Next Article in Journal
Targeting Inflammation and Iron Deficiency in Heart Failure: A Focus on Older Adults
Next Article in Special Issue
The Impact of Urodynamic Findings on Fatigue and Depression in People with Multiple Sclerosis
Previous Article in Journal
High-Throughput Whole-Exome Sequencing and Large-Scale Computational Analysis to Identify the Genetic Biomarkers to Predict the Vedolizumab Response Status in Inflammatory Bowel Disease Patients from Saudi Arabia
Previous Article in Special Issue
Machine Learning for Early Detection of Cognitive Decline in Parkinson’s Disease Using Multimodal Biomarker and Clinical Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hallmarks of Brain Plasticity

by
Yauhen Statsenko
1,2,*,†,
Nik V. Kuznetsov
1,*,† and
Milos Ljubisaljevich
1,3
1
ASPIRE Precision Medicine Institute in Abu Dhabi, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
2
Department of Radiology, College of Medicine and Health Sciences, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
3
Department of Physiology, College of Medicine and Health Sciences, United Arab Emirates University, Al Ain P.O. Box 15551, United Arab Emirates
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2025, 13(2), 460; https://doi.org/10.3390/biomedicines13020460
Submission received: 22 December 2024 / Revised: 15 January 2025 / Accepted: 6 February 2025 / Published: 13 February 2025

Abstract

:
Cerebral plasticity is the ability of the brain to change and adapt in response to experience or learning. Its hallmarks are developmental flexibility, complex interactions between genetic and environmental influences, and structural–functional changes comprising neurogenesis, axonal sprouting, and synaptic remodeling. Studies on brain plasticity have important practical implications. The molecular characteristics of changes in brain plasticity may reveal disease course and the rehabilitative potential of the patient. Neurological disorders are linked with numerous cerebral non-coding RNAs (ncRNAs), in particular, microRNAs; the discovery of their essential role in gene regulation was recently recognized and awarded a Nobel Prize in Physiology or Medicine in 2024. Herein, we review the association of brain plasticity and its homeostasis with ncRNAs, which make them putative targets for RNA-based diagnostics and therapeutics. New insight into the concept of brain plasticity may provide additional perspectives on functional recovery following brain damage. Knowledge of this phenomenon will enable physicians to exploit the potential of cerebral plasticity and regulate eloquent networks with timely interventions. Future studies may reveal pathophysiological mechanisms of brain plasticity at macro- and microscopic levels to advance rehabilitation strategies and improve quality of life in patients with neurological diseases.

Graphical Abstract

1. Glossary

Cerebral plasticity is the ability of the brain to change its activity, structural–functional properties, and adapt in response to in/extrinsic stimuli, experience, learning, or injury.
Activity-dependent plasticity is a form of functional and structural neuroplasticity that arises from cognitive functioning and personal experience [1].
Activity-dependent synaptic plasticity is a modulation of synaptic transmission by repeated nerve impulses [2].
Developmental plasticity is a general term referring to changes in neural connections during development as a result of environmental interactions as well as neural changes induced by learning [3].
Metaplasticity is the ability for synapses to auto-regulate themselves [4].
Natural plasticity is a natural form of plasticity that occurs in physiological conditions due to cyto- and histogenesis, cellular differentiation, the formation of synapses, and the reorganization of the neural circuitry.
Synaptic plasticity refers to the ability of a synapse to change over time through use or disuse [5].
Dendritic structural plasticity is the structural plasticity that occurs at postsynaptic sites in the dendrites and spines of excitatory neurons [5]. Dendritic spines are micron-sized protrusions on the dendritic branches of neurons that host the majority of excitatory synapses in the brain.
Spine plasticity is the biological process by which neuronal activity leads to short- or long-term changes in the morphology and appearance or disappearance of dendritic spines—the specialized protrusions on a neuron’s dendrites that are the sites of excitatory synaptic input. Spine plasticity has been implicated in mediating synaptic plasticity [6].
Post-lesional plasticity occurs after damage to the peripheral or central nervous system, with functional reshaping underlying a partial or complete clinical recovery [2].
Homeostatic plasticity is a mechanism to stabilize the dynamic phenomenon of plasticity and enable the functioning of the system [2]. It refers to the capacity of neurons to regulate their own excitability relative to network activity. The term derives from two opposing concepts, ’homeostasis’ and ’plasticity’; thus, homeostatic plasticity means “staying the same through change” [7].
Cross-modal plasticity refers to the compensation of functional alterations through the recruitment of structures that do not belong to the altered eloquent circuit [8,9,10].
The Hebbian rule states that learning and memory are based on modifications of synaptic strength among neurons that are simultaneously active due to task repetition [11].
Effective connectivity is the experiment- and time-dependent circuit diagram showing the causal influences that neural units exert over one another [12].
Eloquent cortex refers to specific brain areas that directly control function; thus, damage to these areas generally produces major focal neurological deficits. Examples of eloquent cortex are the primary motor cortex (precentral gyrus) and the primary somatosensory cortex (postcentral gyrus).
Neurogenesis is a central mechanism of brain plasticity; it generates new neurons to store and process new information. It is also involved in the formation and consolidation of memories, as well as the development of new skills [13].
Environment plays a crucial role in shaping neural plasticity. Exposure to different environmental factors, including physical, social, and cultural conditions, such as nutrition and stress, education, and lifestyle choices, can impact neural plasticity. They have long-lasting effects on emotional development, cognitive health, and well-being [14,15].
Developmental flexibility is the ability of the brain to adapt and change in response to new experiences and learning throughout its lifespan. It is crucial for cognitive and behavioral development [16].
Synaptic remodeling is the process by which connections between neurons in the brain are changed in response to alterations in neural activity. It plays a key role in cerebral plasticity and the brain’s ability to change and adapt throughout its lifespan [17].
Axonal sprouting is a process by which new axons grow and form connections in the brain. Axonal sprouting is a part of neuroplasticity that mediates the ability to learn and allows neurons to adapt in response to new experiences or changes in the environment. The process can occur in response to injury, disease, or changes in brain activity [18,19].
Oligodendrogenesis is the process of creating new oligodendrocytes and generating myelin around axons, which allows for faster and more efficient communication between neurons. In adults, oligodendrocytes continue to produce myelin important for maintaining healthy brain function [20,21].

2. Introduction

Numerous studies show that experience and lesions of the peripheral or central nervous system can modulate functional cortical organization [2]. Hence, the brain is a dynamic organ, which implies the ability of a network of neural connections to self-modify in response to experience [22]. Brain plasticity (BP) refers to the brain’s ability to optimize the functioning of brain networks through the reorganization of neurosynaptic maps [1,2]. BP is a continuous process through which remodeling of the maps can be short-, middle-, and long-term [23]. The capacity of the brain to change structurally and/or functionally allows an individual to learn, remember, forget, and recover from injury [1,24]. Therefore, BP is a compensatory phenomenon [2]. BP changes throughout its lifespan. It is enhanced in children and reduced in adults [25]. Herein, we summarize views on the pathophysiology of cerebral plasticity at a sub-cellular, cellular, and synaptic map levels.

2.1. Concept of Brain Plasticity

Neuroplasticity is the ability of the brain to change structurally and functionally [24]. Experience may produce multiple dissociable modifications to the neural system (see Figure 1). These refer to an increase in dendritic length and glial cell activity, a change in spine density, synapse formation, and altered metabolic activity. These variations change the brain’s weight, cortical thickness, acetylcholine levels, and dendritic structures. Structural modulation impacts behavior. Age, hormonal profile, trophic factors, stress, and brain pathology also affect the functional outcomes.
The key principle of behavioral neuroscience is that experience can modify brain structure long after brain development is complete [24]. In response to behavioral demands, the mammalian brain can form new synapses, grow dendrites, and create new elements of supportive tissue such as astrocytes and blood vessels [13,26]. Environmental enrichment studies show large changes in various measures of cortical morphology. In these studies, a control group of animals is kept in laboratory cages. Contrarily, the experimental group is placed in large enclosures with visually stimulating objects and an opportunity to interact with the environment. The studies report an increase in the dendritic fields of neurons by 20% relative to cage-reared animals. Dendric space correlates closely with synaptic numbers [27,28,29].
Moreover, experience (environmental enrichment) modulates synapses by modifying the excitatory–inhibitory equilibrium. Specifically, the number of excitatory synapses per neuron increases, and the number of inhibitory synapses decreases. Changes in neuronal morphology require a more active metabolism, blood supply, and support from glial cells, especially from astrocytes [24].
Merely having exercise is not sufficient to induce neuronal changes. A more complex task increases neuronal processing, which results in a more active synapse formation [24]. Environmental enrichment increases both the dendritic length and density of synaptic spines on dendrites. Some authors found an association between extent of dendritic arborization in a cortical language area and amount of education [30]. In another experiment, children with a developmental delay had spindly dendrites with reduced spine density compared to average intelligence children [31].

2.2. Different Types of Plasticity

BP can be classified in different ways. For example, scientists mention ’activity-dependent plasticity’, and the brain’s ability to communicate with itself [1]. The brain’s ability to alter the structural and functional properties of neurons refers to its structural and functional plasticity [32]. Researchers presume that a synapse is the most likely place to identify neural changes associated with behavior [24]. Synaptic plasticity is the ability of a synapse to change over time through use or disuse. Meanwhile, dendritic plasticity occurs at postsynaptic sites in the dendrites and spines of excitatory neurons [5].
In the glossary section, we listed a broader classification of BP into subtypes. Still, physicians mainly focus on post-lesional plasticity which is the ability to adapt after damage to the peripheral or central nervous system due to functional reshaping underlying a partial or complete clinical recovery [2]. Although BP is a dynamic construct, it should be stabilized through a mechanism called homeostatic plasticity; otherwise, the system will not be functional [2].

3. Neuroanatomic and Neurophysiologic Bases of Brain Plasticity

3.1. Plasticity in the Periphery and at the Centrum of the Brain

In local anesthesia, amputation, and peripheral neuropathy, sensory deprivation is the major reason for cerebral reorganization. The adjacent regions of the cortex expand at the expense of the deprived cortex. The suggested mechanism of expansion is as follows. Within certain minutes after trauma, acute reorganization occurs due to the unveiling of latent intracortical connections. In the months that follow, additional remodeling happens. In the primary motor cortex, a peripheral lesion also results in expansion of the cortical areas in the vicinity of the representation of the body part that is injured [2].
After the formation of lesions in the primary somatosensory area of the brain, the damaged representations are redistributed both in remote regions and in the areas adjacent to the injury [2]. Regarding motor function, animal studies have also demonstrated a similar experience-dependent plasticity after the formation of central lesions. Research shows the potential for rehabilitative training to shape remodeling in the adjacent undamaged cortex [2]. Hence, the recruitment of the intact motor cortex is a mechanism of motor recovery [33].

3.2. Natural Plasticity in Different Functional Areas

Changes in plasticity differ among functional areas of the brain. The primary motor cortex controls the kinetic and dynamic parameters of voluntary movement [34], and cortical representations of muscles and movements have a mosaic structure [35]. Motor training reshapes the primary motor cortex, and the acquisition of new motor skills necessitates an extension of activation which is reached by the temporal or durable recruitment of adjacent sites [36].
The primary sensorimotor cortex integrates sensory and motor signals necessary for skilled movement, namely, those involved in cognitive functions such as learning motor skills [37], making calculations [38], and employing mental imagery [39]. Hence, the role of the sensorimotor cortex is more complex than the control of movement. Learning a skill modifies the activity of isolated neurons and brain regions. The synchronous activity of many neurons in the same cortical region may quickly change the time-course of the ensemble of neurons executing the movement [40,41]. The non-primary parts of the somatosensory network also undergo plasticity-related changes, and the effective connectivity within the whole functional network rises [42,43].
The functional areas of language and cognition are cortico–cortical and cortico–subcortical networks that act in parallel. The areas have a hierarchy with both simultaneous and successive activation of the networks. Some of them are essential while others are compensative [44,45,46]. Plasticity implies the modification in the spatio–temporal parameters of the networks functioning.

4. Pathophysiological Mechanisms Underlying Cerebral Plasticity

4.1. Plasticity Mechanisms at the Microlevel

At the microscopic level, many ultrastructural and synaptic changes may take place. During neurodevelopment, these are cyto- and histogenesis with proliferation and elaboration of dendritic and axonal branches; cell migration, formation of synapses, cellular differentiation; precise organization of the circuitry; apoptosis; regression of axons; and the elimination of cells and synapses. At this stage, radial glia control neuronal migration from the subventricular zone to the cortex; thus, they also contribute to developmental plasticity [47]. After the period of neurodevelopment, structural and functional reorganization of the brain may proceed, with the major changes taking place at the synaptic level. The plasticity mechanisms include changes in the activity of isolated neurons, in synaptic efficacy, and in the temporal relations between ensembles of neurons in specific oscillation bands [48]. Combined, these mechanisms can modulate behavior [2,40,49].
The synapse is a dynamic, rather than a static, contract. Beyond its increase in size and number due to learning [50], one can see modulations in synaptic strength, which evidence the presence of plastic properties in these dynamic connections [51]. Once appear at the microscopic level, these modulations account for functional map reshaping at the macroscopic level [52]. They exemplify activity-dependent synaptic plasticity and the auto-regulation of synapses, called ‘metaplasticity‘.
Activity-dependent synaptic plasticity is a leading mechanism of memory formation. Repeated nerve impulses change synaptic transmission; frequent stimuli traveling to the presynaptic membrane may increase or decrease the induced excitation of the postsynaptic neuron. Activity-dependent synaptic plasticity establishes a real-time control over the flow of information within neuronal networks [2]. This type of plasticity explains two opposite phenomena: long-term potentiation and long-term depression. Long-term potentiation is the durable enlargement of synaptic strength followed by brief high-frequency stimulation. Otherwise, such stimulation might lead only to short-term potentiation. The mechanism was demonstrated in the hippocampus and motor cortex, and it may underlie functional plasticity in the motor cortex [53]. Long-term depression plays an important role in learning and memory [54].
Metaplasticity is the ability of synapses to auto-regulate themselves [4,55,56]. Different hypotheses describe the mechanisms of memory formation through the modulation of synapses. According to the synaptic plasticity and memory hypothesis, the induction of activity-dependent synaptic plasticity at the appropriate synapses forms the memory [56,57]. However, little evidence supports the efficacy of activity-dependent synaptic plasticity for storing memory [2]. According to the Hebbian rule, the physiological basis for learning and memory are modifications in synaptic strength among neurons that are simultaneously activated when a task is repeated [11]. This rule is widely accepted in the field of neuroscience [58]. Moreover, scientists have discovered a mechanism essential for balancing the processes of Hebbian learning. It is synaptic stabilization through the regulation of AMPA receptors mediating fast synaptic transmission [59]. This self-regulation of neuronal excitability relative to network activity is called ’homeostatic’ plasticity. The term derives from two opposing concepts and means “staying the same through change”.
The synchronization of episodic electrostimulation of the cerebral ganglia is necessary for massive reorganization of the cortex [60]. ’Effective’ connectivity refers to influences among brain regions. Biomathematical modeling is used to determine how a constrained set of brain regions influence each other in a specific task. Knowledge of these regions comes from neuroanatomy [12]. A study showed a synchronization of activity among different areas involved in sensorimotor function due to training [43,61]. Hence, plasticity may appear as a modification in ’effective’ connectivity within the whole functional network [42].
Another major mechanism of short-term plasticity is the decrease in inhibitory activity in the GABA interneurons that block horizontal connection in regular settings [62]. However, sensory deprivation or learning suppresses the GABA inhibition, which unmasks latent connections and transforms silent synapses into functional ones [11,63]. Tha-lamo–cortical networks facilitate this process [64].
Glia can also affect synaptic transmission, coordinate activity across neuronal networks, and modulate neuronal activity in different ways. These include the release of neurotransmitters and other signaling molecules and neurovascular coupling, which regulates energy metabolism [65]. In addition, glial cells can communicate with each other, thus, they form a glial network that is able to both listen to and communicate with neurosynaptic circuits [66].
At the neuronal level, structural modifications include sprouting of the dendritic spine, growing of the axon, and forming new synapses (neosynaptogenesis). Experience or brain damage may initiate these modifications. Experience-dependent plasticity is based on increased synapse turnover which denotes the accelerated formation and elimination of synapses. This mechanism underlies the adaptive remodeling of neural circuits [67].
Post-injury plasticity is based on the rapid induction of changes in the number, size, and shape of dendritic spines [68,69]. The suggested molecular mechanisms for this are protein synthesis [69], the secretion of growth factors and neurotrophins [70]. AMPA receptors and integrins stabilize morphological changes through a mechanism of ’homeostatic’ plasticity [50,71]. Axons may also spontaneously regenerate and elongate [72]. Glia control the number of synapses [73] and adjust to meet modifications in the brain environment [74]. In both physiological conditions and after injury, changes to the glial cell size and phenotype are quick (within hours) [67,75]. The changes can be conveyed to other glial cells via connexin [76].
Researchers have begun to question the old dogma that the adult mammalian brain cannot develop new neurons. The olfactory bulb, the dentate gyrus, and even the neocortex of adult primates are exceptions to this rule which has turned out not to be absolute [26,77,78]. In vitro, multipotential progenitor cells of adult humans underwent neurogenesis. The cells were isolated from the temporal neocortex, hippocampus, and subcortical white matter [79,80,81]. Studies suggest that these newly created neurons may store memories and contribute to learning via changes to neurosynaptic circuits and the formation of new connections and networks [82]. Post-lesional plasticity can also be arranged by way of neurogenesis, as shown in adult rats. The animals generated endogenous neural precursor cells in situ and differentiated into mature neurons, replacing the damaged ones [83]. This fact supports the idea of neuronal replacement therapies.

4.2. Plasticity Mechanisms at the Macrolevel

At the macroscopic level, functional reorganization is carried out through the mechanisms of diaschisis, functional reorganization of the cortex within eloquent areas and networks, cross-modal plasticity, compensatory strategies, and macroscopic morphological changes. Diaschisis is a general term that describes functional alterations outside of focal brain damage. These are electrophysiological, metabolic, and hemodynamic changes. Although diaschisis underlies initial functional impairment, the same mechanism accounts for spontaneous functional recovery after injury [84,85].
Another mechanism of functional reorganization after brain injury affects the eloquent cortex. Eloquent areas are redundant representations of the same function within the same region. Within eloquent areas, functions have multiple cortical representations within the same region. Therefore, the eloquent site is discrete and, once partially destroyed, it is compensated by adjacent redundant sites that are unmasked post-injury [86,87]. However, in wide lesions, this mechanism does not provide sufficient compensation; therefore, other cortical parts are recruited to restore function [88]. These are regions of the same functional networks, remote ipsi-hemispheric structures, and functional homologous structures in the contralateral hemisphere. If functional compensation is insufficient, suppression of the regions is released step-by-step with the unmasking of each subsequent region [89].
‘Cross-modal plasticity’ refers to the compensation of functional alterations through the recruitment of the structures that do not belong to the eloquent circuit that was altered [8,9,10]. For example, deaf patients may activate the auditory cortex during somatosensory tasks and, in this way, they have better tactile discrimination [90]. For the same reason, these individuals may benefit less from a cochlear implant due to extensive cross-modal plasticity [91]. If unimodal areas cannot be recruited after massive damage, heteromodal associations among cortex areas are activated. Although this activation does not allow for complete functional restoration, this mechanism can be considered an elaboration of compensatory cognitive strategies [92].
Although mainly occuring at the ultrastructural level, neurogenesis may result in macrostructural changes that can be detected with voxel-based morphometry [93]. With this technique, scientists have shown that cortical regions, the cerebellum, the hippocampus, and the density of white matter tracts in the predominant hemisphere can be enlarged to meet professional or educational demands [94,95,96,97,98,99]. In the grey matter, training can induce transient morphological changes [100].

5. Modulation of Experience-Dependent Change

5.1. Modulation by Sex Hormones

Studies report that the brain is more sensitive to experience in females than in males [101,102,103,104]. Some studies suggest that females may have a more densely packed hippocampus with a higher number of neurons relative to its size. However, these disproportions can be manipulated with hormonal replacement therapy [24]. A failure of dendritic growth is a supposed pathophysiologic mechanism in the development of dementia [105].

5.2. Neurodevelopment and Brain Plasticity in Childhood

The superior ability of children to learn a language and to recover from brain trauma demonstrate enhanced brain plasticity compared to adults [1]. During the early years, several mechanisms account for enhanced brain plasticity. First, neurogenesis does not stop immediately after birth, although adult neurogenesis is absent in humans [106]. Second, programmed cell death (apoptosis) may eliminate neurons [107]. Third, the number of synapses may either increase or decrease, and synaptic functioning can be refined by activity-dependent mechanisms [5,25].
In children, plasticity of the brain is maximal, and it can be classified into the following categories: adaptive, impaired, or excessive plasticity, and plasticity that makes the brain vulnerable to injury [1]. The first category refers to adjustments in neuronal circuitry that allow an individual to compensate for injuries to the brain or develop a special skill with practice. The second is linked to cognitive impairment, when genetic or acquired disorders disrupt molecular plasticity pathways. In contrast, excessive plasticity leads to disability through the reorganization of maladaptive neuronal circuits in the developing brain. These new maladaptive brain circuits cause neurologic disorders such as partial seizures following mesial temporal sclerosis or focal dystonia. Finally, brain plasticity can be an ’Achilles’ heel’ and increase the vulnerability of the brain to injury. In energy failure or status epilepticus, the mechanisms regulating plasticity are over-stimulated, which leads to excitotoxic neuronal damage.

5.3. Brain Plasticity in Adulthood

The brain holds the potential for functional and structural rearrangement through- out its lifespan, which has been underestimated recently [108]. In adults, learning induces the elaboration of new circuits and the maintenance of neural networks. In elderly people, natural plasticity may resist negative outcomes of brain aging, which typically results in neurocognitive slowing [22,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126]. In normal aging, the number of synapses increases in the cortex, which allows middle-aged people to compensate for the loss of neurons with age and to maintain the number of synapses throughout their lifetime [24].

6. Molecular Mechanisms of Brain Plasticity

Despite vast molecular profiling and omics studies of brain structures and functions, only certain molecular alterations may serve as biomarkers of plasticity changes in the brain. The newly discovered world of non-coding RNAs (ncRNAs) is constantly expanding into all areas of biomolecular interaction and a variety of cellular processes, including control over the metabolism, gene regulation, and protein turnover. The discovery of the essential role of microRNAs in gene regulation was recently recognized and awarded a Nobel Prize in Physiology or Medicine in 2024.
Logically, multiple ncRNA players are found to be involved in brain plasticity (Table 1). Furthermore, interactions between different types of ncRNAs create multidimensional networks that respond to a range of endogenous and exogenous stimuli. Non-coding RNAs represent a major part of the transcriptome. Various classes of ncRNAs have emerged as critical regulators of transcription, epigenetic processes, and gene silencing. These molecules play an important role in neural brain plasticity, brain homeostasis, and cognitive processes [127]. Non-coding RNAs regulate diverse intracellular and neuronal functions: they modulate chromatin structure, act as chaperones, and contribute to synaptic remodeling and behavior [128].
Neurons are highly compartmentalized because of their morphological and functional complexity. This occurs due to the transport of messenger RNA (mRNA) transcripts to specific subcellular areas, e.g., synaptic regions, for local translation. Increasing evidence shows that highly expressed cerebral ncRNAs participate in the spatial and temporal control of mRNA translation and, therefore, in synaptic plasticity [129].
Non-coding RNAs may contribute to the development of a variety of neuropsychiatric disorders, including schizophrenia, addiction, and fear-related anxiety disorders [127,128]. Moreover, the diversity of ncRNAs and their association with neurodegenerative diseases renders them particularly interesting as putative targets of brain disease [130]. New RNA-based therapeutics can be developed due to this new knowledge of ncRNA regulation and the downstream effects of its interactions in different pathologies.
Table 1. Examples of non-coding RNAs involved in brain plasticity.
Table 1. Examples of non-coding RNAs involved in brain plasticity.
NoName (Acronym)Molecular SpeciesReferences
1Long non-coding RNA (lncRNA)Gomafu, GAS5, MALAT1, HOTAIR[131,132,133,134,135]
2MicroRNA (miRNA)miR-9, miR-34, miR-132[136,137,138]
miR-17-92 cluster[139,140]
miR-144-5p, miR-145, miR-153[141,142,143]
hsa-miR-1-3p,
hsa-miR-335-5p,
hsa-miR-34a-5p
[144]
3Circular RNA (circRNA)ciRS-7, circRMST, circFAT3[145]
circIgfbp2[146]
nearly 1167 cerebral circRNAs[147]
cirC_0000400, cirC_0000331,
cirC_0000406, cirC_0000798
[148]
4Enhancer RNA (eRNA)Bdnf-Enhg1, Bdnf-Enhg2[149]
Evf2[150]
5Long intergenic non-coding RNA
(lincRNA)
linc-Brn1b[151]
Xist[152]
6Piwi-interacting RNA (piRNA)list of 1251 brain-specific piRNAs;
piR-hsa-1281, piR-hsa-1280,
piR-hsa-1282, piR-hsa-27492
[153,154,155]
7Y RNA (yRNA)nELAVL/Y RNA complex hY1, hY4, hY5[156,157,158]

6.1. Long Non-Coding RNAs

Long non-coding RNAs (lncRNAs) act as scaffolds for biomolecule binding and mediate different RNA–protein interactions. LncRNAs are increasingly recognized for their involvement in neurodevelopmental processes, including cell proliferation, neurite outgrowth, synaptogenesis, and neuroplasticity [159]. Neuronal lncRNAs are crucial for orchestrating neurogenesis, tuning neuronal differentiation, and establishing the exact calibration of neuronal excitability [130]. In particular, Malat1 is an lncRNA that is abundant in the nuclei of neurons. It promotes synapse formation by recruiting the serine/arginine splicing factors to the transcription sites of genes involved in synaptogenesis. In vitro, overexpression of Malat1 enhances the number of synapses in hippocampal neurons while its deficiency reduces the number of synapses between dendrites and axons [160,161]. Gomafu is another lncRNA involved in ES cell, neuronal cell, and retinal cell differentiation. A lack of Gomafu led to a hyperactive phenotype and increased sensitivity to the psychostimulant MAP in Gomafu KO mice [131]. The lncRNA Gas5 promotes the neuronal differentiation of hippocampal NSCs and restores learning and memory in rats with cholinergic injury [132]. Furthermore, synapse-specific Gas5 KO impairs fear extinction memory [162].

6.2. MicroRNAs

MicroRNAs (miRNAs) are small (about 20–25 nucleotides in length) endogenous RNAs that regulate gene expression post transcription [163,164]. They are commonly present in specific brain regions and affect nervous system development, plasticity, and function [165]. For example, miR-9 has a critical role in hippocampal synaptic plasticity and memory [136], miR-34 regulates synaptogenesis [137], and miR-132 participates in axon growth, neural migration, and plasticity [138]. In the temperament–character molecular integration network (TCMIN), three miRNAs (hsa-miR-1-3p, hsa-miR-335-5p, and hsa-miR-34a-5p) are sufficient to coordinate interactions between two gene networks. The first network performs the self-regulation of emotional reactivity to extracellular stimuli (e.g., self-regulation of anxiety). The second one interprets meaning (e.g., produces concepts and language) [144].
Potential targeting or therapeutic use is demonstrated for several miRNAs [166]. In particular, the miR-17-92 cluster enhances neuroplasticity [139] and regulates adult hippocampal neurogenesis, anxiety, and depression [140]. miR-144-5p is currently considered a key target in major depressive disorder [141], and miRNA-145 was recently shown to enhance neural repair after spinal cord injury [142]. One of the highly conserved miRNAs in mice and humans, miRNA-153, stabilizes the neurogenesis of neural stem cells and enhances cognitive ability through the Notch signaling pathway [143].

6.3. Circular RNAs

Circular RNAs (circRNAs) are closed structural isoforms of linear mRNA. They are abundant in the brain and play a significant role in the development of the nervous system [167]. Cerebral circRNAs are linked with neurotransmitter function, synaptic activity, and neuronal maturation. The levels of ciRS-7, circRMST, and circFAT3 increase during the differentiation of human embryonic stem cells into rostral and caudal neural progenitor cells [145]. The level of a recently discovered circRNA, circIgfbp2, significantly increases in injured brain tissue. It is involved in neural plasticity. Therefore, circIgfbp2 can be a future therapeutic target for anxiety and sleep disorders after traumatic brain injury [146]. At least four circRNAs (cirC_0000400, cirC_0000331, cirC_0000406, and cirC_0000798) are involved in postoperative neurocognitive disorders [147]. In a rat model, a large number of circRNAs, including 1167 cerebral circRNAs, displayed a developmental-dependent expression pattern. They may have an important biological function in differentiation, development, and aging [148].

6.4. Enhancer RNAs

Enhancer RNAs (eRNAs) are long non-coding RNAs bidirectionally transcribed by RNA polymerase II from enhancer regions of the genome. Generally, eRNAs are not spliced or polyadenylated [168,169,170]. Bdnf-Enhg1 and Bdnf-Enhg2 are characterized as novel enhancers that regulate Bdnf expression in developing neurons [149]. Conserved enhancer Evf2 to functionally and spatially organizes megabase distant genes in the developing forebrain [150].

6.5. Long Intergenic Non-Coding RNAs

Long intergenic non-coding RNAs (lincRNAs) are biochemically identical to other lncRNAs but differ in their genomic organization as they reside in the space between genes [171]. A knockout of linc-Brn1b shows a reduced number of intermediate progenitor cells in the subventricular zone. This suggests that linc-Brn1b can be involved in the development of the cortex [151]. The long non-coding RNA X-inactive specific transcript (XIST) is a promising molecular target for SCI therapy [172]. It may have a significant role in AD [152].

6.6. Piwi-Interacting RNAs

Piwi-interacting RNAs (piRNAs) are a class of Piwi-associated, small (26–32 nucleotide-long) non-coding RNAs. Unlike other small RNAs, they are generated from long genomic clusters [173,174,175]. Piwi-interacting RNAs are a part of a gene regulatory mechanism responsible for establishing stable long-term changes in neurons and the persistence of memory in the brain’s synaptic plasticity [153]. The main molecular function of piRNAs is to regulate transposons. The co-existence of piRNA and retrotransposons might play an important role in brain development and the adult brain [154]. A number of piRNAs across brain transcriptome are associated with Alzheimer’s disease [155].

6.7. Y RNAs

Y RNAs (yRNAs) are a class of non-coding RNAs often found abundantly expressed in the brain and neuronal tissue. Y RNAs are linked to neuronal stress, and they are very often associated with neuronal ELAV-like proteins in Alzheimer’s disease patients [156]. Y RNAs can serve as biomarkers in glioma [157]. A recent study suggests that the strong tendency of yRNAs to bind to nELAVL proteins in response to stress conditions may prevent these proteins from associating with their normal messenger RNA targets [156].

7. Brain Neuroplasticity from a Network Neuroscience Perspective

The initial concept of systematically analyzing the relationships between structure, function, and plasticity of the brain was introduced in 1976 [176]. Now, the concept evolves and expands, creating several promising research areas [177,178]. High-resolution neuroimaging, modeling of neuronal dynamics, and graph theory applications are among the most important approaches. An exhaustive analysis of research tools and advances in brain network neuroscience is beyond the scope of this review. Instead, below are a few points and references for further consideration.

7.1. High-Resolution Neuroimaging Systems and Techniques

High-resolution neuroimaging systems and techniques include functional MRI (fMRI) [179,180,181], voxel-based morphometry (VBM) [182,183,184], diffusion tensor imaging (DTI) [185,186,187], electroencephalography (EEG) [188,189,190], magnetoencephalography (MEG) [191,192,193], optical coherence tomography (OCT) [194,195,196], positron emission tomography (PET), and single-photon-emission tomography (SPET) [197,198,199,200].

7.2. Computational Models of Neuronal Dynamics

Computational models are presented with the classic ‘integrate-and-fire’ model [201,202], the conventional [203] and stochastic versions of Hodgkin–Huxley model [204], and a number of neuronal networks that link cellular mechanisms of neuromodulation to large-scale neural dynamics [205,206,207,208,209,210].

7.3. Graph Theory Applications

Graph theory can be applied to the identification of brain network modules where nodes in graphs can be individual neurons or brain regions [211,212]. The future of graph theory leads to further application of generative models, dynamic networks, multilayer systems, and principles of topology [213]. In particular, graph theory provides insight into early biomarker diagnostics by elucidating the reorganization of brain networks and revealing topological changes associated with neurodegenerative diseases [214].

8. Implications for Medical Practice

8.1. Pharmacology

In a study, the administration of piracetam potentiated post-lesional plasticity, thus playing a neuroprotective role [215]. Transgenic mice models enabled modulation of plasticity. In humans, different drugs can improve brain reshaping after a stroke or brain trauma [216]. The beneficial effects were observed in different functional areas. Norepinephrine, fluoxetine, paroxetine, scopolamine, and lorazepam improved cortical motor plasticity [217,218,219,220,221]. Meanwhile, amphetamine, bromocriptine, and piracetam reactivated brain regions in the left hemisphere and facilitated recovery after aphasia by modulating activity in the language centers [222,223,224].

8.2. Transcranial Magnetic Stimulation

In post-stroke rehabilitation, transcranial magnetic stimulation (TMS) can potentiate motor learning [225]. It can rapidly elevate excitability in the primary motor cortex with long-lasting effects [226]. The same technique modulates sensory maps, as it can eliminate the deficit of spatial awareness in the contralesional space [227,228]. TMS can facilitate cognitive rehabilitation by improving memory and performance in picture-naming, analogic reasoning, and decision-making tasks [225,229,230,231,232]. Its mechanism of action is based on a modulation in effective connectivity [233]. A combination of TMS, pharmacological intervention, and rehabilitation is suggested.

8.3. Surgery

Cortical stimulation is an intervention performed for different indications. High-frequency chronic cortical stimulation efficiently modulates functional networks in movement disorders and chronic pain [234,235,236] This technique improves the functioning of the subcortico–cortical loops which relieve motor, cognitive, and behavioral symptoms in people with Parkinson’s disease [237,238].
Surgical resection redistributes functional activity throughout latent networks. Hence, an incomplete removal of a tumor in eloquent areas reshapes the eloquent maps and extends functional sites. In a few years, during a second surgery, the extended resection will not induce sequelae due to the recruitment of latent networks, unmasked after the first resection [239,240,241,242,243,244,245]. This approach allows for the extension of indications for surgery to ’non-operable’ eloquent regions (sensorimotor and language areas). Still, cortical plasticity may manifest only if subcortical connectivity is not altered. Therefore, a stroke can cause permanent deficits due to damage to the white matter [246]. For the same reason, the resection of subcortical pathways may display sequelae despite the potential for plasticity in the cortex [247,248,249].

8.4. Transplantation

Observations of neural grafts explain how environment and experience can modulate brain function [250]. For example, the transplantation of neuroblasts from the fetal striatum to the same brain region may treat Huntington’s disease. The graft enhances cognitive performance and motor function by strengthening connections in the striato–cortical loop [251]. In people with Parkinson’s disease, the transfer of dopaminergic neural cells to the putamen shows promising results [252]. After basal ganglia infarction, the graft comprising cultured human neuronal cells can reduce motor deficit [253].

9. Conclusions

The brain is a dynamic construct that changes structurally and/or functionally and constitutes interactive distributed glial–neuro–synaptic networks. The behavioral consequences of these changes may vary as a function of their effective connectivity, but the overall system remains stable due to homeostatic plasticity.
New insight into the concept of brain plasticity and homeostasis may provide additional perspectives on functional recovery following brain damage. Knowledge of this phenomenon will enable physicians to exploit the potential of cerebral plasticity and regulate eloquent networks with timely interventions. Future studies may reveal pathophysiological mechanisms of brain plasticity at microscopic and macroscopic levels, which will advance rehabilitation strategies and improve quality of life in patients with neurological disease.
Non-coding RNAs are optimal candidates for elucidating the molecular pathways underlying the phenomenon of brain plasticity. Candidates may signal the development of various neuropsychiatric disorders comprising schizophrenia, addiction, and fear-related anxiety disorders. The diversity of ncRNAs and their association with neurodegenerative disease renders them particularly interesting targets for new therapeutic approaches. New RNA-based therapeutics may arise from novel data on ncRNA regulation and the downstream effects of their interactions.

Author Contributions

Conceptualization—M.L.; writing (original draft preparation)—Y.S. and N.V.K.; writing (review and editing)—N.V.K. and Y.S.; problem investigation—Y.S. and N.V.K.; figure preparation—N.V.K.; final version editing—N.V.K.; supervision—M.L.; project administration—Y.S. and M.L.; funding acquisition—M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by ASPIRE, the technology program management pillar of Abu Dhabi’s Advanced Technology Research Council (ATRC), via the ASPIRE Precision Medicine Research Institute Abu Dhabi (ASPIREPMRIAD), award grant number VRI-20-10.

Institutional Review Board Statement

Ethics approval is not required for this review. The results of the study will be presented at scientific conferences as a poster or presentation in addition to being published in a peer-reviewed journal.

Informed Consent Statement

Patients or the general public are not participants in the study.

Conflicts of Interest

This research was carried out without any financial or commercial ties that may be viewed as having a possible conflict of interest, according to the authors.

Abbreviations

The following abbreviations are used in this manuscript:
ADAlzheimer’s disease
BPbrain plasticity
circRNAcircular RNA
eRNAenhancer RNA
lincRNAlong intergenic non-coding RNA
lncRNAlong non-coding RNA
mRNAmessenger RNA
miRNAmicroRNA
ncRNAnon-coding RNA
piRNAPiwi-interacting RNA
TMStranscranial magnetic stimulation
yRNAY RNA

References

  1. Johnston, M.V. Clinical disorders of brain plasticity. Brain Dev. 2004, 26, 73–80. [Google Scholar] [CrossRef]
  2. Duffau, H. Brain plasticity: From pathophysiological mechanisms to therapeutic applications. J. Clin. Neurosci. 2006, 13, 885–897. [Google Scholar] [CrossRef] [PubMed]
  3. Kolb, B.; Gibb, R. Brain plasticity and behaviour in the developing brain. J. Can. Acad. Child Adolesc. Psychiatry 2011, 20, 265. [Google Scholar]
  4. Fischer, T.M.; Blazis, D.E.; Priver, N.A.; Carew, T.J. Metaplasticity at identified inhibitory synapses in Aplysia. Nature 1997, 389, 860–865. [Google Scholar] [CrossRef] [PubMed]
  5. Forrest, M.P.; Parnell, E.; Penzes, P. Dendritic structural plasticity and neuropsychiatric disease. Nat. Rev. Neurosci. 2018, 19, 215–234. [Google Scholar] [CrossRef] [PubMed]
  6. Nature Portfolio. Spine Plasticity—Nature, Subjects. 2024. Available online: https://www.nature.com/subjects/spine-plasticity (accessed on 22 June 2024).
  7. Wikipedia Contributors. Homeostatic plasticity—Wikipedia, The Free Encyclopedia. 2024. Available online: https://en.wikipedia.org/w/index.php?title=Homeostatic_plasticity&oldid=1216627734 (accessed on 22 June 2024).
  8. Kujala, T.; Alho, K.; Näätänen, R. Cross-modal reorganization of human cortical functions. Trends Neurosci. 2000, 23, 115–120. [Google Scholar] [CrossRef]
  9. Shimojo, S.; Shams, L. Sensory modalities are not separate modalities: Plasticity and interactions. Curr. Opin. Neurobiol. 2001, 11, 505–509. [Google Scholar] [CrossRef]
  10. Bavelier, D.; Neville, H.J. Cross-modal plasticity: Where and how? Nat. Rev. Neurosci. 2002, 3, 443–452. [Google Scholar] [CrossRef]
  11. Rioult-Pedotti, M.S.; Friedman, D.; Donoghue, J.P. Learning-induced LTP in neocortex. Science 2000, 290, 533–536. [Google Scholar] [CrossRef]
  12. Stephan, K.E.; Friston, K.J. Analyzing effective connectivity with functional magnetic resonance imaging. Wiley Interdiscip. Rev. 2010, 1, 446–459. [Google Scholar] [CrossRef] [PubMed]
  13. La Rosa, C.; Parolisi, R.; Bonfanti, L. Brain structural plasticity: From adult neurogenesis to immature neurons. Front. Neurosci. 2020, 14, 512123. [Google Scholar] [CrossRef] [PubMed]
  14. Sale, A.; Berardi, N.; Maffei, L. Environment and brain plasticity: Towards an endogenous pharmacotherapy. Physiol. Rev. 2014, 94, 189–234. [Google Scholar] [CrossRef] [PubMed]
  15. Mandolesi, L.; Gelfo, F.; Serra, L.; Montuori, S.; Polverino, A.; Curcio, G.; Sorrentino, G. Environmental factors promoting neural plasticity: Insights from animal and human studies. Neural Plast. 2017, 2017, 7219461. [Google Scholar] [CrossRef]
  16. Buttelmann, F.; Karbach, J. Development and plasticity of cognitive flexibility in early and middle childhood. Front. Psychol. 2017, 8, 258078. [Google Scholar] [CrossRef] [PubMed]
  17. Stampanoni Bassi, M.; Iezzi, E.; Gilio, L.; Centonze, D.; Buttari, F. Synaptic plasticity shapes brain connectivity: Implications for network topology. Int. J. Mol. Sci. 2019, 20, 6193. [Google Scholar] [CrossRef]
  18. Chen, M.; Zheng, B. Axon plasticity in the mammalian central nervous system after injury. Trends Neurosci. 2014, 37, 583–593. [Google Scholar] [CrossRef]
  19. Marshall, K.L.; Farah, M.H. Axonal regeneration and sprouting as a potential therapeutic target for nervous system disorders. Neural Regen. Res. 2021, 16, 1901–1910. [Google Scholar]
  20. El Waly, B.; Macchi, M.; Cayre, M.; Durbec, P. Oligodendrogenesis in the normal and pathological central nervous system. Front. Neurosci. 2014, 8, 87323. [Google Scholar] [CrossRef] [PubMed]
  21. Fletcher, J.L.; Makowiecki, K.; Cullen, C.L.; Young, K.M. Oligodendrogenesis and myelination regulate cortical development, plasticity and circuit function. Semin. Cell Dev. Biol. 2021, 118, 14–23. [Google Scholar] [CrossRef]
  22. Scott, H. Brain Plasticity Influencing Phantom Limb and Prosthetics. Outstanding Honors Theses, University of South Florida, Tampa, FL, USA, 2011. [Google Scholar]
  23. Duffau, H. New Insights into Functional Mapping in Cerebral Tumor Surgery: Study of The Dynamic Interactions Between the Lesion and the Brain; Focus on Brain Mapping Research; Nova Science Pub Inc.: Hauppauge, NY, USA, 2006; p. 1. [Google Scholar]
  24. Kolb, B.; Whishaw, I.Q. Brain plasticity and behavior. Annu. Rev. Psychol. 1998, 49, 43–64. [Google Scholar] [CrossRef]
  25. Johnston, M.V.; Nishimura, A.; Harum, K.; Pekar, J.; Blue, M.E. Sculpting the developing brain. Adv. Pediatr. 2001, 48, 1–40. [Google Scholar] [CrossRef]
  26. Gould, E.; Reeves, A.J.; Graziano, M.S.; Gross, C.G. Neurogenesis in the neocortex of adult primates. Science 1999, 286, 548–552. [Google Scholar] [CrossRef]
  27. Greenough, W.T. Structural correlates of information storage in the mammalian brain: A review and hypothesis. Trends Neurosci. 1984, 7, 229–233. [Google Scholar] [CrossRef]
  28. Greenough, W.T.; Hwang, H.; Gorman, C. Evidence for active synapse formation or altered postsynaptic metabolism in visual cortex of rats reared in complex environments. Proc. Natl. Acad. Sci. USA 1985, 82, 4549–4552. [Google Scholar] [CrossRef]
  29. Turner, A.M.; Greenough, W.T. Differential rearing effects on rat visual cortex synapses. I. Synaptic and neuronal density and synapses per neuron. Brain Res. 1985, 329, 195–203. [Google Scholar] [CrossRef]
  30. Jacobs, B.; Schall, M.; Scheibel, A.B. A quantitative dendritic analysis of Wernicke’s area in humans. II. Gender, hemispheric, and environmental factors. J. Comp. Neurol. 1993, 327, 97–111. [Google Scholar] [CrossRef]
  31. Purpura, D.P. Dendritic spine” dysgenesis” and mental retardation. Science 1974, 186, 1126–1128. [Google Scholar] [CrossRef] [PubMed]
  32. Wikipedia Contributors. Neuroplasticity—Wikipedia, The Free Encyclopedia. 2024. Available online: https://en.wikipedia.org/wiki/Neuroplasticity (accessed on 22 June 2024).
  33. Nudo, R.J.; Wise, B.M.; SiFuentes, F.; Milliken, G.W. Neural substrates for the effects of rehabilitative training on motor recovery after ischemic infarct. Science 1996, 272, 1791–1794. [Google Scholar] [CrossRef]
  34. Georgopoulos, A.P. News in motor cortical physiology. Physiology 1999, 14, 64–68. [Google Scholar] [CrossRef]
  35. Sanes, J.N.; Schieber, M.H. Orderly somatotopy in primary motor cortex: Does it exist? NeuroImage 2001, 13, 968–974. [Google Scholar] [CrossRef] [PubMed]
  36. Karni, A.; Meyer, G.; Rey-Hipolito, C.; Jezzard, P.; Adams, M.M.; Turner, R.; Ungerleider, L.G. The acquisition of skilled motor performance: Fast and slow experience-driven changes in primary motor cortex. Proc. Natl. Acad. Sci. USA 1998, 95, 861–868. [Google Scholar] [CrossRef] [PubMed]
  37. Hlustik, P.; Solodkin, A.; Gullapalli, R.; Noll, D.C.; Small, S.L. Hand motor skill learning generalizes anatomically and behaviorally. NeuroImage 2000, 11, S866. [Google Scholar] [CrossRef]
  38. Pesenti, M.; Thioux, M.; Seron, X.; De Volder, A. Neuroanatomical substrates of arabic number processing, numerical comparison, and simple addition: A PET study. J. Cogn. Neurosci. 2000, 12, 461–479. [Google Scholar] [CrossRef]
  39. Ganis, G.; Keenan, J.P.; Kosslyn, S.M.; Pascual-Leone, A. Transcranial magnetic stimulation of primary motor cortex affects mental rotation. Cereb. Cortex 2000, 10, 175–180. [Google Scholar] [CrossRef] [PubMed]
  40. Laubach, M.; Wessberg, J.; Nicolelis, M.A. Cortical ensemble activity increasingly predicts behaviour outcomes during learning of a motor task. Nature 2000, 405, 567–571. [Google Scholar] [CrossRef]
  41. Salenius, S.; Hari, R. Synchronous cortical oscillatory activity during motor action. Curr. Opin. Neurobiol. 2003, 13, 678–684. [Google Scholar] [CrossRef] [PubMed]
  42. Buchel, C.; Coull, J.; Friston, K.J. The predictive value of changes in effective connectivity for human learning. Science 1999, 283, 1538–1541. [Google Scholar] [CrossRef] [PubMed]
  43. Andres, F.G.; Gerloff, C. Coherence of sequential movements and motor learning. J. Clin. Neurophysiol. 1999, 16, 520. [Google Scholar] [CrossRef] [PubMed]
  44. Duffau, H.; Capelle, L.; Sichez, N.; Denvil, D.; Lopes, M.; Sichez, J.P.; Bitar, A.; Fohanno, D. Intraoperative mapping of the subcortical language pathways using direct stimulations: An anatomo-functional study. Brain 2002, 125, 199–214. [Google Scholar] [CrossRef] [PubMed]
  45. Duffau, H.; Gatignol, P.; Mandonnet, E.; Peruzzi, P.; Tzourio-Mazoyer, N.; Capelle, L. New insights into the anatomo-functional connectivity of the semantic system: A study using cortico-subcortical electrostimulations. Brain 2005, 128, 797–810. [Google Scholar] [CrossRef]
  46. McClelland, J.L.; Rogers, T.T. The parallel distributed processing approach to semantic cognition. Nat. Rev. Neurosci. 2003, 4, 310–322. [Google Scholar] [CrossRef] [PubMed]
  47. Hatten, M.E. New directions in neuronal migration. Science 2002, 297, 1660–1663. [Google Scholar] [CrossRef]
  48. Gandolfo, F.; Li, C.S.; Benda, B.; Schioppa, C.P.; Bizzi, E. Cortical correlates of learning in monkeys adapting to a new dynamical environment. Proc. Natl. Acad. Sci. USA 2000, 97, 2259–2263. [Google Scholar] [CrossRef] [PubMed]
  49. Jercog, D.; Winke, N.; Sung, K.; Fernandez, M.M.; Francioni, C.; Rajot, D.; Courtin, J.; Chaudun, F.; Jercog, P.E.; Valerio, S.; et al. Dynamical prefrontal population coding during defensive behaviours. Nature 2021, 595, 690–694. [Google Scholar] [CrossRef]
  50. Lamprecht, R.; LeDoux, J. Structural plasticity and memory. Nat. Rev. Neurosci. 2004, 5, 45–54. [Google Scholar] [CrossRef]
  51. Byrne, J.H. Synapses Plastic plasticity. Nature 1997, 389, 791–792. [Google Scholar] [CrossRef]
  52. Buonomano, D.V.; Merzenich, M.M. Cortical plasticity: From synapses to maps. Annu. Rev. Neurosci. 1998, 21, 149–186. [Google Scholar] [CrossRef] [PubMed]
  53. Aroniadou, V.A.; Keller, A. Mechanisms of LTP induction in rat motor cortex in vitro. Cereb. Cortex 1995, 5, 353–362. [Google Scholar] [CrossRef] [PubMed]
  54. Braunewell, K.H.; Manahan-Vaughan, D. Long-term depression: A cellular basis for learning? Rev. Neurosci. 2001, 12, 121–140. [Google Scholar] [CrossRef]
  55. Schulz, A.; Miehl, C.; Berry, M.J., II; Gjorgjieva, J. The generation of cortical novelty responses through inhibitory plasticity. eLife 2021, 10, e65309. [Google Scholar] [CrossRef]
  56. Martin, S.J.; Grimwood, P.D.; Morris, R.G. Synaptic plasticity and memory: An evaluation of the hypothesis. Annu. Rev. Neurosci. 2000, 23, 649–711. [Google Scholar] [CrossRef] [PubMed]
  57. Khona, M.; Fiete, I.R. Attractor and integrator networks in the brain. Nat. Rev. Neurosci. 2022, 23, 744–766. [Google Scholar] [CrossRef]
  58. Widrow, B.; Kim, Y.; Park, D.; Perin, J.K. Nature’s learning rule: The Hebbian-LMS algorithm. In Artificial Intelligence in the Age of Neural Networks and Brain Computing; Elsevier: Amsterdam, The Netherlands, 2024; pp. 11–40. [Google Scholar]
  59. Cruikshank, S.J.; Weinberger, N.M. Evidence for the Hebbian hypothesis in experience-dependent physiological plasticity of neocortex: A critical review. Brain Res. Rev. 1996, 22, 191–228. [Google Scholar] [CrossRef]
  60. Kilgard, M.P.; Merzenich, M.M. Cortical map reorganization enabled by nucleus basalis activity. Science 1998, 279, 1714–1718. [Google Scholar] [CrossRef] [PubMed]
  61. Innocenti, G.M.; Schmidt, K.; Milleret, C.; Fabri, M.; Knyazeva, M.G.; Battaglia-Mayer, A.; Aboitiz, F.; Ptito, M.; Caleo, M.; Marzi, C.A.; et al. The functional characterization of callosal connections. Prog. Neurobiol. 2022, 208, 102186. [Google Scholar] [CrossRef] [PubMed]
  62. Blitz, D.M.; Foster, K.A.; Regehr, W.G. Short-term synaptic plasticity: A comparison of two synapses. Nat. Rev. Neurosci. 2004, 5, 630–640. [Google Scholar] [CrossRef] [PubMed]
  63. Malenka, R.C.; Nicoll, R.A. Silent synapses speak up. Neuron 1997, 19, 473–476. [Google Scholar] [CrossRef]
  64. Ridding, M.; Brouwer, B.; Miles, T.; Pitcher, J.; Thompson, P. Changes in muscle responses to stimulation of the motor cortex induced by peripheral nerve stimulation in human subjects. Exp. Brain Res. 2000, 131, 135–143. [Google Scholar] [CrossRef]
  65. Fields, R.D.; Stevens-Graham, B. New insights into neuron-glia communication. Science 2002, 298, 556–562. [Google Scholar] [CrossRef] [PubMed]
  66. Haydon, P.G. GLIA: Listening and talking to the synapse. Nat. Rev. Neurosci. 2001, 2, 185–193. [Google Scholar] [CrossRef] [PubMed]
  67. Trachtenberg, J.T.; Chen, B.E.; Knott, G.W.; Feng, G.; Sanes, J.R.; Welker, E.; Svoboda, K. Long-term in vivo imaging of experience-dependent synaptic plasticity in adult cortex. Nature 2002, 420, 788–794. [Google Scholar] [CrossRef] [PubMed]
  68. Ivanco, T.L.; Greenough, W.T. Physiological consequences of morphologically detectable synaptic plasticity: Potential uses for examining recovery following damage. Neuropharmacology 2000, 39, 765–776. [Google Scholar] [CrossRef] [PubMed]
  69. by Synaptic, H.D.I. Rapid Dendritic Morphogenesis in CA1. Cold Spring Harb. Symp. Quant. Biol. 1987, 52, 825. [Google Scholar]
  70. Poo, M.m. Neurotrophins as synaptic modulators. Nat. Rev. Neurosci. 2001, 2, 24–32. [Google Scholar] [CrossRef] [PubMed]
  71. Turrigiano, G.G.; Nelson, S.B. Homeostatic plasticity in the developing nervous system. Nat. Rev. Neurosci. 2004, 5, 97–107. [Google Scholar] [CrossRef]
  72. Selzer, M.E. Promotion of axonal regeneration in the injured CNS. Lancet Neurol. 2003, 2, 157–166. [Google Scholar] [CrossRef] [PubMed]
  73. Ullian, E.M.; Sapperstein, S.K.; Christopherson, K.S.; Barres, B.A. Control of synapse number by glia. Science 2001, 291, 657–661. [Google Scholar] [CrossRef]
  74. Shao, Y.; McCarthy, K.D. Plasticity of astrocytes. Glia 1994, 11, 147–155. [Google Scholar] [CrossRef] [PubMed]
  75. Langle, S.L.; Poulain, D.A.; Theodosis, D.T. Neuronal–glial remodeling: A structural basis for neuronal–glial interactions in the adult hypothalamus. J. Physiol. 2002, 96, 169–175. [Google Scholar] [CrossRef]
  76. Zhang, W.; Couldwell, W.T.; Simard, M.F.; Song, H.; Lin, J.H.; Nedergaard, M. Direct gap junction communication between malignant glioma cells and astrocytes. Cancer Res. 1999, 59, 1994–2003. [Google Scholar]
  77. Steindler, D.A.; Pincus, D.W. Stem cells and neuropoiesis in the adult human brain. Lancet 2002, 359, 1047–1054. [Google Scholar] [CrossRef] [PubMed]
  78. Sanai, N.; Tramontin, A.D.; Quinones-Hinojosa, A.; Barbaro, N.M.; Gupta, N.; Kunwar, S.; Lawton, M.T.; McDermott, M.W.; Parsa, A.T.; Manuel-García Verdugo, J.; et al. Unique astrocyte ribbon in adult human brain contains neural stem cells but lacks chain migration. Nature 2004, 427, 740–744. [Google Scholar] [CrossRef] [PubMed]
  79. Pincus, D.; Harrison-Restelli, C.; Barry, J.; Goodman, R.; Fraser, R.; Nedergaard, M.; Goldman, S. In vitro neurogenesis by adult human epileptic temporal neocortex. Clin. Neurosurg. 1997, 44, 17–25. [Google Scholar]
  80. Roy, N.S.; Wang, S.; Jiang, L.; Kang, J.; Benraiss, A.; Harrison-Restelli, C.; Fraser, R.A.; Couldwell, W.T.; Kawaguchi, A.; Okano, H.; et al. In vitro neurogenesis by progenitor cells isolated from the adult human hippocampus. Nat. Med. 2000, 6, 271–277. [Google Scholar] [CrossRef] [PubMed]
  81. Nunes, M.C.; Roy, N.S.; Keyoung, H.M.; Goodman, R.R.; McKhann, G.; Jiang, L.; Kang, J.; Nedergaard, M.; Goldman, S.A. Identification and isolation of multipotential neural progenitor cells from the subcortical white matter of the adult human brain. Nat. Med. 2003, 9, 439–447. [Google Scholar] [CrossRef] [PubMed]
  82. Gross, C.G. Neurogenesis in the adult brain: Death of a dogma. Nat. Rev. Neurosci. 2000, 1, 67–73. [Google Scholar] [CrossRef]
  83. Magavi, S.S.; Macklis, J.D. Induction of neuronal type-specific neurogenesis in the cerebral cortex of adult mice: Manipulation of neural precursors in situ. Dev. Brain Res. 2002, 134, 57–76. [Google Scholar] [CrossRef] [PubMed]
  84. Nguyen, D.K.; Botez, M. Diaschisis and neurobehavior. Can. J. Neurol. Sci. 1998, 25, 5–12. [Google Scholar] [CrossRef] [PubMed]
  85. Seitz, R.J.; Azari, N.P.; Knorr, U.; Binkofski, F.; Herzog, H.; Freund, H.J. The role of diaschisis in stroke recovery. Stroke 1999, 30, 1844–1850. [Google Scholar] [CrossRef]
  86. Duffau, H.; Sichez, J.P.; Lehéricy, S. Intraoperative unmasking of brain redundant motor sites during resection of a precentral angioma: Evidence using direct cortical stimulation. Ann. Neurol. 2000, 47, 132–135. [Google Scholar] [CrossRef]
  87. Duffau, H. Acute functional reorganisation of the human motor cortex during resection of central lesions: A study using intraoperative brain mapping. J. Neurol. Neurosurg. Psychiatry 2001, 70, 506–513. [Google Scholar] [CrossRef] [PubMed]
  88. Rijntjes, M.; Weiller, C. Recovery of motor and language abilities after stroke: The contribution of functional imaging. Prog. Neurobiol. 2002, 66, 109–122. [Google Scholar] [CrossRef]
  89. Krainik, A.; Duffau, H.; Capelle, L.; Cornu, P.; Boch, A.L.; Mangin, J.F.; Le Bihan, D.; Marsault, C.; Chiras, J.; Lehéricy, S. Role of the healthy hemisphere in recovery after resection of the supplementary motor area. Neurology 2004, 62, 1323–1332. [Google Scholar] [CrossRef] [PubMed]
  90. Levänen, S.; Jousmäki, V.; Hari, R. Vibration-induced auditory-cortex activation in a congenitally deaf adult. Curr. Biol. 1998, 8, 869–872. [Google Scholar] [CrossRef]
  91. Lee, D.S.; Lee, J.S.; Oh, S.H.; Kim, S.K.; Kim, J.W.; Chung, J.K.; Lee, M.C.; Kim, C.S. Cross-modal plasticity and cochlear implants. Nature 2001, 409, 149–150. [Google Scholar] [CrossRef]
  92. Rossini, P.M.; Dal Forno, G. Integrated technology for evaluation of brain function and neural plasticity. Phys. Med. Rehabil. Clin. North Am. 2004, 15, 263–306. [Google Scholar] [CrossRef]
  93. Luders, E.; Gaser, C.; Jancke, L.; Schlaug, G. A voxel-based approach to gray matter asymmetries. Neuroimage 2004, 22, 656–664. [Google Scholar] [CrossRef] [PubMed]
  94. Josse, G.; Mazoyer, B.; Crivello, F.; Tzourio-Mazoyer, N. Left planum temporale: An anatomical marker of left hemispheric specialization for language comprehension. Cogn. Brain Res. 2003, 18, 1–14. [Google Scholar] [CrossRef]
  95. Sluming, V.; Barrick, T.; Howard, M.; Cezayirli, E.; Mayes, A.; Roberts, N. Voxel-based morphometry reveals increased gray matter density in Broca’s area in male symphony orchestra musicians. Neuroimage 2002, 17, 1613–1622. [Google Scholar] [CrossRef]
  96. Hutchinson, S.; Lee, L.H.L.; Gaab, N.; Schlaug, G. Cerebellar volume of musicians. Cereb. Cortex 2003, 13, 943–949. [Google Scholar] [CrossRef]
  97. Mackay, C.E.; Roberts, N.; Mayes, A.R.; Downes, J.J.; Foster, J.K.; Mann, D. An exploratory study of the relationship between face recognition memory and the volume of medial temporal lobe structures in healthy young males. Behav. Neurol. 1998, 11, 3–20. [Google Scholar] [CrossRef]
  98. Maguire, E.A.; Gadian, D.G.; Johnsrude, I.S.; Good, C.D.; Ashburner, J.; Frackowiak, R.S.; Frith, C.D. Navigation-related structural change in the hippocampi of taxi drivers. Proc. Natl. Acad. Sci. USA 2000, 97, 4398–4403. [Google Scholar] [CrossRef]
  99. Paus, T.; Zijdenbos, A.; Worsley, K.; Collins, D.L.; Blumenthal, J.; Giedd, J.N.; Rapoport, J.L.; Evans, A.C. Structural maturation of neural pathways in children and adolescents: In vivo study. Science 1999, 283, 1908–1911. [Google Scholar] [CrossRef]
  100. Draganski, B.; Gaser, C.; Busch, V.; Schuierer, G.; Bogdahn, U.; May, A. Changes in grey matter induced by training. Nature 2004, 427, 311–312. [Google Scholar] [CrossRef]
  101. Juraska, J.M. Sex differences in dendritic response to differential experience in the rat visual cortex. Brain Res. 1984, 295, 27–34. [Google Scholar] [CrossRef] [PubMed]
  102. Juraska, J.M. Sex differences in developmental plasticity of behavior and the brain. In Developmental Neuropsychobiology; Academic Press: Cambridge, MA, USA, 1986; pp. 409–422. [Google Scholar]
  103. Juraska, J.M.; Fitch, J.M.; Henderson, C.; Rivers, N. Sex differences in the dendritic branching of dentate granule cells following differential experience. Brain Res. 1985, 333, 73–80. [Google Scholar] [CrossRef]
  104. Juraska, J.M. The structure of the rat cerebral cortex: Effects of gender and the environment. In The Cerebral Cortex of the Rat; Kolb, B., Tees, R.C., Eds.; The MIT Press: Cambridge, MA, USA, 1990. [Google Scholar]
  105. Coleman, P.; Buell, S. Regulation of dendritic extent in developing and aging brain. In Synaptic Plasticity; The Guilford Press: New York, NY, USA, 1985; pp. 311–333. [Google Scholar]
  106. Duque, A.; Arellano, J.I.; Rakic, P. An assessment of the existence of adult neurogenesis in humans and value of its rodent models for neuropsychiatric diseases. Mol. Psychiatry 2022, 27, 377–382. [Google Scholar] [CrossRef] [PubMed]
  107. Raff, M.C.; Barres, B.A.; Burne, J.F.; Coles, H.S.; Ishizaki, Y.; Jacobson, M.D. Programmed cell death and the control of cell survival: Lessons from the nervous system. Science 1993, 262, 695–700. [Google Scholar] [CrossRef]
  108. Johansson, B.B. Brain plasticity in health and disease. Keio J. Med. 2004, 53, 231–246. [Google Scholar] [CrossRef]
  109. Statsenko, Y.; Habuza, T.; Charykova, I.; Gorkom, K.N.V.; Zaki, N.; Almansoori, T.M.; Baylis, G.; Ljubisavljevic, M.; Belghali, M. Predicting age from behavioral test performance for screening early onset of cognitive decline. Front. Aging Neurosci. 2021, 13, 661514. [Google Scholar] [CrossRef] [PubMed]
  110. Statsenko, Y.; Habuza, T.; Gorkom, K.N.V.; Zaki, N.; Almansoori, T.M.; Al Zahmi, F.; Ljubisavljevic, M.R.; Belghali, M. Proportional changes in cognitive subdomains during normal brain aging. Front. Aging Neurosci. 2021, 13, 673469. [Google Scholar] [CrossRef]
  111. Statsenko, Y.; Habuza, T.; Smetanina, D.; Simiyu, G.L.; Uzianbaeva, L.; Neidl-Van Gorkom, K.; Zaki, N.; Charykova, I.; Al Koteesh, J.; Almansoori, T.M.; et al. Brain morphometry and cognitive performance in normal brain aging: Age-and sex-related structural and functional changes. Front. Aging Neurosci. 2022, 13, 713680. [Google Scholar] [CrossRef]
  112. Belghali, M.; Statsenko, Y.; Laver, V. Stroop switching card test: Brief screening of executive functions across the lifespan. Aging Neuropsychol. Cogn. 2022, 29, 14–33. [Google Scholar] [CrossRef]
  113. Statsenko, Y.; Habuza, T.; Uzianbaeva, L.; Gorkom, K.; Belghali, M.; Charykova, I. Correlation between lifelong dynamics of psychophysiological performance and brain morphology. ESNR 2021. Neuroradiology 2021, 63, 41–42. [Google Scholar]
  114. Habuza, T.; Statsenko, Y.; Uzianbaeva, L.; Gorkom, K.; Zaki, N.; Belghali, M. Models of brain cognitive and morphological changes across the life: Machine learning-based approach. ESNR 2021. Neuroradiology 2021, 63, 10–26226. [Google Scholar]
  115. Uzianbaeva, L.; Statsenko, Y.; Habuza, T.; Gorkom, K.; Belghali, M.; Charykova, I. Effects of sex age-related changes in brain morphology. ESNR 2021. Neuroradiology 2021, 63, 42–43. [Google Scholar]
  116. Gorkom, K.; Statsenko, Y.; Habuza, T.; Uzianbaeva, L.; Belghali, M.; Charykova, I. Comparison of brain volumetric changes with functional outcomes in physiologic brain aging. ESNR 2021. Neuroradiology 2021, 63, 43–44. [Google Scholar]
  117. Statsenko, Y.; Habuza, T.; Charykova, I.; Gorkom, K.; Zaki, N.; Almansoori, T.; Ljubisavljevic, M.; Szolics, M.; Al Koteesh, J.; Ponomareva, A.; et al. AI models of age-associated changes in CNS composition identified by MRI. J. Neurol. Sci. 2021, 429, 118303. [Google Scholar] [CrossRef]
  118. Habuza, T.; Zaki, N.; Statsenko, Y.; Elyassami, S. MRI and cognitive tests-based screening tool for dementia. J. Neurol. Sci. 2021, 429, 118964. [Google Scholar] [CrossRef]
  119. Statsenko, Y.; Habuza, T.; Gorkom, K.N.V.; Zaki, N.; Almansoori, T.M. Applying the inverse efficiency score to visual–motor task for studying speed-accuracy performance while aging. Front. Aging Neurosci. 2020, 12, 574401. [Google Scholar] [CrossRef]
  120. Habuza, T.; Zaki, N.; Statsenko, Y.; Alnajjar, F.; Elyassami, S. Predicting the diagnosis of dementia from MRI data: Added value to cognitive tests. In Proceedings of the 7th Annual International Conference on Arab Women in Computing in Conjunction with the 2nd Forum of Women in Research, Sharjah, United Arab Emirates, 25–26 August 2021; pp. 1–7. [Google Scholar]
  121. Habuza, T.; Zaki, N.; Mohamed, E.A.; Statsenko, Y. Deviation from model of normal aging in Alzheimer’s disease: Application of deep learning to structural MRI data and cognitive tests. IEEE Access 2022, 10, 53234–53249. [Google Scholar] [CrossRef]
  122. Statsenko, Y.; Meribout, S.; Habuza, T.; Almansoori, T.M.; Gorkom, K.N.V.; Gelovani, J.G.; Ljubisavljevic, M. Patterns of structure- function association in normal aging and in Alzheimer’s disease: Screening for mild cognitive impairment and dementia with ML regression and classification models. Front. Aging Neurosci. 2023, 14, 943566. [Google Scholar] [CrossRef] [PubMed]
  123. Smetanina, D.; Meribout, S.; Habuza, T.; Simiyu, G.; Ismail, F.; Zareba, K.; Gorkom, K.; Almansoori, T.; Gelovani, J.; Ljubisavljevic, M.; et al. Reference curves of age-related volumetric changes in hippocampus and brain ventricles in healthy population. J. Neurol. Sci. 2023, 455, 121995. [Google Scholar] [CrossRef]
  124. Meribout, S.; Habuza, T.; Smetanina, D.; Simiyu, G.; Ismail, F.; Gorkom, K.; Almansoori, T.; Gelovani, J.; Statsenko, Y.; Ljubisavljevic, M. Rate and onset of cognitive decline and cortical atrophy in normal and pathological ageing. J. Neurol. Sci. 2023, 455, 121426. [Google Scholar] [CrossRef]
  125. Meribout, S.; Habuza, T.; Smetanina, D.; Simiyu, G.; Ismail, F.; Gorkom, K.; Almansoori, T.; Gelovani, J.; Statsenko, Y.; Ljubisavljevic, M. Functional changes in age-related neurocognitive slowing and disease-related cognitive decline: Evidence from global cognitive tests and psychophysiological tests. J. Neurol. Sci. 2023, 455, 121424. [Google Scholar] [CrossRef]
  126. Statsenko, Y.; Meribout, S.; Habuza, T.; Smetanina, D.; Simiyu, G.; Ismail, F.; Gorkom, K.; Almansoori, T.; Gelovani, J.; Ljubisavljevic, M. Structure-function association patterns of the brain in individuals with different level of cognitive impairment. J. Neurol. Sci. 2023, 455, 121462. [Google Scholar] [CrossRef]
  127. Spadaro, P.A.; Bredy, T.W. Emerging role of non-coding RNA in neural plasticity, cognitive function, and neuropsychiatric disorders. Front. Genet. 2012, 3, 132. [Google Scholar] [CrossRef]
  128. Kyzar, E.J.; Bohnsack, J.P.; Pandey, S.C. Current and future perspectives of noncoding RNAs in brain function and neuropsychiatric disease. Biol. Psychiatry 2022, 91, 183–193. [Google Scholar] [CrossRef]
  129. Earls, L.R.; Westmoreland, J.J.; Zakharenko, S.S. Non-coding RNA regulation of synaptic plasticity and memory: Implications for aging. Ageing Res. Rev. 2014, 17, 34–42. [Google Scholar] [CrossRef]
  130. Keihani, S.; Kluever, V.; Fornasiero, E.F. Brain long noncoding RNAs: Multitask regulators of neuronal differentiation and function. Molecules 2021, 26, 3951. [Google Scholar] [CrossRef]
  131. Ip, J.Y.; Sone, M.; Nashiki, C.; Pan, Q.; Kitaichi, K.; Yanaka, K.; Abe, T.; Takao, K.; Miyakawa, T.; Blencowe, B.J.; et al. Gomafu lncRNA knockout mice exhibit mild hyperactivity with enhanced responsiveness to the psychostimulant methamphetamine. Sci. Rep. 2016, 6, 27204. [Google Scholar] [CrossRef]
  132. Zhao, H.; Jin, T.; Cheng, X.; Qin, J.; Zhang, L.; He, H.; Xue, J.; Jin, G. GAS5 which is regulated by Lhx8 promotes the recovery of learning and memory in rats with cholinergic nerve injury. Life Sci. 2020, 260, 118388. [Google Scholar] [CrossRef]
  133. Li, L.; Zhuang, Y.; Zhao, X.; Li, X. Long non-coding RNA in neuronal development and neurological disorders. Front. Genet. 2019, 9, 436598. [Google Scholar] [CrossRef]
  134. Bhan, A.; Soleimani, M.; Mandal, S.S. Long noncoding RNA (lncRNA): Functions in health and disease. In Gene Regulation, Epigenetics and Hormone Signaling; Wiley-Blackwell: Hoboken, NJ, USA, 2017; pp. 169–208. [Google Scholar]
  135. Khalifa, F.N.; Hussein, R.F.; Mekawy, D.M.; Elwi, H.M.; Alsaeed, S.A.; Elnawawy, Y.; Shaheen, S.H. Potential role of the lncRNA “HOTAIR”/miRNA “206”/BDNF network in the alteration in expression of synaptic plasticity gene arc and BDNF level in sera of patients with heroin use disorder through the PI3K/AKT/mTOR pathway compared to the controls. Mol. Biol. Rep. 2024, 51, 293. [Google Scholar] [CrossRef]
  136. Sim, S.E.; Lim, C.S.; Kim, J.I.; Seo, D.; Chun, H.; Yu, N.K.; Lee, J.; Kang, S.J.; Ko, H.G.; Choi, J.H.; et al. The brain-enriched microRNA miR-9-3p regulates synaptic plasticity and memory. J. Neurosci. 2016, 36, 8641–8652. [Google Scholar] [CrossRef]
  137. McNeill, E.M.; Warinner, C.; Alkins, S.; Taylor, A.; Heggeness, H.; DeLuca, T.F.; Fulga, T.A.; Wall, D.P.; Griffith, L.C.; Van Vactor, D. The conserved microRNA miR-34 regulates synaptogenesis via coordination of distinct mechanisms in presynaptic and postsynaptic cells. Nat. Commun. 2020, 11, 1092. [Google Scholar] [CrossRef]
  138. Qian, Y.; Song, J.; Ouyang, Y.; Han, Q.; Chen, W.; Zhao, X.; Xie, Y.; Chen, Y.; Yuan, W.; Fan, C. Advances in roles of miR-132 in the nervous system. Front. Pharmacol. 2017, 8, 770. [Google Scholar] [CrossRef]
  139. Xin, H.; Katakowski, M.; Wang, F.; Qian, J.Y.; Liu, X.S.; Ali, M.M.; Buller, B.; Zhang, Z.G.; Chopp, M. MicroRNA-17–92 cluster in exosomes enhance neuroplasticity and functional recovery after stroke in rats. Stroke 2017, 48, 747–753. [Google Scholar] [CrossRef]
  140. Jin, J.; Kim, S.N.; Liu, X.; Zhang, H.; Zhang, C.; Seo, J.S.; Kim, Y.; Sun, T. miR-17-92 cluster regulates adult hippocampal neurogenesis, anxiety, and depression. Cell Rep. 2016, 16, 1653–1663. [Google Scholar] [CrossRef]
  141. Wu, X.; Zhang, Y.; Wang, P.; Li, X.; Song, Z.; Wei, C.; Zhang, Q.; Luo, B.; Liu, Z.; Yang, Y.; et al. Clinical and preclinical evaluation of miR-144-5p as a key target for major depressive disorder. CNS Neurosci. Ther. 2023, 29, 3598–3611. [Google Scholar] [CrossRef] [PubMed]
  142. Gao, C.; Yin, F.; Li, R.; Ruan, Q.; Meng, C.; Zhao, K.; Zhu, Q. MicroRNA-145-Mediated KDM6A Downregulation Enhances Neural Repair after Spinal Cord Injury via the NOTCH2/Abcb1a Axis. Oxidative Med. Cell. Longev. 2021, 2021, 2580619. [Google Scholar] [CrossRef] [PubMed]
  143. Qiao, J.; Zhao, J.; Chang, S.; Sun, Q.; Liu, N.; Dong, J.; Chen, Y.; Yang, D.; Ye, D.; Liu, X.; et al. MicroRNA-153 improves the neurogenesis of neural stem cells and enhances the cognitive ability of aged mice through the notch signaling pathway. Cell Death Differ. 2020, 27, 808–825. [Google Scholar] [CrossRef] [PubMed]
  144. Del Val, C.; Díaz de la Guardia-Bolívar, E.; Zwir, I.; Mishra, P.P.; Mesa, A.; Salas, R.; Poblete, G.F.; de Erausquin, G.; Raitoharju, E.; Kähönen, M.; et al. Gene expression networks regulated by human personality. Mol. Psychiatry 2024, 29, 2241–2260. [Google Scholar] [CrossRef]
  145. Seeler, S.; Andersen, M.S.; Sztanka-Toth, T.; Rybiczka-Tešulov, M.; van den Munkhof, M.H.; Chang, C.C.; Maimaitili, M.; Venø, M.T.; Hansen, T.B.; Pasterkamp, R.J.; et al. A circular RNA expressed from the FAT3 locus regulates neural development. Mol. Neurobiol. 2023, 60, 3239–3260. [Google Scholar] [CrossRef]
  146. Du, M.; Wu, C.; Yu, R.; Cheng, Y.; Tang, Z.; Wu, B.; Fu, J.; Tan, W.; Zhou, Q.; Zhu, Z.; et al. A novel circular RNA, circIgfbp2, links neural plasticity and anxiety through targeting mitochondrial dysfunction and oxidative stress-induced synapse dysfunction after traumatic brain injury. Mol. Psychiatry 2022, 27, 4575–4589. [Google Scholar] [CrossRef]
  147. Bao, N.; Liu, J.; Peng, Z.; Zhang, R.; Ni, R.; Li, R.; Wu, J.; Liu, Z.; Pan, B. Identification of circRNA-miRNA-mRNA networks to explore the molecular mechanism and immune regulation of postoperative neurocognitive disorder. Aging 2022, 14, 8374. [Google Scholar] [CrossRef] [PubMed]
  148. Mahmoudi, E.; Fitzsimmons, C.; Geaghan, M.P.; Shannon Weickert, C.; Atkins, J.R.; Wang, X.; Cairns, M.J. Circular RNA biogenesis is decreased in postmortem cortical gray matter in schizophrenia and may alter the bioavailability of associated miRNA. Neuropsychopharmacology 2019, 44, 1043–1054. [Google Scholar] [CrossRef]
  149. Brookes, E.; Alan Au, H.Y.; Varsally, W.; Barrington, C.; Hadjur, S.; Riccio, A. A novel enhancer that regulates Bdnf expression in developing neurons. bioRxiv 2021. [Google Scholar] [CrossRef]
  150. Cajigas, I.; Chakraborty, A.; Swyter, K.R.; Luo, H.; Bastidas, M.; Nigro, M.; Morris, E.R.; Chen, S.; VanGompel, M.J.; Leib, D.; et al. The Evf2 ultraconserved enhancer lncRNA functionally and spatially organizes megabase distant genes in the developing forebrain. Mol. Cell 2018, 71, 956–972. [Google Scholar] [CrossRef]
  151. Sauvageau, M.; Goff, L.A.; Lodato, S.; Bonev, B.; Groff, A.F.; Gerhardinger, C.; Sanchez-Gomez, D.B.; Hacisuleyman, E.; Li, E.; Spence, M.; et al. Multiple knockout mouse models reveal lincRNAs are required for life and brain development. eLife 2013, 2, e01749. [Google Scholar] [CrossRef]
  152. Chanda, K.; Mukhopadhyay, D. LncRNA Xist, X-chromosome instability and Alzheimer’s disease. Curr. Alzheimer Res. 2020, 17, 499–507. [Google Scholar] [CrossRef] [PubMed]
  153. Rajasethupathy, P.; Antonov, I.; Sheridan, R.; Frey, S.; Sander, C.; Tuschl, T.; Kandel, E.R. A role for neuronal piRNAs in the epigenetic control of memory-related synaptic plasticity. Cell 2012, 149, 693–707. [Google Scholar] [CrossRef]
  154. Zuo, L.; Wang, Z.; Tan, Y.; Chen, X.; Luo, X. piRNAs and their functions in the brain. Int. J. Hum. Genet. 2016, 16, 53–60. [Google Scholar] [CrossRef]
  155. Qiu, W.; Guo, X.; Lin, X.; Yang, Q.; Zhang, W.; Zhang, Y.; Zuo, L.; Zhu, Y.; Li, C.S.R.; Ma, C.; et al. Transcriptome-wide piRNA profiling in human brains of Alzheimer’s disease. Neurobiol. Aging 2017, 57, 170–177. [Google Scholar] [CrossRef] [PubMed]
  156. Scheckel, C.; Drapeau, E.; Frias, M.A.; Park, C.Y.; Fak, J.; Zucker-Scharff, I.; Kou, Y.; Haroutunian, V.; Ma’ayan, A.; Buxbaum, J.D.; et al. Regulatory consequences of neuronal ELAV-like protein binding to coding and non-coding RNAs in human brain. eLife 2016, 5, e10421. [Google Scholar] [CrossRef]
  157. Wei, Z.; Batagov, A.O.; Schinelli, S.; Wang, J.; Wang, Y.; El Fatimy, R.; Rabinovsky, R.; Balaj, L.; Chen, C.C.; Hochberg, F.; et al. Coding and noncoding landscape of extracellular RNA released by human glioma stem cells. Nat. Commun. 2017, 8, 1145. [Google Scholar] [CrossRef] [PubMed]
  158. Haack, F.; Trakooljul, N.; Gley, K.; Murani, E.; Hadlich, F.; Wimmers, K.; Ponsuksili, S. Deep sequencing of small non-coding RNA highlights brain-specific expression patterns and RNA cleavage. RNA Biol. 2019, 16, 1764–1774. [Google Scholar] [CrossRef]
  159. Zimmer-Bensch, G. Emerging roles of long non-coding RNAs as drivers of brain evolution. Cells 2019, 8, 1399. [Google Scholar] [CrossRef]
  160. Bernard, D.; Prasanth, K.V.; Tripathi, V.; Colasse, S.; Nakamura, T.; Xuan, Z.; Zhang, M.Q.; Sedel, F.; Jourdren, L.; Coulpier, F.; et al. A long nuclear-retained non-coding RNA regulates synaptogenesis by modulating gene expression. EMBO J. 2010, 29, 3082–3093. [Google Scholar] [CrossRef]
  161. Tripathi, V.; Ellis, J.D.; Shen, Z.; Song, D.Y.; Pan, Q.; Watt, A.T.; Freier, S.M.; Bennett, C.F.; Sharma, A.; Bubulya, P.A.; et al. The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol. Cell 2010, 39, 925–938. [Google Scholar] [CrossRef]
  162. Liau, W.S.; Zhao, Q.; Bademosi, A.; Gormal, R.S.; Gong, H.; Marshall, P.R.; Periyakaruppiah, A.; Madugalle, S.U.; Zajaczkowski, E.L.; Leighton, L.J.; et al. Fear extinction is regulated by the activity of long noncoding RNAs at the synapse. Nat. Commun. 2023, 14, 7616. [Google Scholar] [CrossRef]
  163. Lu, T.X.; Rothenberg, M.E. MicroRNA. J. Allergy Clin. Immunol. 2018, 141, 1202–1207. [Google Scholar] [CrossRef]
  164. Starega-Roslan, J.; Krol, J.; Koscianska, E.; Kozlowski, P.; Szlachcic, W.J.; Sobczak, K.; Krzyzosiak, W.J. Structural basis of microRNA length variety. Nucleic Acids Res. 2011, 39, 257–268. [Google Scholar] [CrossRef]
  165. Fiore, R.; Siegel, G.; Schratt, G. MicroRNA function in neuronal development, plasticity and disease. Biochim. Biophys. Acta (BBA)-Gene Regul. Mech. 2008, 1779, 471–478. [Google Scholar] [CrossRef]
  166. Martins, H.C.; Schratt, G. MicroRNA-dependent control of neuroplasticity in affective disorders. Transl. Psychiatry 2021, 11, 263. [Google Scholar] [CrossRef] [PubMed]
  167. Constantin, L. Circular RNAs and neuronal development. In Circular RNAs: Biogenesis and Functions; Springer Nature: Singapore, 2018; pp. 205–213. [Google Scholar]
  168. Kim, T.K.; Hemberg, M.; Gray, J.M.; Costa, A.M.; Bear, D.M.; Wu, J.; Harmin, D.A.; Laptewicz, M.; Barbara-Haley, K.; Kuersten, S.; et al. Widespread transcription at neuronal activity-regulated enhancers. Nature 2010, 465, 182–187. [Google Scholar] [CrossRef]
  169. Arner, E.; Daub, C.O.; Vitting-Seerup, K.; Andersson, R.; Lilje, B.; Drabløs, F.; Lennartsson, A.; Rönnerblad, M.; Hrydziuszko, O.; Vitezic, M.; et al. Transcribed enhancers lead waves of coordinated transcription in transitioning mammalian cells. Science 2015, 347, 1010–1014. [Google Scholar] [CrossRef] [PubMed]
  170. Gray, J.M.; Kim, T.K.; West, A.E.; Nord, A.S.; Markenscoff-Papadimitriou, E.; Lomvardas, S. Genomic views of transcriptional enhancers: Essential determinants of cellular identity and activity-dependent responses in the CNS. J. Neurosci. 2015, 35, 13819–13826. [Google Scholar] [CrossRef] [PubMed]
  171. Roberts, T.C.; Morris, K.V.; Wood, M.J. The role of long non-coding RNAs in neurodevelopment, brain function and neurological disease. Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 20130507. [Google Scholar] [CrossRef] [PubMed]
  172. Gu, S.; Xie, R.; Liu, X.; Shou, J.; Gu, W.; Che, X. Long coding RNA XIST contributes to neuronal apoptosis through the downregulation of AKT phosphorylation and is negatively regulated by miR-494 in rat spinal cord injury. Int. J. Mol. Sci. 2017, 18, 732. [Google Scholar] [CrossRef] [PubMed]
  173. Aravin, A.; Gaidatzis, D.; Pfeffer, S.; Lagos-Quintana, M.; Landgraf, P.; Iovino, N.; Morris, P.; Brownstein, M.J.; Kuramochi-Miyagawa, S.; Nakano, T.; et al. A novel class of small RNAs bind to MILI protein in mouse testes. Nature 2006, 442, 203–207. [Google Scholar] [CrossRef] [PubMed]
  174. Girard, A.; Sachidanandam, R.; Hannon, G.J.; Carmell, M.A. A germline-specific class of small RNAs binds mammalian Piwi proteins. Nature 2006, 442, 199–202. [Google Scholar] [CrossRef] [PubMed]
  175. Betel, D.; Sheridan, R.; Marks, D.S.; Sander, C. Computational analysis of mouse piRNA sequence and biogenesis. PLoS Comput. Biol. 2007, 3, e222. [Google Scholar] [CrossRef]
  176. Paillard, J. Réflexions sur l’usage du concept de plasticité en neurobiology. J. Psychol. 1976, 1, 33–47. [Google Scholar]
  177. Alexandre, F. A global framework for a systemic view of brain modeling. Brain Inform. 2021, 8, 3. [Google Scholar] [CrossRef] [PubMed]
  178. Hille, M.; Kühn, S.; Kempermann, G.; Bonhoeffer, T.; Lindenberger, U. From animal models to human individuality: Integrative approaches to the study of brain plasticity. Neuron 2024, 112, 3522–3541. [Google Scholar] [CrossRef]
  179. De Yoe, E.A.; Bandettini, P.; Neitz, J.; Miller, D.; Winans, P. Functional magnetic resonance imaging (FMRI) of the human brain. J. Neurosci. Methods 1994, 54, 171–187. [Google Scholar] [CrossRef] [PubMed]
  180. Bandettini, P.A. Twenty years of functional MRI: The science and the stories. Neuroimage 2012, 62, 575–588. [Google Scholar] [CrossRef] [PubMed]
  181. Poldrack, R.A.; Mumford, J.A.; Nichols, T.E. Handbook of Functional MRI Data Analysis; Cambridge University Press: Cambridge, UK, 2024. [Google Scholar]
  182. Mechelli, A.; Price, C.J.; Friston, K.J.; Ashburner, J. Voxel-based morphometry of the human brain: Methods and applications. Curr. Med. Imaging 2005, 1, 105–113. [Google Scholar] [CrossRef]
  183. Scarpazza, C.; Stefania De Simone, M. Voxel-based morphometry: Current perspectives. Neurosci. Neuroecon. 2016, 5, 19–35. [Google Scholar] [CrossRef]
  184. Hedderich, D.M.; Opfer, R.; Krüger, J.; Spies, L.; Yakushev, I.; Buchert, R. Clinical validation of artificial intelligence-based single-subject morphometry without normative reference database. J. Alzheimer’s Dis. 2025, 103, 13872877241304607. [Google Scholar] [CrossRef]
  185. Tournier, J.D.; Mori, S.; Leemans, A. Diffusion tensor imaging and beyond. Magn. Reson. Med. 2011, 65, 1532. [Google Scholar] [CrossRef]
  186. Van Hecke, W.; Emsell, L.; Sunaert, S. (Eds.) Diffusion Tensor Imaging: A Practical Handbook; Springer: New York, NY, USA, 2016. [Google Scholar]
  187. Dell’Acqua, F.; Leemans, A.; Dawson, M.; Descoteaux, M. Single-shell diffusion models: From DTI to HARDI. In Handbook of Diffusion MR Tractography; Academic Press: Cambridge, MA, USA, 2025; pp. 177–200. [Google Scholar]
  188. Craik, A.; He, Y.; Contreras-Vidal, J.L. Deep learning for electroencephalogram (EEG) classification tasks: A review. J. Neural Eng. 2019, 16, 031001. [Google Scholar] [CrossRef] [PubMed]
  189. Michel, C.M.; Brunet, D. EEG source imaging: A practical review of the analysis steps. Front. Neurol. 2019, 10, 325. [Google Scholar] [CrossRef]
  190. Müller-Putz, G.R. Electroencephalography. Handbook of clinical neurology 2020, 168, 249–262. [Google Scholar] [PubMed]
  191. Baillet, S. Magnetoencephalography for brain electrophysiology and imaging. Nat. Neurosci. 2017, 20, 327–339. [Google Scholar] [CrossRef] [PubMed]
  192. Brickwedde, M.; Anders, P.; Kühn, A.A.; Lofredi, R.; Holtkamp, M.; Kaindl, A.M.; Grent-‘t-Jong, T.; Krüger, P.; Sander, T.; Uhlhaas, P.J. Applications of OPM-MEG for translational neuroscience: A perspective. Transl. Psychiatry 2024, 14, 341. [Google Scholar] [CrossRef]
  193. Rhodes, N.; Sato, J.; Safar, K.; Amorim, K.; Taylor, M.J.; Brookes, M.J. Paediatric magnetoencephalography and its role in neurodevelopmental disorders. Br. J. Radiol. 2024, 97, 1591–1601. [Google Scholar] [CrossRef]
  194. Boppart, S.A. Optical coherence tomography: Technology and applications for neuroimaging. Psychophysiology 2003, 40, 529–541. [Google Scholar] [CrossRef]
  195. Subhash, H.M. Full-Field and Single-Shot Full-Field Optical Coherence Tomography: A Novel Technique for Biomedical Imaging Applications. Adv. Opt. Technol. 2012, 2012, 435408. [Google Scholar] [CrossRef]
  196. Yang, L.; Chen, P.; Wen, X.; Zhao, Q. Optical coherence tomography (OCT) and OCT angiography: Technological development and applications in brain science. Theranostics 2025, 15, 122. [Google Scholar] [CrossRef] [PubMed]
  197. Poldrack, R.A. Imaging brain plasticity: Conceptual and methodological issues—A theoretical review. Neuroimage 2000, 12, 1–3. [Google Scholar] [CrossRef] [PubMed]
  198. Tai, Y.F.; Piccini, P. Applications of positron emission tomography (PET) in neurology. J. Neurol. Neurosurg. Psychiatry 2004, 75, 669–676. [Google Scholar] [CrossRef] [PubMed]
  199. Veldsman, M.; Egorova, N. Advances in neuroimaging for neurodegenerative disease. In Neurodegenerative Diseases: Pathology, Mechanisms, and Potential Therapeutic Targets; Springer: Cham, Switzerland, 2017; pp. 451–478. [Google Scholar]
  200. Hricak, H.; Mayerhoefer, M.E.; Herrmann, K.; Lewis, J.S.; Pomper, M.G.; Hess, C.P.; Riklund, K.; Scott, A.M.; Weissleder, R. Advances and challenges in precision imaging. Lancet Oncol. 2025, 26, e34–e45. [Google Scholar] [CrossRef] [PubMed]
  201. Lapicque, L. Recherches quantitatives sur l’excitation electrique des nerfs. J. Physiol 1907, 9, 620–635. [Google Scholar]
  202. Abbott, L.F. Lapicque’s introduction of the integrate-and-fire model neuron (1907). Brain Res. Bull. 1999, 50, 303–304. [Google Scholar] [CrossRef] [PubMed]
  203. Hodgkin, A.; Huxley, A. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 1952, 117, 500–544. [Google Scholar] [CrossRef] [PubMed]
  204. Neher, E.; Sakmann, B. Single-channel currents recorded from membrane of denervated frog muscle fibres. Nature 1976, 260, 799–802. [Google Scholar] [CrossRef] [PubMed]
  205. Ibarz, B.; Casado, J.M.; Sanjuán, M.A. Map-based models in neuronal dynamics. Phys. Rep. 2011, 501, 1–74. [Google Scholar] [CrossRef]
  206. Buesing, L.; Bill, J.; Nessler, B.; Maass, W. Neural dynamics as sampling: A model for stochastic computation in recurrent networks of spiking neurons. PLoS Comput. Biol. 2011, 7, e1002211. [Google Scholar] [CrossRef] [PubMed]
  207. Cuppini, C.; Magosso, E.; Ursino, M. Organization, maturation, and plasticity of multisensory integration: Insights from computational modeling studies. Front. Psychol. 2011, 2, 77. [Google Scholar] [CrossRef] [PubMed]
  208. Shine, J.M.; Müller, E.J.; Munn, B.; Cabral, J.; Moran, R.J.; Breakspear, M. Computational models link cellular mechanisms of neuromodulation to large-scale neural dynamics. Nat. Neurosci. 2021, 24, 765–776. [Google Scholar] [CrossRef] [PubMed]
  209. Ramaswamy, S. Data-driven multiscale computational models of cortical and subcortical regions. Curr. Opin. Neurobiol. 2024, 85, 102842. [Google Scholar] [CrossRef]
  210. Po, H.F.; Houben, A.M.; Haeb, A.C.; Jenkins, D.R.; Hill, E.J.; Parri, H.R.; Soriano, J.; Saad, D. Inferring structure of cortical neuronal networks from activity data: A statistical physics approach. Proc. Natl. Acad. Sci. Nexus 2025, 4, pgae565. [Google Scholar] [CrossRef] [PubMed]
  211. Sporns, O. Contributions and challenges for network models in cognitive neuroscience. Nat. Neurosci. 2014, 17, 652–660. [Google Scholar] [CrossRef] [PubMed]
  212. Sporns, O.; Betzel, R.F. Modular brain networks. Annu. Rev. Psychol. 2016, 67, 613–640. [Google Scholar] [CrossRef] [PubMed]
  213. Sporns, O. Graph theory methods: Applications in brain networks. Dialogues Clin. Neurosci. 2018, 20, 111–121. [Google Scholar] [CrossRef] [PubMed]
  214. Ranasinghe, P.V.; Mapa, M.S. Functional connectivity and cognitive decline: A review of rs-fMRI, EEG, MEG, and graph theory approaches in aging and dementia. Explor. Med. 2024, 5, 797–821. [Google Scholar] [CrossRef]
  215. Xerri, C.; Zennou-Azogui, Y. Influence of the postlesion environment and chronic piracetam treatment on the organization of the somatotopic map in the rat primary somatosensory cortex after focal cortical injury. Neuroscience 2003, 118, 161–177. [Google Scholar] [CrossRef] [PubMed]
  216. Goldstein, L.B. Neuropharmacology of TBI-induced plasticity. Brain Inj. 2003, 17, 685–694. [Google Scholar] [CrossRef]
  217. Plewnia, C.; Hoppe, J.; Cohen, L.G.; Gerloff, C. Improved motor skill acquisition after selective stimulation of central nore-pinephrine. Neurology 2004, 62, 2124–2126. [Google Scholar] [CrossRef] [PubMed]
  218. Loubinoux, I.; Pariente, J.; Boulanouar, K.; Carel, C.; Manelfe, C.; Rascol, O.; Celsis, P.; Chollet, F. A single dose of the serotonin neurotransmission agonist paroxetine enhances motor output: Double-blind, placebo-controlled, fMRI study in healthy subjects. NeuroImage 2002, 15, 26–36. [Google Scholar] [CrossRef]
  219. Dam, M.; Tonin, P.; De Boni, A.; Pizzolato, G.; Casson, S.; Ermani, M.; Freo, U.; Piron, L.; Battistin, L. Effects of fluoxetine and maprotiline on functional recovery in poststroke hemiplegic patients undergoing rehabilitation therapy. Stroke 1996, 27, 1211–1214. [Google Scholar] [CrossRef] [PubMed]
  220. Sawaki, L.; Boroojerdi, B.; Kaelin-Lang, A.; Burstein, A.H.; Bütefisch, C.M.; Kopylev, L.; Davis, B.; Cohen, L.G. Cholinergic influences on use-dependent plasticity. J. Neurophysiol. 2002, 87, 166–171. [Google Scholar] [CrossRef] [PubMed]
  221. Ziemann, U.; Lönnecker, S.; Steinhoff, B.J.; Paulus, W. The effect of lorazepam on the motor cortical excitability in man. Exp. Brain Res. 1996, 109, 127–135. [Google Scholar] [CrossRef] [PubMed]
  222. Walker-Batson, D.; Curtis, S.; Natarajan, R.; Ford, J.; Dronkers, N.; Salmeron, E.; Lai, J.; Unwin, D.H. A double-blind, placebo- controlled study of the use of amphetamine in the treatment of aphasia. Stroke 2001, 32, 2093–2098. [Google Scholar] [CrossRef]
  223. Bragoni, M.; Altieri, M.; Di Piero, V.; Padovani, A.; Mostardini, C.; Lenzi, G.L. Bromocriptine and speech therapy in non-fluent chronic aphasia after stroke. Neurol. Sci. 2000, 21, 19–22. [Google Scholar] [CrossRef]
  224. Kessler, J.; Thiel, A.; Karbe, H.; Heiss, W.D. Piracetam improves activated blood flow and facilitates rehabilitation of poststroke aphasic patients. Stroke 2000, 31, 2112–2116. [Google Scholar] [CrossRef] [PubMed]
  225. Rossi, S.; Rossini, P.M. TMS in cognitive plasticity and the potential for rehabilitation. Trends Cogn. Sci. 2004, 8, 273–279. [Google Scholar] [CrossRef] [PubMed]
  226. Stefan, K.; Kunesch, E.; Cohen, L.G.; Benecke, R.; Classen, J. Induction of plasticity in the human motor cortex by paired associative stimulation. Brain 2000, 123 Pt 3, 572–584. [Google Scholar] [CrossRef] [PubMed]
  227. Oliveri, M.; Bisiach, E.; Brighina, F.; Piazza, A.; La Bua, V.; Buffa, D.; Fierro, B. rTMS of the unaffected hemisphere transiently reduces contralesional visuospatial hemineglect. Neurology 2001, 57, 1338–1340. [Google Scholar] [CrossRef] [PubMed]
  228. Brighina, F.; Bisiach, E.; Oliveri, M.; Piazza, A.; La Bua, V.; Daniele, O.; Fierro, B. 1 Hz repetitive transcranial magnetic stimulation of the unaffected hemisphere ameliorates contralesional visuospatial neglect in humans. Neurosci. Lett. 2003, 336, 131–133. [Google Scholar] [CrossRef] [PubMed]
  229. Mottaghy, F.M.; Hungs, M.; Brügmann, M.; Sparing, R.; Boroojerdi, B.; Foltys, H.; Huber, W.; Töpper, R. Facilitation of picture naming after repetitive transcranial magnetic stimulation. Neurology 1999, 53, 1806–1812. [Google Scholar] [CrossRef] [PubMed]
  230. Grafman, J.; Wassermann, E. Transcranial magnetic stimulation can measure and modulate learning and memory. Neuropsychologia 1999, 37, 159–167. [Google Scholar] [CrossRef]
  231. Boylan, L.S. Enhancing analogic reasoning with rTMS over the left prefrontal cortex. Neurology 2001, 57, 1349. [Google Scholar] [CrossRef] [PubMed]
  232. Evers, S.; Böckermann, I.; Nyhuis, P.W. The impact of transcranial magnetic stimulation on cognitive processing: An event-related potential study. Neuroreport 2001, 12, 2915–2918. [Google Scholar] [CrossRef]
  233. Lee, L.; Siebner, H.R.; Rowe, J.B.; Rizzo, V.; Rothwell, J.C.; Frackowiak, R.S.J.; Friston, K.J. Acute remapping within the motor system induced by low-frequency repetitive transcranial magnetic stimulation. J. Neurosci. 2003, 23, 5308–5318. [Google Scholar] [CrossRef] [PubMed]
  234. Canavero, S.; Bonicalzi, V.; Paolotti, R.; Castellano, G.; Greco-Crasto, S.; Rizzo, L.; Davini, O.; Maina, R. Therapeutic extradural cortical stimulation for movement disorders: A review. Neurol. Res. 2003, 25, 118–122. [Google Scholar] [CrossRef]
  235. Tsubokawa, T.; Katayama, Y.; Yamamoto, T.; Hirayama, T.; Koyama, S. Chronic motor cortex stimulation in patients with thalamic pain. J. Neurosurg. 1993, 78, 393–401. [Google Scholar] [CrossRef] [PubMed]
  236. Canavero, S.; Bonicalzi, V. Extradural cortical stimulation for central pain. In Operative Neuromodulation; Acta Neurochirurgica Supplements; Springer: Vienna, Austria, 2007; Volume 97, pp. 27–36. [Google Scholar]
  237. Benabid, A. Deep brain stimulation for Parkinson’s disease. Curr. Opin. Neurobiol. 2003, 13, 696–706. [Google Scholar] [CrossRef] [PubMed]
  238. Lozano, A.M.; Dostrovsky, J.; Chen, R.; Ashby, P. Deep brain stimulation for Parkinson’s disease: Disrupting the disruption. Lancet Neurol. 2002, 1, 225–231. [Google Scholar] [CrossRef] [PubMed]
  239. Duffau, H.; Capelle, L.; Lopes, M.; Faillot, T.; Sichez, J.P.; Fohanno, D. The insular lobe: Physiopathological and surgical considerations. Neurosurgery 2000, 47, 801–810, discussion 810–811. [Google Scholar] [CrossRef]
  240. Duffau, H.; Bauchet, L.; Lehéricy, S.; Capelle, L. Functional compensation of the left dominant insula for language. Neuroreport 2001, 12, 2159–2163. [Google Scholar] [CrossRef] [PubMed]
  241. Duffau, H.; Denvil, D.; Capelle, L. Absence of movement disorders after surgical resection of glioma invading the right striatum. J. Neurosurg. 2002, 97, 363–369. [Google Scholar] [CrossRef]
  242. Duffau, H.; Capelle, L.; Denvil, D.; Sichez, N.; Gatignol, P.; Lopes, M.; Mitchell, M.C.; Sichez, J.P.; Van Effenterre, R. Functional recovery after surgical resection of low grade gliomas in eloquent brain: Hypothesis of brain compensation. J. Neurol. Neurosurg. Psychiatry 2003, 74, 901–907. [Google Scholar] [CrossRef] [PubMed]
  243. Duffau, H.; Capelle, L.; Denvil, D.; Gatignol, P.; Sichez, N.; Lopes, M.; Sichez, J.P.; Van Effenterre, R. The role of dominant premotor cortex in language: A study using intraoperative functional mapping in awake patients. NeuroImage 2003, 20, 1903–1914. [Google Scholar] [CrossRef]
  244. Duffau, H.; Khalil, I.; Gatignol, P.; Denvil, D.; Capelle, L. Surgical removal of corpus callosum infiltrated by low-grade glioma: Functional outcome and oncological considerations. J. Neurosurg. 2004, 100, 431–437. [Google Scholar] [CrossRef]
  245. Duffau, H.; Denvil, D.; Capelle, L. Long term reshaping of language, sensory, and motor maps after glioma resection: A new parameter to integrate in the surgical strategy. J. Neurol. Neurosurg. Psychiatry 2002, 72, 511–516. [Google Scholar]
  246. Naeser, M.A.; Palumbo, C.L.; Helm-Estabrooks, N.; Stiassny-Eder, D.; Albert, M.L. Severe nonfluency in aphasia. Role of the medial subcallosal fasciculus and other white matter pathways in recovery of spontaneous speech. Brain 1989, 112 Pt 1, 1–38. [Google Scholar] [CrossRef]
  247. Holodny, A.I.; Schwartz, T.H.; Ollenschleger, M.; Liu, W.C.; Schulder, M. Tumor involvement of the corticospinal tract: Diffusion magnetic resonance tractography with intraoperative correlation. J. Neurosurg. 2001, 95, 1082. [Google Scholar] [CrossRef] [PubMed]
  248. Peraud, A.; Meschede, M.; Eisner, W.; Ilmberger, J.; Reulen, H.J. Surgical resection of grade II astrocytomas in the superior frontal gyrus. Neurosurgery 2002, 50, 966–975, discussion 975–977. [Google Scholar] [PubMed]
  249. Duffau, H. Intraoperative cortico–subcortical stimulations in surgery of low-grade gliomas. Expert Rev. Neurother. 2005, 5, 473–485. [Google Scholar] [CrossRef] [PubMed]
  250. Döbrössy, M.D.; Dunnett, S.B. The influence of environment and experience on neural grafts. Nat. Rev. Neurosci. 2001, 2, 871–879. [Google Scholar] [CrossRef]
  251. Bachoud-Lévi, A.C.; Rémy, P.; Nguyen, J.P.; Brugières, P.; Lefaucheur, J.P.; Bourdet, C.; Baudic, S.; Gaura, V.; Maison, P.; Haddad, B.; et al. Motor and cognitive improvements in patients with Huntington’s disease after neural transplantation. Lancet 2000, 356, 1975–1979. [Google Scholar] [CrossRef] [PubMed]
  252. Lindvall, O. Stem cells for cell therapy in Parkinson’s disease. Pharmacol. Res. 2003, 47, 279–287. [Google Scholar] [CrossRef] [PubMed]
  253. Buchan, A.M.; Warren, D.; Burnstein, R. Transplantation of cultured human neuronal cells for patients with stroke. Neurology 2001, 56, 821–822. [Google Scholar] [CrossRef]
Figure 1. Hallmarks of brain plasticity: developmental flexibility, complex interactions between genetic and environmental influences, and structural–functional changes comprising neurogenesis, axonal sprouting, and synaptic remodeling.
Figure 1. Hallmarks of brain plasticity: developmental flexibility, complex interactions between genetic and environmental influences, and structural–functional changes comprising neurogenesis, axonal sprouting, and synaptic remodeling.
Biomedicines 13 00460 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Statsenko, Y.; Kuznetsov, N.V.; Ljubisaljevich, M. Hallmarks of Brain Plasticity. Biomedicines 2025, 13, 460. https://doi.org/10.3390/biomedicines13020460

AMA Style

Statsenko Y, Kuznetsov NV, Ljubisaljevich M. Hallmarks of Brain Plasticity. Biomedicines. 2025; 13(2):460. https://doi.org/10.3390/biomedicines13020460

Chicago/Turabian Style

Statsenko, Yauhen, Nik V. Kuznetsov, and Milos Ljubisaljevich. 2025. "Hallmarks of Brain Plasticity" Biomedicines 13, no. 2: 460. https://doi.org/10.3390/biomedicines13020460

APA Style

Statsenko, Y., Kuznetsov, N. V., & Ljubisaljevich, M. (2025). Hallmarks of Brain Plasticity. Biomedicines, 13(2), 460. https://doi.org/10.3390/biomedicines13020460

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop