Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Mar 1.
Published in final edited form as: Neurobiol Dis. 2009 Nov 24;37(3):549–557. doi: 10.1016/j.nbd.2009.11.013

Cell cycle inhibition as a strategy for treatment of central nervous system diseases which must not block normal neurogenesis

Da-Zhi Liu 1,*, Bradley P Ander 1, Frank R Sharp 1
PMCID: PMC2823995  NIHMSID: NIHMS161771  PMID: 19944161

Abstract

Classically, the cell cycle is regarded as the central process leading to cellular proliferation. However, increasing evidence over the last decade supports the notion that neuronal cell cycle re-entry results in post-mitotic death. A mature neuron that re-enters the cell cycle can neither advance to a new G0 quiescent state nor revert to its earlier G0 state. This presents a critical dilemma to the neuron from which death may be an unavoidable, but necessary, outcome for adult neurons attempting to complete the cell cycle. In contrast, tumor cells that undergo aberrant cell cycle re-entry divide and can survive. Thus, cell cycle inhibition strategies are of interest in cancer treatment, but may also represent an important means of protecting neurons. In this review, we put forth the concept of the “expanded cell cycle” and summarize the cell cycle proteins, signal transduction events and mitogenic molecules that can drive a neuron into the cell cycle in various CNS diseases. We also discuss the pharmacological approaches that interfere with the mitogenic pathways and prevent mature neurons from attempting cell cycle re-entry, protecting them from cell death. Lastly, future attempts at blocking the cell cycle to rescue mature neurons from injury should be designed so as to not block normal neurogenesis.

Keywords: neuron, cell cycle, mitogen, mitogenic pathway, cell cycle re-entry, cell stress, CNS diseases, neurogenesis

Introduction

Postmortem studies over the last decade have revealed pathological evidence of aberrant expression of cell cycle related molecules in the neurons of the hippocampus, subiculum, locus coeruleus and dorsal raphe nuclei. Direct proof of DNA replication was also identified in brains of patients with Alzheimer's disease (AD) (McShea et al., 1997; Nagy et al., 1997; Vincent et al., 1997; Busser et al., 1998; Yang et al., 2001), epilepsy (Nagy and Esiri, 1998), Parkinson's disease (PD) (Jordan-Sciutto et al., 2003) and amyotrophic lateral sclerosis (ALS) (Ranganathan and Bowser, 2003). These important discoveries stimulated new hypotheses and studies challenging the traditional concept that post-mitotic neurons are terminally differentiated and maintained in the G0 quiescent phase. Although still controversial, evidence of cell cycle re-entry in neurons related to pathological conditions has been further confirmed in a number of reports, including experiments in primary neuron cultures (Copani et al., 1999; Copani et al., 2006; Zhang et al., 2006; Rao et al., 2007; Liu et al., 2008) and animal models of disease including AD (Herrup et al., 2004; Khurana et al., 2006; Yang et al., 2006), ALS (Nguyen et al., 2003), stroke (Imai et al., 2002; O'Hare et al., 2002), traumatic brain injury (TBI) (Di Giovanni et al., 2005) and cerebral hypoxia-ischemia (Kuan et al., 2004). Furthermore, our recent genomic and bioinformatic studies have consistently placed “cell cycle” as among the top ranked functional categories for gene transcripts that are altered in rats one day following any of several different types of insults to the central nervous system (CNS) (e.g., cerebral ischemia, intracerebral hemorrhage and kainate-induced seizures) (Tang et al., 2001; Tang et al., 2002; Liu et al., 2009b).

These examples demonstrate aberrant cell cycle re-entry as a hallmark of CNS diseases with dying neurons, and while cell cycle re-entry in more commonly associated with tumor cells, there are very different consequences of this event between these tissues. In contrast to neurons, when tumor cells re-enter the cell cycle, they survive and may continue to proliferate in the presence of an oncogene. For reasons not completely understood, a mature neuron that re-enters the cell cycle is neither able to advance to a new G0 quiescent state nor revert to its earlier G0 state. This presents a critical dilemma to the neuron from which death may be an unavoidable, but necessary, outcome for these mature neurons attempting to complete the cell cycle. Since re-entry into the cell cycle by neurons has been associated with many diseases and linked inextricably to death, the cell cycle represents a viable target for treatments and therapies, so long as the consequences on other cell types, such as neuroprogenitor cells, are considered.

As a way to describe potential therapeutic targets, we propose the “expanded cell cycle” — one which includes not only the traditional cell cycle proteins, but also the mitogenic molecules and the signaling pathways that interact with them. The “expanded cell cycle” includes some of the current targets for treating CNS diseases. It provides a composite perspective encompassing a broad range of molecules representing potential targets and thus approaches that can serve as treatments for CNS diseases by sharing a common outcome — cell cycle inhibition. A detailed description of the cell cycle and its components follows before we discuss the specific interactions that neurons have with cell cycle proteins.

Classical proteins and regulators of the cell cycle

The cell cycle is the series of events that lead to cell replication (Norbury and Nurse, 1992; Dirks and Rutka, 1997; Malumbres and Barbacid, 2001; Schwartz and Shah, 2005). In brief, the release of cells from a quiescent state (G0) results in their entry into the first gap phase (G1), during which the cells prepare for DNA replication in the synthetic phase (S). This is followed by the second gap phase (G2) and mitosis phase (M). When cells cease proliferating, either due to the presence of specific anti-mitogenic signals, or the absence of pro-mitogenic signals, they exit the cycle and enter the G0 quiescent phase. A majority of types of newly divided G0 cells can re-enter the cell cycle after passing specified checkpoints, whereas some types of cells, such as neurons, cannot. Because such a large number of molecules involved in the cell cycle have been discovered and characterized, we will provide a brief overview of these below.

Cyclin-dependent kinases and cyclins

Cyclin-dependent kinases (Cdks) are a group of serine/threonine kinases that form active heterodimeric complexes following binding to their regulatory subunits, cyclins (Malumbres and Barbacid, 2001). There are two main families of cyclins: (1) mitotic cyclins (e.g., cyclin A, cyclin B) and (2) G1 cyclins (e.g., cyclin C, cyclin D, cyclin E) (Dirks and Rutka, 1997). Several Cdks — mainly Cdk4, Cdk6, Cdk2, Cdk1, and possibly Cdk3 — cooperate to drive cells through the cell cycle (Malumbres and Barbacid, 2001). For example, Cdk4 and Cdk6 form active complexes with the D-type cyclins (cyclins D1, D2 and D3), which are thought to be involved in early G1 (Baldin et al., 1993; Ohtsubo and Roberts, 1993; Quelle et al., 1993; Resnitzky et al., 1994). The complexes of Cdk2 with cyclins E1 and E2 are required to complete G1 and initiate S phase (Resnitzky et al., 1994; Ohtsubo et al., 1995), whereas Cdk2 with cyclin A control S/G transition (Dirks and Rutka, 1997). Translocation of cyclin B with Cdk1 from cytoplasm into the nucleus heralds the onset of mitosis (together, cyclin B and Cdk1 are also called `mitosis promoting factor'), and the destruction of cyclin B is required for exit from mitosis (King et al., 1994; Pines, 1995). The role of Cdk3 is still obscure, mainly due to its low expression levels (Malumbres and Barbacid, 2001).

Cyclin-dependent kinase inhibitors

There are two subclasses of cyclin-dependent kinase inhibitors (CdkIs) - the Ink4 family (including p16Ink4a, p15Ink4b, p18Ink4c and p19Ink4d) that prevents the formation of cyclin/Cdk complexes; and the Cip/Kip family (including p21Cip1, p27Kip1 and p57Kip2) that inhibits the kinase activity of the already formed cyclin/cdk complexes (Shankland and Wolf, 2000; Nagata et al., 2003; Copani and Nicoletti, 2005). Thus, these inhibitors regulate the cell cycle via assessing damage and arresting progress at any of several defined checkpoints (Sherr and Roberts, 1995; Dirks and Rutka, 1997; Carnero et al., 2000; Yang and Herrup, 2007).

Cdk substrates

The primary substrates of Cdk4/6 and Cdk2 in G1 progression are members of the retinoblastoma protein (Rb) family, including p107 and p130 (Lundberg and Weinberg, 1998; Harbour et al., 1999; Ezhevsky et al., 2001). Rb family members are phosphorylated by activated Cdk4/6/cyclin D and Cdk2/cyclin E complexes (Morgan, 1997; Adams, 2001). The pRb is released from the transcription factor complex E2F/DP, which then activates genes required for transition to the S phase (Malumbres and Barbacid, 2001).

Cell cycle re-entry in post-mitotic neurons results in death

Under physiological conditions, neurons are subjected to a variety of stimuli and signals. These include mitogenic signals that promote re-entry into the cell cycle, and also a series of anti-mitogenic factors that strive to maintain the neuron at rest (Copani and Nicoletti, 2005). However once brain injuries occur, this balance is lost. For example, some cell cycle proteins (e.g., cyclin D, PCNA and GADD34) are produced in mature neurons very soon after experimental rat brain ischemia (Guegan et al., 1997; Li et al., 1997; Katchanov et al., 2001; Imai et al., 2002). In addition, expression of cell cycle proteins was also observed in the brains of AD patients who had mild cognitive impairment (Yang et al., 2001), and 6–8 months before the onset of amyloid beta (Aβ) deposition in the Aβ precursor protein (APP) transgenic mouse models of AD (Yang et al., 2006; Varvel et al., 2008). These findings suggest that the initiation of cell cycle protein expression is an early event in these disease processes that may eventually lead to the death of mature neurons.

However, the expression of cell cycle proteins is not always associated with cell cycle re-entry by neurons. Recent studies have demonstrated that some core cell cycle proteins serve diverse post-mitotic functions that span various developmental stages of a neuron, including neuronal migration, axonal elongation, axonal pruning, dendrite morphogenesis, and synaptic maturation and plasticity (Frank and Tsai, 2009; Kim et al., 2009). Additionally, we, and others, have observed sporadic expression of cyclin D in unperturbed normal primary neurons, but there was no active Cdk4 detected in those neurons (Rao et al., 2007; Liu et al., 2008). Since G0/G1 transition is dependent on cyclin D/Cdk4 complex formation, cyclin D expression without active Cdk4 means that the control neurons could not re-enter the cell cycle (Rao et al., 2007; Liu et al., 2008). When subjected to a mitogenic stimulus like thrombin, the neurons did re-enter the cell cycle, ultimately dying via apoptosis (Rao et al., 2007; Liu et al., 2008).

This supports the idea of a “two hit hypothesis”, similar to that first proposed by Zhu et al. and Yang et al. (Zhu et al., 2004; Yang and Herrup, 2007; Zhu et al., 2007). In this case the two conditions that must be met in order for aberrant cell cycle re-entry to occur in neurons are: (1) an elevation in cell cycle proteins and (2) an increase in promitogenic signals. Thus, even though mature neurons may express some cell cycle proteins, the amount produced is not sufficient on its own to drive the mature neuron to re-enter the cell cycle. The final death of the neurons likely requires the stimulus of additional pro-mitogenic molecules, such as thrombin, Aβ, reactive oxygen species (ROS), nitric oxide (NO), and others, which when elevated will trigger the mitogenic signal cascades in the injured neurons. Once mitogenic signaling is stimulated beyond a certain threshold, neurons appear to exit their quiescent state and re-enter the cell cycle.

The extent to which neurons proceed into the cell cycle and the phases whereby cell cycle re-entry leads to apoptosis vary in response to the mitogenic stimuli. Thrombin-treated neurons do not proceed to S phase, but rather die via apoptosis after the G0/G1 transition (Rao et al., 2007; Liu et al., 2008). This may be facilitated by the thrombin-induced decreases in G1/S regulators cyclin C and Cdk3, Cdk2 and Cdk1 (Rao et al., 2008). However, Aβ challenged neurons can proceed through S phase before dying (Copani et al., 1999). Yang et al. further confirmed the G1/S transition of cell cycle reactivated neurons by using fluorescent in situ hybridization (FISH) on samples from AD patients (Yang et al., 2003). This was a direct indication that DNA replication does occur in the mature neurons, as the FISH methodology differentiates between DNA replication and other synthetic events such as DNA repair.

However, these neurons that re-enter the cell cycle become trapped in S phase. They neither finish dividing nor revert to their G0 quiescent state (Katchanov et al., 2001; Yang et al., 2003; Kuan et al., 2004; Pallas and Camins, 2006). The most plausible explanation is that several factors involved in cell cycle progression are lost or inhibited in mature neurons. On the other hand, a neuron that re-enters the cell cycle cannot revert to an earlier G0 either, because the transitions through the mitotic cell cycle are irreversible processes (Novak et al., 2007). Presumably, the failure to complete the cell cycle in these stressed neurons triggers specific apoptotic or other cell death mechanisms to rid the tissue of these cells. Since cell cycle inhibition can prevent a neuron from reentering the cell cycle, and neuron death (or loss) is a general consequence in neurological diseases (Brasnjevic et al., 2008), cell cycle inhibition appears to be a candidate strategy for the treatment of these diseases.

Aberrant cell cycle re-entry in disease

Aberrant cell cycle re-entry is the hallmark of many tumor cells (Bartkova et al., 1996; Smith et al., 2007; Ishikawa et al., 2009). The cell cycle inhibitors (e.g., Cdk inhibitors) have been widely studied as cancer therapeutics. They have been used to inhibit growth of several types of tumor cells in numerous preclinical studies, both in vitro and in vivo (Drees et al., 1997; Schwartz et al., 1997; Arguello et al., 1998; Erickson et al., 1998; Chien et al., 1999; Shapiro et al., 1999; Tirado et al., 2005). Several Cdk inhibitors (Flavopiridol, UCN-01, Roscovitine, AT7519) have advanced to human clinical trials for evaluation as treatment for a broad range of solid tumors and hematological malignancies such as chronic lymphocytic leukemia (CLL) (Senderowicz, 1999; Senderowicz and Sausville, 2000; Whittaker et al., 2007; Wyatt et al., 2008).

Although tumor cells undergo uncontrolled proliferation, many tumors originate from adult tissues, in which the majority of cells are in the G0 quiescent phase (Malumbres and Barbacid, 2001). Thus, the cells that go on to form tumors and mature neurons share a common G0 state of quiescence. However, if tumor cells re-enter the cell cycle, they survive and often proliferate, whereas mature neurons will die. Therefore, cell cycle inhibition helps protect neurons, but kills tumor cells. This is strongly supported by experiments that show Cdk inhibition prevents the death of nerve growth factor (NGF)-differentiated PC12 neuronal cells, but promotes the death of naive PC12 (pheochromocytoma) tumor cells (Park et al., 1996; Park et al., 1997). Thus, cell cycle inhibition may be useful for treatment of CNS diseases wherein cell cycle re-entry occurs. This is very plausible at present since there are many innovative approaches being tested as cancer therapies, such as miRNA-based gene therapy techniques (Tong and Nemunaitis, 2008).

In neurons, Cdk activation occurs prior to the release of cytochrome c, mitochondrial dysfunction, and caspase activation on the path to neuronal death (Endres et al., 1998; Fink et al., 1998; Stefanis et al., 1999; Greene et al., 2007). Cdk inhibition provides long-term rescue from death and prevents release of cytochrome c, loss of mitochondrial transmembrane potential, and prevents caspase activation and processing (Stefanis et al., 1999). In contrast, general caspase inhibitors do not affect early cytochrome c release and do not prevent the loss of mitochondrial transmembrane potential (Stefanis et al., 1999). Moreover, caspase inhibitors protect neurons from this rapid apoptotic death, but do not prevent them from undergoing delayed death in which nuclear features of apoptosis are absent (Stefanis et al., 1999). Thus, cell cycle re-entry (Cdk4 activation, G0/G1 transition) lies upstream of these other pathways producing neuronal death, and mature neurons that re-enter the cell cycle are likely programmed to die not only via caspase-mediated pathways, but also via non-caspase-mediated pathways.

Though no clinical trials of Cdk inhibitors are reported in the treatment of CNS diseases, preclinical experiments demonstrate that Cdk inhibitors (Flavopiridol, Olomoucine, Roscovitine, or Quinazolines) improve behavioral outcomes and increase neuronal survival in a series of CNS disease models such as AD (Copani et al., 2001; Jorda et al., 2003; Verdaguer et al., 2004b; Kruman, 2006; Mark Robert Barvian et al., 2006), PD (Kruman, 2006), stroke (Osuga et al., 2000; Wang et al., 2002), TBI (Di Giovanni et al., 2005; Hilton et al., 2008), spinal cord injury (SCI) (Di Giovanni et al., 2003; Tian et al., 2006), excitotoxic stress (Park et al., 2000; Verdaguer et al., 2003b; Verdaguer et al., 2003a; Verdaguer et al., 2004a) and optic nerve transection (Lefevre et al., 2002) (Table 1). Various side effects may arise because of the non-specificity of those Cdk inhibitors, which also makes it difficult to rule out actions on other molecules (Bain et al., 2003; Sausville, 2003; Blagden and de Bono, 2005; Sridhar et al., 2006; Bain et al., 2007). Fortunately, efforts are underway to develop compounds with increased selectivity for specific Cdks (Hirai et al., 2005; Sridhar et al., 2006).

Table 1.

Agents that interfere with mitogenic signaling pathways and molecules of the “expanded cell cycle”.

Target Agent Stage/CNS Diseases
Tau solvent TRx-0014 Clinical use in AD (TauRx-Therapeutics-Ltd., 2007; Williams, 2009).
Antioxidants Edaravone
NXY-059
Coenzyme Q10
Vitamin E
Melatonin
Trolox
SOD
NAC
PBN
Clinical use in stroke in Asia (Wang and Shuaib, 2007; Abe, 2008)
Clinical trials in stroke (Wang and Shuaib, 2007), AD (Berman and Brodaty, 2004; Vina et al., 2004; Park and Jin, 2008), ALS (Ferrante et al., 2005; Levy et al., 2006; Yoshino and Kimura, 2006).
Experimental trials in ALS (Ito et al., 2008), PD (Yuan et al., 2008), SCI (Tanabe et al., 2009).
Glutamatergic modulators Riluzole
Ceftriaxone
Talampanel
MK-801
NBQX
Clinical use in ALS (Andrews, 2009).
Clinical trials in ALS (Andrews, 2009).
Experimental trials in ICH (Ardizzone et al., 2004), stroke (Meden et al., 1993).
NMDA-receptor modulators Memantine Clinical use in AD (Danysz and Parsons, 2003; Molinuevo et al., 2005; Robinson and Keating, 2006; Iraqi and Hughes, 2009).
Clinical trials in ALS (Andrews, 2009).
Off-label use in psychiatric disorders (Zdanys and Tampi, 2008).
IL-1 inhibitors Kineret
Rytvela
Experimental trials in hypoxic-ischemic newborn brain injury (Quiniou et al., 2008).
TNF-α inhibitors Thalidomide
CNI-1493
Experimental trials in AD (Li et al., 2004; Tweedie et al., 2007); PD (Li et al., 2004; Tweedie et al., 2007); stroke (Meistrell et al., 1997; Nawashiro et al., 1997; Barone and Parsons, 2000; Li et al., 2004; Tweedie et al., 2007).
Thrombin inhibitors Hirudin Experimental trials in ICH (Xue et al., 2006; Sun et al., 2008; Liu et al., 2009a).
Thrombin receptor-1 antagonist BMS-200261 Experimental trials in stroke (Junge et al., 2003); PD (Hamill et al., 2007).
Secretase inhibitors OM 99-1
OM 99-2
BMS 289948
BMS 299897
Experimental trials in AD (Roberds et al., 2001; Ghosh et al., 2002; Selkoe, 2003; Anderson et al., 2005).
i/eNOS inhibitors l-NAME
AMT
NPA
Experimental trials in ICH (Lee da et al., 2006); SCI (Tanabe et al., 2009).
Cox-2 inhibitors Rofecoxib
Celecoxib
NS-398
Clinical trials in AD (Etminan et al., 2003; Warner and Mitchell, 2004).
Experimental trials in ICH (Lee da et al., 2006).
Ras inhibitors Exoenzyme C3
Lovastatin
Experimental trials in motor neuron disorders (Donovan et al., 1997; Smirnova et al., 2001).
Ca2+ channel blockers Flunarizine Clinical trials in stroke (Franke et al., 1996).
Calpain inhibitors MDL-28170 Experimental trials in AD (Jordan et al., 1997; Di Rosa et al., 2002; Battaglia et al., 2003).
Src Inhibitors PP1
PP2
Experimental trials in ICH (Ardizzone et al., 2007; Liu et al., 2008; Liu et al., 2009a); AD (Williamson et al., 2002).
PI3K inhibitors LY 294002 Experimental trials in AD (Zhao et al., 2004).
JAK/Stat Inhibitors EGCG Experimental trials in AD, PD, HIV associated Dementia, multiple sclerosis (MS), ALS, or Pick's Disease (Tan, 2008).
GSK-3β inhibitors Lithium
Kenpaullone
Indirubin
SB 216763
SB 415286
Experimental trials in AD (Cross et al., 2001; Bhat et al., 2004; Jope and Johnson, 2004; Su et al., 2004; Dunn et al., 2005; Pallas and Camins, 2006).
Clinical trials in ALS (Fornai et al., 2008; Andrews, 2009).
ERK1/2 kinase pathway PD98059 Experimental trials in ICH (Fujimoto et al., 2007; Ohnishi et al., 2007).
P38 kinase pathway SB203580
SB239063
Experimental trials in ICH (Fujimoto et al., 2007); PD (Karunakaran et al., 2008); stroke (Barone et al., 2001).
JUN kinase pathway CEP-1347
Colostrinin
SP600125
Experimental trials in ICH (Fujimoto et al., 2007; Ohnishi et al., 2007); AD (Leszek et al., 2002; Wang et al., 2004); stroke (Kuan and Burke, 2005); PD (Wang et al., 2004; Kuan and Burke, 2005).
CDK inhibitors Flavopiridol
Olomoucine
Roscovitine
Quinazolines
Experimental trials in AD (Copani et al., 2001; Jorda et al., 2003; Verdaguer et al., 2004b; Kruman, 2006; Mark Robert Barvian et al., 2006); PD (Kruman, 2006); stroke (Osuga et al., 2000; Wang et al., 2002); TBI (Di Giovanni et al., 2005; Hilton et al., 2008); SCI (Di Giovanni et al., 2003; Tian et al., 2006); excitotoxic stress(Park et al., 2000; Verdaguer et al., 2003b; Verdaguer et al., 2003a; Verdaguer et al., 2004a); optic nerve transection (Lefevre et al., 2002).
NF-κB pathway KINK-1 Experimental trials in stroke (Herrmann et al., 2005).

4-Amino-5-(4-methylphenyl)-7-(t-butyl)pyrazolo(3,4-d)pyrimidine (PP1) 4-Amino-5-(4-chlorophenyl)-7(t-butyl)pyrazol(3,4-d)pyramide (PP2)

2-Amino-5,6-dihydro-6-methyl-4H-1,3-thiazine hydrochloride (AMT)

2,3-Dihydroxy-6-nitro-7-sulfamoyl-benzo[f]quinoxaline-2,3-dione (NBQX)

Epigallocatechin-3-gallate (EGCG)

Methythionium chloride (TRx-0014)

5-Methyl-10,11-dihydro-5H-dibenzo[a,d]cyclohepten-5,10-imine (MK 801)

N-acytyl-L-cysteine (NAC)

N(G)-nitro-l-arginine methyl ester (l-NAME)

N(omega)-propyl-l-arginine (NPA)

Phenyl-N-tert-butylnitrone (PBN)

Superoxide dismutase (SOD)

“Expanded cell cycle” molecules and pathways

Clearer understanding of the mechanisms for neuronal cell cycle re-entry could lead to identification of new pharmacological targets for prevention of cell death and disease progression that lay outside the traditional cell cycle members (Table 1). One approach would be to develop targets based on a broadened or expanded sense of the cell cycle – one which includes not only the cell cycle proteins mentioned above, but also mitogenic molecules and the signaling pathways that interact with them.

Mitogenic molecules can function either as physiological signals or initiators of pathological events depending on their concentrations and activation states (Donovan and Cunningham, 1998; Striggow et al., 2000; D'Autreaux and Toledano, 2007; Chiang et al., 2008; Liu et al., 2008). Increases in the level and activation of these molecules are an indication of increased mitogenic potential, particularly within the injured brain (Akiyama et al., 1992; Hua et al., 2003; Xi et al., 2003; Grammas et al., 2006; van Groen et al., 2006; Xi et al., 2006; Sokolova and Reiser, 2008). This growing list of mitogenic molecules, besides thrombin, Aβ, ROS and NO, noted above, includes excitatory amino acids such as glutamate, various inflammatory cytokines such as interleukin-1 (IL-1), IL-2, IL-6, IL-18, prostaglandin E2 (PG E2), lipopolysaccharide (LPS), tumor necrosis factor-α (TNF-α) and others (Copani et al., 1999; Monfort et al., 2002; Paris et al., 2002; Li et al., 2003; Ishikawa et al., 2006; Caltagarone et al., 2007; Liu et al., 2008; Ojala et al., 2008; Xing et al., 2008). A wide variety of mitogenic molecules are recruited even by a single CNS disease. Each molecule often has a specific ligand-receptor interaction, but may affect multiple downstream signaling pathways (Demoulin and Renauld, 1998; Nathan, 2003; Caltagarone et al., 2007; Sturm et al., 2008; Sun et al., 2008).

The mitogenic signaling of one molecule is often modified or augmented by another. For example, mitochondrial failure results in the release of ROS, which enhance Aβ production. Intracellular Aβ accumulation in turn promotes ROS generation, producing a vicious cycle (Standridge, 2006). The signaling can also be accelerated by one molecule on its own, such as the autocrine cycling of NO, mediated by the inducible enzymes NO/ Ras/ Raf/ MEK-1/ ERK 1, 2/ NF-κB/ eNOS/ NO (Copani and Nicoletti, 2005). These kinds of positive feedback make it possible to elevate molecules abruptly, either as a normal physiological response to disease, or as the cause of disease-induced damage itself.

The actions of mitogenic molecules are both diverse and overlapping, which provides for functional redundancy within mitogenic signaling transduction pathways. As biological cofactors that are enhanced by specific pathological conditions, mitogenic molecules activate specific pathways to mediate cell cycle re-entry and neuronal death. Examples of some mitogenic pathways that overlap and commonly lead to cell cycle re-entry include: (1) FAK/ Src/ Ras/ Raf/ MEK1, 2/ ERK1, 2 → cell cycle re-entry (Copani et al., 1999; Williamson et al., 2002; Wang and Reiser, 2003b; Ohnishi et al., 2007; Varvel et al., 2008); (2) Ras /Rac1/ MEK3, 6/ P38 → cell cycle re-entry (Guan et al., 2005; Segarra et al., 2006); (3) PLC/ IP3/ PKC/ JNK → cell cycle re-entry (Dwivedi and Pandey, 1999; Lopez-Bergami and Ronai, 2008); (4) PI3K/ Akt/ mTOR/ Tau → cell cycle re-entry (Zhu et al., 2000; Khurana et al., 2006; Xing et al., 2008); and (5) JAK/ STAT → cell cycle re-entry (Goody et al., 2007; Tan, 2008). In addition, many molecules, including Ca2+, ROS, NO and PGE2, etc., can directly or indirectly increase the intensity of mitogenic signaling (Fiebich et al., 2001; de Bernardo et al., 2004; Copani and Nicoletti, 2005; Qian et al., 2006; Li et al., 2009; Mao et al., 2009).

MicroRNAs, which are endogenous, non-coding, single-stranded RNA molecules of 19–25 nucleotides in length, have recently attracted attention due in part to the fact that each miRNA can potentially regulate hundreds of genes. It is predicted that over one third of all human genes may be regulated by miRNAs (Bartel, 2004; Esquela-Kerscher and Slack, 2006; Chen and Rajewsky, 2007; Filip, 2007; Guarnieri and DiLeone, 2008). Several miRNAs modulate the major proliferation pathways through direct interaction with transcripts of critical regulators such as Ras, PI3K or ABL, members of the retinoblastoma family, cyclin-Cdk complexes and cell cycle inhibitors of the p27, Ink4 or Cip/Kip families (Gillies and Lorimer, 2007; Bueno et al., 2008). A complex interaction between miRNAs and E2F family members also exists to modulate cell cycle-dependent transcription during cellular proliferation (Brosh et al., 2008).

Agents that interfere with molecules and pathways of the “expanded cell cycle”

In theory any part of the “expanded cell cycle” could be a potential target for drug discovery. For example, an intracerebral hemorrhage would activate thrombin through the coagulation cascade and thrombin would go on to activate src family kinase members (Wang and Reiser, 2003a; Liu et al., 2008). Src family kinases will activate MAPK which will activate cdk4/cyclinD complexes and promote cell cycle re-entry (Wang and Reiser, 2003b; Fujimoto et al., 2007; Liu et al., 2008). Thus, these molecules (thrombin, src kinase, and MAPK), while not considered traditional components of the cell cycle, would all be part of the “expanded cell cycle”. Similarly, other protein kinases (including JAK, Akt, PKC, JNK, ERK, P38, GSK-3β, Cdks, etc.) are also important molecules in the mitogenic pathways leading to neuronal cell cycle re-entry. However, unlike the Cdk-specific inhibitors noted above, many of these kinase inhibitors are currently approved for human use, primarily for the treatment of cancer (Fabian et al., 2005). Since the theory of neuronal cell cycle re-entry was proposed, some of the kinase inhibitors have recently been examined experimentally in the treatment of CNS diseases (Table 1). However, these experiments have been challenging because many kinases play important roles in essential biological processes and many of the kinase inhibitors lack specificity for their targets.

Treatments using antioxidants, NMDA-receptor modulators, cytokine inhibitors, i/eNOS inhibitors, COX-2 inhibitors, and others have often worked fairly well in animal models of brain disease, but have generally failed individually in clinical trials with a few exceptions (Richard Green et al., 2003; Robinson and Keating, 2006; Abe, 2008). Many of these evaluations occurred before cell cycle re-entry was implicated as a mechanism for neuronal death. Even now, their direct effects on the cell-cycle have not been comprehensively studied, and combinations of some of these compounds may be useful for the purpose of cell cycle inhibition experimentally and/or clinically as treatment for CNS diseases.

It is now clear that neurogenesis occurs in the brain of adult mammals (Liu et al., 1998; Brandt and Storch, 2008; Neundorfer, 2008; Shen et al., 2008). This neurogenesis may be associated with maintenance or restoration of neurological function in animal models of CNS diseases, suggesting that neurogenesis is functionally important to recovery (Palmer et al., 2000; Ohab et al., 2006; Abdipranoto et al., 2008; Chesnokova and Pechnick, 2008; Liu et al., 2009a). Neurogenesis arises from brain progenitor cells, rather than from differentiated adult neurons.

Therapies directed at any component inhibiting the cell cycle must be as specific as possible considering cell cycle re-entry contributes to both the death of mature neurons and the genesis of neuroprogenitor cells in adult brain. Therefore, any therapeutics that prevent neuronal death by blocking mitogenic signaling may have limited benefit because they may also prevent neurogenesis. This may provide at least a partial explanation for the questionable efficacy of some currently approved drugs, such as the NMDA receptor modulator Memantine, in the clinical treatment of AD, since NMDA receptor activation has been shown to enhance progenitor cell proliferation and lead to increased neurogenesis (Suzuki et al., 2006). This is consistent with the clinical reports that cognitive dysfunction arises when cell cycle inhibition strategies are used in cancer therapeutics (Dietrich et al., 2008; Konat et al., 2008).

This cognitive dysfunction may also be explained by the fact that current cell cycle inhibition strategies are not cell-specific and also block the proliferation of important brain progenitor cells, thus impairing adult brain neurogenesis. Thus, it appears that cell cycle inhibition strategies could help protect neurons and improve disease and injury outcomes, as long as they do not interfere with the growth of other important cells in the brain. If drugs that block the cell cycle are used to prevent neuronal death in CNS diseases, it is likely that compounds would need to directly (or indirectly) block neuronal cell cycle re-entry and yet not affect the ongoing process of neurogenesis. This will only be possible if the signaling mechanisms are different in adult progenitor cells that divide in the adult brain, versus adult neurons that re-enter the cell cycle.

Conclusions

Cell cycle inhibition protects mature neurons from death. However, it is likely that to truly protect neurons, the best strategy may be to ensure they do not leave the G0 phase at all, since the mere entrance into the initial cell cycle may lead to unavoidable cell death. Moreover, future studies aimed at understanding the respective cell cycle pathways of mature neurons and neuronal progenitors are probably necessary before choosing the best drug targets for treating CNS diseases.

Acknowledgements

This study was supported by NIH grant NS054652 (FRS). We thank Mr. James C. Hathaway for helpful discussion and editing of the manuscript.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Abdipranoto A, Wu S, Stayte S, Vissel B. The role of neurogenesis in neurodegenerative diseases and its implications for therapeutic development. CNS Neurol Disord Drug Targets. 2008;7:187–210. doi: 10.2174/187152708784083858. [DOI] [PubMed] [Google Scholar]
  2. Abe K. Neuroprotective therapy for ischemic stroke with free radical scavenger and gene-stem cell therapy. Rinsho Shinkeigaku. 2008;48:896–898. doi: 10.5692/clinicalneurol.48.896. [DOI] [PubMed] [Google Scholar]
  3. Adams PD. Regulation of the retinoblastoma tumor suppressor protein by cyclin/cdks. Biochim Biophys Acta. 2001;1471:M123–133. doi: 10.1016/s0304-419x(01)00019-1. [DOI] [PubMed] [Google Scholar]
  4. Akiyama H, Ikeda K, Kondo H, McGeer PL. Thrombin accumulation in brains of patients with Alzheimer's disease. Neurosci Lett. 1992;146:152–154. doi: 10.1016/0304-3940(92)90065-f. [DOI] [PubMed] [Google Scholar]
  5. Anderson JJ, Holtz G, Baskin PP, Turner M, Rowe B, Wang B, Kounnas MZ, Lamb BT, Barten D, Felsenstein K, McDonald I, Srinivasan K, Munoz B, Wagner SL. Reductions in beta-amyloid concentrations in vivo by the gamma-secretase inhibitors BMS-289948 and BMS-299897. Biochem Pharmacol. 2005;69:689–698. doi: 10.1016/j.bcp.2004.11.015. [DOI] [PubMed] [Google Scholar]
  6. Andrews J. Amyotrophic lateral sclerosis: clinical management and research update. Curr Neurol Neurosci Rep. 2009;9:59–68. doi: 10.1007/s11910-009-0010-0. [DOI] [PubMed] [Google Scholar]
  7. Ardizzone TD, Zhan X, Ander BP, Sharp FR. SRC kinase inhibition improves acute outcomes after experimental intracerebral hemorrhage. Stroke. 2007;38:1621–1625. doi: 10.1161/STROKEAHA.106.478966. [DOI] [PubMed] [Google Scholar]
  8. Ardizzone TD, Lu A, Wagner KR, Tang Y, Ran R, Sharp FR. Glutamate receptor blockade attenuates glucose hypermetabolism in perihematomal brain after experimental intracerebral hemorrhage in rat. Stroke. 2004;35:2587–2591. doi: 10.1161/01.STR.0000143451.14228.ff. [DOI] [PubMed] [Google Scholar]
  9. Arguello F, Alexander M, Sterry JA, Tudor G, Smith EM, Kalavar NT, Greene JF, Jr., Koss W, Morgan CD, Stinson SF, Siford TJ, Alvord WG, Klabansky RL, Sausville EA. Flavopiridol induces apoptosis of normal lymphoid cells, causes immunosuppression, and has potent antitumor activity In vivo against human leukemia and lymphoma xenografts. Blood. 1998;91:2482–2490. [PubMed] [Google Scholar]
  10. Bain J, McLauchlan H, Elliott M, Cohen P. The specificities of protein kinase inhibitors: an update. Biochem J. 2003;371:199–204. doi: 10.1042/BJ20021535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bain J, Plater L, Elliott M, Shpiro N, Hastie CJ, McLauchlan H, Klevernic I, Arthur JS, Alessi DR, Cohen P. The selectivity of protein kinase inhibitors: a further update. Biochem J. 2007;408:297–315. doi: 10.1042/BJ20070797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Baldin V, Lukas J, Marcote MJ, Pagano M, Draetta G. Cyclin D1 is a nuclear protein required for cell cycle progression in G1. Genes Dev. 1993;7:812–821. doi: 10.1101/gad.7.5.812. [DOI] [PubMed] [Google Scholar]
  13. Barone FC, Parsons AA. Therapeutic potential of anti-inflammatory drugs in focal stroke. Expert Opin Investig Drugs. 2000;9:2281–2306. doi: 10.1517/13543784.9.10.2281. [DOI] [PubMed] [Google Scholar]
  14. Barone FC, Irving EA, Ray AM, Lee JC, Kassis S, Kumar S, Badger AM, Legos JJ, Erhardt JA, Ohlstein EH, Hunter AJ, Harrison DC, Philpott K, Smith BR, Adams JL, Parsons AA. Inhibition of p38 mitogen-activated protein kinase provides neuroprotection in cerebral focal ischemia. Med Res Rev. 2001;21:129–145. doi: 10.1002/1098-1128(200103)21:2<129::aid-med1003>3.0.co;2-h. [DOI] [PubMed] [Google Scholar]
  15. Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116:281–297. doi: 10.1016/s0092-8674(04)00045-5. [DOI] [PubMed] [Google Scholar]
  16. Bartkova J, Zemanova M, Bartek J. Expression of CDK7/CAK in normal and tumor cells of diverse histogenesis, cell-cycle position and differentiation. Int J Cancer. 1996;66:732–737. doi: 10.1002/(SICI)1097-0215(19960611)66:6<732::AID-IJC4>3.0.CO;2-0. [DOI] [PubMed] [Google Scholar]
  17. Battaglia F, Trinchese F, Liu S, Walter S, Nixon RA, Arancio O. Calpain inhibitors, a treatment for Alzheimer's disease: position paper. J Mol Neurosci. 2003;20:357–362. doi: 10.1385/JMN:20:3:357. [DOI] [PubMed] [Google Scholar]
  18. Berman K, Brodaty H. Tocopherol (vitamin E) in Alzheimer's disease and other neurodegenerative disorders. CNS Drugs. 2004;18:807–825. doi: 10.2165/00023210-200418120-00005. [DOI] [PubMed] [Google Scholar]
  19. Bhat RV, Budd Haeberlein SL, Avila J. Glycogen synthase kinase 3: a drug target for CNS therapies. J Neurochem. 2004;89:1313–1317. doi: 10.1111/j.1471-4159.2004.02422.x. [DOI] [PubMed] [Google Scholar]
  20. Blagden S, de Bono J. Drugging cell cycle kinases in cancer therapy. Curr Drug Targets. 2005;6:325–335. doi: 10.2174/1389450053765824. [DOI] [PubMed] [Google Scholar]
  21. Brandt MD, Storch A. Neurogenesis in the adult brain: from bench to bedside? Fortschr Neurol Psychiatr. 2008;76:517–529. doi: 10.1055/s-2008-1038218. [DOI] [PubMed] [Google Scholar]
  22. Brasnjevic I, Hof PR, Steinbusch HW, Schmitz C. Accumulation of nuclear DNA damage or neuron loss: molecular basis for a new approach to understanding selective neuronal vulnerability in neurodegenerative diseases. DNA Repair (Amst) 2008;7:1087–1097. doi: 10.1016/j.dnarep.2008.03.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Brosh R, Shalgi R, Liran A, Landan G, Korotayev K, Nguyen GH, Enerly E, Johnsen H, Buganim Y, Solomon H, Goldstein I, Madar S, Goldfinger N, Borresen-Dale AL, Ginsberg D, Harris CC, Pilpel Y, Oren M, Rotter V. p53-Repressed miRNAs are involved with E2F in a feed-forward loop promoting proliferation. Mol Syst Biol. 2008;4:229. doi: 10.1038/msb.2008.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Bueno MJ, de Castro IP, Malumbres M. Control of cell proliferation pathways by microRNAs. Cell Cycle. 2008;7:3143–3148. doi: 10.4161/cc.7.20.6833. [DOI] [PubMed] [Google Scholar]
  25. Busser J, Geldmacher DS, Herrup K. Ectopic cell cycle proteins predict the sites of neuronal cell death in Alzheimer's disease brain. J Neurosci. 1998;18:2801–2807. doi: 10.1523/JNEUROSCI.18-08-02801.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Caltagarone J, Jing Z, Bowser R. Focal adhesions regulate Abeta signaling and cell death in Alzheimer's disease. Biochim Biophys Acta. 2007;1772:438–445. doi: 10.1016/j.bbadis.2006.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Carnero A, Hudson JD, Price CM, Beach DH. p16INK4A and p19ARF act in overlapping pathways in cellular immortalization. Nat Cell Biol. 2000;2:148–155. doi: 10.1038/35004020. [DOI] [PubMed] [Google Scholar]
  28. Chen K, Rajewsky N. The evolution of gene regulation by transcription factors and microRNAs. Nat Rev Genet. 2007;8:93–103. doi: 10.1038/nrg1990. [DOI] [PubMed] [Google Scholar]
  29. Chesnokova V, Pechnick RN. Antidepressants and Cdk inhibitors: releasing the brake on neurogenesis? Cell Cycle. 2008;7:2321–2326. doi: 10.4161/cc.6446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chiang PK, Lam MA, Luo Y. The many faces of amyloid beta in Alzheimer's disease. Curr Mol Med. 2008;8:580–584. doi: 10.2174/156652408785747951. [DOI] [PubMed] [Google Scholar]
  31. Chien M, Astumian M, Liebowitz D, Rinker-Schaeffer C, Stadler WM. In vitro evaluation of flavopiridol, a novel cell cycle inhibitor, in bladder cancer. Cancer Chemother Pharmacol. 1999;44:81–87. doi: 10.1007/s002800050948. [DOI] [PubMed] [Google Scholar]
  32. Copani A, Nicoletti F. Cell-cycle mechanisms and neuronal cell death. Kluwer Academic/Plenum; New York: 2005. [Google Scholar]
  33. Copani A, Uberti D, Sortino MA, Bruno V, Nicoletti F, Memo M. Activation of cell-cycle-associated proteins in neuronal death: a mandatory or dispensable path? Trends Neurosci. 2001;24:25–31. doi: 10.1016/s0166-2236(00)01663-5. [DOI] [PubMed] [Google Scholar]
  34. Copani A, Condorelli F, Caruso A, Vancheri C, Sala A, Giuffrida Stella AM, Canonico PL, Nicoletti F, Sortino MA. Mitotic signaling by beta-amyloid causes neuronal death. Faseb J. 1999;13:2225–2234. [PubMed] [Google Scholar]
  35. Copani A, Hoozemans JJ, Caraci F, Calafiore M, Van Haastert ES, Veerhuis R, Rozemuller AJ, Aronica E, Sortino MA, Nicoletti F. DNA polymerase-beta is expressed early in neurons of Alzheimer's disease brain and is loaded into DNA replication forks in neurons challenged with beta-amyloid. J Neurosci. 2006;26:10949–10957. doi: 10.1523/JNEUROSCI.2793-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Cross DA, Culbert AA, Chalmers KA, Facci L, Skaper SD, Reith AD. Selective small-molecule inhibitors of glycogen synthase kinase-3 activity protect primary neurones from death. J Neurochem. 2001;77:94–102. doi: 10.1046/j.1471-4159.2001.t01-1-00251.x. [DOI] [PubMed] [Google Scholar]
  37. D'Autreaux B, Toledano MB. ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nat Rev Mol Cell Biol. 2007;8:813–824. doi: 10.1038/nrm2256. [DOI] [PubMed] [Google Scholar]
  38. Danysz W, Parsons CG. The NMDA receptor antagonist memantine as a symptomatological and neuroprotective treatment for Alzheimer's disease: preclinical evidence. Int J Geriatr Psychiatry. 2003;18:S23–32. doi: 10.1002/gps.938. [DOI] [PubMed] [Google Scholar]
  39. de Bernardo S, Canals S, Casarejos MJ, Solano RM, Menendez J, Mena MA. Role of extracellular signal-regulated protein kinase in neuronal cell death induced by glutathione depletion in neuron/glia mesencephalic cultures. J Neurochem. 2004;91:667–682. doi: 10.1111/j.1471-4159.2004.02744.x. [DOI] [PubMed] [Google Scholar]
  40. Demoulin JB, Renauld JC. Signalling by cytokines interacting with the interleukin-2 receptor gamma chain. Cytokines Cell Mol Ther. 1998;4:243–256. [PubMed] [Google Scholar]
  41. Di Giovanni S, Knoblach SM, Brandoli C, Aden SA, Hoffman EP, Faden AI. Gene profiling in spinal cord injury shows role of cell cycle in neuronal death. Ann Neurol. 2003;53:454–468. doi: 10.1002/ana.10472. [DOI] [PubMed] [Google Scholar]
  42. Di Giovanni S, Movsesyan V, Ahmed F, Cernak I, Schinelli S, Stoica B, Faden AI. Cell cycle inhibition provides neuroprotection and reduces glial proliferation and scar formation after traumatic brain injury. Proc Natl Acad Sci U S A. 2005;102:8333–8338. doi: 10.1073/pnas.0500989102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Di Rosa G, Odrijin T, Nixon RA, Arancio O. Calpain inhibitors: a treatment for Alzheimer's disease. J Mol Neurosci. 2002;19:135–141. doi: 10.1007/s12031-002-0024-4. [DOI] [PubMed] [Google Scholar]
  44. Dietrich J, Monje M, Wefel J, Meyers C. Clinical patterns and biological correlates of cognitive dysfunction associated with cancer therapy. Oncologist. 2008;13:1285–1295. doi: 10.1634/theoncologist.2008-0130. [DOI] [PubMed] [Google Scholar]
  45. Dirks PB, Rutka JT. Current concepts in neuro-oncology: the cell cycle--a review. Neurosurgery. 1997;40:1000–1013. doi: 10.1097/00006123-199705000-00025. discussion 1013–1005. [DOI] [PubMed] [Google Scholar]
  46. Donovan FM, Cunningham DD. Signaling pathways involved in thrombin-induced cell protection. J Biol Chem. 1998;273:12746–12752. doi: 10.1074/jbc.273.21.12746. [DOI] [PubMed] [Google Scholar]
  47. Donovan FM, Pike CJ, Cotman CW, Cunningham DD. Thrombin induces apoptosis in cultured neurons and astrocytes via a pathway requiring tyrosine kinase and RhoA activities. J Neurosci. 1997;17:5316–5326. doi: 10.1523/JNEUROSCI.17-14-05316.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Drees M, Dengler WA, Roth T, Labonte H, Mayo J, Malspeis L, Grever M, Sausville EA, Fiebig HH. Flavopiridol (L86-8275): selective antitumor activity in vitro and activity in vivo for prostate carcinoma cells. Clin Cancer Res. 1997;3:273–279. [PubMed] [Google Scholar]
  49. Dunn N, Holmes C, Mullee M. Does lithium therapy protect against the onset of dementia? Alzheimer Dis Assoc Disord. 2005;19:20–22. doi: 10.1097/01.wad.0000155068.23937.9b. [DOI] [PubMed] [Google Scholar]
  50. Dwivedi Y, Pandey GN. Effects of treatment with haloperidol, chlorpromazine, and clozapine on protein kinase C (PKC) and phosphoinositide-specific phospholipase C (PI-PLC) activity and on mRNA and protein expression of PKC and PLC isozymes in rat brain. J Pharmacol Exp Ther. 1999;291:688–704. [PubMed] [Google Scholar]
  51. Endres M, Namura S, Shimizu-Sasamata M, Waeber C, Zhang L, Gomez-Isla T, Hyman BT, Moskowitz MA. Attenuation of delayed neuronal death after mild focal ischemia in mice by inhibition of the caspase family. J Cereb Blood Flow Metab. 1998;18:238–247. doi: 10.1097/00004647-199803000-00002. [DOI] [PubMed] [Google Scholar]
  52. Erickson S, Sangfelt O, Heyman M, Castro J, Einhorn S, Grander D. Involvement of the Ink4 proteins p16 and p15 in T-lymphocyte senescence. Oncogene. 1998;17:595–602. doi: 10.1038/sj.onc.1201965. [DOI] [PubMed] [Google Scholar]
  53. Esquela-Kerscher A, Slack FJ. Oncomirs - microRNAs with a role in cancer. Nat Rev Cancer. 2006;6:259–269. doi: 10.1038/nrc1840. [DOI] [PubMed] [Google Scholar]
  54. Etminan M, Gill S, Samii A. Effect of non-steroidal anti-inflammatory drugs on risk of Alzheimer's disease: systematic review and meta-analysis of observational studies. Bmj. 2003;327:128. doi: 10.1136/bmj.327.7407.128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Ezhevsky SA, Ho A, Becker-Hapak M, Davis PK, Dowdy SF. Differential regulation of retinoblastoma tumor suppressor protein by G(1) cyclin-dependent kinase complexes in vivo. Mol Cell Biol. 2001;21:4773–4784. doi: 10.1128/MCB.21.14.4773-4784.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Fabian MA, et al. A small molecule-kinase interaction map for clinical kinase inhibitors. Nat Biotechnol. 2005;23:329–336. doi: 10.1038/nbt1068. [DOI] [PubMed] [Google Scholar]
  57. Ferrante KL, et al. Tolerance of high-dose (3,000 mg/day) coenzyme Q10 in ALS. Neurology. 2005;65:1834–1836. doi: 10.1212/01.wnl.0000187070.35365.d7. [DOI] [PubMed] [Google Scholar]
  58. Fiebich BL, Schleicher S, Spleiss O, Czygan M, Hull M. Mechanisms of prostaglandin E2-induced interleukin-6 release in astrocytes: possible involvement of EP4-like receptors, p38 mitogen-activated protein kinase and protein kinase C. J Neurochem. 2001;79:950–958. doi: 10.1046/j.1471-4159.2001.00652.x. [DOI] [PubMed] [Google Scholar]
  59. Filip A. MiRNA--new mechanisms of gene expression control. Postepy Biochem. 2007;53:413–419. [PubMed] [Google Scholar]
  60. Fink K, Zhu J, Namura S, Shimizu-Sasamata M, Endres M, Ma J, Dalkara T, Yuan J, Moskowitz MA. Prolonged therapeutic window for ischemic brain damage caused by delayed caspase activation. J Cereb Blood Flow Metab. 1998;18:1071–1076. doi: 10.1097/00004647-199810000-00003. [DOI] [PubMed] [Google Scholar]
  61. Fornai F, Longone P, Cafaro L, Kastsiuchenka O, Ferrucci M, Manca ML, Lazzeri G, Spalloni A, Bellio N, Lenzi P, Modugno N, Siciliano G, Isidoro C, Murri L, Ruggieri S, Paparelli A. Lithium delays progression of amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A. 2008;105:2052–2057. doi: 10.1073/pnas.0708022105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Frank CL, Tsai LH. Alternative functions of core cell cycle regulators in neuronal migration, neuronal maturation, and synaptic plasticity. Neuron. 2009;62:312–326. doi: 10.1016/j.neuron.2009.03.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Franke CL, Palm R, Dalby M, Schoonderwaldt HC, Hantson L, Eriksson B, Lang-Jenssen L, Smakman J. Flunarizine in stroke treatment (FIST): a double-blind, placebo-controlled trial in Scandinavia and the Netherlands. Acta Neurol Scand. 1996;93:56–60. doi: 10.1111/j.1600-0404.1996.tb00171.x. [DOI] [PubMed] [Google Scholar]
  64. Fujimoto S, Katsuki H, Ohnishi M, Takagi M, Kume T, Akaike A. Thrombin induces striatal neurotoxicity depending on mitogen-activated protein kinase pathways in vivo. Neuroscience. 2007;144:694–701. doi: 10.1016/j.neuroscience.2006.09.049. [DOI] [PubMed] [Google Scholar]
  65. Ghosh AK, Hong L, Tang J. Beta-secretase as a therapeutic target for inhibitor drugs. Curr Med Chem. 2002;9:1135–1144. doi: 10.2174/0929867023370149. [DOI] [PubMed] [Google Scholar]
  66. Gillies JK, Lorimer IA. Regulation of p27Kip1 by miRNA 221/222 in glioblastoma. Cell Cycle. 2007;6:2005–2009. doi: 10.4161/cc.6.16.4526. [DOI] [PubMed] [Google Scholar]
  67. Goody RJ, Beckham JD, Rubtsova K, Tyler KL. JAK-STAT signaling pathways are activated in the brain following reovirus infection. J Neurovirol. 2007;13:373–383. doi: 10.1080/13550280701344983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Grammas P, Samany PG, Thirumangalakudi L. Thrombin and inflammatory proteins are elevated in Alzheimer's disease microvessels: implications for disease pathogenesis. J Alzheimers Dis. 2006;9:51–58. doi: 10.3233/jad-2006-9105. [DOI] [PubMed] [Google Scholar]
  69. Greene LA, Liu DX, Troy CM, Biswas SC. Cell cycle molecules define a pathway required for neuron death in development and disease. Biochim Biophys Acta. 2007;1772:392–401. doi: 10.1016/j.bbadis.2006.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Guan J, Luo Y, Denker BM. Purkinje cell protein-2 (Pcp2) stimulates differentiation in PC12 cells by Gbetagamma-mediated activation of Ras and p38 MAPK. Biochem J. 2005;392:389–397. doi: 10.1042/BJ20042102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Guarnieri DJ, DiLeone RJ. MicroRNAs: a new class of gene regulators. Ann Med. 2008;40:197–208. doi: 10.1080/07853890701771823. [DOI] [PubMed] [Google Scholar]
  72. Guegan C, Levy V, David JP, Ajchenbaum-Cymbalista F, Sola B. c-Jun and cyclin D1 proteins as mediators of neuronal death after a focal ischaemic insult. Neuroreport. 1997;8:1003–1007. doi: 10.1097/00001756-199703030-00037. [DOI] [PubMed] [Google Scholar]
  73. Hamill CE, Caudle WM, Richardson JR, Yuan H, Pennell KD, Greene JG, Miller GW, Traynelis SF. Exacerbation of dopaminergic terminal damage in a mouse model of Parkinson's disease by the G-protein-coupled receptor protease-activated receptor 1. Mol Pharmacol. 2007;72:653–664. doi: 10.1124/mol.107.038158. [DOI] [PubMed] [Google Scholar]
  74. Harbour JW, Luo RX, Dei Santi A, Postigo AA, Dean DC. Cdk phosphorylation triggers sequential intramolecular interactions that progressively block Rb functions as cells move through G1. Cell. 1999;98:859–869. doi: 10.1016/s0092-8674(00)81519-6. [DOI] [PubMed] [Google Scholar]
  75. Herrmann O, Baumann B, de Lorenzi R, Muhammad S, Zhang W, Kleesiek J, Malfertheiner M, Kohrmann M, Potrovita I, Maegele I, Beyer C, Burke JR, Hasan MT, Bujard H, Wirth T, Pasparakis M, Schwaninger M. IKK mediates ischemia-induced neuronal death. Nat Med. 2005;11:1322–1329. doi: 10.1038/nm1323. [DOI] [PubMed] [Google Scholar]
  76. Herrup K, Neve R, Ackerman SL, Copani A. Divide and die: cell cycle events as triggers of nerve cell death. J Neurosci. 2004;24:9232–9239. doi: 10.1523/JNEUROSCI.3347-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. 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]
  78. Hirai H, Kawanishi N, Iwasawa Y. Recent advances in the development of selective small molecule inhibitors for cyclin-dependent kinases. Curr Top Med Chem. 2005;5:167–179. doi: 10.2174/1568026053507688. [DOI] [PubMed] [Google Scholar]
  79. Hua Y, Wu J, Keep RF, Hoff JT, Xi G. Thrombin exacerbates brain edema in focal cerebral ischemia. Acta Neurochir Suppl. 2003;86:163–166. doi: 10.1007/978-3-7091-0651-8_34. [DOI] [PubMed] [Google Scholar]
  80. Imai H, Harland J, McCulloch J, Graham DI, Brown SM, Macrae IM. Specific expression of the cell cycle regulation proteins, GADD34 and PCNA, in the peri-infarct zone after focal cerebral ischaemia in the rat. Eur J Neurosci. 2002;15:1929–1936. doi: 10.1046/j.1460-9568.2002.02025.x. [DOI] [PubMed] [Google Scholar]
  81. Iraqi A, Hughes TL. Nightmares and memantine: a case report and review of literature. J Am Med Dir Assoc. 2009;10:77–78. doi: 10.1016/j.jamda.2008.10.001. [DOI] [PubMed] [Google Scholar]
  82. Ishikawa H, Tsuyama N, Obata M, M MK. Mitogenic signals initiated via interleukin-6 receptor complexes in cooperation with other transmembrane molecules in myelomas. J Clin Exp Hematop. 2006;46:55–66. doi: 10.3960/jslrt.46.55. [DOI] [PubMed] [Google Scholar]
  83. Ishikawa M, Ogihara Y, Miura M. Visualization of radiation-induced cell cycle-associated events in tumor cells expressing the fusion protein of Azami Green and the destruction box of human Geminin. Biochem Biophys Res Commun. 2009 doi: 10.1016/j.bbrc.2009.08.160. [DOI] [PubMed] [Google Scholar]
  84. Ito H, Wate R, Zhang J, Ohnishi S, Kaneko S, Ito H, Nakano S, Kusaka H. Treatment with edaravone, initiated at symptom onset, slows motor decline and decreases SOD1 deposition in ALS mice. Exp Neurol. 2008;213:448–455. doi: 10.1016/j.expneurol.2008.07.017. [DOI] [PubMed] [Google Scholar]
  85. Jope RS, Johnson GV. The glamour and gloom of glycogen synthase kinase-3. Trends Biochem Sci. 2004;29:95–102. doi: 10.1016/j.tibs.2003.12.004. [DOI] [PubMed] [Google Scholar]
  86. Jorda EG, Verdaguer E, Canudas AM, Jimenez A, Bruna A, Caelles C, Bravo R, Escubedo E, Pubill D, Camarasa J, Pallas M, Camins A. Neuroprotective action of flavopiridol, a cyclin-dependent kinase inhibitor, in colchicine-induced apoptosis. Neuropharmacology. 2003;45:672–683. doi: 10.1016/s0028-3908(03)00204-1. [DOI] [PubMed] [Google Scholar]
  87. Jordan-Sciutto KL, Dorsey R, Chalovich EM, Hammond RR, Achim CL. Expression patterns of retinoblastoma protein in Parkinson disease. J Neuropathol Exp Neurol. 2003;62:68–74. doi: 10.1093/jnen/62.1.68. [DOI] [PubMed] [Google Scholar]
  88. Jordan J, Galindo MF, Miller RJ. Role of calpain- and interleukin-1 beta converting enzyme-like proteases in the beta-amyloid-induced death of rat hippocampal neurons in culture. J Neurochem. 1997;68:1612–1621. doi: 10.1046/j.1471-4159.1997.68041612.x. [DOI] [PubMed] [Google Scholar]
  89. Junge CE, Sugawara T, Mannaioni G, Alagarsamy S, Conn PJ, Brat DJ, Chan PH, Traynelis SF. The contribution of protease-activated receptor 1 to neuronal damage caused by transient focal cerebral ischemia. Proc Natl Acad Sci U S A. 2003;100:13019–13024. doi: 10.1073/pnas.2235594100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Karunakaran S, Saeed U, Mishra M, Valli RK, Joshi SD, Meka DP, Seth P, Ravindranath V. 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]
  91. Katchanov J, Harms C, Gertz K, Hauck L, Waeber C, Hirt L, Priller J, von Harsdorf R, Bruck W, Hortnagl H, Dirnagl U, Bhide PG, Endres M. Mild cerebral ischemia induces loss of cyclin-dependent kinase inhibitors and activation of cell cycle machinery before delayed neuronal cell death. J Neurosci. 2001;21:5045–5053. doi: 10.1523/JNEUROSCI.21-14-05045.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Khurana V, Lu Y, Steinhilb ML, Oldham S, Shulman JM, Feany MB. TOR-mediated cell-cycle activation causes neurodegeneration in a Drosophila tauopathy model. Curr Biol. 2006;16:230–241. doi: 10.1016/j.cub.2005.12.042. [DOI] [PubMed] [Google Scholar]
  93. Kim AH, Puram SV, Bilimoria PM, Ikeuchi Y, Keough S, Wong M, Rowitch D, Bonni A. A centrosomal Cdc20-APC pathway controls dendrite morphogenesis in postmitotic neurons. Cell. 2009;136:322–336. doi: 10.1016/j.cell.2008.11.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. King RW, Jackson PK, Kirschner MW. Mitosis in transition. Cell. 1994;79:563–571. doi: 10.1016/0092-8674(94)90542-8. [DOI] [PubMed] [Google Scholar]
  95. Konat GW, Kraszpulski M, James I, Zhang HT, Abraham J. Cognitive dysfunction induced by chronic administration of common cancer chemotherapeutics in rats. Metab Brain Dis. 2008;23:325–333. doi: 10.1007/s11011-008-9100-y. [DOI] [PubMed] [Google Scholar]
  96. Kruman I, Schwartz E. Methods of neuroprotection by cyclin-dependent kinase inhibition. US 20080182853. 2006 [Google Scholar]
  97. Kuan CY, Burke RE. Targeting the JNK signaling pathway for stroke and Parkinson's diseases therapy. Curr Drug Targets CNS Neurol Disord. 2005;4:63–67. doi: 10.2174/1568007053005145. [DOI] [PubMed] [Google Scholar]
  98. Kuan CY, Schloemer AJ, Lu A, Burns KA, Weng WL, Williams MT, Strauss KI, Vorhees CV, Flavell RA, Davis RJ, Sharp FR, Rakic P. Hypoxia-ischemia induces DNA synthesis without cell proliferation in dying neurons in adult rodent brain. J Neurosci. 2004;24:10763–10772. doi: 10.1523/JNEUROSCI.3883-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Lee da Y, Park KW, Jin BK. Thrombin induces neurodegeneration and microglial activation in the cortex in vivo and in vitro: proteolytic and non-proteolytic actions. Biochem Biophys Res Commun. 2006;346:727–738. doi: 10.1016/j.bbrc.2006.05.174. [DOI] [PubMed] [Google Scholar]
  100. Lefevre K, Clarke PG, Danthe EE, Castagne V. Involvement of cyclin-dependent kinases in axotomy-induced retinal ganglion cell death. J Comp Neurol. 2002;447:72–81. doi: 10.1002/cne.10215. [DOI] [PubMed] [Google Scholar]
  101. Leszek J, Inglot AD, Janusz M, Byczkiewicz F, Kiejna A, Georgiades J, Lisowski J. Colostrinin proline-rich polypeptide complex from ovine colostrum--a long-term study of its efficacy in Alzheimer's disease. Med Sci Monit. 2002;8:PI93–96. [PubMed] [Google Scholar]
  102. Levy G, Kaufmann P, Buchsbaum R, Montes J, Barsdorf A, Arbing R, Battista V, Zhou X, Mitsumoto H, Levin B, Thompson JL. A two-stage design for a phase II clinical trial of coenzyme Q10 in ALS. Neurology. 2006;66:660–663. doi: 10.1212/01.wnl.0000201182.60750.66. [DOI] [PubMed] [Google Scholar]
  103. Li N, Wang C, Wu Y, Liu X, Cao X. Ca2+/Calmodulin-dependent Protein Kinase II Promotes Cell Cycle Progression by Directly Activating MEK1 and Subsequently Modulating p27 Phosphorylation. J Biol Chem. 2009;284:3021–3027. doi: 10.1074/jbc.M805483200. [DOI] [PubMed] [Google Scholar]
  104. Li R, Yang L, Lindholm K, Konishi Y, Yue X, Hampel H, Zhang D, Shen Y. Tumor necrosis factor death receptor signaling cascade is required for amyloid-beta protein-induced neuron death. J Neurosci. 2004;24:1760–1771. doi: 10.1523/JNEUROSCI.4580-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Li Y, Chopp M, Powers C, Jiang N. Immunoreactivity of cyclin D1/cdk4 in neurons and oligodendrocytes after focal cerebral ischemia in rat. J Cereb Blood Flow Metab. 1997;17:846–856. doi: 10.1097/00004647-199708000-00003. [DOI] [PubMed] [Google Scholar]
  106. Li Y, Liu L, Barger SW, Griffin WS. Interleukin-1 mediates pathological effects of microglia on tau phosphorylation and on synaptophysin synthesis in cortical neurons through a p38-MAPK pathway. J Neurosci. 2003;23:1605–1611. doi: 10.1523/JNEUROSCI.23-05-01605.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Liu DZ, Cheng XY, Ander BP, Xu H, Davis RR, Gregg JP, Sharp FR. 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]
  108. Liu DZ, Ander BP, Xu H, Shen Y, Kaur P, Deng W, Sharp FR. Blood brain barrier breakdown and repair after thrombin-induced brain injury. Annals of Neurology. 2009a doi: 10.1002/ana.21924. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Liu DZ, Tian Y, Ander BP, Xu H, Stamova B, Zhan X, Turner RJ, Jickling G, Sharp FR. Brain and blood microRNA expression profiling of ischemic stroke, intracerebral hemorrhage and kainate seizures. Journal of Cerebral Blood Flow & Metabolism. 2009b doi: 10.1038/jcbfm.2009.186. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Liu J, Solway K, Messing RO, Sharp FR. Increased neurogenesis in the dentate gyrus after transient global ischemia in gerbils. J Neurosci. 1998;18:7768–7778. doi: 10.1523/JNEUROSCI.18-19-07768.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Lopez-Bergami P, Ronai Z. Requirements for PKC-augmented JNK activation by MKK4/7. Int J Biochem Cell Biol. 2008;40:1055–1064. doi: 10.1016/j.biocel.2007.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Lundberg AS, Weinberg RA. Functional inactivation of the retinoblastoma protein requires sequential modification by at least two distinct cyclin-cdk complexes. Mol Cell Biol. 1998;18:753–761. doi: 10.1128/mcb.18.2.753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Malumbres M, Barbacid M. To cycle or not to cycle: a critical decision in cancer. Nat Rev Cancer. 2001;1:222–231. doi: 10.1038/35106065. [DOI] [PubMed] [Google Scholar]
  114. Mao LM, Tang QS, Wang JQ. Regulation of extracellular signal-regulated kinase phosphorylation in cultured rat striatal neurons. Brain Res Bull. 2009;78:328–334. doi: 10.1016/j.brainresbull.2008.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Barvian Mark Robert, Dobrusin Ellen Myra, Kaltenbronn James Stanley, Toogood Peter L., Winters Roy Thomas, Sidhu Inderjit S., Singh Rajeshwar, Bathini Yadagiri, Micetich RG. Quinazolines and their use for inhibiting cyclin-dependent kinase enzymes. US 6982260. 2006 [Google Scholar]
  116. McShea A, Harris PL, Webster KR, Wahl AF, Smith MA. Abnormal expression of the cell cycle regulators P16 and CDK4 in Alzheimer's disease. Am J Pathol. 1997;150:1933–1939. [PMC free article] [PubMed] [Google Scholar]
  117. Meden P, Overgaard K, Sereghy T, Boysen G. Enhancing the efficacy of thrombolysis by AMPA receptor blockade with NBQX in a rat embolic stroke model. J Neurol Sci. 1993;119:209–216. doi: 10.1016/0022-510x(93)90136-m. [DOI] [PubMed] [Google Scholar]
  118. Meistrell ME, 3rd, Botchkina GI, Wang H, Di Santo E, Cockroft KM, Bloom O, Vishnubhakat JM, Ghezzi P, Tracey KJ. Tumor necrosis factor is a brain damaging cytokine in cerebral ischemia. Shock. 1997;8:341–348. [PubMed] [Google Scholar]
  119. Molinuevo JL, Llado A, Rami L. Memantine: targeting glutamate excitotoxicity in Alzheimer's disease and other dementias. Am J Alzheimers Dis Other Demen. 2005;20:77–85. doi: 10.1177/153331750502000206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Monfort P, Kosenko E, Erceg S, Canales JJ, Felipo V. Molecular mechanism of acute ammonia toxicity: role of NMDA receptors. Neurochem Int. 2002;41:95–102. doi: 10.1016/s0197-0186(02)00029-3. [DOI] [PubMed] [Google Scholar]
  121. Morgan DO. Cyclin-dependent kinases: engines, clocks, and microprocessors. Annu Rev Cell Dev Biol. 1997;13:261–291. doi: 10.1146/annurev.cellbio.13.1.261. [DOI] [PubMed] [Google Scholar]
  122. Nagata M, Tomari S, Kanemoto K, Usui J, Lemley KV. Podocytes, parietal cells, and glomerular pathology: the role of cell cycle proteins. Pediatr Nephrol. 2003;18:3–8. doi: 10.1007/s00467-002-0995-y. [DOI] [PubMed] [Google Scholar]
  123. Nagy Z, Esiri MM. Neuronal cyclin expression in the hippocampus in temporal lobe epilepsy. Exp Neurol. 1998;150:240–247. doi: 10.1006/exnr.1997.6753. [DOI] [PubMed] [Google Scholar]
  124. Nagy Z, Esiri MM, Cato AM, Smith AD. Cell cycle markers in the hippocampus in Alzheimer's disease. Acta Neuropathol. 1997;94:6–15. doi: 10.1007/s004010050665. [DOI] [PubMed] [Google Scholar]
  125. Nathan C. Specificity of a third kind: reactive oxygen and nitrogen intermediates in cell signaling. J Clin Invest. 2003;111:769–778. doi: 10.1172/JCI18174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Nawashiro H, Martin D, Hallenbeck JM. Inhibition of tumor necrosis factor and amelioration of brain infarction in mice. J Cereb Blood Flow Metab. 1997;17:229–232. doi: 10.1097/00004647-199702000-00013. [DOI] [PubMed] [Google Scholar]
  127. Neundorfer B. Does the neurogenesis in the adult brain show the way into the future? Fortschr Neurol Psychiatr. 2008;76:511. doi: 10.1055/s-2008-1038252. [DOI] [PubMed] [Google Scholar]
  128. Nguyen MD, Boudreau M, Kriz J, Couillard-Despres S, Kaplan DR, Julien JP. Cell cycle regulators in the neuronal death pathway of amyotrophic lateral sclerosis caused by mutant superoxide dismutase 1. J Neurosci. 2003;23:2131–2140. doi: 10.1523/JNEUROSCI.23-06-02131.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Norbury C, Nurse P. Animal cell cycles and their control. Annu Rev Biochem. 1992;61:441–470. doi: 10.1146/annurev.bi.61.070192.002301. [DOI] [PubMed] [Google Scholar]
  130. Novak B, Tyson JJ, Gyorffy B, Csikasz-Nagy A. Irreversible cell-cycle transitions are due to systems-level feedback. Nat Cell Biol. 2007;9:724–728. doi: 10.1038/ncb0707-724. [DOI] [PubMed] [Google Scholar]
  131. O'Hare M, Wang F, Park DS. Cyclin-dependent kinases as potential targets to improve stroke outcome. Pharmacol Ther. 2002;93:135–143. doi: 10.1016/s0163-7258(02)00183-3. [DOI] [PubMed] [Google Scholar]
  132. Ohab JJ, Fleming S, Blesch A, Carmichael ST. A neurovascular niche for neurogenesis after stroke. J Neurosci. 2006;26:13007–13016. doi: 10.1523/JNEUROSCI.4323-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Ohnishi M, Katsuki H, Fujimoto S, Takagi M, Kume T, Akaike A. Involvement of thrombin and mitogen-activated protein kinase pathways in hemorrhagic brain injury. Exp Neurol. 2007 doi: 10.1016/j.expneurol.2007.03.030. [DOI] [PubMed] [Google Scholar]
  134. Ohtsubo M, Roberts JM. Cyclin-dependent regulation of G1 in mammalian fibroblasts. Science. 1993;259:1908–1912. doi: 10.1126/science.8384376. [DOI] [PubMed] [Google Scholar]
  135. Ohtsubo M, Theodoras AM, Schumacher J, Roberts JM, Pagano M. Human cyclin E, a nuclear protein essential for the G1-to-S phase transition. Mol Cell Biol. 1995;15:2612–2624. doi: 10.1128/mcb.15.5.2612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Ojala JO, Sutinen EM, Salminen A, Pirttila T. Interleukin-18 increases expression of kinases involved in tau phosphorylation in SH-SY5Y neuroblastoma cells. J Neuroimmunol. 2008;205:86–93. doi: 10.1016/j.jneuroim.2008.09.012. [DOI] [PubMed] [Google Scholar]
  137. Osuga H, Osuga S, Wang F, Fetni R, Hogan MJ, Slack RS, Hakim AM, Ikeda JE, Park DS. Cyclin-dependent kinases as a therapeutic target for stroke. Proc Natl Acad Sci U S A. 2000;97:10254–10259. doi: 10.1073/pnas.170144197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Pallas M, Camins A. Molecular and biochemical features in Alzheimer's disease. Curr Pharm Des. 2006;12:4389–4408. doi: 10.2174/138161206778792967. [DOI] [PubMed] [Google Scholar]
  139. Palmer TD, Willhoite AR, Gage FH. Vascular niche for adult hippocampal neurogenesis. J Comp Neurol. 2000;425:479–494. doi: 10.1002/1096-9861(20001002)425:4<479::aid-cne2>3.0.co;2-3. [DOI] [PubMed] [Google Scholar]
  140. Paris D, Townsend KP, Obregon DF, Humphrey J, Mullan M. Pro-inflammatory effect of freshly solubilized beta-amyloid peptides in the brain. Prostaglandins Other Lipid Mediat. 2002;70:1–12. doi: 10.1016/s0090-6980(02)00111-9. [DOI] [PubMed] [Google Scholar]
  141. Park DS, Farinelli SE, Greene LA. Inhibitors of cyclin-dependent kinases promote survival of post-mitotic neuronally differentiated PC12 cells and sympathetic neurons. J Biol Chem. 1996;271:8161–8169. doi: 10.1074/jbc.271.14.8161. [DOI] [PubMed] [Google Scholar]
  142. Park DS, Morris EJ, Greene LA, Geller HM. G1/S cell cycle blockers and inhibitors of cyclin-dependent kinases suppress camptothecin-induced neuronal apoptosis. J Neurosci. 1997;17:1256–1270. doi: 10.1523/JNEUROSCI.17-04-01256.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Park DS, Obeidat A, Giovanni A, Greene LA. Cell cycle regulators in neuronal death evoked by excitotoxic stress: implications for neurodegeneration and its treatment. Neurobiol Aging. 2000;21:771–781. doi: 10.1016/s0197-4580(00)00220-7. [DOI] [PubMed] [Google Scholar]
  144. Park KW, Jin BK. Thrombin-induced oxidative stress contributes to the death of hippocampal neurons: role of neuronal NADPH oxidase. J Neurosci Res. 2008;86:1053–1063. doi: 10.1002/jnr.21571. [DOI] [PubMed] [Google Scholar]
  145. Pines J. Cyclins and cyclin-dependent kinases: a biochemical view. Biochem J. 1995;308(Pt 3):697–711. doi: 10.1042/bj3080697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Qian JY, Leung A, Harding P, LaPointe MC. PGE2 stimulates human brain natriuretic peptide expression via EP4 and p42/44 MAPK. Am J Physiol Heart Circ Physiol. 2006;290:H1740–1746. doi: 10.1152/ajpheart.00904.2005. [DOI] [PubMed] [Google Scholar]
  147. Quelle DE, Ashmun RA, Shurtleff SA, Kato JY, Bar-Sagi D, Roussel MF, Sherr CJ. Overexpression of mouse D-type cyclins accelerates G1 phase in rodent fibroblasts. Genes Dev. 1993;7:1559–1571. doi: 10.1101/gad.7.8.1559. [DOI] [PubMed] [Google Scholar]
  148. Quiniou C, Kooli E, Joyal JS, Sapieha P, Sennlaub F, Lahaie I, Shao Z, Hou X, Hardy P, Lubell W, Chemtob S. Interleukin-1 and ischemic brain injury in the newborn: development of a small molecule inhibitor of IL-1 receptor. Semin Perinatol. 2008;32:325–333. doi: 10.1053/j.semperi.2008.07.001. [DOI] [PubMed] [Google Scholar]
  149. Ranganathan S, Bowser R. Alterations in G(1) to S phase cell-cycle regulators during amyotrophic lateral sclerosis. Am J Pathol. 2003;162:823–835. doi: 10.1016/S0002-9440(10)63879-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Rao HV, Thirumangalakudi L, Grammas P. Cyclin C and cyclin dependent kinases 1, 2 and 3 in thrombin-induced neuronal cell cycle progression and apoptosis. Neurosci Lett. 2008 doi: 10.1016/j.neulet.2008.12.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Rao HV, Thirumangalakudi L, Desmond P, Grammas P. Cyclin D1, cdk4, and Bim are involved in thrombin-induced apoptosis in cultured cortical neurons. J Neurochem. 2007;101:498–505. doi: 10.1111/j.1471-4159.2006.04389.x. [DOI] [PubMed] [Google Scholar]
  152. Resnitzky D, Gossen M, Bujard H, Reed SI. Acceleration of the G1/S phase transition by expression of cyclins D1 and E with an inducible system. Mol Cell Biol. 1994;14:1669–1679. doi: 10.1128/mcb.14.3.1669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Richard Green A, Odergren T, Ashwood T. Animal models of stroke: do they have value for discovering neuroprotective agents? Trends Pharmacol Sci. 2003;24:402–408. doi: 10.1016/S0165-6147(03)00192-5. [DOI] [PubMed] [Google Scholar]
  154. Roberds SL, et al. BACE knockout mice are healthy despite lacking the primary beta-secretase activity in brain: implications for Alzheimer's disease therapeutics. Hum Mol Genet. 2001;10:1317–1324. doi: 10.1093/hmg/10.12.1317. [DOI] [PubMed] [Google Scholar]
  155. Robinson DM, Keating GM. Memantine: a review of its use in Alzheimer's disease. Drugs. 2006;66:1515–1534. doi: 10.2165/00003495-200666110-00015. [DOI] [PubMed] [Google Scholar]
  156. Sausville EA. Cyclin-dependent kinase modulators studied at the NCI: pre-clinical and clinical studies. Curr Med Chem Anticancer Agents. 2003;3:47–56. doi: 10.2174/1568011033353560. [DOI] [PubMed] [Google Scholar]
  157. Schwartz GK, Shah MA. Targeting the cell cycle: a new approach to cancer therapy. J Clin Oncol. 2005;23:9408–9421. doi: 10.1200/JCO.2005.01.5594. [DOI] [PubMed] [Google Scholar]
  158. Schwartz GK, Farsi K, Maslak P, Kelsen DP, Spriggs D. Potentiation of apoptosis by flavopiridol in mitomycin-C-treated gastric and breast cancer cells. Clin Cancer Res. 1997;3:1467–1472. [PubMed] [Google Scholar]
  159. Segarra J, Balenci L, Drenth T, Maina F, Lamballe F. Combined signaling through ERK, PI3K/AKT, and RAC1/p38 is required for met-triggered cortical neuron migration. J Biol Chem. 2006;281:4771–4778. doi: 10.1074/jbc.M508298200. [DOI] [PubMed] [Google Scholar]
  160. Selkoe DJ. Folding proteins in fatal ways. Nature. 2003;426:900–904. doi: 10.1038/nature02264. [DOI] [PubMed] [Google Scholar]
  161. Senderowicz AM. Flavopiridol: the first cyclin-dependent kinase inhibitor in human clinical trials. Invest New Drugs. 1999;17:313–320. doi: 10.1023/a:1006353008903. [DOI] [PubMed] [Google Scholar]
  162. Senderowicz AM, Sausville EA. Preclinical and clinical development of cyclin-dependent kinase modulators. J Natl Cancer Inst. 2000;92:376–387. doi: 10.1093/jnci/92.5.376. [DOI] [PubMed] [Google Scholar]
  163. Shankland SJ, Wolf G. Cell cycle regulatory proteins in renal disease: role in hypertrophy, proliferation, and apoptosis. Am J Physiol Renal Physiol. 2000;278:F515–529. doi: 10.1152/ajprenal.2000.278.4.F515. [DOI] [PubMed] [Google Scholar]
  164. Shapiro GI, Koestner DA, Matranga CB, Rollins BJ. Flavopiridol induces cell cycle arrest and p53-independent apoptosis in non-small cell lung cancer cell lines. Clin Cancer Res. 1999;5:2925–2938. [PubMed] [Google Scholar]
  165. Shen J, Xie L, Mao X, Zhou Y, Zhan R, Greenberg DA, Jin K. Neurogenesis after primary intracerebral hemorrhage in adult human brain. J Cereb Blood Flow Metab. 2008;28:1460–1468. doi: 10.1038/jcbfm.2008.37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Sherr CJ, Roberts JM. Inhibitors of mammalian G1 cyclin-dependent kinases. Genes Dev. 1995;9:1149–1163. doi: 10.1101/gad.9.10.1149. [DOI] [PubMed] [Google Scholar]
  167. Smirnova IV, Citron BA, Arnold PM, Festoff BW. Neuroprotective signal transduction in model motor neurons exposed to thrombin: G-protein modulation effects on neurite outgrowth, Ca(2+) mobilization, and apoptosis. J Neurobiol. 2001;48:87–100. [PubMed] [Google Scholar]
  168. Smith PJ, Marquez N, Wiltshire M, Chappell S, Njoh K, Campbell L, Khan IA, Silvestre O, Errington RJ. Mitotic bypass via an occult cell cycle phase following DNA topoisomerase II inhibition in p53 functional human tumor cells. Cell Cycle. 2007;6:2071–2081. doi: 10.4161/cc.6.16.4585. [DOI] [PubMed] [Google Scholar]
  169. Sokolova E, Reiser G. Prothrombin/thrombin and the thrombin receptors PAR-1 and PAR-4 in the brain: localization, expression and participation in neurodegenerative diseases. Thromb Haemost. 2008;100:576–581. [PubMed] [Google Scholar]
  170. Sridhar J, Akula N, Pattabiraman N. Selectivity and potency of cyclin-dependent kinase inhibitors. Aaps J. 2006;8:E204–221. doi: 10.1208/aapsj080125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Standridge JB. Vicious cycles within the neuropathophysiologic mechanisms of Alzheimer's disease. Curr Alzheimer Res. 2006;3:95–108. doi: 10.2174/156720506776383068. [DOI] [PubMed] [Google Scholar]
  172. Stefanis L, Park DS, Friedman WJ, Greene LA. Caspase-dependent and - independent death of camptothecin-treated embryonic cortical neurons. J Neurosci. 1999;19:6235–6247. doi: 10.1523/JNEUROSCI.19-15-06235.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Striggow F, Riek M, Breder J, Henrich-Noack P, Reymann KG, Reiser G. The protease thrombin is an endogenous mediator of hippocampal neuroprotection against ischemia at low concentrations but causes degeneration at high concentrations. Proc Natl Acad Sci U S A. 2000;97:2264–2269. doi: 10.1073/pnas.040552897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Sturm EM, Schratl P, Schuligoi R, Konya V, Sturm GJ, Lippe IT, Peskar BA, Heinemann A. Prostaglandin E2 inhibits eosinophil trafficking through E-prostanoid 2 receptors. J Immunol. 2008;181:7273–7283. doi: 10.4049/jimmunol.181.10.7273. [DOI] [PubMed] [Google Scholar]
  175. Su Y, Ryder J, Li B, Wu X, Fox N, Solenberg P, Brune K, Paul S, Zhou Y, Liu F, Ni B. Lithium, a common drug for bipolar disorder treatment, regulates amyloid-beta precursor protein processing. Biochemistry. 2004;43:6899–6908. doi: 10.1021/bi035627j. [DOI] [PubMed] [Google Scholar]
  176. Sun Z, Zhao Z, Zhao S, Sheng Y, Zhao Z, Gao C, Li J, Liu X. Recombinant hirudin treatment modulates aquaporin-4 and aquaporin-9 expression after intracerebral hemorrhage in vivo. Mol Biol Rep. 2008 doi: 10.1007/s11033-008-9287-3. [DOI] [PubMed] [Google Scholar]
  177. Suzuki M, Nelson AD, Eickstaedt JB, Wallace K, Wright LS, Svendsen CN. Glutamate enhances proliferation and neurogenesis in human neural progenitor cell cultures derived from the fetal cortex. Eur J Neurosci. 2006;24:645–653. doi: 10.1111/j.1460-9568.2006.04957.x. [DOI] [PubMed] [Google Scholar]
  178. Tan J, et al. NEURODEGENERATIVE DISEASE TREATMENT USING JAK/STAT INHIBITION. PCT/US2008/055646. 2008 [Google Scholar]
  179. Tanabe M, Nagatani Y, Saitoh K, Takasu K, Ono H. Pharmacological assessments of nitric oxide synthase isoforms and downstream diversity of NO signaling in the maintenance of thermal and mechanical hypersensitivity after peripheral nerve injury in mice. Neuropharmacology. 2009;56:702–708. doi: 10.1016/j.neuropharm.2008.12.003. [DOI] [PubMed] [Google Scholar]
  180. Tang Y, Lu A, Aronow BJ, Sharp FR. Blood genomic responses differ after stroke, seizures, hypoglycemia, and hypoxia: blood genomic fingerprints of disease. Ann Neurol. 2001;50:699–707. doi: 10.1002/ana.10042. [DOI] [PubMed] [Google Scholar]
  181. Tang Y, Lu A, Aronow BJ, Wagner KR, Sharp FR. Genomic responses of the brain to ischemic stroke, intracerebral haemorrhage, kainate seizures, hypoglycemia, and hypoxia. Eur J Neurosci. 2002;15:1937–1952. doi: 10.1046/j.1460-9568.2002.02030.x. [DOI] [PubMed] [Google Scholar]
  182. TauRx-Therapeutics-Ltd. TRx0014 in Patients With Mild or Moderate Alzheimer's Disease. 2007 ClinicalTrials.gov Identifier: NCT00515333 http://clinicaltrialsgov/ct2/show/NCT00515333?spons=%22TauRx+Therapeutics+Ltd%22&spons_ex=Y&rank=1.
  183. Tian DS, Yu ZY, Xie MJ, Bu BT, Witte OW, Wang W. Suppression of astroglial scar formation and enhanced axonal regeneration associated with functional recovery in a spinal cord injury rat model by the cell cycle inhibitor olomoucine. J Neurosci Res. 2006;84:1053–1063. doi: 10.1002/jnr.20999. [DOI] [PubMed] [Google Scholar]
  184. Tirado OM, Mateo-Lozano S, Notario V. Roscovitine is an effective inducer of apoptosis of Ewing's sarcoma family tumor cells in vitro and in vivo. Cancer Res. 2005;65:9320–9327. doi: 10.1158/0008-5472.CAN-05-1276. [DOI] [PubMed] [Google Scholar]
  185. Tong AW, Nemunaitis J. Modulation of miRNA activity in human cancer: a new paradigm for cancer gene therapy? Cancer Gene Ther. 2008;15:341–355. doi: 10.1038/cgt.2008.8. [DOI] [PubMed] [Google Scholar]
  186. Tweedie D, Sambamurti K, Greig NH. TNF-alpha inhibition as a treatment strategy for neurodegenerative disorders: new drug candidates and targets. Curr Alzheimer Res. 2007;4:378–385. doi: 10.2174/156720507781788873. [DOI] [PubMed] [Google Scholar]
  187. van Groen T, Kiliaan AJ, Kadish I. Deposition of mouse amyloid beta in human APP/PS1 double and single AD model transgenic mice. Neurobiol Dis. 2006;23:653–662. doi: 10.1016/j.nbd.2006.05.010. [DOI] [PubMed] [Google Scholar]
  188. Varvel NH, Bhaskar K, Patil AR, Pimplikar SW, Herrup K, Lamb BT. Abeta oligomers induce neuronal cell cycle events in Alzheimer's disease. J Neurosci. 2008;28:10786–10793. doi: 10.1523/JNEUROSCI.2441-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Verdaguer E, Jimenez A, Canudas AM, Jorda EG, Sureda FX, Pallas M, Camins A. Inhibition of cell cycle pathway by flavopiridol promotes survival of cerebellar granule cells after an excitotoxic treatment. J Pharmacol Exp Ther. 2004a;308:609–616. doi: 10.1124/jpet.103.057497. [DOI] [PubMed] [Google Scholar]
  190. Verdaguer E, Jorda EG, Stranges A, Canudas AM, Jimenez A, Sureda FX, Pallas M, Camins A. Inhibition of CDKs: a strategy for preventing kainic acid-induced apoptosis in neurons. Ann N Y Acad Sci. 2003a;1010:671–674. doi: 10.1196/annals.1299.122. [DOI] [PubMed] [Google Scholar]
  191. Verdaguer E, Jorda EG, Canudas AM, Jimenez A, Pubill D, Escubedo E, Camarasa J, Pallas M, Camins A. Antiapoptotic effects of roscovitine in cerebellar granule cells deprived of serum and potassium: a cell cycle-related mechanism. Neurochem Int. 2004b;44:251–261. doi: 10.1016/s0197-0186(03)00147-5. [DOI] [PubMed] [Google Scholar]
  192. Verdaguer E, Jorda EG, Canudas AM, Jimenez A, Sureda FX, Rimbau V, Pubill D, Escubedo E, Camarasa J, Pallas M, Camins A. 3-Amino thioacridone, a selective cyclin-dependent kinase 4 inhibitor, attenuates kainic acid-induced apoptosis in neurons. Neuroscience. 2003b;120:599–603. doi: 10.1016/s0306-4522(03)00424-x. [DOI] [PubMed] [Google Scholar]
  193. Vina J, Lloret A, Orti R, Alonso D. Molecular bases of the treatment of Alzheimer's disease with antioxidants: prevention of oxidative stress. Mol Aspects Med. 2004;25:117–123. doi: 10.1016/j.mam.2004.02.013. [DOI] [PubMed] [Google Scholar]
  194. Vincent I, Jicha G, Rosado M, Dickson DW. Aberrant expression of mitotic cdc2/cyclin B1 kinase in degenerating neurons of Alzheimer's disease brain. J Neurosci. 1997;17:3588–3598. doi: 10.1523/JNEUROSCI.17-10-03588.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Wang CX, Shuaib A. Neuroprotective effects of free radical scavengers in stroke. Drugs Aging. 2007;24:537–546. doi: 10.2165/00002512-200724070-00002. [DOI] [PubMed] [Google Scholar]
  196. Wang F, Corbett D, Osuga H, Osuga S, Ikeda JE, Slack RS, Hogan MJ, Hakim AM, Park DS. 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]
  197. Wang H, Reiser G. The role of the Ca2+-sensitive tyrosine kinase Pyk2 and Src in thrombin signalling in rat astrocytes. J Neurochem. 2003a;84:1349–1357. doi: 10.1046/j.1471-4159.2003.01637.x. [DOI] [PubMed] [Google Scholar]
  198. Wang H, Reiser G. Thrombin signaling in the brain: the role of protease-activated receptors. Biol Chem. 2003b;384:193–202. doi: 10.1515/BC.2003.021. [DOI] [PubMed] [Google Scholar]
  199. Wang LH, Besirli CG, Johnson EM., Jr. Mixed-lineage kinases: a target for the prevention of neurodegeneration. Annu Rev Pharmacol Toxicol. 2004;44:451–474. doi: 10.1146/annurev.pharmtox.44.101802.121840. [DOI] [PubMed] [Google Scholar]
  200. Warner TD, Mitchell JA. Cyclooxygenases: new forms, new inhibitors, and lessons from the clinic. Faseb J. 2004;18:790–804. doi: 10.1096/fj.03-0645rev. [DOI] [PubMed] [Google Scholar]
  201. Whittaker SR, Te Poele RH, Chan F, Linardopoulos S, Walton MI, Garrett MD, Workman P. The cyclin-dependent kinase inhibitor seliciclib (Rroscovitine; CYC202) decreases the expression of mitotic control genes and prevents entry into mitosis. Cell Cycle. 2007;6:3114–3131. doi: 10.4161/cc.6.24.5142. [DOI] [PubMed] [Google Scholar]
  202. Williams M. Progress in Alzheimer's disease drug discovery: an update. Curr Opin Investig Drugs. 2009;10:23–34. [PubMed] [Google Scholar]
  203. Williamson R, Scales T, Clark BR, Gibb G, Reynolds CH, Kellie S, Bird IN, Varndell IM, Sheppard PW, Everall I, Anderton BH. Rapid tyrosine phosphorylation of neuronal proteins including tau and focal adhesion kinase in response to amyloid-beta peptide exposure: involvement of Src family protein kinases. J Neurosci. 2002;22:10–20. doi: 10.1523/JNEUROSCI.22-01-00010.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Wyatt PG, et al. Identification of N-(4-piperidinyl)-4-(2,6-dichlorobenzoylamino)-1H-pyrazole-3-carboxamide (AT7519), a novel cyclin dependent kinase inhibitor using fragment-based X-ray crystallography and structure based drug design. J Med Chem. 2008;51:4986–4999. doi: 10.1021/jm800382h. [DOI] [PubMed] [Google Scholar]
  205. Xi G, Reiser G, Keep RF. The role of thrombin and thrombin receptors in ischemic, hemorrhagic and traumatic brain injury: deleterious or protective? J Neurochem. 2003;84:3–9. doi: 10.1046/j.1471-4159.2003.01268.x. [DOI] [PubMed] [Google Scholar]
  206. Xi G, Keep RF, Hoff JT. Mechanisms of brain injury after intracerebral haemorrhage. Lancet Neurol. 2006;5:53–63. doi: 10.1016/S1474-4422(05)70283-0. [DOI] [PubMed] [Google Scholar]
  207. Xing B, Xin T, Hunter RL, Bing G. Pioglitazone inhibition of lipopolysaccharideinduced nitric oxide synthase is associated with altered activity of p38 MAP kinase and PI3K/Akt. J Neuroinflammation. 2008;5:4. doi: 10.1186/1742-2094-5-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Xue M, Hollenberg MD, Yong VW. Combination of thrombin and matrix metalloproteinase-9 exacerbates neurotoxicity in cell culture and intracerebral hemorrhage in mice. J Neurosci. 2006;26:10281–10291. doi: 10.1523/JNEUROSCI.2806-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Yang Y, Herrup K. Cell division in the CNS: protective response or lethal event in post-mitotic neurons? Biochim Biophys Acta. 2007;1772:457–466. doi: 10.1016/j.bbadis.2006.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Yang Y, Geldmacher DS, Herrup K. DNA replication precedes neuronal cell death in Alzheimer's disease. J Neurosci. 2001;21:2661–2668. doi: 10.1523/JNEUROSCI.21-08-02661.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Yang Y, Mufson EJ, Herrup K. Neuronal cell death is preceded by cell cycle events at all stages of Alzheimer's disease. J Neurosci. 2003;23:2557–2563. doi: 10.1523/JNEUROSCI.23-07-02557.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Yang Y, Varvel NH, Lamb BT, Herrup K. Ectopic cell cycle events link human Alzheimer's disease and amyloid precursor protein transgenic mouse models. J Neurosci. 2006;26:775–784. doi: 10.1523/JNEUROSCI.3707-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Yoshino H, Kimura A. Investigation of the therapeutic effects of edaravone, a free radical scavenger, on amyotrophic lateral sclerosis (Phase II study) Amyotroph Lateral Scler. 2006;7:241–245. doi: 10.1080/17482960600881870. [DOI] [PubMed] [Google Scholar]
  214. Yuan WJ, Yasuhara T, Shingo T, Muraoka K, Agari T, Kameda M, Uozumi T, Tajiri N, Morimoto T, Jing M, Baba T, Wang F, Leung H, Matsui T, Miyoshi Y, Date I. Neuroprotective effects of edaravone-administration on 6-OHDA-treated dopaminergic neurons. BMC Neurosci. 2008;9:75. doi: 10.1186/1471-2202-9-75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Zdanys K, Tampi RR. A systematic review of off-label uses of memantine for psychiatric disorders. Prog Neuropsychopharmacol Biol Psychiatry. 2008;32:1362–1374. doi: 10.1016/j.pnpbp.2008.01.008. [DOI] [PubMed] [Google Scholar]
  216. Zhang Y, Qu D, Morris EJ, O'Hare MJ, Callaghan SM, Slack RS, Geller HM, Park DS. The Chk1/Cdc25A pathway as activators of the cell cycle in neuronal death induced by camptothecin. J Neurosci. 2006;26:8819–8828. doi: 10.1523/JNEUROSCI.2593-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Zhao L, Teter B, Morihara T, Lim GP, Ambegaokar SS, Ubeda OJ, Frautschy SA, Cole GM. Insulin-degrading enzyme as a downstream target of insulin receptor signaling cascade: implications for Alzheimer's disease intervention. J Neurosci. 2004;24:11120–11126. doi: 10.1523/JNEUROSCI.2860-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Zhu X, Raina AK, Perry G, Smith MA. Alzheimer's disease: the two-hit hypothesis. Lancet Neurol. 2004;3:219–226. doi: 10.1016/S1474-4422(04)00707-0. [DOI] [PubMed] [Google Scholar]
  219. Zhu X, Lee HG, Perry G, Smith MA. Alzheimer disease, the two-hit hypothesis: an update. Biochim Biophys Acta. 2007;1772:494–502. doi: 10.1016/j.bbadis.2006.10.014. [DOI] [PubMed] [Google Scholar]
  220. Zhu X, Rottkamp CA, Boux H, Takeda A, Perry G, Smith MA. Activation of p38 kinase links tau phosphorylation, oxidative stress, and cell cycle-related events in Alzheimer disease. J Neuropathol Exp Neurol. 2000;59:880–888. doi: 10.1093/jnen/59.10.880. [DOI] [PubMed] [Google Scholar]

RESOURCES