Neuroprotection by monoamine oxidase B inhibitors: a therapeutic strategy for Parkinson's disease

11
Neuroprotection by monoamine oxidase B inhibitors: a therapeutic strategy for Parkinson’s disease? { Rinat Tabakman, { Shimon Lecht, and Philip Lazarovici* Summary Parkinsonism (PD) is a neurodegenerative disorder of the brain resulting in dopamine deficiency caused by the progressive death of dopaminergic neurons. PD is char- acterized by a combination of rigidity, poverty of move- ment, tremor and postural instability. Selegiline is a selective and irreversible propargylamine type B mono- amine oxidase (MAO-B) inhibitor. This drug, which in- hibits dopamine metabolism, has been effectively used in the treatment of PD. However, its therapeutic effects are compromised by its many neurotoxic metabolites. To circumvent this obstacle, a novel MAO-B inhibitor, rasagi- line, was developed. Paradoxically, the neuroprotective mechanism of propargylamines in different neuronal models appears to be independent of MAO-B inhibition. Recent investigations into the neuroprotective mechan- ism of propargylamines indicate that glyceraldehyde-3- phosphate dehydrogenase (GAPDH), MAO-B and/or other unknown proteins may represent pivotal proteins in the survival of the injured neurons. Delineation of the mechanism(s) involved in the neuroprotective effects exerted by MAO-B inhibitors may provide the key to preventive novel therapeutic modalities. BioEssays 26:80–90, 2004. ß 2003 Wiley Periodicals, Inc. Introduction Parkinson’s disease (PD) is a progressive movement disorder that affects 1 –2%of the adult population over 60 years of age. The main symptoms are tremor at rest, muscular rigidity and a decrease in the frequency of voluntary movements (hypoki- nesia). (1) For many years, it has been known that PD synd- rome is due to a disorder of the basal ganglia, brain structures regulating motor activity and innervated by one of the brain’s major dopaminergic pathways. The neurochemical basis for the disease was discovered in 1960 by Hornykiewics, (2) who showed that the dopamine content of the substantia nigra (SN) and corpus striatum in postmortem PD brains was extremely low (less than 10% of normal). Pathologically, PD is charac- terized by progressive degeneration of pigmented brain stem nuclei, mostly the pars compacta of the SN, along with the formation of characteristic eosinophilic cytoplasmic inclusions known as Lewy bodies. (3) The disease progresses slowly for many years. The clinical symptoms are due to the degenera- tion (death) of dopaminergic neurons in the SN, resulting in a dramatic decline in dopamine level. Diagnosis can be made when at least 60% of the dopaminergic SNc neurons are lost. (4) Many factors are considered to contribute to the pathogenesis of PD: genetic, (5) age-related, (6) enviromental toxins, (7) and oxidative stress. (8,9) In the past decade, novel hypotheses have been forwarded suggesting important pathological con- tributions to dopaminergic neuronal cell death: glutamate neurotoxicity, (10) mitochondrial abnormalities, disturbances in intracellular calcium homeostasis, altered iron metabolism and apoptosis. (9,11–14) Although the genes responsible for a few rare familial cases have been uncovered, (15–17) the 80 BioEssays 26.1 BioEssays 26:80–90, ß 2003 Wiley Periodicals, Inc. Department of Pharmacology, School of Pharmacy, Faculty of Medicine, The Hebrew University of Jerusalem, Jerusalem, Israel. *Correspondence to: Philip Lazarovici, Department of Pharmacology and Experimental Therapeutics, School of Pharmacy, Faculty of Medicine, The Hebrew University, Jerusalem, 91120, Israel. E-mail: [email protected] y This review is part of a PhD thesis to be submitted to the Hebrew University of Jerusalem by TR. z The scientific views and opinions expressed in this article are solely those of the authors and are not to be construed as having any commercial interest. DOI 10.1002/bies.10378 Published online in Wiley InterScience (www.interscience.wiley.com). Abbreviations: PD, Parkinson’s disease; UPDRS, unified Parkinson’s disease rating scale; SN, substantia nigra; DA, dopaminergic; MAO-B, type B monoamine oxidase; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; SOD, superoxide dismutase; NADþ, nicotinamide adenine dinucleotide; FAD, flavin adenine dinucleotide; BDNF, brain- derived neurotrophic factor; NGF, nerve growth factor; GDNF, glial derived neurotrophic factor; DMS, desmethylselegiline; MPPþ, 1- methyl-4-phenyl pyridinium; MPTP, N-methyl-4-phenyl-1,2,3,6-tetra- hydropyridine; 6-OH-DA, 6-hydroxydopamine; OGD, oxygen–glucose deprivation; PET, positron emission tomography; fMRI, functional magnetic resonance imaging; IC 50 , 50% inhibitory concentration; Ki, inhibition constant; PC12, pheochromocytoma cells. Challenges

Transcript of Neuroprotection by monoamine oxidase B inhibitors: a therapeutic strategy for Parkinson's disease

Neuroprotection by monoamineoxidase B inhibitors:a therapeutic strategyfor Parkinson’s disease?{

Rinat Tabakman,{ Shimon Lecht, and Philip Lazarovici*

SummaryParkinsonism (PD) is a neurodegenerative disorder of thebrain resulting in dopamine deficiency caused by theprogressive death of dopaminergic neurons. PD is char-acterized by a combination of rigidity, poverty of move-ment, tremor and postural instability. Selegiline is aselective and irreversible propargylamine type B mono-amine oxidase (MAO-B) inhibitor. This drug, which in-hibits dopaminemetabolism, hasbeen effectively used inthe treatment of PD. However, its therapeutic effects arecompromised by its many neurotoxic metabolites. Tocircumvent thisobstacle, anovelMAO-B inhibitor, rasagi-line, was developed. Paradoxically, the neuroprotectivemechanism of propargylamines in different neuronalmodels appears to be independent of MAO-B inhibition.

Recent investigations into the neuroprotective mechan-ism of propargylamines indicate that glyceraldehyde-3-phosphate dehydrogenase (GAPDH), MAO-B and/orother unknown proteins may represent pivotal proteinsin the survival of the injured neurons. Delineation of themechanism(s) involved in the neuroprotective effectsexerted by MAO-B inhibitors may provide the key topreventive novel therapeutic modalities. BioEssays26:80–90, 2004. � 2003 Wiley Periodicals, Inc.

Introduction

Parkinson’s disease (PD) is a progressive movement disorder

that affects 1–2%of the adult population over 60 years of age.

The main symptoms are tremor at rest, muscular rigidity and a

decrease in the frequency of voluntary movements (hypoki-

nesia).(1) For many years, it has been known that PD synd-

rome is due to a disorder of the basal ganglia, brain structures

regulating motor activity and innervated by one of the brain’s

major dopaminergic pathways. The neurochemical basis for

the disease was discovered in 1960 by Hornykiewics,(2) who

showed that the dopamine content of the substantia nigra (SN)

and corpus striatum in postmortem PD brains was extremely

low (less than 10% of normal). Pathologically, PD is charac-

terized by progressive degeneration of pigmented brain stem

nuclei, mostly the pars compacta of the SN, along with the

formation of characteristic eosinophilic cytoplasmic inclusions

known as Lewy bodies.(3) The disease progresses slowly for

many years. The clinical symptoms are due to the degenera-

tion (death) of dopaminergic neurons in the SN, resulting in a

dramatic decline in dopamine level. Diagnosis can be made

when at least 60% of the dopaminergic SNc neurons are lost.(4)

Many factors are considered to contribute to the pathogenesis

of PD: genetic,(5) age-related,(6) enviromental toxins,(7) and

oxidative stress.(8,9) In the past decade, novel hypotheses

have been forwarded suggesting important pathological con-

tributions to dopaminergic neuronal cell death: glutamate

neurotoxicity,(10) mitochondrial abnormalities, disturbances in

intracellular calcium homeostasis, altered iron metabolism

and apoptosis.(9,11–14) Although the genes responsible for

a few rare familial cases have been uncovered,(15–17) the

80 BioEssays 26.1 BioEssays 26:80–90, � 2003 Wiley Periodicals, Inc.

Department of Pharmacology, School of Pharmacy, Faculty of

Medicine, The Hebrew University of Jerusalem, Jerusalem, Israel.

*Correspondence to: Philip Lazarovici, Department of Pharmacology

and Experimental Therapeutics, School of Pharmacy, Faculty of

Medicine, The Hebrew University, Jerusalem, 91120, Israel.

E-mail: [email protected] review is part of a PhD thesis to be submitted to the Hebrew

University of Jerusalem by TR.zThe scientific views and opinions expressed in this article are solely

those of the authors and are not to be construed as having any

commercial interest.

DOI 10.1002/bies.10378

Published online in Wiley InterScience (www.interscience.wiley.com).

Abbreviations: PD, Parkinson’s disease; UPDRS, unified Parkinson’s

disease rating scale; SN, substantia nigra; DA, dopaminergic; MAO-B,

type B monoamine oxidase; GAPDH, glyceraldehyde-3-phosphate

dehydrogenase; SOD, superoxide dismutase; NADþ, nicotinamide

adenine dinucleotide; FAD, flavin adenine dinucleotide; BDNF, brain-

derived neurotrophic factor; NGF, nerve growth factor; GDNF, glial

derived neurotrophic factor; DMS, desmethylselegiline; MPPþ, 1-

methyl-4-phenyl pyridinium; MPTP, N-methyl-4-phenyl-1,2,3,6-tetra-

hydropyridine; 6-OH-DA, 6-hydroxydopamine; OGD, oxygen–glucose

deprivation; PET, positron emission tomography; fMRI, functional

magnetic resonance imaging; IC50, 50% inhibitory concentration; Ki,

inhibition constant; PC12, pheochromocytoma cells.

Challenges

molecular basis for the more prevalent idiopathic PD remains

unknown. Novel technologies including DNA-microarray(18)

and single nucleotide polymorphisms(19) are expected to

pinpoint the pathogenesis of this disorder. In addition, new

techniques in neuroimaging such as positron emission tomo-

graphy (PET) and functional magnetic resonance imaging

(fMRI) may provide better means of assessing SN function.(20)

In this review, we briefly summarize the emerging knowledge

regarding the complex mechanism of action of type B mono-

amine oxidase (MAO-B) inhibitors, an important family of

drugs used to treat PD. While the search for MAO-B inhibitors

has turned up some promising propargylamine candidates,

their mode of action in ameliorating PD cannot be explained

solely by MAO-B inhibition. Today, the basis for their neuro-

protective mechanism of action is being actively pursued. We

hypothesize that the beneficial role of MAO-B inhibitors in PD

is due to a shift from neurotoxicity to neuroprotection of the

injured dopaminergic neurons.

Apoptosis and Parkinson’s disease

Since 1996, when apoptotic death was discovered in

dopaminergic cells of Parkinsonian patients,(21) many addi-

tional studies have suggested the involvement of an apoptotic

mechanism in this disorder. Apoptosis, a delayed form of cell

death associated with the activation of a ‘‘genetic program’’, is

an important biological process controlling cell number in

various tissues. This mechanism plays an important role in the

development of the nervous system(22) and is also triggered

following pathological events such as stroke and neurode-

generative diseases.(14) The SN of PD patients shows an

increase in caspase activity and in the pro-apoptotic Bax

protein, as well as nuclear translocation of glyceraldehyde-3-

phosphate dehydrogenase (GAPDH), indicating the activation

of apoptotic signals.(23–25) However, it is not clear whether the

degeneration associated with PD involves a fast ‘event’ in

which a significant number of dopaminergic neurons deterio-

rate as a result of unknown factor(s),(26) or a very slow one that

may last several decades.(27) Furthermore, other postmortem

studies, based on nick-end labeling of fragmented DNA,(28)

Bcl-2 and Bax apoptotic protein markers,(27) failed to show a

significant contribution of apoptosis to SN dopaminergic

neuronal cell death. An additional complication adding to the

cell death controversy in PD is the finding that, in postmortem

tissue, apoptotic markers are present in the glia but not in SN

neurons.(29,30) In general, nigrostriatal dopaminergic neurons

are lost with age. This phenomenon is dramatically acceler-

ated in Parkinson’s patients. The grounds for the vulnerability

of the nigrostriatal dopaminergic neurons are unknown and the

precise mechanism of cell death awaits further investigation.

The gene-expression microarray technique is expected to

make a significant contribution to the identification of gene

products involved in cell death and leading to PD.(18)

Therapeutical approaches

to Parkinson’s disease

As the cause of the death of the dopaminergic neurons is not

known, drugs that can arrest and/or reduce and/or delay the

cell death process in dopaminergic neurons are not available.

However, significant symptomatic relief has been obtained by

the use of different dopaminergic drugs that fall into the follow-

ing categories: dopamine precursors (levodopa); compounds

mimicking (agonists) the action of dopamine (bromocriptine);

agents that prevent dopamine degradation (MAO-B inhibitors

such as rasagiline).

Levodopa, a dopamine precursor that crosses the blood–

brain barrier, is initially effective in most patients, but often

loses its efficacy after several years of treatment, the period

varying from one patient to another. Complications associated

with its use are involuntary movements, which occur in many

patients within several years, and an unpredictable ‘‘on–off’’

effect in the course of treatment.(31) Other adverse effects

include queasiness (nausea), a drop in blood pressure

(hypotension), and, occasionally, psychotic symptoms.(31)

Levodopa treatment requires a certain number of live dopami-

nergic neurons to convert this metabolic precursor into the

neurotransmitter dopamine (DA). As a result, the amount of

synaptic DA is increased and the DA deficiency characteristic

of PD is corrected. The disadvantage of this therapeutic

approach is the augmented cell death of large numbers of DA

neurons during PD progression, leading to less DA synthesis

during levodopa treatment and rendering it less effective.

Drugs that mimic DA (agonists) are less effective than

levodopa.(32) They activate postsynaptic DA receptors in the

SN and improve motor function independently of DA neuron

degeneration.

The first generation of MAO inhibitors, originally synthe-

sized in the 1950s, was used as antidepressants due to their

inhibitory effect on the metabolism of monoamine neurotrans-

mitters. These inhibitors are both irreversible and non-selec-

tive. However, their use as antidepressants was restricted

because of hepatotoxicity(33) and due to an increase in blood

pressure (hypertensive response) following the ingestion of

some foods and drinks containing tyramine,(34) which are nor-

mally metabolized in the gastrointestinal tract by MAO.(34,35)

As a result of MAO inhibition, the ingested tyramine enters the

circulation and is actively taken up by peripheral adrenergic

neurons, displacing stored noradrenaline and giving rise to a

hypertensive response that can be fatal.(36) This side-effect

pharmacological phenomenon(37) was named the ‘‘cheese

effect’’, since many cheeses are rich in tyramine. The next

discovery was made by Johnston,(38) who showed that the

MAO enzyme exists in two forms: A and B. This led to the

synthesis of a new generation of MAO inhibitors exhibiting

greater selectivity toward the individual isoforms. MAO-A

inhibitors are potent antidepressants but, since MAO-A is the

major isoform in the intestine, they induce the ‘‘cheese effect’’.

Challenges

BioEssays 26.1 81

The finding that MAO-B inhibitors have weaker antidepressant

activity, do not induce the ‘‘cheese effect,’’(39) but can amelio-

rate PD symptoms, was the basis for modern Parkinson

therapeutics,(39,40) using the MAO-B inhibitor selegiline and

rasagiline. Selegiline has been approved for use in the USA as

an adjunct to levodopa but not for monotherapy; rasagiline is a

new addition to the MAO-B inhibitor group.

Monoamine oxidases and inhibitors

in Parkinson’s disease

MAO (EC 1.4.3.4) oxidatively deaminates monoamine neuro-

transmitters (norepinephrine, epinephrine, serotonin and

dopamine), as well as exogeneous amines (e.g., tyramine).

The end products of the enzymatic reaction are aldehydes and

hydrogen peroxide, both toxic to DA neurons. The physio-

logical role of MAO is to terminate the action of several

neurotransmitters and to detoxify exogenous monoamines.

Two MAO subtypes, MAO-A and MAO-B, were defined in

1968,(38) based on their differential sensitivity to irreversible

inhibitors. 20 years later, the corresponding cDNAs of these

two subtypes were cloned. Sequence comparison indicated

72.6% amino acid homology.(41,42) The genes encoding the

two subtypes have been localized to human chromosome

Xp11.23-Xp11.4.(43) TheMAO-AandMAO-Bgenesarehomo-

logous at the level of intron–exon organization and in co-factor

FAD-binding site, and both enzymes are located at the outer

mitochondrial membrane.(44) Along with the similarities, MAO-

A and MAO-B display differences in their numbers of amino

acids, 527 and 520, and molecular weights 59.7 kDa and

58.8 kDa, respectively. The in vivo substrate affinity in humans

also differs: MAO-A metabolizes mainly serotonin, norepi-

nephrine and epinephrine, whereas MAO-B metabolizes

dopamine and phenylethylamine.(44) Both MAO-B and MAO-

A are widely distributed in the human body, the ratio between

them varying in different tissues. The highest value (>69) is

found in platelets; lower ratios were detected in the human

brain, the highest of which (5.5) was found in the SN.(45) The

selectivity toward different endogenous substrates, together

with tissue-specific localization, determines the clinical poten-

tial of each isoform. Inhibition of MAO activity in the brain

increases the synaptic level of the neurotransmitters seroto-

nin, noradrenaline and dopamine. This mechanism was

exploited to treat depression, first with nonselective MAO

inhibitors and, more recently, with selective MAO-A inhibitors.

Today however, inhibitors of MAO-A are not the drugs of first

choice in the treatment of depression, mainly because of the

development of novel antidepressants that act by different

pharmacological mechanisms, without eliciting the cheese

effect.

Many selective MAO-B inhibitors, potentially applicable to

PD treatment, have been synthesized. They are categorized

according to chemical structure, each including several

derivatives (Table 1). The most-potent and selective group

consists of propargylamine derivatives. These are irreversible

inhibitors, which would ensure continuous dopaminergic

stimulation, considered a major therapeutic goal. Lazabemide,

a 2-aminoethyl carboxamide derivative, which is a reversible,

highly selective antioxidant, was developed as a MAO-B

inhibitor for treatment of PD.(46) Clinical trials with this com-

pound were discontinued in phase III of the clinical study due

to the abnormal liver functions that developed in patients

undergoing this treatment. To date, of the variety of selective

MAO-B inhibitors investigated, only one, selegiline (l-depre-

nyl) (Table 1), has been in clinical use since the mid 1970s.(47)

Rasagiline (N-propargyl-1-R-aminoindan) (Table 1), a novel,

potent, selective, and irreversible inhibitor of MAO-B, absorb-

ed after oral administration,(48,49) is currently being evaluated

in phase III clinical trial.(40) In contrast to selegiline, rasagiline is

not metabolized to potentially toxic amphetaminic metabolites

and its major metabolite, 1-R-aminoindan, has exhibited

beneficial effects in animal models of PD.(50) Because of its

selectivity towards MAO-B, rasagiline does not trigger the

‘‘cheese effect’’,(51) and was found to improve motor and

cognitive function in PD animal models.(50) A recent clinical

trial demonstrated the beneficial influence of rasagiline mono-

therapy in early stages of Parkinson’s disease.(40) To evaluate

the safety and efficacy of rasagiline, a multicenter clinical trial

was carried out in which a total of 404 patients in early stages of

PD were included. The participants were randomly assigned to

groups receiving rasagiline at a dose of 1–2 mg per day or

corresponding placebo. A one-week escalation period was

followed by a 25-week maintenance period. The efficacy of

treatment was evaluated according to the Unified Parkinson’s

Disease Rating Scale (UPDRS); rasagiline monotherapy at

both doses proved effective.(40)

Although inhibition of MAO-B is of prime importance in

controlling PD symptoms, it appears that this inhibition under-

lies many additional effects in the brain. An interesting analogy

can be drawn between PD patients chronically administered

MAO-B inhibitors and knockout mice lacking the gene encod-

ing MAO-B. Chen et al. demonstrated significant upregulation

(�30%) of D2 receptors in the striatum and the functional

supersensitivity of D1 receptors in the mouse nucleus accum-

bens.(52) This upregulation resembles similar regulation of the

dopamine receptors observed in PD patients chronically

treated with MAO-B inhibitors. It has been suggested that

MAO-Bmakes asmaller contribution thanMAO-A todopamine

metabolism in the mouse, as compared with man,(53) making it

difficult to conclude that similar changes occur in MAO-B

inhibitor-treated patients. However, subjects in whom the

MAO-B gene is deleted (atypical Norrie disease) have normal

levels of DA metabolites in their plasma.(54) The parallelism

between MAO-B knockout mice and humans with a deleted

MAO-B gene, suggests a similarity in the lack of change in

DA metabolites levels This additional circumstantial evidence

leads us to postulate that upon treatment with MAO-B

Challenges

82 BioEssays 26.1

inhibitors, DA receptors undergo alterations seen in the MAO-

B knockout mice. However, changes in dopamine receptors

are not the sole effect of the MAO-B inhibitors. Recent studies

show the potential for an additional beneficial activity:

neuroprotection. Cumulative in vivo and in vitro evidence

points to the neuroprotective properties of propargylamine

MAO-B inhibitors, such as selegiline and rasagiline.(55–57)

MAO-B inhibitors and neuroprotection

in different neurotoxic models

Since PD progresses very slowly and over many years, its

diagnosis poses a tremendous challenge. Because the trigger

for dopaminergic cell death is unknown, the therapeutic

strategy in the last 15 years has been based on the DA

precursor levodopa, which raises the DA level. The adverse

effects of the treatment, the fact that this compound does

not arrest dopaminergic cell death and disease progression

and the possible contribution of apoptotic cell death to

Parkinson’s disease, prompted the search for new neuropro-

tective approaches. A neuroprotective therapeutic or disease-

modifying modality may be defined as an intervention that

delays or prevents neuronal cell death and thus affects

disease progression.

Selegiline aroused considerable interest as a potential

neuroprotective drug. As summarized in Tables 2 and 3,

preclinical studies point to the neuroprotective effect of the

Table 1. Classification of MAO-B inhibitors of clinical relevance, according to chemical structure, potency,

mechanism of action and selectivity

Chemical group Chemical structureMAO-B inhibition

(IC50 nM)Mechanismof inhibition

Selectivitya

MAOB/MAOA References

Propargylamine derivatives

Selegiline

6 Irreversible 233 95

Rasagiline 30 Irreversible 100 95

Allylamine derivatives 100 Irreversible 50 95

2-aminoethyl carboxamide

derivatives

Lazabemide

37 Reversible 26568 95

N-allenic indolalkylamine

derivatives

25–25000 Irreversible >1 96

4-substituted

cubylcarbinyl amines

120–260b Irreversible — 97

aRatio between MAO-B and MAO-A (IC50). High value indicates greater selectivity towards MAO-B.bKi.

Challenges

BioEssays 26.1 83

propargylamine MAO-B inhibitors selegiline and rasagiline.

This is illustrated by in vitro studies using a variety of

dopaminergic and nondopaminergic neurons taken from

different species and exposed to a range of neurotoxic insults

(excitotoxicity, oxidative stress, depolarization, growth factor

withdrawal and oxygen glucose deprivation). For example,

rasagiline prolongs the survival of cultured primary human and

rat DA neurons under serum-free conditions,(58) reduces

glutamate toxicity in hippocampal neurons,(59) protects nerve

growth factor (NGF)-differentiated pheochromocytoma PC12

cells from oxygen-glucose-deprivation (OGD-ischemia) in-

sult,(57,60) prevents deficits in behavioral parameters following

hypoxia in adult and senescent rats,(61) protects from experi-

mental focal ischemia in the rat,(62) and from closed head injury

in the mouse.(63) The last three are invivomodels. Although the

neuroprotective in vitro effect of selegiline, and even more so

of rasagiline, is clear-cut, the lackof reliable pathophysiological

models for the disease has led to some skepticism among the

Parkinson medical community. The PD models using neuronal

cultures treated with 1-methyl-4-phenyl pyridinium (MPPþ, a

metabolite of the neurotoxin N-methyl-4-phenyl-1,2,3,6-tetra-

hydropyridine (MPTP)) and 6-hydroxydopamine (6-OH-DA)

Table 2. Neuroprotective in vitro effects of MAO-B inhibitors selegiline and rasagiline

MAO-B Inhibitor Insult ModelNeurotoxicitymechanism

Neuroprotectivemechanism References

Selegiline Excitotoxicity Mesencephalic

dopaminergic, primary

neurons

Depolarization and

calcium overload

Yes

Independent of MAO-B

98

Hypocampal glia and

neurons

Oxidative stress Yes, release of NGF 99

MPPþ Dopaminergic neurons

mesencephalic and

striatal cells

Oxidative stress Yes 100

BSO L-buthionine-(S,R)-

sulfoximine

Mesencephalic neurons Glutathione depletion Yes

Independent of MAO-B

101

Serum and NGF

withdrawal

NGF-partially

differentiated rat

PC12 cells

Oxidative stress and

growth factor

starvation

Yes

Independent of

MAO-B, induction

of transcription

102

Natural cell death

combined with serum

deprivation

Mesencephalic neurons Apoptosis and

oxidative stress

Yes

Serum-enhanced

49

103

104

Serum deprivation and

hypoxia

E1A-NR3 immortalized

retinal neurons

Oxidative stress and

hypoxia

Yesa 105

SIN-1 Neuroblastoma

SH-SY5Y cells

Oxidative stress Yes 77

Rasagiline Natural cell death

combined with serum

deprivation

Mesencephalic neurons Apoptosis and

oxidative stress

Yes

Serum-enhanced

49

103

SIN-1 Neuroblastoma

SH-SY5Y cells

Oxidative stress Yesb 75

76

77

6-OHDA, SIN-1 Neuroblastoma SH-SY5Y

cells

Oxidative stress Yesb

Independent of MAO-B

76

N-methyl(R)salsolinol,

6-OHDA, peroxynitrite

Neuroblastoma

SH-SY5Y cells

Oxidative stress Yesb,c

Independent of MAO-B

106

N-methyl(R)salsolinol SH-SY5Y cells

overexpressing Bcl-2

Oxidative stress Yesd 107

Serum and NGF

withdrawal

Partially differentiated

PC12 cells

Free radical-induced

apoptosis

Yese 48

Oxidative stress Independent of MAO-B

OGD NGF-differentiated

PC12 cells

Oxygen-glucose

deprivation

Yes

Independent of MAO-B

57

60

aRegulation of apoptosis-related gene expression.bStabilization of mitochondrial potential.cSuppresion of caspases and DNA fragmentation.dPrevention of nuclear accumulation of GAPDH.eIncreased gene expression of anti-apoptotic targets (proteins).

Challenges

84 BioEssays 26.1

to trigger cell death, are the most common, and less

disputable. They are the main models used for screening

and developing novel drugs for PD treatment, although the

animals injected with the neurotoxin MPTP do not exhibit Lewy

bodies, a hallmark of PD.

While a variety of neurological deficits are triggered by the

above-mentioned insults in in vivo animal models, mainly the

MPTP model partially mimics the PD syndrome. In rats fed

selegiline, the lifespan of the animals was increased, in parallel

with enhanced catecholaminergic activity in the brain.(64,65)

This phenomenon was unrelated to MAO-B inhibition and

was a topic of controversy. The neuroprotective effects of the

MAO-B inhibitors selegiline and rasagiline have also been

investigated in mice, gerbils and monkeys (Table 3), in brain-

ischemia, head-trauma, and under MPTP toxicity. As opposed

to the clearly neuroprotective effects of propargylamines in

in vitro cellular systems and in vivo experimental animal

models, clinical studies with selegiline have failed to distin-

guish between the contribution of MAO-B inhibition and

MAO-B induced neuroprotection to the amelioration of clinical

symptoms.(66) Although selegiline delayed the need for levo-

dopa, its effect on disease progression has not been proved.

As for rasagiline, a placebo (control) study revealed that

patients with early PD who were treated with rasagiline for

12 months, showed less decline in neurological symptoms

(UPDRS score) than patients whose rasagiline treatment

was delayed for 6 months. These clinical results cannot be

explained by the purely symptomatic effect of rasagiline,(67)

and they may represent a neuroprotective effect.

Insights into the neuroprotection mechanism:

lack of correlation between neuroprotection

and MAO-B inhibition

The neuroprotective effect of propargylamines, demonstrated

in different insult models, both in vitro and in vivo, suggests that

it is not due to MAO inhibition (Tables 2 and 3). The finding that

propargylamines induce neuroprotection in neuronal cell

cultures lacking MAO-B on the one hand, and the absence

of any neuroprotective effect by clorgyline, a MAO-A inhibitor,

on the other, also support the lack of correlation between

neuroprotection and MAO inhibition.(57) Furthermore, the

propargylamine concentration required for neuroprotection

both in vitro and in vivo is 1 mM, whereas the IC50 or Ki value for

in vitro MAO-B inhibition is in the range of 6–30 nM. As shown

in Tables 2 and 3, the major insults evoking neurotoxicity and,

in many cases, leading to neuronal apoptosis, are depolariza-

tion, calcium overload and oxidative stress. Conceptually,

therefore, the neuroprotective effect of propargylamines might

be grounded in the prevention of cell death or the activation of

pro-survival pathways minimizing cellular stress.

Rasagiline and selegiline exert their antioxidant effect

by upregulating the enzymes involved.(68) Selegiline(69) and

rasagiline,(70) promote free radical scavenging by activating

superoxide dismutase (SOD1 and SOD2) and catalase or by

increasing the SOD protein level,(71) upon chronic adminis-

tration to rats. In vitro experiments in our laboratory using cyclic

voltammetry further substantiated the antioxidant properties of

these compounds. Selegiline displays antioxidant effects

in vivo, which are probably not related to MAO-B inhibition

since selegiline reduced the free radical level at a pM con-

centrations too low to inhibit MAO-B activity.(72) The neuro-

protective effect of selegiline could also be mediated indirectly

via antagonistic modulation of the polyamine’s binding site on

the NMDA receptor, thereby reducing NMDA-receptor-gener-

ated excitotoxicity.(73) Moreover, rasagiline and selegiline act

directly as anti-apoptotic agents, probably by interfering with

the cellular apoptotic cascade (pro-survival) (Table 2). Several

studies have shown that treatment of cells with rasagiline

causes upregulation of putative anti-apoptotic and antioxidant

proteins such as Bcl-2, SOD, glutathione and BCLXL.(74) It has

been suggested that rasagiline protects cells from apoptosis

by stabilizing the mitochondrial membrane potential(75,76)

since such stabilization showed a causal relationship to inhibi-

tion of caspase activity and prevention of DNA fragmentation.

Table 3. Neuroprotective in vivo effects of MAO-B inhibitors selegiline and rasagiline

MAO-B Inhibitor Insult ModelNeurotoxicitymechanism

Neuroprotectivemechanism References

Selegiline Permanent or transient

occlusion of cerebral artery

Rat

Mouse

Gerbil

Focal or whole

brain ischemia

Yes 99

108

109

Selegiline Unilateral MPPþ neurotoxicity Sprague-Dawley rat Oxidative stress Yes 110

Independent of MAO-B

Rasagiline, Selegiline MPTP Monkey Oxidative stress Yes 111

Rasagiline Permanent middle cerebral

artery occlusion

Wistar rat Brain ischemia Yes Independent of MAO-B 62

Rasagiline Closed head injury Mouse Mechanical brain

trauma

Yes 48

Challenges

BioEssays 26.1 85

In any event, the propargyl group of rasagiline and selegiline

is essential to MAO-B inhibition and the anti-apoptotic

effect.(77)

Molecular neuroprotective mechanism

of MAO-B inhibitors

The precise target and details of the molecular mechanisms

involved in the neuroprotective effect exerted by propargy-

lamines are not known. A major obstacle in mechanistic

studies on selegiline is that the molecule is largely (/80%)

converted into amphetamines, some of which display neuro-

toxic activity.(78) Cumulative evidence suggests that brain

injury following amphetamine and methamphetamine admin-

istration is due to increased free radical formation and mito-

chondrial damage, leading to a failure in cellular energy

metabolism followed by secondary excitotoxicity-induced

seizures.(78) The neuroprotective activity of selegiline derives

from a minor metabolite, desmethylselegiline (DMS), which

was shown to be neuroprotective in vitro.(79) However, the

amphetamine-like major metabolites may cause cell da-

mage(80) and thus could reverse the potential beneficial

effects of DMS. Indeed, the addition of L-methamphetamine

to oxygen–glucose-deprived PC-12 cells substantially in-

creased cell death.(57) When both L-methamphetamine and

selegiline were added to PC-12 cells exposed to OGD, the

protective effect of the latter was reduced.(57) In contrast to

selegiline, whose neuroprotective activity derives from its

minor metabolite, DMS, that of rasagiline is mediated solely by

the parent drug(57,81) and its major metabolite, 1-R-amino-

indan, is not toxic.(57)

An interesting attempt to identify the target of the propargy-

lamine anti-apoptotic effect was made by Kragten et al. in

1998.(82) They synthesized a selegiline derivative (CGP 3466)

that is neuroprotective but does not inhibit MAO and is not

metabolized to amphetamines. Using affinity binding, labeling

and BIAcore technology (in which CGP 3466 was covalently

bound to the surface of a flow cell with a CM5 sensor chip(82)),

they detected in rat brain lysates four protein bands of

38,43,50 and 299 kDa, all putative endogeneous binding sites

for CGP 3466. One of these proteins (38 kDa) was identified

as glyceraldehyde-3-phosphate dehydrogenase (GAPDH).

Using GAPDH antisense oligonucleotides, the investigators

established the importance of this enzyme in mediating the

neuroprotective effect of CGP 3466. Although the precise

role of GAPDH in neuronal apoptosis and in the neuroprotec-

tive mechanism of propargylamines awaits further investiga-

tion, GAPDH is the first selective target reported for these

compounds. GAPDH binds nicotinamide adenine dinucleotide

(NADþ)(82) at a site structurally similar to that of the flavin

adenine dinucleotide (FAD)-binding enzymes, to which MAO-

B belongs. Therefore, we propose that propargylamines bind

to GAPDH or other proteins with a tertiary structure similar to

that of the active site of MAO-B, thereby initiating an anti-

apoptotic process in the cells. It is very tempting to suggest

that the active site of these proteins possesses an FAD-

binding domain such as in MAO-B.(83) Recently, the structure

of human MAO-B was determined by crystalization at 3A

resolution. Electron density analysis revealed that pargyline, a

selegiline analog, binds covalently to N5 of the flavine

nucleotide.(83) Therefore, it is conceivable that the insertion

of propargylamines into the FAD pocket(83) of certain proteins

with a structure similar to that of MAO-B will be followed by

covalent binding to FAD or NAD, as in MAO-B, resulting in

neuroprotection.

Hydrogen peroxide (H2O2), which is a by-product of the

enzymatic reaction catalyzed by MAO-B, has been shown to

inhibit GAPDH. This inactivates the glycolytic and mitochon-

drial pathways of ADP phosphorylation, resulting in a drop in

intracellular ATP level and cell death. It was shown that MAO-

B inhibition by propargylamines confers neuroprotection

from H2O2-induced injury due to GAPDH damage.(84) It

was recently demonstrated that the level of GAPDH and its

nuclear accumulation increase in apoptotic models of serum

withdrawal, growth factor deprivation and MPPþ-induced

neurotoxicity, pointing to the enzyme’s essential role in apop-

tosis.(85,86) Furthermore, propargylamines prevented both

these phenomena, suggesting that this activity rather than

MAO-B inhibition, may contribute to the neuroprotective

effect.(85) Another important finding was the subcellular redu-

ction in GAPDH activity in Alzheimer’s disease and Hunting-

ton’s disease.(87) The investigators posited that intracellular

formation of complexes between GAPDH and proteins chara-

cteristic of neurodegenerative disorders, such as a-synuclein

in PD, b-amyloid in Alzheimer’s or huntingtin in Huntington’s,

may represent an emerging cellular phenotype of neurode-

generative disorders The major challenges in this field will be

to further clarify the apoptotic role of GAPDH, the pathophy-

siological meaning of GAPDH complexes, to identify novel

anti-apoptotic and/or pro-survival FAD- or NAD-binding

proteins and to delineate the interaction between these novel

enzymes and propargylamines in neuroprotection.

Neuroprotection strategy perspectives

In the present review, we have focused on the neuroprotection

of MAO-B inhibitors, their chemical characterization and their

cellular and molecular mechanisms of action. It is impossible in

so brief a review to cite all the earlier work recently summarized

by Magyar and Vizi.(88) We have included a few examples to

illustrate the newly emerging apoptotic concept of dopami-

nergic neuronal degeneration and have described selected

studies on the neuroprotective contribution of MAO-B inhi-

bitors such as propargylamines. The majority of in vitro and

in vivo PD studies attribute the neurodegeneration of SN

dopaminergic neurons to a long series of insults such as:

mitochondrial impairment and oxidative stress, Ca2þoverload,

iron metabolism, neurotoxins, excitotoxicity, neurotrophins,

Challenges

86 BioEssays 26.1

oxygen and glucose deprivation, depolarization and inflam-

mation. It is conceivable that the neuroprotective effect

of propargylamines is due to a shift in the neurotoxicity-

neuroprotection balance towards survival (Fig. 1). MAO-B,

GAPDH and other FAD/NAD-binding proteins represent

pivotal targets of the balance, as reflected by the neuropro-

tective effect of selegiline and rasagiline. Therefore, a pharma-

ceutical approach towards neuroprotection would be to

synthesize novel MAO-B/GAPDH inhibitors. Such inhibitors

should help to launch mechanistic studies aimed at elucidating

the molecular interactions of these compounds in anti-apo-

ptotic and/or survival signaling pathways. A crucial need in PD

drug development is the establishment of new in vitro and

in vivo models of the disease. Once these models are avail-

able, it will be necessary to identify neuronal dopaminergic

apoptotic pathways, as well as survival pathways. One of the

most remarkable changes recently observed in the nigro-

striatal region of the PD brain is the decreased level of neuro-

trophins supporting dopaminergic neuron survival, such

as brain-derived neurotrophic factor (BDNF), NGF(89) and

glial-derived neurotrophic factor (GDNF).(90) Therefore, re-

combinant neurotrophins or compounds increasing the local

production of neurotrophins,(91) and/or neurotrophin receptor-

agonists could prove beneficial in the treatment of this

disorder. Development of MAO-B inhibitors with neurotro-

phin-like activity may be a future strategy in PD therapy. This

possibility is supported by the recent finding that rasagiline

increases GDNF production and release by neuroblastoma

cells.(92) Despite the uncertainties and difficulties attendant

upon neuroprotective therapy in PD, clinical trials with innovat-

ive neuroprotective agents must proceed for neuroprotection

to become a reality.

In addition to the use of low molecular weight drugs for PD

treatment, modern therapeutic tactics focus on cell and gene

therapy. Cell therapy includes implantation in the SN of em-

bryonic or neuronal stem cells, which can synthesize and

release dopamine to correct the PD syndrome.(93) Gene

therapy in PD should aim both at supplementing the low SN

dopamine level by introducing the genes encoding dopamine-

synthesizing enzymes into non-dopaminergic cells in the

striatum, and at supporting the survival of dopaminergic

neurons by preventing apoptosis through the introduction of

genes blocking this process.(94)

We anticipate that, in coming years, the cellular pathophy-

siology of Parkinson’s disease will be clarified, allowing the

design of new drugs and novel therapeutic approaches.

Figure 1. Schematic depicting the balance be-

tween neurotoxicity and neuroprotection in Par-

kinson’s disease. Upper part: Normal population of

SN dopaminergic neurons, consisting mainly of

live neurons (blue) and a few dead neurons (red).

Enzymes (green) (targeted by propargylamines)

pivot the balance between cell survival (#) and cell

death ("). Insults shift the balance towards

neurotoxicity; rasagiline and selegiline tip the

balance towards neuroprotection.

Challenges

BioEssays 26.1 87

References1. Sethi KD. Clinical aspects of Parkinson’s disease. Curr Opin Neurol

2002;15:457–460.

2. Hornykiewicz O, Kish SJ. Biochemical pathophysiology of Parkinson’s

disease. Adv Neurol 1987;45:19–34.

3. Jellinger K. Pathology of Parkinson’s syndrome. In: Calne DB, editor.

Handbook of Experimental Pharmacology. Berlin: Springer-Verlag; 1989.

p 47–112.

4. Bernheimer H, Birkmayer W, Hornykiewicz O, Jellinger K, Seitelberger F.

Brain dopamine and the syndromes of Parkinson and Huntington. Clini-

cal, morphological and neurochemical correlations. J Neurol Sci 1973;

20:415–455.

5. Riess O, Kuhn W, Kruger R. Genetic influence on the development of

Parkinson’s disease. J Neurol 2000;247:69–74.

6. Yahr MD, Bergman KJ. Advances in Neurology. New York: Raven Press;

1986. p 317–321.

7. Tanner CM. The role of environmental toxins in the etiology of Parkinson’s

disease. TINS 1989;12:49–54.

8. Fahn S, Cohen G. The oxidant stress hypothesis in Parkinson’s disease.

Evidence supporting it. Ann Neurol 1992;32:804–812.

9. Gerlach M, Riederer P, Youdim MBH. Molecular mechanism for

neurodegeneration: synergism between reactive oxygen species,

calcium and excitotoxic amino acids. In: Battistin L, Scarlato G, Caraceni

T, Ruggieri S, editors. Advances in Neurology. Philadelphia: Lippincott-

Raven; 1996. p 177–194.

10. Headley PM, Grillner S. Excitatory amino acid neurotoxicity and synaptic

transmission: the evidence for a physiological function. In: Lodge D,

Collingridge G, editors. The pharmacology of excitatory amino acids: A

TIPS special report. Cambridge: Elsevier; 1990. p 30–36.

11. De Erausquin GA, Costa E, Hanbauer I. Calcium homoestasis, free

radical formation, and trophic factor dependence mechanisms in

Parkinson’s disease. Pharmacol Rev 1994;46:467–482.

12. Beal MF. Ageing, energy, and oxidative stress in neurodegenerative

diseases. Ann Neurol 1995;38:357–366.

13. Gerlach M, Ben-Shachar D, Riederer P, Youdim MBH. Altered brain

metabolism of iron as a cause of neurodegenerative diseases?

J Neurochem 1994;63:793–807.

14. Thompson CB. Apoptosis in the pathogenesis and treatment of disease.

Science 1995;267:1456–1462.

15. Kitada T, Asakawa S, Hattori N, Matsumine H, Yamamura Y, Minoshima

S, Yokochi M, Mizuno Y, Shimizu N. Mutations in the parkin gene cause

autosomal recessive juvenile parkinsonism. Nature 1998;392:605–

608.

16. Leroy E, et al. The ubiquitin pathway in Parkinson’s disease. Nature

1998;392:451–452.

17. Polymeropoulos MH, et al. Mutation in the alpha-synuclein gene

identified in families with Parkinson’s disease. Science 1997;276:2045–

2047.

18. Mandel S, Weinreb O, Youdim MB. Using cDNA microarray to assess

Parkinson’s disease models and the effects of neuroprotective drugs.

TIPS 2003;24:184–191.

19. Momose Y, Murata M, Kobayashi K, Tachikawa M, Nakabayashi Y,

Kanazawa I, Toda T. Association studies of multiple candidate genes for

Parkinson’s disease using single nucleotide polymorphisms. Ann Neurol

2002;51:133–136.

20. Ceballos-Baumann AO. Functional imaging in Parkinson’s disease: acti-

vation studies with PET, fMRI and SPECT. J Neurol 2003;250(Suppl 1):

I15–I23.

21. Mochizuki H, Goto K, Mon H, Mizuno Y. Histochemical detection of

apoptosis in Parkinson’s disease. J Neurol Sci 1996;137:120–123.

22. Oppenheim RW, Flavell RA, Vinsant S, Prevette D, Kuan CY, Rakic P.

Programmed cell death of developing mammalian neurons after genetic

deletion of caspases. J Neurosci 2001;21:4752–4760.

23. Hartmann A, et al. Caspase-3: A vulnerability factor and final effector in

apoptotic death of dopaminergic neurons in Parkinson’s disease. Proc

Natl Acad Sci USA 2000;97:2875–2880.

24. Mogi M, Togari A, Kondo T, Mizuno Y, Komure O, Kuno S, Ichinose H,

Nagatsu T. Caspase activities and tumor necrosis factor receptor R1

(p55) level are elevated in the substantia nigra from parkinsonian brain.

J Neural Transm 2000;107:335–341.

25. Tatton NA. Increased caspase 3 and Bax immunoreactivity accompany

nuclear GAPDH translocation and neuronal apoptosis in Parkinson’s

disease. Exp Neurol 2000;166:29–43.

26. Clarke G, Collins RA, Leavitt BR, Andrews DF, Hayden MR, Lumsden CJ,

McInnes RR. A one-hit model of cell death in inherited neuronal

degenerations. Nature 2000;406:195–199.

27. Wullner U, Kornhuber J, Weller M, Schulz JB, Loschmann PA, Riederer P,

Klockgether T. Cell death and apoptosis regulating proteins in

Parkinson’s disease—a cautionary note. Acta Neuropathol 1999;97:

408–412.

28. Kosel S, Egensperger R, von Eitzen U, Mehraein P, Graeber MB. On the

question of apoptosis in the parkinsonian substantia nigra. Acta

Neuropathol (Berl) 1997;93:105–108.

29. Jellinger KA, Stadelmann CH. The enigma of cell death in neurodegen-

erative disorders. J Neural Transm Suppl 2000;60:21–36.

30. Banati RB, Daniel SE, Blunt SB. Glial pathology but absence of apoptotic

nigral neurons in long-standing Parkinson’s disease. Mov Disord 1998;

13:221–227.

31. Rajput AH, Fenton ME, Birdi S, Macaulay R, George D, Rozdilsky B, Ang

LC, Senthilselvan A, Hornykiewicz O. Clinical-pathological study of

levodopa complications. Mov Disord 2002;17:289–296.

32. Hristova AH, Koller WC. Early Parkinson’s disease: what is the best

approach to treatment. Drugs Aging 2000;17:165–181.

33. Yu PH, Tipton KF. Deuterium isotope effect of phenelzine on the inhibition

of rat liver mitochondrial monoamine oxidase activity. Biochem Pharma-

col 1989;38:4245–4251.

34. Davis DS, Tasuhara H, Boobis AR, George CF. The effects of reversible

inhibitors of monoamine oxidase on tyramine deamination in dog intes-

tine. In: Tipton KF, Dostert P, Strolin-Benedetti M, editors. Monoamine

Oxidase and Disease. London: Academic Press; 1984. p 443–448.

35. Hasan F, McCrodden JM, Kennedy NP, Tipton KF. The involvement of

intestinal monoamine oxidase in the transport and metabolism of

tyramine. J Neural Transm 1988;26(Suppl):1–9.

36. Blackwell B. Hypertensive crisis due to monoamine oxidase inhibitors.

Lancet 1963;ii:849–851.

37. Blackwell B, Marley E, Price J. Hypertensive interactions between

monoamine oxidase inhibitors and food stuffs. Br J Psychiatry 1967;113:

349–365.

38. Johnston JP. Some observations upon a new inhibitor of monoamine

oxidase in the brain tissue. Biochem Pharmacol 1968;17:1285–1297.

39. Parkinson Study Group. Effects of deprenyl on the progression of

disability in early Parkinson’s disease. New Engl J Med 1989;321:1364–

1371.

40. Parkinson Study Group. A controlled trial of rasagiline in early Parkinson’s

disease: the TEMPO Study. Arch Neurol 2002;59:1937–1943.

41. Bach AW, Lan NC, Johnson DL, Abell CW, Bembenek ME, Kwan SW,

Seeburg PH, Shih JC. cDNA cloning of human liver monoamine oxidase

A and B: molecular basis of differences in enzymatic properties. Proc

Natl Acad Sci USA 1988;85:4934–4938.

42. Hsu YP, Weyler W, Chen S, Sims KB, Rinehart WB, Utterback MC, Powell

JF, Breakefield XO. Structural features of human monoamine oxidase A

elucidated from cDNA and peptide sequences. J Neurochem 1988;51:

1321–1324.

43. Chen ZY, et al. Structure of the human gene for monoamine oxidase type

A. Nucleic Acids Res 1991;19:4537–4541.

44. Shih JC, Chen K, Ridd MJ. Monoamine Oxidase: From Genes to

Behavior. Annu Rev Neurosci 1999;22:197–217.

45. Hall TR, Uruena G. Pharmacology and physiology of monoamine oxidase

activity in vertebrates—a comparative study. Comp Biochem Physiol B

1983;76:393–397.

46. Parkinson Study Group. Effect of lazabemide on the progression of

disability in early Parkinson’s disease. Ann Neurol 1996;40:99–107.

47. Birkmayer W, Riederer P, Youdim MB, Linauer W. The potentiation of the

anti-akinetic effect after L-dopa treatment by an inhibitor of MAO-B,

Deprenyl. J Neural Transm 1975;36:303–326.

48. Finberg JP, Lamensdorf I, Commissiong JW, Youdim MB. Pharmacology

and neuroprotective properties of rasagiline. J Neural Transm 1996;48:

95–101.

49. Youdim MBH, Wadia A, Tatton W, Weinstock M. The anti-Parkinson

drug rasagiline and its cholinesterase inhibitor derivatives exert

Challenges

88 BioEssays 26.1

neuroprotection unrelated to MAO inhibition in cell culture and in vivo.

Ann NY Acad Sci 2001;939:450–458.

50. Speiser Z, Levy R, Cohen S. Effects of N-propargyl-1-(R)aminoindan

(rasagiline) in models of motor and cognition disorders. J Neural Transm

1998;Suppl 52:287–300.

51. Youdim MB, Finberg JP. Pharmacological actions of l-deprenyl (selegi-

line) and other selective monoamine oxidase B inhibitors. Clin Pharmacol

Ther 1994;56:725–733.

52. Chen L, He M, Sibille E, Thompson A, Sarnyai Z, Baker H, Shippenberg

T, Toth M. Adaptive changes in postsynaptic dopamine receptors

despite unaltered dopamine dynamics in mice lacking monoamine

oxidase B J Neurochem 1999;73:647–655.

53. Garrick NA, Murphy DL. Species differences in the deamination of

dopamine and other substrates for monoamine oxidase in brain.

Psychopharmacology 1980;72:27–33.

54. Lenders JWM, et al. Specific genetic deficiencies of the A and B

isoenzymes of monoamine oxidase are characterized by distinct neuro-

chemical and clinical phenotypes. J Clin Invest 1996;97:1010–1019.

55. Finnegan KT, Skratt JJ, Irwin I, DeLanney LE, Langston JW. Protection

against DSP-4-induced neurotoxicity by deprenyl is not related to its

inhibition of MAO B. Eur J Pharmacol 1990;184:119–126.

56. Yu PH, Davis BA, Fang J, Boulton AA. Neuroprotective effects of some

monoamine oxidase-B inhibitors against DSP-4-induced noradrenaline

depletion in the mouse hippocampus. J Neurochem 1994;62:697–704.

57. Abu-Raya S, Tabakman R, Blaugrund E, Trembovler V, Lazarovici P.

Neuroprotective and neurotoxic effects of monoamine oxidase-B

inhibitors and derived metabolites under ischemia in PC12 cells. Eur J

Pharmacol 2002;434:109–116.

58. Goggi J, Theofilopoulos S, Riaz SS, Jauniaux E, Stern GM, Bradford HF.

The neuronal survival effects of rasagiline and deprenyl on fetal human

and rat ventral mesencephalic neurons in culture. Neuroreport

2000;11:3937–3941.

59. Yoles E, Shwartz M. N-Propargyl-1 (R)-aminoindan (TVP-1012), a

putative neuroprotective agent, enhances in vitro neuronal survival after

glutamate toxicity. Proc Am Soc Neurosci 1995;21:562.

60. Abu-Raya S, Blaugrund E, Trembovler V, Shilderman-Bloch E, Shohami

E, Lazarovici P. Rasagiline, a monoamine oxidase-B inhibitor, protects

NGF-differentiated PC12 cells against oxygen-glucose deprivation.

J Neurosci Res 1999;58:456–463.

61. Speiser Z, Katzir O, Rehavi M, Zabarski T, Cohen S. Sparing by

rasagiline (TVP-1012) of cholinergic functions and behavior in the

postnatal anoxia rat. Pharmacol Biochem Behav 1998;60:387–393.

62. Speiser Z, Mayk A, Eliash S, Cohen S. Studies with rasagiline, a MAO-B

inhibitor, in experimental focal ischemia in the rat. J Neural Transm

1999;106:593–606.

63. Huang W, Chen Y, Shohami E, Weinstock M. Neuroprotective effect of

rasagiline, a selective monoamine oxidase-B inhibitor, against closed

head injury in the mouse. Eur J Pharmacol 1999;366:127–135.

64. Knoll J, Miklya I. Enhanced catecholaminergic and serotoninergic

activity in rat brain from weaning to sexual maturity: rationale for

prophylactic (-)deprenyl (selegiline) medication. Life Sci 1995;56:611–

620.

65. Knoll J. The pharmacology of selegiline ((-)deprenyl). New aspects. Acta

Neurol Scand Suppl 1989;126:83–91.

66. Parkinson Study Group. Effect of tocopherol and deprenyl on the

progression of disability in early Parkinson’s disease. New Engl J Med

1989;328:176–183.

67. Siderowf A, and the Parkinson Study Group. Earlier treatment with

rasagiline may attenuate (UPDRS) progression of PD. Movement

Disorders 2001;16:981.

68. Kitani K, Kanai S, Ivy GO, Carrillo MC. Pharmacological modifications of

endogenous antioxidant enzymes with special reference to the effects of

deprenyl: a possible antioxidant strategy. Mech Ageing Dev 1999;111:

211–221.

69. Kitani K, Miyasaka K, Kanai S, Carrillo MC, Ivy GO. Upregulation of

antioxidant enzyme activities by deprenyl. Implications for life span

extension. Ann N Y Acad Sci 1996;786:391–409.

70. Carrillo MC, Kanai S, Nokubo M, Kitani K. (-) deprenyl induces activities

of both superoxide dismutase and catalase but not of glutathione

peroxidase in the striatum of young male rats. Life Sci 1991;48:517–521.

71. Lai CT, Zuo DM, Yu PH. Is brain superoxide dismutase activity increased

following chronic treatment with 1-deprenyl? J Neural Transm Suppl

1994;41:221–229.

72. Wu RM, Chiuech CC, Pert A, Murphy DL. Apparent antioxidant effect of

L-deprenyl on hydroxyl radical formation and nigral injury elicited by

MPPþ in vivo. Eur J Pharmacol 1993;243:241–248.

73. Gerlach M, Desser H, Youdim MB, Riederer P. New horizons in molecular

mechanisms underlying Parkinson’s disease and in our understanding of

the neuroprotective effects of selegiline. J Neural Transm Suppl 1996;48:

7–21.

74. Naoi M, Maruyama W, Youdium MBH. Molecular mechanism underlying

neuroprotection by rasagiline. J Neurochem 2002;81:4.

75. Maruyama W, Yamamoto T, Kitani K, Carrillo MC, Youdim M, Naoi M.

Mechanism underlying anti-apoptotic activity of a (-) deprenyl-related

propargylamine, rasagiline. Mech Ageing Dev 2000;116:181–191.

76. Maruyama W, Takahashi T, Youdim M, Naoi M. The anti-Parkinson drug,

rasagiline, prevents apoptotic DNA damage induced by peroxynitrite in

human dopaminergic neuroblastoma SH-SY5Y cells. J Neural Transm

2002;109:467–481.

77. Maruyama W, Naoi M. Neuroprotection by (-)-deprenyl and related

compounds. Mech Ageing Dev 1999;111:189–200.

78. Bowyer JF, Peterson SL. Neuronal degeneration in the forebrain

produced by amphetamine, methamphetamine and fenfluramine. In:

Lester DS, Slikker W Jr, Lazarovici P, editors. Cellular and Molecular

Mechanisms of Toxin Action: Site-Selective Neurotoxicity. London: Taylor

and Francis; 2002. p 207–232.

79. Mytilineou C, Radcliffe PM, Olanow CW. L-(-)-desmethylselegiline, a

metabolite of selegiline [L-(-)-deprenyl], protects mesencephalic dopa-

mine neurons from excitoxicity in vitro. J Neurochem 1997;68:434–

436.

80. Oh C, Murray B, Bhattacharya N, Holland D, Tatton WG. (-)-Deprenyl

alters the survival of adult murine facial motoneurons after axotomy:

increases in vulnerable C57BL strain but decreases in motor neuron

degeneration mutants. J Neurosci Res 1994;38:64–74.

81. Maruyama W, Youdim MBH, Naoi M. Antiapoptotic properties of

rasagiline N-propargylamine-1(R)-aminoindan, and its optical (S)-iso-

mer, TV1022. Ann NY Acad Sci 2001;939:320–329.

82. Kragten E, Lalande I, Zimmermann K, Roggo S, Schindler P, Muller D,

Van Oostrum J, Waldmeier P, Furst P. Glyceraldehyde-3-phosphate

dehydrogenase, the putative target of the anti-apoptotic compounds

CGP 3466 and R-(-)-deprenyl. J Biol Chem 1998;273:5821–5828.

83. Binda C, Newton-Vinson P, Hubalek F, Edmondson DE, Mattevi A.

Structure of human monoamine oxidase B, a drug target for the treatment

of neurological disorders. Nat Struct Biol 2002;9:22–26.

84. Hyslop PA, Hinshaw DB, Halsey WA Jr, Schraufstatter IU, Sauerheber

RD, Spragg RG, Jackson JH, Cochrane CG. Mechanisms of oxidant-

mediated cell injury. The glycolytic and mitochondrial pathways of ADP

phosphorylation are major intracellular targets inactivated by hydrogen

peroxide. J Biol Chem 1988;263:1665–1675.

85. Carlile GW, Chalmers-Redman RM, Tatton NA, Pong A, Borden KE,

Tatton WG. Reduced apoptosis after nerve growth factor and serum

withdrawal: conversion of tetrameric glyceraldehyde-3-phosphate dehy-

drogenase to a dimer. Mol Pharmacol 2000;57:2–12.

86. Fukuhara Y, Takeshima T, Kashiwaya Y, Shimoda K, Ishitani R,

Nakashima K. GAPDH knockdown rescues mesencephalic dopaminer-

gic neurons from MPPþ-induced apoptosis. Neuroreport 2001;12: 2049–

2052.

87. Mazzola JL, Sirover MA. Alteration of intracellular structure and function

of glyceraldehyde-3-phosphate dehydrogenase: a common phenotype

of neurodegenerative disorders? Neurotoxicology 2002;23:603–609.

88. Magyar K, Vizi ES. Milestones in monoamine oxidase research: dis-

covery of (-)-deprenyl. Budapest: Medicina Publishing House Co; 2000.

89. Mogi M, Togari A, Kondo T, Mizuno Y, Komure O, Kuno S, Ichinose H,

Nagatsu T. Brain-derived growth factor and nerve growth factor

concentrations are decreased in the substantia nigra in Parkinson’s

disease. Neurosci Lett 1999;270:45–48.

90. Brundin P. GDNF treatment in Parkinson’s disease: time for controlled

clinical trials? Brain 2002;125:2149–2151.

91. Yamaguchi K, Sasano A, Urakami T, Tsuji T, Kondo K. Stimulation of

nerve growth factor production by pyrroloquinoline quinone and its

Challenges

BioEssays 26.1 89

derivatives in vitro and in vivo. Biosci Biotechnol Biochem 1993;57:

1231–1233.

92. Maruyama W. Neuroprotection by propargylamines in PD Suppression of

apoptosis and induction of prosurvival genes Neurotoxicol and Teratol

2002;4:675–682.

93. Zigova T, Snyder EY, Sanberg PR. Neural Stem Cells for Brain and Spinal

Cord Repair. Totowa: Humana Press; 2003.

94. Nagatsu T. Parkinson’s disease: changes in apoptosis-related factors

suggesting possible gene therapy. J Neural Transm 2002;109:731–745.

95. Bentue-Ferrer D, Menard G, Allain H. Monoamine oxidase B inhibitors:

current status and future potential. CNS drugs 1996;6:217–236.

96. Perez V, Marco JL, Fernandez-Alvarez E, Unzeta M. Kinetic studies of N-

allenic analogues of tryptamine as monoamine oxidase inhibitors. J

Pharm Pharmacol 1996;48:718–722.

97. Zhou JP, Li J, Upadhyaya S, Eaton PE, Silverman RB. 4-substituted

cubylcarbinylamines: a new class of mechanism-based monoamine

oxidase B inactivators. J Med Chem 1997;40:1165–1168.

98. Mytilineou C, Radcliffe PM, Olanow CW. L-(-)-Desmethylselegiline, a

metabolite of selegiline [L-(-)-Deprenyl], protects mesencephalic dopa-

mine neurons from excitotoxicity in vitro. J Neurochem 1997;68:434–436.

99. Semkova I, Wolz P, Schilling M, Krieglstein J. Selegiline enhances NGF

synthesis and protects central nervous system neurons from excitotoxic

and ischemic damage. Eur J Pharmacol 1996;315:19–30.

100. Koutsilieri E, Chen TS, Rausch WD, Riederer P. Selegiline is neuropro-

tective in primary brain cultures treated with 1-methyl-4-phenylpyridi-

nium. Europ J Pharmacol 1996;306:181–186.

101. Mytilineou C, Leonardi EK, Radcliffe P, Heinonen EH, Han SK, Werner P,

Cohen G, Olanow CW. Deprenyl and desmethylselegiline protect

mesencephalic neurons from toxicity induced by glutathione depletion.

J Pharmacol Exp Therap 1998;284:700–706.

102. Tatton WG, Ju WY, Holland DP, Tai C, Kwan M. (-)-Deprenyl reduces

PC12 cell apoptosis by inducing new protein synthesis. J Neurochem

1994;63:1572–1575.

103. Finberg JP, Takeshima T, Johnston JM, Commissiong JW. Increased

survival of dopaminergic neurons by rasagiline, a monoamine oxidase B

inhibitor. NeuroReport 1998;9:703–707.

104. Roy E, Bedard PJ. Deprenyl increases survival of rat foetal nigral neurons

in culture. NeuroReport 1993;4:1183–1186.

105. Xu L, Ma J, Seigel GM, Ma J. L-deprenyl, blocking apoptosis and

regulating gene expression in cultured retinal neurons. Biochem

Pharmacol 1999;58:1183–1190.

106. Maruyama W, Akao Y, Youdim MB, Naoi M. Neurotoxins induce

apoptosis in dopamine neurons: protection by N-propargylamine-1(R)-

and (S)-aminoindan, rasagiline and TV1022. In: Riederer P, Calne DB,

Horowski R, Mizuno Y, Olanow CW, Poewe W, Youdim MBH, editors.

Advances in Research on Neurodegeneration. Wien: Springer-Verlag;

2000. p 171–186.

107. Maruyama W, Akao Y, Youdim MB, Davis BA, Naoi M. Transfection-

enforced Bcl-2 overexpression and an anti-Parkinson drug, rasagiline,

prevent nuclear accumulation of glyceraldehyde-3-phosphate dehydro-

genase induced by an endogenous dopaminergic neurotoxin, N-methyl

(R) salsolinol. J Neurochem 2001;78:727–735.

108. Lahtinen H, Koistinaho J, Kauppinen R, Haapalinna A, Keinanen R,

Sivenius J. Selegiline treatment after transient global ischemia in gerbils

enhances the survival of CA1 pyramidal cells in the hippocampus. Brain

Res 1997;757:260–267.

109. Knollema S, Aukema W, Hom H, Korf J, Ter Horst GJ. L-deprenyl

reduces brain damage in rats exposed to transient hypoxia-ischemia.

Stroke 1995;26:1883–1887.

110. Wu RM, Chen RC, Chiueh CC. Effect of MAO-B inhibitors on MPPþ

toxicity in vivo. Ann NY Acad Sci 2000;899:255–261.

111. Kupsch A, Sautter J, Gotz ME, Breithaupt W, Schwarz J, Youdim MBH,

Riederer P, Gerlach M, Oertel WH. Monoamine oxidase-inhibition and

MPTP-induced neurotoxicity in the non-human primate: comparison of

rasagiline (TVP 1012) with selegiline. J Neural Transm 2001;108:985–

1009.

Challenges

90 BioEssays 26.1