Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Feb 19.
Published in final edited form as: Nat Rev Drug Discov. 2009 Nov;8(11):892–909. doi: 10.1038/nrd2999

Targeting protein kinases in central nervous system disorders

Laura K Chico *,, Linda J Van Eldik *,§, D Martin Watterson *,
PMCID: PMC2825114  NIHMSID: NIHMS174179  PMID: 19876042

Abstract

Protein kinases are a growing drug target class in disorders in peripheral tissues, but the development of kinase-targeted therapies for central nervous system (CNS) diseases remains a challenge, largely owing to issues associated specifically with CNS drug discovery. However, several candidate therapeutics that target CNS protein kinases are now in various stages of preclinical and clinical development. We review candidate compounds and discuss selected CNS protein kinases that are emerging as important therapeutic targets. In addition, we analyse trends in small-molecule properties that correlate with key challenges in CNS drug discovery, such as blood–brain barrier penetrance and cytochrome P450-mediated metabolism, and discuss the potential of future approaches that will integrate molecular-fragment expansion with pharmacoinformatics to address these challenges.


Protein kinases regulate diverse cellular functions through the orchestrated propagation and amplification of cellular stimuli into distinct biological responses through coordinated signal transduction cascades. With several hundred kinases encoded in the human genome, almost every signal transduction process is influenced by interconnected phosphorylation events. Deregulation of kinase activity has been implicated in various diseases, ranging from vascular disorders and inflammatory diseases to neurological disorders and cancer1,2. This has generated intense interest in the pursuit of protein kinases as drug targets.

However, most kinase-targeted drugs and potential kinase targets that have been investigated are for non-central nervous system (CNS) disorders. CNS disease indications for kinase-targeted drugs seem to be lagging behind those for other disease areas, such as cancer (FIG. 1), and ~25% of publications related to CNS disorders are for CNS cancers. This pattern is also seen in pharmaceutical industry pipelines that are publicly disclosed. For example, Novartis, the manufacturer of the kinase inhibitor cancer therapeutics imatinib (Gleevec) and nilotinib (Tasigna), has various oncology candidates and CNS-targeted therapies in clinical development. Half of the oncology pipeline are kinase inhibitors, but none of the disclosed CNS candidates seems to target protein kinases. Nevertheless, there is increasing interest in the development of kinase-targeted therapeutics for CNS indications3-11, and the established success in other disease indications provides encouragement for the development of small-molecule kinase modulators for CNS disorders.

Figure 1. Trends in drug discovery and development efforts targeting protein kinases.

Figure 1

The number of publications since 1990 on small-molecule protein kinase inhibitors that have been evaluated in animal efficacy models and clinical trials for oncology indications was determined by a literature search with SciFinder Scholar. The number steadily increased until the recent plateau in 2007. The orange bar denotes the number of publications on central nervous system (CNS) disease indications in 2008. This activity level is comparable to that for the cancer field in 1999. The curve shows several small inflection points that correlate with milestones related to the kinase inhibitor imatinib (Gleevec; Novartis). For example, the first reports of in vivo efficacy of imatinib appeared in 1995 and 1996, and shortly thereafter the slope of the curve showed a small increase. An increased slope is also seen after the first clinical trials with imatinib were initiated in 1998 and after drug approval in 2001. Before 2001, only one kinase inhibitor was approved as a drug (fasudil; approved in Japan in 1995 for cerebral vasospasm). Since 2001, eight additional kinase inhibitors have received approval for oncology indications: gefitinib (Iressa; AstraZeneca) was approved in the United States in 2003; erlotinib (Tarceva; OSI/Genentech/Roche) was approved in the United States and Europe in 2004; sorafenib (Nexavar; Bayer/Onyx) was approved in the United States and Europe in 2005; sunitinib (Sutent; Pfizer) and dasatinib (Sprycel; Bristol–Myers Squibb) were approved in the United States and Europe in 2006; and temsirolimus (Torisel; Wyeth) was approved in the United States, and lapatinib (Tykerb; GlaxoSmithKline) and nilotinib (Tasigna; Novartis) were approved in the United States and Europe in 2007.

This Review focuses on the special challenge of targeting protein kinases for CNS disease indications. A recent review2 provides an up-to-date overview of targeting protein kinases in cancer, presents an in-depth analysis of the structural basis of protein kinase inhibitor recognition and selectivity, and discusses various general approaches to kinase inhibitor design. Therefore, the broader aspects of protein kinase inhibitor interactions in drug discovery are not a focus of this article. In addition, not all protein kinase targets can be covered in a single review. Instead, we provide an overview of promising protein kinase targets for CNS disease indications in which the disclosed state of small-molecule inhibitor development for these targets is at the in vivo stage of evaluation (FIG. 2; TABLES 1,2; see Supplementary information S1, S2 (box, table)), and summarize selected case studies to illustrate the issues that often arise during kinase or CNS drug discovery programmes. We conclude with a discussion of the challenges of developing small-molecule therapeutic candidates that are capable of penetrating the blood–brain barrier (BBB) and the potential of emerging approaches, such as molecular-properties-driven fragment expansion, to address these challenges.

Figure 2. Selected chemical structures of small-molecule kinase inhibitors in clinical trials for CNS indications.

Figure 2

See TABLE 1 for further information.

Table 1. Protein kinase targets for CNS indications: small-molecule compounds in clinical trials.

Kinase Type Disease indication Clinical trial
identifier (Phase)
Example compound (company)
BCR–ABL Y Glioma NCT00021229 (I–II),
NCT00010049 (I–II)*
Imatinib (Novartis)
NCT00423735 (II) Dasatinib (Bristol–Myers Squibb)
KIT Y Glioma NCT00021229 (I–II)
NCT00010049 (I–II)*
Imatinib (Novartis)
NCT00329719 (I–II) Sorafenib (Bayer)
Neurofibromatosis type 1 NCT00727233 (I)
Glioma NCT00459381 (II) Pazopanib (GlaxoSmithKline)
NCT00423735 (II) Dasatinib (Bristol–Myers Squibb)
Meningioma NCT00589784 (II) Sunitinib (Pfizer)
Glioma NCT00387933 (I) Vatalanib (Novartis)
Von Hippel–Lindau-related
haemangioblastoma
NCT00052013 (II)
Meningioma NCT00348790 (II)
EGFR1 Y Glioma NCT00025675 (II) Gefitinib (AstraZeneca)
NCT00054496 (II) Erlotinib (Genentech)
EGFR2 Y Glioma NCT00107237 (I–II) AEE-788 (Novartis)
FLT3 Y Meningioma NCT00589784 (II) Sunitinib (Pfizer)
GSK3 S–T Alzheimer’s disease NCT00088387 (II) Lithium (several companies)
No identifier (II) NP-12 (Noscira)
Depression No identifier (Phase
unreported)
NP031115 (Noscira)
MLK1, MLK2
and MLK3
S–T Parkinson’s disease NCT00040404 (II–III) CEP-1347 (Cephalon)
mTOR S–T Glioma NCT00831324 (II) Everolimus (Novartis)
NCT00047073 (I–II)
NCT00457808 (I/II)
Sirolimus
(Wyeth)
NCT00022724 (I–II) Temsirolimus (Wyeth)
p38α S–T Neuropathic pain NCT00390845 (II) SB-681323 (GlaxoSmithKline)
PDGFRβ Y Glioma NCT00021229 (I–II)
NCT00010049 (I–II)*
Imatinib (Novartis)
NCT00329719 (I–II) Sorafenib (Bayer)
Neurofibromatosis type 1 NCT00727233 (I)
Glioma NCT00459381 (II) Pazopanib (GlaxoSmithKline)
NCT00423735 (II) Dasatinib (Bristol-Myers Squibb)
Meningioma NCT00589784 (II) Sunitinib (Pfizer)
Glioma NCT00387933 (I) Vatalanib (Novartis)
Von Hippel–Lindau-related
haemangioblastoma
NCT00052013 (II)
Meningioma NCT00348790 (II)
PDHK S–T Glioma NCT00703859 (I) Dichloroacetate (several companies)
PKC S–T Bipolar disorder NCT00026585 (II)
NCT00411203 (III)
Tamoxifen (AstraZeneca)
RAF S–T Glioma NCT00329719 (I/II) Sorafenib (Bayer)
Neurofibromatosis type 1 NCT00727233 (I)
ROCK S–T Raynaud’s phenomenon NCT00498615 (III) Fasudil (Asahi Kasei)
SRC Y Glioma NCT00423735 (II) Dasatinib (Bristol–Myers Squibb)
VEGFR1 Y Glioma NCT00459381 (II) Pazopanib (GlaxoSmithKline)
NCT00006247 (I) Semaxanib/SU5416 (Pharmacia)§
NCT00387933 (I) Vatalanib (Novartis)
Von Hippel–Lindau-related
haemangioblastoma
NCT00052013 (II)
Meningioma NCT00348790 (II)
VEGFR2 Y Glioma NCT00107237 (I–II) AEE-788 (Novartis)
NCT00329719 (I–II) Sorafenib (Bayer)
Neurofibromatosis type 1 NCT00727233 (I)
Glioma NCT00459381 (II) Pazopanib (GlaxoSmithKline)
NCT00006247 (I) Semaxanib/SU5416 (Pharmacia)§
Meningioma NCT00589784 (II) Sunitinib (Pfizer)
Glioma NCT00387933 (I) Vatalanib (Novartis)
Von Hippel–Lindau-related
haemangioblastoma
NCT00052013 (II)
Meningioma NCT00348790 (II)
VEGFR3 Y Glioma NCT00459381 (II) Sorafenib (Bayer)
Neurofibromatosis type 1 NCT00727233 (I)
Glioma NCT00459381 (II) Pazopanib (GlaxoSmithKline)
NCT00387933 (I) Vatalanib (Novartis)
Von Hippel–Lindau-related
haemangioblastoma
NCT00052013 (II)
Meningioma NCT00348790 (II)
*

The trial has been completed but the therapy was not efficacious.

The trial was terminated early owing to a failure to delay disability.

§

Development has been discontinued owing to a lack of clinical efficacy in colorectal cancer trials. EGFR, epidermal growth factor receptor; ERK, extracellular signal-regulated kinase; FLT, FMS-related tyrosine kinase; GSK, glycogen synthase kinase; JAK3, Janus kinase 3; JNK, c-Jun N-terminal protein kinase; MEK, mitogen-activated protein kinase (MAPK)–ERK kinase; MLK, mixed lineage kinase; mTOR, mammalian target of rapamycin; PDGFR, platelet-derived growth factor receptor; PDHK, pyruvate dehydrogenase kinase; PKC, protein kinase C; ROCK, Rho-associated protein kinase; S–T, serine–threonine; VEGFR, vascular endothelial growth factor receptor; Y, tyrosine.

Table 2. Kinase targets for CNS indications: small molecules with preclinical in vivo efficacy data.

Kinase Type Disease indication Example compounds Refs
AMPK S–T Multiple sclerosis Arasine 144
Cerebral ischaemia Compound C 145
CDK S–T Traumatic brain injury Roscovitine 147
Cerebral ischaemia Flavopiridol and indolinone A 148,149
DAPK1 S–T Acute brain injury 11-(3-imino-5,6-dihydrobenzo[h]cinnolin-2(3H)-yl)-
N-(6-phenylpyridazin-3-yl)undecanamide
41,43
Alzheimer’s disease 11-(3-imino-5,6-dihydrobenzo[h]cinnolin-2(3H)-yl)-
N-(6-phenylpyridazin-3-yl)undecanamide
42
GSK3 S–T Alzheimer’s disease SB-216763 150
AR-AO14418 150
Lithium 150,151
Indirubin-3′-monoxime 150
Alsterpaullone 150
Chir98014 150
NP-12* 34
Amyotrophic lateral sclerosis AR-AO14418 26
Bipolar disorder Multiple 3-benzofuranyl-4-indolylmalemide 152
Cerebral ischaemia Compound 1 153
Chir025* 154
TDZD-8 30
AR-AO14418 28
Depression NP031115* 32
HIV-associated dementia Valproic acid 155
Parkinson’s disease AR-AO14418 156
Indirubin-3′-oxime ((E)-3-(hydroxyimino)-2,3′-biindolin-
2′-one)
156
Neurocognitive deficits after
cranial irradiation
SB-415286 157
SB-216763 157
Shock TDZD-8 30
Spinal cord injury TDZD-8 31
Traumatic brain injury Lithium 158
JAK3 S–T Amyotrophic lateral sclerosis WHI-P131 159
JNK1,
JNK2
and
JNK3
S–T Subarachnoid haemorrhage SP600125 87
Parkinson’s disease SP600125 160
Cerebral ischaemia AS601245 86
MEK1
and
MEK2
S–T Behavioral disorders and drug
abuse
SL 327 161
Cerebral ischaemia U0126 82
PD-98059 162,163
Traumatic brain injury PD-98059 85
Neuropathic pain PD 198306 164
MLK1,
MLK2
and
MLK3
S–T Excitotoxic injury CEP-1347 165
Hearing loss CEP-1347 89
Huntington’s disease CEP-1347 90,166
CEP-11004 90
Motor neuron disease CEP-1347 167
Parkinson’s disease CEP-5104, CEP-6331 and CEP-1347 94
MLCK S–T Cerebral vasospasm ML-9 168
mTOR Y Autism Rapamycin 169
Huntington’s disease Temsirolimus 170
Tuberous sclerosis Rapamycin 169
p38α
MAPK
S–T Alzheimer’s disease MW01-2-069ASRM 58
Amyotrophic lateral sclerosis SB203580 59
Cerebral ischaemia SB203580 60
SB239063 61; 62
Neuropathic pain FR167653 171
Parkinson’s disease SB239063 70
Spinal cord injury SB203580 172
Cerebral ischaemia VX-745 173
Traumatic brain injury SB203580 85
PKA S–T Alzheimer’s disease H89 174
Memory impairment Halofantrine 175
Morphine tolerance KT 5270 176
Neuropathic pain Triterpene mixture 177
PKC S–T Alzheimer’s disease GF-109203X 178
Bipolar disorder Tamoxifen 179
Brain tumours LY 333531 180
Cerebral ischaemia Staurosporine 181
GF-109203X 178
Ebselen 182
Cerebral vasospasm H-7 and calphostin C 183
Cheleritrine and rottlerin 184
Neuropathic pain Staurosporine and calphostin C 181
Opioid dependence H-7 185
Parkinson’s disease Rottlerin 186
PLK2 S–T Parkinson’s disease BI2536 187
ROCK S–T Alzheimer’s disease Y-27632 188
Cerebral ischaemia Fasudil and hydroxyfasudil 189, 190
Multiple sclerosis Fasudil 191
Epilepsy Y-27632 192
Spinal canal stenosis Fasudil 171
Pain Y-27632 193
Spinal cord injury Y-27632 194
SNRPE S–T Cerebral ischaemia SB-699393 146
SRC Y Brain injury PP-1 195
Intracerebral haemorrhage PP-2 196
TGFβR Y Glioblastoma SX-007* 197
*

The structure has not been disclosed. AMPK, AMP-activated kinase; CDK, cyclin-dependent kinase; DAPK1, death-associated protein kinase 1; GSK3, glycogen synthase kinase 3; JAK3, Janus kinase 3; MAPK, mitogen-activated protein kinase; MEK, MAPK–ERK kinase; MLCK, myosin light chain kinase; MLK, mixed lineage kinase; mTOR, mammalian target of rapamycin; PKA, protein kinase A; PKC, protein kinase C; PLK2, polo-like kinase 2; ROCK, Rho-associated protein kinase; SNRPE, small nuclear ribonucleoprotein polypeptide E (also known as B-RAF); TGFβR, transforming growth factor-β receptor.

Examples of CNS protein kinase targets

Glycogen synthase kinase 3 (GSK3)

GSK3 is a serine–threonine protein kinase family consisting of two isoforms, GSK3α and GSK3β, that are similar in their catalytic domain structures, substrate specificity, tissue distribution and molecular mechanisms of regulation (reviewed in REFS 4-7). GSK3 regulates a wide range of cellular functions, such as glucose metabolism, gene expression, cytoskeletal organization, vesicular transport, cell growth, motility, differentiation and cell survival (FIG. 3), which are mediated through diverse substrates and signalling pathways. The activity of GSK3 can be regulated by several integrated mechanisms, such as coordinated phosphorylation of the enzyme and its substrates, dynamic regulation of the subcellular distribution of the kinase and the formation of complexes of the enzyme with scaffolding or anchoring proteins that control access to specific substrates5-7 (BOX 1).

Figure 3. Protein kinase targets as components of intracellular signal transduction networks.

Figure 3

Glycogen synthase kinase 3 (GSK3) is an example of a widely distributed target that modulates diverse cellular functions. Death-associated protein kinase 1 (DAPK1) has a more restricted cellular role, but both are components of an integrated web of pathways. Selective biological responses are possible because such networks can transduce various cellular stimuli into an integrated response, through mechanisms that are incompletely understood. The target kinases represent network nodes, the quantitative importance of which is altered under conditions of deregulation. The activity of GSK3 is regulated by several mechanisms that allow selective phosphorylation of substrates in response to distinct input signals. The regulatory mechanisms include modulation of GSK3 activity by inhibitory serine phosphorylation and activating tyrosine phosphorylation of the enzyme, coordinated phosphorylation of its substrates, dynamic regulation of the subcellular distribution of GSK3 and the assembly of GSK3 into distinct multi-protein complexes for selective substrate phosphorylation. DAPK1 regulates signalling pathways that are important to cell survival and translational control through phosphorylations of a highly selective set of substrate proteins in response to its activation by distinct cellular stimuli, such as increased Ca2+ levels, which are sensed through its calmodulin (CAM) subunit. Examples of central nervous system functions that have been proposed for DAPK1 and GSK3 are shown in orange. Aβ, amyloid-β; BAX, BCL2-associated X protein; CAMK1, Ca2+–calmodulin-dependent protein kinase 1; CAMKK, CAMK kinase; C/EBP, CCAAT/enhancer-binding protein; CREB, cyclic AMP response element binding protein; ECM, extracellular matrix; EIF2B, eukaryotic initiation factor 2B; ERK, extracellular signal-regulated kinase; HSF1, heat shock factor 1; IRS1, insulin receptor substrate 1; JNK, Jun I-terminal kinase; MAP1B, microtubule-associated protein 1B; MAPK, mitogen-activated protein kinase; MLC, myosin light chain; MRPL13, mitochondrial ribosomal protein L13; MRPS6, mitochondrial ribosomal protein S6; MYC, myelocytomatosis viral oncogene homologue; NFAT, nuclear factor of activated T cells; NF-κB, nuclear factor κB; phospho–S, phosphorylated serine; phospho–Y, phosphorylated tyrosine; PI3K, phosphoinositide 3-kinase; PKA, protein kinase A; PKC, protein kinase C; PKD, protein kinase D; PYK2, proline-rich tyrosine kinase 2; TP53, tumour protein p53; RPS6KA1, ribosomal protein S6 kinase, 90kDa, polypeptide 1.

Box 1 | Regulation of GSK3 activity.

Glycogen synthase kinase 3 (GSK3) activity can be regulated by several integrated mechanisms, such as coordinated phosphorylation of the enzyme and its substrates, dynamic regulation of its subcellular distribution and the formation of complexes of the enzyme with scaffolding or anchoring proteins that control access to specific substrates (reviewed in REFS 5-7). For example, exposure of cells to growth factors, insulin, receptor ligands and other stimuli that activate intracellular kinases — such as phosphoinositide 3-kinase–AKT, protein kinase A and protein kinase C — leads to inhibition of the constitutively active GSK3 through an inhibitory phosphorylation of an amino-terminal serine in the enzyme (ser21 in GSK3α and ser9 in GSK3β). In contrast to inhibition of enzyme activity by serine phosphorylation, GSK3 enzymatic activity is enhanced by tyrosine phosphorylation (tyr279 in GSK3α and tyr216 in GSK3β). The mechanisms regulating tyrosine phosphorylation of GSK3 are not well defined, but may occur through autophosphorylation and/or by other kinases6.

In addition to the regulation of GSK3 pathways by phosphorylation of the kinase, physiological responses can also be regulated by the phosphorylation state of GSK3 substrates. The mechanism of action for this mode of regulation is the requirement that most GSK3 substrates must be pre-phosphorylated, or ‘primed’, to be phosphorylated effectively by GSK3. Another mechanism of GSK3 pathway regulation is to limit the substrates upon which GSK3 acts by dynamic regulation of the subcellular localization of the enzyme, such as in the nucleus and mitochondria. The association of GSK3 with distinct multiprotein complexes in the cell, including in subcellular organelles, ensures that the action of GSK3 in response to particular stimuli is directed towards specific substrates. These integrated mechanisms allow coordinated and restricted substrate-specific phosphorylation, thereby enabling GSK3 to act on a diverse array of protein substrates while producing selective biological responses to discrete stimuli.

The involvement of GSK3 in numerous cellular processes, coupled with the complexity of the mechanisms for its regulation, underscore the importance of appropriate functioning of GSK3-modulated signalling. Not surprisingly, upregulation of GSK3 activity has been linked to numerous human pathological conditions, including diabetes, muscle hypertrophy, cancer, Alzheimer’s disease, stroke, sleep disorders, and neuropsychiatric and mood disorders (reviewed in REFS 4,9,12-14). Although regulation of GSK3 is generally through modulation of enzymatic activity, there are also examples of increased GSK3 activity in Alzheimer’s disease through increased levels of the protein or changes in splicing of the gene15,16. The involvement of GSK3 in several human pathological conditions raised interest in the kinase as a potential drug discovery target. For example, several studies have linked GSK3 deregulation to Alzheimer’s disease pathophysiology, and GSK3 seems to contribute to both the amyloid and tau pathologies that characterize the disease5,12,17,18.

The potential of GSK3 as an Alzheimer’s disease drug discovery target has been raised by studies showing that overexpression of GSK3β in transgenic mouse models results in increased tau phosphorylation and deficits in spatial learning, and that inhibition of GSK3β activity leads to neuroprotective effects, decreased amyloid-β production and a reduction in tau hyperphosphorylation4,12,18. Lithium, a mood stabilizer that is widely used in the treatment of bipolar disorder, is a GSK3 inhibitor19 and is currently in an alternative-use clinical trial for Alzheimer’s disease (TABLE 1). However, lithium is associated with adverse effects, especially in the elderly. For example, in an open-label clinical study20 assessing the feasibility and tolerability of lithium at therapeutic doses administered for up to 1 year in patients with mild to moderate Alzheimer’s disease, a significant percentage of patients had contraindications and there was a high discontinuation rate from the trial. The occurrence of some adverse effects and the risk of serious toxicity in elderly patients could limit the potential usefulness of lithium therapy for Alzheimer’s disease. Lithium apparently reduces GSK3 activity through both a direct inhibition of enzyme activity, by competing with Mg2+ ions, as well as by increasing the inhibitory phosphorylation of GSK3 that inactivates the enzyme21,22. Whether or not lithium exerts its therapeutic effects through GSK3 inhibition is not known, as lithium affects additional enzymes. However, studies with lithium created interest in the kinase as a CNS drug target and provided a rationale for the subsequent development of new, more selective GSK3 inhibitors. More selective maleimide-based lithium mimetics, for example, are active in a mouse model of mania23 and in a cell-based model of Parkinson’s disease24.

Currently, chemically diverse small-molecule GSK3 inhibitors have shown efficacy in preclinical animal models of various CNS disorders (TABLE 2). These include indirubins, paullones, aminopyrazoles, thiazoles, and thiadiazolidinones (TDZDs). The thiazole AR-A014418 is efficacious in animal models of Alzheimer’s disease9, Parkinson’s disease25, amyotrophic lateral sclerosis (ALS)26, depression27 and ischaemic stroke28. Small heterocyclic TDZDs are being explored for their potential as therapeutics for Alzheimer’s disease and other ‘tauopathies’29, ischaemic injury30, spinal cord trauma31, mood disorders32 and excitotoxic conditions33. For example, a TDZD compound, NP12, is efficacious in an Alzheimer’s disease mouse model34, has successfully completed a Phase I clinical trial in Europe, and is currently in a Phase II clinical trial for Alzheimer’s disease (TABLE 1). Thus, several GSK3 inhibitors are well into the drug development pipeline for CNS indications35, and the final validation of GSK3 as a target for CNS disorders awaits the outcomes on therapeutic efficacy in human clinical trials.

Targeting GSK3 for CNS drug discovery exemplifies key considerations in kinase inhibitor drug discovery. GSK3 is a widely distributed protein kinase with many substrates and signalling pathway interactions. This raised concerns that chronic inhibition of GSK3 might lead to oncogenic side effects related to its proliferative targets, such as β-catenin or certain transcription factors. However, chronic treatment in animals and humans has not elicited these potential side effects36,37, which is consistent with the prevailing view that oncogenesis is more than a ‘single hit’ response. GSK3-targeted therapeutic approaches show how inhibitor treatment can preferentially modulate one pathway over another through the use of low doses to attain the desired efficacy. The selective response is also related to the common issue of what proportion of the total kinase activity in a tissue is required to mediate a physiological effect or, in the case of inhibitor treatment, to mediate a pharmacological effect9,38. The levels of GSK3 inhibition that are required to obtain neurological effects seem to be much lower than those required for stabilization of β-catenin, and approximately 20–25% inhibition of GSK3 is sufficient for therapeutic efficacy in CNS diseases7,37. The issues relating to kinase inhibitor development discussed in this section for GSK3 as a case study are also relevant to the cyclin-dependent kinases (CDKs), protein kinase A and protein kinase C (TABLE 2).

Death-associated protein kinase 1 (DAPK1)

DAPK1 is a pro-apoptotic, calmodulin-regulated, serine–threonine protein kinase that acts early in the apoptosis pathways before the cell is committed to death39,40. DAPK1 was recently identified as a CNS drug discovery target that is implicated in post-injury synaptic dysfunction, an end point of pathology progression that can be attenuated in animal models with DAPK1 inhibitor treatments within clinically relevant time frames. DAPK1, like GSK3, is a protein kinase with several biological roles. It provides an early example of the use of the co-crystal structure of an inactive small-molecule fragment bound to the target kinase as the starting point for bioavailable protein kinase inhibitor design41, allowing in vivo target validation studies to be performed in animal models41-43.

DAPK1 was first identified as a CNS drug discovery target in animal models of CNS injury. More recent human genetics studies show a link between human disease and DAPK1 that is consistent with the animal model studies. DAPK1 is a druggable kinase target in animal models of both acute and sustained brain injury41-43. Its enzyme activity in tissue homogenates increases after injury during a therapeutically relevant time frame40, and small-molecule intervention during this time modifies disease progression, as assessed by subsequent neurological outcomes41. Specifically, a single treatment of rats with a small-molecule DAPK1 inhibitor (TABLE 2) 6 hours after cerebral ischaemia attenuated the loss of brain tissue, measured 1 week later41. Consistent results were obtained with the same DAPK1 inhibitor in a different cerebral ischaemia animal model43. More recently, human clinical genetics studies44 identified that the DAPK1 gene locus is linked to late-onset Alzheimer’s disease susceptibility. Specifically, two single nucleotide polymorphisms (SNPs) were associated with DAPK1 allele-specific expression and late-onset Alzheimer’s disease. It is not yet known whether the SNPs in the human DAPK1 gene alter the expression or the activity of the kinase. A meta-analysis of genetic variation in case-control samples provided further evidence that DAPK1 genetic mutations may affect disease susceptibility. The clinical linkage and animal pathophysiology studies are consistent with pharmacological studies in animal models42.

The identification of important protein substrates of DAPK1 in the CNS indicated molecular mechanisms by which DAPK1 could be involved in CNS pathophysiology progression (FIG. 3). For example, the finding that DAPK1 can phosphorylate and inactivate the neuronal survival protein Ca2+–calmodulin-dependent protein kinase kinase 1 (CAMKK1)45, and that phosphorylation of ribosomal protein S6 by DAPK1 can alter neuronal protein biosynthesis46, suggest that DAPK1 is involved in stress- or injury-induced phosphorylation cascades that can lead to synaptic dysfunction or neuronal death. DAPK1 has also been reported to interact with numerous other signalling proteins and pathways, such as DAPK3, protein kinase D, myosin light chains, beclin, syntaxin 1A and extracellular signal-regulated kinase 1 (ERK1; also known as MAPK3), some of which may be related to its pro-apoptotic functions47-52. The available evidence suggests that DAPK1 may have several physiological roles, and can promote cell death and synaptic dysfunction by inhibiting survival pathways once activated by various stress or injury stimuli. Although initial target validation evidence with bioavailable kinase inhibitors that are not drug candidates supports DAPK1 as a drug discovery target for neurological disorders, no clinically promising small-molecule DAPK1 inhibitor drugs have yet been disclosed.

The mitogen-activated protein kinase (MAPK) family

The MAPKs are serine–threonine protein kinases that integrate and process extracellular stimuli through a series of intracellular signalling complexes and phosphorylation cascades that lead to coordinated and distinct responses to many diverse stimuli (reviewed in REF. 53). In its simplest form, the MAPK cascade consists of a three-tiered set of protein kinases: a MAPK (ERKs, c-Jun N-terminal kinases (JNKs) and p38s), and two upstream components (a MAPK kinase and a MAPK kinase kinase) that activate the MAPKs by a series of activating phosphorylations. The activated MAPK can phosphorylate a number of substrates, which leads to stimulus-specific responses. This description of MAPK cascades as linear, isolated pathways is oversimplified: MAPK pathways can influence and be influenced by other signalling pathways, by interactions with scaffolding proteins and by specific localizations within cells54. In addition, each MAPK family consists of several isoforms that can have distinct functions54. For example, the p38α MAPK isoform (also known as MAPK14) has a crucial role in the activation of inflammatory cell signalling cascades, and its position at the interface of diverse signalling pathways allows it to modulate downstream biological responses through several regulatory mechanisms (recently reviewed in REFS 55-57). Several kinases in the MAPK signalling cascades have emerged as potential CNS therapeutic targets, including ERK1, ERK2, JNK1–JNK3, MAPK–ERK kinase 1 (MEK1), MEK2, mixed-lineage kinase 1 (MLK1)–MLK3 and p38α (TABLE 2).

The p38 MAPK family is being investigated as a potential therapeutic target for many CNS disorders. Diverse chemical classes of small-molecule p38α MAPK inhibitors have been tested in preclinical animal models of CNS diseases, including Alzheimer’s disease, ALS, cerebral ischaemia and neuropathic pain58-63. The p38α MAPK is an especially interesting case study for several reasons. First, the enzyme is an established therapeutic target for peripheral inflammatory disorders such as rheumatoid arthritis64, and the clinical correlations between kinase activation and CNS disorders make p38α MAPK a logical target for future CNS drug discovery and development programmes. For example, the p38 MAPK pathway is activated in human brain tissue in several CNS disorders, including Alzheimer’s disease65-68, Down’s syndrome68, Parkinson’s disease69,70, tauopathies71,72 and gliomas73. Second, the use of drug-resistant p38α knock-in mice (BOX 2) provides a rare precedent for in vivo target validation of a small-molecule drug64. Third, inhibition of increased production of proinflammatory cytokines or other injurious mediators offers the potential for an extended pharmacodynamic effect, due to the modulation of a biosynthetic process74. Fourth, the development of p38α MAPK inhibitors is a case study of how molecular properties of small-molecule inhibitors can affect bioavailability, including uptake by the brain58. Fifth, it provides an example of an opportunity to use complementary aspects of enzyme and small-molecule structures to generate selectivity among enzyme isoforms and the rest of the kinome (BOX 2)75-79.

Box 2 | Generation of small-molecule p38 MAPK inhibitors.

The p38α mitogen-activated protein kinase (MAPK) is a serine–threonine protein kinase that plays a crucial role in the regulation of inflammatory cell signalling cascades. Its position at the interface of diverse signalling pathways allows it to modulate downstream biological responses through several regulatory mechanisms (reviewed recently in REFS 55-57). The p38α and p38β MAPK isoforms have key structural differences in their active sites compared with most other protein kinases and other p38 MAPK isoforms. This can be exploited to generate selective small-molecule inhibitors that bind within the ATP fold58,77. Specifically, p38α and p38β have a hydrophobic pocket with a ‘gatekeeper’ threonine, an amino acid with a small side chain, whereas most other kinases have a larger amino acid, such as methionine, at this locus78. The small side chain allows bulky substituents on diverse small molecules to occupy the pocket, which would be restricted in kinases with the larger gatekeeper amino acid. This structural feature can be combined with a nearby interaction that is potentially common to many kinases, such as H-bond interactions available close-by in three-dimensional space, to generate selective small-molecule inhibitors76,79.

The three-dimensional selectivity filter has been used64 to generate knock-in mouse strains that express only a mutant p38(T106M) kinase that had normal p38 MAPK enzyme activity but could no longer bind small-molecule p38 MAPK inhibitors that exploit this structural feature (this is an example of designed drug resistance). These p38-knock-in mice provided valuable tools to show that the in vivo efficacy of an experimental therapeutic agent was through inhibition of its target kinase — in this case, p38α MAPK64.

The advancement of p38 MAPK inhibitors into clinical trials for diseases of peripheral tissues had early problems with unacceptable safety profiles80. However, it is likely that the early adverse effects were due to chemical toxicity or off-target interactions that seem to be compound specific rather than target related, and the more recent success emerging from the use of new chemotypes or scaffolds is consistent with this proposal (reviewed in REF. 55). Insights into the mechanisms by which p38α MAPK contributes to proinflammatory cytokine overproduction in the CNS suggest that this kinase could also be a valid target for neurological disorders (reviewed in REF. 3). For example, the p38 MAPK pathway is activated in neurons and glia in the brains of patients with early-stage Alzheimer’s disease and in rodent models of neurodegenerative disease, as assessed by staining for phosphorylated (activated) p38 MAPK. Furthermore, increased p38 MAPK expression and activity have been linked to glial proinflammatory cytokine production and neuronal tau phosphorylation and synaptic dysfunction65-67,81. Initial causative links between p38α MAPK activation and CNS pathophysiology have been provided by preclinical studies using small-molecule p38α MAPK inhibitors in animal models of various CNS disorders, as discussed above (TABLE 2). The available data make a compelling argument for further exploring p38α MAPK as a therapeutic target for the treatment of CNS diseases.

Other branches of the MAPK pathways are also being investigated as potential CNS targets. For example, compounds such as PD-98059, U0126 and PD198306, which block ERK1 and ERK2 activation through inhibition of upstream kinases MEK1 and MEK2 have shown efficacy in preclinical animal models of cerebral ischaemia, traumatic brain injury and neuropathic pain82-85. Similarly, inhibitors of the JNK pathway are efficacious in various animal models of neurodegenerative disorders, including cerebral ischaemia, subarachnoid haemorrhage, motor neuron disease, excitotoxic cell death, age-dependent hearing loss, Huntington’s disease and Parkinson’s disease86-94. However, an inhibitor of MLK that blocks JNK activation (CEP-1347) failed to show efficacy in the recent Parkinson Research Evaluation of CEP-1347 Trial (PRECEPT) of early Parkinson’s disease95. It is unclear why the drug was ineffective in delaying disability. Various explanations have been proposed, such as failure of the inhibitor to reach therapeutic levels in the CNS, an insufficient therapeutic window for rescue of neuronal apoptosis or the targeting of a less optimal step or kinase pathway than the intended target. Another possible explanation was that additional signalling pathways need to be targeted to effectively prevent dopaminergic neuron death96,97. An interesting suggestion has been made in relation to this76, that therapy with an MLK inhibitor in combination with a GSK3 inhibitor (to activate the phosphoinositide 3-kinase pathway) may be a more effective therapy for neurodegenerative diseases than using a single kinase inhibitor approach. The clinical effectiveness of imatinib, which is a multi-target kinase inhibitor, suggests that combining single-target kinase inhibitors could be a reasonable approach.

Rho-associated protein kinase 1 (ROCK1) and myosin light chain kinase (MLCK)

ROCK1 and MLCK are actomyosin-associated serine–threonine protein kinases that play a part in tissue barrier dysfunction by increasing myosin light chain phosphorylation and by mediating downstream cytoskeletal changes98,99. Both kinases are drug discovery targets in disorders that have microvascular dysfunction as their pathological basis100,101. Microvascular dysfunction is a co-morbid variable in many CNS disorders102.

The Rho–ROCK1 pathway has been implicated in numerous disorders, including stroke and asthma, and ROCK1 inhibitors have shown efficacy in animal models and in human disease (reviewed in REFS 8,10). ROCK1 is the target of the first kinase inhibitor drug to be approved, fasudil (AT-877; Asahi Kasei)103. The approved indication was for cerebral vasospasm, a microvascular disorder. Currently, fasudil is in clinical trials in the United States for Raynaud’s phenomenon (TABLE 1). At higher concentrations (in the micromolar range), fasudil can function as an MLCK inhibitor104.

Studies with knockout mice and with MLCK inhibitors in wild-type mice in several animal disease models have provided extensive preclinical target validation evidence for this kinase in disorders in which tissue barrier dysfunction is part of the pathology101,105,106. For example, studies106,107 using both an isoform-selective MLCK-knockout mouse and a selective inhibitor in wild-type mice showed that MLCK inhibition can provide protection against tissue injury. Diverse animal model studies, using either genetic or in vivo chemical biology approaches, showed that MLCK inhibition provided protection from disease-relevant stressors in diseases that have tissue barrier dysfunction as part of pathology progression. These include ventilator and inflammation-induced lung injury101,106,107, immune-mediated intestinal diarrhoea108, severe-burn-induced microvascular barrier injury105 and endotoxic shock109. The accumulating data that MLCK inhibitors can protect against tissue injury provide the target validation impetus to extend the investigation to CNS disorders that have tissue barrier dysfunction as a component of pathology, such as multiple sclerosis, stroke and traumatic brain injury.

In summary, both ROCK1 and MLCK are promising kinase targets for CNS disorders involving BBB dysfunction or microvascular pathologies. Like GSK3, they are examples of widely distributed protein kinases. The successful clinical use of fasudil provides a case study showing that ubiquitous kinases, like many widely distributed targets, can be therapeutically modulated if they are validated as targets for the disorder and appropriate dosing regimes are employed. In addition, distinct kinase isoforms of ROCK1 and MLCK seem to be crucial for drug action100,106, which is reminiscent of the p38 MAPK isoforms.

Selected protein kinase targets that lack in vivo pre-clinical proof of concept validation

Evidence relating to the neurobiology of disease has highlighted various CNS protein kinases as candidate therapeutic targets, but in vivo evidence of outcome modulation by small-molecule intervention has not yet been disclosed. Furthermore, the clinical evidence is often an analysis of post-mortem human tissue, which has its limitations. However, it is anticipated that current research will soon provide initial target validation evidence, adding to the rapidly growing body of protein kinase targets for CNS drug discovery. Several of these are worth mentioning specifically as they are examples of possible future targets in the disease areas discussed above. For example, leucine-rich repeat kinase 2 (LRRK2) can localize to Lewy bodies in human Parkinson’s disease brain tissue, and LRRK2 genetic variance is linked to familial autosomal-dominant, late-onset Parkinson’s disease and some forms of sporadic Parkinson’s disease110,111. The gene encoding dual-specificity tyrosine phosphorylation-regulated kinase 1A is located in the Down’s syndrome region of chromosome 21, and overexpression is thought to contribute to abnormal brain development and disease pathogenesis112. Members of the p21-activated kinase (PAK) family of kinases, especially PAK1, have altered expression in neurodegenerative diseases and are upregulated in gliomas113. Expression of the casein kinase 1 (CK1) family of serine–threonine protein kinases, such as CK1δ, is increased in Alzheimer’s disease brain samples, and these kinases can phosphorylate the Alzheimer’s disease -related protein presenilin114. Furthermore, CK1 is a priming kinase for GSK3β and is an upstream regulator of CDK5 (REF. 114) — two other protein kinases that are implicated in Alzheimer’s disease115.

Challenges in targeting kinases in CNS disorders

BBB penetrance

The greatest challenge facing any CNS-targeted drug discovery programme is effective penetration of the BBB. It is estimated116 that only ~2% of small-molecule drugs are able to effectively cross the BBB. The physicochemical properties of a drug considerably influence passive diffusion across biological membranes117-119 and the potential to serve as a substrate for the P-glycoprotein (PGP) efflux transporter120. Molecular weight, polar surface area (PSA) and lipophilicity (LogP) are key molecular properties that correlate with and may have an important role in influencing the BBB penetrance of a molecule. For example, an analysis (FIG. 4) of the mean LogP, molecular weight, and PSA values of CNS-penetrant small molecules (see Supplementary information S1,S3 (box, table)) compared with those of kinase inhibitor drugs (see Supplementary information S1,S4 (box, table)) or drugs approved and marketed for all disease indications121 reveals that kinase inhibitor drugs tend to have higher mean values for these parameters than CNS-penetrant compounds. PGP substrates also tend to have higher average molecular weight, PSA and LogP values than molecules which are not PGP substrates (see Supplementary information S3 (table)).

Figure 4. Key molecular-property trends for small molecules.

Figure 4

Comparisons of mean values of three key molecular properties, lipophilicity (LogP) (a), molecular weight (MW) (b) and polar surface areas (PSA) (c), are shown for CNS-penetrant small molecules (CNS; see Supplementary information S3 (table)), approved and marketed drugs for any indication (MKT; data from Vieth121) and kinase inhibitor drugs for all indications (KID; see Supplementary information S4 (table)). One-way ANOVA was used to evaluate significant differences to CNS mean values. Mean values for LogP (d), MW (e) and PSA (f) were also calculated for compounds in the CNS dataset (see Supplementary information S3 (table)) that are P-glycoprotein substrates (PGP+) compared with compounds that have been experimentally validated as non-P-glycoprotein substrates (PGP). Student’s t-test was used to compare mean values. *p < 0.05. **p < 0.001.

The trend for discrepancy between the molecular properties of CNS-penetrant compounds and those of current kinase inhibitor drugs raises concerns about the outcome of pending clinical trials and the prevailing approach of trying to use drugs that were developed and approved for peripheral-tissue diseases to treat CNS disorders. It is not known how the characteristics of small molecules affect the targeting of protein kinases in CNS disorders, but case studies provide some insights into emerging trends. For example, imatinib is an effective non-CNS cancer therapeutic122 that has now entered trials for glioma (TABLE 1), but such CNS clinical trials have generally failed123,124. This may be related to the observations that the molecular weight and PSA of imatinib (493.60 and 86.28 Å2, respectively) are greater than those of other CNS-penetrant compounds (FIG. 4a,b,c), and that imatinib is a PGP substrate125. Imatinib has a LogBB of −1.50, showing that it has poor brain uptake. LogBB values greater than −1 are considered to have reasonable brain penetration126. The follow-on drug dasatinib (Sprycel; Bristol–myers Squibb) (TABLE 1) has a greater molecular weight and PSA than imatinib, but does not seem to be a good substrate for the PGP efflux transporter127. Dasatinib has an incrementally improved LogBB of −1.30, which will hopefully be sufficient for CNS efficacy128. These borderline cases are difficult to evaluate, but the more extreme cases of higher molecular weight and PSA of many kinase inhibitor drugs (FIG. 4) suggest that future efforts at targeting protein kinases in CNS disorders might benefit from a primary focus on CNS disorders at the outset of the discovery programme.

Forecasting the potential uptake of small molecules by the brain in specific CNS drug discovery programmes remains an elusive goal. The consideration of several physicochemical characteristics of protein kinase-targeted drugs represents a promising strategy to achieve this. However, this is a context-dependent approach owing to the complex nature of in vivo absorption, tissue distribution, diffusion across the BBB and PGP substrate status. only tenuous conclusions can be drawn from a single medicinal chemistry refinement programme. For example, early studies of benzodiazepines129 found correlations among brain uptake, efficacy and blood levels of the drugs with high lipophilicity. However, the authors were careful to note the complicating factors of increased metabolism to active metabolites of some analogues and a potential bias in the data set towards some physical properties such as lipophilicity. As more results became available, it was clear that there was not a simple correlation between LogBB and LogP. For example, flunitrazepam, diazepam and midazolam have LogP values that increase linearly with respect to their corresponding LogBB values (see Supplementary information S3 (table)). However, the benzodiazepine oxazepam has a lower LogP value of 1.35 and a higher LogBB value of 0.61 than the three other benzodiazepine compounds. Similarly, the elevated LogP of alprazolam (3.13) does not give a corresponding increase in LogBB. The refinement of antihistamines provides another example. Diphenhydramine, a first-generation antihistamine, generates sedative effects, which are attributed to its ability to penetrate the BBB130. However, these effects are absent with the second-generation antihistamine fexofenadine (Allegra; Sanofi–Aventis). A possible explanation is that efflux of fexofenadine by PGP130,131 limits its brain uptake and sedative effects.

The trend analysis of molecular properties presented in FIG. 4 shows that CNS-penetrant compounds have much lower PSA values than other drug classes (for example, marketed drugs and kinase inhibitors). Certain molecular drug target classes, such as kinase inhibitors, are characterized by high mean PSA values132. This raises the question of whether the CNS kinase target class is not amenable to a lower PSA, or whether there has not been sufficient focus on such properties during medicinal chemistry refinement. There is the concern that decreasing key properties such as PSA to improve BBB penetrance could produce a better substrate for cytochrome P450 2D6 (CYP2D6)133. This is an undesired outcome as the polymorphic nature of this drug-metabolizing enzyme results in differences in metabolism among individuals. Like CNS drugs, CYP2D6 substrates are characterized by lower mean PSA values (35.25 Å2) than non-CYP2D6 substrates (57.49 Å2)133. This is an interesting correlation, as a higher proportion of CNS drugs are CYP2D6 substrates than non-CNS drugs133.

In summary, the complexity and challenge of penetrating the BBB make it difficult to apply existing drugs for peripheral-tissue disorders to the CNS, even when the drugs are safe and efficacious. Studying the molecular basis of CNS drug failures for well validated targets might allow additional BBB-related issues to be elucidated beyond physicochemical properties that affect drug uptake or PGP substrate status. The technology now available for protein kinase inhibitor design allows for the introduction of CNS-relevant considerations at the early ligand design stage concerning key physicochemical properties, especially lower molecular weight and PSA.

Cytochrome P450 (CYP) substrate status

The impact of CYP-mediated metabolism on drug safety and efficacy makes it a key consideration in drug discovery, irrespective of the target or tissue site of action. Together, CYP2D6 and CYP3A4 account for the metabolism of ~70% of marketed drugs134. However, the prevalence of CYP2D6-mediated metabolism of existing CNS drugs and the variable patient responses due to functional genetic polymorphisms in the CYP gene119,135 emphasize the importance of avoiding CYP2D6 metabolism in the discovery of new CNS therapeutics. The CYP2D6-mediated conversion of codeine to its active drug form, morphine, provides an example of unwanted, variable patient responses. Patients with a ‘slow metabolizer’ phenotype can experience reduced analgesic effects resulting from diminished morphine production, whereas a ‘rapid metabolizer’ phenotype increases toxicity risks as excessive levels of morphine can be produced136. Increasing LogP and minimizing PSA are sometimes used to improve brain uptake of small molecules, but this can also increase the likelihood that the small molecule will serve as a CYP2D6 substrate133. Therefore, medicinal chemistry optimization to improve brain uptake must be done carefully to minimize the probability of creating good CYP2D6 substrates. A promising trend is that early testing and avoiding the generation of good CYP2D6 substrates have become standard practices in medicinal chemistry refinement.

CYP3A4-mediated metabolism remains a pervasive issue in kinase inhibitor drug discovery. For example, nearly half of the kinase inhibitor drugs in TABLE 1 and a high percentage of all kinase inhibitor drugs, regardless of disease area, are CYP3A4 substrates (see Supplementary information S4 (table)). One example is imatinib137, which is often used clinically in conjunction with anti-convulsant therapy123 — a therapy that can increase CYP3A4 activity. As a result, the pharmacokinetics of imatinib can be altered. Although it is difficult to alter CYP substrate status by changing a single molecular property, CYP3A4 substrates have a higher average LogP than other CYP substrates, providing another reason to minimize lipophilicity and molecular weight during medicinal chemistry refinement. Although treatment of patients can be guided by genotyping or the adjustment of dosing for each drug, a preferable strategy is to reduce variability and safety issues at the medicinal chemistry refinement stage by a consideration of the features that determine whether a drug will be a good substrate for CYP2D6 or CYP3A4.

The kinase ATP site — selectivity and affinity

The kinase catalytic core structure (FIG. 5) has a bi-lobe architecture comprising a smaller N-terminal domain consisting mainly of β-sheets and a larger C-terminal domain consisting mainly of α-helices. Between the two domains is the ATP-binding fold and the linker hinge region that connects the two domains. The adenine ring of ATP interacts with the hinge region, and the ribose and triphosphate groups of ATP bind in a channel that extends to the peptide-substrate-binding region. The ATP fold contains a set of key conserved regions that are essential for substrate recognition and catalysis. For example, an activation loop containing a conserved DFG motif is important in regulating kinase activity, and a P-loop containing a glycine-rich motif forms the ‘roof’ of the ATP-binding site. Although the canonical protein kinase ATP-binding site contains many conserved regions and features, there are also distinctive pockets and residues that differentiate the structural landscapes of the various kinase ATP folds. Consequently, a mixture of conserved and unique features makes targeting the protein kinase ATP fold an attractive strategy for designing selective kinase inhibitors.

Figure 5. Structure of the protein kinase catalytic core and bound inhibitor in the ATP fold.

Figure 5

Protein kinases are characterized by a bi-lobe architecture consisting of an amino-terminal β-sheet lobe and a larger α-helical carboxy-terminal domain, with a canonical fold that contains the nucleotide-binding site located between the two domains. Other conserved protein kinase features include the P-loop (shown in magenta), which contains a glycine-rich motif (GXGXXG), and the activation loop (shown in purple), which is characterized by an initial DFG and terminal APE motif. Upon ATP binding, a reorientation of the C-helix (shown in light blue) and the glycine-rich loop is often observed. This structure shows SYK tyrosine kinase in complex with the therapeutic kinase inhibitor imatinib (Protein Data Bank identification number: 1XBB).

Most kinase inhibitor drugs and drug candidates occupy the ATP fold, but gain selectivity by exploiting space and contacts with the enzyme that are not used by ATP. In general, inhibitor selectivity is achieved through sets of contacts and binding modes that are unique for individual kinases or small groups of kinases2. Crystal structures of kinases in complex with an inhibitor (see FIG. 5 for the crystal structure of imatinib in complex with SYK tyrosine kinase) show that the compounds can occupy a similar three-dimensional space as ATP, but interact with distinct amino-acid residues. Kinase inhibitors can bind to conserved enzyme structural features that are used by ATP, and to nearby three-dimensional topologies that are unique or restricted to a subclass of kinases, such as a hydrophobic pocket with restricted accessibility. This steric array of binding modes allows a selective and high-affinity interaction of the small molecule with only a small subset of an organism’s proteome. One complementary inhibitor design strategy involves targeting regions outside the ATP-binding site using non-competitive or allosteric inhibitors2. For example, non-competitive inhibitors have been identified for MEK1 and MEK2 (REFS 138,139) that exploit a region adjacent to the canonical ATP-binding site and show high target selectivity. Although the development of allosteric inhibitors has so far not had the clinical success of ATP-competitive inhibitors, this may be due in part to the lack of knowledge about small-molecule allosteric pockets, such as that of MEK1 and MEK2, in other kinases.

A primary drug discovery goal is to design an inhibitor with sufficient affinity to allow competition with the two kinase substrates, ATP and the substrate protein, the apparent affinities of which are well above 1 μM. The aim is to inhibit the phosphorylation reaction, the rate-limiting step of which is the release of the products ADP and phosphorylated protein140. Kinase targets are usually components of complex, interconnected signal transduction cascades comprising many protein kinases, with pathway redundancy and crosstalk between pathways (reviewed in REF. 91). The presence of such complex intracellular networks raises the possibility that broad specificity of a kinase inhibitor for a cluster of kinase targets, rather than high affinity for a single target, may provide the desired pharmacological effectiveness in vivo. Imatinib is an example of this paradigm: its apparent inhibition constant for the target kinases is just below the operational goal of less than 1 μM and its pharmacological targets are a discrete set of protein kinases122. The key challenge to the development of protein kinase inhibitors for CNS disorders is the design of inhibitors that exhibit sufficient affinity and selectivity for a molecular target, or set of molecular targets, and have the appropriate molecular property profiles for CNS penetrance.

Conclusions and trends

The targeting of protein kinases in CNS disorders is a promising field at an early stage of development. Although there are challenging barriers to overcome, these have the potential to be addressed with current approaches. Analysis of current protein kinase inhibitor drugs and CNS-penetrant small molecules reveals addressable challenges related to the bioavailability and pharmacokinetics of kinase inhibitor drug candidates. It is crucial to consider the physicochemical properties of small molecules that influence the interplay of pharmacokinetics and pharmacodynamics at an early stage in CNS drug discovery. Current technologies increase the potential for success by allowing the generation of compounds that share molecular and pharmacological properties with compounds that were previously successful, yet are unique chemical structures.

Greater emphasis is increasingly being placed on finding approaches that allow early assessment of a candidate drug’s potential for long-term success. This is based on the realization that, across several molecular-target classes, the percentage change in compound properties made during medicinal chemistry refinement of initial hits to leads is substantially higher than during refinement of lead compounds to clinical candidates. The trend indicates the impact of the starting hit on the drug discovery and development process. Such challenges are now addressed earlier in CNS drug discovery programmes than they were previously. The strategy represents a major challenge in early-stage screening and substantially reduces the medicinal chemistry effort, especially in terms of the refinement of less desirable hits. The new approaches also offer the potential to improve the integration of ligand discovery with drug discovery. For example, activity-based high-throughput screening (HTS) has been a primary source of new hits for drug discovery. However, as noted by Lipinski117, hits and leads that are derived from such programmes tend to be far from optimal in their drug-like physicochemical properties. The outcome requires that considerable effort be expended on detailed, quantitative structure–activity relationship analyses on hits from HTS activity screens to define the minimal structure that is required for activity. The analyses are needed at this stage owing to the usual changes in molecular properties, such as increased molecular weight and lipophilicity, that occur as a drug development programme progresses.

The significant difference in mean values for physicochemical properties of CNS-penetrant small molecules compared with those for current protein kinase inhibitor drugs indicates an even greater divide between hits and CNS drug candidates than for other molecular targets and disease areas. Based on analyses of the molecular properties of the CNS-penetrant small-molecules database (see Supplementary information S3 (table)) and kinase inhibitor drugs and drug candidates (see Supplementary information S2,S3,S4 (tables)), it seems that kinase inhibitor drugs for CNS indications will require a more restricted profile of molecular properties than the Lipinski ‘Rule of Five’ — a widely used filter to prioritize medicinal chemistry refinement efforts117. Most of the brain-penetrant small molecules (see Supplementary information S3 (table)) have a molecular weight < 400, LogP < 4 and PSA < 80Å2. Compounds outside this cluster of values have a higher likelihood of having undesired CYP metabolism and PGP efflux and a lower mean LogBB. Therefore, the likelihood of success of a CNS-focused kinase inhibitor discovery programme might be increased by a careful approach to the screening and selection of hit compounds, and by the monitoring of key molecular properties during recursive medicinal chemistry refinement.

One emerging and promising approach with the potential to address the challenges in targeting protein kinases for CNS disorders is a form of fragment-based drug discovery (BOX 3), referred to as fragment expansion. Recent reviews highlight the success in developing fragments into clinical candidates (for example, see REFS 141-143). Fragment expansion approaches tend to yield fewer but higher quality compounds than other drug discovery approaches, and result in less time being spent on non-essential structure–activity studies. This ‘smart chemistry’ approach also allows in vivo proof of concept testing using animal models of preclinical safety, pharmacology and efficacy at an early stage of the drug discovery timeline.

Box 3 | Fragment-based discovery strategies for CNS kinase inhibitors.

Molecular-fragment-based approaches start with weakly active or inactive chemical scaffolds that are amenable to chemical diversification at multiple points141. Structure-assisted, fragment-based drug discovery uses small, usually weakly binding, molecular fragments or scaffolds as a starting point for finding novel hits and lead compounds. Studies usually begin with a high-resolution X-ray crystal structure of the kinase domain of a potential central nervous system (CNS) target containing an inactive molecular fragment bound to the active site of the kinase. The assumption is that the fragment makes specific interactions with the protein that are typically retained through the process of fragment expansion and optimization.

Part a of the figure shows a typical protein kinase ATP fold containing a bound ATP analogue, and examples of fragments or scaffolds that are used in the discovery of kinase inhibitors. The ATP fold is the region of protein kinases that has been successfully targeted to generate drugs and candidate compounds. Part b shows two widely used fragment-based discovery strategies: fragment expansion and fragment linkage. The fragment expansion variation of the approach uses chemical diversification of the core fragment to incrementally grow the starting fragment in multiple directions, thereby sampling regions of ‘chemical space’ that could result in favourable interactions with the protein. The assumption is that the custom set of interactions between a minimally expanded ligand and its target protein will optimize activity and selectivity. Linking two fragments that bind in different locations is an alternative strategy. Both approaches are compatible with inclusion of a selection filter that is based on small-molecule physical property considerations. Computed molecular properties of the proposed products are examined to select candidates for synthesis that are similar to known CNS-penetrant compounds. The aim is to increase the probability of producing kinase inhibitors that are more CNS drug-like than existing compounds.

The consideration of CNS drug characteristics for designed compounds complements the inhibitor design method that maximizes target interactions. The combination thereby reduces the total number of compounds to be made initially in the discovery phase. Once an initial set of compounds is synthesized based on the drug-like ligand design and hits are obtained, selected pharmacology-driven in vitro screens can be used as an initial experimental filter. The approach provides a medicinal chemistry refinement strategy that is driven by a recursive design–synthesis–evaluation–design cycle.

graphic file with name nihms-174179-f0001.jpg

A compound discovery and refinement paradigm that allows early testing in vivo is attractive for targeting protein kinases in CNS disorders owing to the inherent molecular and tissue-related challenges. The complexity of protein kinase-mediated signal transduction cascades and the limited knowledge about these highly integrated intracellular pathways necessitate in vivo testing of hypotheses early in the discovery process. Early testing of the potential for a given protein kinase to be druggable requires in vivo experimental probing with bioavailable compounds. The initial design of molecules with the appropriate characteristics and the use of pharmacological filters as an adjunct to target activity screens increase the probability that the fewer compounds emerging from the early stages will be more drug-like and stable and will have BBB penetrance (see FIG. 6 for an example flowsheet). Implementing procedures to reduce the risk of compound failure at an early stage will result in a ‘failing fast in vivo’ paradigm, but the recursive nature of the process will allow lessons to be learned from well designed failures, as well as from successes, which should rapidly enhance the process as the project progresses. The net result is that fewer compounds are synthesized and less time is required to produce clinical candidates.

Figure 6. Example of a potential compound discovery paradigm for CNS kinase targets.

Figure 6

Focusing on small-molecule kinase inhibitors with key molecular property characteristics and promising profiles in early-stage pharmacological screens increases the potential for positive outcomes in the subsequent preclinical development of investigational new drugs. This substantially reduces early discovery demands on synthetic chemistry and biology efforts and the later attrition rate in development. Medicinal chemistry refinement takes into consideration molecular properties, suitability for central nervous system (CNS) penetration and ligand activity. Pharmacology-related screens are used to measure brain uptake, assess microsomal stability, determine whether the compound is a cytochrome P450 2D6 (CYP2D6) or CYP3A4 substrate and ensure that there are no adverse effects. HTS, high-throughput screening; IC50, half-maximal inhibitory concentration; SAR, structure–activity relationship.

In summary, protein kinases hold tremendous promise as therapeutic targets for CNS disorders, with an increasing number of protein kinases emerging as viable targets for major neurological diseases that lack effective therapies. The development of kinase-targeted therapeutics for CNS diseases must address the usual key drug discovery hurdles, such as first-pass metabolism, target affinity and selectivity, and therapeutic index. Although kinase inhibitor drug discovery in general has successfully overcome many of the challenges and progressed considerably over the past decade, there are special challenges related to CNS drug discovery that must be addressed at the start of the drug discovery process. Consideration of important CNS challenges in molecular design and the early-stage integration of key feasibility screens should enhance the likelihood of success in developing new kinase-targeted therapeutics for CNS disorders.

Supplementary Material

S1
s2
s3
s4

Acknowledgements

The authors acknowledge funding from the National Institutes of Health (AG031311, NS056051 and AG013939.

Glossary

Blood–brain barrier

The barrier formed by tight junctions of brain endothelial cells that restrict the passage of drugs between the bloodstream and the brain.

Pharmacodynamics

The biochemical and physiological effects of an administered drug on the body, and the relationships between drug dose and efficacy or toxicity.

Kinome

The subset of genes that encode protein kinases in the genome of an organism.

P-glycoprotein

An efflux transporter at the blood–brain barrier that can prevent drug substrate accumulation in the brain.

Polar surface area

The total surface area of all polar atoms (usually oxygen and nitrogen), including attached hydrogen ions.

LogP

A measure of a drug’s lipophilicity, it is the logarithm of the octanol–water partition coefficient. Higher values are associated with greater lipophilicity, whereas lower values are characteristic of more water-soluble compounds.

LogBB

The steady-state ratio of drug concentration in the brain versus drug concentration in the blood.

Pharmacokinetics

‘What the body does to the drug’, including the extent and rate of drug absorption, distribution, metabolism and excretion.

References

  • 1.Catapano LA, Manji HK. Kinases as drug targets in the treatment of bipolar disorder. Drug Disc. Today. 2008;13:295–302. doi: 10.1016/j.drudis.2008.02.007. [DOI] [PubMed] [Google Scholar]
  • 2.Zhang J, Yang PL, Gray NS. Targeting cancer with small molecule kinase inhibitors. Nature Rev. Cancer. 2009;9:28–39. doi: 10.1038/nrc2559.A recent detailed review of the structural basis of target recognition and molecular mechanism of protein kinase inhibitors.
  • 3.Borders AS, de Almeida L, Van Eldik LJ, Watterson DM. The p38α mitogen-activated protein kinase as a central nervous system drug discovery target. BMC Neurosci. 2008;9:S12. doi: 10.1186/1471-2202-9-S2-S12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Martinez A, Perez DI. GSK-3 inhibitors: a ray of hope for the treatment of Alzheimer’s disease? J. Alzheimer. Dis. 2008;15:181–191. doi: 10.3233/jad-2008-15204. [DOI] [PubMed] [Google Scholar]
  • 5.Jope RS, Yuskaitis CJ, Beurel E. Glycogen synthase kinase-3 (GSK3): inflammation, diseases, and therapeutics. Neurochem. Res. 2007;32:577–595. doi: 10.1007/s11064-006-9128-5.A review of the link between GSK3β and pathology progression in CNS diseases
  • 6.Forde JE, Dale TC. Glycogen synthase kinase 3: a key regulator of cellular fate. Cell. Mol. Life Sci. 2007;64:1930–1944. doi: 10.1007/s00018-007-7045-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Kockeritz L, Doble B, Patel S, Woodgett JR. Glycogen synthase kinase-3 — an overview of an over-achieving protein kinase. Curr. Drug Targets. 2006;7:1377–1388. doi: 10.2174/1389450110607011377. [DOI] [PubMed] [Google Scholar]
  • 8.Kubo T, Yamaguchi A, Iwata N, Yamashita T. The therapeutic effects of Rho-ROCK inhibitors on CNS disorders. Ther. Clin. Risk Manag. 2008;4:605–615. doi: 10.2147/tcrm.s2907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Martinez A. Preclinical efficacy on GSK-3 inhibitors: towards a future generation of powerful drugs. Med. Res. Rev. 2008;28:773–796. doi: 10.1002/med.20119.A recent review documenting the efficacy of various classes of GSK3 inhibitors in preclinical models of CNS disorders.
  • 10.Mueller BK, Mack H, Teusch N. Rho kinase, a promising drug target for neurological disorders. Nature Rev. Drug Discov. 2005;4:387–398. doi: 10.1038/nrd1719.A review of Rho kinase as a CNS therapeutic target.
  • 11.Schumacher AM, Velentza AV, Watterson DM. Death associated protein kinase as a potential therapeutic target. Exp. Opin. Ther. Targets. 2002;6:497–506. doi: 10.1517/14728222.6.4.497. [DOI] [PubMed] [Google Scholar]
  • 12.Hooper C, Killick R, Lovestone S. The GSK3 hypothesis of Alzheimer’s disease. J. Neurochem. 2008;104:1433–1439. doi: 10.1111/j.1471-4159.2007.05194.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Jope RS, Roh M-S. Glycogen synthase kinase-3 (GSK3) in psychiatric diseases and therapeutic interventions. Curr. Drug Targets. 2006;7:1421–1434. doi: 10.2174/1389450110607011421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Lovestone S, Killick R, Di Forti M, Murray R. Schizophrenia as a GSK-3 dysregulation disorder. TRENDS Neurosci. 2007;30:142–149. doi: 10.1016/j.tins.2007.02.002. [DOI] [PubMed] [Google Scholar]
  • 15.Hye A, et al. Glycogen synthase kinase-3 is increased in white cells early in Alzheimer’s disease. Neurosci. Lett. 2005;373:1–4. doi: 10.1016/j.neulet.2004.10.031. [DOI] [PubMed] [Google Scholar]
  • 16.Kwok JBJ, et al. Glycogen synthase kinase-3β and tau genes interact in Alzheimer’s disease. Ann. Neurol. 2008;64:446–454. doi: 10.1002/ana.21476. [DOI] [PubMed] [Google Scholar]
  • 17.Parihar MS, Hemnani T. Alzheimer’s disease pathogenesis and therapeutic interventions. J. Clin. Neurosci. 2004;11:456–467. doi: 10.1016/j.jocn.2003.12.007. [DOI] [PubMed] [Google Scholar]
  • 18.Huang H-C, Klein PS. Multiple roles for glycogen synthase kinase-3 as a drug target in Alzheimer’s disease. Curr. Drug Targets. 2006;7:1389–1397. doi: 10.2174/1389450110607011389. [DOI] [PubMed] [Google Scholar]
  • 19.Klein PS, Melton DA. A molecular mechanism for the effect of lithium on development. Proc. Natl Acad. Sci. USA. 1996;93:8455–8459. doi: 10.1073/pnas.93.16.8455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Macdonald A, et al. A feasibility and tolerability study of lithium in Alzheimer’s disease. Intl J. Geriatr. Psychiatry. 2008;23:704–711. doi: 10.1002/gps.1964. [DOI] [PubMed] [Google Scholar]
  • 21.Jope RS. Lithium and GSK-3: one inhibitor, two inhibitory actions, multiple outcomes. Trends Pharmacol. Sci. 2003;24:441–443. doi: 10.1016/S0165-6147(03)00206-2. [DOI] [PubMed] [Google Scholar]
  • 22.Zhang F, Phiel CJ, Spece L, Gurvich N, Klein PS. Inhibitory phosphorylation of glycogen synthase kinase-3 (GSK-3) in response to lithium. Evidence for autoregulation of GSK-3. J. Biol. Chem. 2003;278:33067–33077. doi: 10.1074/jbc.M212635200. [DOI] [PubMed] [Google Scholar]
  • 23.Kozikowski AP, et al. Structure-based design leads to the identification of lithium mimetics that block mania-like effects in rodents. Possible new GSK-3β therapies for bipolar disorders. J. Am. Chem. Soc. 2007;129:8328–8332. doi: 10.1021/ja068969w. [DOI] [PubMed] [Google Scholar]
  • 24.Kozikowski AP, et al. Highly potent and specific GSK-3β inhibitors that block tau phosphorylation and decrease α-synuclein protein expression in a cellular model of Parkinson’s disease. Chem. Med. Chem. 2006;1:256–266. doi: 10.1002/cmdc.200500039. [DOI] [PubMed] [Google Scholar]
  • 25.Wang W, et al. Inhibition of glycogen synthase kinase-3β protects dopaminergic neurons from MPTP toxicity. Neuropharmacol. 2007;52:1678–1684. doi: 10.1016/j.neuropharm.2007.03.017. [DOI] [PubMed] [Google Scholar]
  • 26.Koh S, et al. Inhibition of glycogen synthase kinase-3 suppresses the onset of symptoms and disease progression of G93A-SOD1 mouse model of ALS. Exp. Neurol. 2007;205:336–346. doi: 10.1016/j.expneurol.2007.03.004. [DOI] [PubMed] [Google Scholar]
  • 27.Gould TD, Einat H, Bhat R, Manji HK. AR-A014418, a selective GSK-3 inhibitor, produces antidepressant-like effects in the forced swim test. Int. J. Neuropsychopharmacol. 2004;7:387–390. doi: 10.1017/S1461145704004535. [DOI] [PubMed] [Google Scholar]
  • 28.Koh SH, Yoo AR, Chang DI, Hwang SJ, Kim SH. Inhibition of GSK-3 reduces infarct volume and improves neurobehavioral functions. Biochem. Biophys. Res. Commun. 2008;371:894–899. doi: 10.1016/j.bbrc.2008.05.006. [DOI] [PubMed] [Google Scholar]
  • 29.Martinez A, Alonso M, Castro A, Prez C, Moreno FJ. First non-ATP competitive glycogen synthase kinase 3β (GSK-3β) inhibitors: thiadiazolidinones (TDZD) as potential drugs for the treatment of Alzheimer’s disease. J. Med. Chem. 2002;45:1292–1299. doi: 10.1021/jm011020u. [DOI] [PubMed] [Google Scholar]
  • 30.Collino M, et al. Treatment with the glycogen synthase kinase-3β inhibitor, TDZD-8, affects transient cerebral ischemia/reperfusion injury in the rat hippocampus. Shock. 2008;30:299–307. doi: 10.1097/SHK.0b013e318164e762. [DOI] [PubMed] [Google Scholar]
  • 31.Cuzzocrea S, et al. Glycogen synthase kinase-3β nhibition reduces secondary damage in experimental spinal cord trauma. J. Pharmacol. Exp. Ther. 2006;318:79–89. doi: 10.1124/jpet.106.102863. [DOI] [PubMed] [Google Scholar]
  • 32.Rosa AO, et al. Antidepressant-like effect of the novel thiadiazolidinone NP031115 in mice. Prog. Neuro. Psychopharmacol. Biol. Psych. 2008;32:1549–1556. doi: 10.1016/j.pnpbp.2008.05.020. [DOI] [PubMed] [Google Scholar]
  • 33.Luna-Medina R, et al. NP031112, a thiadiazolidinone compound, prevents inflammation and neurodegeneration under excitotoxic conditions: potential therapeutic role in brain disorders. J. Neurosci. 2007;27:5766–5776. doi: 10.1523/JNEUROSCI.1004-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Sereno L, et al. A novel GSK-3β inhibitor reduces Alzheimer’s pathology and rescues neuronal loss in vivo. Neurobiol. Dis. 2009;35:359–367. doi: 10.1016/j.nbd.2009.05.025. [DOI] [PubMed] [Google Scholar]
  • 35.Medina M, Castro A. Glycogen synthase kinase-3 (GSK-3) inhibitors reach the clinic. Curr. Opin. Drug Discov. Devel. 2008;11:533–543.A recent review of the drug development status of GSK3 inhibitors.
  • 36.Gould TD, Gray NA, Manji HK. Effects of a glycogen synthase kinase-3 inhibitor, lithium, in adenomatous polyposis coli mutant mice. Pharmacol. Res. 2003;48:49–53. [PubMed] [Google Scholar]
  • 37.Cohen Y, Chetrit A, Cohen Y, Sirota P, Modan B. Cancer morbidity in psychiatric patients: influence of lithium carbonate treatment. Med. Oncol. 1998;15:32–36. doi: 10.1007/BF02787342. [DOI] [PubMed] [Google Scholar]
  • 38.O’Brien WT, et al. Glycogen synthase kinase-3β haploinsufficiency mimics the behavioral and molecular effects of lithium. J. Neurosci. 2004;24:6791–6798. doi: 10.1523/JNEUROSCI.4753-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Bialik S, Kimchi A. The death-associated protein kinases: structure, function, and beyond. Ann. Rev. Biochem. 2006;75:189–210. doi: 10.1146/annurev.biochem.75.103004.142615. [DOI] [PubMed] [Google Scholar]
  • 40.Schumacher AM, Velentza AV, Watterson DM, Wainwright MS. DAPK catalytic activity in the hippocampus increases during the recovery phase in an animal model of brain hypoxic-ischemic injury. Biochim. Biophys. Acta. 2002;1600:128–137. doi: 10.1016/s1570-9639(02)00453-3. [DOI] [PubMed] [Google Scholar]
  • 41.Velentza AV, et al. An aminopyridazine-based inhibitor of a pro-apoptotic protein kinase attenuates hypoxia-ischemia induced brain injury. Bioorg. Med. Chem. Lett. 2003;13:3465–3470. doi: 10.1016/s0960-894x(03)00733-9.This study provides an example of kinase co-crystallography with an inactive fragment as the starting point for the design of bioavailable kinase inhibitors, and the first in vivo demonstration of DAPK as a discovery target for CNS disorders.
  • 42.Craft JM, Watterson DM, Frautschy SA, Van Eldik LJ. Aminopyridazines inhibit β-amyloid-induced glial activation and neuronal damage in vivo. Neurobiol. Aging. 2004;25:1283–1292. doi: 10.1016/j.neurobiolaging.2004.01.006. [DOI] [PubMed] [Google Scholar]
  • 43.Shamloo M, et al. Death-associated protein kinase is activated by dephosphorylation in response to cerebral ischemia. J. Biol. Chem. 2005;280:42290–42299. doi: 10.1074/jbc.M505804200. [DOI] [PubMed] [Google Scholar]
  • 44.Li Y, et al. DAPK1 variants are associated with Alzheimer’s disease and allele-specific expression. Hum. Mol. Genet. 2006;15:2560–2568. doi: 10.1093/hmg/ddl178. [DOI] [PubMed] [Google Scholar]
  • 45.Schumacher AM, Schavocky JP, Velentza AV, Mirzoeva S, Watterson DM. A calmodulin-regulated protein kinase linked to neuron survival is a substrate for the calmodulin-regulated death-associated protein kinase. Biochemistry. 2004;43:8116–8124. doi: 10.1021/bi049589v. [DOI] [PubMed] [Google Scholar]
  • 46.Schumacher AM, Velentza AV, Watterson DM, Dresios J. Death-associated protein kinase phosphorylates mammalian ribosomal protein S6 and reduces protein synthesis. Biochemistry. 2006;45:13614–13621. doi: 10.1021/bi060413y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Shani G, et al. Death-associated protein kinase phosphorylates ZIP kinase forming a unique kinase hierarchy to activate its cell death functions. Mol. Cell Biol. 2004;24:8611–8626. doi: 10.1128/MCB.24.19.8611-8626.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Eisenberg-Lerner A, Kimchi A. DAP kinase regulates JNK signaling by binding and activating protein kinase D under oxidative stress. Cell Death Diff. 2007;14:1908–1915. doi: 10.1038/sj.cdd.4402212. [DOI] [PubMed] [Google Scholar]
  • 49.Tian JH, Das S, Sheng ZH. Ca2+-dependent phosphorylation of syntaxin-1A by the death-associated protein (DAP) kinase regulates its interaction with Munc18. J. Biol. Chem. 2003;278:26265–26274. doi: 10.1074/jbc.M300492200. [DOI] [PubMed] [Google Scholar]
  • 50.Bialik S, Bresnick AR, Kimchi A. DAP-kinase-mediated morphological changes are localization dependent and involve myosin-II phosphorylation. Cell Death Diff. 2004;11:631–644. doi: 10.1038/sj.cdd.4401386. [DOI] [PubMed] [Google Scholar]
  • 51.Zalckvar E, et al. DAP-kinase-mediated phosphorylation on the BH3 domain of beclin 1 promotes dissociation of beclin 1 from Bcl-XL and induction of autophagy. EMBO Reports. 2009;10:285–292. doi: 10.1038/embor.2008.246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Chen CH, et al. Bidirectional signals transduced by DAPK-ERK interaction promote the apoptotic effect of DAPK. EMBO J. 2005;24:294–304. doi: 10.1038/sj.emboj.7600510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Kyriakis JM, Avruch J. Mammalian mitogen-activated protein kinase signal transduction pathways activated by stress and inflammation. Physiol. Rev. 2001;81:808–869. doi: 10.1152/physrev.2001.81.2.807. [DOI] [PubMed] [Google Scholar]
  • 54.Shaul YD, Seger R. The MEK/ERK cascade: from signaling specificity to diverse functions. Biochim. Biophys. Acta. 2006;1773:1213–1226. doi: 10.1016/j.bbamcr.2006.10.005. [DOI] [PubMed] [Google Scholar]
  • 55.Cuenda A, Rousseau S. p38 MAP-kinases pathway regulation, function and role in human diseases. Biochim. Biophys. Acta. 2007;1773:1358–1375. doi: 10.1016/j.bbamcr.2007.03.010. [DOI] [PubMed] [Google Scholar]
  • 56.Krishna M, Narang H. The complexity of mitogen-activated protein kinases (MAPKs) made simple. Cell Mol. Life Sci. 2008;65:3525–3544. doi: 10.1007/s00018-008-8170-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Schindler JF, Monahan JB, Smith WG. p38 pathway kinases as anti-inflammatory drug targets. J. Dent. Res. 2007;86:800–811. doi: 10.1177/154405910708600902. [DOI] [PubMed] [Google Scholar]
  • 58.Munoz L, et al. A novel p38α MAPK inhibitor suppresses brain proinflammatory cytokine up-regulation and attenuates synaptic dysfunction and behavioral deficits in an Alzheimer’s disease mouse model. J. Neuroinflammation. 2007;4:21. doi: 10.1186/1742-2094-4-21.A proof of concept study showing that selective kinase inhibitors can attenuate brain injury if designed to be bioavailable and BBB-penetrant.
  • 59.Dewil M, de la Cruz VF, Van Den Bosch L, Robberecht W. Inhibition of p38 mitogen activated protein kinase activation and mutant SOD1G93A-induced motor neuron death. Neurobiol. Dis. 2007;26:332–341. doi: 10.1016/j.nbd.2006.12.023. [DOI] [PubMed] [Google Scholar]
  • 60.Piao CS, Kim JB, Han PL, Lee JK. Administration of the p38 MAPK inhibitor SB203580 affords brain protection with a wide therapeutic window against focal ischemic insult. J. Neurosci. Res. 2003;73:537–544. doi: 10.1002/jnr.10671. [DOI] [PubMed] [Google Scholar]
  • 61.Barone FC, et al. SB 239063, a second-generation p38 mitogen-activated protein kinase inhibitor, reduces brain injury and neurological deficits in cerebral focal ischemia. J. Pharmacol. Exp. Ther. 2001;296:312–321. [PubMed] [Google Scholar]
  • 62.Legos JJ, et al. SB 239063, a novel p38 inhibitor, attenuates early neuronal injury following ischemia. Brain Res. 2001;892:70–77. doi: 10.1016/s0006-8993(00)03228-5. [DOI] [PubMed] [Google Scholar]
  • 63.Ito T, et al. Glial phosphorylated p38 MAP kinase mediates pain in a rat model of lumbar disc herniation and induces motor dysfunction in a rat model of lumbar spinal canal stenosis. Spine. 2007;32:159–167. doi: 10.1097/01.brs.0000251437.10545.e9. [DOI] [PubMed] [Google Scholar]
  • 64.O’Keefe SJ, et al. Chemical genetics define the roles of p38α and p38β in acute and chronic inflammation. J. Biol. Chem. 2007;282:34553–34671. doi: 10.1074/jbc.M704236200.A precedent for in vivo target validation of a small-molecule kinase inhibitor using mice strains engineered to be drug resistant but to retain physiological functions.
  • 65.Hensley K, et al. p38 kinase is activated in the Alzheimer’s disease brain. J. Neurochem. 1999;72:2053–2058. doi: 10.1046/j.1471-4159.1999.0722053.x. [DOI] [PubMed] [Google Scholar]
  • 66.Sheng JG, et al. Interleukin-1 promotion of MAPK-p38 overexpression in experimental animals and in Alzheimer’s disease: potential significance for tau protein phosphorylation. Neurochem. Intl. 2001;39:341–348. doi: 10.1016/s0197-0186(01)00041-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Sun A, Liu M, Nguyen XV, Bing G. p38 MAP kinase is activated at early stages in Alzheimer’s disease brain. Exp. Neurol. 2003;183:394–405. doi: 10.1016/s0014-4886(03)00180-8. [DOI] [PubMed] [Google Scholar]
  • 68.Swatton JE, et al. Increased MAP kinase activity in Alzheimer’s and Down syndrome but not in schizophrenia human brain. Eur. J. Neurosci. 2004;19:2711–2719. doi: 10.1111/j.0953-816X.2004.03365.x. [DOI] [PubMed] [Google Scholar]
  • 69.Ferrer I, et al. Active, phosphorylation-dependent mitogen-activated protein kinase (MAPK/ERK), stress-activated protein kinase/c-Jun N-terminal kinase (SAPK/JNK), and p38 kinase expression in Parkinson’s disease and dementia with Lewy bodies. J. Neural Transm. 2001;108:1383–1396. doi: 10.1007/s007020100015. [DOI] [PubMed] [Google Scholar]
  • 70.Karunakaran S, et al. Selective activation of p38 mitogen-activated protein kinase in dopaminergic neurons of substantia nigra leads to nuclear translocation of p53 in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated mice. J. Neurosci. 2008;28:12500–12509. doi: 10.1523/JNEUROSCI.4511-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Atzori C, et al. Activation of the JNK/p38 pathway occurs in diseases characterized by tau protein pathology and is related to tau phosphorylation but not to apoptosis. J. Neuropath. Exp. Neurol. 2001;60:1190–1197. doi: 10.1093/jnen/60.12.1190. [DOI] [PubMed] [Google Scholar]
  • 72.Ferrer I, Blanco R, Carmona M, Puig B. Phosphorylated mitogen-activated protein kinase (MAPK/ERK-P), protein kinase of p38kDa (p38-P), stress-activated protein kinase (SAPK/JNK-P), and calcium/calmodulin-dependent kinase II (CaM kinase II) are differentially expressed in tau deposits in neurons and glial cells in tauopathies. J. Neural Transm. 2001;108:1397–1415. doi: 10.1007/s007020100016. [DOI] [PubMed] [Google Scholar]
  • 73.Demuth T, et al. MAP-ing glioma invasion: mitogen-activated protein kinase kinase 3 and p38 drive glioma invasion and progression and predict patient survival. Mol. Cancer Ther. 2007;6:1212–1222. doi: 10.1158/1535-7163.MCT-06-0711. [DOI] [PubMed] [Google Scholar]
  • 74.Lloyd E, Somera-Molina KC, Van Eldik LJ, Watterson DM, Wainwright MS. Suppression of acute proinflammatory cytokine and chemokine upregulation by post-injury administration of a novel small molecule improves long-term neurology outcome in a mouse model of traumatic brain injury. J. Neuroinflammation. 2008;5:28. doi: 10.1186/1742-2094-5-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Dominguez C, Powers D, Tamayo N. p38 MAP kinase inhibitors: many are made, but few are chosen. Curr. Opin. Drug Discov. Devel. 2005;8:421–430. [PubMed] [Google Scholar]
  • 76.McInnes C, Fischer PM. Strategies for the design of potent and selective kinase inhibitors. Curr. Pharmaceut. Design. 2005;11:1845–1863. doi: 10.2174/1381612053764850. [DOI] [PubMed] [Google Scholar]
  • 77.Wrobleski ST, Doweyko AM. Structural comparison of p38 inhibitor-protein complexes: a review of recent p38 inhibitors having unique binding interactions. Curr. Topics Med. Chem. 2005;5:1005–1016. doi: 10.2174/1568026054985894. [DOI] [PubMed] [Google Scholar]
  • 78.Alaimo PM, Knight ZA, Shokat KM. Targeting the gatekeeper residue in phosphoinositide 3-kinases. Bioorg. Med. Chem. 2005;13:2825–2836. doi: 10.1016/j.bmc.2005.02.021. [DOI] [PubMed] [Google Scholar]
  • 79.Herberich B, et al. Discovery of highly selective and potent p38 inhibitors based on a phthalazine scaffold. J. Med. Chem. 2008;51:6271–6279. doi: 10.1021/jm8005417. [DOI] [PubMed] [Google Scholar]
  • 80.Dambach DM. Potential adverse effects associated with inhibition of p38α/β MAP kinases. Curr. Topics Med. Chem. 2005;5:929–939. doi: 10.2174/1568026054985911. [DOI] [PubMed] [Google Scholar]
  • 81.Giovannini MG, et al. β-amyloid-induced inflammation and cholinergic hypofunction in the rat brain in vivo: involvement of the p38 MAPK pathway. Neurobiol. Dis. 2002;11:257–274. doi: 10.1006/nbdi.2002.0538. [DOI] [PubMed] [Google Scholar]
  • 82.Wang ZQ, Wu DC, Huang FP, Yang GY. Inhibition of MEK/ERK 1/2 pathway reduces pro-inflammatory cytokine interleukin-1 expression in focal cerebral ischemia. Brain Res. 2003;996:55–66. doi: 10.1016/j.brainres.2003.09.074. [DOI] [PubMed] [Google Scholar]
  • 83.Ciruela A, et al. Identification of MEK1 as a novel target for the treatment of neuropathic pain. Br. J. Pharmacol. 2003;138:751–756. doi: 10.1038/sj.bjp.0705103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Farrokhnia N, Ericsson A, Terent A, Lennmyr F. MEK-inhibitor U0126 in hyperglycaemic focal ischaemic brain injury in the rat. Eur. J. Clin. Invest. 2008;38:679–685. doi: 10.1111/j.1365-2362.2008.01990.x. [DOI] [PubMed] [Google Scholar]
  • 85.Mori T, et al. Mitogen-activated protein kinase inhibition in traumatic brain injury: in vitro and in vivo effects. J. Cerebral Blood Flow Metab. 2002;22:444–452. doi: 10.1097/00004647-200204000-00008. [DOI] [PubMed] [Google Scholar]
  • 86.Carboni S, et al. AS601245, a c-Jun NH2-terminal kinase (JNK) inhibitor, reduces axon/dendrite damage and cognitive deficits after global cerebral ischemia in gerbils. Br. J. Pharmacol. 2008;153:157–165. doi: 10.1038/sj.bjp.0707574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Yatsushige H, Ostrowski RP, Tsubokawa T, Colohan A, Zhang JH. Role of c-Jun N-terminal kinase in early brain injury after subarachnoid hemorrhage. J. Neurosci. Res. 2007;85:1436–1448. doi: 10.1002/jnr.21281. [DOI] [PubMed] [Google Scholar]
  • 88.Saporito MS, et al. Preservation of cholinergic activity and prevention of neuron death by CEP-1347/KT7515 following excitotoxic injury of the nucleus basalis magnocellularis. Neuroscience. 1998;86:461–472. doi: 10.1016/s0306-4522(98)00059-1. [DOI] [PubMed] [Google Scholar]
  • 89.Pirvola U, et al. Rescue of hearing, auditory hair cells, and neurons by CEP-1347/KT7515, an inhibitor of c-Jun N-terminal kinase activation. J. Neurosci. 2000;20:43–50. doi: 10.1523/JNEUROSCI.20-01-00043.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Apostol BL, et al. CEP-1347 reduces Huntington-associated neurotoxicity and restores BDNF levels in R6/2 mice. Mol. Cell. Neurosci. 2008;39:8–20. doi: 10.1016/j.mcn.2008.04.007. [DOI] [PubMed] [Google Scholar]
  • 91.Wang LH, Besirli CG, Johnson EM. Mixed-lineage kinases: a target for the prevention of neurodegeneration. Ann. Rev. Pharmacol. Toxicol. 2004;44:451–474. doi: 10.1146/annurev.pharmtox.44.101802.121840. [DOI] [PubMed] [Google Scholar]
  • 92.Wang W, Ma C, Mao Z, Li M. JNK inhibition as a potential strategy in treating Parkinson’s disease. Drug News Perspect. 2004;17:646–654. doi: 10.1358/dnp.2004.17.10.873916. [DOI] [PubMed] [Google Scholar]
  • 93.Resnick L, Fennell M. Targeting JNK3 for the treatment of neurodegenerative disorders. Drug Discov. Today. 2004;9:932–939. doi: 10.1016/S1359-6446(04)03251-9. [DOI] [PubMed] [Google Scholar]
  • 94.Hudkins RL, et al. Mixed lineage kinase 1 and mixed-lineage kinase 3 subtype-selective dihydronaphthyl-[3,4-a]pyrrolo[3,4-c]carbazole-5-ones: optimization, mixed-lineage kinase 1 crystallography, and oral in vivo activity in 1-methyl-4-phenyltetrahydropyridine models. J. Med. Chem. 2008;51:5680–5689. doi: 10.1021/jm8005838. [DOI] [PubMed] [Google Scholar]
  • 95.The Parkinson Study Group PRECEPT Investigators Mixed lineage kinase inhibitor CEP-1347 fails to delay disability in early Parkinson disease. Neurology. 2007;69:1480–1490. doi: 10.1212/01.wnl.0000277648.63931.c0. [DOI] [PubMed] [Google Scholar]
  • 96.Waldmeier P, Bozyczko-Coyne D, Williams M, Vaught JL. Recent clinical failures in Parkinson’s disease with apoptosis inhibitors underline the need for a paradigm shift in drug discovery for neurodegenerative diseases. Biochem. Pharmacol. 2006;72:1197–1206. doi: 10.1016/j.bcp.2006.06.031. [DOI] [PubMed] [Google Scholar]
  • 97.Wang LH, et al. Mixed lineage kinase inhibitor CEP-1347 fails to delay disability in early Parkinson disease. Neurology. 2008;71:462–463. doi: 10.1212/01.wnl.0000324506.93877.5e. [DOI] [PubMed] [Google Scholar]
  • 98.Bogatcheva NV, Adyshev D, Mambetsariev B, Moldobaeva N, Verin AD. Involvement of microtubules, p38, and Rho kinases pathway in 2-methoxyestradiol-induced lung vascular barrier dysfunction. Am. J. Physiol. Lung Cell Mol. Physiol. 2007;292:L487–L499. doi: 10.1152/ajplung.00217.2006. [DOI] [PubMed] [Google Scholar]
  • 99.Breslin JW, et al. Involvement of ROCK-mediated endothelial tension development in neutrophil-stimulated microvascular leakage. Am. J. Physiol. Heart Circ. Physiol. 2006;290:H741–H750. doi: 10.1152/ajpheart.00238.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Olson MF. Applications for ROCK kinase inhibition. Curr. Opin. Cell Biol. 2008;20:242–248. doi: 10.1016/j.ceb.2008.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Behanna HA, Watterson DM, Ralay Ranaivo H. Development of a novel bioavailable inhibitor of the calmodulin-regulated protein kinase MLCK: a lead compound that attenuates vascular leak. Biochim. Biophys. Acta. 2006;1763:1266–1274. doi: 10.1016/j.bbamcr.2006.08.007. [DOI] [PubMed] [Google Scholar]
  • 102.Zlokovic BV. The blood-brain barrier in health and chronic neurodegenerative disorders. Neuron. 2008;57:178–201. doi: 10.1016/j.neuron.2008.01.003. [DOI] [PubMed] [Google Scholar]
  • 103.Uehata M, et al. Calcium sensitization of smooth muscle mediated by Rho-associated protein kinase in hypertension. Nature. 1997;389:990–994. doi: 10.1038/40187. [DOI] [PubMed] [Google Scholar]
  • 104.Tamura M, et al. Development of specific Rho-kinase inhibitors and their clinical application. Biochim. Biophys. Acta. 2005;1754:245–252. doi: 10.1016/j.bbapap.2005.06.015. [DOI] [PubMed] [Google Scholar]
  • 105.Reynoso R, et al. A role for long chain myosin light chain kinase (MLCK-210) in microvascular hyperpermeability during severe burns. Shock. 2007;28:589–595. doi: 10.1097/SHK.0b013e31804d415f. [DOI] [PubMed] [Google Scholar]
  • 106.Wainwright MS, et al. Protein kinase involved in lung injury susceptibility: evidence from enzyme isoform genetic knockout and in vivo inhibitor treatment. Proc. Natl Acad. Sci. USA. 2003;100:6233–6238. doi: 10.1073/pnas.1031595100.The first demonstration by genetic knockout and small-molecule inhibitor approaches that MLCK is a drug discovery target for tissue barrier dysfunction.
  • 107.Rossi JL, Velentza AV, Steinhorn DM, Watterson DM, Wainwright MS. MLCK 210 gene knockout or kinase inhibition preserves lung function following endotoxin-induced lung injury in mice. Am. J. Physiol. Lung Cell Mol. Physiol. 2007;292:L1327–L1334. doi: 10.1152/ajplung.00380.2006. [DOI] [PubMed] [Google Scholar]
  • 108.Clayburgh DR, et al. Epithelial myosin light chain kinase-dependent barrier dysfunction mediates T cell activation-induced diarrhea in vivo. J. Clin. Invest. 2005;115:2702–2715. doi: 10.1172/JCI24970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Ralay Ranaivo H, et al. Protection against endotoxic shock as a consequence of reduced nitrosative stress in MLCK210 null mice. Am. J. Pathol. 2007;170:439–446. doi: 10.2353/ajpath.2007.060219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Gandhi PN, Chen SG, Wilson-Delfosse AL. Leucine-rich repeat kinase 2 (LRRK2): a key player in the pathogenesis of Parkinson’s disease. J. Neurosci. Res. 2009;87:1283–1295. doi: 10.1002/jnr.21949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Mata IF, Wedemeyer WJ, Farrer MJ, Taylor JP, Gallo KA. LRRK2 in Parkinson’s disease: protein domains and functional insights. Trends Neurosci. 2006;29:286–293. doi: 10.1016/j.tins.2006.03.006. [DOI] [PubMed] [Google Scholar]
  • 112.Wegiel J, et al. The role of overexpressed DYRK1A protein in the early onset of neurofibrillary degeneration in Down syndrome. Acta Neuropathol. 2008;116:391–407. doi: 10.1007/s00401-008-0419-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Nikolic M. The Pak1kinase: an important regulator of neuronal morphology and function in the developing forebrain. Mol. Neurobiol. 2008;37:187–202. doi: 10.1007/s12035-008-8032-1. [DOI] [PubMed] [Google Scholar]
  • 114.Knippschild W, et al. The casein kinase 1 family: participation in multiple cellular processes in eukaryotes. Cell. Signal. 2005;17:675–689. doi: 10.1016/j.cellsig.2004.12.011. [DOI] [PubMed] [Google Scholar]
  • 115.Gong CS, Iqbal K. Hyperphosphorylation of microtubule-associated protein tau: a promising therapeutic target for Alzheimer disease. Curr. Med. Chem. 2008;15:2321–2328. doi: 10.2174/092986708785909111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Pardridge WM. The blood-brain barrier: bottleneck in brain drug development. NeuroRx. 2005;2:3–14. doi: 10.1602/neurorx.2.1.3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Lipinski CA, Lombardo F, Dominy BW, Feeney PJ. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2001;46:3–26. doi: 10.1016/s0169-409x(00)00129-0.The seminal paper describing the ‘Rule of Five’ for drug-like molecular properties.
  • 118.Levin VA. Relationship of octanol/water partition coefficient and molecular weight to rat brain capillary permeability. J. Med. Chem. 1980;23:682–684. doi: 10.1021/jm00180a022. [DOI] [PubMed] [Google Scholar]
  • 119.Leeson PD, Springthorpe B. The influence of drug-like concepts on decision-making in medicinal chemistry. Nature Rev. Drug Discov. 2007;6:881–890. doi: 10.1038/nrd2445. [DOI] [PubMed] [Google Scholar]
  • 120.Nies AT. The role of membrane transporters in drug delivery to brain tumors. Cancer Lett. 2007;254:11–29. doi: 10.1016/j.canlet.2006.12.023. [DOI] [PubMed] [Google Scholar]
  • 121.Vieth M, et al. Characteristic physical properties and structural fragments of marketed oral drugs. J. Med. Chem. 2004;47:224–232. doi: 10.1021/jm030267j. [DOI] [PubMed] [Google Scholar]
  • 122.Capdeville R, Buchdunger E, Zimmermann J, Matter A. Glivec (STI571, imatinib), a rationally developed, targeted anticancer drug. Nature Rev. Drug Discov. 2002;1:493–502. doi: 10.1038/nrd839.A case study and historical review of the development of imatinib and the validation of protein kinases as drug targets.
  • 123.Wen PY, et al. Phase I/II study of imatinib mesylate for recurrent malignant gliomas: North American Brain Tumor Consortium Study 99–08. Clin. Cancer Res. 2006;12:4899–4907. doi: 10.1158/1078-0432.CCR-06-0773. [DOI] [PubMed] [Google Scholar]
  • 124.Raymond E, et al. Phase II study of imatinib in patients with recurrent gliomas of various histologies: a European Organisation for Research and Treatment of Cancer Brain Tumor Group Study. J. Clin. Oncol. 2008;26:4659–4665. doi: 10.1200/JCO.2008.16.9235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Dai H, Marbach P, Lemaire M, Hayes M, Elmquist WF. Distribution of STI-571 to the brain is limited by P-glycoprotein-mediated efflux. J. Pharmacol. Exp. Ther. 2003;304:1085–1092. doi: 10.1124/jpet.102.045260. [DOI] [PubMed] [Google Scholar]
  • 126.Rishton GM, LaBonte K, Williams AJ, Kassam K, Kolovanov E. Computational approaches to the prediction of blood-brain barrier permeability: a comparative analysis of central nervous system drugs versus secretase inhibitors for Alzheimer’s disease. Curr. Opin. Drug Discov. Devel. 2006;9:303–313. [PubMed] [Google Scholar]
  • 127.Kamath AV, Wang J, Lee FY, Marathe PH. Preclinical pharmacokinetics and in vitro metabolism of dasatinib (BMS-354825): a potent oral multi-targeted kinase inhibitor against SRC and BCR-ABL. Cancer Chemother. Pharmacol. 2008;61:365–376. doi: 10.1007/s00280-007-0478-8. [DOI] [PubMed] [Google Scholar]
  • 128.Porkka K, et al. Dasatinib crosses the blood-brain barrier and is an efficient therapy for central nervous system Philadelphia chromosome-positive leukemia. Blood. 2008;112:1005–1012. doi: 10.1182/blood-2008-02-140665. [DOI] [PubMed] [Google Scholar]
  • 129.Arendt RM, Greenblatt DJ, Liebisch DC, Luu MD, Paul SM. Determinants of benzodiazepine brain uptake: lipophilicity versus binding affinity. Psychopharmacology. 1987;93:72–76. doi: 10.1007/BF02439589. [DOI] [PubMed] [Google Scholar]
  • 130.Obradovic T, Dobson G, Shingaki T, Kungu T, Hidalgo I. Assessment of the first and second generation antihistamines brain penetration and role of P-glycoprotein. Pharm. Res. 2007;24:318–327. doi: 10.1007/s11095-006-9149-4. [DOI] [PubMed] [Google Scholar]
  • 131.Mahar Doan K, et al. Passive permeability and P-glycoprotein-mediated efflux differentiate central nervous system (CNS) and non-CNS marketed drugs. J. Pharmacol. Exp. Ther. 2002;303:1029–1037. doi: 10.1124/jpet.102.039255. [DOI] [PubMed] [Google Scholar]
  • 132.Morphy R. The influence of target family and functional activity on the physicochemical properties of pre-clinical compounds. J. Med. Chem. 49:2969–2978. doi: 10.1021/jm0512185. [DOI] [PubMed] [Google Scholar]
  • 133.Chico LK, et al. Molecular properties and CYP2D6 substrates: central nervous system therapeutics case study and pattern analysis of a substrate database. Drug Metab. Disp. 2009 Aug 6; doi: 10.1124/dmd.109.028134. doi:10.1124/dmd.109.028134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Rendic S. Summary of information on human CYP enzymes: human P450 metabolism data. Drug Metab. Rev. 2002;34:83–448. doi: 10.1081/dmr-120001392. [DOI] [PubMed] [Google Scholar]
  • 135.Kramer JA, Sagartz JE, Morris DL. The application of discovery toxicology and pathology towards the design of safer pharmaceutical lead candidates. Nature Rev. Drug Discov. 2007;6:636–649. doi: 10.1038/nrd2378. [DOI] [PubMed] [Google Scholar]
  • 136.Kirchheiner J, et al. Pharmacokinetics of codeine and its metabolite morphine in ultra-rapid metabolizers due to CYP2D6 duplication. Pharmacogenomics. 2007;7:257–265. doi: 10.1038/sj.tpj.6500406. [DOI] [PubMed] [Google Scholar]
  • 137.Peng B, Lloyd P, Schran H. Clinical pharmacokinetics of imatinib. Clin. Pharmacokinet. 2005;44:879–894. doi: 10.2165/00003088-200544090-00001. [DOI] [PubMed] [Google Scholar]
  • 138.Ohren JF, et al. Structures of human MAP kinase kinase 1 (MEK1) and MEK2 describe novel noncompetitive kinase inhibition. Nature Struct. Mol. Biol. 2004;11:1192–1197. doi: 10.1038/nsmb859. [DOI] [PubMed] [Google Scholar]
  • 139.VanScyoc WS, Holdgate GA, Sullivan JE, Ward WH. Enzyme kinetics and binding studies on inhibitors of MEK protein kinase. Biochemistry. 2008;47:5017–5027. doi: 10.1021/bi701811x. [DOI] [PubMed] [Google Scholar]
  • 140.Adams JA. Kinetic and catalytic mechanisms of protein kinases. Chem. Rev. 2001;101:2271–2290. doi: 10.1021/cr000230w. [DOI] [PubMed] [Google Scholar]
  • 141.Congreve M, Chessari G, Tisi D, Woodhead AJ. Recent developments in fragment-based drug discovery. J. Med. Chem. 2008;51:3661–3680. doi: 10.1021/jm8000373.This paper provides an update on the fragment-based approach to drug discovery and case studies of success with the approach.
  • 142.Chessari G, Woodhead AJ. From fragment to clinical candidate – a historical perspective. Drug Discov. Today. 2009;14:668–675. doi: 10.1016/j.drudis.2009.04.007. [DOI] [PubMed] [Google Scholar]
  • 143.de Kloe GE, Bailey D, Leurs R, de Esch IJP. Transforming fragments into candidates: small becomes big in medicinal chemistry. Drug Discov. Today. 2009;14:630–646. doi: 10.1016/j.drudis.2009.03.009. [DOI] [PubMed] [Google Scholar]
  • 144.Nath N, et al. 5-aminoimidazole-4-4-carboxamide ribonucleoside: a novel immunomodulator with therapeutic efficacy in experimental autoimmune encephalomyelitis. J. Immunol. 2005;175:566–574. doi: 10.4049/jimmunol.175.1.566. [DOI] [PubMed] [Google Scholar]
  • 145.McCullough LD, et al. Pharmacological inhibition of AMP-activated protein kinase provides neuroprotection in stroke. J. Biol. Chem. 2005;280:20493–20502. doi: 10.1074/jbc.M409985200. [DOI] [PubMed] [Google Scholar]
  • 146.Takle AK, et al. The identification of potent, selective and CNS penetrant furan-based inhibitors of B-Raf kinase. Bioorg. Med. Chem. Lett. 2008;18:4373–4376. doi: 10.1016/j.bmcl.2008.06.070. [DOI] [PubMed] [Google Scholar]
  • 147.Hilton GD, Stoica BA, Byrnes KR, Faden AI. Roscovitine reduces neuronal loss, glial activation, and neurologic deficits after brain trauma. J. Cereb. Blood Flow Metab. 2008;28:1845–1859. doi: 10.1038/jcbfm.2008.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Wang F, et al. Inhibition of cyclin-dependent kinases improves CA1 neuronal survival and behavioral performance after global ischemia in the rat. J. Cereb. Blood Flow Metab. 2002;22:171–182. doi: 10.1097/00004647-200202000-00005. [DOI] [PubMed] [Google Scholar]
  • 149.Weishaupt JH, et al. Inhibition of CDK5 is protective in necrotic and apoptotic paradigms of neuronal cell death and prevents mitochondrial dysfunction. Mol. Cell Neurosci. 2003;24:489–502. doi: 10.1016/s1044-7431(03)00221-5. [DOI] [PubMed] [Google Scholar]
  • 150.Selenica ML, et al. Efficacy of small-molecule glycogen synthase kinase-3 inhibitors in the postnatal rat model of tau hyperphosphorylation. Br. J. Pharmacol. 2007;152:959–979. doi: 10.1038/sj.bjp.0707471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Muñoz-Montaño JR, Moreno FJ, Avila J, Diaz-Nido J. Lithium inhibits Alzheimer’s disease-like tau protein phosphorylation in neurons. FEBS Lett. 1997;411:183–188. doi: 10.1016/s0014-5793(97)00688-1. [DOI] [PubMed] [Google Scholar]
  • 152.Kozikowski AP, et al. Structure-based design leads to the identification of lithium mimetics that block mania-like effects in rodents. possible new GSK-3β herapies for bipolar disorders. J. Am. Chem. Soc. 2007;129:8328–8332. doi: 10.1021/ja068969w. [DOI] [PubMed] [Google Scholar]
  • 153.Cowper-Smith CD, Anger GJA, Magal E, Norman MH, Robertson GS. Delayed administration of a potent cyclin dependent kinase and glycogen synthase kinase 3β inhibitor produces long-term neuroprotection in a hypoxia-ischemia model of brain inury. Neurosci. 2008;155:864–875. doi: 10.1016/j.neuroscience.2008.05.051. [DOI] [PubMed] [Google Scholar]
  • 154.Kelly S, et al. Glycogen synthase kinase 3β inhibitor Chir025 reduces neuronal death resulting from oxygen-glucose deprivation, glutamate excitotoxicity, and cerebral ischemia. Exp. Neurol. 2004;188:378–386. doi: 10.1016/j.expneurol.2004.04.004. [DOI] [PubMed] [Google Scholar]
  • 155.Dewhurst S, Maggirwar SB, Schifitto G, Gendelman HE, Gelbard HA. Glycogen synthase kinase 3 beta (GSK-3β) as a therapeutic target in neuroAIDS. J. Neuroimmune. Pharmacol. 2006;2:93–96. doi: 10.1007/s11481-006-9051-1. [DOI] [PubMed] [Google Scholar]
  • 156.Wang W, et al. Inhibition of glycogen synthase kinase-3β protects dopaminergic neurons from MPTP toxicity. Neuropharmacology. 2007;52:1678–1684. doi: 10.1016/j.neuropharm.2007.03.017. [DOI] [PubMed] [Google Scholar]
  • 157.Thotala DK, Hallahan DE, Yazlovitskaya EM. Inhibition of glycogen synthase kinase 3 beta attenuates neurocognitive dysfunction resulting from cranial irradiation. Cancer Res. 2008;68:5859–5868. doi: 10.1158/0008-5472.CAN-07-6327. [DOI] [PubMed] [Google Scholar]
  • 158.Shapira M, et al. Role of glycogen synthase kinase-3β in early depressive behavior induced by mild traumatic brain injury. Mol. Cell. Neurosci. 2007;34:571–577. doi: 10.1016/j.mcn.2006.12.006. [DOI] [PubMed] [Google Scholar]
  • 159.Trieu VN, Liu R, Liu XP, Uckun FM. A specific inhibitor of janus kinase-3 increases survival in a transgenic mouse model of amyotrophic lateral sclerosis. Biochem. Biophys. Res. Commun. 2000;267:22–25. doi: 10.1006/bbrc.1999.1905. [DOI] [PubMed] [Google Scholar]
  • 160.Wang W, et al. SP600125, a new JNK inhibitor, protects dopaminergic neurons in the MPTP model of Parkinson’s disease. Neurosci. Res. 2004;48:195–202. doi: 10.1016/j.neures.2003.10.012. [DOI] [PubMed] [Google Scholar]
  • 161.Carr KD, et al. Effects of the MEK inhibitor, SL-327, on rewarding, motor- and cellular-activating effects of d-amphetamine and SKF-82958, and their augmentation by food restriction in rat. Psychopharmacology. 2009;201:495–506. doi: 10.1007/s00213-008-1313-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Alessandrini A, Namura S, Moskowitz MA, Bonventre JV. MEK1 protein kinase inhibition protects against damage resulting from focal cerebral ischemia. Proc. Natl Acad. Sci. USA. 1999;96:12866–12869. doi: 10.1073/pnas.96.22.12866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Gu Z, Jiang Q, Zhang G. Extracellular signal-regulated kinase and c-Jun N-terminal protein kinase in ischemic tolerance. NeuroReport. 2001;12:3487–3491. doi: 10.1097/00001756-200111160-00023. [DOI] [PubMed] [Google Scholar]
  • 164.Ciruela A, et al. Identification of MEK1 as a novel target for the treatment of neuropathic pain. Br. J. Pharmacol. 2003;138:751–756. doi: 10.1038/sj.bjp.0705103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Saporito MS, Hudkins RL, Maroney AC. Discovery of CEP-1347/KT-7515, an inhibitor of the JNK/SAPK pathway for the treatment of neurodegenerative diseases. Prog. Med. Chem. 2002;40:23–62. doi: 10.1016/s0079-6468(08)70081-x. [DOI] [PubMed] [Google Scholar]
  • 166.Conforti P, et al. Blood level of brain-derived neutrophic factor mRNA is progressively reduced in rodent models of Huntington’s disease: restoration by the neuroprotective compound CEP-1347. Mol. Cell Neurosci. 2008;39:1–7. doi: 10.1016/j.mcn.2008.04.012. [DOI] [PubMed] [Google Scholar]
  • 167.Glicksman MA, et al. CEP-1347/KT7515 prevents motor neuronal programmed cell death and injury-induced dedifferentiation in vivo. J. Neurobiol. 1998;35:361–370. [PubMed] [Google Scholar]
  • 168.Kokubu K, Tani E, Nakano M, Minami N, Shindo H. Effects of ML-9 on experimental delayed cerebral vasospasm. J. Neurosurg. 1989;71:916–922. doi: 10.3171/jns.1989.71.6.0916. [DOI] [PubMed] [Google Scholar]
  • 169.Meikle L, et al. Response of a neuronal model of tuberous sclerosis to mammalian target of rapamycin (mTOR) inhibitors: effects on mTORC1 and Akt signaling lead to improved survival and function. J. Neurosci. 2008;28:5422–5432. doi: 10.1523/JNEUROSCI.0955-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Ravikumar B, et al. Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nature Genet. 2004;36:585–595. doi: 10.1038/ng1362. [DOI] [PubMed] [Google Scholar]
  • 171.Ito T, et al. Glial phosphorylated p38 MAP kinase mediates pain in a rat model of lumbar disc herniation and induces motor dysfunction in a rat model of lumbar spinal canal stenosis. Spine. 2007;32:159–167. doi: 10.1097/01.brs.0000251437.10545.e9. [DOI] [PubMed] [Google Scholar]
  • 172.Stirling DP, Liu J, Plunet W, Steeves JD, Tetzlaff W. SB203580, a p38 mitogen-activated protein kinase inhibitor, fails to improve functional outcome following a moderate spinal cord injury in rat. Neuroscience. 2008;155:128–137. doi: 10.1016/j.neuroscience.2008.05.007. [DOI] [PubMed] [Google Scholar]
  • 173.Ferraccioli GF. VX-745 Vertex Pharmaceuticals. Curr. Opin. Anti-Inflamm. Immun. Invest. Drugs. 2000;2:74–77. [PubMed] [Google Scholar]
  • 174.Su Y, Ryder J, Ni B. Inhibition of Aβ production and APP maturation by a specific PKA inhibitor. FEBS Lett. 2003;546:407–410. doi: 10.1016/s0014-5793(03)00645-8. [DOI] [PubMed] [Google Scholar]
  • 175.Zhao WQ, et al. Inhibitors of cAMP-dependent protein kinase impair long-term memory formation in day-old chicks. Neurobiol. Learn. Mem. 1995;64:106–118. doi: 10.1006/nlme.1995.1049. [DOI] [PubMed] [Google Scholar]
  • 176.Smith FL, Javed RR, Smith PA, Dewey WL, Gabra BH. PKC and PKA inhibitors reinstate morphine-induced behaviors in morphine tolerant mice. Pharmacol. Res. 2006;54:474–480. doi: 10.1016/j.phrs.2006.09.007. [DOI] [PubMed] [Google Scholar]
  • 177.Otuki MF, et al. Antinociceptive properties of mixture of α-amyrin and β-amyrin triterpenes: evidence for participation of protein kinase C and protein kinase A pathways. J. Pharmacol. Exp. Ther. 2005;313:310–318. doi: 10.1124/jpet.104.071779. [DOI] [PubMed] [Google Scholar]
  • 178.Liu SJ, et al. Overactivation of glycogen synthase kinase-3 by inhibition of phosphoinositol-3 kinase and protein kinase C leads to hyperphosphorylation of tau and impairment of spatial memory. J. Neurochem. 2003;87:1333–1344. doi: 10.1046/j.1471-4159.2003.02070.x. [DOI] [PubMed] [Google Scholar]
  • 179.Zarate CA, Jr, et al. Efficacy of a protein kinase C inhibitor (tamoxifen) in the treatment of acute mania: a pilot study. Bipolar Disord. 2007;9:561–570. doi: 10.1111/j.1399-5618.2007.00530.x. [DOI] [PubMed] [Google Scholar]
  • 180.Teicher BA, et al. Enzymatic rationale and preclinical support for a potent protein kinase Cβ inhibitor in cancer therapy. Adv. Enzyme. Regul. 1999;39:313–327. doi: 10.1016/s0065-2571(98)00026-0. [DOI] [PubMed] [Google Scholar]
  • 181.da Cunha JM, Rae GA, Ferreira SH, Cunha Fde Q. Endothelins induce ETB receptor-mediated mechanical hypernociception in rat hindpaw: roles of cAMP and protein kinase C. Eur. J. Pharm. 2004;501:87–94. doi: 10.1016/j.ejphar.2004.08.004. [DOI] [PubMed] [Google Scholar]
  • 182.Parnham M, Sies H. Ebselen: prospective therapy for cerebral ischaemia. Exp. Opin. Investig. Drugs. 2000;9:607–619. doi: 10.1517/13543784.9.3.607. [DOI] [PubMed] [Google Scholar]
  • 183.Minami N, Tani E, Maeda Y, Yamaura I, Fukami M. Effects of inhibitors of protein kinase C and calpain in experimental delayed cerebral vasospasm. J. Neurosurg. 1992;76:111–118. doi: 10.3171/jns.1992.76.1.0111. [DOI] [PubMed] [Google Scholar]
  • 184.Nishizawa S, et al. Attenuation of canine cerebral vasospasm after subarachnoid hemorrhage by protein kinase C inhibitors despite augmented phosphorylation of myosin light chain. J. Vasc. Res. 2003;40:169–178. doi: 10.1159/000070714. [DOI] [PubMed] [Google Scholar]
  • 185.Tokuyama S, Ho IK, Yamamoto T. A protein kinase inhibitor, H-7, blocks naloxone-precipitated changes in dopamine and its metabolites in the brains of opioid-dependent rats. Brain Res. Bull. 2000;52:363–369. doi: 10.1016/s0361-9230(00)00273-2. [DOI] [PubMed] [Google Scholar]
  • 186.Zhang D, Kanthasamy A, Yang Y, Anantharam V, Kanthasamy A. Protein kinase Cδ negatively regulates tyrosine hydroxylase activity and dopamine synthesis by enhancing protein phosphatase-2A activity in dopaminergic neurons. J. Neurosci. 2007;27:5349–5362. doi: 10.1523/JNEUROSCI.4107-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Inglis K, et al. Polo-like kinase 2 (PLK2) phosphorylates α-synuclein at serine 129 in central nervous system. J. Biol. Chem. 2009;284:2598–2602. doi: 10.1074/jbc.C800206200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Zhou Y, et al. Nonsteroidal anti-inflammatory drugs can lower amyloidogenic Aβ42 by inhibiting Rho. Science. 2003;302:1215–1217. doi: 10.1126/science.1090154. [DOI] [PubMed] [Google Scholar]
  • 189.Satoh S, Toshima Y, Ikegaki I, Iwasaki M, Asano T. Wide therapeutic time window for fasudil neuroprotection against ischemia-induced delayed neuronal death in gerbils. Brain Res. 2007;1128:175–180. doi: 10.1016/j.brainres.2006.10.027. [DOI] [PubMed] [Google Scholar]
  • 190.Satoh S, et al. Pharmacological profile of hydroxy fasudil as a selective rho kinase inhibitor on ischemic brain damage. Life Sci. 2001;69:1441–1453. doi: 10.1016/s0024-3205(01)01229-2. [DOI] [PubMed] [Google Scholar]
  • 191.Sun X, et al. The selective Rho-kinase inhibitor Fasudil is protective and therapeutic in experimental autoimmune encephalomyelitis. J. Neuroinflamm. 2006;180:126–134. doi: 10.1016/j.jneuroim.2006.06.027. [DOI] [PubMed] [Google Scholar]
  • 192.Inan S, Büyükafsar K. Antiepileptic effects of two Rho-kinase inhibitors, Y-27632 and fasudil, in mice. Br. J. Pharmacol. 2008;155:44–51. doi: 10.1038/bjp.2008.225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Büyükafsar K, et al. Rho-kinase inhibitor, Y-27632, has an antinociceptive effect in mice. Eur. J. Pharmacol. 2006;541:49–52. doi: 10.1016/j.ejphar.2006.04.042. [DOI] [PubMed] [Google Scholar]
  • 194.Fournier AE, Takizawa BT, Strittmatter SM. Rho kinase inhibition enhances axonal regeneration in the injured CNS. J. Neurosci. 2003;23:1416–1423. doi: 10.1523/JNEUROSCI.23-04-01416.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Jadhav V, Matchett G, Hsu FP, Zhang JH. Inhibition of Src tyrosine kinase and effect on outcomes in a new in vivo model of surgically induced brain injury. J. Neurosurg. 2007;106:680–686. doi: 10.3171/jns.2007.106.4.680. [DOI] [PubMed] [Google Scholar]
  • 196.Liu DZ, et al. Src kinase inhibition decreases thrombin-induced injury and cell cycle re-entry in striatal neurons. Neurobiol. Dis. 2008;30:201–211. doi: 10.1016/j.nbd.2008.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Tran TT, et al. Inhibiting TGF-β signaling restores immune surveillance in the SMA-560 glioma model. Neuro. Oncol. 2007;9:259–270. doi: 10.1215/15228517-2007-010. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

S1
s2
s3
s4

RESOURCES