Научная статья на тему 'ROLE OF THE MBP PROTEIN IN MYELIN FORMATION AND DEGRADATION IN THE BRAIN'

ROLE OF THE MBP PROTEIN IN MYELIN FORMATION AND DEGRADATION IN THE BRAIN Текст научной статьи по специальности «Биологические науки»

CC BY
162
15
i Надоели баннеры? Вы всегда можете отключить рекламу.
Журнал
Biological Communications
WOS
Scopus
ВАК
RSCI
Область наук
Ключевые слова
MYELIN / MBP / MULTIPLE SCLEROSIS / OLIGODENDROCYTE / AXON / MAMMALIAN BRAIN / AMYLOID

Аннотация научной статьи по биологическим наукам, автор научной работы — Shenfeld Aleksandr, Galkin Alexey

The compact myelin sheath functions as an insulator for efficient conduction of nerve impulses. The formation of myelin sheaths around the axons of the most actively functioning neurons continues not only at the stage of brain development, but also in the process of learning and acquiring certain skills. Pathological or age-related disruption in myelin results in nerve conduction failure and neurodegeneration. Myelin Basic Protein (MBP) is the main constituent of the myelin sheath, representing about 30 % of the total myelin proteins in the central nervous system. Deletion in the MBP coding gene in mutant mice causes a severe neurological phenotype associated with rapid death of newborns. In this review, we discuss the current understanding of the role of the MBP protein in the formation of compact myelin and in neurodegeneration associated with demyelination.

i Надоели баннеры? Вы всегда можете отключить рекламу.
iНе можете найти то, что вам нужно? Попробуйте сервис подбора литературы.
i Надоели баннеры? Вы всегда можете отключить рекламу.

Текст научной работы на тему «ROLE OF THE MBP PROTEIN IN MYELIN FORMATION AND DEGRADATION IN THE BRAIN»

REVIEW COMMUNICATIONS

CELL BIOLOGY

Role of the MBP protein in myelin formation and degradation in the brain

Aleksandr Shenfeld12 and Alexey Galkin13

1Vavilov Institute of General Genetics, Russian Academy of Sciences, Saint Petersburg Branch,

Universitetskaya nab., 7-9, Saint Petersburg, 199034, Russian Federation

laboratory of Amyloid Biology, Saint Petersburg State University,

Universitetskaya nab., 7-9, Saint Petersburg, 199034, Russian Federation

3Department of Genetics and Biotechnology, Faculty of Biology, Saint Petersburg State

University, Universitetskaya nab., 7-9, Saint Petersburg, 199034, Russian Federation

Address correspondence and requests for materials to Alexey Galkin, apgalkin@mail.ru

Abstract

The compact myelin sheath functions as an insulator for efficient conduction of nerve impulses. The formation of myelin sheaths around the axons of the most actively functioning neurons continues not only at the stage of brain development, but also in the process of learning and acquiring certain skills. Pathological or age-related disruption in myelin results in nerve conduction failure and neurodegeneration. Myelin Basic Protein (MBP) is the main constituent of the myelin sheath, representing about 30 % of the total myelin proteins in the central nervous system. Deletion in the MBP coding gene in mutant mice causes a severe neurological phenotype associated with rapid death of newborns. In this review, we discuss the current understanding of the role of the MBP protein in the formation of compact myelin and in neurodegeneration associated with demyelination.

Keywords: myelin, MBP, multiple sclerosis, oligodendrocyte, axon, mammalian brain, amyloid

Citation: Shenfeld, A. and Galkin, A. 2022. Role of the MBP protein in myelin formation and degradation in the brain. Bio. Comm. 67(2): 127-138. https://doi. org/10.21638/spbu03.2022.206

Authors' information: Aleksandr Shenfeld, PhD Student, Research Assistant, orcid. org/0000-0002-7286-2203; Alexey Galkin, Dr. of Sci. in Biology, Professor, orcid.org/0000-0002-7362-8857

Manuscript Editor: Michael Firsov, Sechenov Institute of Evolutionary Physiology and Biochemistry, Russian Academy of Sciences, Saint Petersburg, Russia

Received: November 16, 2021;

Revised: January 24, 2022;

Accepted: February 1, 2022.

Copyright: © 2022 Shenfeld and Galkin. This is an open-access article distributed under the terms of the License Agreement with Saint Petersburg State University, which permits to the authors unrestricted distribution, and self-archiving free of charge.

Funding: No funding information provided.

Ethics statement: This paper does not contain any studies involving human participants or animals performed by any of the authors.

Competing interests: The authors have declared that no competing interests exist.

Introduction

Myelin is a specialized sheath that forms around the axons of many actively working neurons and functions as an insulator in conducting nerve impulses between the body of the neuron and its target (Hartline and Colman, 2007). Compact myelin sheaths are formed by specialized glial cells — oligodendrocytes in the central nervous system (CNS) and Schwann cells in the peripheral nervous system (PNS). They are wrapped around the axon in a spiral fashion (Monje, 2018). Myelin of the central and peripheral nervous systems are functionally and structurally similar. The main difference between them is that in the PNS, one myelinating Schwann cell envelops one axon by itself; while in the CNS a single oligodendrocyte forms numerous flattened processes that can simultaneously myelinate different axons, while the bodies of the glial cells do not participate in myelination. Myelin of nerve fiber is not presented as a continuous sheath along the axon; in contrast, it forms separate segments or "internodes". Each myelinated segment of axon is flanked by narrow portions uncovered by myelin, called nodes of Ranvier, which are critical to the nerve functioning (Nave and Werner, 2014) (Fig. 1A). Myelinated nerve fibers, along with other glial cells and blood vessels, form the so-called white matter (WM) of the brain.

Myelin facilitates the conduction of nerve impulses by serving as an electrical insulator. Unmyelinated fibers have voltage-gated sodium channels along the entire length of the membrane, so the action potential travels as local circuits of ion current continuously (de Col, Messlinger and Carr, 2008). In contrast, in myelinated axons, the axolemma of nodes of Rannier contains a high density of Na+-channels against a relatively small number of channels on the membrane

Fig. 1. (A) The diagram showing compact myelin structure formed by the oligodendrocyte processes (blue) that are wrapped around the axon (purple). (B) Compact myelin is represented by a series of alternating major dense (MDL) and intraperiod (IPL) lines. (C) Proposed model of MBP binding to lipid bilayers within myelin composition. A cross section of a flattened process of an oligodendrocyte in the area of compact myelin formation is shown. The N and C-terminal sequences of MBP (red) bind to lipids (blue) of the opposite process membranes of the oligodendrocyte. (D) Amyloid model of MBP protein folding within myelin sheath. According to this model MBP molecules not only bind opposite membranes, but also form amyloid fibrils inside the process of the oligodendrocyte.

of internodes. When the membrane at the node is excited, the local currents flow further to the axolemma in the internodes; here, due to the insulating properties of the myelin sheath, the impulse does not attenuate, but "jumps" to the membrane at the next node. Thus, highresistance sheath contributes to the preservation of the action potential in the nerve fiber, and the excess of Na+-channels in the membrane of Ranvier's nodes amplifies the transmitted signal each time, which leads to an increase in the speed of local circuit spreading (Huxle and Stampfli, 1949). Active excitation of the axonal membrane jumps from one node to the next; this causes a fast saltatory propagation of the impulse (from the Latin saltare — "to jump"), in contrast to unmyelinated nerve fibers, where action potentials are transferred in orders of magnitude slower (Ritchie, 1982).

According to transmission electron microscopy (TEM) data, mature myelin contains "compact" and "non-compact" areas (Stadelmann, Timmler, Barrantes-

Freer and Simons, 2019). Non-compact myelin is present in the form of cytoplasm-filled membrane cavities in the outer and inner layers of the myelin sheaths as well as in the paranodal loops — specialized structures located on lateral margins of the myelin segments (Peters, 1960). These compartments perform a signaling and trophic function, supplying energy substrates from the myelin sheath to the axon, including lactate and py-ruvate (Funfschilling et al., 2012). Compact myelin is characterized by a unique molecular composition providing its insulating properties. The dry mass of myelin contains a high proportion of lipid (70-85 %), whereas the protein content is very low — 15-30 %, of which myelin basic protein (MBP) is one of the two most abundant proteins (Norton and Cammer, 1984). In the CNS, MBP constitute about 30 % of total myelin protein by weight, while in the PNS its quantity is minimal, and its deficiency is compensated for by adhesive PNS-speci-fic P0 protein (Martini et al., 1995). MBP is an essential

structural component of mature compact myelin, since disruptions in its production cause severe demyelination of the CNS. For instance, a mouse shiverer mutation, which results from a large deletion in the MBP gene, impairs the structural integrity of the myelin sheath and causes the classic symptoms of demyelination, accompanied by whole body tremors, seizures, progressive ataxia, hindlimb paresis and early death (Bourre et al., 1980; Jacque, Delassalle, Raoul and Baumann, 1983; Fitzner et al., 2006). Similar myelin deficits also result from a rat Long Evans shaker (les) mutation upon the insertion of a 5.7 kb endogenous retrotransposon in the non-coding region (intron 3) of the MBP gene, which alters the normal splicing dynamics of MBP mRNA and decreases its expression level (O'Connor et al., 1999).

In TEM, a compact myelin sheath is visualized as a series of alternating dark and pale lines, separated by unstained lipid-rich layers (Fig. 1B) (Stoeckenius, 1959; Peters, 1960). A pale line, or intraperiod, corresponds to closely apposing extracellular surfaces of the oligoden-drocyte plasma membrane. The dark, or major dense line (MDL) represents the inner protein-rich layer between the cytoplasmic sides of oligodendroglia membranes. MBP is located between cytoplasmic membrane faces in MDL, stabilizing the multilayer myelin structure through its anchoring to opposite membrane sides with N- and C- termini (Fig. 1C) (Harauz and Libich, 2009). MBP is assumed to be retained within this layer by electrostatic forces with membrane lipids. This molecule is characterized by a high positive net charge caused by its large number of charged residues throughout the protein's sequence, mostly arginine and lysine (Boggs et al., 2004). Its high overall positive charge and low hydro-phobicity, which maximize intramolecular electrostatic repulsion, causes an internal conformational disorder of protein in solution, which allows it to interact with the negatively charged phospholipid-rich cytoplasmic surface of myelin membranes (Raasakka et al., 2017). Thus, the adhesive role of MBP is supposed to be fundamental in myelin assembly.

Heterogeneity of MBP isoforms

Human MBP protein is produced by the Golli-MBP gene complex, which localizes on Chromosome 18q (de Ferra et al., 1985), consists of 11 exons and has two transcription initiation sites (Campagnoni et al., 1993). This gene produces two subfamilies of proteins — the "classical" myelin-specific MBP isoforms which are present in myelinating oligodendrocytes in the CNS and Schwann cells in PNS; and "Golli" MBP variants (the "Golli" prefix is an abbreviation for gene expressed in the oligodendrocyte lineage) that are found, besides the neuroglia, in the immune tissues (Campagnoni et al., 1993; Fritz and Kalvakolanu, 1995). Golli-MBP protein

molecules (33-35 kDa), unlike classical MBP isoforms, include a 130-amino acid golli domain that is encoded by I-III exons, whose mRNAs are transcribed from upstream promoter (Feng et al., 2000). These isoforms are not normal components of the myelin sheath since they are mainly localized in the nuclei and cell bodies of oligodendrocyte precursors cells (OPC). Mainly the expression of Golli-MBP isoforms is detected during embryogenesis and followed by the expression of the classic MBP. The exact function of the Golli isoforms is not fully understood. However, several studies have shown that Golli-MBP mediates the entry of Ca2+ ions into oligo-dendrocytes through voltage-dependent ion channels in a depolarized state, and it is also involved in the differentiation and migration of OPC during myelination (Jacobs et al., 2005). The nuclear localization of Golli-MBP is explained by its binding to transcription factors (Fer-nandes et al., 2004); however, its role in the regulation of gene expression is still controversial.

Classical MBP protein is represented by six major isoforms in mice and four in humans (Harauz and Boggs, 2013). These proteins are derived from alternative splicing of a single MBP mRNA, consisting of seven downstream exons of the Golli gene complex. The production level of various MBP isoforms is determined by the developmental stage of oligodendrocytes, for instance, isoforms encoded by mRNA with exon-II, which correspond to 17.22, 20.2, and 21.5 kDa in mice, are produced in the onset of myelin formation (Smith et al., 2013). The exon-II negative MBP variants (14 and 18.5 kDa in mice), in contrast, present later in development (Pedraza, Fidler, Staugaitis, and Colman, 1997). Also, exon-II containing MBP isoforms are not involved in the compact myelin formation. At the same time, other exon-II negative variants of MBP (14 and 18.5 kDa in mice) are present in myelin sheath composition (Al-linquant, Staugaitis, D'Urso and Colman, 1991). Probably this heterogeneity of MBP isoforms is necessary for developmental regulation of myelinogenesis.

Among other things, the diversity of MBP post-translational modifications is another determining factor of its heterogeneity, the role of which is unknown or poorly understood (Kim et al., 2003). Some protein post-translational modifications are reversible while others are not. To date, it is well known that MBP in vivo undergoes five forms of post-translational modification: acetylation, methylation, phosphorylation, citrullina-tion and deamidation (Boggs, 2006). Modification of the MBP protein contributes to the formation of "charge isomers", which always give many bands on electropho-resis. Normally, human MBP 18.5-kDa occurs as a series of C1-C8 isomers in order of decreasing total positive charge (Zand et al., 1998). The C2 isomer corresponds to deamidated MBP (Kim et al., 2003), while C3-C6 can be modified by combinations of deamidation, phosphory-

lation, and citrullination (deimination); the last two types of modification are expected to have particular functional significance (Boggs, 2006). Covalent attachment of phosphate groups to serine and threonine residues of MBP occurs in response to various extracellular signals, with the greatest changes observed in mature myelin (Zand et al., 1998). Phosphorylation of MBP in the myelin sheath changes in response to neural action potential, and in oligodendrocyte culture in response to extracellular ligands and depolarization (Murray and Steck, 1984; Boggs, 2006). It is assumed that phosphorylation of MBP isoforms containing the sequence encoded by exon-II can inhibit their transport into the nucleus (Pedraza et al., 1997). Reducing the cationic charge of MBP by phosphorylation influences protein conformations as well as protein — lipid and protein — protein interactions. Phosphorylated MBP has a decreased ability to assemble actin in vitro (Boggs et al., 2006), but an enhanced ability to polymerize and bundle tubulin (Harauz and Libich, 2009). It is believed that MBP phosphorylation is not a spontaneous process, since the degree of protein modification changes during the development of the organism, ageing and pathological processes. For instance, the level of MBP phosphoryla-tion in the brain of patients with multiple sclerosis (MS) is significantly reduced (Kim et al., 2003).

Removal of arginyl residues by peptidylarginine deiminase (PAD) leads to the formation of citrulline — an important irreversible modification of MBP (Kim et al., 2003). In the C8 isoform, at least 6 of the 19 Arg residues are citrullinated, which reduces the positive charge of the protein from +19 to +13 — this, in turn, leads to disruption of protein folding, weakening of the contact of MBP molecules with phospholipids, and degradation by proteases such as cathepsin-D (Vassall et al., 2016). The proportion of MBP protein presented in a citrulli-nated isoform is significantly higher in children than in adults. These data give reasons to believe that they play a regulatory role in the formation of the myelin sheath rather than in its functioning. For example, a healthy brain normally contains 20 % of citrullinated forms of MBP, while an increase of their quantity can serve as a marker of the development of multiple sclerosis (Boggs, 2006). In addition, MBP has recently been identified as a target for citrullination in the brains of prion-infected patients with Creutzfeldt-Jakob disease and patients with Alzheimer's disease (Ishigami et al., 2005; Jang et al., 2010).

Although post-translational modifications of MBP are distributed along the entire length of the protein sequence, their highest frequency is limited to regions that are presumably disordered. Post-translational modification hotspots are mostly represented in the N- and C-terminal regions of MBP, which correspond to exons I and V-VII, respectively (Harauz and Libich, 2009).

These modifications, especially phosphorylation and citrullination, contribute to a decrease in the total protein charge. The stability of the multilayer proteolipid myelin sheath is determined by the initial electrostatic interactions of MBP with membrane lipids (Raasakka et al., 2017; Vassall, Bamm and Harauz, 2015). In this regard, it can be assumed that due to a decrease in the forces of intramolecular electrostatic repulsion mediated by modification, the protein undergoes conforma-tional changes, which can affect its ability to bind to the membrane, and these changes, in turn, affect both the formation of myelin and its stabilization. For instance, in the MS brain, deiminated as well as methylated variants of MBP are often found (Kim et al., 2003). For this reason, some researchers classify demyelination as a "post-translational disease".

Role of MBP in myelination dynamics throughout life

White matter generation begins in late embryonic and the first six postnatal weeks in rodents. Myelin-forming oligodendrocytes arise from multipotent NG2-progeni-tor cells, also called oligodendrocyte precursor cells (OPCs), which persist throughout life within the adult brain parenchyma and make up around 5 % of the glial cell population (Dawson, Levine, and Reynolds, 2000; Levine, Reynolds, and Fawcett, 2001). Since myelin acts as an insulator for the transmission of action potential along the axon, an active process of myelination occurs in the early postembryonic stages of development during active cognitive processes; this is necessary for accelerating information transfer in neural circuits and sustaining axons' activity (Hasegawa et al., 1992). Myelin sheaths are formed only around the axons of actively functioning neurons, thus in adult cortex, areas remain with unmyelinated or partially myelinated axons. The learning process of a child and an adult, associated with the activation of certain neurons, is accompanied by de novo myelination. (Monje, 2018; Williamson and Lyons, 2018). For instance, human studies provide compelling evidence of a link between adaptive myelination and motor learning, like juggling (Scholz, Klein, Behrens and Johansen-Berg, 2009) or piano practicing (Steele et al., 2013). Elevated motor circuit activity entails not only generation of new oligodendrocytes (McKenzie et al., 2014) but also increased expression of MBP in white matter areas underlying the motor cortex (Sampaio-Baptista et al., 2013). MBP production is proposed to modulate WM plasticity in response to learning-induced neuronal activity by regulating the myelin sheath thickness (Martini and Schachner, 1997). It is worth noting that activity-regulated myelination is not restricted to the motor system, and also includes non-motor learning, as it was revealed in the studies of effects of adult

secondary language learning (Schlegel, Rudelson and Tse, 2012). Thus, de novo myelination is supposed to be an important form of the neuron circuits plasticity.

Myelination, as a developmentally regulated process, in general requires coordinated expression of many genes, including MBP. As mentioned above, MBP is an upstream product of the Golli-MBP gene complex, expression of which is differently regulated both during development and within specific tissues (Campagnoni et al., 1993). Golli-MBP mRNAs are prenatally expressed in oligodendrocytes (Campagnoni et al., 1993) as well as some neuronal populations (Landry et al., 1996, Pribyl et al., 1996). Also, its messenger transcript variants were found in the human fetal spinal cord, thymus, and spleen (Pribyl et al., 1993). For this reason, the function of Golli-MBP isoforms extends beyond the regulation of myelin formation and includes oligodendrocyte proliferation and migration (Jacobs et al., 2005; Paez et al., 2011). Due to this functional ambiguity and spatial distribution in early-developmental stages, Golli-MBP has been called a "molecular link" between the nervous and immune system (Pribyl et al.,1993). Expression of the classic MBP isoforms, in contrast, dominates in later stages of CNS development. In the human prenatal brain development, there were two distinguished periods of MBP expression: one in onset and the other during myelination (Zecevic, Andjelkovic, Matthieu and Tosic, 1998). In the mouse brain the peak accumulation of mRNAs encoding MBP occurred in the third postnatal week, thereafter its level declined to a much lower steady state (Zeller et al., 1984). Notably, in the mouse CNS a positive correlation was observed between the MBP production level and CNS myelin sheath thickness, which emphasizes MBP as an essential limiting factor in myelination (Shine et al., 1992). However, the ratios of MBP isoforms vary during development too. The level of expression of exon-II positive MBP messenger RNAs, coding 21.5 and 17 kDa polypeptides, dominates at early postnatal ages, while exon-II negative isoforms, i.e., 18.5 and 14 kDa, are present in the more mature brain (Carson, Nielson and Barbarese et al., 1983). Since exon-II containing MBP molecules are not an integral part of compact myelin, it is believed that these isoforms may play a regulatory role in the process of oligodendro-cyte proliferation via a mechanism that relies on their dynamic nuclear import and export (Pedraza, Fidler, Staugaitis and Colman, 1997). Moreover, these mRNAs serve as precursors in biogenesis of myelin-forming exon-II negative MBP isoforms, which are obtained via splicing out exon-II of the primary transcript (Campagnoni et al., 1993). Among other things, an increase of exon-II containing MBP transcripts coincides with the earliest histological signs of remyelination — the process of myelin sheath recovery of previously demyelin-ated nerve fibers. A 13-fold higher level of expression

of exon-II containing MBP transcripts was noted at the beginning of remyelination relative to control white matter mRNAs level. In the later stages of myelin recovery these molecules were replaced by exon-II negative spliced variants (Campagnoni et al., 1993). Thus, the dynamics of the expression of different forms of MBP during remyelination coincide with myelination at the early stages of CNS development. Nevertheless, despite an abundance of MBP isoforms produced through development, it is supposed that the shortest 14 kDa MBP variant might be sufficient for normal development and functioning of myelin. Studies on homozygous shiverer (shi/shi) mice have shown that transgenic production of the small 14 kDa MBP isoform partly recovered individuals from the shiverer phenotype with elimination of tremor and increased lifespan (Kimura et al., 1989).

The content of myelin across the lifespan is not constant. In the human brain, total white matter volume reaches its peak by the age of 30 and then declines with ageing (Miller et al., 2012), correlating with cognitive impairment that mostly affects working executive memory and the slowing of processing speed (Peters, 2002). Under the microscope, age-related degenerative abnormalities in myelin sheaths are revealed as myelin balloons (Feldman and Peters, 1998) and splits of the lamellae at the major dense line (Peters, 2002). Reasons for this are not clear, however, this could be a consequence of a complex of factors, such as reduced OPC differentiation (Sim, Zhao, Penderis and Franklin, 2002), differential vulnerability of oligodendrocytes to accumulation of DNA damage and oxidative stress (Tse and Herrup, 2017), altered lipid metabolism (Lefevre-Arbogast et al., 2021), and decreased nutrient and energy availability for myelinating cells (Sams, 2021). In addition, senescence-dependent abnormalities in myelin are often caused by a decrease in the expression level of various genes, including MBP, which serves as a robust indicator of brain age. Studies on the human visual cortex of healthy people have shown that classic MBP isoforms gradually increase up to 42 years and then decline during ageing, while Golli-MBP protein production, in contrast, increases (Siu, Balsor, Jones and Murphy, 2015). Significant age-related reduction in MBP levels was revealed in specific layers of the hippocampus (Ahn et al., 2017) — the part of the limbic system critical for learning and memory. In addition, overall age-related decline of MBP was identified in the corpus callosum and in the dorsal column of the spinal cord in aged rats (Xie, Zhang, Fu and Chen, 2013). Taking into consideration the importance of MBP in sustaining myelin, this age-related protein downregulation obstructs further myelin formation de novo and myelin remodeling, which leads to reduction in the conduction velocity of nerve fibers and diminution in neural connectivity.

It is important to note that myelin breakdown through ageing is associated with astrocytes and mi-

croglia activation, which are involved, among other things, in the degradation of myelin components (Xie, Zhang, Fu and Chen, 2013). The inclusions of MBP and proteolipid protein (PLP) were identified within microglial cells of ageing wild-type mice (i. e., up to 24 months old). Also, western blot analysis of purified microglia from 12-month-old mice revealed sarcosyl-insoluble high molecular weight aggregates of MBP (Safaiyan et al., 2016; Theriault and Rivest, 2016). Based on these results, it was assumed that microglia participate in myelin debris uptake that accumulates in the ageing brain. Lysosomal degradation plays a special role within this, because in studies on RAB7 knockout mice with impaired lysosomal activity, MBP-rich puncta in microglia appeared earlier than in wild-type ageing mice (i. e., at 9 months old rather than 18 months old). A significant accumulation of MBP aggregates in ageing microglia indicates impaired clearance capacity. It is still not completely known whether this is a consequence of the senescence of microglia itself or if the accumulation of MBP aggregates is caused by the overloading of myelin debris in the cellular environment, which normal cells cannot cope with utilizing. In some ways this impaired microglial degradative capacity is reminiscent of aberrant accumulation of amyloid proteins in microglial cells during amyloid-induced neurodegenerative disorders, such as Alzheimer's (Paresce, Chung and Maxfield, 1997) or Huntington diseases (Franklin, Clarke and Patani, 2021), which can be explained by the high stability and resistance of amyloid fibrils to enzymatic pro-teolysis (Kushnirov, Dergalev and Alexandrov, 2020; Schonfelder et al., 2021).

Models of MBP folding within compact myelin

There are various data and ideas about how exactly MBP connects the opposite membranes of myelin and contributes to its compaction. Difficulties in the study of these issues are primarily related to the fact that the formation of compact myelin occurs only in vivo, when the processes of oligodendrocytes are wrapped around axons. All attempts to obtain three-dimensional crystals of 18.5 kDa MBP suitable for X-ray diffractometry have failed, because after removal from tissues, the native form of the protein adopts a random coil conformation and persists as a population of structurally non-identical molecules (Sedzik and Kirschner, 1992). Spectroscopic studies showed that MBP in aqueous solution has a disordered conformation while its secondary structure changes in the presence of lipids or detergents: SDS (sodium dodecyl sulfate) or DPC (dodecylphosphocho-line) (Polverini et al., 1999). In vitro studies show that different MBP isoforms can interact with a variety of proteins, including cytoskeletal proteins (Bamm et al.,

2011; Boggs et al., 2014). At the same time, analysis of the structure of myelin membranes shows that MBP displaces other proteins in the major dense line (Zuchero et al., 2015). The mRNA of membrane-associated MBP isoforms is trafficked to the oligodendrocyte processes plasmalemma, where it is translated and quickly connects to the inner membrane sheets (Bakhti, Aggarwal and Simons, 2014; Seiberlich et al., 2015).

The interaction of MBP with the membrane is mainly based on electrostatic forces between the positively charged amino acid residues of MBP and the negatively charged head groups of the inner leaflet lipids, phospha-tidylserine and phosphatidylinositol 4,5-bisphosphate (Riccio et al., 2000; Nawaz et al., 2009). By binding to the cytosolic membrane surfaces, opposite charges are neutralized, allowing other forces such as hydrogen bonding and hydrophobic factors to be unmasked. Membrane binding switches the properties of MBP, thereby promoting self-interaction into a tightly packed protein phase that forms the major dense line and binds the cytoplasmic surfaces of the bilayers tightly together (Kattnig et al, 2012; Muruganandam et al., 2013). Such a phase transition from a soluble to a polymerized pool of molecules is frequently observed for many structurally disordered proteins, in particular those engaged in RNA binding (Calabretta and Richard, 2015).

An important biochemical feature of the compact myelin composition is a low ratio (0.25) of proteins to lipids, in comparison with the plasma membranes of other cells (ratios ranging from 1.0 to 4.0) (Agarwal et al., 2011). Moreover, studies on the oligodendrocyte cell culture of mutant shiverer mice have shown that the amount of protein in the underdeveloped myelin sheets is significantly higher than in the mature myelin of wild-type individuals (Aggarwal et al., 2011). It is believed that an MBP-formed molecular sieve serves as a diffusion barrier for most cytosolic proteins, including 2'3'-cyclic nucleotide 3'-phosphodiesterase (CNPase) and myelin-associated glycoprotein (MAG) (Aggarwal et al., 2011), and other cellular compartments, to generate oligodendrocyte flat processes. For efficient myelin compaction, it is reasonable to coat the entire cytosolic surface of the oligodendrocyte membrane with MBP molecules, which requires accurate regulation of targeted delivery of large amounts of protein. The membrane-associated translation of the MBP protein in the processes of oligodendrocytes promotes its rapid adhesion to the membrane and conformational stabilization, which is accompanied by flattening of the processes (Simons and Nave, 2016). The MBP assembling within sheaths during myelination, in turn, leads to the disassembly of the actin cytoskeleton (Zuchero et al., 2015). The binding of MBP molecules to membrane phospho-lipids causes the competitive displacement of gelzolin and cofilin factors, which, being in the cytoplasm after

dissociation with the membrane, triggers depolymer-ization of the actin cytoskeleton (Zuchero et al., 2015). These intracellular molecular rearrangements ultimately lead to the formation of lipid-saturated isolating myelin sheets containing mainly extracellular PLP and intracel-lular MBP proteins (Aggarwal et al., 2013).

Transmission electron microscopy studies of MBP/ C1 particles adhering to the lipid monolayer combined with three-dimensional reconstruction showed that MBP particles form subtle variations of the basic "C" shape with a diameter of approximately 11 nm in a low salt buffer which mimics the physiological conditions (Beniac et al., 1997). The authors also proposed a model of the molecule topology, which is consistent with earlier three-dimensional models, according to which the MBP comprised five ^-sheets in antiparallel configuration and a large proportion of irregular coil (Ridsdale et al., 1997). Later, spectroscopic studies, including multidimensional NMR spectroscopy, of both the full-length 18.5 kDa MBP and their peptides revealed three segments with a strong propensity to form a-helices within a phospholipid environment. Note that all of these models describing the putative structure of MBP are based on the study of the structure of a protein not bound to the membrane. According to in vitro studies based on membrane mimetic model systems, including MD simulations, it has been suggested that the MBP is packed into the major dense line of myelin in a hairpin conformation where different a-helical segments interact simultaneously with apposing membrane leaflets (Harauz and Libich, 2009). However, this model does not provide a clear mechanism or dynamics of the global protein folding in myelin itself.

According to another model, the binding of MBP to opposite membrane surfaces induces its phase transition into amyloid conformation (Fig. 1D) (Aggarwal et al., 2013). The authors demonstrated that myelin bundles colocalize with amyloid-specific dye Thioflavin S in the brain of wild-type mice, whereas MBP-deficient shiverer mice with underdeveloped myelin bundles were not stained with Thioflavin S. Note that Thioflavin staining is not rigorous and sufficient evidence of the amyloid structure of the protein. Also, the authors of this work demonstrated that 14 kDa MBP variant and its two short fragments form fibrils in vitro. Despite TEM-visible fibrillar morphology, the presence of a cross-beta structure, which is a distinctive fundamental feature of amyloid folding, has not been studied. The authors suggested that the attachment of MBP to the membrane promotes it to disorder-to-order transition due to the neutralization of the positive charge in MBP, which induces its self-assembly by loss of electrostatic repulsion. This may contribute to the formation of amyloid mesh-work, in other words a "molecular sieve", the generation of which causes extrusion of cytoplasm excess, the con-

vergence of leaflets together, and formation of the major dense line as a result. Earlier it was shown that hexahis-tidine tagged 18.5 kDa MBP forms "tubular" structures on binary monolayers containing nickel-chelating lipids (Ishiyama et al., 2001). However, the amyloid nature of these structures has not been tested either. The hypothesis postulating the formation of MBP amyloid fibrils looks attractive. The presence of an amyloid network within the flattened processes of oligodendrocytes can promote reliable isolation of axons and rapid transmission of nerve impulses. A number of studies convincingly show that proteins can function in the brain in an amyloid form and perform vital functions (Maji et al., 2009; Sopova et al., 2019; Galkin and Sysoev, 2021).

Role of MBP in pathological demyelination

The diseases characterized by damage of myelinated covers of nerve fibers in the human CNS are usually grouped into a family of so-called demyelinating diseases. These pathologies can be caused by different factors including inflammatory processes, viral infection, acquired metabolic derangements, ischaemic damage, etc. (Love, 2006). Among demyelinating diseases, the most common is multiple sclerosis (MS) — the autoimmune pathology that commonly affects young and middle-aged adults. MS-mediated myelin degradation leads to disruption of neuron circuits, resulting in a range of neurological and psychiatric symptoms, including cognitive and motor impairment, all of which are determined by the locations of the lesions within the nervous system. The exact etiology of MS is currently unclear, but risk factors include bacterial and viral infections (Marrodan, Alessandro, Farez and Correale, 2019). Apparently, abnormal immune response to an infectious agent causes activation of microglial cells that destroy the blood-brain barrier and myelin sheaths. The destruction of myelin leads to the induction of an autoimmune reaction to a number of proteins, in particular MBP (Lehmann, Rottlaender and Kuerten, 2015). MBP-specific antibodies in the blood of patients are one of the main indicators of active demyelination in MS (Cohen, Herndon and McKhann, 1976; Whitaker, 1977). Traumatic brain injuries also provoke myelin destruction, causing the accumulation of MBP in the cerebrospi-nal fluid and blood, which leads to the development of MS. Myelin damage in MS promotes proteolysis of MBP, resulting in its cleavage into separate fragments, some of which are resistant to further cleavage. Interestingly, the most immunogenic epitope of this protein in MS (Whitaker, 1997) can form amyloid fibrils according to bioinformatic predictions (Aggarwal et al., 2013). It can be assumed that the MBP fragments forming amyloid aggregates are resistant to proteolysis and provoke an autoimmune response during myelin degradation.

It should be noted that MS is accompanied by modifications of the MBP. C1, the least modified and most cationic isoform, is the most abundant form of MBP in healthy adult humans; C8, a less cationic isoform with extensive citrullination of arginyl residues, has an increased level in the brain in individuals with MS (Mos-carello, Mastronardi and Wood, 2006). Moreover, the severity of MS strongly correlates with the deimination level — 18 % against 45 % in a normal and chronic MS brain, respectively. Furthermore, in the rapidly progressive and aggressive Marburg variant of MS, the MBP citrullination level reaches 90 % (Musse, Boggs and Harauz, 2006). As mentioned earlier, the high positive net charge provides adhesion and stabilization of molecules within phospholipid composition of compact multilayers, which is necessary for myelin retention. When MBP become less cationic it is supposed to be unable to carry out this function. In all likelihood, modifications of MBP in MS are not a cause but a consequence of the disease.

Conclusion

The myelin sheaths are unique structures that ensure the functioning of the most actively working neurons. Modern knowledge about the structure of myelin is far from complete, since this structure is formed only in vivo as a result of the interaction of processes of oligodendrocytes with axons. MBP is a key protein for myelin compaction and function. In this review, we discussed current data and hypotheses about the functional role of MBP, its modifications and conformational changes in the human brain in health and disease. Based on the data obtained in vivo, it can be concluded that this protein binds the opposite membranes of the flattened processes of oligodendrocytes and displaces other proteins from the region of the major dense line. Moreover, binding to membrane lipids provokes aggregation of MBP, which contributes to the compaction of myelin. So far, we can only assume which conformational changes in MBP occur during the formation of compact myelin. In our opinion, the most reasonable and attractive hypothesis is that MBP molecules inside the processes of oligoden-drocytes form an amyloid network, which contributes to the reliable isolation of axons associated with myelin. MBP also plays an important role in the development of multiple sclerosis, which is characterized by myelin degradation and induction of an autoimmune response. It is interesting to note that in multiple sclerosis, the most immunogenic epitope of MBP has potentially am-yloidogenic properties. Further studies of the structural properties of MBP in vivo will provide progress in understanding the organization and functioning of myelin sheaths in health and disease.

References

Aggarwal, S., Yurlova, L., Snaidero, N., Reetz, C., Frey, S., Zimmermann, J., Pähler, G., Janshoff, A., Friedrichs, N., Müller, D.J., Goebel, C., and Simons, M. 2011. A size barrier limits protein diffusion at the cell surface to generate lipid-rich myelin-membrane sheets. Developmental Cell 21(3):445-56. https://doi.org/10.1016/j.devcel.2011.08.001 Aggarwal, S., Snaidero, N., Pähler, G., Frey, S., Sánchez, P., Zweckstetter, M., Janshoff, A., Schneider, A., Weil, M.T., Schaap, I.A.T., Görlich, D., and Simons, M. 2013. Myelin membrane assembly is driven by a phase transition of myelin basic proteins into a cohesive protein meshwork. PLOS Biology 11(6):e1001577. https://doi.org/10.1371/ journal.pbio.1001577 Ahn, J. H., Lee, T. K., Park, J. K., Cho, J. H., Kim, I. H., Lee, J. C., Hong, S., Jeon, Y.H., Kang, I.J., Lee, Y.J., Won, M.H., and Lee, C.H.2017. Age-dependent differences in myelin basic protein expression in the hippocampus of young, adult and aged gerbils. Laboratory Animal Research 33(3):237-243. https://doi.org/10.5625/ lar.2017.33.3.237 Allinquant, B., Staugaitis, S. M., D'Urso, D., and Colman, D. R. 1991. The ectopic expression of myelin basic protein iso-forms in Shiverer oligodendrocytes: implications for myelinogenesis. Journal of Cell Biology (2):393-403. https:// doi.org/10.1083/jcb.113.2.393 Bakhti, M., Aggarwal, S., and Simons, M.2014. Myelin architecture: zippering membranes tightly together. Cellular and Molecular Life Sciences 71(7):1265-1277. https://doi. org/10.1007/s00018-013-1492-0 Bamm, V.V., De Avila, M., Smith, G.S.T., Ahmed, M.A.M., and Harauz, G.2011. Structured functional domains of myelin basic protein: Cross talk between actin polymerization and Ca2+-dependent calmodulin interaction. Biophysical Journal 101(5):1248-1256. https://doi. org/10.1016/j.bpj.201 1.07.035 Beniac, D. R., Luckevich, M. D., Czarnota, G.J, Tompkins, T. A., Ridsdale, R.A., Ottensmeyer, F. P., Moscarello, M.A., and Harauz, G.1997. Three-dimensional structure of myelin basic protein. I.Reconstruction via angular reconstitution of randomly oriented single particles. Journal of Biological Chemistry 272(7):4261-4268. https://doi. org/10.1074/jbc.272.7.4261 Boggs, J. M., Rangaraj, G., Hill, C. M. D., Bates, I. R., Heng, Y.-M., and Harauz, G. 2005. Effect of arginine loss in myelin basic protein, as occurs in its deiminated charge isoform, on mediation of actin polymerization and actin binding to a lipid membrane in vitro. Biochemistry 44(9):3524-3534. https://doi.org/10.1021/bi0473760 Boggs, J.M.2006. Myelin basic protein: A multifunctional protein. Cellular and Molecular Life Sciences 63(17):1945-1961. https://doi.org/10.1007/s00018-006-6094-7 Boggs, J.M., Homchaudhuri, L., Ranagaraj, G., Liu, Y., Smith, G. S., and Harauz, G. 2014. Interaction of myelin basic protein with cytoskeletal and signaling proteins in cultured primary oligodendrocytes and N19 oligoden-droglial cells. BMC Research Notes 7(1):1-14. https://doi. org/10.1186/1756-0500-7-387 Bourre, L.M., Jacque, C., Delassalle, A., Nguyen-Legros, J., Dumont, O., Lachapelle, F., Raoul, M., Alvarez, C., and Baumann, N.1980. Density profile and basic protein measurements in the myelin range of particulate material from normal developing mouse brain and from neurological mutants (Jimpy; quaking; Trembler; shiverer and its mld allele) obtained by zonal centrifugation. Journal of Neurochemistry 35(2):458-464. https://doi. org/10.1111/j.1471-4159.1980.tb06287.x

Calabretta, S. and Richard, S.2015. Emerging roles of disordered sequences in RNA-binding proteins. Trends in Biochemical Sciences: Cell Press 40(11):662-672. https://doi. org/10.1016/j.tibs.2015.08.012 Campagnoni, A. T., Pribyl, T. M., Campagnoni, C. W., Kampf, K., Amur-Umarjee, C. F., Handley, V. W., Newman, S. L., Gar-bay, B., and Kitamura, K.1993. Structure and developmental regulation of Golli-mbp, a 105 kilobase gene that encompasses the myelin basic protein gene and is expressed in cells in the oligodendrocyte lineage in the brain. Journal of Biological Chemistry 268:4930-4938. https://doi.org/10.1016/S0021-9258(18)53485-2 Carson, J. H., Nielson, M. L., and Barbarese, E. 1983. Developmental regulation of myelin basic protein expression in mouse brain. Developmental Biology 96(2):485-492. https://doi.org/10.1016/0012-1606(83)90185-9 Cohen, S.R., Herndon, R.M., and McKhann, G.M.1976. Myelin basic protein in cerebrospinal fluid as an indicator of active demyelination. Transactions of the American Neurological Association 101:45-47. Dawson, M. R. L., Levine, J. M., and Reynolds, R. 2000. NG-2-ex-pressing cells in the central nervous system: are they oligodendroglial progenitors? Journal of Neuroscience Research 61(5):471-479. https://doi.org/10.1002/1097-4547(20000901)61:5<471::AID-JNR1>3.0.C0;2-N de Col, R., Messlinger, K., and Carr, R.W.2008. Conduction velocity is regulated by sodium channel inactivation in unmyelinated axons innervating the rat cranial meninges. The Journal of Physiology 586(4):1089-1103. https:// doi.org/10.1113/jphysiol.2007.145383 de Ferra, F., Engh, H., Hudson, L., Kamholz, J., Puckett, C., Mo-lineaux, S., and Lazzarini, R.A.1985. Alternative splicing accounts for the four forms of myelin basic protein. Cell 43(3 Pt 2):721-727. https://doi.org/10.1016/0092-8674(85)90245-4 Feldman, M.L. and Peters, A.1998. Ballooning of myelin sheaths in normally aged macaques. Journal of Neurocytology 27(8):605-614. https://doi. org/10.1023/a:1006926428699 Feng, J.M., Fernandes, A.O., Campagnoni, C.W., Hu, Y.H., and Campagnoni, A. T. 2004. The golli-myelin basic protein negatively regulates signal transduction in T lymphocytes. Journal of Neuroimmunology 152(1-2):57-66. https://doi.org/10.1016/j.jneuroim.2004.03.021 Fernandes, A.O., Campagnoni, C.W., Kampf, K., Feng, J.M., Handley, V.W., Schonmann, V., and Campagnoni, A.T. 2004. identification of a protein that interacts with the golli-myelin basic protein and with nuclear LIM interac-tor in the nervous system. Journal of Neuroscience Research 75(4):461-471. https://doi.org/10.1002/jnr.10882 Fitzner, D., Schneider, A., Kippert, A., Möbius, W., Willig, K. I., Hell, S.W., and Simons, M.2006. Myelin basic protein-dependent plasma membrane reorganization in the formation of myelin. EMBO Journal 25(21):5037-5048. https://doi.org/10.1038/sj.emboj.7601376 Franklin, H., Clarke, B.E., and Patani, R.2021. Astrocytes and microglia in neurodegenerative diseases: Lessons from human in vitro models. Progress in Neuro-biology 200:101973. https://doi.org/10.1016Zj.pneuro-bio.2020.101973 Fritz, R.B. and Kalvakolanu, I.1995. Thymic expression of the golli-myelin basic protein gene in the SJL/J mouse. Journal of Neuroimmunology 57(1-2):93-99. https://doi. org/10.1016/0165-5728(94)00167-m Fünfschilling, U., Supplie, L.M., Mahad, D., Boretius, S., Saab, A. S., Edgar, J., Brinkmann, B. G., Kassmann, C. M., Tzvetanova, I.D., Möbius, W., Diaz, F., Meijer, D., Su-ter, U., Hamprecht, B., Sereda, M.W., Moraes, C.T.,

Frahm, J., Goebbels, S., and Nave, K.A.2012. Glycolytic oligodendrocytes maintain myelin and long-term axo-nal integrity. Nature 485(7399):517-521. https://doi. org/10.1038/nature11007 Galkin, A. P. and Sysoev, E. 1.2021. Stress response is the main trigger of sporadic amyloidoses. International Journal of Molecular Sciences 22(8):4092. https://doi.org/10.3390/ ijms22084092

Harauz, G. and Boggs, J. M. 2013. Myelin management by the 18.5-kDa and 21.5-kDa classic myelin basic protein iso-forms. Journal of Neurochemistry 125(3):334-361. https:// doi.org/10.1111/jnc.12195 Harauz, G. and Libich, D.2009. The classic basic protein of myelin — conserved structural motifs and the dynamic molecular barcode involved in membrane adhesion and protein-protein interactions. Current Protein & Peptide Science 10(3):196-215. https://doi. org/10.2174/138920309788452218 Hartline, D. K. and Colman, D. R. 2007. Rapid conduction and the evolution of giant axons and myelinated fibers. Current Biology 17(1):R29-35. https://doi.org/10.1016/j. cub.2006.11.042 Hasegawa, M., Houdou, S., Mito, T., Takashima, S., Asanu-ma, K., and Ohno, T.1992. Development of myelina-tion in the human fetal and infant cerebrum: a myelin basic protein immunohistochemical study. Brain & Development 14(1):1-6. https://doi.org/10.1016/s0387-7604(12)80271-3 Huxley, A. F. and Stampfli, R.1949. Evidence for saltatory conduction in peripheral myelinated nerve fibres. The Journal of Physiology 108(3):315-339. https://doi. org/10.1113/jphysiol.1949.sp004335 Ishigami, A., Ohsawa, T., Hiratsuka, M., Taguchi, H., Kobayas-hi, S., Saito, Y., Murayama, S., Asaga, H., Toda, T., Kimu-ra, N., and Maruyama, N. 2005. Abnormal accumulation of citrullinated proteins catalyzed by peptidylarginine deiminase in hippocampal extracts from patients with Alzheimer's disease. Journal of Neuroscience Research 80(1):120-128. https://doi.org/10.1002/jnr.20431 Ishiyama, N., Bates, I. R., Hill, C. M., Wood, D. D., Matharu, P., Viner, N.J., Moscarello, M. A., and Harauz, G.2001. The effects of deimination of myelin basic protein on structures formed by its interaction with phosphoinositide-containing lipid monolayers. Journal of Structural Biology 136(1):30-45. https://doi.org/10.1006/jsbi.2001.4421 Jacobs, E.C., Pribyl, T.M., Kampf, K., Campagnoni, C., Col-well, C. S., Reyes, S. D., and Campagnoni, A. T. 2005. Region-specific myelin pathology in mice lacking the golli products of the myelin basic protein gene. Journal of Neuroscience 25(30):7004-7013. https://doi.org/10.1523/ JNEUROSCI. 0288-05.2005 Jacque, C., Delassalle, A., Raoul, M., and Baumann, N. 1983. Myelin basic protein deposition in the optic and sciatic nerves of dysmyelinating mutants quaking, jimpy, Trembler, mld, and shiverer during development. Journal of Neurochemistry 41(5):1335-1340. https://doi. org/10.1111/j.1471-4159.1983.tb00830.x Jang, B., Jin, J.K., Jeon, Y.C., Cho, H.J., Ishigami, A., Choi, K. C., and Choi, E.K.2010. Involvement of peptidylarginine deiminase-mediated post-translational citrullination in pathogenesis of sporadic Creutzfeldt-Jakob disease. Acta Neuropathologica 119(2):199-210. https://doi. org/10.1007/s00401-009-0625-x Kattnig, D. R., Bund, T., Boggs, J. M., Harauz, G., and Hinder-berger, D. 2012. Lateral self-assembly of 18.5-kDa myelin basic protein (MBP) charge component-C1 on membranes. Biochimica et Biophysica Acta 1818(1 1):2636-2647. https://doi.org/10.1016/j.bbamem.2012.06.010

Kim, J.K., Mastronardi, F.G., Wood, D.D., Lubman, D.M., Zand, R., and Moscarello, M.A.2003. Multiple sclerosis: an important role for post-translational modifications of myelin basic protein in pathogenesis. Molecular & Cellular Proteomics 2(7):453-462. https://doi.org/10.1074/ mcp.M200050-MCP200 Kimura, M., Sato, M., Akatsuka, A., Nozawa-Kimura, S., Taka-hashi, R., Yokoyama, M., Nomura, T., and Katsuki, T. 1989. Restoration of myelin formation by a single type of myelin basic protein in transgenic shiverer mice. Proceedings of the National Academy of Sciences USA 86(14):5661-5665. https://doi.org/10.1073/pnas.86.14.5661 Kushnirov, V.V., Dergalev, A.A., and Alexandrov, A.I.2020. Proteinase K resistant cores of prions and amyloids. Prion 14(1):11-19. https://doi.org/10.1080/19336896.20 19.1704612

Landry, C.F., Ellison, J.A., Pribyl, T.M., Campagnoni, C., Kampf, K., and Campagnoni, A.T.1996. Myelin basic protein gene expression in neurons: developmental and regional changes in protein targeting within neuronal nuclei, cell bodies, and processes. Journal of Neuroscience 16(8):2452-2462. https://doi.org/10.1523/JNEUR0-SCI. 16-08-02452.1996 Lefevre-Arbogast, S., Hejblum, B. P., Helmer, C., Klose, C., Ma-nach, C., Low, D.Y., Urpi-Sarda, M., Andres-Lacueva, C., González-Domínguez, R., Aigner, L., Altendorfer, B., Lucassen, P.J., Ruigrok, S. R., De Lucia, C., Du Preez, A., Proust-Lima, C., Thuret, S., Korosi, A., Samieri, C.2021. Early signature in the blood lipidome associated with subsequent cognitive decline in the elderly: A case-control analysis nested within the Three-City cohort study. EBioMedicine 64:103216. https://doi.org/10.1016Zj.ebi-om.2021.103216 Lehmann, P.V., Rottlaender, A., and Kuerten, S.2015. The autoimmune pathogenesis of multiple sclerosis. Die Pharmazie 70(1):5-11. Levine, J. M., Reynolds, R., and Fawcett, J. W. 2001. The oligodendrocyte precursor cell in health and disease. Trends in Neurosciences 24(1):39-47. https://doi.org/10.1016/ s0166-2236(00)01691-x Love, S.2006. Demyelinating diseases. Journal of Clinical Pathology 59(11):1151-1 159. https://doi.org/10.1136/ jcp.2005.031195 Maji, S.K., Perrin, M.H., Sawaya, M.R., Jessberger, S., Riss-man, R.A., Singru, P.S., and Vale, W.2009. Functional amyloids as natural storage of peptide hormones in pituitary secretory granules. Science 325(5938):328-332. https://doi.org/10.1126/science.1173155 Marrodan, M., Alessandro, L., Farez, M.F., and Correale, J. 2019. The role of infections in multiple sclerosis. Multiple Sclerosis Journal 25(7):891-901. https://doi. org/10.1177/1352458518823940 Martini, R., Mohajeri, M.H., Kasper, S., Giese, K.P., and Schachner, M. 1995. Mice doubly deficient in the genes for P0 and myelin basic protein show that both proteins contribute to the formation of the major dense line in peripheral nerve myelin. Journal of Neuroscience 15(6):4488-4495. https://doi.org/10.1523/JNEUR0-SCI. 15-06-04488.1995 Martini, R. and Schachner, M. 1997. Molecular bases of myelin formation as revealed by investigations on mice deficient in glial cell surface molecules. Glia 19(4):298-310. https:// doi.org/10.1002/(SICI)1098-1 136(199704)19:4<298::AID-GLIA3>3.0.C0;2-U McKenzie, I.A., Ohayon, D., Li, H., Faria, L.P., Emery, B., To-hyama, K., and Richardson, W.D.2014. Motor skill learning requires active central myelination. Sci-

ence 346(6207):318-322. https://doi.org/10.1126/sci-ence.1254960

Miller, D.J., Duka, T., Stimpson, C. D., Schapiro, S.J., Baze, W. B., McArthur, M.J., Fobbs, A.J., Sousa, A. M. M., Sestan, N., and Wildman, D.E.2012. Prolonged myelination in human neocortical evolution. Proceedings of the National Academy of Sciences USA 109(41):16480-16485. https:// doi.org/10.1073/pnas.1117943109 Monje, M.2018. Myelin plasticity and nervous system function. Annual Review of Neuroscience 41:61-76. https://doi. org/10.1146/annurev-neuro-080317-061853 Moscarello, M.A., Mastronardi, F.G., and Wood, D.D.2006. The role of citrullinated proteins suggests a novel mechanism in the pathogenesis of multiple sclerosis. Neurochemical Research 32(2):251-256. https://doi. org/10.1007/s11064-006-9144-5 Murray, N. and Steck, A.J. 1984. Impulse conduction regulates myelin basic protein phosphorylation in rat optic nerve. Journal of Neurochemistry 43(1):243-248. https://doi. org/10.1111/j.1471-4159.1984.tb06702.x Muruganandam, G., Burck, J., Ulrich, A.S., Kursula, I., and Kursula, P. 2013. Lipid membrane association of myelin proteins and peptide segments studied by oriented and synchrotron radiation circular dichroism spectroscopy. The Journal of Physical Chemistry B 117(48):14983-14993. https://doi.org/10.1021/jp4098588 Musse, A. A., Boggs, J. M., and Harauz, G. 2006. Deimination of membrane-bound myelin basic protein in multiple sclerosis exposes an immunodominant epitope. Proceedings of the National Academy of Sciences USA 103(12):4422-4427. https://doi.org/10.1073/pnas.0509158103 Nawaz, S., Kippert, A., Saab, A.S., Werner, H.B., Lang, T. Nave, K.A., and Simons, M.2009. Phosphatidylinositol 4,5-bisphosphate-dependent interaction of myelin basic protein with the plasma membrane in oligodendrog-lial cells and its rapid perturbation by elevated calcium. Journal of Neuroscience 29(15):4794-4807. https://doi. org/10.1523/JNEUR0SCI. 3955-08.2009 Nave, K.A. and Werne, H. B. 2014. Myelination of the nervous system: mechanisms and functions. Annual Review of Cell and Developmental Biology 30:503-533. https://doi. org/10.1146/annurev-cellbio-100913-013101 Norton, W.T. and Cammer, W.1984. Isolation and characterization of myelin. Myelin 1967:147-195. https://doi. org/10.1007/978-1-4757-1830-0_5 O'Connor, L.T., Goetz, B.D., Kwiecien, J.M., Delaney, K.H., Fletch, A.L., and Duncan, I.D.1999. Insertion of a ret-rotransposon in Mbp disrupts mRNA splicing and my-elination in a new mutant rat. Journal of Neuroscience 19(9):3404-3413. https://doi.org/10.1523/JNEUR0-SCI. 19-09-03404.1999 Paez, P. M., Fulton, D., Spreuer, V., Handley, V., and Campagnoni, A.T.2011. Modulation of canonical transient receptor potential channel 1 in the proliferation of oligodendrocyte precursor cells by the golli products of the myelin basic protein gene. Journal of Neuroscience 31:3625-3637. https://doi.org/10.1523/JNEUR0-SCI. 4424-10.2011 Paresce, D.M., Chung, H., and Maxfield, F.R.1997. Slow degradation of aggregates of the Alzheimer's disease amyloid beta-protein by microglial cells. Journal of Biological Chemistry 272(46):29390-29397. https://doi. org/10.1074/jbc.272.46.29390 Pedraza, L., Fidler, L., Staugaitis, S. M., and Colman, D. R. 1997. The active transport of myelin basic protein into the nucleus suggests a regulatory role in myelination. Neuron 18(4):579-589. https://doi.org/10.1016/s0896-6273(00)80299-8

Peters, A. 1960. The structure of myelin sheaths in the central nervous system of Xenopus laevis (Daudin). The Journal of Biophysical and Biochemical Cytology 7(1):121-126. https://doi.org/10.1083/jcb.7.1.121 Peters, A.2002. The effects of normal aging on myelin and nerve fibers: a review. Journal of Neurocytology 31(8-9):581-593. https://doi.org/10.1023/a:1025731309829 Polverini, E., Fasano, A., Zito, F., Riccio, P., and Cavatorta, P. 1999. Conformation of bovine myelin basic protein purified with bound lipids. European Biophysics Journal 28(4):351-355. https://doi.org/10.1007/s002490050218 Pribyl, T. M., Campagnoni, C. W., Kampf, K., Kashima, T., Hand-ley, V.W., McMahon, J., and Campagnoni, A.T. 1993. The human myelin basic protein gene is included within a 179-kilobase transcription unit: expression in the immune and central nervous systems. Proceedings of the National Academy of Sciences USA 90(22):10695-10699. https://doi.org/10.1073/pnas.90.22.10695 Pribyl, T.M., Campagnoni, C.W., Kampf, K., Ellison, L.A., Landry, C.F., Kashima, T., McMahon, J., and Campagnoni, A.T. 1996. Expression of the myelin basic protein gene locus in neurons and oligodendrocytes in the human fetal central nervous system. Journal of Comparative Neurology 374(3):342-353. https://doi. org/10.1002/(SICI)1 096-9861(19961 021)374:3<342::AID-CNE3>3.0.CO;2-1 Raasakka, A., Ruskamo, S., Kowal, J., Barker, R., Baumann, A., Martel, A., and Kursula, P. 2017. Membrane association landscape of myelin basic protein portrays formation of the myelin major dense line. Scientific Reports 7(1):1-18. https://doi.org/10.1038/s41598-017-05364-3 Riccio, P., Fasano, A., Borenshtein, N., Bleve-Zacheo, T., and Kirschner, D.A.2000. Multilamellar packing of myelin modeled by lipid-bound MBP.Journal of Neuroscience Research 59(4):513-521. https://doi. org/10.1002/(SICI)1097-4547(20000215)59:4<513::AID-JNR6>3.0.CO;2-M Ridsdale, R.A., Beniac, D. R., Tompkins, T.A., Moscarello, M.A., and Harauz, G.1997. Three-dimensional structure of myelin basic protein. II.Molecular modeling and considerations of predicted structures in multiple sclerosis. Journal of Biological Chemistry 272(7):4269-4275. https:// doi.org/10.1074/jbc.272.7.4269 Ritchie, J. M. 1982. On the relation between fibre diameter and conduction velocity in myelinated nerve fibres. Proceedings of the Royal Society B: Biological Sciences 217(1206):29-35. https://doi.org/10.1098/rspb.1982.0092 Sampaio-Baptista, C., Khrapitchev, A., Foxley, S., Schlagheck, T., Scholz, J., Jbabdi, S., DeLuca, G.C., Miller, K. L., Taylor, A., Thomas, N., Kleim, J., Sibson, N.R., Banner-man, D., and Johansen-Berg, H. 2013. Motor skill learning induces changes in white matter microstructure and myelination. Journal of Neuroscience 33:19499-19503. https://doi.org/10.1523/JNEUROSCI. 3048-13.2013 Sams, E.C.2021. Oligodendrocytes in the aging brain. Neuronal Signal 5(3):NS20210008. https://doi.org/10.1042/ NS20210008

Safaiyan, S., Kannaiyan, N., Snaidero, N., Brioschi, S., Biber, K., Yona, S., Edinger, A. I., Jung, S.J., Rossner, M.J., and Simons, M. 2016. Age-related myelin degradation burdens the clearance function of microglia during aging. Nature Neuroscience 19(8):995-998. https://doi.org/10.1038/ nn.4325

Scholz, J., Klein, M. C., Behrens, T. E.J., and Johansen-Berg, H. 2009. Training induces changes in white-matter architecture. Nature Neuroscience 12:1370-1371. https://doi. org/10.1038/nn.2412

Schonfelder, J., Pfeiffer, B., Pradhan, T., Bijzet, J., Hazen-berg, B.P.C., Schonland, S.O., Hegenbart, U., Reif, B., Haupt, C., and Fandrich, M.2021. Protease resistance of ex vivo amyloid fibrils implies the proteolytic selection of disease-associated fibril morphologies. Amyloid 28(4):243-251. https://doi.org/10.1080/13506129.2021. 1960501

Schlegel, A., Rudelson, J., and Tse, P.2012. White matter structure changes as adults learn a second language. Journal of Cognitive Neuroscience 24:1664-1670. https://doi. org/10.1162/jocn_a_00240 Sedzik, J. and Kirschner, D.A.1992. Is myelin basic protein crystallizable? Neurochemical Research 17(2):157-166. https://doi.org/10.1007/BF00966794 Seiberlich, V., Bauer, N.G., Schwarz, L., Ffrench-Constant, C., Goldbaum, O., and Richter-Landsberg, C.2015. Down-regulation of the microtubule associated protein Tau impairs process outgrowth and myelin basic protein mRNA transport in oligodendrocytes. Glia 63(9):1621-1635. https://doi.org/10.1002/glia.22832 Shine, H.D., Readhead, C., Popko, B., Hood, L., and Sid-man, R.L.1992. Morphometric analysis of normal, mutant, and transgenic CNS: correlation of myelin basic protein expression to myelinogenesis. Journal of Neurochemistry 58(1):342-349. https://doi. org/10.1111/j.1471-4159.1992.tb09316.x Sim, F.J., Zhao, C., Penderis, J., and Franklin, R.J. M. 2002. The age-related decrease in CNS remyelination efficiency is attributable to an impairment of both oligodendrocyte progenitor recruitment and differentiation. Journal of Neuroscience 22(7):2451-2459. https://doi.org/10.1523/ JNEUROSCI. 22-07-02451.2002 Simons, M. and Nave, K.A.2016. Oligodendrocytes: myelination and axonal support. Cold Spring Harbor Perspectives in Biology 8(1):a020479. https://doi.org/10.1101/cshper-spect.a020479

Siu, C.R., Balsor, J.L., Jones, D.R., and Murphy, K.M.2015. Classic and Golli Myelin Basic Protein have distinct developmental trajectories in human visual cortex. Frontiers in Neuroscience 9:138. https://doi.org/10.3389/ fnins.2015.00138 Smith, G. S. T., Samborska, B., Hawley, S. P., Klaiman, J. M., Gil-lis, T.E., Jones, N., and Harauz, G.2013. Nucleus-localized 21.5-kDa myelin basic protein promotes oligodendrocyte proliferation and enhances neurite outgrowth in coculture, unlike the plasma membrane-associated 18.5-kDa isoform. Journal of Neuroscience Research 91(3):349-362. https://doi.org/10.1002/jnr.23166 Sopova, J.V., Koshel, E.I., Belashova, T.A., Zadorsky, S.P., Sergeeva, A.V., Siniukova, V.A., and Galkin, A.P.2019. RNA-binding protein FXR1 is presented in rat brain in amyloid form. Scientific Reports 9(1):1-14. https://doi. org/10.1126/science.1173155 Stadelmann, C., Timmler, S., Barrantes-Freer, A., and Simons, M.2019. Myelin in the central nervous system: structure, function, and pathology. Physiological Reviews 99(3):1381-1431. https://doi.org/10.1152/phys-rev.00031.2018

Steele, C., Bailey, J., Zatorre, R., and Penhune, V.2013. Early musical training and white-matter plasticity in the corpus callosum: evidence for a sensitive period. Journal of Neuroscience 33:1282-1290. https://doi.org/10.1523/ JNEUROSCI. 3578-12.2013 Stoeckenius, W.1959. An electron microscope study of myelin figures. The Journal of Biophysical and Biochemical Cytology 5(3):491-500. https://doi.org/10.1083/jcb.53.491

Thériault, P. and Rivest, S. 2016. Microglia: senescence impairs clearance of myelin debris. Current Biology 26(16):R772-775. https://doi.org/10.1016/jxub.2016.06.066 Tse, K.H. and Herrup, K.2017. Re-imagining Alzheimer's disease — the diminishing importance of amyloid and a glimpse of what lies ahead. Journal of Neurochemistry 143(4):432-444. https://doi.org/10.1111/jnc.14079 Vassall, K.A., Bamm, V.V., and Harauz, G.2015. MyelStones: The executive roles of myelin basic protein in myelin assembly and destabilization in multiple sclerosis. Biochemical Journal 472(1):17-32. https://doi.org/10.1042/ BJ20150710

Williamson, J.M. and Lyons, D.A.2018. Myelin dynamics throughout life: an ever-changing landscape? Frontiers in Cellular Neuroscience 12:424. https://doi.org/10.3389/ fncel.2018.00424 Whitaker, J.N.1977. Myelin encephalitogenic protein fragments in cerebrospinal fluid of persons with multiple sclerosis. Neurology 27(10):91 1-920. https://doi. org/10.1212/wnl.27.10.911 Whitaker, J. N. 1997. Myelin basic protein in cerebrospinal fluid and other body fluids. Multiple SclerosisJournal 4(1):16-21. https://doi.org/10.1 177/135245859800400105 Xie, F., Zhang, J. C., Fu, H., and Chen, J. 2013. Age-related decline of myelin proteins is highly correlated with activa-

tion of astrocytes and microglia in the rat CNS. International Journal of Molecular Medicine 32(5):1021-1028. https://doi.org/10.3892/ijmm.2013.1486

Zand, R., Li, M.X., Jin, X., and Lubman, D.1998. Determination of the sites of posttranslational modifications in the charge isomers of bovine myelin basic protein by capillary electrophoresis- mass spectroscopy. Biochemistry 37(8):2441-2449. https://doi.org/10.1021/bi972347t

Zecevic, N., Andjelkovic, A., Matthieu, J. M., and Tosic, M. 1998. Myelin basic protein immunoreactivity in the human embryonic CNS. Developmental Brain Research 105(1):97-108. https://doi.org/10.1016/S0165-3806(97)00176-4

Zeller, N.K., Hunkeler, M.J., Campagnoni, A.T., Sprague, J., and Lazzarini, R. A. 1984. Characterization of mouse myelin basic protein messenger RNAs with a myelin basic protein cDNA clone. Proceedings of the National Academy of Sciences USA 81(1):18-22. https://doi.org/10.1073/ pnas.81.1.18

Zuchero, J.B., Fu, M.M., Sloan, S.A., Ibrahim, A., Olson, A., Zaremba, A., Dugas, J. C., Wienbar, S., Caprariello, A. W., Kantor, C., Leonoudakis, D., Lariosa-Willingham, K., Kronenberg, G., Gertz, K., Soderling, S.H., Miller, R.H., and Barres, B.A. 2015. CNS myelin wrapping is driven by actin disassembly. Developmental Cell 34(2):152-167. https://doi.org/10.1016Zj.devcel.2015.06.011

i Надоели баннеры? Вы всегда можете отключить рекламу.