110,111], opening of VGCCs [102,110,112], as well as release from intracellular stores [16,103,113]. Accumulation of [Ca2+]i in astrocytes has also been demonstrated to come from release from intracellular stores [99,113] as well as the activation of the sodium calcium exchanger [98]. Many in vitro studies have now demonstrated that disruption of calcium homeostasis can injure cells or lead to cell death and have implicated a variety of cellular mechanisms (Figure 2). These include in activation of apoptotic pathways [114,115,116,117,118,119], mitochondrial dysfunction [120,121,122,123,124], free radical production [125] lipid peroxidation [126,127] and osmotic disturbances [128]. In cultures, mechanical strain injury not only increased acute [Ca2+]i, but also triggered delayed depolarization lasting up to 24 h following injury [129]. Consequences of increased calcium uptake include significant neuronal [100,105,112,130,131] and astrocytic [98,101,132] cell death as well as persistent neuronal dysfunction [133,134,135,136,137,138]. Second insults such as hypoxia [112,139], and ischemia [139,140,141,142,143] can lead to additional and extended accumulation of [Ca2+]i which is associated with further diminished outcome [112].
Figure 2. Schematic describing the role of VGCC in TBI-induced neuronal and astorcytic cell death.
Figure 2. Schematic describing the role of VGCC in TBI-induced neuronal and astorcytic cell death.
Pharmaceuticals 06 00788 g002
In vivo studies, although far fewer in number, also support the involvement of calcium influx following traumatic injury to the spinal cord and brain. Extracellular calcium levels fall dramatically immediately after spinal cord injury [95,144], while total tissue levels of calcium increase [107]. Nilsson and colleagues, using calcium sensitive electrodes, identified immediate and dramatic decreases in extracellular calcium at the focus of injury in a weight-drop model of cortical contusion in rats. The authors concluded that massive Ca2+ entry into the intracellular compartment occurring at the site of injury was responsible for this phenomenon [145]. Dienel [146] used 45Ca2+ autoradiography to map increased cellular levels of calcium after transient forebrain ischemia in rats, and argued that such results were due to an accumulation of intracellular calcium after injury. Following lateral fluid percussion, assessment of 45Ca2+ using autoradiography indicated not only an acute increase of calcium in the injured cortex and hippocampus, but a sustained accumulation lasting for several days post-injury. Furthermore, a more diffuse 45Ca2+ accumulation was also detected in the thalamus that could be observed for up to two weeks following injury [147,148]. Using the lateral fluid percussion model of TBI, DeLorenzo and colleagues acutely isolated injured hippocampal neurons in culture and imaged calcium fluxes using the calcium sensitive dye Fura-2-AM. In these preparations [Ca2+]i levels remained elevated for at least one week following fluid percussion TBI and calcium homeostasis was impaired for at least 30 days, indicating that TBI has enduring effects on Ca2+ dynamics [149,150]. These pre-clinical studies document the significant increase in [Ca2+]i that follows TBI, and point to the possibility that pharmacological interventions aimed at reducing the rise in [Ca2+]I, such as antagonism of VGCC (Figure 2), could be a rational and effective therapeutic strategy to reduce cell death and dysfunction.

3.2. L-Type VGCC Antagonists to Treat Traumatic Brain Injury

The L-type VGCCs require strong depolarization for activation, and are typically related to long-lasting calcium currents [21]. L-type VGCCs are prominent in the vertebrate cardiac, skeletal and smooth muscle. In the central nervous system, they are located on the cell body and proximal and distal dendrites [66,151]. There are three classes of L-type VGCC antagonists including phenalkylamines (verapamil, D-600), benzothiazepine (dlitiazem) and dihydropyridines (nimodipine, nifedipine, nitrendipine) [21]. Pre-clinical rodent [152,153,154,155] and primate [156] studies demonstrated the utility of nimodipine to improve cerebral blood flow, reduce infarct volume and improve behavioral outcome following an ischemic injury. Importantly, improved outcome was observed in the absence of significant drug side effects. Clinical trials to translate nimodipine to treat the acute effects of ischemia found significant improvements in both blood flow and outcome following injury [157,158,159].
Despite the promising clinical evidence for the potential of L-type antagonists to improve outcome following an ischemic insult, there is significantly less data evaluating these compounds in traumatic brain injury. Initial in vitro studies found that nimodipine reduced cell death following treatment with high concentrations of glutamate [160]. In in vitro cell culture models of uniaxial and bi-axial mechanical strain injury, a cocktail of a sodium channel blocker (tetrodotoxin, TTX) NMDA antagonist ((2R)-amino-5-phosphonopentanoate; APV), AMPA antagonist (6-cyano-7-nitroquinoxaline-2,3-dione; CNQX), and nimodipine applied prior to strain significantly reduced the average accumulation of [Ca2+]i in neurons as well as the total number of neurons experiencing significant [Ca2+]i load [130]. In a different model of mechanical injury, nifedipine significantly reduced cell death and improved cell function in an organotypic hippocampal slice culture [97]. In initial in vivo studies, verapamil administered following a lateral fluid percussion traumatic brain injury in rats significantly improved cerebral blood flow in the injured cortex leading the authors to conclude that L-type VGCC play a role in vasoconstriction and the loss of vasoreactivity following TBI [161]. In a separate study, nimodipine treatment was able to reduce lipid peroxidation caused by traumatic brain injury in rodents [162]. Considering the extensive literature on the accumulation of [Ca2+]i and the loss of Ca2+ homeostasis, there is sparse pre-clinical data evaluating the potential of L-type VGCC antagonists to improve outcome following TBI.
Even with such limited pre-clinical evidence, L-type VGCC antagonists have been tested in several clinical trials in patients with severe TBI examining their potential to improve outcome when administered systemically acutely following injury. In initial clinical trials (HIT 1) it was observed that intravenous delivery of nimodipine had modest effects on outcome with only a trend toward improved outcome [163,164]. The authors also noted that the drug was well tolerated and patients experienced few side effects. In a follow up study (HIT II) [165] these same authors reached a similar conclusion that nimodipine had modest effects on patients with severe TBI. However the authors made an additional observation; nimodipine was more efficacious in patients with evidence of traumatic subarachnoid hemorrhage from their initial CT scan. However subsequent studies have failed to find a similar beneficial effect even within this subpopulation [166,167]. A more recent study reported that cerebral metabolism and outcome were improved in a small patient population by intravenous nimodipine treatment [168,169]. Therefore, there continues to be considerable uncertainty about a beneficial effect of nimodipine treatment for acute traumatic brain injury [170]. In summary, to date, antagonists of L-type VGCC have failed to provide convincing evidence that blockade of L-type VGCCs is an effective therapeutic strategy for TBI.

3.3. N-Type VGCC Antagonists to Treat Traumatic Brain Injury

N-type VGCCs are activated by strong depolarization, and are found on the dendrites, soma, and axon terminals of neurons [67,171,172,173]. These VGCCs located at axon terminals play a critical role in neurotransmitter release throughout the central nervous system [172,174], and may play a dominant role in norepinephrine release from sympathetic neurons [175]. In 1985 Olivera and colleagues described small peptides isolated from the venoms of cone snails that were very specific and potent inhibitors of N-type VGCCs [25,69,81]. One sub-class, the ω-conotoxins, MVIIA was found to be highly selective for the N-type VGCC, and a synthetic form of the toxin called ziconotide (also known as SNX-111, Prialt®) was developed and has been used in studies of pain systems, ischemia and TBI [176,177,178,179]. Similar to L-type antagonists, initial characterization of ziconotide was on improving outcome in rodent models of ischemia [180,181,182]. Unlike L-type antagonists, however, the focus of ziconotide treatment was neuroprotection and not blood flow.
Because of their role in neurotransmitter release, N-type VGCCs represent an attractive target for neuroprotection through their ability to reduce excitatory neurotransmitter release and there by prevent or limit excitotoxic cell damage by glutamate. However, there have been relatively few pre-clinical studies of the potential of N-type VGCC antagonists in experimental models of TBI. A recent in vitro study, using a mechanical strain-injury model of TBI in mixed neuronal glial cultures, examined the effects of the selective N-type VGCC blocker SNX-185 on [Ca2+]i imaged with Fura-2-AM and neuronal and astrocyte survival after injury by immunohistochemistry. SNX-185 is a synthetic form of ω-conotoxin TVIA and is similar to MVIIA (i.e., ziconotide) in its selectivity for N-type VGCCs. Addition of SNX-185 to the culture media before or immediately (i.e., <5 sec) after mechanical strain-injury significantly reduced the rise in [Ca2+]i and improved survival of injured neurons and astrocytes following mild, moderate or severe mechanical strain injury [102]. Concentrations of 100 and 1000 nM SNX-185 in the culture media appeared to be equally effective. Delayed addition of SNX-185 to the media 5 min after mechanical strain injury did not improve cell survival nor did it prevent the rise in [Ca2+]i. However, these results would be expected because in the in vitro mechanical strain-injury model the peak increase in [Ca2+]i occurs within a few seconds to a minute after injury, before the 5 min delayed treatment with SNX-185. However, mechanical injury does lead to extended elevations in [Ca2+]i [16,104,105]. Interestingly, treatment with SNX-185 facilitated the return of injured neurons to more normal calcium homeostasis. Specifically, elevated levels of [Ca2+]i following strain injury returned to near baseline levels (i.e., 100 nM) significantly sooner in injured neurons when SNX-185 was added to the bath media 5 min after injury.
Because [Ca2+]i can remain elevated for long periods of time after injury, these results suggest that that N-type VGCC blockers may prevent the rise in [Ca2+]i for cells that may be injured but show a delayed rise in [Ca2+]i and facilitate the return of calcium homeostasis. However, it remains to be demonstrated experimentally whether calcium homeostasis is normal or whether or not the return to baseline calcium levels would be beneficial to injured cells. Microdialysis samples taken from these cultures demonstrated a significant reduction in extracellular glutamate [102] suggesting that, in part, SNX-185 was decreasing calcium accumulation by blocking presynaptic N-type calcium channels and thereby reducing pre-synaptic glutamate release. Recently, these same authors tested whether SNX-185 would have similar effects in an in vitro model of mechanical injury with hypoxia as a second insult [112]. Hypoxia was produced by changing the cell culture CO2 level from 5% (normoxic) to 20% (i.e., hypoxic) after mechanical strain injury, and then carrying out calcium imaging with Fura-2-AM 3, 6 and 24 h later. Compared to the normoxic condition, hypoxia further increased the rise of [Ca2+] after injury at all three imaging time points, and resulted in greater cell death across all mechanical strain injury severities. Addition of 1µM SNX-185 5 min after the second insult reduced the added effects of hypoxia on [Ca2+] as well as the effects of hypoxia on cell survival, resulting in great numbers of surviving neurons. These results point to the importance of secondary insults in the pathophysiology of TBI and further support continued research and development of N-type VGCC antagonists to improve outcome in TBI patients at risk for second insult. In an in vitro model of axonal injury in which the opening of stretch-activated sodium channels in axons was followed by an increase in calcium accumulation, damage to the injured axon was prevented by either blocking sodium channels with tetratrodotoxin or blockade of VGCCs by the omega-conotoxin MVIIC [106].
In vivo studies with ziconotide, the synthetic version of MVIIA, have also demonstrated neuroprotective activity in rodent models of TBI. When administered i.v. 1 h following lateral fluid percussion TBI, ziconotide significantly reduced the increase in 45Ca2 accumulation in the dorsal hippocampus for up to 48 h following injury, with less dramatic effects observed in the cortex [183]. In a follow up study the same authors found that intravenous ziconotide dramatically reduced 45Ca2+ accumulation in the ipsilateral cortex (by 75%), but also several other brain regions including ipsilateral thalamus and hippocampus (up to 50%) [184]. In the rat model of controlled cortical impact TBI, an accumulation of Ca2+ was found in mitochondria along with significant mitochondria respiratory dysfunction for up to 48 h following injury [124]. Intravenous administration of ziconotide within 4 h from the time of injury significantly improved mitochondrial respiration in the injured brain [185,186,187]. Mitochondrial function was restored even when ziconoitde was delivered 10 h following the lateral fluid percussion injury [185]. Repeated intravenous injections of ziconotide at 3, 6, and 12 h after an impact-acceleration model of diffuse axonal traumatic brain injury in rats significantly improved cognitive and sensory-motor function [179]. In a more recent study, direct stereotaxic injection of SNX-185 into the dorsal hippocampus 5 min following lateral fluid percussion brain injury significantly reduced neuronal loss in the ipsilateral hippocampus and improved motor and cognitive function [188]. However, in a recent study, extracellular glutamate was significantly elevated in the striatum for up to 48 h following midline fluid percussion TBI in rats [189], and the rise in glutamate was not blocked by the N-type VGCC blocker GVIA, suggesting that the elevated striatal glutamate following TBI may not be calcium-dependent.
Although the preclinical data were limited and the exact neuroprotective mechanisms of ziconotide were not well understood, a phase II human safety, efficacy, and feasibility trial of intravenous ziconitide in severe TBI patients was initiated. Unfortunately the study was halted before completion due to complicating cardiovascular (i.e., hypotension) effects of systemic ziconotide treatment, likely due to the effects of N-type antagonists on heart rate and blood pressure [190,191], and the results of this trial have not been published. Therefore, the potential for clinical development of N-type VGCC blockers remains an open question, one that should be further explored given the established role of calcium in cell injury following TBI, the established efficacy of N-type VGCC blockers in limiting intracellular calcium accumulation, and the strength of the preclinical data in in vitro and in vivo studies.

4. Conclusions

4.1. Lessons to be Learned from Ziconotide Development for Chronic Pain

To date there has been a failure to translate pre-clinical therapeutic strategies from bench to the bedside to treat patients with TBI. A great deal has been written about this dilemma and several recommendations have come from careful analyses of recent failures [10]. First, experimental treatments should target known mechanisms of secondary injury following TBI and the dosages used should be adequate to block the targeted injury mechanism. Second, the effectiveness of the treatment should be demonstrated in animal models of brain injury, and the therapeutic window and optimal duration of treatment should be established. Third, valid outcome measures should be used. Finally, safety of the treatment for eventual use in patients with TBI should be demonstrated.
These recommendations and the development and clinical approval of intrathecal ziconotide for treatment of intractable pain offer lessons for the development of drugs, such as ziconotide, for use in TBI. As in TBI studies, initial findings in rodents [176,177,178,192] and in primates [176] indicated the potential for ziconotide to reduce chronic pain in the absence of significant side effects. However, initial trials to translate ziconotide to the bedside were not successful due to significant peripheral side effects of the calcium channel blocker in a limited clinical study, including substantial hypotension following systemic administration [193]. However, rather than abandoning zicontide for pain management, the intrathecal route of administration was adopted that provided pain relief at a lower dose and did not produce significant cardiovascular or central CNS effects. Currently intrathecal ziconotide, now available as Prialt, has been used successfully in patients to treat acute post-operative pain [194], AIDS-related pain, cancer-related pain [195], and severe chronic pain [196]. In the most recent trials, ziconotide was found to be both efficacious and safe for continuous administration over at least a three-year period [197,198]. Based on the evidence reviewed above it would seem reasonable to continue development of N-type VGCC blockers, including ziconotide and SNX-185, to improve motor and cognitive function following TBI.

4.2. Summary

Based on the limited pre-clinical data, we feel that specific L- and N-type VGCC antagonists continue to have translational potential for pharmacotherapy in TBI patients. Furthermore, there have been no specific studies addressing the role that P/Q, R and T-type VGCC play in the pathophysiology of traumatic brain injury. The mechanisms of action for antagonists for each of these channels have been well defined, and their bioavailability and pharmacokinetics have been characterized. There is substantial preclinical data supporting further development of N-type VGCC blockers for use in TBI [199]. To date, however, only a very limited number of pre-clinical experiments have evaluated the potential of either L- or N-type VGCC antagonists to improve outcome following TBI. There is also evidence that R-type VGCC may increase the constriction of arteries improving outcome in rodent models of subarachnoid hemorrhage [200] and that the R-type antagonist SNX-482 can improve blood flow in the days following injury [201]. Future studies should also include other N-type VGCC blockers such as SNX-185, antagonists of other VGCC (i.e., P/Q, R or T-type channels), the use of lower doses to avoid or reduce complications due to systemic side effects. In addition, as demonstrated in the treatment for chronic pain, intrathecal delivery of ziconotide can reduce the effect of channel antagonists on peripheral targets such as the heart and vasculature. This is relevant not only for N-type channels but also R- [200,202], T- [202], P/Q- [202,203,204] and L-type channels [202,205] that also may alter global heart rate and/or blood flow when administered systemically. There may also be potential for VGCC antagonists to improve function following TBI in cases where patients endure second insults, such as hypoxia, ischemia or seizures. However, based on the recommendation from Bullock and colleagues [10], it is clear that further pre-clinical evaluation is necessary in order to design a well-informed clinical trial.

Conflict of interest

The authors hold no conflict of interest.

References

  1. Faul, M.; Xu, L.; Wald, M.M.; Coronado, V.G. Traumatic Brain Injury in the United States: Emergency Department Visits, Hospitalizations and Deaths 2002–2006; Centers for Disease Control and Prevention: Atlanta, GA, USA, 2010. [Google Scholar]
  2. Frost, R.B.; Farrer, T.J.; Primosch, M.; Hedges, D.W. Prevalence of traumatic brain injury in the general adult population: A meta-analysis. Neuroepidemiology 2012, 40, 154–159. [Google Scholar]
  3. Brooks, D.N.; Aughton, M.E. Cognitive recovery during the first year after severe blunt head injury. Int. Rehabil. Med. 1979, 1, 166–172. [Google Scholar]
  4. Brooks, N.; McKinlay, W.; Symington, C.; Beattie, A.; Campsie, L. Return to work within the first seven years of severe head injury. Brain Inj. 1987, 1, 5–19. [Google Scholar] [CrossRef]
  5. Benedictus, M.R.; Spikman, J.M.; van der Naalt, J. Cognitive and behavioral impairment in traumatic brain injury related to outcome and return to work. Arch. Phys. Med. Rehabil. 2010, 91, 1436–1441. [Google Scholar] [CrossRef]
  6. Bootes, K.; Chapparo, C. Difficulties with multitasking on return to work after TBI: A critical case study. Work 2010, 36, 207–216. [Google Scholar]
  7. Gilworth, G.; Eyres, S.; Carey, A.; Bhakta, B.B.; Tennant, A. Working with a brain injury: Personal experiences of returning to work following a mild or moderate brain injury. J. Rehabil. Med. 2008, 40, 334–339. [Google Scholar] [CrossRef]
  8. Shigaki, C.L.; Johnstone, B.; Schopp, L.H. Financial and vocational outcomes 2 years after traumatic brain injury. Disabil. Rehabil. 2009, 31, 484–489. [Google Scholar] [CrossRef]
  9. Lewin, J.; Sumners, D. Anorexia due to brain injury. Brain Inj. 1992, 6, 199–201. [Google Scholar] [CrossRef]
  10. Bullock, M.R.; Lyeth, B.G.; Muizelaar, J.P. Current status of neuroprotection trials for traumatic brain injury: Lessons from animal models and clinical studies. Neurosurgery 1999, 45, 207–217; discussion 217–220. [Google Scholar]
  11. Petersen, O.H.; Michalak, M.; Verkhratsky, A. Calcium signalling: Past, present and future. Cell Calcium 2005, 38, 161–169. [Google Scholar] [CrossRef]
  12. Carafoli, E. Intracellular calcium homeostasis. Annu. Rev. Biochem. 1987, 56, 395–433. [Google Scholar] [CrossRef]
  13. Miller, R.J. The control of neuronal Ca2+ homeostasis. Prog. Neurobiol. 1991, 37, 255–285. [Google Scholar] [CrossRef]
  14. Racay, P.; Kaplan, P.; Lehotsky, J. Control of Ca2+ homeostasis in neuronal cells. Gen. Physiol. Biophys. 1996, 15, 193–210. [Google Scholar]
  15. Racay, P.; Lehotsky, J. Intracellular and molecular aspects of Ca(2+)-mediated signal transduction in neuronal cells. Gen. Physiol. Biophys. 1996, 15, 273–289. [Google Scholar]
  16. Weber, J.T. Calcium homeostasis following traumatic neuronal injury. Curr. Neurovasc. Res. 2004, 1, 151–171. [Google Scholar] [CrossRef]
  17. Schanne, F.A.; Kane, A.B.; Young, E.E.; Farber, J.L. Calcium dependence of toxic cell death: A final common pathway. Science 1979, 206, 700–702. [Google Scholar]
  18. Berridge, M.J.; Irvine, R.F. Inositol phosphates and cell signalling. Nature 1989, 341, 197–205. [Google Scholar] [CrossRef]
  19. Carafoli, E.; Longoni, S. The plasma membrane in the control of the signaling function of calcium. Soc. Gen. Physiol. Ser. 1987, 42, 21–29. [Google Scholar]
  20. Campbell, K.P.; Knudson, C.M.; Imagawa, T.; Leung, A.T.; Sutko, J.L.; Kahl, S.D.; Raab, C.R.; Madson, L. Identification and characterization of the high affinity [3h]ryanodine receptor of the junctional sarcoplasmic reticulum Ca2+ release channel. J. Biol. Chem. 1987, 262, 6460–6463. [Google Scholar]
  21. Hille, B. Ion Channels of Excitable Membranes, 3rd ed; Sinauer Associates, Inc.: Sunderland, MA, USA, 2001; Chapter xviii; pp. 814, 818. [Google Scholar]
  22. Catterall, W.A. Structure and function of voltage-gated ion channels. Annu. Rev. Biochem. 1995, 64, 493–531. [Google Scholar] [CrossRef]
  23. Dunlap, K.; Luebke, J.I.; Turner, T.J. Exocytotic Ca2+ channels in mammalian central neurons. Trends Neurosci. 1995, 18, 89–98. [Google Scholar] [CrossRef]
  24. Miljanich, G.P.; Ramachandran, J. Antagonists of neuronal calcium channels: Structure, function, and therapeutic implications. Annu. Rev. Pharmacol. Toxicol. 1995, 35, 707–734. [Google Scholar] [CrossRef]
  25. Olivera, B.M.; Miljanich, G.P.; Ramachandran, J.; Adams, M.E. Calcium channel diversity and neurotransmitter release: The omega-conotoxins and omega-agatoxins. Annu. Rev. Biochem. 1994, 63, 823–867. [Google Scholar] [CrossRef]
  26. Tsien, R.W.; Nilius, B. Cardiac calcium currents at the level of single channels. Experientia 1987, 43, 1169–1172. [Google Scholar] [CrossRef]
  27. Cain, S.M.; Snutch, T.P. Voltage-gated calcium channels and disease. BioFactors 2011, 37, 197–205. [Google Scholar] [CrossRef]
  28. Dolphin, A.C. Calcium channel auxiliary alpha2delta and beta subunits: Trafficking and one step beyond. Nat. Rev. Neurosci. 2012, 13, 542–555. [Google Scholar] [CrossRef]
  29. Kim, H.L.; Kim, H.; Lee, P.; King, R.G.; Chin, H. Rat brain expresses an alternatively spliced form of the dihydropyridine-sensitive L-type calcium channel alpha 2 subunit. Proc. Natl. Acad. Sci. USA 1992, 89, 3251–3255. [Google Scholar] [CrossRef]
  30. Perez-Reyes, E.; Wei, X.Y.; Castellano, A.; Birnbaumer, L. Molecular diversity of L-type calcium channels. Evidence for alternative splicing of the transcripts of three non-allelic genes. J. Biol. Chem. 1990, 265, 20430–20436. [Google Scholar]
  31. Allen, S.E.; Darnell, R.B.; Lipscombe, D. The neuronal splicing factor nova controls alternative splicing in N-type and P-type Cav2 calcium channels. Channels 2010, 4, 483–489. [Google Scholar]
  32. Williams, M.E.; Marubio, L.M.; Deal, C.R.; Hans, M.; Brust, P.F.; Philipson, L.H.; Miller, R.J.; Johnson, E.C.; Harpold, M.M.; Ellis, S.B. Structure and functional characterization of neuronal alpha 1e calcium channel subtypes. J. Biol. Chem. 1994, 269, 22347–22357. [Google Scholar]
  33. Senatore, A.; Spafford, J.D. Gene transcription and splicing of T-type channels are evolutionarily-conserved strategies for regulating channel expression and gating. PLoS One 2012, 7, e37409. [Google Scholar] [CrossRef]
  34. Zhang, H.Y.; Liao, P.; Wang, J.J.; Yu de, J.; Soong, T.W. Alternative splicing modulates diltiazem sensitivity of cardiac and vascular smooth muscle Ca(v)1.2 calcium channels. Br. J. Pharmacol. 2010, 160, 1631–1640. [Google Scholar] [CrossRef]
  35. Zuccotti, A.; Clementi, S.; Reinbothe, T.; Torrente, A.; Vandael, D.H.; Pirone, A. Structural and functional differences between l-type calcium channels: Crucial issues for future selective targeting. Trends Pharmacol. Sci. 2011, 32, 366–375. [Google Scholar] [CrossRef]
  36. Helton, T.D.; Kojetin, D.J.; Cavanagh, J.; Horne, W.A. Alternative splicing of a beta4 subunit proline-rich motif regulates voltage-dependent gating and toxin block of Cav2.1 Ca2+ channels. J. Neurosci. 2002, 22, 9331–9339. [Google Scholar]
  37. Catterall, W.A. Voltage-gated calcium channels. Cold Spring Harbor Perspectives in Biology 2011, 3, a003947. [Google Scholar] [CrossRef]
  38. Ertel, E.A.; Campbell, K.P.; Harpold, M.M.; Hofmann, F.; Mori, Y.; Perez-Reyes, E.; Schwartz, A.; Snutch, T.P.; Tanabe, T.; Birnbaumer, L.; et al. Nomenclature of voltage-gated calcium channels. Neuron 2000, 25, 533–535. [Google Scholar] [CrossRef]
  39. Sochivko, D.; Pereverzev, A.; Smyth, N.; Gissel, C.; Schneider, T.; Beck, H. The Ca(v)2.3 Ca(2+) channel subunit contributes to R-type Ca(2+) currents in murine hippocampal and neocortical neurones. J. Physiol. 2002, 542, 699–710. [Google Scholar] [CrossRef]
  40. Sharman, J.L.; Benson, H.E.; Pawson, A.J.; Lukito, V.; Mpamhanga, C.P.; Bombail, V.; Davenport, A.P.; Peters, J.A.; Spedding, M.; Harmar, A.J.; et al. IUPHAR-DB: Updated database content and new features. Nucl. Acids Res. 2013, 41, 1083–1088. [Google Scholar] [CrossRef]
  41. Pringos, E.; Vignes, M.; Martinez, J.; Rolland, V. Peptide neurotoxins that affect voltage-gated calcium channels: A close-up on omega-agatoxins. Toxins 2011, 3, 17–42. [Google Scholar] [CrossRef]
  42. Bowman, D.; Alexander, S.; Lodge, D. Pharmacological characterisation of the calcium channels coupled to the plateau phase of kcl-induced intracellular free Ca2+ elevation in chicken and rat synaptosomes. Neuropharmacology 1993, 32, 1195–1202. [Google Scholar] [CrossRef]
  43. Stephens, G.J.; Page, K.M.; Burley, J.R.; Berrow, N.S.; Dolphin, A.C. Functional expression of rat brain cloned alpha1e calcium channels in cos-7 cells. Pflugers Arch. 1997, 433, 523–532. [Google Scholar] [CrossRef]
  44. McDonough, S.I.; Swartz, K.J.; Mintz, I.M.; Boland, L.M.; Bean, B.P. Inhibition of calcium channels in rat central and peripheral neurons by omega-conotoxin mviic. J. Neurosci. 1996, 16, 2612–2623. [Google Scholar]
  45. Ertel, E.A.; Warren, V.A.; Adams, M.E.; Griffin, P.R.; Cohen, C.J.; Smith, M.M. Type iii omega-agatoxins: A family of probes for similar binding sites on L- and N-type calcium channels. Biochemistry 1994, 33, 5098–5108. [Google Scholar] [CrossRef]
  46. Arroyo, G.; Aldea, M.; Fuentealba, J.; Albillos, A.; Garcia, A.G. Snx482 selectively blocks P/Q Ca2+ channels and delays the inactivation of Na+ channels of chromaffin cells. Eur. J. Pharmacol. 2003, 475, 11–18. [Google Scholar] [CrossRef]
  47. Lu, Z.J.; Pereverzev, A.; Liu, H.L.; Weiergraber, M.; Henry, M.; Krieger, A.; Smyth, N.; Hescheler, J.; Schneider, T. Arrhythmia in isolated prenatal hearts after ablation of the Cav2.3 (alpha1e) subunit of voltage-gated Ca2+ channels. Cell. Phys. Biochem. 2004, 14, 11–22. [Google Scholar]
  48. Xu, W.; Lipscombe, D. Neuronal Ca(v)1.3alpha(1) l-type channels activate at relatively hyperpolarized membrane potentials and are incompletely inhibited by dihydropyridines. J. Neurosci. 2001, 21, 5944–5951. [Google Scholar]
  49. Catterall, W.A.; Perez-Reyes, E.; Snutch, T.P.; Striessnig, J. International Union Of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharm. Rev. 2005, 57, 411–425. [Google Scholar]
  50. Stengel, W.; Jainz, M.; Andreas, K. Different potencies of dihydropyridine derivatives in blocking T-type but not L-type Ca2+ channels in neuroblastoma-glioma hybrid cells. Eur. J. Pharmacol. 1998, 342, 339–345. [Google Scholar] [CrossRef]
  51. Furukawa, T.; Nukada, T.; Namiki, Y.; Miyashita, Y.; Hatsuno, K.; Ueno, Y.; Yamakawa, T.; Isshiki, T. Five different profiles of dihydropyridines in blocking T-type Ca(2+) channel subtypes (Ca(v)3.1 (alpha(1g)), Ca(v)3.2 (alpha(1h)), and Ca(v)3.3 (alpha(1i))) expressed in Xenopus oocytes. Eur. J. Pharmacol. 2009, 613, 100–107. [Google Scholar]
  52. Horiba, M.; Muto, T.; Ueda, N.; Opthof, T.; Miwa, K.; Hojo, M.; Lee, J.K.; Kamiya, K.; Kodama, I.; Yasui, K. T-type Ca2+ channel blockers prevent cardiac cell hypertrophy through an inhibition of calcineurin-nfat3 activation as well as L-type Ca2+ channel blockers. Life Sci. 2008, 82, 554–560. [Google Scholar] [CrossRef]
  53. Masumiya, H.; Kase, J.; Tanaka, Y.; Tanaka, H.; Shigenobu, K. Frequency-dependent blockade of T-type Ca2+ current by efonidipine in cardiomyocytes. Life Sci. 2000, 68, 345–351. [Google Scholar] [CrossRef]
  54. Kwan, Y.W.; Bangalore, R.; Lakitsh, M.; Glossmann, H.; Kass, R.S. Inhibition of cardiac L-type calcium channels by quaternary amlodipine: Implications for pharmacokinetics and access to dihydropyridine binding site. J. Mol. Cell Cardiol. 1995, 27, 253–262. [Google Scholar] [CrossRef]
  55. Furukawa, T.; Yamakawa, T.; Midera, T.; Sagawa, T.; Mori, Y.; Nukada, T. Selectivities of dihydropyridine derivatives in blocking Ca(2+) channel subtypes expressed in Xenopus oocytes. J. Pharmacol. Exp. Ther. 1999, 291, 464–473. [Google Scholar]
  56. Freeze, B.S.; McNulty, M.M.; Hanck, D.A. State-dependent verapamil block of the cloned human Ca(v)3.1 T-type Ca(2+) channel. Mol. Pharmacol. 2006, 70, 718–726. [Google Scholar]
  57. Kuga, T.; Sadoshima, J.; Tomoike, H.; Kanaide, H.; Akaike, N.; Nakamura, M. Actions of Ca2+ antagonists on two types of Ca2+ channels in rat aorta smooth muscle cells in primary culture. Circ. Res. 1990, 67, 469–480. [Google Scholar] [CrossRef]
  58. Xi, Q.; Angus, J.A. Evidence against an action of mibefradil at N-type voltage-operated calcium channels. Naunyn Schmiedebergs Arch. Pharmacol. 2001, 364, 430–436. [Google Scholar] [CrossRef]
  59. Aczel, S.; Kurka, B.; Hering, S. Mechanism of voltage- and use-dependent block of class a Ca2+ channels by mibefradil. Br. J. Pharmacol. 1998, 125, 447–454. [Google Scholar] [CrossRef]
  60. Perez-Reyes, E.; Lory, P. Molecular biology of T-type calcium channels. CNS Neurol. Disord. Drug Targets 2006, 5, 605–609. [Google Scholar] [CrossRef]
  61. Perez-Reyes, E.; Van Deusen, A.L.; Vitko, I. Molecular pharmacology of human Cav3.2 T-type Ca2+ channels: Block by antihypertensives, antiarrhythmics, and their analogs. J. Pharmacol. Exp. Ther. 2009, 328, 621–627. [Google Scholar]
  62. McDonough, S.I.; Bean, B.P. Mibefradil inhibition of T-type calcium channels in cerebellar purkinje neurons. Mol. Pharmacol. 1998, 54, 1080–1087. [Google Scholar]
  63. Cain, S.M.; Snutch, T.P. T-type calcium channels in burst-firing, network synchrony, and epilepsy. Biochim. Biophys. Acta 2013, 1828, 1572–1578. [Google Scholar] [CrossRef]
  64. Takemura, M.; Kiyama, H.; Fukui, H.; Tohyama, M.; Wada, H. Autoradiographic visualization in rat brain of receptors for omega-conotoxin gvia, a newly discovered calcium antagonist. Brain Res. 1988, 451, 386–389. [Google Scholar] [CrossRef]
  65. Takemura, M.; Kiyama, H.; Fukui, H.; Tohyama, M.; Wada, H. Distribution of the omega-conotoxin receptor in rat brain. An autoradiographic mapping. Neuroscience 1989, 32, 405–416. [Google Scholar] [CrossRef]
  66. Timmermann, D.B.; Westenbroek, R.E.; Schousboe, A.; Catterall, W.A. Distribution of high-voltage-activated calcium channels in cultured gamma-aminobutyric acidergic neurons from mouse cerebral cortex. J. Neurosci. Res. 2002, 67, 48–61. [Google Scholar] [CrossRef]
  67. Westenbroek, R.E.; Sakurai, T.; Elliott, E.M.; Hell, J.W.; Starr, T.V.; Snutch, T.P.; Catterall, W.A. Immunochemical identification and subcellular distribution of the alpha 1a subunits of brain calcium channels. J. Neurosci. 1995, 15, 6403–6418. [Google Scholar]
  68. Nakanishi, S.; Fujii, A.; Kimura, T.; Sakakibara, S.; Mikoshiba, K. Spatial distribution of omega-agatoxin IVA binding sites in mouse brain slices. J. Neurosci. Res. 1995, 41, 532–539. [Google Scholar] [CrossRef]
  69. McCleskey, E.W.; Fox, A.P.; Feldman, D.H.; Cruz, L.J.; Olivera, B.M.; Tsien, R.W.; Yoshikami, D. Omega-conotoxin: Direct and persistent blockade of specific types of calcium channels in neurons but not muscle. Proc. Natl. Acad. Sci. USA 1987, 84, 4327–4331. [Google Scholar] [CrossRef]
  70. Rhee, J.S.; Ishibashi, H.; Akaike, N. Calcium channels in the gabaergic presynaptic nerve terminals projecting to meynert neurons of the rat. J. Neurochem. 1999, 72, 800–807. [Google Scholar] [CrossRef]
  71. Zaman, T.; Lee, K.; Park, C.; Paydar, A.; Choi, J.H.; Cheong, E.; Lee, C.J.; Shin, H.S. Cav2.3 channels are critical for oscillatory burst discharges in the reticular thalamus and absence epilepsy. Neuron 2011, 70, 95–108. [Google Scholar] [CrossRef]
  72. Muller, R.; Struck, H.; Ho, M.S.; Brockhaus-Dumke, A.; Klosterkotter, J.; Broich, K.; Hescheler, J.; Schneider, T.; Weiergraber, M. Atropine-sensitive hippocampal theta oscillations are mediated by Cav2.3 R-type Ca(2+) channels. Neuroscience 2012, 205, 125–139. [Google Scholar]
  73. Weiergraber, M.; Kamp, M.A.; Radhakrishnan, K.; Hescheler, J.; Schneider, T. The Ca(v)2.3 voltage-gated calcium channel in epileptogenesis—Shedding new light on an enigmatic channel. Neurosci. Biobehav. Rev. 2006, 30, 1122–1144. [Google Scholar] [CrossRef]
  74. Klugbauer, N.; Lacinova, L.; Marais, E.; Hobom, M.; Hofmann, F. Molecular diversity of the calcium channel alpha2delta subunit. J. Neurosci. 1999, 19, 684–691. [Google Scholar]
  75. Rettig, J.; Sheng, Z.H.; Kim, D.K.; Hodson, C.D.; Snutch, T.P.; Catterall, W.A. Isoform-specific interaction of the alpha1a subunits of brain Ca2+ channels with the presynaptic proteins syntaxin and snap-25. Proc. Natl. Acad. Sci. USA 1996, 93, 7363–7368. [Google Scholar] [CrossRef]
  76. Sheng, Z.H.; Yokoyama, C.T.; Catterall, W.A. Interaction of the synprint site of N-type Ca2+channels with the c2b domain of synaptotagmin I. Proc. Natl. Acad. Sci. USA 1997, 94, 5405–5410. [Google Scholar] [CrossRef]
  77. Verkhratsky, A.; Steinhauser, C. Ion channels in glial cells. Brain Res. Brain Res. Rev. 2000, 32, 380–412. [Google Scholar] [CrossRef]
  78. Moreno Davila, H. Molecular and functional diversity of voltage-gated calcium channels. Ann. N. Y. Acad. Sci. 1999, 868, 102–117. [Google Scholar] [CrossRef]
  79. Masumiya, H.; Shijuku, T.; Tanaka, H.; Shigenobu, K. Inhibition of myocardial L- and T-type Ca2+ currents by efonidipine: Possible mechanism for its chronotropic effect. Eur. J. Pharmacol. 1998, 349, 351–357. [Google Scholar] [CrossRef]
  80. Galetin, T.; Tevoufouet, E.E.; Sandmeyer, J.; Matthes, J.; Nguemo, F.; Hescheler, J.; Weiergraber, M.; Schneider, T. Pharmacoresistant Ca(v) 2.3 (E-type/R-type) voltage-gated calcium channels influence heart rate dynamics and may contribute to cardiac impulse conduction. Cell Biochem. Funct. 2012. [Google Scholar] [CrossRef]
  81. Olivera, B.M.; Gray, W.R.; Zeikus, R.; McIntosh, J.M.; Varga, J.; Rivier, J.; de Santos, V.; Cruz, L.J. Peptide neurotoxins from fish-hunting cone snails. Science 1985, 230, 1338–1343. [Google Scholar]
  82. Adams, D.J.; Berecki, G. Mechanisms of conotoxin inhibition of N-type (Ca2.2) calcium channels. Biochim. Biophys. Acta 2013, in press. [Google Scholar]
  83. D'Ascenzo, M.; Vairano, M.; Andreassi, C.; Navarra, P.; Azzena, G.B.; Grassi, C. Electrophysiological and molecular evidence of L-(Cav1), N- (Cav2.2), and R- (Cav2.3) type Ca2+ channels in rat cortical astrocytes. Glia 2004, 45, 354–363. [Google Scholar]
  84. Newcomb, R.; Szoke, B.; Palma, A.; Wang, G.; Chen, X.; Hopkins, W.; Cong, R.; Miller, J.; Urge, L.; Tarczy-Hornoch, K.; et al. Selective peptide antagonist of the class E calcium channel from the venom of the tarantula hysterocrates gigas. Biochemistry 1998, 37, 15353–15362. [Google Scholar] [CrossRef]
  85. Wang, G.; Dayanithi, G.; Newcomb, R.; Lemos, J.R. An R-type Ca(2+) current in neurohypophysial terminals preferentially regulates oxytocin secretion. J. Neurosci. 1999, 19, 9235–9241. [Google Scholar]
  86. Furukawa, T.; Nukada, T.; Miura, R.; Ooga, K.; Honda, M.; Watanabe, S.; Koganesawa, S.; Isshiki, T. Differential blocking action of dihydropyridine Ca2+ antagonists on a T-type Ca2+ channel (alpha1g) expressed in Xenopus oocytes. J. Cardiovasc. Pharmacol. 2005, 45, 241–246. [Google Scholar] [CrossRef]
  87. Clozel, J.P.; Ertel, E.A.; Ertel, S.I. Discovery and main pharmacological properties of mibefradil (RO 40-5967), the first selective T-type calcium channel blocker. J. Hypertens. Suppl. 1997, 15, S17–S25. [Google Scholar] [CrossRef]
  88. Bernal, J.; Lee, J.H.; Cribbs, L.L.; Perez-Reyes, E. Full reversal of Pb++ block of L-type Ca++ channels requires treatment with heavy metal antidotes. J. Pharmacol. Exp. Ther. 1997, 282, 172–180. [Google Scholar]
  89. Lee, J.H.; Gomora, J.C.; Cribbs, L.L.; Perez-Reyes, E. Nickel block of three cloned T-type calcium channels: Low concentrations selectively block alpha1h. Biophys. J. 1999, 77, 3034–3042. [Google Scholar] [CrossRef]
  90. Nelson, M.T.; Woo, J.; Kang, H.W.; Vitko, I.; Barrett, P.Q.; Perez-Reyes, E.; Lee, J.H.; Shin, H.S.; Todorovic, S.M. Reducing agents sensitize c-type nociceptors by relieving high-affinity zinc inhibition of T-type calcium channels. J. Neurosci. 2007, 27, 8250–8260. [Google Scholar] [CrossRef]
  91. Shcheglovitov, A.; Vitko, I.; Lazarenko, R.M.; Orestes, P.; Todorovic, S.M.; Perez-Reyes, E. Molecular and biophysical basis of glutamate and trace metal modulation of voltage-gated Ca(v)2.3 calcium channels. J. Gen. Physiol. 2012, 139, 219–234. [Google Scholar] [CrossRef]
  92. Siesjo, B.K. Calcium and cell death. Magnesium 1989, 8, 223–237. [Google Scholar]
  93. Siesjo, B.K.; Agardh, C.D.; Bengtsson, F. Free radicals and brain damage. Cerebrovasc. Brain Metab. Rev. 1989, 1, 165–211. [Google Scholar]
  94. Siesjo, B.K.; Bengtsson, F.; Grampp, W.; Theander, S. Calcium, excitotoxins, and neuronal death in the brain. Ann. N. Y. Acad. Sci. 1989, 568, 234–251. [Google Scholar] [CrossRef]
  95. Young, W. Role of calcium in central nervous system injuries. J Neurotrauma 1992, 9 (Suppl 1), S9–S25. [Google Scholar]
  96. Nilsson, P.; Laursen, H.; Hillered, L.; Hansen, A.J. Calcium movements in traumatic brain injury: The role of glutamate receptor-operated ion channels. J Cereb. Blood. Flow. Metab. 1996, 16, 262–270. [Google Scholar]
  97. Cater, H.L.; Gitterman, D.; Davis, S.M.; Benham, C.D.; Morrison, B., 3rd; Sundstrom, L.E. Stretch-induced injury in organotypic hippocampal slice cultures reproduces in vivo post-traumatic neurodegeneration: Role of glutamate receptors and voltage-dependent calcium channels. J. Neurochem. 2007, 101, 434–447. [Google Scholar]
  98. Floyd, C.L.; Gorin, F.A.; Lyeth, B.G. Mechanical strain injury increases intracellular sodium and reverses Na+/Ca2+ exchange in cortical astrocytes. Glia 2005, 51, 35–46. [Google Scholar] [CrossRef]
  99. Floyd, C.L.; Rzigalinski, B.A.; Weber, J.T.; Sitterding, H.A.; Willoughby, K.A.; Ellis, E.F. Traumatic injury of cultured astrocytes alters inositol (1,4,5)-trisphosphate-mediated signaling. Glia 2001, 33, 12–23. [Google Scholar] [CrossRef]
  100. LaPlaca, M.C.; Thibault, L.E. An in vitro traumatic injury model to examine the response of neurons to a hydrodynamically-induced deformation. Ann. Biomed. Eng. 1997, 25, 665–677. [Google Scholar] [CrossRef]
  101. Rzigalinski, B.A.; Weber, J.T.; Willoughby, K.A.; Ellis, E.F. Intracellular free calcium dynamics in stretch-injured astrocytes. J. Neurochem. 1998, 70, 2377–2385. [Google Scholar]
  102. Shahlaie, K.; Lyeth, B.G.; Gurkoff, G.G.; Muizelaar, J.P.; Berman, R.F. Neuroprotective effects of selective N-type VGCC blockade on stretch-injury-induced calcium dynamics in cortical neurons. J. Neurotrauma 2010, 27, 175–187. [Google Scholar] [CrossRef]
  103. Staal, J.A.; Dickson, T.C.; Gasperini, R.; Liu, Y.; Foa, L.; Vickers, J.C. Initial calcium release from intracellular stores followed by calcium dysregulation is linked to secondary axotomy following transient axonal stretch injury. J. Neurochem. 2010, 112, 1147–1155. [Google Scholar] [CrossRef]
  104. Weber, J.T.; Rzigalinski, B.A.; Ellis, E.F. Traumatic injury of cortical neurons causes changes in intracellular calcium stores and capacitative calcium influx. J. Biol. Chem. 2001, 276, 1800–1807. [Google Scholar] [CrossRef]
  105. Weber, J.T.; Rzigalinski, B.A.; Willoughby, K.A.; Moore, S.F.; Ellis, E.F. Alterations in calcium-mediated signal transduction after traumatic injury of cortical neurons. Cell Calcium 1999, 26, 289–299. [Google Scholar] [CrossRef]
  106. Wolf, J.A.; Stys, P.K.; Lusardi, T.; Meaney, D.; Smith, D.H. Traumatic axonal injury induces calcium influx modulated by tetrodotoxin-sensitive sodium channels. J. Neurosci. 2001, 21, 1923–1930. [Google Scholar]
  107. Young, W.; Koreh, I. Potassium and calcium changes in injured spinal cords. Brain Res. 1986, 365, 42–53. [Google Scholar] [CrossRef]
  108. Ahmed, S.M.; Weber, J.T.; Liang, S.; Willoughby, K.A.; Sitterding, H.A.; Rzigalinski, B.A.; Ellis, E.F. Nmda receptor activation contributes to a portion of the decreased mitochondrial membrane potential and elevated intracellular free calcium in strain-injured neurons. J. Neurotrauma 2002, 19, 1619–1629. [Google Scholar] [CrossRef]
  109. Geddes, D.M.; LaPlaca, M.C.; Cargill, R.S., 2nd. Susceptibility of hippocampal neurons to mechanically induced injury. Exp. Neurol. 2003, 184, 420–427. [Google Scholar] [CrossRef]
  110. Geddes-Klein, D.M.; Serbest, G.; Mesfin, M.N.; Cohen, A.S.; Meaney, D.F. Pharmacologically induced calcium oscillations protect neurons from increases in cytosolic calcium after trauma. J. Neurochem. 2006, 97, 462–474. [Google Scholar] [CrossRef]
  111. LaPlaca, M.C.; Thibault, L.E. Dynamic mechanical deformation of neurons triggers an acute calcium response and cell injury involving the n-methyl-d-aspartate glutamate receptor. J. Neurosci. Res. 1998, 52, 220–229. [Google Scholar] [CrossRef]
  112. Shahlaie, K.; Gurkoff, G.G.; Lyeth, B.G.; Muizelaar, J.P.; Berman, R.F. Neuroprotective effects of SNX-185 in an in vitro model of TBI with a second insult. Restor. Neurol. Neurosci. 2013, 31, 141–153. [Google Scholar]
  113. Chen, T.; Willoughby, K.A.; Ellis, E.F. Group I metabotropic receptor antagonism blocks depletion of calcium stores and reduces potentiated capacitative calcium entry in strain-injured neurons and astrocytes. J. Neurotrauma 2004, 21, 271–281. [Google Scholar] [CrossRef]
  114. Buki, A.; Okonkwo, D.O.; Wang, K.K.; Povlishock, J.T. Cytochrome C release and caspase activation in traumatic axonal injury. J. Neurosci. 2000, 20, 2825–2834. [Google Scholar]
  115. Conti, A.C.; Raghupathi, R.; Trojanowski, J.Q.; McIntosh, T.K. Experimental brain injury induces regionally distinct apoptosis during the acute and delayed post-traumatic period. J. Neurosci. 1998, 18, 5663–5672. [Google Scholar]
  116. Eldadah, B.A.; Faden, A.I. Caspase pathways, neuronal apoptosis, and CNS injury. J. Neurotrauma 2000, 17, 811–829. [Google Scholar] [CrossRef]
  117. Keane, R.W.; Kraydieh, S.; Lotocki, G.; Alonso, O.F.; Aldana, P.; Dietrich, W.D. Apoptotic and antiapoptotic mechanisms after traumatic brain injury. J. Cereb. Blood. Flow. Metab. 2001, 21, 1189–1198. [Google Scholar]
  118. Newcomb, J.K.; Zhao, X.; Pike, B.R.; Hayes, R.L. Temporal profile of apoptotic-like changes in neurons and astrocytes following controlled cortical impact injury in the rat. Exp. Neurol. 1999, 158, 76–88. [Google Scholar] [CrossRef]
  119. Raghupathi, R.; Graham, D.I.; McIntosh, T.K. Apoptosis after traumatic brain injury. J. Neurotrauma 2000, 17, 927–938. [Google Scholar] [CrossRef]
  120. Robertson, C.L.; Puskar, A.; Hoffman, G.E.; Murphy, A.Z.; Saraswati, M.; Fiskum, G. Physiologic progesterone reduces mitochondrial dysfunction and hippocampal cell loss after traumatic brain injury in female rats. Exp. Neurol. 2006, 197, 235–243. [Google Scholar] [CrossRef]
  121. Robertson, C.L.; Saraswati, M.; Fiskum, G. Mitochondrial dysfunction early after traumatic brain injury in immature rats. J. Neurochem. 2007, 101, 1248–1257. [Google Scholar] [CrossRef]
  122. Singh, I.N.; Sullivan, P.G.; Deng, Y.; Mbye, L.H.; Hall, E.D. Time course of post-traumatic mitochondrial oxidative damage and dysfunction in a mouse model of focal traumatic brain injury: Implications for neuroprotective therapy. J. Cereb. Blood. Flow. Metab. 2006, 26, 1407–1418. [Google Scholar] [CrossRef]
  123. White, R.J.; Reynolds, I.J. Mitochondria accumulate Ca2+ following intense glutamate stimulation of cultured rat forebrain neurones. J. Physiol. 1997, 498 (Pt 1), 31–47. [Google Scholar]
  124. Xiong, Y.; Gu, Q.; Peterson, P.L.; Muizelaar, J.P.; Lee, C.P. Mitochondrial dysfunction and calcium perturbation induced by traumatic brain injury. J. Neurotrauma 1997, 14, 23–34. [Google Scholar] [CrossRef]
  125. Mustafa, A.G.; Singh, I.N.; Wang, J.; Carrico, K.M.; Hall, E.D. Mitochondrial protection after traumatic brain injury by scavenging lipid peroxyl radicals. J. Neurochem 2010, 114, 271–280. [Google Scholar]
  126. Braughler, J.M.; Duncan, L.A.; Goodman, T. Calcium enhances in vitro free radical-induced damage to brain synaptosomes, mitochondria, and cultured spinal cord neurons. J. Neurochem 1985, 45, 1288–1293. [Google Scholar] [CrossRef]
  127. Braughler, J.M.; Hall, E.D. Involvement of lipid peroxidation in CNS injury. J. Neurotrauma 1992, 9 (Suppl 1), S1–S7. [Google Scholar]
  128. Odland, R.M.; Sutton, R.L. Hyperosmosis of cerebral injury. Neurological Res. 1999, 21, 500–508. [Google Scholar]
  129. Tavalin, S.J.; Ellis, E.F.; Satin, L.S. Mechanical perturbation of cultured cortical neurons reveals a stretch-induced delayed depolarization. J. Neurophysiol. 1995, 74, 2767–2773. [Google Scholar]
  130. Geddes-Klein, D.M.; Schiffman, K.B.; Meaney, D.F. Mechanisms and consequences of neuronal stretch injury in vitro differ with the model of trauma. J. Neurotrauma 2006, 23, 193–204. [Google Scholar] [CrossRef]
  131. Lusardi, T.A.; Wolf, J.A.; Putt, M.E.; Smith, D.H.; Meaney, D.F. Effect of acute calcium influx after mechanical stretch injury in vitro on the viability of hippocampal neurons. J. Neurotrauma 2004, 21, 61–72. [Google Scholar] [CrossRef]
  132. Rzigalinski, B.A.; Liang, S.; McKinney, J.S.; Willoughby, K.A.; Ellis, E.F. Effect of Ca2+ on in vitro astrocyte injury. J. Neurochem. 1997, 68, 289–296. [Google Scholar]
  133. Goforth, P.B.; Ellis, E.F.; Satin, L.S. Mechanical injury modulates AMPA receptor kinetics via an NMDA receptor-dependent pathway. J. Neurotrauma 2004, 21, 719–732. [Google Scholar] [CrossRef]
  134. Goforth, P.B.; Ellis, E.F.; Satin, L.S. Enhancement of AMPA-mediated current after traumatic injury in cortical neurons. J. Neurosci 1999, 19, 7367–7374. [Google Scholar]
  135. Goforth, P.B.; Ren, J.; Schwartz, B.S.; Satin, L.S. Excitatory synaptic transmission and network activity are depressed following mechanical injury in cortical neurons. J. Neurophys. 2011, 105, 2350–2363. [Google Scholar] [CrossRef]
  136. Spaethling, J.; Le, L.; Meaney, D.F. NMDA receptor mediated phosphorylation of GluR1 subunits contributes to the appearance of calcium-permeable AMPA receptors after mechanical stretch injury. Neurobiol. Dis. 2012, 46, 646–654. [Google Scholar] [CrossRef]
  137. Spaethling, J.M.; Klein, D.M.; Singh, P.; Meaney, D.F. Calcium-permeable AMPA receptors appear in cortical neurons after traumatic mechanical injury and contribute to neuronal fate. J. Neurotrauma 2008, 25, 1207–1216. [Google Scholar] [CrossRef]
  138. Gurkoff, G.G.; Shahlaie, K.; Lyeth, B.G. In vitro mechanical strain trauma alters neuronal calcium responses: Implications for posttraumatic epilepsy. Epilepsia 2012, 53 (Suppl 1), 53–60. [Google Scholar]
  139. DeWitt, D.S.; Jenkins, L.W.; Prough, D.S. Enhanced vulnerability to secondary ischemic insults after experimental traumatic brain injury. New. Horiz. 1995, 3, 376–383. [Google Scholar]
  140. Bullock, R.; Zauner, A.; Woodward, J.J.; Myseros, J.; Choi, S.C.; Ward, J.D.; Marmarou, A.; Young, H.F. Factors affecting excitatory amino acid release following severe human head injury. J. Neurosurg. 1998, 89, 507–518. [Google Scholar] [CrossRef]
  141. Katayama, Y.; Kawamata, T.; Tamura, T.; Hovda, D.A.; Becker, D.P.; Tsubokawa, T. Calcium-dependent glutamate release concomitant with massive potassium flux during cerebral ischemia in vivo. Brain Res. 1991, 558, 136–140. [Google Scholar] [CrossRef]
  142. Kochanek, P.M. Ischemic and traumatic brain injury: Pathobiology and cellular mechanisms. Crit. Care Med. 1993, 21, S333–S335. [Google Scholar] [CrossRef]
  143. Engel, D.C.; Slemmer, J.E.; Vlug, A.S.; Maas, A.I.; Weber, J.T. Combined effects of mechanical and ischemic injury to cortical cells: Secondary ischemia increases damage and decreases effects of neuroprotective agents. Neuropharmacology 2005, 49, 985–995. [Google Scholar]
  144. Stokes, B.T.; Fox, P.; Hollinden, G. Extracellular calcium activity in the injured spinal cord. Exp. Neurol. 1983, 80, 561–572. [Google Scholar] [CrossRef]
  145. Nilsson, P.; Hillered, L.; Olsson, Y.; Sheardown, M.J.; Hansen, A.J. Regional changes in interstitial K+ and Ca2+ levels following cortical compression contusion trauma in rats. J. Cereb. Blood Flow Metab. 1993, 13, 183–192. [Google Scholar] [CrossRef]
  146. Dienel, G.A. Regional accumulation of calcium in postischemic rat brain. J. Neurochem. 1984, 43, 913–925. [Google Scholar] [CrossRef]
  147. Fineman, I.; Hovda, D.A.; Smith, M.; Yoshino, A.; Becker, D.P. Concussive brain injury is associated with a prolonged accumulation of calcium: A 45Ca autoradiographic study. Brain Res. 1993, 624, 94–102. [Google Scholar] [CrossRef]
  148. Osteen, C.L.; Moore, A.H.; Prins, M.L.; Hovda, D.A. Age-dependency of 45calcium accumulation following lateral fluid percussion: Acute and delayed patterns. J. Neurotrauma 2001, 18, 141–162. [Google Scholar] [CrossRef]
  149. Deshpande, L.S.; Sun, D.A.; Sombati, S.; Baranova, A.; Wilson, M.S.; Attkisson, E.; Hamm, R.J.; DeLorenzo, R.J. Alterations in neuronal calcium levels are associated with cognitive deficits after traumatic brain injury. Neurosci. Lett. 2008, 441, 115–119. [Google Scholar] [CrossRef]
  150. Sun, D.A.; Deshpande, L.S.; Sombati, S.; Baranova, A.; Wilson, M.S.; Hamm, R.J.; DeLorenzo, R.J. Traumatic brain injury causes a long-lasting calcium (Ca2+)-plateau of elevated intracellular Ca levels and altered Ca2+ homeostatic mechanisms in hippocampal neurons surviving brain injury. Eur. J. Neurosci. 2008, 27, 1659–1672. [Google Scholar] [CrossRef]
  151. Hell, J.W.; Westenbroek, R.E.; Warner, C.; Ahlijanian, M.K.; Prystay, W.; Gilbert, M.M.; Snutch, T.P.; Catterall, W.A. Identification and differential subcellular localization of the neuronal class C and class D L-type calcium channel alpha 1 subunits. J. Cell. Biol. 1993, 123, 949–962. [Google Scholar] [CrossRef]
  152. Hoffmeister, F.; Benz, U.; Heise, A.; Krause, H.P.; Neuser, V. Behavioral effects of nimodipine in animals. Arzneimittel Forschung 1982, 32, 347–360. [Google Scholar]
  153. Lazarewicz, J.W.; Pluta, R.; Puka, M.; Salinska, E. Diverse mechanisms of neuronal protection by nimodipine in experimental rabbit brain ischemia. Stroke 1990, 21, IV108–110. [Google Scholar]
  154. Symon, L.; Harris, R.J.; Branston, N.M. Calcium ions and calcium antagonists in ischaemia. Acta Neurochir. (Wien) 1982, 63, 267–275. [Google Scholar]
  155. Babu, C.S.; Ramanathan, M. Post-ischemic administration of nimodipine following focal cerebral ischemic-reperfusion injury in rats alleviated excitotoxicity, neurobehavioural alterations and partially the bioenergetics. Int. J. Dev. Neurosci. 2011, 29, 93–105. [Google Scholar] [CrossRef]
  156. Steen, P.A.; Gisvold, S.E.; Milde, J.H.; Newberg, L.A.; Scheithauer, B.W.; Lanier, W.L.; Michenfelder, J.D. Nimodipine improves outcome when given after complete cerebral ischemia in primates. Anesthesiology 1985, 62, 406–414. [Google Scholar] [CrossRef]
  157. Gaab, M.R.; Haubitz, I.; Brawanski, A.; Korn, A.; Czech, T. Acute effects of nimodipine on the cerebral blood flow and intracranial pressure. Neurochirurgia 1985, 28 (Suppl 1), 93–99. [Google Scholar]
  158. Gelmers, H.J. The effects of nimodipine on the clinical course of patients with acute ischemic stroke. Acta Neurol. Scand. 1984, 69, 232–239. [Google Scholar] [CrossRef]
  159. Paci, A.; Ottaviano, P.; Trenta, A.; Iannone, G.; de Santis, L.; Lancia, G.; Moschini, E.; Carosi, M.; Amigoni, S.; Caresia, L. Nimodipine in acute ischemic stroke: A double-blind controlled study. Acta Neurol. Scand. 1989, 80, 282–286. [Google Scholar] [CrossRef]
  160. Krieglstein, J.; Lippert, K.; Poch, G. Apparent independent action of nimodipine and glutamate antagonists to protect cultured neurons against glutamate-induced damage. Neuropharmacology 1996, 35, 1737–1742. [Google Scholar] [CrossRef]
  161. Maeda, T.; Lee, S.M.; Hovda, D.A. Restoration of cerebral vasoreactivity by an l-type calcium channel blocker following fluid percussion brain injury. J. Neurotrauma 2005, 22, 763–771. [Google Scholar] [CrossRef]
  162. Ercan, M.; Inci, S.; Kilinc, K.; Palaoglu, S.; Aypar, U. Nimodipine attenuates lipid peroxidation during the acute phase of head trauma in rats. Neurosurg. Rev. 2001, 24, 127–130. [Google Scholar] [CrossRef]
  163. Bailey, I.; Bell, A.; Gray, J.; Gullan, R.; Heiskanan, O.; Marks, P.V.; Marsh, H.; Mendelow, D.A.; Murray, G.; Ohman, J.; et al. A trial of the effect of nimodipine on outcome after head injury. Acta Neurochir. (Wien.) 1991, 110, 97–105. [Google Scholar]
  164. Teasdale, G.; Bailey, I.; Bell, A.; Gray, J.; Gullan, R.; Heiskanan, O.; Marks, P.V.; Marsh, H.; Mendelow, D.A.; Murray, G.; et al. A randomized trial of nimodipine in severe head injury: Hit I. British/finnish co-operative head injury trial group. J. Neurotrauma 1992, 9, S545–S550. [Google Scholar]
  165. Group, E.S. A multicenter trial of the efficacy of nimodipine on outcome after severe head injury. The european study group on nimodipine in severe head injury. J. Neurosurg. 1994, 80, 797–804. [Google Scholar] [CrossRef]
  166. Murray, G.D.; Teasdale, G.M.; Schmitz, H. Nimodipine in traumatic subarachnoid haemorrhage: A re-analysis of the hit I and hit II trials. Acta Neurochir. (Wien.) 1996, 138, 1163–1167. [Google Scholar]
  167. Vergouwen, M.D.; Vermeulen, M.; Roos, Y.B. Effect of nimodipine on outcome in patients with traumatic subarachnoid haemorrhage: A systematic review. Lancet Neurol. 2006, 5, 1029–1032. [Google Scholar] [CrossRef]
  168. Aslan, A.; Gurelik, M.; Cemek, M.; Buyukokuroglu, M.; Goksel, H.M.; Eser, O. Nimodipine can diminish oxidative stress in patients with severe head trauma. J. Neurosurg. Sci. 2012, 56, 247–253. [Google Scholar]
  169. Aslan, A.; Gurelik, M.; Cemek, M.; Goksel, H.M.; Buyukokuroglu, M.E. Nimodipine can improve cerebral metabolism and outcome in patients with severe head trauma. Pharm. Res. 2009, 59, 120–124. [Google Scholar] [CrossRef]
  170. Langham, J.; Goldfrad, C.; Teasdale, G.; Shaw, D.; Rowan, K. Calcium channel blockers for acute traumatic brain injury. Cochrane Database Syst. Rev. 2003, CD000565. [Google Scholar]
  171. Williams, M.E.; Brust, P.F.; Feldman, D.H.; Patthi, S.; Simerson, S.; Maroufi, A.; McCue, A.F.; Velicelebi, G.; Ellis, S.B.; Harpold, M.M. Structure and functional expression of an omega-conotoxin-sensitive human N-type calcium channel. Science 1992, 257, 389–395. [Google Scholar]
  172. Dubel, S.J.; Starr, T.V.; Hell, J.; Ahlijanian, M.K.; Enyeart, J.J.; Catterall, W.A.; Snutch, T.P. Molecular cloning of the alpha-1 subunit of an omega-conotoxin-sensitive calcium channel. Proc. Natl. Acad. Sci. USA 1992, 89, 5058–5062. [Google Scholar]
  173. Volsen, S.G.; Day, N.C.; McCormack, A.L.; Smith, W.; Craig, P.J.; Beattie, R.; Ince, P.G.; Shaw, P.J.; Ellis, S.B.; Gillespie, A.; et al. The expression of neuronal voltage-dependent calcium channels in human cerebellum. Brain Res. Mol. Brain Res. 1995, 34, 271–282. [Google Scholar] [CrossRef]
  174. Reynolds, I.J.; Wagner, J.A.; Snyder, S.H.; Thayer, S.A.; Olivera, B.M.; Miller, R.J. Brain voltage-sensitive calcium channel subtypes differentiated by omega-conotoxin fraction gvia. Proc. Natl. Acad. Sci. USA 1986, 83, 8804–8807. [Google Scholar] [CrossRef]
  175. Hirning, L.D.; Fox, A.P.; McCleskey, E.W.; Olivera, B.M.; Thayer, S.A.; Miller, R.J.; Tsien, R.W. Dominant role of N-type Ca2+ channels in evoked release of norepinephrine from sympathetic neurons. Science 1988, 239, 57–61. [Google Scholar]
  176. Bowersox, S.; Mandema, J.; Tarczy-Hornoch, K.; Miljanich, G.; Luther, R.R. Pharmacokinetics of SNX-111, a selective N-type calcium channel blocker, in rats and cynomolgus monkeys. Drug Metab. Dispos. 1997, 25, 379–383. [Google Scholar]
  177. Bowersox, S.S.; Gadbois, T.; Singh, T.; Pettus, M.; Wang, Y.X.; Luther, R.R. Selective N-type neuronal voltage-sensitive calcium channel blocker, SNX-111, produces spinal antinociception in rat models of acute, persistent and neuropathic pain. J. Pharmacol. Exp. Ther. 1996, 279, 1243–1249. [Google Scholar]
  178. Bowersox, S.S.; Luther, R. Pharmacotherapeutic potential of omega-conotoxin MVIIA (SNX-111), an N-type neuronal calcium channel blocker found in the venom of conus magus. Toxicon 1998, 36, 1651–1658. [Google Scholar]
  179. Berman, R.F.; Verweij, B.H.; Muizelaar, J.P. Neurobehavioral protection by the neuronal calcium channel blocker ziconotide in a model of traumatic diffuse brain injury in rats. J. Neurosurg. 2000, 93, 821–828. [Google Scholar] [CrossRef]
  180. Colbourne, F.; Li, H.; Buchan, A.M.; Clemens, J.A. Continuing postischemic neuronal death in CA1: Influence of ischemia duration and cytoprotective doses of NBQX and SNX-111 in rats. Stroke 1999, 30, 662–668. [Google Scholar] [CrossRef]
  181. Perez-Pinzon, M.A.; Yenari, M.A.; Sun, G.H.; Kunis, D.M.; Steinberg, G.K. SNX-111, a novel, presynaptic N-type calcium channel antagonist, is neuroprotective against focal cerebral ischemia in rabbits. J. Neurol. Sci. 1997, 153, 25–31. [Google Scholar] [CrossRef]
  182. Valentino, K.; Newcomb, R.; Gadbois, T.; Singh, T.; Bowersox, S.; Bitner, S.; Justice, A.; Yamashiro, D.; Hoffman, B.B.; Ciaranello, R.; et al. A selective N-type calcium channel antagonist protects against neuronal loss after global cerebral ischemia. Proc. Natl. Acad. Sci. USA 1993, 90, 7894–7897. [Google Scholar] [CrossRef]
  183. Hovda, D.A.; Fu, K.; Badie, H.; Samii, A.; Pinanong, P.; Becker, D.P. Administration of an omega-conopeptide one hour following traumatic brain injury reduces 45calcium accumulation. Acta Neurochir. Suppl. (Wien.) 1994, 60, 521–523. [Google Scholar]
  184. Samii, A.; Badie, H.; Fu, K.; Luther, R.R.; Hovda, D.A. Effects of an N-type calcium channel antagonist (SNX-111; ziconotide) on calcium-45 accumulation following fluid-percussion injury. J. Neurotrauma 1999, 16, 879–892. [Google Scholar] [CrossRef]
  185. Verweij, B.H.; Muizelaar, J.P.; Vinas, F.C.; Peterson, P.L.; Xiong, Y.; Lee, C.P. Improvement in mitochondrial dysfunction as a new surrogate efficiency measure for preclinical trials: Dose-response and time-window profiles for administration of the calcium channel blocker ziconotide in experimental brain injury. J. Neurosurg. 2000, 93, 829–834. [Google Scholar]
  186. Verweij, B.H.; Muizelaar, J.P.; Vinas, F.C.; Peterson, P.L.; Xiong, Y.; Lee, C.P. Mitochondrial dysfunction after experimental and human brain injury and its possible reversal with a selective N-type calcium channel antagonist (SNX-111). Neurol. Res. 1997, 19, 334–339. [Google Scholar]
  187. Xiong, Y.; Peterson, P.L.; Verweij, B.H.; Vinas, F.C.; Muizelaar, J.P.; Lee, C.P. Mitochondrial dysfunction after experimental traumatic brain injury: Combined efficacy of SNX-111 and U-101033E. J. Neurotrauma 1998, 15, 531–544. [Google Scholar] [CrossRef]
  188. Lee, L.L.; Galo, E.; Lyeth, B.G.; Muizelaar, J.P.; Berman, R.F. Neuroprotection in the rat lateral fluid percussion model of traumatic brain injury by SNX-185, an N-type voltage-gated calcium channel blocker. Exp. Neurol. 2004, 190, 70–78. [Google Scholar]
  189. Hinzman, J.M.; Thomas, T.C.; Quintero, J.E.; Gerhardt, G.A.; Lifshitz, J. Disruptions in the regulation of extracellular glutamate by neurons and glia in the rat striatum two days after diffuse brain injury. J. Neurotrauma 2012, 29, 1197–1208. [Google Scholar] [CrossRef]
  190. Pruneau, D.; Angus, J.A. Omega-conotoxin GVIA, the N-type calcium channel inhibitor, is sympatholytic but not vagolytic: Consequences for hemodynamics and autonomic reflexes in conscious rabbits. J. Cardiovasc. Pharmacol. 1990, 16, 675–680. [Google Scholar] [CrossRef]
  191. Wright, C.E.; Angus, J.A. Prolonged cardiovascular effects of the N-type Ca2+ channel antagonist omega-conotoxin GVIA in conscious rabbits. J. Cardiovasc. Pharmacol. 1997, 30, 392–399. [Google Scholar] [CrossRef]
  192. Malmberg, A.B.; Yaksh, T.L. Effect of continuous intrathecal infusion of omega-conopeptides, N-type calcium-channel blockers, on behavior and antinociception in the formalin and hot-plate tests in rats. Pain 1995, 60, 83–90. [Google Scholar] [CrossRef]
  193. Penn, R.D.; Paice, J.A. Adverse effects associated with the intrathecal administration of ziconotide. Pain 2000, 85, 291–296. [Google Scholar] [CrossRef]
  194. Atanassoff, P.G.; Hartmannsgruber, M.W.; Thrasher, J.; Wermeling, D.; Longton, W.; Gaeta, R.; Singh, T.; Mayo, M.; McGuire, D.; Luther, R.R. Ziconotide, a new N-type calcium channel blocker, administered intrathecally for acute postoperative pain. Reg. Anesth. Pain Med. 2000, 25, 274–278. [Google Scholar]
  195. Staats, P.S.; Yearwood, T.; Charapata, S.G.; Presley, R.W.; Wallace, M.S.; Byas-Smith, M.; Fisher, R.; Bryce, D.A.; Mangieri, E.A.; Luther, R.R.; et al. Intrathecal ziconotide in the treatment of refractory pain in patients with cancer or aids: A randomized controlled trial. JAMA 2004, 291, 63–70. [Google Scholar] [CrossRef]
  196. Miljanich, G.P. Ziconotide: Neuronal calcium channel blocker for treating severe chronic pain. Curr. Med. Chem. 2004, 11, 3029–3040. [Google Scholar] [CrossRef]
  197. Smith, H.S.; Deer, T.R. Safety and efficacy of intrathecal ziconotide in the management of severe chronic pain. Ther. Clin. Risk Manag. 2009, 5, 521–534. [Google Scholar] [CrossRef]
  198. Webster, L.R.; Fisher, R.; Charapata, S.; Wallace, M.S. Long-term intrathecal ziconotide for chronic pain: An open-label study. J. Pain Symptom Manage. 2009, 37, 363–372. [Google Scholar] [CrossRef]
  199. Wheaton, P.; Mathias, J.L.; Vink, R. Impact of pharmacological treatments on outcome in adult rodents after traumatic brain injury: A meta-analysis. J. Psychopharm. 2011, 25, 1581–1599. [Google Scholar] [CrossRef]
  200. Ishiguro, M.; Wellman, T.L.; Honda, A.; Russell, S.R.; Tranmer, B.I.; Wellman, G.C. Emergence of a R-type Ca2+ channel (Cav 2.3) contributes to cerebral artery constriction after subarachnoid hemorrhage. Circ. Res. 2005, 96, 419–426. [Google Scholar]
  201. Wang, F.; Yin, Y.H.; Jia, F.; Jiang, J.Y. Antagonism of r-type calcium channels significantly improves cerebral blood flow after subarachnoid hemorrhage in rats. J. Neurotrauma 2010, 27, 1723–1732. [Google Scholar] [CrossRef]
  202. Hansen, P.B. Functional and pharmacological consequences of the distribution of voltage-gated calcium channels in the renal blood vessels. Acta Physiol. 2013, 207, 690–699. [Google Scholar] [CrossRef]
  203. Hansen, P.B.; Jensen, B.L.; Andreasen, D.; Friis, U.G.; Skott, O. Vascular smooth muscle cells express the alpha(1a) subunit of a P-/Q-type voltage-dependent Ca(2+)channel, and it is functionally important in renal afferent arterioles. Circ. Res. 2000, 87, 896–902. [Google Scholar] [CrossRef]
  204. Hansen, P.B.; Poulsen, C.B.; Walter, S.; Marcussen, N.; Cribbs, L.L.; Skott, O.; Jensen, B.L. Functional importance of L- and P/Q-type voltage-gated calcium channels in human renal vasculature. Hypertension 2011, 58, 464–470. [Google Scholar] [CrossRef]
  205. Canova, D.; Roatta, S.; Micieli, G.; Bosone, D. Cerebral oxygenation and haemodynamic effects induced by nimodipine in healthy subjects. Funct. Neurol. 2012, 27, 169–176. [Google Scholar]

Share and Cite

MDPI and ACS Style

Gurkoff, G.; Shahlaie, K.; Lyeth, B.; Berman, R. Voltage-Gated Calcium Channel Antagonists and Traumatic Brain Injury. Pharmaceuticals 2013, 6, 788-812. https://doi.org/10.3390/ph6070788

AMA Style

Gurkoff G, Shahlaie K, Lyeth B, Berman R. Voltage-Gated Calcium Channel Antagonists and Traumatic Brain Injury. Pharmaceuticals. 2013; 6(7):788-812. https://doi.org/10.3390/ph6070788

Chicago/Turabian Style

Gurkoff, Gene, Kiarash Shahlaie, Bruce Lyeth, and Robert Berman. 2013. "Voltage-Gated Calcium Channel Antagonists and Traumatic Brain Injury" Pharmaceuticals 6, no. 7: 788-812. https://doi.org/10.3390/ph6070788

Article Metrics

Back to TopTop