Novel therapies in development for the treatment of traumatic brain injury

12
Review 2002 © Ashley Publications Ltd ISSN 1354-3784 1375 Ashley Publications www.ashley-pub.com 1. Introduction 2. Secondary injury 3. Attenuating secondary injury 4. Conclusion 5. Expert opinion Monthly Focus: Central & Peripheral Nervous System Novel therapies in development for the treatment of traumatic brain injury Robert Vink 1† and Alan J Nimmo 2 1† Department of Pathology, The University of Adelaide, Australia & 2 School of Pharmacy and Molecular Sciences, James Cook University, Australia In industrialised countries, the mean per capita incidence of traumatic brain injury (TBI) that results in a hospital presentation is 250 per 100,000. In Europe and North America alone, this translates to > 2 million TBI presenta- tions annually. Approximately 25% of these presentations are admitted for hospitalisation. Despite the significance of these figures, there is no single interventional pharmacotherapy that has shown efficacy in the treatment of clinical TBI. This lack of efficacy in clinical trials may be due, in part, to the inherent heterogeneity of the traumatic brain injury population. However, it is the multifactorial nature of secondary injury that also poses a major hurdle, particularly for those therapies that have been designed to specifi- cally target an individual injury factor. It is now becoming increasingly rec- ognised that any successful TBI therapy may have to simultaneously affect multiple injury factors, somewhat analogous to other broad spectrum inter- ventions. Recent efforts in experimental TBI have therefore focussed on developing novel pharmacotherapies that may affect multiple injury factors and thus improve the likelihood of a successful outcome. While a number of interventions are noteworthy in this regard, this review will focus on three novel compounds that show particular promise: magnesium, substance P antagonists and cyclosporin A. Keywords: cyclosporin A, head injury, magnesium, neuropeptides, neuroprotection, neurotrauma, substance P Expert Opin. Investig. Drugs (2002) 11(10):1375-1386 1. Introduction Traumatic brain injury (TBI) is the biggest killer of individuals under 44 years of age. Those individuals who survive TBI are often left with permanent neurological defi- cits that adversely affect their quality of life and, in many instances, prevent their return to the workforce. It is now recognised that neuronal cell death resulting from TBI is caused through both primary and secondary injury mechanisms. Primary injury occurs at the time of the traumatic event and may generally be considered to be the mechanical disruption caused by TBI, including, amongst others, laceration, contusion, shearing and axonal stretching. By its nature, primary injury can only be prevented and tremendous advances have been made in this regard with the use of safety devices to prevent TBI, including seatbelts, airbags and helmets. Secondary injury, on the other hand, is the delayed biochemical and physiological events that are initiated by the primary traumatic event, but manifest over the ensuing minutes to days afterward. It is this delayed injury cascade that is thought to be associated with the development of many of the neurological deficits observed after TBI. Since this secondary injury is delayed, there exists a window of opportunity to identify the factors that make up the secondary injury cascade and treat with ‘antifactors’ that

Transcript of Novel therapies in development for the treatment of traumatic brain injury

Review

Ashley Publicationswww.ashley-pub.com

1. Introduction

2. Secondary injury

3. Attenuating secondary injury

4. Conclusion

5. Expert opinion

Monthly Focus: Central & Peripheral Nervous System

Novel therapies in development for the treatment of traumatic brain injuryRobert Vink1† and Alan J Nimmo2

1†Department of Pathology, The University of Adelaide, Australia & 2School of Pharmacy and Molecular Sciences, James Cook University, Australia

In industrialised countries, the mean per capita incidence of traumatic braininjury (TBI) that results in a hospital presentation is 250 per 100,000. InEurope and North America alone, this translates to > 2 million TBI presenta-tions annually. Approximately 25% of these presentations are admitted forhospitalisation. Despite the significance of these figures, there is no singleinterventional pharmacotherapy that has shown efficacy in the treatmentof clinical TBI. This lack of efficacy in clinical trials may be due, in part, to theinherent heterogeneity of the traumatic brain injury population. However,it is the multifactorial nature of secondary injury that also poses a majorhurdle, particularly for those therapies that have been designed to specifi-cally target an individual injury factor. It is now becoming increasingly rec-ognised that any successful TBI therapy may have to simultaneously affectmultiple injury factors, somewhat analogous to other broad spectrum inter-ventions. Recent efforts in experimental TBI have therefore focussed ondeveloping novel pharmacotherapies that may affect multiple injury factorsand thus improve the likelihood of a successful outcome. While a number ofinterventions are noteworthy in this regard, this review will focus on threenovel compounds that show particular promise: magnesium, substance Pantagonists and cyclosporin A.

Keywords: cyclosporin A, head injury, magnesium, neuropeptides, neuroprotection, neurotrauma, substance P

Expert Opin. Investig. Drugs (2002) 11(10):1375-1386

1. Introduction

Traumatic brain injury (TBI) is the biggest killer of individuals under 44 years of age.Those individuals who survive TBI are often left with permanent neurological defi-cits that adversely affect their quality of life and, in many instances, prevent theirreturn to the workforce. It is now recognised that neuronal cell death resulting fromTBI is caused through both primary and secondary injury mechanisms. Primaryinjury occurs at the time of the traumatic event and may generally be considered tobe the mechanical disruption caused by TBI, including, amongst others, laceration,contusion, shearing and axonal stretching. By its nature, primary injury can only beprevented and tremendous advances have been made in this regard with the use ofsafety devices to prevent TBI, including seatbelts, airbags and helmets. Secondaryinjury, on the other hand, is the delayed biochemical and physiological events thatare initiated by the primary traumatic event, but manifest over the ensuing minutesto days afterward. It is this delayed injury cascade that is thought to be associatedwith the development of many of the neurological deficits observed after TBI. Sincethis secondary injury is delayed, there exists a window of opportunity to identify thefactors that make up the secondary injury cascade and treat with ‘antifactors’ that

2002 © Ashley Publications Ltd ISSN 1354-3784 1375

Novel therapies in development for the treatment of traumatic brain injury

could prevent, or at least attenuate, the injury process andresult in significant improvement in functional outcome [1].

Despite the identification of several promising candi-dates in experimental animal studies of TBI, the clinicaltrials that have attempted to pharmacologically inhibitindividual secondary injury factors have met with very lim-ited success. Interested readers are referred to some excel-lent analyses of these trials that have been reviewedelsewhere [2-4]. Various factors have been proposed toaccount for the lack of efficacy of these compounds anddiscussions of the inherent limitations of the experimentalanimal models, inadequate preclinical testing and hetero-geneity of the clinical TBI population can be found else-where [4,5]. However, it is the complexity of the secondaryinjury process and its role in the development of irreversi-ble brain injury that is also largely responsible for thisapparent failure to develop effective pharmacotherapies.Indeed, while a number of individual secondary injury fac-tors have been identified, their relationship to functionaloutcome and their inter-relationships with one another areyet to be fully characterised. Moreover, it is becomingincreasingly recognised that secondary injury is a multifac-torial process with different injury factors contributing tothe injury process at different time-points following theinsult. For example, it has long been recognised that theneuronal cell death that occurs in the first 24 h after TBIoccurs through a process of necrosis involving swelling ofmitochondria and other organelles and subsequent mem-brane degeneration [6]. Inhibition of necrosis was, there-fore, envisioned as being the aim of neuroprotectivetherapies. However, it has since been recognised that neu-ronal cell death after TBI can continue for days after TBIthrough a process known as apoptosis [7], whose identifyingfeatures include DNA condensation and fragmentation,cell shrinkage and the ultimate formation of apoptoticbodies [6]. What further complicates matters is the observa-tion that inhibition of necrosis has the potential to exacer-bate apoptosis and vice versa [8,9].

It is therefore unlikely that a single injury factor isresponsible for all of the observed functional deficits fol-lowing TBI. Rather, an intervention that simultaneouslyinhibits or at least attenuates a number of secondary injuryfactors is more likely to achieve a successful outcome. Fewinterventions have been identified that have effects on adiverse range of secondary injury factors. Nonetheless,three compounds with multifactorial effects have beenidentified in experimental brain injury studies and haveeither entered or are about to enter clinical trials in TBI.This review will briefly summarise some of the more rele-vant secondary injury mechanisms associated with thedevelopment of neuronal cell death following TBI. Theproperties that characterise magnesium, substance P antag-onists and cyclosporin A (CsA) as novel ‘broad spectrum’interventional pharmacotherapies for use in TBI, will thenbe discussed.

2. Secondary injury

2.1 ExcitotoxicityOne of the first recorded events after TBI is the release of exci-tatory neurotransmitters such as glutamate. Glutamate is foundin almost every brain region and in almost half of all brain syn-apses. Its release has been noted as early as minutes after experi-mental brain injury [10] with a maximum value being achieved10 min after injury. While experimental studies suggest that theglutamate concentration falls rapidly after this peak, increasedlevels during the first few days after severe clinical injury havebeen reported [11]. These sustained high levels of glutamateseem to be correlated more to structural damage rather thanvesicular release. Once released, glutamate can bind to anumber of different receptor subtypes, with the NMDA (N-methyl D-aspartate) receptor widely shown to be associatedwith neurotoxicity [12]. With significant quantities of glutamatereleased immediately after the traumatic event, binding to theNMDA receptor is known to promote substantial calciuminflux with resultant calcium overload. Not only does this causedepolarisation, the high intracellular calcium concentration isknown to activate a plethora of calcium-dependent enzymes,including proteases, lipases and endonucleases, whose uncon-trolled activation form part of an autodestructive cascade.Glutamate excitotoxicity is, however, not only mediated by theNMDA receptor, but also by the non-NMDA glutamate recep-tors [8]. Thus, glutamate excitotoxicity may involve a number ofdifferent mechanisms, each dependent on a different receptor.Such complexity makes effective pharmacological blockade ofglutamate excitotoxicity problematic.

2.2 Calcium-mediated eventsAccumulation of excessive levels of calcium has long beenconsidered part of the final common pathway of neuronal celldeath [13]. This has been largely due to the recognition thatcalcium is a ubiquitous second messenger whose uncontrolledaccumulation can initiate a series of destructive enzymaticreactions. Entry of calcium is via one of three voltage sensitivecalcium channels and by ligand gated calcium channels. All ofthese calcium channels have their own specific antagonists.Furthermore, there are Na+/Ca2+ antiporters that normallyextrude calcium from the cell, but at high internal Na+ con-centration, have been shown to translocate Ca2+ into the cell[14]. Add to this the presence of intracellular Ca2+ stores thatcan release their ion stores upon stimulation by second mes-sengers [15] activated by TBI and we have a highly complex,integrated system for the regulation of intracellular Ca2+ con-centration. It is, therefore, perhaps not unexpected that phar-macotherapies directed at inhibiting one of these processes inan effort to attenuate intracellular Ca2+ accumulation havedemonstrated limited success.

2.3 MagnesiumA role for magnesium in the pathophysiology of TBI was firstproposed in the late 1980s, following the observation that

1376 Expert Opin. Investig. Drugs (2002) 11(10)

Vink & Nimmo

brain intracellular free magnesium concentration declines aftera traumatic event and that this decline is correlated to func-tional outcome [16]. This observation has now been replicatedby a number of research groups in a number of different exper-imental models that simulate features of clinical TBI [17-19]. Assuch, free magnesium decline seems to be a ubiquitous featureof injury to the central nervous system (CNS). What is partic-ularly relevant to TBI is that magnesium is involved in the reg-ulation of a number of physiological and biochemicalprocesses within the cell, each of which plays an important rolein normal cell function. For example, magnesium is essentialfor normal cellular bioenergetics, including both substratelevel and oxidative phosphorylation [20]. Indeed, magnesium isa mandatory cofactor in all energy producing and consumingreactions involving carbohydrate, lipid, nucleic acid and pro-tein metabolism. Over 300 of the enzymes involved in theseprocesses are magnesium dependent and have a direct require-ment for optimum magnesium levels [21]. For example, the for-mation of the initiation complex in protein synthesis dependson magnesium ions [22] and any decline in magnesium willinhibit protein synthesis. Indeed, reducing levels of magne-sium to < 0.25 mM have been shown to inhibit both RNAaggregation and DNA synthesis [23]. Magnesium is also man-datory for ATPase function. Therefore, reducing magnesiumlevels will inhibit ion homeostasis maintained by the Na+/K+

and Ca2+ ATPase complexes. In addition to direct effects onenzymes, magnesium also has effects on plasma membraneintegrity [24] and ion channel activity. Particularly notable isthe fact that magnesium is an endogenous antagonist of Ca2+

channels [25], including the glutamate NMDA channel [26].Moreover, by blocking presynaptic Ca2+ channels, magnesiuminhibits glutamate release [27]. Finally, magnesium is known toblock the mitochondrial permeability transition pore impli-cated in apoptosis [28].

Clearly, there are numerous avenues by which magnesiumcan regulate normal physiology, thus making it a particularlyattractive target for pharmacologies aimed at attenuating mul-tiple secondary injury factors.

2.4 Free radicalsFree radicals are highly reactive molecules produced as a nor-mal by-product of oxidative metabolism. Under normal con-ditions their concentration is tightly controlled byendogenous antioxidant mechanisms, including glutathioneand the enzyme superoxide dismutase. However, traumaticinjury profoundly increases the production of free radicals,particularly when iron released from haemoglobin is present[29-31]. At high concentrations, these reactive oxygen species(ROS) cause considerable damage to proteins, lipids andDNA, with rapid cell death likely if the damage continuesunabated. Even if cell death is not rapid, there is evidence thatcontinued exposure to ROS opens the mitochondrial transi-tion pore, thus promoting cell death via apoptosis [32].

Administration of antioxidants has been shown to be effec-tive in experimental models of traumatic injury to the CNS

[33], particularly when there is an ischaemic or haemorrhagiccomponent to the insult. However, on their own, they havenot shown efficacy in clinical trials of TBI.

2.5 Mitochondrial damageMitochondrial integrity is critical to cell survival, given thehigh energy requirements of neurons. While most studieshave examined mitochondrial function in ischaemia, thereare now a number of studies that have focussed on TBI [34].Initial studies have shown that TBI induces impairments inthe rate of respiration, with mild-to-moderate injury pro-ducing slight but insignificant changes in respiratory rateand in coupling between electron transport and ATP synthe-sis [35], while more severe injuries produce significantimpairments of mitochondrial respiration and respiratorycoupling [36]. These alterations in mitochondrial respirationwere associated with alterations in respiratory chain cyto-chrome oxidase expression and activity [37]. In addition to animpaired ability of mitochondria to carry out oxidativephosphorylation, TBI profoundly impairs the ability ofmitochondria to actively transport calcium [38]. Not onlydoes this impact on the cytosolic Ca2+ concentration and theassociated Ca2+ induced cell death, it also promotes the per-meability changes of the mitochondrial permeability transi-tion pore [39]. This pore is integrally involved in apoptoticcell death as well as with uncoupling and inhibition of oxi-dative phosphorylation and stimulating the generation ofmitochondrial ROS [34].

2.6 OedemaIn a recent clinical study, oedema was found to be responsi-ble for 50% of all deaths recorded in young victims of braininjury [40]. This, in itself, emphasises the importance ofbrain oedema in determining outcome following TBI.There has been a considerable amount of research done inan effort to understand the mechanisms associated with theformation of oedema, and how this relates to functionaloutcome following TBI [41]. There are primarily two formsof oedema initiated after TBI. The first is vasogenicoedema, which is apparent very early after the traumaticevent and is associated with an increased permeability of theblood–brain barrier (BBB). Water is found to accumulate inthe extracellular space as osmotic substances escape fromthe vasculature. The second form of oedema is cytotoxicoedema, which develops when osmotic pressure in theintracellular space increases to beyond that observed in thesurrounding space. This form of oedema is usually associ-ated with some form of neurotoxic event and, although itdevelops at later time points than vasogenic oedema, it hasbeen reported to account for most of the brain swellingafter TBI [42]. Nonetheless, it is the initial vasogenic oedemathat is thought to be permissive for the subsequent develop-ment of cytotoxic oedema [43]. Clearly, inhibition of oedemais a highly desirable objective in the pharmacological man-agement of TBI.

Expert Opin. Investig. Drugs (2002) 11(10) 1377

Novel therapies in development for the treatment of traumatic brain injury

3. Attenuating secondary injury

3.1 MagnesiumHaving discussed the pathophysiology of magnesium declineafter TBI above (Section 1.3), the evidence that suggests thatadministration of magnesium may be neuroprotective willnow be addressed. Initial studies administered intravenousmagnesium sulfate prior to injury in an effort to prevent themagnesium decline after trauma [16]. In addition to prevent-ing the brain free magnesium decline, it was noted that thefunctional outcome of the animals was significantly betterthan that observed in vehicle treated controls. The influenceof brain magnesium status on outcome following TBI waslater confirmed in studies demonstrating that dietary magne-sium depletion prior to injury exacerbated neurological defi-cits after injury, whereas preinjury administration ofmagnesium sulfate attenuated these deficits [44]. Since thoseearly prophylactic studies, numerous reports have since con-firmed that postinjury administration of magnesium saltsresults in a significant improvement in both motor and cogni-tive outcome after either TBI [45-49] or a lesion inducinginjury to the sensorimotor cortex [50]. While these studieshave used either the sulfate or chloride salts of magnesium,recent dose/response studies confirm that both salts areequally effective when given by intravenous or intramuscularinjection [51]. Indeed, both salts were shown to enter the brainintracellular compartment and increase free magnesium con-centration after TBI [51]. Even a bolus dose given as late as24 h after TBI had a significant effect on neurological out-come, although administration within 12 h of injury was rec-ommended for maximum efficacy [52]. It should be noted thatthere are also some contradictory reports in the literature, par-ticularly in hypoxia/ischaemia studies, suggesting that magne-sium may not be neuroprotective. However, these reportstend to use an intravenous dose between 100 – 300 mg/kg,far in excess of the optimal dose reported in the dose/responsestudies [51,53].

In addition to neurological outcome, a variety of outcomeparameters have shown improvement with magnesiumadministration. Reduction in magnesium after TBI wasshown to be associated with profound oedema formation [54].Not surprisingly, administration of magnesium after TBIreduced brain oedema [45,55], although it was only recentlydemonstrated that this might be via a direct attenuation ofBBB opening [56]. Up until now, it was generally believed thatit was the effect of magnesium on the activity of the Na+/K+

ATPase that accounted for this reduction in brain swellingafter TBI.

Recent studies have also demonstrated a reduction in post-traumatic lesion size with magnesium treatment [57] similar tothat described when magnesium is administered after ischae-mia [58]. This reduction in lesion volume with magnesiumtreatment after TBI was associated with a reduction in acutecytoskeletal alterations, a condition that the authors attributeto activation of calpain [59]. Calpain is one of the calcium-acti-

vated proteases activated after TBI, whose substrates includecytoskeletal proteins, receptor proteins, signal transductionenzymes and transcription factors. Moreover, becausecytoskeletal damage was exacerbated when injury was inducedin a magnesium deficient animal, the authors suggest that themagnesium reduces cytoskeletal injury by inhibiting calpain[59]. Certainly, a decrease in magnesium concentration canresult in an increased calcium concentration, as has beenshown in the hippocampal CA1 region [60]. Presumably thisoccurs via increased glutamate NMDA channel influx, sincetraumatic stretch injury has been shown to reduce the magne-sium blockade of the NMDA channel [61]. Thus, magnesiumcould inhibit calpain activation by both reducing Ca2+ influxand also by directly inhibiting calpain itself.

The potentially protective effects of magnesium are notlimited to calcium-mediated events, oedema formation andinhibition of neurotransmitter release and binding. Magne-sium is a known membrane stabiliser [24] and has beenreported to reduce membrane lipolysis [62]. This has anumber of positive effects on ion channels as reflected in amagnesium-induced inhibition of membrane depolarisa-tion and cortical spreading depression [63] and improve-ment in spinal somatosensory evoked potentials followingspinal cord injury [64]. By reducing lipid peroxidation [64],magnesium has also been shown to decrease hypoxia-induced increases in oxygen free radicals [65] and attenuatethe production of ROS [66,67]. These observations are con-sistent with the improvement in endogenous antioxidantlevels [68] and the reduction in malondialdehyde levels [69]

reported following post-traumatic magnesium administra-tion. The positive effect on membranes is also reflected inthe ability of magnesium to preserve mitochondrial mem-brane potential [70]. This presumably will have positiveeffects on oxidative phosphorylation and explain the reduc-tion in lactate levels observed when magnesium is adminis-tered following TBI [69]. Therefore, by improvingmembrane stability and decreasing the production of ROS,magnesium may preserve the mitochondrial capacity foroxidative phosphorylation and facilitate the synthesis of theadenine nucleotides. However, while the mitochondria areprimarily concerned with energy metabolism, their role inapoptotic cell death following nervous system insults hasnow become firmly established [34].

The potential role of magnesium in apoptosis hasrecently been reviewed [71] and this discussion will, there-fore, focus on aspects that are relevant to TBI. Magnesiumhas the ability to block the mitochondrial permeabilitytransition pore [67,72]. It is the increased permeability of thispore that is associated with the release of cytochrome c andinduction of the caspases that ultimately induce apoptoticcell death. The permeability transition is influenced by anumber of factors, including ROS, adenine nucleotide con-centration, calcium and the Bcl2 protein family. Magne-sium has an influence on all of these factors, some of whichhave already been mentioned.

1378 Expert Opin. Investig. Drugs (2002) 11(10)

Vink & Nimmo

The evidence that it influences the Bcl2 family has onlyrecently been reported. In hypoxia, it was shown that admin-istration of magnesium reduces the concentration of pro-apoptotic Bax and increases the concentration of antiapop-totic Bcl2 [73]. It is the balance between these two proteinsthat has been shown to be associated with the mitochondrialpore transition. In addition to both directly and indirectlyaffecting the mitochondrial permeability transition pore,magnesium is also capable of preventing induction of pro-apoptotic p53 [74] and inhibiting DNA fragmentation factor[75] following TBI. The net result is that magnesium adminis-tered after injury to the brain decreases nuclear oxidative dam-age and DNA fragmentation [76].

While this discussion of magnesium treatment in TBI hasfocussed on biochemical events, magnesium also has anumber of beneficial effects on physiological parameters. Themost significant in terms of TBI are effects on the cerebrovas-culature. Administration of magnesium reverses delayed cere-brovasospasm after subarachnoid haemorrhage [77]. Moreover,it has been widely demonstrated that magnesium inhibits cer-ebral vasospasms [78-80]. These properties, in part, account forthe reversal of vasoconstriction and increased cerebral blood-flow that has been recorded in rats following magnesiumadministration [81,82]. Although it is thought that the ability ofmagnesium to block calcium channels may, in part, accountfor these effects, other mechanisms of action have not beenruled out [83].

The multifactorial nature of magnesium’s effects on thebrain has resulted in a number of ongoing Phase III clinicaltrials in various nervous system disorders, including TBI(Magnesium Sulfate for Neuroprotection after BrainTrauma), cerebral palsy (Randomised Clinical Trial of theBeneficial Effects of Antenatal Magnesium Sulfate [BEAM]),stroke (Intravenous MAGnesium Efficacy in Stroke [Images],Magnetic Resonance in Intravenous Magnesium Efficacy inStroke [MR Images], Field Administration of Stroke Ther-apy-Magnesium [Fast-Mag]), subarachnoid haemorrhage(Magnesium and Acetylsalicylic Acid in Subarachnoid Hem-orrhage [MASH]) and intracranial haemorrhage (MAGne-sium Efficacy and Safety in Treating Intra-CranialHemorrhage [MAGESTIC]).

3.2 Substance P antagonistsDespite the small number of reports that have appeared inbrain injury studies, substance P antagonists are receivingincreased attention as neuroprotective agents. This is, in part,because of their recent identification as an effective pharmaco-therapy in clinical depression, emesis and neuropathic pain[84] and the subsequent interest in their potential in otherpathologies of the nervous system. There is now considerableevidence to suggest that substance P plays a significant role intissue injury. For example, in cardiac studies, magnesium defi-ciency has been shown to promote substance P induced neu-rogenic inflammation [85]. Many in vivo actions of substanceP are mediated by neurokinin 1 (NK1) receptors, and selective

NK1 antagonists [86] were subsequently used to inhibit sub-stance P action and reduce the neurogenic inflammation.NK1 receptor blockade was reported to reduce in vivo pro-oxi-dant stress (including glutathione loss), prenecrotic perivascu-lar inflammatory infiltration, circulating histamine, PGE2and lipid peroxidation products [85]. These authors proposethat substance P release may, in fact, be one of the earliestpathophysiological events associated with injury, leading tostimulation of inflammatory cytokines and subsequent stimu-lation of free radical mechanisms of injury. Significant sub-stance P release has been detected in the nervous systemfollowing both peripheral nerve injury [87] and traumatic spi-nal cord injury [88], an effect that was blocked by inhibitors ofserotonin synthesis [88]. In the peripheral nervous system(PNS), NK1 receptor immunoreactivity has been shown toincrease following injury [89], an observation also demon-strated for glia following CNS injury [90,91]. This increase wasnot observed in undamaged areas. The increased expression ofthese receptors on the astrocytes after injury may, therefore, belinked to their transformation to reactive astrocytes. In brainischaemia, induction of NK1 receptors in the endotheliumwas shown to contribute to oedema formation [92], an obser-vation that has been well characterised in peripheral tissueinjury [93]. Accordingly, the NK1 receptor was proposed as apotential therapeutic target in the treatment of stroke [92].Targeting this receptor has indeed proven beneficial, withinhibition of the substance P NK1 receptor following experi-mental ischaemia reducing infarct volume and improvingfunctional outcome [94]. Similar beneficial effects of NK1antagonists have additionally been reported in intestinalischaemia/reperfusion injury [95] and in TBI [96]. In the latter,the improvement in both cognitive and motor outcome withadministration of an NKI antagonist was correlated with areduction in post-traumatic oedema formation. Since thereare high numbers of NK1 receptors in the hippocampus andstriatum [97] and the fact that NK1 antagonists improved cog-nitive outcome after TBI, it appears there may also be a rolefor substance P in learning and memory.

There are a variety of other mechanisms by which sub-stance P antagonists may confer a neuroprotective effect. Fol-lowing middle cerebral artery occlusion, there is a markedexpression of substance P and NK1 receptors in glutamater-gic pyramidal cells [92]. Such colocalisation of NMDA andNK1 receptors had previously been reported in spinal cord[98] and brain [99]. Moreover, this colocalisation may facilitateglutamate-mediated neurotoxicity, a hypothesis supported bythe observation that substance P potentiates cellularresponses to NMDA, an effect that can be blocked by sub-stance P antagonists [100] and has been implicated in themodulation of locomotor behaviour. Substance P is capableof regulating the action of other neurotransmitters, includingdopamine release [101] and acetylcholine release [102] and byopening inward cation channels [103], notably calcium, it mayalso modulate the presynaptic release and postsynapticactions of a number of other neurotransmitters. Substance P

Expert Opin. Investig. Drugs (2002) 11(10) 1379

Novel therapies in development for the treatment of traumatic brain injury

also induces endothelial cells to produce nitric oxide [104],

which, in itself, has been implicated as a secondary injuryfactor. The priming of polymorphonuclear cells for oxidativemetabolism (superoxide production) by substance P [105] alsoprovides a source of ROS known to exacerbate the injuryprocess. Finally, the evidence showing that NK1 antagonistsreduce cFos expression [106] suggests that there are alsomolecular effects of the compounds still to be characterised.

To date, substance P antagonists have been shown to haveantidepressant, antiemetic and antinociceptive properties,together with an ability to reduce neurogenic inflammation,BBB permeability, oedema, lesion volume and improve func-tional outcome in both ischaemia and TBI. Such multifacto-rial effects make it an ideal candidate for further investigation.

3.3 Cyclosporin ACsA is a highly lipophilic member of the immunophilin familyused for over 2 decades as an immunosuppressant drug. Its useas a treatment in neurological disorders has been recentlyreviewed [107] and, as such, the following comments will specif-ically relate to its potential use in TBI.

Early studies had demonstrated that CsA was able toreduce cell death and reduce lesion volume following ischae-mia–reperfusion injury [108,109]. Although the precise mecha-nisms were unknown, CsA inhibits the mitochondrialpermeability transition pore opening [28,110] and it was thismechanism that was considered particularly relevant to out-come. In studies of TBI, administration of CsA was againshown to reduce lesion volume, along with post-traumaticcytoskeletal changes and axonal injury as identified usingamyloid precursor protein immunoreactivity [111,112]. Thecompound’s ability to preserve cholinergic neurons [113]

implied a potential benefit in some cognitive disorders, apotential supported by findings demonstrating a positiveeffect of CsA on hippocampal plasticity after TBI [114]. How-ever, studies to date have failed to illustrate a benefit on cog-nitive outcome after TBI despite a robust protective effect onmotor and sensorimotor function [115].

The ability of CsA to improve post-traumatic mitochon-drial membrane potential, as well as decreasing intramito-chondrial calcium levels and ROS [116], once more supportedthe compounds ability to reduce mitochondrial permeabilitytransition pore opening as a primary mechanism of action.However, later studies have since demonstrated that the com-pound does not always inhibit apoptosis, suggesting that neu-roprotection may also be conferred by mechanisms other thanjust inhibition of mitochondrial permeability transition poreopening [117]. This was later confirmed when FK-506,another neuroprotective member of the immunophilin family,but without the ability to block mitochondrial permeabilitytransition pore opening, was also successfully used to attenu-ate axonal damage after TBI [118]. Obviously, alternativemechanisms were responsible. Aside from well-knownantiparasitic, fungicidal and anti-inflammatory properties,CsA also has the ability to directly inhibit calcineurin, a cal-

cium activated phosphatase strongly implicated in secondaryinjury. The compound has also been reported to inhibit lipidperoxidation after spinal cord injury at least as effectively asmethylprednisolone [119]. Finally, the compound is a weaksubstance P NK1 receptor blocker [120].

Thus, CsA has a number of beneficial effects on secondaryinjury factors that may contribute to its neuroprotective prop-erties. Dose response curves in experimental studies have sup-ported a reasonable therapeutic window of effectiveness [121],and Phase III clinical trials of CsA treatment of TBI havecommenced in a collaboration between the University of Flor-ida and the Medical College of Virginia.

3.4 Other compoundsThere are a number of other multifactorial compounds thathave been proposed as potentially effective therapies for thetreatment of TBI. Some of these compounds are new andrequire further research, others have been used experimentallyfor many years, but have failed to garner widespread support,while others have primarily been used in studies of ischaemiarather than TBI. The following is a short description of someof these compounds that are of particular interest to TBI.

3.4.1 DexanabinolDexanabinol (HU-211) is a non-psychotropic analogue of tet-rahydrocannabinol that has properties as an NMDA receptorantagonist [122], a free radical scavenger, an antioxidant [123] andan inhibitor of cytokine TNF-α [124]. Although most studieshave utilised models of ischaemia to demonstrate the neuropro-tective properties of this compound, some studies have exam-ined its utility in TBI [124-126]. In summary, these studies havedemonstrated that dexanabinol reduces BBB permeability,oedema and lesion volume, while improving both motor andcognitive outcome after trauma [127]. Dexanabinol has only beenstudied under conditions of very severe TBI containing anischaemic component. It is therefore unclear whether the drugmay be efficacious in mild-to-moderate TBI that make up mostclinical trauma cases. Nonetheless, the drug is in Phase III clini-cal trial (Efficacy Assessment of Dexanabinol Treatment of TBIVictims) and it is hopeful that the treatment may be of benefitto severely injured patients.

3.4.2 AM-36AM-36 is an arylalkylpiperazine compound designed to blockthe polyamine site of the NMDA receptor and simultaneouslypossess antioxidant activity [128]. By targeting the polyaminesite, the inventors avoided the deleterious side effects associatedwith other glutamate NMDA channel blockers, such as MK-801. In addition to its ability to block NMDA channel activityand lipid peroxidation, the compound has also shown proper-ties as a sodium channel blocker and an inhibitor of apoptosis[129]. Thus, the compound is a true multifactorial drug. Whenadministered < 3 h after middle cerebral artery occlusion, AM-36 significantly attenuated infarct volume and improved motorand sensorimotor deficits [130]. The effects on functional out-

1380 Expert Opin. Investig. Drugs (2002) 11(10)

Vink & Nimmo

come correlated with histological improvement. Whilst so farlimited to studies of ischaemia, the multiple modes of action onboth necrosis and apoptosis bode well for future studies utilis-ing other injury models, including TBI.

3.4.3 Thyrotrophin releasing hormoneThe suggestion that thyrotrophin releasing hormone (TRH)and its analogues may prove useful in the treatment of injury tothe CNS has been promoted for many years [131]. Experimentalstudies have shown that TRH and the early TRH analogues,such as CG3703 and YM-14673, were able to improve motoroutcome [132]. Some of this improvement could be attributedto physiological and metabolic actions, including improvingmagnesium status [133], bioenergetic state [133] and promotingrecovery in cerebral bloodflow following brain trauma [134].However, other physiological actions of these compounds,including autonomic, analeptic and endocrine effects, wereundesirable from a clinical point of view. Accordingly, thesecompounds did not attract much support as neuroprotectiveagents and their clinical use, to date, has been limited to thepromotion of recovery from disturbances of consciousness [135].Newly developed TRH analogues have been designed to largelyeliminate the undesirable physiological actions of this class ofdrug while preserving the neuroprotective effects [136]. In doingso, the compound has become even more of a multifactorialdrug, now with nootropic actions, but without the adverse sideeffects. However, with the removal of many of the physiologicaleffects, it has become very difficult to establish how these drugsactually confer their neuroprotection and, accordingly, nomechanism of action has been proposed. Until these mecha-nisms of action have been identified, there will be a continuingreluctance to utilise these compounds.

4. Conclusion

A number of experimental studies in TBI have identifiedthat magnesium, substance P antagonists and CsA haveeffects on multiple secondary injury factors. However, in aneffort to develop a designer ‘magic bullet’ drug, the unique,

broad spectrum properties of these drugs were overlooked infavour of developing designer drugs targeting a single mech-anism of injury thought to be primarily responsible for irre-versible tissue injury. It has now been accepted that a singleinjury factor is highly unlikely to be responsible for all of theinjury that occurs after trauma, particularly in view of thefact that cellular injury can occur through different mecha-nisms incorporating both necrosis and apoptosis. Accord-ingly, the focus has shifted back onto drugs that havemultifactorial effects. A great deal of evidence had alreadybeen accumulated in favour of magnesium as a neuroprotect-ant, which accounts for the large number of clinical trialsthat are currently underway using this compound. However,there is considerable effort at present to identify other suchcompounds and recent focus is on the substance P antago-nists and CsA. Despite the earlier failures in developing aneffective pharmacological intervention for TBI, it is hopedthat these new efforts will be successful.

5. Expert opinion

It is clear that TBI results in a multifactorial secondaryinjury cascade that results in neuronal cell death. It shouldalso be clear that the relative contribution of each secondaryinjury factor to the injury cascade will be largely influencedby the type of injury that has occurred. For example, verysevere injuries that have either a haemorrhagic or ischaemiccomponent would be expected to have far more free radicalproduction than those without these secondary complica-tions. It is this heterogeneity of injury combined with thecomplexity of the secondary injury cascade that is particu-larly challenging to those seeking to develop efficaciouspharmacological therapies.

Somewhat surprisingly, most, if not all, clinical trials in TBIto date have focussed on interventions that target one specificinjury factor, be it inhibition of the NMDA channel or someother individual factor. The reasons for this are somewhatunclear. However, it is encouraging to note that the more recentresearch efforts have focussed on treatment strategies that affect

Table 1. Interaction between magnesium, substance P antagonists and cyclosporin A and the various secondary injury factors and injury outcome.

Compound Injury factors Functional outcome

Lesion volume

Excitotoxicity Calcium-mediated events

Magnesium ROS Mitochondrial events

Oedema

Magnesium [26,27,60] [16,59,60] [16,51] [65-69] [28,67,70,72,73] [45,55,56] [16,44-52] [57,58]

Substance P antagonists

[100] [103] – [85,105] – [92,96] [94,96] [94]

CsA – [116] – [116,19] [28,110,116] – [115] [108,109,111,112]

CsA: Cyclosporin A: ROS: Reactive oxygen species.

Expert Opin. Investig. Drugs (2002) 11(10) 1381

Novel therapies in development for the treatment of traumatic brain injury

multiple injury factors simultaneously. Some investigators havechosen to develop cocktails of individual drugs to achieve thisgoal and, in some cases, this has met with reasonable success[137]. However, in other cases, the combination of drugs has notbeen as effective as the individual drugs given alone [138], sug-gesting that care needs to be taken to ensure the properties ofeach drug do not counteract the other. This is one of the defi-ciencies in polypharmacy, the other being that multiple drugshave multiple side effects. For the pharmaceutical industry, thisis a most serious consideration.

The alternative to using polypharmacy is to identify poten-tial broad spectrum, therapeutic candidates that are multifac-torial in nature. While there are a number of promisingcandidates, this review has focussed on three novel com-pounds that have shown very promising results in TBI studiesto date. That is not to say that the basic research over thepotential of magnesium, substance P antagonists and CsA astherapeutic agents for use in TBI is complete. Not all of them

have been tested in multiple models of TBI, or have had com-plete dose/response studies performed or therapeutic windowof efficacy established. Nonetheless, despite these deficiencies,there is considerable excitement surrounding the prospect thatone of these agents may be the first successful pharmacother-apy to be used in TBI. All of these drugs attenuate multipleinjury factors (summarised in Table 1) and have had neuro-protective properties demonstrated in different experimentalmodels of TBI. Finally, the neuroprotection with all of thesedrugs has been demonstrated not only in terms of lesion vol-ume, but also in terms of functional outcome. It is thisapproach that will lead to the identification of a novel therapyfor the treatment of TBI.

Acknowledgement

RV and AJN are supported, in part, by the AustralianNational Health and Medical Research Council.

BibliographyPapers of special note have been highlighted as either of interest (•) or of considerable interest (••) to readers.

1. MCINTOSH TK, JUHLER M, WIELOCH T: Novel pharmacologic strategies in the treatment of experimental traumatic brain injury: 1998. J. Neurotrauma (1998) 15:731-769.

2. DOPPENBERG EMR, BULLOCK R: Clinical neuro-protection trials in severe traumatic brain injury: lessons from previous studies. J. Neurotrauma (1997) 14:71-80.

3. MAAS AIR: Neuroprotective agents in traumatic brain injury. Expert Opin. Investig. Drugs (2001) 10:753-767.

4. NARAYAN RK, MICHEL ME, ANSELL B et al.: Clinical trials in head injury. J. Neurotrauma (2002) 19:503-557.

•• Comprehensive and critical review of the positive and negative aspects of previous clinical trials in TBI.

5. FADEN AI: Neuroprotection and traumatic brain injury: the search continues. Arch. Neurol. (2001) 58:1553-1555.

6. PETTMAN B, HENDERSON CE: Neuronal cell death. Neuron (1998) 20:633-647.

• Excellent review of the mechanisms associated with neuronal cell death.

7. CONTI AC, RAGHUPATHI R, TROJANOWSKI JQ, MCINTOSH TK: Experimental brain injury induces regionally distinct apoptosis during the acute and delayed post-traumatic period. J. Neurosci.

(1998) 18:5663-5672.

8. LEA PMT, FADEN AI: Traumatic brain injury: developmental differences in glutamate receptor response and the impact on treatment. Ment. Retard. Dev. Disabil. Res. Rev. (2001) 7:235-248.

9. POHL D, BITTIGAU P, ISHIMARU MJ et al.: N-methyl-D-aspartate antagonists and apoptotic cell death triggered by head trauma in developing rat brain. Proc. Natl. Acad. Sci. USA (1999) 96:2508-2513.

10. FADEN AI, DEMEDIUK P, PANTER SS, VINK R: Excitatory amino acids, N-methyl-D-asparate receptors and traumatic brain injury. Science (1989) 244:798-800.

11. BULLOCK R, ZAUNER A, WOODWARD JJ et al.: Factors affecting excitatory amino acid release following severe human head injury. J. Neurosurg. (1998) 89:507-518.

12. CHOI DW: Excitotoxic cell death. J. Neurobiol. (1992) 23:1261-1276.

13. SIESJO BK: Historical overview: calcium, ischaemia and death of brain cells. Ann. NY Acad. Sci. (1988) 522.

14. NACHSEN DA, SANCHEZ-ARMASS S, WEINSTEIN AM: The regulation of cytosolic calcium in rat brain synaptosomes by sodium-dependent calcium efflux. J. Physiol. (1986) 381:17-28.

15. BERRIDGE MJ, IRVINE RF: Inositol triphosphates and cell signalling. Nature (1989) 341:197-205.

16. VINK R, MCINTOSH TK, DEMEDIUK

P, WEINER MW, FADEN AI: Decline in intracellular free magnesium concentration is associated with irreversible tissue injury following brain trauma. J. Biol. Chem. (1988) 263:757-761.

•• One of the first manuscripts describing a decline in brain intracellular free magnesium concentration following injury to the brain.

17. HEATH DL, VINK R: Traumatic brain axonal injury produces sustained decline in intracellular free magnesium concentration. Brain Res. (1996) 738:150-153.

18. SMITH DH, CECIL KM, MEANEY DF et al.: Magnetic resonance spectroscopy of diffuse brain trauma in the pig. J. Neurotrauma (1998) 15:665-674.

19. SUZUKI M, NISHINA M, ENDO M et al.: Decrease in cerebral free magnesium concentration following closed head injury and effects of VA-045 in rats. Gen. Pharmacol. (1997) 28:119-121.

20. EBEL H, GUNTHER T: Magnesium metabolism: a review. J. Clin. Chem. Clin. Biochem. (1980) 18:257-270.

21. BIRCH NJ (Ed.): Magnesium and the Cell. Academic Press, London (1993).

22. TERASAKI M, RUBIN H: Evidence that intracellular magnesium is present in cells at a regulatory concentration for protein synthesis. Proc. Natl. Acad. Sci. USA (1985) 82:7324-7326.

23. RUBIN H: Magnesium deprivation reproduces the coordinate effects of serum removal or cortisol addition on transport

1382 Expert Opin. Investig. Drugs (2002) 11(10)

Vink & Nimmo

and metabolism in chick embryo fibroblasts. J. Cell. Physiol. (1976) 89:613-626.

24. BARA M, GUIET-BARA A: Potassium, magnesium and membranes. Review of present status and new findings. Magnesium (1984) 3:212-225.

25. ISERI LT, FRENCH JH: Magnesium: nature’s physiologic calcium blocker. Am. Heart J. (1984) 108:188-193.

26. MAYER ML, WESTBROOK GL, GUTHRIE PB: Voltage-dependent block by Mg2+ of NMDA responses in spinal cord neurons. Nature (1984) 309:261-263.

27. ROTHMAN SM: Synaptic activity mediates death of hypoxic neurons. Science (1983) 220:536-537.

28. SZABO I, ZORATTI M: The giant channel of the inner mitochondrial membrane is inhibited by cyclosporin A. J. Biol. Chem. (1991) 266:3376-3379.

29. BRAUGHLER JM, HALL ED: Involvement of lipid peroxidation in CNS injury. In: Central Nervous System Trauma Status Report – 1991. Jane JA, Torner JC, Anderson DK, Young W (Eds), Mary Ann Liebert, Inc., New York (1992):1-7.

30. CERNAK I, SAVIC VJ, KOTUR J, PROKIC V, VELJOVIC M, GRBOVIC D: Characterization of plasma magnesium concentration and oxidative stress following graded traumatic brain injury in humans. J. Neurotrauma (2000) 17:53-68.

• The first extensive clinical study examining the inter-relationship between blood free magnesium and calcium concentration following stress and TBI.

31. LEWEN A, MATZ P, CHAN PH: Free radical pathways in CNS injury. J. Neurotrauma (2000) 17:871-890.

32. HALESTRAP AP, WOODFIELD KY, CONNERN CP: Oxidative stress, thiol reagents and membrane potential modulate the mitochondrial permeability transition by affecting nucleotide binding to the adenine nucleotide translocase. J. Biol. Chem. (1997) 272:3346-3354.

33. HALL ED, YONKERS PA, ANDRUS PK, COX JW, ANDERSON DK: Biochemistry and pharmacology of lipid antioxidants in acute brain and spinal cord injury. J. Neurotrauma (1992) 9:S425-S442.

34. FISKUM G: Mitochondrial participation in ischemic and traumatic neural cell death. J. Neurotrauma (2000) 17:843-855.

35. VINK R, HEAD VA, ROGERS PJ, MCINTOSH TK, FADEN AI:

Mitochondrial metabolism following traumatic brain injury in rats. J. Neurotrauma (1990) 7:21-27.

36. XIONG Y, GU Q, PETERSEN PL, MUIZELAAR JP, LEE CP: Mitochondrial dysfunction and calcium perturbation induced by traumatic brain injury. J. Neurotrauma (1997) 14:23-34.

37. HARRIS LK, BLACK RT, GOLDEN KM, REEVES TM, POVLISHOCK JT, PHILLIPS LL: Traumatic brain injury-induced changes in gene expression and functional activity of mitochondrial cytochrome c oxidase. J. Neurotrauma (2001) 18:993-1009.

38. VERWEIJ BH, MUIZELAAR JP, VINAS EC, PETERSON PL, XIONG Y, LEE CP: Mitochondrial dysfunction after experimental and human brain injury and its possible reversal with a selective N-type calcium channel anatagonist (SNX-111). Neurol. Res. (1997) 19:334-339.

39. LEMASTERS JL, NIEMINEN AL, QIAN T et al.: The mitochondrial permeability transition in cell death: a common mechanism in necrosis, apoptosis and autophagy. Biochim. Biophys. Acta (1998) 1366:177-196.

40. FEICKERT HJ, DROMMER S, HEYER R: Severe head injury in children: impact of risk factors. J. Trauma (1999) 47:33-38.

41. MARMAROU A: Traumatic brain edema – an overview. Acta Neurochir. Suppl. (1994) 60:421-424.

42. BARZO P, MARMAROU A, FATOUROS P, HAYASAKI K, CORWIN F: Contribution of vasogenic and cellular edema to traumatic brain swelling measured by diffusion-weighted imaging. J. Neurosurgery (1997) 87:900-907.

43. BEAUMONT A, MARMAROU A, HAYASAKI K et al.: The permissive nature of blood–brain barrier (BBB) opening in edema formation following traumatic brain injury. Acta Neurochir. Suppl. (2000) 76:125-129.

44. MCINTOSH TK, FADEN AI, YAMAKAMI I, VINK R: Magnesium deficiency exacerbates and pretreatment improves outcome following traumatic brain injury in rats: 31P magnetic resonance spectroscopy and behavioural studies. J. Neurotrauma (1988) 5:17-31.

45. FELDMAN Z, GUREVITCH B, ARTRU AA et al.: Effect of magnesium given 1 h after head trauma on brain edema and neurological outcome. J. Neurosurgery

(1996) 85:131-137.

46. HEATH DL, VINK R: Magnesium sulfate improves neurologic outcome following severe closed head injury in rats. Neurosci. Lett. (1997) 228:175-178.

47. MCINTOSH TK, VINK R, YAMAKAMI I, FADEN AI: Magnesium protects against neurological deficit after brain injury. Brain Res. (1989) 482:252-260.

48. SMITH DH, OKIYAMA K, GENNARELLI TA, MCINTOSH TK: Magnesium and ketamine attenuate cognitive dysfunction following experimental brain injury. Neurosci. Lett. (1993) 157:211-214.

49. VINK R, MCINTOSH TK: Pharmacological and physiological effects of magnesium on experimental traumatic brain injury. Magnesium Res. (1990) 3:163-169.

50. HOANE MR, IRISH SL, MARKS BB, BARTH TM: Preoperative regimens of magnesium facilitate recovery of function and prevent subcortical atrophy following lesions of the rat sensorimotor cortex. Brain Res. Bull. (1998) 45:45-51.

51. HEATH DL, VINK R: Optimization of magnesium therapy following severe diffuse axonal brain injury in rats. J. Pharmacol. Exp. Ther. (1999) 288:1311-1316.

•• Dose/response characterisation of different magnesium salts in TBI.

52. HEATH DL, VINK R: Delayed therapy with magnesium up to 24 h following traumatic brain injury improves motor outcome. J. Neurosurg. (1999) 90:504-509.

53. MUIR KW: New experimental and clinical data on the efficacy of pharmacological magnesium infusions in cerebral infarcts. Magnes. Res. (1998) 11:43-56.

54. SOARES HD, THOMAS M, CLOHERTY K, MCINTOSH TK: Development of prolonged focal cerebral edema and regional cation changes following experimental brain injury in the rat. J. Neurochem. (1992) 58:1845-1852.

55. OKIYAMA K, SMITH DH, GENNARELLI TA, SIMON RP, LEACH M, MCINTOSH TK: The sodium channel blocker and glutamate release inhibitor BW1003C87 and magnesium attenuate regional cerebral edema following experimental brain injury in the rat. J. Neurochem. (1995) 64:802-809.

56. KAYA M, KUCUK M, KALAYCI RB et al.: Magnesium sulfate attenuates increased blood–brain barrier permeability during

Expert Opin. Investig. Drugs (2002) 11(10) 1383

Novel therapies in development for the treatment of traumatic brain injury

insulin-induced hypoglycemia in rats. Can. J. Physiol. Pharmacol. (2001) 79:793-798.

57. BAREYRE FM, SAATMAN KE, RAGHUPATHI R, MCINTOSH TK: Postinjury treatment with magnesium chloride attenuates cortical damage after traumatic brain injury in rats. J. Neurotrauma (2000) 17:1029-1039.

58. YANG Y, QIU L, FAYYAZ A, SHUAIB A: Survival and histological evaluation of therapeutic window of postischaemia treatment with magnesium sulfate in embolic stroke model of rat. Neurosci. Lett. (2000) 285:119-122.

59. SAATMAN KE, BAREYRE FM, GRADY MS, MCINTOSH TK: Acute cytoskeletal alterations and cell death induced by experimental brain injury are attenuated by magnesium treatment and exacerbated by magnesium deficiency. J. Neuropathol. Exp. Neurol. (2001) 60:183-194.

60. ZHANG AM, FAN SH, CHENG TPO, ALTURA BT, WONG RKS, ALTURA BM: Extracellular Mg2+ modulates intracellular Ca2+ in acutely isolated hippocampal CA1 pyramidal cells of the guinea-pig. Brain Res. (1996) 728:204-208.

61. ZHANG L, RZIGALINSKI BA, ELLIS EF, SATIN LS: Reduction of voltage-dependent Mg2+ blockade of NMDA current in mechanically injured neurons. Science (1996) 274:1921-1923.

• A manuscript demonstrating that the NMDA receptor magnesium block is reduced following TBI.

62. GASPAROVIC C, BERGHMANS K: Ca2+- and Mg2+-modulated lipolysis in neonatal rat brain slices observed by one- and two-dimensional NMR. J. Neurochem. (1998) 71:1727-1732.

63. VAN DER HEL WS, VAN DEN BERGH WM, NICOLAY K, TULLEKEN KA, DIJKHUIZEN RM: Suppression of cortical spreading depressions after magnesium treatment in the rat. Neuroreport (1998) 9:2179-2182.

64. SUZER T, COSKUN E, ISLEKEL H, TAHTA K: Neuroprotective effect of magnesium on lipid peroxidation and axonal function after experimental spinal cord injury. Spinal Cord (1999) 37:480-484.

65. MAULIK D, ZANELLI S, NUMAGAMI Y, OHNISHI ST, MISHRA OP, DELIVORIA-PAPADOPOULOS M: Oxygen free radical generation during in-utero hypoxia in the fetal guinea-pig brain: the effects of maturity and of magnesium

sulfate administration. Brain Res. (1999) 817:117-122.

66. REGAN RF, JASPER E, GUO YP, PANTER SS: The effect of magnesium on oxidative neuronal injury in vitro. J. Neurochem. (1998) 70:77-85.

67. KOWALTOWSKI AJ, NAIA-DA-SILVA ES, CASTILHO RF, VERCESI AE: Ca2+-stimulated mitochondrial reactive oxygen species generation and permeability transition are inhibited by dibucaine or Mg2+. Arch. Biochem. Biophys. (1998) 359:77-81.

68. USTUN ME, DUMAN A, OGUN CO, VATANSEV H, AK A: Effects of nimodipine and magnesium sulfate on endogenous antioxidant levels in brain tissue after experimental head trauma. J. Neurosurg. Anesthesiol. (2001) 13:227-232.

69. USTUN ME, GURBILEK M, AK A, VATANSEV H, DUMAN A: Effects of magnesium sulfate on tissue lactate and malondialdehyde levels in experimental head trauma. Intensive Care Med. (2001) 27:264-268.

70. SHARIKABAD MN, OSTBYE KM, BRORS O: Increased [Mg2+]o reduces Ca2+ influx and disruption of mitochondrial membrane potential during reoxygenation. Am. J. Physiol. Heart Circ. Physiol. (2001) 281:H2113-H2123.

71. WOLF FI, CITTADINI A: Magnesium in cell proliferation and differentiation. Front. Biosci. (1999) 4:D607-617.

• A review of the potential role of magnesium in apoptosis.

72. BIBAN C, TASSANI V, TONINELLO A, SILIPRANDI D, SILIPRANDI N: The alterations in the energy linked properties induced in rat liver mitochondria by acetylsalicylate are prevented by cyclosporin A or Mg2+. Biochem. Pharmacol. (1995) 50:497-500.

73. RAVISHANKAR S, ASHRAF QM, FRITZ K, MISHRA OP, DELIVORIA-PAPADOPOULOS M: Expression of Bax and Bcl-2 proteins during hypoxia in cerebral cortical neuronal nuclei of newborn piglets: effect of administration of magnesium sulfate. Brain Res. (2001) 18:23-29.

74. MUIR JK, RAGHUPATHI R, EMERY DL, BAREYRE FM, MCINTOSH TK: Postinjury magnesium treatment attenuates traumatic brain injury-induced cortical induction of p53 mRNA in rats. Exp. Neurol. (1999) 159:584-593.

75. ZHANG C, RAGHUPATHI R, LAPLACA MC, BAREYRE FM, MCINTOSH TK: Changes in DNA fragmentation factor (DFF) following experimental brain trauma in the rat: effect of post-traumatic magnesium treatment. J. Neurotrauma (1998) 15:904.

76. MAULIK D, QAYYUM I, POWELL SR, KARANTZA M, MISHRA OP, DELIVORIA-PAPADOPOULOS M: Posthypoxic magnesium decreases nuclear oxidative damage in the fetal guinea-pig brain. Brain Res. (2001) 26:130-136.

77. RAM Z, SADEH M, SHACKED I, SAHAR A, HADANI M: Magnesium sulfate reverses experimental delayed cerebral vasospasm after subarachnoid hemorrhage in rats. Stroke (1991) 22:922-927.

78. ALTURA BT, ALTURA BM: The role of magnesium in aetiology of strokes and cerebrovasospasm. Magnesium (1982) 1:277-291.

79. EMA M, GEBREWOLD A, ALTURA BT, ALTURA BM: Magnesium sulfate prevents alcohol-induced spasms of cerebral blood vessels: an in situ study on the brain microcirculation from male versus female rats. Magnes. Trace Elem. (1992) 10(2-4):269-80.

80. PYNE GJ, CADOUX-HUDSON TA, CLARK JF: Magnesium protection against in vitro cerebral vasospasm after subarachnoid haemorrhage. Br. J. Neurosurg. (2001) 15:409-415.

81. KEMP PA, GARDINER SM, BENNNETT T, RUBIN PC: Magnesium sulfate reverses the carotid vasoconstriction caused by endothelin-I, angiotensin-II and neuropeptide Y but not that caused by N(G)-nitro-L-arginine methyl ester, in conscious rats. Clin. Sci. (1993) 85:175-181.

82. KEMP PA, GARDINER SM, MARCH JE, RUBIN PC, BENNETT T: Assessment of the effects of endothelin-1 and magnesium sulfate on regional blood flows in conscious rats, by the coloured microsphere reference technique. Br. J. Pharmacol. (1999) 126:621-626.

83. NIGAM S, AVERDUNK R, GUNTHER T: Alteration of prostanoid metabolism in rats with magnesium deficiency. Prost. Leukotri. Med. (1986) 23:1-10.

84. RUPNIAK NM, KRAMER MS: Discovery of the antidepressant and antiemetic efficacy of substance P receptor (NK1) antagonists. Trends Pharmacol. Sci. (1999) 20:485-490.

1384 Expert Opin. Investig. Drugs (2002) 11(10)

Vink & Nimmo

• A review of substance P antagonists and how they were discovered to play a role in CNS function.

85. KRAMER JH, PHILLIPS TM, WEGLICKE WB: Magnesium deficiency enhanced postischemic myocardial injury is reduced by substance P receptor blockade. J. Cardiol. (1997) 29:97-110.

86. FONG TM, HUANG RRC, STADER CD: Localization of agonist and antagonist binding domains of the human neurokinin-1 receptor. J. Biol. Chem. (1992) 267:25664-25667.

87. MALCANGIO M, RAMER MS, JONES MG, MCMAHON SB: Abnormal substance P release from the spinal cord following injury to primary sensory neurons. Eur. J. Neurosci. (2000) 12:397-399.

88. SHARMA HS, NYBERG F, OLSSON Y, DEY PK: Alteration of substance P after trauma to the spinal cord: an experimental study in the rat. Neuroscience (1990) 38:205-212.

89. CRUCE WL, LOVELL JA, CRISP T, STUESSE SL: Effect of ageing on the substance P receptor, NK-1, in the spinal cord of rats with peripheral nerve injury. Somatosens. Motor Res. (2001) 18:66-75.

90. LIN RC: Reactive astrocytes express substance-P immunoreactivity in the adult forebrain after injury. Neuroreport (1995) 7:310-312.

91. MANTYH PW, JOHNSON DJ, BOEHMER CG et al.: Substance P receptor binding sites are expressed by glia in vivo after neuronal injury. Proc. Natl. Acad. Sci. USA (1989) 86:5193-5197.

92. STUMM R, CULMSEE C, SCHAFER MK, KRIEGLSTEIN J, WEIHE E: Adaptive plasticity in tachykinin and tachykinin receptor expression after focal cerebral ischaemia is differentially linked to GABAergic and glutamatergic cerebrocortical circuits and cerebrovenular endothelium. J. Neurosci. (2001) 21:798-811.

93. WOIE K, KOLLER M, HEYERAAS KJ, REED RK: Neurogenic inflammation in rat trachea is accompanied by increased negativity of interstitial fluid pressure. Circ. Research (1993) 73:839-845.

94. YU Z, CHENG G, HUANG X, LI K, CAO X: Neurokinin-1 receptor antagonist SR140333: a novel type of drug to treat ischaemia. Neuroreport (1997) 8:2117-2119.

95. SOUZA DG, MENDONCA VA, DE ACMS, POOLE S, TEIXEIRA MM: Role of tachykinin NK receptors on the local and remote injuries following ischaemia and reperfusion of the superior mesenteric artery in the rat. Br. J. Pharmacol. (2002) 135:303-312.

96. VINK R, HU X, BENNETT C, CERNAK I, NIMMO AJ: Inhibition of neurogenic inflammation improves motor and cognitive outcome following diffuse traumatic brain injury. Rest. Neurol. Neurosci. (2000) 16:164.

97. HUSTON JP, HASENOHRL RU: The role of neuropeptides in learning: focus on the neurokinin substance P. Behav. Brain Res. (1995) 66:117-127.

98. BENOLIEL R, TANAKA M, CAUDLE RM, IADAROLA MJ: colocalization of N-methyl-D-aspartate receptors and substance P (neurokinin-1) receptors in rat spinal cord. Neurosci. Lett. (2000) 291:61-64.

99. YAO R, RAMESHWAR P, DONNELLY RJ, SIEGEL A: Neurokinin-1 expression and colocalization with glutamate and GABA in the hypothalamus of the cat. Brain Res. Mol. Brain Res. (1999) 71:149-158.

100. PARKER D, ZHANG W, GRILLNER S: Substance P modulates NMDA responses and causes long-term protein synthesis-dependent modulation of the lamprey locomotor network. J. Neurosci. (1998) 18:4800-4813.

101. GAUCHY C, DESBAN M, GLOWINSKI J, KEMEL ML: Distinct regulations by septide and the neurokinin-1 tachykinin receptor agonist [pro9]substance P of the N-methyl-D-aspartate-evoked release of dopamine in striosome- and matrix-enriched areas of the rat striatum. Neuroscience (1996) 73:929-939.

102. ANDERSON JJ, CHASE TN, ENGBER TM: Substance P increases release of acetylcholine in the dorsal striatum of freely moving rats. Brain Res. (1993) 623:189-194.

103. SHEN KZ, NORTH RA: Substance P opens cation channels and closes potassium channels in rat locus coeruleus neurons. Neuroscience (1992) 50:345-353.

104. PERSSON MG, HEDQVIST P, GUSTAFSSON LE: Nerve induced tachykinin-mediated vasodilation in skeltal muscle is dependent on nitric oxide formation. Eur. J. Pharmacol. (1991) 205:295-301.

105. HAFSTROM I, GYLLENHAMMER H, PALMBLAD J, RINGERTZ B: Substance P activates and modulates neutrophil oxidative metabolism and aggregation. J. Rheumatol. (1998) 16:1033-1037.

106. BEREITER DA, BEREITER DF, TONNESSEN BH, MACLEAN DB: Selective blockade of substance P or neurokinin A receptors reduces the expression of c-fos in trigeminal subnucleus caudalis after corneal stimulation in the rat. Neuroscience (1998) 83:525-534.

107. GOLD BG: Neuroimmunophilin ligands: evaluation of their therapeutic potential for the treatment of neurological disorders. Expert Opin. Investig. Drugs (2000) 9:2331-2342.

108. SHIGA Y, ONODERA H, MATSUO Y, KOGURE K: Cyclosporin A protects against ischaemia–reperfusion injury in the brain. Brain Res. (1992) 595:145-148.

109. UCHINO H, ELMER E, UCHINO K, LINDVALL O, SIESJO BK: Cyclosporin A dramatically ameliorates CA1 hippocampal damage following transient forebrain ischaemia in the rat. Acta Physiol. Scand. (1995) 155:469-471.

110. BRUSTOVETSKY N, DUBINSKY JM: Limitations of cyclosporin A inhibition of the permeability transition in CNS mitochondria. J. Neurosci. (2000) 20:8229-8237.

111. BUKI A, OKONKWO DO, POVLISHOCK JT: Postinjury cyclosporin A adminsitration limits axonal damage and disconnection in traumatic brain injury. J. Neurotrauma (1999) 16:511-521.

112. SCHEFF SW, SULLIVAN PG: Cyclosporin A significantly ameliorates cortical damage following experimental traumatic brain injury in rodents. J. Neurotrauma (1999) 16:783-792.

113. BORLONGAN CV, STAHL CE, KEEP MF, ELMER E, WATANABE S: Cyclosporin-A enhances choline acetyltransferase immunoreactivity in the septal region of adult rats. Neurosci. Lett. (2000) 279:73-76.

114. ALBENSI BC, SULLIVAN PG, THOMPSON MB, SCHEFF SW, MATTSON MP: Cyclosporin ameliorates traumatic brain-injury-induced alterations of hippocampal synaptic plasticity. Exp. Neurol. (2000) 162:385-389.

115. RIESS P, BAREYRE FM, SAATMAN KE et al.: Effects of chronic, postinjury Cyclosporin A administration on motor and

Expert Opin. Investig. Drugs (2002) 11(10) 1385

Novel therapies in development for the treatment of traumatic brain injury

sensorimotor function following severe, experimental traumatic brain injury. Restor. Neurol. Neurosci. (2001) 18:1-8.

• A manuscript describing the effects of cyclosporin A on functional outcome following TBI.

116. SULLIVAN PG, THOMPSON MB, SCHEFF SW: Cyclosporin A attenuates acute mitochondrial dysfunction following traumatic brain injury. Exp. Neurol. (1999) 160:226-234.

117. HAGL C, TATTON NA, KHALADJ N et al.: Involvement of apoptosis in neurological injury after hypothermic circulatory arrest: a new target for therapeutic intervention? Ann. Thorac. Surg. (2001) 72:1457-1464.

118. SINGLETON RH, STONE JR, OKONKWO DO, PELLICANE AJ, POVLISHOCK JT: The immunophilin ligand FK506 attenuates axonal injury in an impact-acceleration model of traumatic brain injury. J Neurotrauma (2001) 18:607-614.

119. DIAZ-RUIZ A, RIOS C, DUARTE I et al.: Lipid peroxidation inhibition in spinal cord injury: cyclosporin-A versus methylprednisolone. Neuroreport (2000) 11:1765-1767.

120. GITTER BD, WATERS DC, THRELKELD PG, LOVELACE AM, MATSUMOTO K, BRUNS RF: Cyclosporin A is a substance P (tachykinin NK1) receptor antagonist. Eur. J. Pharmacol. (1995) 289:439-446.

121. SULLIVAN PG, RABCHEVSKY AG, HICKS RR, GIBSON TR, FLETCHER-TURNER A, SCHEFF SW: Dose-response curve and optimal dosing regimen of cyclosporin A after traumatic brain injury in rats. Neuroscience (2000) 101:289-295.

•• This manuscript describes dose response curves for cyclosporin A administration in TBI.

122. FEIGENBAUM JJ, BERGMANN F, RICHMOND SA et al.: Nonpsychotropic cannabinoid acts as a functional N-methyl-D-aspartate receptor blocker. Proc. Natl. Acad. Sci. USA (1989) 86:9584-9987.

123. ESHHAR N, STRIEM S, KOHEN R, TIROSH O, BIEGON A: Neuroprotectant and antioxidant activities of HU-211, a novel NMDA receptor antagonist. Eur. J. Pharmacol. (1995) 283:19-29.

124. SHOHAMI E, GALLILY R, MECHOULAM R, BASS R, BEN HUR T:

Cytokine production in the brain following closed head injury: dexanabinol (HU-211) is a novel TNF-alpha inhibitor and an effective neuroprotectant. J. Neuroimmunol. (1997) 72:169-177.

•• Offers an excellent review of the potential for dexanabinol in TBI.

125. NADLER V, BIEGON A, BEIT-YANNAI E, ADAMCHIK J, SHOHAMI E: 45Ca accumulation in rat brain after closed head injury; attenuation by the novel neuroprotective agent HU-211. Brain Res. (1995) 685:1-11.

126. SHOHAMI E, NOVIKOV M, MECHOULAM R: A nonpsychotropic cannabinoid, HU-211, has cerebroprotective effects after closed head injury in the rat. J. Neurotrauma (1993) 10:109-119.

127. POP E: Dexanabinol pharmos. Curr. Opin. Investig. Drugs (2000) 1:494-503.

• A review of the use of dexanabinol in studies of TBI.

128. JARROTT B, CALLAWAY JK, JACKSON WR, BEART PM: Development of a novel arylalkylpiperazine compound (AM-36) as a hybrid neuroprotective drug. Drug Dev. Res. (1999) 46:261-267.

129. CALLAWAY JK, BEART PM, JARROTT B, GIARDINA SF: Incorporation of sodium channel blocking and free radical scavenging activities into a single drug, AM-36, results in profound inhibition of neuronal apoptosis. Br. J. Pharmacol. (2001) 132:1691-1698.

130. CALLAWAY JK, KNIGHT MJ, WATKINS DJ, BEART PM, JARROTT B: Delayed treatment with AM-36, a novel neuroprotective agent, reduces neuronal damage after endothelin-1-induced middle cerebral artery occlusion in conscious rats. Stroke (1999) 30:2704-2712.

•• Demonstrates the functional efficacy of a novel drug specifically designed to be multifactorial in its neuroprotective actions.

131. FADEN AI: Role of thyrotropin-releasing hormone and opiate receptor antagonists in limiting central nervous system injury. In: Advances in Neurology, Vol. 47: Functional Recovery in Neurological Disease. Waxman SG (Eds), Raven Press, New York (1988):531-546.

132. MCINTOSH TK, VINK R, FADEN AI: An analog of thyrotropin-releasing hormone improves outcome after traumatic brain

injury: 31P NMR studies. Amer. J. Physiol. (1988) 254:R785-R792.

133. VINK R, MCINTOSH TK, FADEN AI: Treatment with the thyrotropin-releasing hormone analog CG3703 restores magnesium homeostasis following traumatic brain injury in rats. Brain Res. (1988) 460:184-188.

134. DEWITT DS, PROUGH DS, UCHIDA T, DEAL DD, VINES SM: Effects of nalmefene, CG3703, tirilazad or dopamine on cerebral blood flow, oxygen delivery and electroencephalographic activity after traumatic brain injury and hemorrhage. J Neurotrauma (1997) 14:931-941.

135. MAEJIMA S, KATAYAMA Y: Neurosurgical trauma in Japan. World J. Surg. (2001) 25:1205-1209.

136. FADEN AI, FOX GB, FAN L et al.: Novel TRH analog improves motor and cognitive recovery after traumatic brain injury in rodents. Am. J. Physiol. (1999) 277:R1196-R1204.

137. SCHMID-ELSAESSER R, ZAUSINGER S, HUNGERHUBER E, BAETHMANN A, REULEN HJ: Neuroprotective effects of combination therapy with tirilizad and magnesium in rats subjected to reversible focal cerebral ischaemia. Neurosurg. (1999) 44:163-172.

• A description of one of the successful uses of polypharmacy in an effort to attenuate multiple secondary injury factors following ischaemia.

138. GULUMA KZ, SAATMAN KE, BROWN A, RAGHUPATHI R, MCINTOSH TK: Sequential pharmacotherapy with magnesium chloride and basic fibroblast growth factor after fluid percussion brain injury results in less neuromotor efficacy than that achieved with magnesium alone. J. Neurotrauma (1999) 16:311-312.

AffiliationRobert Vink1† and Alan J Nimmo2

1Department of Pathology, The University of Adelaide, Adelaide, South Australia 5005, Australia2School of Pharmacy and Molecular Sciences, James Cook University, Australia†Author for correspondenceTel: +61 8 8222 3092;Fax: +61 8 8222 3093;E-mail: [email protected]

1386 Expert Opin. Investig. Drugs (2002) 11(10)