REVIEW ARTICLE | OPEN ACCESS DOI: 10.23937/2469-5866/1410029

The Role of Prolyl Oligopeptidase in Microtubule-Associated Processes and Cognitive Impairment

Nuria Trallero#, Ariadna Anunciación-Llunell#, Roger Prades and Teresa Tarragó*

Iproteos S.L., Barcelona, Spain

Abstract

Prolyl oligopeptidase (POP) is a cytosolic serine protease with prominent expression in the brain. Inhibition of this enzyme leads to cognition-enhancing and neuroprotective effects in animal models with cognitive deficits. However, the biological function of POP remains unknown. Although in the past it was though that its catalytic activity was responsible for its physiological role, lately it has been hypothesized that POP is involved in the inositol pathway and that it interacts with several proteins, including α-tubulin, thereby implying that its function may be related to protein-protein interactions. In this review, we analyze the destabilization of microtubules in neurological diseases such as schizophrenia, Parkinson's, Alzheimer's and Huntington's disease. Given the interaction of POP with α-tubulin, we discuss the relevance of this protease in the modulation of synaptic processes. In this context, we also examine the potential of POP as a promising target for the treatment of cognitive impairment.

Keywords

Microtubules, Neurological disease, Prolyl oligopeptidase, Protein-protein interaction, Synapsis

Prolyl Oligopeptidase Overview

Prolyl oligopeptidase (POP, EC3.4.21.26), also known as prolyl endopeptidase, is a cytosolic serine protease (81 kDa) that hydrolyzes post-proline bonds of small peptides (less than 30 residues long). It was discovered in the human uterus in the early 1970s [1] and was first described as an oxytocin-hydrolyzing enzyme and further characterized as a peptidase that cleaves small substrates involved in learning and mnemonic processes, such as substance P, neurotensin, arginine-vasopressin and thyrotropin-releasing hormone, among others [2]. Some studies described that the administration of a POP inhibitor increases the levels of substance P and arginine-vasopressin in the hippocampus and frontal cortex [3]. However, it has been proposed that the hydrolytic activity of POP does not drive its physiological role, which remains elusive. In this regard, given that POP interacts with other proteins and it modulates the inositol triphosphate pathway, several alternative hypotheses have been put forward to explain POP function [4]. Despite further research is needed to elucidate the exact role of this enzyme in vivo, in this review, we focus on the known POP interactors and their relevance in POP function Regarding structure, POP is a monomeric protein that has an overall cylindrical shape formed by two domains. The catalytic domain is a typical α/β-hydrolase, whereas the structural domain is a seven-bladed β-propeller, which acts as an empty cylinder, restricting the size and orientation of the substrate [5,6] (Figure 1). Although the enzymatic mechanism of POP is still not fully understood, electron microscopy, NMR and X-ray crystallography studies have unveiled a state of equilibrium between open and closed conformations of this enzyme in solution. This equilibrium can be modulated by direct active site POP inhibitors [5,7].

Figure 1: POP conformations in solution. A) Closed conformation of porcine POP (PDB ID: 1QFS) covalently bound to inhibitor Z-pro-prolinal (ZPP), the first POP inhibitor to be designed. β-propeller is shown in blue, α/β-hydrolase domain in orange and ZPP in green; B) Open conformation of Aemonas punctate POP (PDB ID: 3IUJ). View Figure 1

POP is ubiquitously expressed in the human body, but there is an increased concentration in the central nervous system (CNS) [8,9]. In this regard, this enzyme is expressed in cortical and hippocampal glutamatergic neurons, in ɣ-aminobutyric acid (GABA) ergic and cholinergic interneurons of the thalamus and cortex, and in Purkinje cells [10,11]. Moreover, after an inflammatory insult, POP expression is elevated in microglia and astrocytes [12].

Given the expression of POP in the CNS, its involvement in neurodegenerative and neuropsychiatric diseases has been addressed. Several studies report that POP activity is altered in patients with Alzheimer's disease (AD), Lewy body dementia, Parkinson's disease (PD), Huntington's disease (HD), and schizophrenia, among others [9]. POP activity has also been found to be altered in the serum of patients with mood disorders, such as depression and bipolar disorder [13]. However, care must be taken when interpreting these studies since many of them measured POP in plasma or post-mortem brain samples. In addition, experimental data show that POP inhibition has neuroprotective, anti-amnesic and cognition-enhancing properties in scopolamine-treated rats, whereas it decreases extracellular acetylcholine concentrations in the cortex and hippocampus of these animals [14,15]. The cognition-enhancing properties of POP inhibitors have been further demonstrated in other animal models with cognitive impairment [16-18] and in healthy human volunteers [19-22]. Furthermore, in cortical and cerebellar granule cells, POP inhibitors exert neuroprotective effects [23,24]. Three of these inhibitors, namely S-17092, JTP-4819 and Z-321, reached clinical stages for the treatment of AD [19-22]. However, unfortunately, the development of these drugs did not progress beyond these stages.

Involvement of Prolyl Oligopeptidase in Protein-Protein Interactions

The presence of POP in the cytosol, together with its conformational dynamics and capacity to interact with other proteins, has led to the hypothesis that its biological function is related to protein-protein interactions. This notion was first put forward by Schulz, et al. in 2005, who proposed that POP regulates a number of cell functions independently of its peptidase activity [25]. In addition, it has been observed that POP inhibitors induce changes in the tertiary structure of the enzyme and can modify its interactome. The main POP interactors described in the literature are the cytosolic proteins α-synuclein, glyceraldehyde-3-phosphate dehydrogenase (GAPDH), growth-associated protein 43 (GAP43) and α-tubulin.

α-synuclein

The abnormal fibrillary aggregation of α-synuclein protein in the cytoplasm of neurons and glia is the main characteristic of synucleinopathies, including PD, Lewy body dementia and multiple-system atrophy, among others [26]. The contribution of POP to enhancing α-synuclein aggregation was reported for the first time by Brandt, et al. in 2008 [27]. Since then, it has been shown that POP co-localizes with α-synuclein in the substantia nigra of PD patients [28] and later that the two proteins directly interact. This interaction was revealed through microscale thermophoresis and surface plasmon resonance studies [29]. In addition, POP inhibition reverses α-synuclein aggregation and promotes autophagy-the catabolic pathway used for clearing insoluble α-synuclein oligomers and fibrils [30]. These findings suggest that modulation of the α-synuclein-POP interaction using POP inhibitors could provide a new approach for the treatment of synucleinopathies.

GAPDH

GAPDH is a glycolytic enzyme involved in energy production pathways. However, this protein also participates in other non-glycolytic processes, including programmed cell death [30]. In this regard, it has been shown that GAPDH translocation and accumulation in the nucleus serves as an initiation signal for apoptosis [31,32]. Puttonnen, et al. and some years later Matsuda, et al. reported that POP inhibitors prevent GAPDH translocation into the nucleus under oxidative stress [33,34]. Later on, using co-localization, co-immunoprecipitation and proximity ligation experiments, Matsuda, et al. confirmed the interaction of POP with GAPDH [33]. This information strongly supports the notion that POP mediates GAPDH nuclear translocation and, consequently, that it may be involved in the regulation of apoptosis.

GAP43

Among other CNS-related functions, GAP43 is involved in growth cone formation and axon guidance [35,36], although the underlying molecular mechanism is still not well understood. Valproic acid, carbamazepine and lithium treatment induce an increase in the average spread area of growth cones in sensory neuron cultures, and this increase is reversed by the administration of POP inhibitors [37]. This finding suggests that POP is also related to growth cone development. In this regard, Di Daniel, et al. proposed that this function of POP might be explained by its interaction with GAP43. They demonstrated binding between POP and GAP43 using co-immunoprecipitation and yeast two-hybrid assays [38]. Years later, this study was questioned by Szeltner, et al. who reported partial co-localization of POP and GAP43, without strong physical binding. In this regard, those authors reported a weak and transient interaction between the two proteins, as demonstrated by a glutaraldehyde cross-linking assay [39]. Despite this contradiction, it is plausible that POP participates in growth cone development via direct or indirect interaction with GAP43.

α-tubulin

Repeated assemblies of α- and β-tubulin heterodimers comprise the structure of microtubules (MTs), which, together with microfilaments and intermediate filaments, form the cell cytoskeleton. Therefore, MTs are involved in a large number of cellular processes, including mitosis, cell motility, maintenance of cell shape, neurite growth, intracellular transport and vesicle secretion [40]. Using co-localization and yeast two-hybrid assays, Schulz, et al. demonstrated POP binding to the C-terminus region of α-tubulin [25]. They also showed that POP inhibition promotes protein and peptide release in U-343 cells. These results suggest that POP participates in MT-associated processes, such as intracellular trafficking, protein secretion and axonal transport.

Prolyl Oligopeptidase and Microtubules as Key Players in Synaptic Processes

Since several POP inhibitors have been demonstrated to improve memory and cognition in animals, the possible implication of POP in synaptic function has been studied using alternative approaches.

Firstly, in the same publication that proposed POP interaction with GAP43, the authors also generated POP knockout (KO) mice. These animals were used to demonstrate that the absence of POP causes a decrease in the number of collapsed growth cones and an increase in their spread area in explants when compared with wild-type (WT) mice [38]. Likewise, they observed that WT mouse explants treated with valproic acid, carbamazepine or lithium (all of them considered mood stabilizers) also show a decrease in the number of growth cones. However, treatment with POP inhibitors reversed the effect of these three mood stabilizers. Höfling, et al. also determined that POP KO mice show an increased expression of polysialylated-neural cell adhesion molecule (PSA-NCAM) [41], which has been associated with enhanced brain plasticity [42]. Alternatively, D'Agostino, et al. used POP gene trap mice to study the involvement of POP in synapsis. They observed that these mice present reduced hippocampal long-term potentiation, impaired learning and memory processes related to the hippocampus, decreased spine density in the hippocampus, especially in the CA1 region, and reduced GAP43 expression [43].

Given these observations, it is plausible that POP participates in the modulation of neuronal synapses. Although the molecular mechanisms underlying such modulation are not well understood, the interaction POP-GAP43 and POP-α-tubulin could provide an explanation. Here we outline some basic ideas about α-tubulin and, consequently, MT involvement in synaptic processes and in neurological diseases associated with POP dysregulation.

As mentioned previously, MTs are the major component of the cell cytoskeleton, and as such they are present in all eukaryotic cell types. The key properties of MTs are their dynamic behavior and directionality. The orientation of α- and β-tubulin heterodimers that make up the MT structure is particularly important for the polarity and stability of these cytoskeletal components. The (-) end of a MT has an exposed α-tubulin subunit, whereas the (+) end-where MT elongation takes place-has an exposed β-tubulin. Regarding neuronal MTs, which are characterized by the formation of the bundles required for the growth and maintenance of neurites [40,44], it has been described that axonal MTs are uniformly directed, with (+) ends orientated to axon tips, whereas in dendrites, MTs present mixed orientations (Figure 2). An additional difference between the MTs of axons and dendrites is their predominant microtubule-associated proteins (MAPs). Tau is found mostly in axons, while MAP2 predominates in dendrites. Despite these variations, there is general consensus that MT dynamics play a key role in ensuring the correct function of axons and dendrites, and consequently, in memory, learning and other cognitive processes. Indeed, MTs, in cooperation with other proteins, are involved in the regulation of intracellular cargo trafficking in axons and dendrites, as well as in the modulation of dendritic spine morphology and synaptic plasticity [45,46].

Figure 2: Microtubule distribution along axons and dendrites. MTs are oriented uniformly in axons and Tau protein is the predominant MAP, whereas in dendrites MTs show a mixed orientation and MAP2 protein is bound to them. View Figure 2

Neurodegenerative diseases are commonly characterized by a decrease in MT mass, as well as disrupted polarity patterns and impaired axonal transport, which are caused, in part, by a reduction in MT stability [40,47]. These findings support the notion that the modulation of MT dynamics could provide a potential therapy through which to improve the cognitive impairment associated with neurological diseases. Given the reported interaction between α-tubulin and POP and the implication of these proteins in synaptic processes, it cannot be ruled out that POP, and therefore POP inhibition, regulates MT dynamics (Figure 3). However, the interaction between POP and α-tubulin and the possible implication of this protease in MT stability requires further attention.

Figure 3: Destabilization of microtubules in neurological diseases. The diagram shows the difference in MT conformation between healthy subjects and patients with neurological diseases. The hypothetical recovery of the normal MT structure after POP inhibition is also shown. View Figure 3

Prolyl Oligopeptidase in Neurological Diseases

POP has been proposed to be a promising target for the cognitive impairment present in some CNS diseases, including schizophrenia, PD and AD, among others. This proposal has come about on the basis of three main observations: i) The brains of individuals with these conditions present altered levels of POP [13]; ii) The hypothesized implication of POP in spine density and formation [38,41-43] and; iii) The promising discoveries about molecules that interact with POP [25,29,33,38].

Schizophrenia

According to the WHO, schizophrenia is a neuropsychiatric disease that affects more than 1% of the population worldwide. Individuals with this condition can manifest three types of symptoms: Positive, negative and cognitive. Several antipsychotics for the treatment of positive and negative symptoms are available on the market. However, to date, no drug has been approved to treat cognitive impairment. POP emerged as a promising target to tackle the cognitive symptoms of schizophrenia [48]. Successful results regarding this indication were published by Prades, et al. In that study, three distinct mouse models of schizophrenia-which were based on the administration of subchronic phencyclidine (PCP) or acute dizocilpine (MK801), and on maternal immune activation induced by polyinosinic: Polycytidylic (PIC) during pregnancy-were treated with the POP inhibitor IPR19. Those authors showed that this inhibitor reversed the impaired cognitive symptomatology in a number of cognition tests that evaluate working and visual memory [49]. Therefore, it was concluded that POP inhibition exerts cognition-enhancing effects in animal models of schizophrenia.

Patients with schizophrenia show alterations in dendritic spines. These abnormalities have been observed in multiple brain regions, specifically in layer 3 of the neocortex, where pyramidal cells present a lower density of the smallest spines. It has been hypothesized that these spine deficits appear during early childhood and adolescence, probably as a result of disturbances in molecular mechanisms such as spine formation, pruning, and/or maintenance [50,51]. POP inhibitors may be able to restore dendritic spine deficits and therefore contribute to improving cognitive symptomatology.

Parkinson's disease

PD is a neurodegenerative disorder characterized by loss of neurons in the substantia nigra and an accumulation of Lewy bodies, which comprise aggregated forms of α-synuclein. As previously mentioned, α-synuclein co-localizes and interacts with POP in the substantia nigra [28,29]. Nevertheless, no differences in POP activity or expression between PD and control subjects have been observed [28].

MTs are involved in PD, as demonstrated by observations of the interaction of tubulin with α-synuclein [52]. In addition, α-synuclein has been linked to MT stabilization, although the mechanism of action is not well understood. It has been proposed that α-synuclein is a MAP and that it therefore modulates the stabilization, polymerization and dynamics of MTs. Another hypothesis is that the interaction between α-synuclein and GSK-3β modulates Tau phosphorylation and consequently MT stabilization. Nevertheless, these two hypotheses are not necessarily exclusive [53].

Myöhänen, et al. demonstrated that POP is involved in α-synuclein aggregation. In this regard, those authors showed that POP inhibitor KYP-2047 stimulates the clearance of α-synuclein aggregates by enhancing macroautophagy in the A30P transgenic mouse model of PD [30].

Alzheimer disease

The hallmark of AD is the accumulation of misfolded β-amyloid peptide plaques in the extracellular space, as well as the aggregation of phosphorylated Tau protein in neurons-a process that results in neurofibrillary tangles. The brains of AD patients show altered expression and activity of POP, and this protease has also been observed to co-localize with β-amyloid protein both intra- and extracellularly in this organ [9,29]. Tau protein co-localizes with POP inside cells in AD and control brains [26].

MTs are involved in the development of AD, as demonstrated by their reduced length and density [54], which indicate a possible alteration of synaptic processes. In this regard, in early stages of AD, β-amyloid alters changes in dendritic spine shape, whereas in later stages the cellular response to β-amyloid toxicity leads to a reduction in MT density and length and to the loss of dendritic spines [55].

Huntington's disease

HD is an inherited neurodegenerative disorder characterized by motor, psychiatric and cognitive deficits [56]. Cognitive impairment is one of the earliest symptoms of this disease. HD is caused by an abnormal CAG repeat expansion within exon 1 of the human huntingtin gene (HTT) [57], which encodes the huntingtin protein. Mutated huntingtin forms oligomers and globular intermediates that lead to aggregates, which in turn promote neuronal dysfunction and neurodegeneration [58].

HD is another example of a pathological process in which POP activity is decreased in brain samples of patients respect to controls [59]. Therefore, HD presents abnormalities in MTs and the cytoskeleton, most of these abnormalities being related to MAP dysregulation. Post-mortem HD brain samples present hyperphosphorylated Tau aggregates, a phenomenon observed in AD [60]. Dendritic atrophy induced by the dysregulation of MAP2, probably caused by splicing altered events, has also been reported [61].

Future Perspectives

Patients with neurological disorders, including AD, PD, HD and schizophrenia, present altered POP expression patterns. Moreover, the last two decades have brought about the design and testing of inhibitors of this serine protease for the treatment of the cognitive impairment related to these pathologies. POP inhibitors have widely demonstrated in vitro and in vivo cognition-enhancing and neuroprotective properties. However, although some of these drug candidates reached clinical stages, the biological mechanisms underlying the effects of POP are still unclear. Even though, the catalytic activity of the protease has been traditionally proposed as the function responsible for the biological activity of POP, a new promising hypothesis attribute the in vivo functionalities of this protease to its involvement in key physiologically relevant protein-protein interactions.

One of the most promising POP interactors is α-tubulin, a component of the heterodimer that forms the MT structure. In this regard, several neurological diseases are characterized by alterations in the stability and formation of MTs, as well as deficits in spine density.

On the basis of the findings reviewed herein, POP emerges as potential new target for the treatment of the cognitive impairment associated with some neurological diseases and psychiatric diseases. Research efforts should now be channeled into studying the binding of POP to α-tubulin, the potential involvement of POP in the cytoskeleton and neurite growth, and the effect of POP inhibitors as MT stabilizers.

Notes

N.T., A.A-Ll, R.P. and T.T. are employees of Iproteos, S.L. T.T. is founder of Iproteos, S.L.

Author Contributions

N.T. and A.A-Ll. drafted the manuscript, R.P. and T.T. revised the manuscript. All the contributors revised, gave their approval to the final version of the manuscript, and agreed to be accountable for all aspects of the work.

References

  1. Walter R, Shlank H, Glass JD, Schwartz IL, Kerenyi TD (1971) Leucylglycinamide released from oxytocin by human uterine enzyme. Science 173: 827-829.
  2. Kovács GL, De Wied D (1994) Peptidergic modulation of learning and memory processes. Pharmacol Rev 46: 269-291.
  3. Toide K, Shinoda M, Fujiwara T, Iwamoto Y (1997) Effect of a Novel Prolyl Endopeptidase Inhibitor, JTP-4819, on Spatial Memory and Central Cholinergic Neurons in Aged Rats. Pharmacol Biochem Behav 56: 427-434.
  4. Williams RSB, Eames M, Ryves WJ, Viggars J, Harwood AJ (1999) Loss of a prolyl oligopeptidase confers resistance to lithium by elevation of inositol (1,4,5) trisphosphate. EMBO J 18: 2734-2745.
  5. Millet O, Tarragó T, Gairí M, Kotev M, López A, Giralt E, et al. (2016) Active-Site-Directed Inhibitors of Prolyl Oligopeptidase Abolish Its Conformational Dynamics. Chembiochem 17: 913-917.
  6. Fülöp V, Böcskei Z, Polgar L (1998) Prolyl oligopeptidase: An unusual beta-propeller domain regulates proteolysis. Cell 94: 161-170.
  7. López A, Vilaseca M, Madurga S, Varese M, Tarragó T, et al. (2016) Analyzing slowly exchanging protein conformations by ion mobility mass spectrometry: Study of the dynamic equilibrium of prolyl oligopeptidase. J Mass Spectrom 51: 504-511.
  8. Wilk S (1983) Prolyl endopeptidase. Life Sci 33: 2149-2147.
  9. Goossens F, De Meester I, Vanhoof G, Scharpé S (1996) Distribution of prolyl oligopeptidase in human peripheral tissues and body fluid. Eur J Clin Chem Clin Biochem 34: 17-22.
  10. Myöhänen TT, Kääriäinen TM, Jalkanen AJ, Piltonen M, Männistö PT (2009) Localization of prolyl oligopeptidase in the thalamic and cortical projection neurons: A retrograde neurotracing study in the rat brain. Neurosci Lett 450: 201-205.
  11. Myöhänen TT, Venäläinen JI, Garcia-Horsman JA, Piltonen M, Männistö PT (2008) Cellular and subcellular distribution of rat brain prolyl oligopeptidase and its association with specific neuronal neurotransmitters. J Comp Neurol 507: 1694-1708.
  12. Natunen TA, Gynther M, Rostalski H, Jaako K, Jalkanen AJ (2019) Extracellular prolyl oligopeptidase derived from activated microglia is a potential neuroprotection target. Basic Clin Pharmacol Toxicol 124: 40-49.
  13. García-Horsman JA, Männistö PT, Venäläinen JI (2007) On the role of prolyl oligopeptidase in health and disease. Neuropeptides 41: 1-24.
  14. Yoshimoto T, Kado K, Matsubara F, Koriyama N, Kaneto H, et al. (1987) Specific inhibitors for prolyl endopeptidase and their anti-amnesic effect. J Pharmacobiodyn 10: 730-735.
  15. Jalkanen AJ, Leikas JV, Forsberg MM (2014) Prolyl oligopeptidase inhibition decreases extracellular acetylcholine levels in rat hippocampus and prefrontal cortex. Neurosci Lett 579: 110-113.
  16. Schneider JS, Giardiniere M, Morain P (2002) Effects of the prolyl endopeptidase inhibitor S 17092 on cognitive deficits in Chronic low dose MPTP-treated monkeys. Neuropsychopharmacology 26: 176-182.
  17. Toide K, Shinoda M, Miyazaki A (1998) A novel prolyl endopeptidase inhibitor, JTP-4819. Its behavioral and neurochemical properties for the treatment of Alzheimer's disease. Rev Neurosci 9: 17-29.
  18. Kato A, Fukunari A, Sakai Y, Nakajima T (1997) Prevention of amyloid-like deposition by a selective prolyl endopeptidase inhibitor, Y-29794, in senescence-accelerated mouse. J Pharmacol Exp Ther 283: 328-335.
  19. Umemura K, Kondo K, Ikeda Y, Kobayashi T, Urata Y, et al. (1997) Pharmacokinetics and safety of JTP-4819, a novel specific orally active prolyl endopeptidase inhibitor, in healthy male volunteers. Br J Clin Pharmacol 43: 613-618.
  20. Morain P, Robin JL, De Nanteuil G, Jochemsen R, Heidet V, et al. (2000) Pharmacodynamic and pharmacokinetic profile of S 17092, a new orally active prolyl endopeptidase inhibitor, in elderly healthy volunteers. A phase I study. Br J Clin Pharmacol 50: 350-359.
  21. Umemura K, Kondo K, Ikeda Y, Nishimoto M, Hiraga Y, et al. (1999) Pharmacokinetics and safety of Z-321, a novel specific orally active prolyl endopeptidase inhibitor, in healthy male volunteers. J Clin Pharmacol 39: 4462-4470.
  22. Morain P, Boeijinga PH, Demazières A, De Nanteuil G, Luthringer R (2007) Psychotropic profile of S 17092, a prolyl endopeptidase inhibitor, using quantitative EEG in young healthy volunteers. Neuropsychobiology 55: 176-183.
  23. Katsube N, Sunaga K, Aishita H, Chuang DM, Ishitani R (1999) ONO-1603, a potential antidementia drug, delays age-induced apoptosis and suppresses overexpression of glyceraldehyde-3-phosphate dehydrogenase in cultured central nervous system neurons. J Pharmacol Exp Ther 288: 6-13.
  24. Katsube N, Sunaga K, Chuang DM, Ishitani R (1996) ONO-1603, a potential antidementia drag, shows neuroprotective effects and increases m3-muscarinic receptor mRNA levels in differentiating rat cerebellar granule neurons. Neurosci Lett 214: 151-154.
  25. Schulz I, Zeitschel U, Rudolph T, Ruiz-Carrillo D, Rahfeld JU, et al. (2005) Subcellular localization suggests novel functions for prolyl endopeptidase in protein secretion. J Neurochem 94: 970-979.
  26. Spillantini MG, Crowther RA, Jakes R, Hasegawa M, Goedert M (1998) Alpha-Synuclein in filamentous inclusions of Lewy bodies from Parkinson's disease and dementia with lewy bodies. Proc Natl Acad Sci U S A 95: 6469-6473.
  27. Brandt I, Gérard M, Sergeant K, Devreese B, Baekelandt V, et al. (2008) Prolyl oligopeptidase stimulates the aggregation of α-synuclein. Peptides 29: 1472-1478.
  28. Hannula MJ, Myöhänen TT, Tenorio-Laranga J, Männistö PT, Garcia-Horsman JA (2013) Prolyl oligopeptidase colocalizes with α-synuclein, β-amyloid, tau protein and astroglia in the post-mortem brain samples with Parkinson's and Alzheimer's diseases. Neuroscience 242: 140-150.
  29. Savolainen MH, Yan X, Myöhänen TT, Huttunen HJ (2015) Prolyl oligopeptidase enhances α-Synuclein dimerization via direct protein-protein interaction. J Biol Chem 290: 5117-5126.
  30. Savolainen MH, Richie CT, Harvey BK, Männistö PT, Maguire-Zeiss KA, et al. (2014) The beneficial effect of a prolyl oligopeptidase inhibitor, KYP-2047, on alpha-synuclein clearance and autophagy in A30P transgenic mouse. Neurobiol Dis 68: 1-15.
  31. Saunders PA, Chalecka-Franaszek E, Chuang DM (1997) Subcellular distribution of glyceraldehyde-3-phosphate dehydrogenase in cerebellar granule cells undergoing cytosine arabinoside-induced apoptosis. J Neurochem 69: 1820-1828.
  32. Ishitani R, Tanaka M, Sunaga K, Katsube N, Chuang DM (1998) Nuclear localization of overexpressed glyceraldehyde-3-phosphate dehydrogenase in cultured cerebellar neurons undergoing apoptosis. Mol Pharmacol 53: 701-707.
  33. Matsuda T, Sakaguchi M, Tanaka S, Yoshimoto T, Takaoka M (2013) Prolyl oligopeptidase is a glyceraldehyde-3-phosphate dehydrogenase-binding protein that regulates genotoxic stress-induced cell death. Int J Biochem Cell Biol 45: 850-857.
  34. Puttonen KA, Lehtonen Š, Raasmaja A, Männistö PT (2006) A prolyl oligopeptidase inhibitor, Z-Pro-Prolinal, inhibits glyceraldehyde-3-phosphate dehydrogenase translocation and production of reactive oxygen species in CV1-P cells exposed to 6-hydroxydopamine. Toxicol Vitr 20: 1446-1454.
  35. Oestreicher AB, De Graan PNE, Gispen WH, Verhaagen J, Schrama LH (1997) B-50, the growth associated protein-43: Modulation of cell morphology and communication in the nervous system. Prog Neurobiol 53: 627-686.
  36. Aigner L, Caroni P (1993) Depletion of 43-kD growth-associated protein in primary sensory neurons leads to diminished formation and spreading of growth cones. J Cell Biol 123: 417-429.
  37. Harwood AJ, Mudge AW, Cheng L, Williams RSB (2002) A common mechanism of action for three mood-stabilizing drugs. Nature 417: 292-295.
  38. Di Daniel E, Glover CP, Grot E, Chan MK, Sanderson TH, et al. (2009) Prolyl oligopeptidase binds to GAP-43 and functions without its peptidase activity. Mol Cell Neurosci 41: 373-382.
  39. Szeltner Z, Morawski M, Juhász T, Szamosi I, Liliom K, et al. (2010) GAP43 shows partial co-localisation but no strong physical interaction with prolyl oligopeptidase. Biochim Biophys Acta - Proteins Proteomics 1804: 2162-2176.
  40. Matamoros AJ, Baas PW (2016) Microtubules in health and degenerative disease of the nervous system. Brain Res Bull 126: 217-225.
  41. Höfling C, Kulesskaya N, Jaako K, Peltonen I, Männistö PT, et al. (2016) Deficiency of prolyl oligopeptidase in mice disturbs synaptic plasticity and reduces anxiety-like behaviour, body weight, and brain volume. Eur Neuropsychopharmacol 26: 1048-1061.
  42. Uryu K, Butler AK, Chesselet MF (1999) Synaptogenesis and ultrastructural localization of the polysialylated neural cell adhesion molecule in the developing striatum. J Comp Neurol 405: 216-232.
  43. Diano S, Jeong JK, Kim JD, D'Agostino G, Calignano A, et al. (2012) Prolyl endopeptidase-deficient mice have reduced synaptic spine density in the CA1 region of the hippocampus, impaired LTP, and spatial learning and memory. Cereb Cortex 23: 2007-2014.
  44. Conde C, Cáceres A (2009) Microtubule assembly, organization and dynamics in axons and dendrites. Nat Rev Neurosci 10: 319-332.
  45. Pellegrini L, Wetzel A, Grannó S, Heaton G, Harvey K (2017) Back to the tubule: Microtubule dynamics in Parkinson's disease. Cell Mol Life Sci 74: 409-434.
  46. Camera P, Defilippi P, Jaworski J, Spangler SA, Gouveia SM, et al. (2009) Dynamic microtubules regulate dendritic spine morphology and synaptic plasticity. Neuron 61: 85-100.
  47. Brunden KR, Lee VMY, Smith AB, Trojanowski JQ, Ballatore C (2017) Altered microtubule dynamics in neurodegenerative disease: Therapeutic potential of microtubule-stabilizing drugs. Neurobiol Dis 105: 328-335.
  48. López A, Mendieta L, Prades R, Royo S, Tarragó T, et al. (2013) Peptide POP inhibitors for the treatment of the cognitive symptoms of schizophrenia. Future Med Chem 5: 1509-1523.
  49. Prades R, Munarriz-Cuezva E, Urigüen L, Gil-Pisa I, Gómez L, et al. (2017) The prolyl oligopeptidase inhibitor IPR19 ameliorates cognitive deficits in mouse models of schizophrenia. Eur Neuropsychopharmacol 27: 180-191.
  50. Glausier JR, Lewis DA (2013) Dendritic spine pathology in schizophrenia. Neuroscience 251: 90-107.
  51. MacDonald ML, Alhassan J, Newman JT, Richard M, Gu H, et al. (2017) Selective loss of smaller spines in Schizophrenia. Am J Psychiatry 174: 586-594.
  52. Zhou RM, Huang YX, Li XL, Chen C, Shi Q, et al. (2010) Molecular interaction of α-synuclein with tubulin influences on the polymerization of microtubule in vitro and structure of microtubule in cells. Mol Biol Rep 37: 3183-3192.
  53. Carnwath T, Mohammed R, Tsiang D (2018) The direct and indirect effects of α-synuclein on microtubule stability in the pathogenesis of parkinson's disease. Neuropsychiatr Dis Treat 14: 1685-1695.
  54. Cash AD, Aliev G, Siedlak SL, Nunomura A, Fujioka H, et al. (2003) Microtubule reduction in Alzheimer's disease and aging is independent of τ filament formation. Am J Pathol 162: 1623-1627.
  55. Pchitskaya EI, Zhemkov VA, Bezprozvanny IB (2018) Dynamic microtubules in Alzheimer's disease: Association with dendritic spine pathology. Biochem 83: 1068-1074.
  56. Ross CA, Aylward EH, Wild EJ, Langbehn DR, Long JD, et al. (2014) Huntington disease: Natural history, biomarkers and prospects for therapeutics. Nat Rev Neurol 10: 204-216.
  57. Gusella JF, Wexler NS, Conneally PM, Naylor SL, Anderson MA, et al. (1983) A polymorphic DNA marker genetically linked to Huntington's disease. Nature 306: 234-238.
  58. Arrasate M, Finkbeiner S (2012) Protein aggregates in Huntington's disease. Exp Neurol 238: 1-11.
  59. Mantle D, Falkous G, Ishiura S, Blanchard PJ, Perry EK (1996) Comparison of proline endopeptidase activity in brain tissue from normal cases and cases with Alzheimer's disease, Lewy body dementia, Parkinson's disease and Huntington's disease. Clin Chim Acta 249: 129-139.
  60. Barker RA, Cisbani G, Spillantini MG, Vuono R, de Silva R, et al. (2015) The role of tau in the pathological process and clinical expression of Huntington's disease. Brain 138: 1907-1918.
  61. Cabrera JR, Lucas JJ (2017) MAP2 splicing is altered in Huntington's disease. Brain Pathol 27: 181-189.

Citation

Trallero N, Anunciación-Llunell A, Prades R, Tarragó T (2019) The Role of Prolyl Oligopeptidase in Microtubule-Associated Processes and Cognitive Impairment. Int J Brain Disord Treat 5:029. doi.org/10.23937/2469-5866/1410029