Artigos Científicos

A role for synaptic plasticity in the adolescent development of executive function

L D Selemon

3 de junho de 2013

Translational Psychiatry (2013) 3, e238

A role for synaptic plasticity in the adolescent development of executive function

L D Selemon1

1Department of Neurobiology, Yale University School of Medicine, New Haven, CT, USA

Correspondence: Dr LD Selemon, Department of Neurobiology, PO Box 208001, New Haven, CT 06520-8001, USA. E-mail: ldselemon@aol.com

Top

Abstract

Adolescent brain maturation is characterized by the emergence of executive function mediated by the prefrontal cortex, e.g., goal planning, inhibition of impulsive behavior and set shifting. Synaptic pruning of excitatory contacts is the signature morphologic event of late brain maturation during adolescence. Mounting evidence suggests that glutamate receptor-mediated synaptic plasticity, in particular long term depression (LTD), is important for elimination of synaptic contacts in brain development. This review examines the possibility (1) that LTD mechanisms are enhanced in the prefrontal cortex during adolescence due to ongoing synaptic pruning in this late developing cortex and (2) that enhanced synaptic plasticity in the prefrontal cortex represents a key molecular substrate underlying the critical period for maturation of executive function. Molecular sites of interaction between environmental factors, such as alcohol and stress, and glutamate receptor mediated plasticity are considered. The accentuated negative impact of these factors during adolescence may be due in part to interference with LTD mechanisms that refine prefrontal cortical circuitry and when disrupted derail normal maturation of executive function. Diminished prefrontal cortical control over risk-taking behavior could further exacerbate negative outcomes associated with these behaviors, as for example addiction and depression. Greater insight into the neurobiology of the adolescent brain is needed to fully understand the molecular basis for heightened vulnerability during adolescence to the injurious effects of substance abuse and stress.

Keywords: 

alcohol; depression; dopamine; long term depression (LTD); prefrontal cortex; substance abuse

Top

Adolescent development of executive function

Adolescence is rather inexactly defined as the period beginning with the onset of puberty and ending with the shouldering of adult responsibilities.1 It is a time of increased propensity to engage in risky behaviors that include experimentation with alcohol, tobacco, drugs and sexual behavior. Dahl1 has called the adolescent brain a ‘natural tinderbox’ because gonadal hormones are actively stimulating affective and appetitive behaviors, such as sexual drive, increased emotional intensity, and risk taking, yet the brain systems that regulate and moderate these emotional and appetitive urges are not yet mature.

The prefrontal cortex (PFC) mediates executive functions, i.e., internally guided behavior, goal planning, and impulse control, that form the essence of rational thinking and serve to counter appetitive urges and check risk-taking behavior.2, 3The PFC is the last brain region to mature,4, 5, 6, 7 and therefore not surprisingly the frontal lobe capacity for internally guided behavior, working memory, and organizational skills do not reach full adult functional capacity until mid to late adolescence.8, 9, 10, 11, 12

Crews et al.13 have drawn parallels between adolescence and early sensory critical periods, which are dependent on plasticity of developing sensory connectivity and allow for environmental (sensory) modulation of maturing sensory connections. Specifically, they have suggested that in adolescence PFC circuitry may be endowed with similar plasticity and responsiveness to environmental factors, and as a consequence with heightened vulnerability to the detrimental effects of substance abuse and stress.13

This review examines the literature on adolescent development across species and focuses on the role that glutamate-receptor mediated plasticity may play in maturation of PFC circuitry in adolescence. It is postulated that adolescence represents a phase of increased activity of long term depression (LTD) mechanisms that predispose to synaptic elimination and further that termination of this LTD-permissive phase marks the transition to adulthood. Finally, consideration is given to the possibility that greater vulnerability to substances of abuse and stress may represent an interaction between these environmental factors and the LTD mechanisms of plasticity that are accentuated during adolescence. The hypothesis put forward in this review, while speculative, is intended to spark further research into possible molecular mechanisms associated with adolescent development of the PFC. Certainly synaptic plasticity has been studied much less extensively in the PFC than in the hippocampus; nonetheless, mounting evidence suggests that both long term potentiation (LTP) and LTD play an important role in cognitive functioning mediated by the PFC and perhaps when disturbed in diseases related to malfunction of this cortex.14

Top

Preadolescent development and sensory critical periods

The specificity and topography of brain wiring are not entirely genetically preprogrammed but instead established via dynamic processes occurring in the developing brain. Adolescence represents the final epoch in a series of developmental stages that transform the immature brain into its adult form. In order to fully understand adolescent development, it is important to appreciate how it differs from earlier preadolescent maturation.

The developmental mechanisms that account for major remodeling of connectivity occur before the onset of adolescence, i.e., before postnatal day 28 (PD28) in rodents, 9 months in cats, and 3 years in non-human primates15, 16, 17 and include prominent degeneration of neurons and axons.18, 19 Indeed, the immature mammalian brain is distinguished from its adult counterpart by the presence of connections between brain areas that are not interconnected in the mature brain and by overlap of terminal fields that are segregated in the adult brain. For example, in the newborn hamsters and rats, uncrossed retinocollicular projections, i.e. from the retina to the ipsilateral superior colliculus (SC), not only occupy a much expanded territory in the SC relative to that of the adult brain but also originate from nasal as well as temporal retinal ganglion cells.20, 21, 22 Retraction of the terminal projections is associated with loss of these nasal, ipsilaterally-projecting ganglion cells.22 More generally, in the central nervous system overproduction of neurons with ensuing neuronal death is a common mechanism employed by the developing brain to ensure that the appropriate balance of projection and receptive neurons is attained.19, 23, 24, 25

A second, pervasive form of degeneration in the developing brain is degeneration limited to axonal connections leaving the neurons of origin intact. For instance, in the central nervous system, cortical callosal projections that are widespread in kittens and young rats are constricted to the adult patterning by retraction of callosal axons without cell loss.26, 27, 28 Quantitative analysis of axon number in major tracts underscores the magnitude of this form of degeneration as the number of axons in the young non-human primate brain ranges from twice (optic tract) to 3.5 times (corpus callosum) the number in the adult brain.29, 30, 31 Both forms of degeneration, involving loss of neurons or loss of axons, necessarily are associated with dissolution of established synapses.32 However, these early developmental events are occurring at a time when, overall, synapses are increasing in density.33, 34, 35, 36, 37, 38 The classic example or early connectional remodeling, that of reduction of polyneuronal input on a single muscle fiber to a single axon, illustrates how synaptic number may increase as the surviving single axon sprouts a much more elaborate terminal plexus.18, 39 Likewise, in the central nervous system regression of inappropriate synapses is more than compensated by growth and expansion of appropriate terminal fields.40

A wealth of evidence has established that reorganization of connections throughout the brain is activity dependent and therefore mediated by a Hebbian mechanism.41, 42, 43, 44, 45 Although normal regression of connections in the visual system can proceed in the absence of visual input41 there exists a period of plasticity during postnatal development that permits rewiring in response to altered sensory environments.43, 46, 47 It is noteworthy that critical periods for sensory plasticity occur in the same preadolescent period in which remodeling of connectivity occurs.34, 48, 49

Top

Adolescence: synaptic elimination and excitatory/inhibitory balance

The maturational event most consistently linked to the adolescent stage of development is reduction of synaptic density or ‘synaptic pruning.’ Quantitative analyses of synapses in the non-human primate uncovered a synchronous increase in synaptic density in multiple cortical areas that peaks during the postnatal 3rd month, declines slowly (10%) until~2 years of age with a steeper decline (40%) occurring between 2.7 and 5 years (adulthood).35, 36, 37, 38 In the human cortex the timing of peak synaptic density is staggered in different regions, but the basic pattern of peak synaptic density in early childhood followed by robust synapse elimination throughout early (auditory cortex) or mid-adolescence (PFC) is in basic agreement with non-human primate studies.4, 50 More recent data have established that synapse elimination in humans does not end in adolescence but continues at a lower rate into early adulthood.51 In addition, in human cortex the synaptic related proteins synaptophysin and postsynaptic density protein-95 (PSD-95) show similar patterns of peak in childhood and decline through adolescence,52 though it should be noted that a recent study found increasing concentrations of synaptic–related molecules throughout the adolescent epoch.53 Nonetheless, most evidence points to synaptic pruning as the signature late maturational process associated with adolescence. Other species have been studied less extensively but exhibit a comparable pattern. Peak synaptic density was observed by the 7th postnatal week in the cat.34 In rat, recent data suggest that peak spine density in the PFC is present at PD31 with spine density decreasing thereafter until PD 57 or PD60, i.e., early adulthood.33

Synaptic elimination in adolescence is widely thought to account for the decline in gray matter volume detected via longitudinal magnetic resonance imaging (MRI) of human subjects. Although reduction of synaptic connectivity might be accompanied by retraction of glial and neuronal processes, elimination of neuronal cell bodies occurs much earlier in development.54 One of the first longitudinal MRI studies of human subjects detected divergent developmental growth patterns in gray and white matter volumes: white matter volume increased linearly until approximately age 22 whereas cortical gray matter volume in the frontal and parietal lobes peaked just prior to adolescence (~10–12 years) and then declined to adult volumes.5 Cross-sectional studies of children and adolescents, including one recent large multicenter study, also show opposing patterns for gray and white matter.55, 56, 57 Interestingly, changing cortical volumes over this age range are most prominent in the frontal and parietal lobes.8, 58, 59 Indeed, a recent study indicates that there is a progression in which higher cortical association areas like the PFC are last to show reduction of gray matter volume.7

The functional significance of synaptic elimination during adolescence, though still enigmatic, probably involves adjustment of the excitatory/inhibitory balance on individual neurons and within networks. The main argument in support of this hypothesis stems from the specificity of the loss: excitatory synapses are selectively degenerated whereas inhibitory synapses are spared.35, 37 Even loss of chandelier axon boutons in the PFC, a finding that was originally interpreted as loss of inhibitory synapses,60 now supports the elimination of excitatory input in light of new physiologic data.61 Furthermore, recent evidence has established that D2 dopamine receptors on interneurons undergo a profound maturational change during adolescence.62, 63, 64 Prior to adolescence, D2 stimulation elicits either no effect or only weak inhibition on interneurons. However, in adult animals stimulation of D2 receptors is strongly excitatory and therefore results in robust firing of interneurons and potent inhibition of their pyramidal cell targets. As a result, inhibition gains a position of ascendency in adolescence via increased dopamine-mediated firing of interneurons as well as a relative gain in inhibitory/excitatory synapse ratio. In the PFC, neurophysiologic studies have established a critical role for inhibitory synapses in mediating information flow through local networks.65, 66 Moreover, fast-spiking interneurons mediate gamma oscillations which are essential to cortical computation in many areas of the cortex and to cognitive processing in the PFC.67, 68 Thus, the correct balance of inhibition and excitation seems to be critical for normative executive function, and conversely, disturbance of this balance is thought to be a fundamental component of psychiatric illness.69, 70

Top

Molecular mechanisms associated with synaptic stabilization and synaptic pruning

Synapse stabilization and synapse elimination are primary players in the maturational processes associated with preadolescent and adolescent development. The transition of nascent synapses into mature synapses represents the first step in synapse stabilization. N-methyl-D-aspartate receptors (NMDAR) are localized very early to the postsynaptic membrane, but transition to a more mature synapse state is characterized by recruitment of alpha-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid receptors (AMPAR) to the synapse.71, 72, 73,74 Expression of AMPARs on the postsynaptic membrane is induced by NMDAR-mediated long term potentiation (LTP), the same mechanism originally described in hippocampus for learning and memory.73, 74, 75, 76 A second NMDAR-mediated process, LTD, results when afferent stimulation fails to activate a target neuron.76 In many respects LTP and LTD are opposite processes although they engage distinct intracellular signaling mechanisms.77, 78, 79, 80 Essentially stimulation of NMDARs can induce activity-dependent strengthening of synapses via LTP or weakening via LTD, and AMPAR insertion or removal from the postsynaptic membrane is the conduit for this change in synaptic strength.81, 82Importantly, LTP and LTD do not just strengthen or weaken synaptic connections (short term plasticity) but actually trigger the addition or loss of synapses (long term plasticity) even in the adult brain.83, 84, 85, 86, 87, 88, 89, 90, 91, 92

Well before the role of NMDAR-mediated LTP in establishing mature synaptic connections was recognized, Constantine-Paton et al.44 postulated that activity-dependent remodeling of connectivity in the developing brain might be mediated by NMDARs because these receptors are perfectly suited to detection of synchronized pre- and postsynaptic activation. Growing evidence now supports the idea that LTP and LTD are required for generation of whisker barrel field maps in the primary somatosensory cortex and ocular dominance columns in the primary visual cortex, both involving reorganization of thalamic inputs to layer 4.49, 93, 94,95, 96 In development, as in learning and memory, plasticity is bidirectional, i.e., synchronized activity of afferent inputs may trigger LTP and resulting synapse maturation and stabilization; conversely, asynchronous activity may diminish synaptic strength via LTD and predispose the synapse to elimination.97

Recently, changes in the NMDAR have been linked to critical periods of early developmental plasticity. NMDAR subunit composition shifts from NR2B predominant to NR2A prevalent forms in early development in the visual and somatosensory cortices.98, 99, 100 Moreover, the shift in NR2B to NR2A shows a rough correspondence to critical periods for sensory plasticity: the beginning of the critical period is marked by an increase in NR2A expression, and the end of the critical period is associated with a decrease in NR2B expression.100, 101Importantly, the switch is not locked to a specific age but in fact can be delayed by sensory deprivation, suggesting that it is controlled by activity.93, 102, 103, 104,105 In turn, the changeover from NR2B to NR2A receptor subtypes controls the sensitivity of these connections to stimulation by NMDARs. For example, in the primary visual cortex of the ferret, NR2B levels are already high at eye opening and decline in layer 4 at the end of the critical period for plasticity of ocular dominance columns but remain high in layer 2/3.106 Correspondingly, physiologic studies in the cat visual cortex have shown that cortical layer 4 neurons, but not layer 2/3 cells, exhibit reduced sensitivity of visual and spontaneous activity to an NMDAR antagonist at the end of the critical period.107 Together these findings suggest that the switch from NR2B to NR2A dominated receptors terminates the critical period of experience-dependent plasticity for establishment of ocular dominance columns in the visual cortex.

NMDAR-mediated LTP and LTD may also constitute the molecular underpinnings for adolescent synaptic pruning, albeit with greater emphasis on LTD and synaptic elimination. How could the same mechanism account for two very different developmental processes? Perhaps the period of adolescence corresponds to a widespread shift in the balance of LTP/LTD mechanisms and a corresponding prevalence of synaptic elimination over synaptic addition. In rat hippocampal slices, an increased NR2A/NR2B ratio has been linked to decreased spine motility and increased synaptic stabilization, suggesting a role in the NMDAR subunit composition in halting synaptogenesis.108 Furthermore, the NR2A state is less conducive to LTP. This is because calcium/calmodulin-dependent protein kinase II (CaMKII), which has a well established role in LTP,109, 110 binds preferentially to the NR2B subunit.110, 111, 112 Accordingly, NR2B expression on the postsynaptic membrane has been shown to be necessary for LTP induction, while a role for NR2A in LTP is not well established.113, 114, 115 Moreover, NR2A expression is enhanced by ligand binding to NMDARs and therefore is activity-modulated whereas NR2B expression is not dependent on previous activity.116 The NR2A subunit is therefore thought to be responsible for metaplasticity of synapses, i.e., a change in the likelihood of subsequent synaptic plasticity.117, 118 With age and activity, NR2A subunits become incorporated into the postsynaptic membrane, replacing NR2B subunits.116 The resulting increased NR2A/NR2B ratio translates into a higher threshold for induction of LTP and conversely a state that is more favorable for induction of LTD.118, 119

The role of plasticity in the neocortex is not as well established as in the hippocampus. However, NMDAR-mediated LTP and LTD have been described in the visual neocortex120 and at multiple synapses in the PFC.121, 122, 123 Notably, LTD mediated through metabotropic glutamate receptors (mGluRs) has emerged as a major alterative to NMDAR-mediated LTD in widespread areas of the brain124, 125,126 and therefore deserves consideration as a possible molecular basis for synaptic pruning in the PFC. In this regard, mGluR plasticity has been described at the thalamocortical synapse in the somatosensory cortex,127 perhaps indicating that this form of plasticity is also present at the mediodorsal thalamic synapses in the PFC. However, at the thalamocortical synapse, mGluR LTD acts presynaptically to decrease transmitter release and depress synaptic activity.127Such a mechanism would be unlikely to result in synapse loss and spine involution and therefore would not be a strong candidate for LTD-facilitated synaptic pruning during adolescence. Furthermore, mGluR LTD at postsynaptic sites in the hippocampus has been associated with large spines containing an abundance of AMPAR.128 Unlike the hippocampus where large mushroom spines are in the majority, thin, filopodial spines predominate in the PFC.129 Thus, strong evidence for mGluR in plasticity related to PFC synaptic pruning is presently lacking; nonetheless, possible involvement of mGLuR-mediated LTD in prefrontal adolescent maturation cannot be discounted.

Many questions remain to be answered about the role of metaplasticity in the PFC as well. As the NR2A subtype promotes a LTD-receptive state in the synapse and LTD is associated with synaptic elimination, it would be interesting to know whether and when the NR2B to NR2A switch occurs in the PFC and how it relates to the synaptic pruning that refines connectivity associated with cognitive control of behavior. If the LTD-receptive state is a hallmark of adolescent development, a reasonable presumption is that there exists an additional molecular switch that greatly curtails the LTD-receptive state of adolescence into the much less receptive state of adulthood. This switch, although presently unidentified, would transform the synapse into a state that is less receptive to alterations in AMPAR expression on the postsynaptic membrane. Given that synaptic pruning continues into early adulthood albeit at a lower level than that of adolescence,33, 51 it seems likely that the transition phase is gradual rather than abrupt resulting in a much less plastic state by the end of the third decade in humans.

Top

Adolescence development of cognitive function and synaptic plasticity

Executive functions governed by the PFC exhibit a prolonged period of maturation reaching fulfillment only in late adolescence.11, 130 Volumetric changes occurring during adolescence have been correlated with improved cognitive performance, e.g., verbal and spatial memory performance is positively correlated with gray matter thinning in the frontal lobes.6 General intelligence has also been shown to bear a relationship to the trajectory of frontal cortical gray matter thinning, such that subjects with superior intelligence show a robust early adolescent increase in gray matter volume followed by equally robust thinning during later adolescence.131 However, too much cortical thinning during adolescence has been associated with diseased states such as Attention Deficit Hyperactivity Disorder (ADHD).132 Thus, there is an optimal level of synaptic pruning that is essential to normal development of adult cognitive function.

One recent study addressed the role of AMPAR expression and LTD in the development of PFC function in the mouse. Vazdarjanova et al.133 utilized a transgenic mouse that over-expresses calcyon, a protein which mediates activity-dependent AMPAR internalization, and found that calcyon over-expression over the lifetime of the mouse resulted in marked impairment of contextual fear extinction (CFE) and working memory capacity, both dependent on normal PFC function. Most relevant to this discussion, adolescence was the critical period for production of these deficits. When over-expression was silenced specifically during the adolescent epoch, normal CFE function was rescued.133 One possible explanation for these findings is that AMPAR internalization and associated functions like LTD are more sensitive to regulation during adolescence and this regulation is turned off, or at least greatly diminished, in the adult brain. Whether overactive LTD during adolescence translates into altered synaptic number in the PFC or elsewhere is currently not known. However, it is interesting that upregulated calcyon expression has been found in schizophrenia, a neurodevelopmental disease in which PFC gray matter deficits are prominent.134,135, 136

In the PFC, synaptic plasticity is highly modulated by dopamine receptor, especially the D1 receptor.14, 122, 137 This is not surprising since D1 receptor stimulation has been shown to trigger phosphorylation of AMPAR, which in turn promotes trafficking of these receptors to the external membrane.138, 139 The D1 receptor is therefore strategically positioned to effect changes in AMPAR synaptic expression and ultimately in synaptic strength and/or number. In the adult non-human primate, long term sensitizing regimens of amphetamine decrease spine density on pyramidal cells in the PFC and have detrimental effects on working memory performance.140 Moreover, these effects seem to be due to changes at the D1 receptor because both cognitive and morphologic effects on PFC pyramidal neurons can be reversed by long term treatment with a D1 antagonist.141 If AMPAR-mediated LTD expression is in a state of greater sensitivity to modulation in adolescence, then D1 receptor-stimulated interference with this mechanism could be magnified during adolescence resulting in exaggerated consequences at the synapse. Other known modulators of synaptic plasticity, e.g. D2,139muscarinic,142 and cannabinoid143 receptors, might have similarly increased potency during the adolescent period.

Top

Adolescence vulnerability to environmental factors

Adolescence has been described as a period of accentuated opportunity and of enhanced vulnerability.1 It has long been recognized that early onset of substance abuse is associated with greater propensity for problem drug use later in life.144, 145, 146, 147 In recent years, the period of adolescent plasticity has been shown to temporally correlate with the time of greatest vulnerability to addiction.148 Some have postulated that addiction conscripts the learning and memory pathways in a maladaptive fashion,149, 150 but the question of why addiction is more devastating in adolescence than in adulthood remains unanswered. Adolescence is also associated with onset of mental illness, as for example depression rates rise in adolescence especially for females,151 and the prodromal phase of psychosis, including early onset schizophrenia, surfaces during the adolescent window.152 Despite the fact that adolescents are bigger and stronger than younger children, mortality rates increase more than 200% from childhood mainly due to accidents, suicide, substance abuse, and eating disorders.1

One of the most studied environmental effects in adolescence is alcohol abuse. In adults, brain toxicity has been documented as a consequence of chronic alcohol abuse: cortical gray matter thinning is most prominent in the PFC153 and associated with changes in neuronal and glia density in both the orbitofrontal154and superior frontal cortices.155 Alarmingly, the detrimental effects of alcohol consumption seem to be magnified in adolescence. Studies in human subjects have shown that impairment of memory function is more pronounced following even acute exposure to alcohol in younger (ages 21–24) than in older (ages 25–29) subjects.156 In adolescent rats, ethanol administration selectively impairs spatial memory whereas adult rats are unaffected by the same doses.157Moreover, ethanol consumption in rats that simulates binge drinking results in more widespread pathology in adolescent animals than in adults.158

The basis of the enhanced vulnerability to alcohol in adolescence is undoubtedly complex and involves interaction with multiple neurotransmitter systems.159 With regard to neuroplasticity, there are well documented effects of alcohol on the glutamate system. Acutely, ethanol inhibits NMDAR neurotransmission whereas long term exposure results in homeostatic upregulation of NMDAR signaling.159, 160There is also growing evidence to suggest that ethanol has a greater effect on glutamate neurotransmission during adolescence than in later life. Ethanol exposure at low doses in adolescent rats is associated with inhibition of NMDAR-mediated EPSCs in the CA1 region of the hippocampus while high doses are required to inhibit EPSCs in adults.161 Ethanol also blocks LTP in CA1 neurons of the hippocampus in adolescent but not adult rats.162 Thus, even acute alcohol consumption in adolescence could disrupt mechanisms of Hebbian plasticity, and more chronic alcohol consumption in adolescence may induce homeostatic upregulation of glutamate neurotransmission that could result in long term changes in synapse number and dendritic spine morphology.160 Homeostatic regulation of synaptic activity, i.e., increases or decreases synaptic scaling across the whole population of synapses, is also thought to be mediated by increased or decreased expression of AMPAR receptors on the post-synaptic membrane.163 This suggests a potential site of interaction between developmental plasticity and homeostatic plasticity since both involve trafficking of AMPARs. Furthermore, sites of homeostatic plasticity correlate with lamina that exhibit plasticity during critical periods in the visual and somatosensory cortices, suggesting a possible mechanism for heightened vulnerability of selected circuitry during different phases of development.163 If synaptic plasticity in adolescence is primarily occurring in the neural circuitry that mediates executive processing, then disruption of synaptic plasticity at this time might result in enduring deficits in control of emotion, logical thinking and inhibition of impulsivity. In turn this lack of executive control could exacerbate the addictive tendencies and result in more severe alcoholism.

The adolescent brain is also more responsive to stress than the adult brain164 and as a consequence may be more vulnerable to depression.151 Analogous to the manner in which alcohol has age-specific effects that may be depend on which regions of the brain are most plastic, a recent study has shown that the effects of sexual abuse, presumably the stress associated with the abuse, produces different brain pathology at childhood and adolescent ages.165 Notably, frontal gray matter volume deficits were most pronounced in adult subjects who experienced sexual abuse at ages 14–16.165

The neural pathways that mediate and modulate the stressful effects on cognitive function in the PFC involve monoamine signaling.164 Given the prominence of dopamine neurotransmission in mediating stress, the development of dopamine innervation of the PFC during late maturation might provide insight into the enhanced sensitivity to stress at this age. In the non-human primate, dopamine innervation of the middle PFC layers peaks near the onset of puberty and then decreases rapidly to adult levels while innervation of other layers is stable throughout the postnatal period.166 D1 receptor levels also peak and decline to adult levels around the beginning of puberty.167 These findings which indicate that the adult D1 receptor pattern is reached early do not appear to support a role for dopamine in adolescent enhancement of plasticity. However, in rodent prefrontal cortex, cell specificity has been observed in the distribution of D1 receptors with pyramidal cell neurons, but not interneurons, expressing higher levels of D1 receptors in adolescence than in adulthood.168 These rodent data suggest that changes in D1 receptor expression might accentuate dopamine signaling in adolescence and thereby account for greater plasticity during this critical period. However, a credible alternative explanation is that the LTD-receptive state of adolescence is more sensitive to modulators like dopamine and that critical differences might be found in the mechanisms of glutamate receptor-mediated synaptic plasticity in the adolescent brain compared its adult counterpart.

Top

Clinical considerations

Identifying the molecular basis for synaptic pruning in adolescence could have wide ranging clinical ramifications. If NMDA-mediated LTD were proved to underlie reduction of connectivity, then the intracellular pathways associated with LTD processes, including those that mediate AMPAR internalization, could be targeted to curtail excessive synaptic pruning in diseases such as schizophrenia and ADHD. Because the D1 receptor is a key modulator of synaptic plasticity in the PFC and can even determine polarity of plasticity, i.e. high dopamine levels can predispose prefrontal synapses to LTD over LTP,137 treatment with dopaminergic antagonists or drugs that target intracellular dopamine signaling might also be useful in decreasing overactive LTD mechanisms. Along these same lines, drugs that impact the D1 receptor or its signaling pathways could ameliorate the impact of stress on the adolescent brain of individuals at risk for depression. Likewise, the involvement of glutamate receptors, including mGluRs,126 in drug and alcohol addiction raises the possibility that pharmacologic targeting of glutamate signaling might have the potential to diminish the long term consequences of substance abuse in adolescence. In the same manner that the discovery of aberrant mGluR5 mechanisms in Fragile X syndrome has spawned new therapeutic approaches for treating this disease,169, 170, 171 greater insight into the molecular substrates of adolescent maturation of the prefrontal cortex might lead to similar novel drug development for disorders and environmental exposures linked to abnormal adolescent development.

Top

Conclusions

The adolescent epoch is a time when refinement of connectivity establishes the proper excitatory/inhibitory balance in the PFC, and it is a critical period for normal maturation of executive functioning. Adolescence is postulated to be a time when LTD-driven synaptic pruning is occurring at a high rate in regions that govern higher cognitive function like the PFC. Further, the transition to adulthood is hypothesized to be marked by changes in the synapse that make the mature neuron less sensitive to AMPAR internalization, less likely to undergo LTD and thus less likely to undergo retraction of synaptic contacts.

Top

Conflict of interest

The author declares no conflict of interest.

Top

References

  1. Dahl RE. Adolescent brain development: A period of vulnerabilities and opportunities. Ann NY Acad Sci 2004; 1021: 1–22. | Article | PubMed |
  2. Goldman-Rakic PS. Circuitry of the prefrontal cortex and the regulation of representational knowledge. In Plum F, Moutcastle V, (eds) Handbook of Physiology vol. 5. American Physiological Society: Bethesda, MD, 373–4171987.
  3. Fuster JM. Frontal lobe and cognitive development. J Neurocytol 2002; 31: 373–385. | Article | PubMed |
  4. Huttenlocher PR, Dabholkar AS. Regional differences in synaptogenesis in human cerebral cortex. J Comp Neurol 1997; 387: 167–178. | Article | PubMed | ISI | CAS |
  5. Giedd JN, Blumenthal J, Jeffries NE, Castellanos FX, Liu H, Zijdenbos A et al. Brain development during childhood and adolescence: A longitudinal MRI study. Nature Neurosci 1999; 2: 861–863.
  6. Sowell ER, Delis D, Stiles J, Jernigan TL. Improved memory functioning and frontal lobe maturation between childhood and adolescence: A structural MRI study. J Internatl Neuropsychol Soc 2001; 7: 312–322. | Article |
  7. Gotay N, Giedd JN, Lusk L, Hayashi KM, Greenstein D, Vaituzis AC et al. Dynamic mapping of human cortical development during childhood through early adulthood. Proc Natl Acad Sci USA 2004; 101: 8174–8179. | Article | PubMed | CAS |
  8. Sowell ER, Trauner DA, Gamat A, Jernigan TL. Development of cortical and subcortical brain structures in childhood and adolescence: A structural MRI study. Develop Med Child Neurol 2002; 44: 4–16. | Article | PubMed |
  9. De Luca DR, Wood SJ, Anderson V, Buchanan JA, Proffitt TM, Mahony K et al. Normative data from the CANTAB. I: Development of function over the lifespan. J Clin Exp Neuropsychol 2003; 25: 242–254. | Article | PubMed | ISI |
  10. Luna B, Garver KE, Urban TA, Lazar NA, Sweeney JA. Maturation of cognitive processes from late childhood to adulthood. Child Develop 2004;75: 1357–1372. | Article | PubMed |
  11. Luciana M, Conklin HM, Hooper CJ, Yarger RS. The development of nonverbal working memory and control processes in adolescents. Child Develop 2005;76: 697–712. | Article | PubMed |
  12. Best JR, Miller PH. A developmental perspective on executive function. Child Develop 2010; 81: 1641–1660. | Article | PubMed |
  13. Crews F, He J, Hodge C. Adolescent cortical development: A critical period of vulnerability for addiction. Pharmacol Biochem Behav 2007; 86: 189–199. | Article | PubMed |
  14. Goto Y, Yang CR, Otani S. Functional and dysfunctional synaptic plasticity in prefrontal cortex: Roles in psychiatric disorders. Biol Psychiatry 2010; 67: 199–207. | Article | PubMed | ISI |
  15. Spear LP. Adolescent brain development and animal models. Ann NY Acad Sci2004; 1021: 23–26. | Article | PubMed |
  16. Plant TM. A study of the role of the postnatal testes in determining the ontogeny of gonadotropin secretion in the male rhesus monkey (Macaca mulatta). Endocrinol 1985; 116: 1341–1350. | Article |
  17. Butterwick RF, McConnell M, Markwell PJ, Watson TD. Influence of age and sex on plasma lipid and lipoprotein concentrations and associated enzyme activities in cats. Am J Vet Res 2001; 62: 331–336. | Article | PubMed |
  18. Purves D, Lichtman JW. Elimination of synapses in the developing nervous system. Science 1980; 210: 153–157. | Article | PubMed | ISI | CAS |
  19. Cowan WM, Fawcett JW, O’Leary DDM, Stanfield BB. Regressive events in neurogenesis. Science 1984; 225: 1258–1265. | Article | PubMed | ISI | CAS |
  20. Land PW, Lund RD. Development of the rat’s uncrossed retinotectal pathway and its relation to plasticity studies. Science 1979; 205: 698–700. | Article | PubMed | ISI | CAS |
  21. Frost DO, So K-F, Schneider GE. Postnatal development of retinal projections in Syrian hamsters: A study using autoradiographic and anterograde degeneration techniques. Neuroscience 1979; 4: 1649–1677. | Article | PubMed | ISI | CAS |
  22. Insausti R, Blackemore C, Cowan WM. Ganglion cell death during development of ipsilateral retino-collicular projection in golden hamster.Nature 1984; 308: 362–365. | Article | PubMed |
  23. Williams RW, Herrup K. The control of neuron number. Ann Rev Neurosci1988; 11: 423–453. | Article | PubMed |
  24. Williams RW, Rakic P. Elimination of neurons from the rhesus monkey’s lateral geniculate nucleus during development. J Comp Neurol 1988; 272: 424–436. | Article | PubMed |
  25. Lotto RB, Asavartikra P, Vail L, Price DJ. Target-derived neurotrophic factors regulate the death of developing forebrain neurons after a change in their trophic requirements. J Neurosci 2001; 21: 3904–3910. | PubMed | ISI | CAS |
  26. Innocenti GM. Growth and reshaping of axons in the establishment of visual callosal connections. Science 1981; 212: 824–827. | Article | PubMed | ISI | CAS |
  27. O’Leary DDM, Stanfield BB, Cowan WM. Evidence that the early postnatal restriction of the cells of origin of the callosal projection is due to the elimination of axonal collaterals rather than to the death of neurons.Develop Brain Res 1981; 1: 607–617. | Article |
  28. Ivy GO, Killackey HP. Ontogenetic changes in the projections of neocortical neurons. J Neurosci 1982; 2: 735–743. | PubMed | ISI | CAS |
  29. Rakic P, Riley KP. Overproduction and elimination of retinal axons in the fetal rhesus monkey. Science 1983; 219: 1441–1444. | Article | PubMed | ISI | CAS |
  30. LaMantia AS, Rakic P. Axon overproduction and elimination in the corpus callosum of the developing rhesus monkey. J Neurosci 1990; 10: 2156–2175. | PubMed | ISI | CAS |
  31. LaMantia AS, Rakic P. Axon overproduction and elimination in the anterior commissure of the developing rhesus monkey. J Comp Neurol 1994; 340: 328–336. | Article | PubMed |
  32. Campbell G, Shatz CJ. Synapses formed by identified retinogeniculate axons during segregation of eye input. J Neurosci 1992; 12: 1847–1858. | PubMed | ISI | CAS |
  33. Gourley SL, Olevska A, Sloan Warren M, Taylor JR, Koleske AJ. Arg kinase regulates prefrontal dendritic spine refinement and cocaine-induced plasticity. J Neurosci 2012; 32: 2314–2323. | Article | PubMed |
  34. Cragg BG. The development of synapses in the visual system of the cat. J Comp Neurol 1972; 160: 147–166. | Article |
  35. Rakic P, Bourgeois J-P, Eckenhoff MF, Zecevic N, Goldman-Rakic PS. Concurrent overproduction of synapses in diverse regions of the primate cerebral cortex. Science 1986; 232: 232–235. | Article | PubMed | ISI | CAS |
  36. Zecevic N, Bourgeois J-P, Rakic P. Changes in synaptic density in motor cortex of rhesus monkey during fetal and postnatal life. Develop Brain Res1989; 50: 11–32. | Article |
  37. Bourgeois J-P, Rakic P. Changes of synaptic density in the primary visual cortex of the macaque monkey from fetal to adult stage. J Neurosci 1993;13: 2801–2820. | PubMed | ISI | CAS |
  38. Bourgeois J-P, Goldman-Rakic PS, Rakic P. Synaptogenesis in the prefrontal cortex of rhesus monkeys. Cereb Cortex 1994; 4: 78–96. | Article | PubMed | ISI | CAS |
  39. Redfern PA. Neuromuscular transmission in new-born rats. J Physiol 1970;209: 701–709. | PubMed | CAS |
  40. Sretavan D, Shatz CJ. Prenatal development of individual retinogeniculate axons during the period of segregation. Nature 1984; 308: 845–848. | Article | PubMed | ISI | CAS |
  41. Changeux J-P, Danchin A. Selective stabilisation of developing synapses as a mechanism for the specification of neuronal networks. Nature 1976; 264: 705–712. | Article | PubMed | ISI | CAS |
  42. Stryker MP, Harris WA. Binocular impulse blockade prevents the formation of ocular dominance columns in cat visual cortex. J Neurosci 1986; 6: 2117–2133. | PubMed | ISI | CAS |
  43. Shatz CJ, Stryker MP. Ocular dominance in layer IV of the cat’s visual cortex and the effects of monocular deprivation. J Physiol 1978; 281: 267–283. | PubMed | ISI | CAS |
  44. Constantine-Paton M, Cline HT, Debski E. Patterned activity, synaptic convergence, and the NMDA receptor in developing visual pathways. Ann Rev Neurosci 1990; 13: 129–154. | Article | PubMed |
  45. Shatz CJ. Impulse activity and the patterning of connections during CNS development. Neuron 1990; 5: 745–756. | Article | PubMed | ISI | CAS |
  46. Hubel H, Wiesel TN, LeVay S. Plasticity of ocular dominance columns in monkey striate cortex. Philos Trans Roy Soc Lond. Ser B Biol Nerv Syst1977; 278: 377–409. | Article |
  47. LeVay S, Wiesel TN, Hubel DH. The development of ocular dominance columns in normal and visually deprived monkeys. J Comp Neurol 1980; 191: 1–51. | Article | PubMed | ISI | CAS |
  48. Hubel DH, Wiesel TN. The period of susceptibility to the physiological effects of unilateral eye closure in kittens. J Physiol 1970; 206: 419–436. | PubMed | ISI | CAS |
  49. Foeller E, Feldman DE. Synaptic basis for developmental plasticity in somatosensory cortex. Curr Opin Neurobiol 2004; 14: 89–95. | Article | PubMed | ISI | CAS |
  50. Huttenlocher PR. Synaptic density in human frontal cortex – Developmental changes and effects of aging. Brain Res 1979; 163: 195–205. | Article | PubMed | ISI | CAS |
  51. Petanjek A, Judas M, Simic G, Roko Rasin M, Uylings HBM, Rakic P et al. Extraordinary neoteny of synaptic spines in the human prefrontal cortex.Proc Nat Acad Sci USA 2011; 108: 13281–13286. | Article | PubMed |
  52. Glantz LA, Gilmore JH, Hamer RM, Lieberman JA, Jarskog LF. Synaptophysin and postsynaptic density protein 95 in the human prefrontal cortex from mid-gestation into early adulthood. Neuroscience 2007; 149: 582–591. | Article | PubMed | CAS |
  53. Webster MJ, Elashoff M, Weickert CS. Molecular evidence that cortical synaptic growth predominates in the first decade of life in humans. Internat J Develop Neurosci 2011; 29: 225–236. | Article |
  54. Finlay BL, Slattery M. Local differences in the amount of early cell death in the neocortex predict adult local specializations. Science 1983; 219: 1349–1351. | Article | PubMed | ISI | CAS |
  55. Pfefferbaum A, Mathalon DH, Sullivan EV, Rawles JM, Zipursky RB, Lim KO. A quantitative magnetic resonance imaging study of changes in brain morphology from infancy to late adulthood. Arch Neurol 1994; 51: 874–887. | Article | PubMed | ISI | CAS |
  56. Reiss AL, Abrams MT, Singer HS, Ross JL, Denckla MB. Brain development, gender and IQ in children: A volumetric imaging study. Brain 1996; 119: 1763–1774. | Article | PubMed | ISI |
  57. Brain Development Cooperative Group. Total and regional brain volumes in a population-based normative sample from 4 to 18 years: The NIH MRI study of normal brain development. Cereb Cortex 2012; 22: 1–12.
  58. Jernigan TL, Trauner DA, Hesselink JR, Tallal PA. Maturation of human cerebrum observed in vivo during adolescence. Brain 1991; 114: 2037–2049. | Article | PubMed | ISI |
  59. Sowell ER, Thompson PM, Holmes CJ, Batth R, Jernigan TL, Toga AW. Localizing age-related changes in brain structure between childhood and adolescence using statistical parametric mapping. NeuroImage 1999; 9: 587–597. | Article | PubMed | ISI | CAS |
  60. Anderson SA, Classey JD, Conde F, Lund JS, Lewis DA. Synchonous development of pyramidal neuron dendritic spines and parvalbumin-immunoreactive chandelier neuron axon terminals in layer II of monkey prefrontal cortex. Neuroscience 1995; 67: 7–22. | Article | PubMed | ISI | CAS |
  61. Woodruff AR, Anderson SA, Yuste R. The enigmatic function of chandelier cells. Front Neurosci 2010; 4: 201. | Article | PubMed |
  62. Gorelova N, Seamans JK, Yang CR. Mechanisms of dopamine activation of fast-spiking interneurons that exert inhibition in rat prefrontal cortex. J Neurophysiol 2002; 88: 3150–3166. | Article | PubMed | CAS |
  63. Tseng KY, O’Donnell P. D2 dopamine receptors recruit a GABA component for their attenuation of excitatory synaptic transmission in adult rat prefrontal cortex. Synapse 2007; 61: 843–850. | Article | PubMed | CAS |
  64. Tseng KY, O’Donnell P. Dopamine modulation of prefrontal cortical interneurons changes during adolescence. Cereb Cortex 2007; 17: 1235–1240. | Article | PubMed |
  65. Rao SG, Williams GV, Goldman-Rakic PS. Isodirectional turning of adjacent interneurons and pyramidal cells during working memory: Evidence for microcolumnar organization in PFC. J Neurophsyiol 1999; 81: 1903–1916.
  66. Constantinidis C, Williams GV, Goldman-Rakic PS. A role for inhibition in shaping the temporal flow of information in the prefrontal cortex. Nat Neurosci 2002; 5: 175–180. | Article | PubMed | ISI | CAS |
  67. Cho RY, Konecky RO, Carter CS. Impairments in frontal cortical γ synchrony and cognitive control in schizophrenia. Proc Natl Acad Sci USA 2006; 103: 19878–19883. | Article | PubMed | CAS |
  68. Fries P. Neuronal gamma-band synchronization as a fundamental process in cortical computation. Ann Rev Neurosci 2009; 32: 209–224. | Article | PubMed |
  69. Lewis DA, Hashimoto T, Volk DW. Cortical inhibitory neurons and schizophrenia. Nat Rev Neurosci 2005; 6: 312–324. | Article | PubMed | ISI | CAS |
  70. O’Donnell P. Adolescent onset of cortical disinhibition in schizophrenia: Insights from animal models. Schizophr bull 2011; 37: 484–492. | Article | PubMed |
  71. Wu G, Malinow R, Cline HT. Maturation of a central glutamatergic synapse.Science 1996; 274: 972–976. | Article | PubMed | ISI | CAS |
  72. Durand GM, Koyalchuk Y, Konnerth A. Long-term potentitation and functional synapse induction in developing hippocampus. Nature 1996; 381: 71–75. | Article | PubMed | ISI | CAS |
  73. Isaac JT, Crair MC, Nicoll RA, Malenka RC. Silent synapses during development of thalamocoritcal inputs. Neuron 1997; 18: 269–280. | Article | PubMed | ISI | CAS |
  74. Zhu JJ, Esteban JA, Hayashi Y, Malinow R. Postnatal synaptic potentiation: Delivery of GluR4-contatining AMPA receptors by spontaneous activity. Nat Neurosci 2000; 3: 1098–1106. | Article | PubMed | ISI | CAS |
  75. Liao D, Hessler NA, Malinow R. Activation of postsynaptically silent synapses during pairing-induced LTP in CA1 region of hippocampal slice. Nature 1995;375: 400–404. | Article | PubMed | ISI | CAS |
  76. Dudek SM, Bear MF. Long-term depression in area CA1 of hippocampus and effects of N-methy-D-aspartate receptor blockade. Proc Natl Acad Sci USA1992; 89: 4363–4367. | Article | PubMed | CAS |
  77. Barria A, Muller D, Derkach V, Griffith LC, Soderling TR. Regulatory phosphorylation of AMPA-type glutamate receptors by CaM-Kii during long term potentiation. Science 1997; 276: 2042–2045. | Article | PubMed | ISI | CAS |
  78. Leonard AS, Lim IA, Hemsworth DE, Horne MC, Hell JW. Calcium/calmodulin-dependent protein kinase II is associated with the N-methyl-D-aspartate receptor. Proc Natl Acad Sci USA 1999; 96: 3239–3244. | Article | PubMed | CAS |
  79. Lee HK, Kameyama K, Huganir RL, Bear MF. NMDA induces long-term synaptic depression and dephosphorylation of the GluR1 subunit of AMPA receptors in hippocampus. Neuron 1998; 21: 1151–1162. | Article | PubMed | ISI | CAS |
  80. Lee HK, Barbarosie M, Kameyama K, Bear MF, Huganir RL. Regulation of distinct AMPA receptor phosphorylation sites during bidirectional synaptic plasticity. Nature 2000; 405: 955–959. | Article | PubMed | ISI | CAS |
  81. Malenka RC. Synaptic plasticity in the hippocampus: LTP and LTD. Cell1994; 79: 535–538. | Article | PubMed | ISI |
  82. Malinow R, Malenka RC. AMPA receptor trafficking and synaptic plasticity.Ann Rev Neurosci 2002; 25: 103–126. | Article | PubMed | CAS |
  83. Maletic-Savatic M, Malinow R, Svoboda K. Rapid dendritic morphogenesis in CA1 hippocampal dendrites induced by synaptic activity. Science 1999;283: 1923–1927. | Article | PubMed | ISI | CAS |
  84. Engert F, Bonhoeffer T. Dendritic spine changes associated with hippocampal synaptic plasticity. Nature 1999; 399: 66–70. | Article | PubMed | ISI | CAS |
  85. Toni N, Buchs PA, Nikonenko I, Bron CR, Muller D. LTP promotes formation of multiple spine synapses between a single axon terminal and a dendrite.Nature 1999; 402: 421–425. | Article | PubMed | ISI | CAS |
  86. Matsuzaki M, Honkura N, Ellis-Davies GC, Kasai H. Structural basis of long-term potentiation in single dendritic spines. Nature 2004; 429: 761–766. | Article | PubMed | ISI | CAS |
  87. Nagerl UV, Eberhorn N, Cambridge SB, Bonhoeffer T. Bidirectional activity-dependent morphological plasticity in hippocampal neurons. Neuron 2004;44: 759–767. | Article | PubMed | ISI |
  88. Zhou Q, Homma KJ, Poo MM. Shrinkage of dendritic spines associated with long-term depression of hippocampal synapses. Neuron 2004; 44: 749–757. | Article | PubMed | ISI | CAS |
  89. Tominaga-Yoshino K, Kondo S, Tamotsu S, Ogura A. Repetitive activation of protein kinase A induces slow and persistent potentiation associated with synaptogenesis in cultured hippocampus. Neurosci Res 2002; 44: 357–367. | Article | PubMed |
  90. Shinoda Y, Kamikubo Y, Egashira Y, Tominaga-Yoshino K, Ogura A. Repetition of mGluR-dependent LTD causes slowly developing persistent reduction in synaptic strength accompanied by synaptic elimination. Brain Res 2005; 1042: 99–107. | Article | PubMed |
  91. Kamikubo Y, Egashira Y, Tanaka T, Shinoda Y, Tominaga-Yoshino K, Ogura A. Long-lasting synaptic loss after repeated induction of LTD: independence to the means of LTD induction. Eur J Neurosci 2006; 24: 1606–1616. | Article | PubMed | ISI |
  92. Bastrikova N, Gardner GA, Reece JM, Jeromin A, Dudek SM. Synapse elimination accompanies functional plasticity in hippocampal neurons. Proc Natl Acad Sci USA 2008; 105: 3123–3127. | Article | PubMed |
  93. Crair MC, Malenka. A critical period for long-term potentiation at thalamocortical synapses. Nature 1995; 375: 325–328. | Article | PubMed | ISI | CAS |
  94. Allen CB, Celikel T, Feldman DE. Long-term depression induced by sensory deprivation during cortical map plasticity in vivo. Nat Neurosci 2003; 6: 291–299. | Article | PubMed | ISI | CAS |
  95. Lu H-C, She W-C, Plas DT, Neumann PE, Janz R, Crair MC. Adenylyl chclase I regulates AMPA receptor trafficking during mouse cortical ‘barrel’ map development. Nat Neurosci 2003; 6: 939–947. | Article | PubMed | ISI | CAS |
  96. Yoon B-J, Smith GB, Heynen AJ, Neve RL, Bear MF. Essential role for a long-term depression mechanism in ocular dominance plasticity. Proc Nat Acad Sci USA 2009; 106: 9860–9865. | Article | PubMed |
  97. Bear MF. Synaptic plasticity: From theory to reality. Philos Trans Biol Sci2003; 358: 649–655. | Article |
  98. Kato N, Artola A, Singer W. Developmental changes in the susceptibility to long-term potentiation of neurons in rat visual cortex slices. Develop Brain Res 1991; 60: 43–50. | Article |
  99. Flint AC, Maisch Us, Weishaupt JH, Kriegstein AR, Monver H. NR2A subunit expression shortens NMDA receptor synaptic currents in developing neocortex. J Neurosci 1997; 17: 2469–2476. | PubMed | ISI | CAS |
  100. Roberts EB, Romoa AS. Enhanced NR2A subunit expression and decreased NMDA receptor decay time at the onset of ocular dominance plasticity in the ferret. J Neurophysiol 1999; 81: 2587–2591. | PubMed | ISI | CAS |
  101. Sheng M, Cummings J, Roldan LA, Jan YN, Jan LY. Changing subunit composition of heteromeric NMDA receptors during development of rat cortex. Nature 1994; 368: 144–147. | Article | PubMed | ISI | CAS |
  102. Carmignoto G, Vicini S. Activity-dependent decrease in NMDA receptor responses during development of the visual cortex. Science 1992; 258: 1007–1011. | Article | PubMed | ISI | CAS |
  103. Fox K, Daw N, Sato H, Czepita D. The effect of visual experience on development of NMDA receptor synaptic transmission in kitten visual cortex.J Neurosci 1992; 12: 2672–2684. | PubMed | ISI | CAS |
  104. Nase G, Weishaupt J, Stern P, Singer W, Monver H. Genetic and epigenetic regulation of NMDA receptor expression in the rat visual cortex. Eur J Neurosci 1999; 11: 4320–4326. | Article | PubMed | ISI | CAS |
  105. Quinlan EM, Olstein DH, Bear MF. Bidirectional, experience-dependent regulation of N-methyl-D-aspartate receptor subunit composition in the rat visual cortex during postnatal development. Proc Natl Acad Sci USA 1999;96: 12876–12880. | Article | PubMed | CAS |
  106. Erisir A, Harris JL. Decline of the critical period of visual plasticity is concurrent with the reduction of NR2B subunit of the synaptic NMDA receptor in layer 4. J Neurosci 2003; 23: 5208–5218. | PubMed | ISI | CAS |
  107. Fox K, Sato H, Daw N. The location and function of NMDA receptors in cat and kitten visual cortex. J Neurosci 1989; 9: 2443–2454. | PubMed | ISI | CAS |
  108. Gambrill AC, Barria A. NMDA receptor subunit composition controls synaptogenesis and synapse stabilization. Proc Natl Acad Sci USA 2011;108: 5855–5860. | Article | PubMed |
  109. Barria A, Muller D, Derkach V, Griffith LC, Soderling TR. Regulatory phosphorylation of AMPA-type glutamate receptors by CaM-KII during long-term potentiation. Science 1997; 276: 2042–2045. | Article | PubMed | ISI | CAS |
  110. Leonard AS, Lim IA, Hemworth DE, Horne MC, Hell JW. Calcium/calmodulin-dependent protein kinase II is associated with the N-methyl-D-aspartate receptor. Proc Natl Acad Sci USA 1999; 96: 3239–3244. | Article | PubMed | CAS |
  111. Strack S, Colbran RJ. Autophosphorylation-dependent targeting of calcium/calmodulin-dependent protein kinase II by the NR2B subunit of the N-methyl-D-aspartate receptor. J Biol Chem 1998; 273: 20689–20692. | Article | PubMed | ISI | CAS |
  112. Gardoni F, Schrama LH, van Dalen JJ, Gispen WH, Cattabeni F, Di Luca M. AlphaCaMKII binding to the C-terminal tail of NMDA receptor subunit NR2A and its modulation by autophosphorylation. FEBS Lett 1999; 456: 394–398. | Article | PubMed | ISI | CAS |
  113. Barria A, Malinow R. NMDA receptor subunit composition controls synaptic plasticity by regulating binding to CaMKII. Neuron 2005; 48: 289–301. | Article | PubMed | ISI | CAS |
  114. Zhao JP, Constantine-Paton M. NR2A-/- mice lack long-term potentiation but retain NMDA receptor and L-type Ca2+ channel-dependent long-term depression in the juvenile superior colliculus. J Neurosci 2007; 27: 13649–13654. | Article | PubMed |
  115. Foster KA, McLaughlin N, Edbauer D, Phillips M, Bolton A, Constantine-Paton M et al. Distinct roles of NR2A and NR2B cytoplasmic tails in long-term potentiation. J Neurosci 2010; 30: 2676–2685. | Article | PubMed | CAS |
  116. Barria A, Malinow R. Subunit-specific NMDA receptor trafficking to synapses.Neuron 2002; 35: 345–353. | Article | PubMed | ISI | CAS |
  117. Abraham WC, Bear MF. Metaplasticity: the plasticity of synaptic plasticity.Trends Neurosci 1996; 19: 126–130. | Article | PubMed | ISI | CAS |
  118. Philpot BD, Cho KK, Bear MF. Obligatory role of NR2A for metaplasticity in visual cortex. Neuron 2007; 53: 495–502. | Article | PubMed | ISI | CAS |
  119. Yashiro K, Philpot BD. Regulation of NMDA receptor subunit expression and its implications for LTD, LTP, and metaplasticity. Neuropharmacol 2008; 55: 1081–1094. | Article |
  120. Bear MF, Kirkwood A. Neocortical long-term potentiation. Curr Opin Neurobiol 1993; 3: 197–202. | Article | PubMed | CAS |
  121. Herry C, Vouimba R-M, Carcia R. Plasticity in the mediodorsal thalamo-prefrontal cortical transmission in behaving mice. J Neurophysiol 1999; 82: 2827–2832. | PubMed | ISI | CAS |
  122. Gurden H, Tassin JP, Jay TM. Integrity of the mesocortical dopaminergic system is necessary for complete expression of in vivo hippocampal-prefrontal cortex long-term potentiation. Neuroscience 1999; 94: 1019–1027. | Article | PubMed | ISI | CAS |
  123. Bueno-Junior LS, Lopes-Agular C, Ruggiero RN, Romcy-Pereira RN, Leite JP. Muscarinic and nicotinic modulation of thalamo-prefrontal cortex synaptic plasticity in vivo. PLoS One 2012; 7: e47484. | Article | PubMed |
  124. Bellone C, Luscher C, Mameli M. Mechanisms of synaptic depression triggered by metabotropic glutamate receptors. Cell Mol Life Sci 2008; 65: 2913–2923. | Article | PubMed | ISI | CAS |
  125. Gladding CM, Fitzjohn SM, Molnar E. Metabotropic glutamate receptor-mediated long-term depression: Molecular mechanisms. Pharmacol Rev2009; 61: 395–412. | Article | PubMed | ISI |
  126. Luscher C, Huber KM. Group 1 mGluR-dependent synaptic long-term depression: Mechanisms and implications for circuitry and disease. Neuron2010; 65: 445–459. | Article | PubMed | ISI | CAS |
  127. Mateo Z, Porter JT. Group II metabotropic glutamate receptors inhibit glutamate release at thalamocortical synapses in the developing somatosensory cortex. Neuroscience 2007; 146: 1062–1072. | Article | PubMed | ISI | CAS |
  128. Holbro N, Grunditz A, Oertner TG. Differential distribution of endoplasmic reticulcum controls metabotropic signaling and plasticity at hippocampal synapses. Proc Nat Acad Sci USA 2009; 106: 15055–15060. | Article | PubMed |
  129. Dimitriu D, Hao J, Hara Y, Kaufman J, Janssen WGM, Lou W et al. Selective changes in thin spine density and morphology in monkey prefrontal cortex correlate with aging-related cognitive impairment. J Neurosci 2010; 30: 7507–7515. | Article | PubMed | ISI | CAS |
  130. Andersen P. Assessment and development of executive function (EF) during childhood. Child Neuropsychol 2002; 8: 71–82. | Article | PubMed | ISI |
  131. Shaw P, Greenstein D, Lerch J, Clasen L, Lenroot R, Gogtay N et al. Intellectual ability and cortical development in children and adolescents.Nature 2006; 440: 676–679. | Article | PubMed | ISI | CAS |
  132. Shaw P, Sharp WS, Morrison M, Eckstrand K, Greenstein DK, Clasen LS et al. Psychostimulant treatment and the developing cortex in attention deficit hyperactivity disorder. Am J Psychiatry 2009; 166: 58–63. | Article | PubMed | ISI |
  133. Vazdarjanova A, Bunting K, Muthusamy N, Bergson C. Calcyon upregulation in adolescence impairs response inhibition and working memory in adulthood.Molec Psychiatry 2011; 16: 672–684. | Article |
  134. Selemon LD. Regionally diverse cortical pathology in schizophrenia: Clues to the etiology of the disease. Schizophr Bull 2001; 27: 349–377. | Article | PubMed | CAS |
  135. Koh P, Bergson C, Undie A, Goldman-Rakic PS, Lidow M. Up-regulation of the D1 dopamine receptor-interacting protein, calcyon, in patients with schizophrenia. Arch Gen Psychiatry 2003; 60: 311–319. | Article | PubMed | ISI | CAS |
  136. Bai J, He F, Novikova S, Undie A, Dracheva S, Haroutunian V et al. Abnormalities in the dopamine system in schizophrenia may lie in altered levels of dopamine-interacting proteins. Biol Psychiatry 2004; 56: 427–440. | Article | PubMed | ISI |
  137. Law-Tho D, Desce JM, Crepel F. Dopamine favours the emergence of long-term depression versus potentiation in slices of rat prefrontal cortex.Neurosci Lett 1995; 188: 125–128. | Article | PubMed | CAS |
  138. Snyder GL, Allen PB, Fienberg AA, Valle CG, Huganir RL, Nairn AC et al. Regulation of phosphorylation of the GluR1 AMPA receptor in the neostriatum by dopamine and psychostimulants in vivo. J Neurosci 2000; 20: 4480–4488. | PubMed | ISI | CAS |
  139. Sun X, Zhao Y, Wolf ME. Dopamine receptor stimulation modulates AMPA receptor synaptic insertion in prefrontal cortex neurons. J Neurosci 2005;25: 7342–7351. | Article | PubMed | ISI | CAS |
  140. Selemon LD, Begovic’ A, Goldman-Rakic PS, Castner SA. Amphetamine sensitization alters dendritic morphology in prefrontal cortical pyramidal neurons in the non-human primate. Neuropsychopharmacology 2007; 32: 919–931. | Article | PubMed | CAS |
  141. Selemon LD, Begovic’ A, Williams GV, Castner SA. Reversal of neuronal and cogntivie consequences of amphetamine sensitization following chronic treatment with a D1 antagonist. Pharmacol Biochem Behav 2010; 96: 325–332. | Article | PubMed |
  142. Caruana DA, Warburton EC, Bashir ZI. Induction of activity-dependent LTD requires muscarinic receptor activation in medial prefrontal cortex. J Neurosci 2011; 31: 18464–18478. | Article | PubMed |
  143. Auclair N, Otani S, Soubrie P, Crepel F. Cannabinoids modulate synaptic strength and plasticity at glutamatergic synapses of rat prefrontal cortex pyramidal neurons. J Neurophysiol 2000; 83: 3287–3293. | PubMed | ISI | CAS |
  144. Anthony JC, Petronis KR. Early-onset drug use and risk of later drug problems. Drug Alcohol Depend 1995; 40: 9–15. | Article | PubMed | CAS |
  145. Adriani W, Spiker S, Deroche-Gamonet V, Laviola G, Le Moal M, Smit AB et al. Evidence for enhanced neurobehavioral vulnerability to nicotine during periadolescence in rats. J Neurosci 2003; 23: 4712–4716. | PubMed | ISI | CAS |
  146. Kandel DB. Epidemiological and psychosocial perspectives on adolescent drug use. J Am Acad Child Psychiatry 1982; 21: 4328–4347.
  147. Taioli E, Wynder EL. Effect of the age at which smoking begins on frequency of smoking in adulthood. New Engl J Med 1991; 325: 968–969. | Article | PubMed |
  148. Chambers RA, Taylor JR, Potenza MN. Developmental neurocircuitry of motivation in adolescence: A critical period of addiction vulnerability. Am J Psychiatry 2003; 160: 1041–1052. | Article | PubMed | ISI |
  149. Kelley AE. Memory and addiction: Shared neural circuity and molecular mechanisms. Neuron 2004; 44: 161–179. | Article | PubMed | ISI | CAS |
  150. Hyman SE. Addiction: A disease of learning and memory. Am J Psychiatry2005; 162: 1414–1422. | Article | PubMed | ISI |
  151. Andersen SL, Stress TeicherMH. sensitive periods and maturational events in adolescent depression. Trends Neurosci 2008; 31: 183–191. | Article | PubMed | ISI | CAS |
  152. Hafner H, Maurer K, Loffler W, Riecher-Rossler A. The influence of age and sex on the onset and early course of schizophrenia. Brit J Psychiatry 1993;162: 80–86. | Article |
  153. Pfefferbaum A, Sullivan EV, Rosenbloom MJ, Mathalon DH, Lim KO. A controlled study of cortical gray matter and ventricular changes in alcoholic men over a 5-year interval. Arch Gen Psychiatry 1998; 55: 905–912. | Article | PubMed | CAS |
  154. Miguel-Hidalgo JJ, Overholser JC, Meltzer HY, Stockmeier CA, Rajkowska G. Reduced glial and neuronal packing density in the orbitofrontal cortex in alcohol dependence and its relationship with suicide and duration of alcohol dependence. Alcohol Clin Exp Res 2006; 30: 1845–1855. | Article | PubMed |
  155. Harper C, Kril J. Patterns of neuronal loss in the cerebral cortex in chronic alcoholic patients. J Neurol Sciences 1989; 92: 81–89. | Article |
  156. Acheson SK, Stein RM, Swartzwelder HS. Impairment of semantic and figural memory by acute ethanol: Age-dependent effects. Alcohol Clin Exp Res1998; 22: 1437–1442. | Article | PubMed |
  157. Markwiese BJ, Acheson SK, Levin ED, Wilson WA, Swartzwelder HS. Differential effects of ethanol on memory in adolescent and adult rats.Alcohol Clin Exp Res 1998; 22: 416–421. | Article | PubMed |
  158. Crews FT, Braun CJ, Hoplight B, Switzer RC, Knapp DJ. Binge ethanol consumption causes differential brain damage in young adolescent rats compared with adult rats. Alcohol Clin Exp Res 2000; 24: 1712–1723. | Article | PubMed |
  159. Fadda F, Rossetti ZL. Chronic ethanol consumption: From neuroadaptation to neurodegeneration. Progr Neurobiol 1998; 56: 385–431. | Article | PubMed |
  160. Carpenter-Hyland EP, Chandler LJ. Adaptive plasticity of NMDA receptors and dendritic spines: Implications for enhanced vulnerability of the adolescent brain to alcohol addiction. Pharmacol Biochem Behav 2007; 86: 200–208. | Article | PubMed | CAS |
  161. Swartzwelder HS, Wilson WA, Tayyeb MI. Differential sensitivity of NMDA receptor-mediated synaptic potentials to ethanol in immature versus mature hippocampus. Alcohol Clin Exp Res 1995; 19: 320–323. | Article | PubMed |
  162. Swartzwelder HS, Wilson WA, Tayyeb MI. Age-dependent inhibition of long-term potentiation by ethanol in immature versus mature hippocampus.Alcohol Clin Exp Res 1995; 19: 1480–1484. | Article | PubMed |
  163. Turrigiano GG, Nelson SB. Homeostatic plasticity in the developing nervous system. Nat Rev Neurosci 2004; 5: 97–107. | Article | PubMed | ISI | CAS |
  164. Arnsten AFT, Shansky RM. Adolescence: Vulnerable period for stress-induced prefrontal cortical function? Ann NY Acad Sci 2004; 1021: 143–147. | Article | PubMed |
  165. Andersen LS, Tomada A, Vincow ES, Valente E, Polcari A, Teicher MH. Preliminary evidence for sensitive periods in the effect of childhood sexual abuse on regional brain development. J Neuropsychiatry Clin Neurosci 2008;20: 292–301. | Article | PubMed | ISI |
  166. Rosenberg DR, Lewis DA. Postnatal maturation of the dopaminergic innervation of monkey prefrontal and motor cortices: A tyrosine hydroxylase immunohistochemical analysis. J Comp Neurol 1995; 358: 383–400. | Article | PubMed | ISI | CAS |
  167. Lidow MS, Rakic P. Scheduling of monoaminergic neurostransmitter receptor expression in the primate neocortex during postnatal development. Cereb Corex 1992; 2: 401–416. | Article |
  168. Brenhouse HC, Sonntag KC, Andersen SL. Transient D1 dopamine receptor expression on prefrontal cortex projection neurons: Relationship to enhanced motivational salience of drug cues in adolescence. J Neurosci2008; 28: 2375–2382. | Article | PubMed | ISI |
  169. Bassell GJ, Warren ST. Fragile X syndrome: Loss of local mRNA regulation alters synaptic development and function. Neuron 2008; 60: 201–214. | Article | PubMed | ISI | CAS |
  170. Berry-Kravis E, Sumis A, Hervey C, Nelson M, Porges SW, Weng N et al. Open-label treatment trial of lithium to target the underlying defect in fragile X syndrome. J Dev Behav Pediatr 2008; 29: 293–302. | Article | PubMed | ISI |
  171. Berry-Kravis E, Hessl D, Coffey S, Hervey C, Schneider A, Yuhas J et al. A pilot open label, single dose trial of fenobam in adults with fragile X syndrome. J Med Genet 2009; 46: 266–271. | Article | PubMed | ISI | CAS |
Top

Acknowledgements

I thank Dr Keith Young for his pre-submission reading of this manuscript and helpful comments.


Artigo original:

http://www.nature.com/tp/journal/v3/n3/full/tp20137a.html

O portal Psiquiatria Infantil.com.br já recebeu

47.012.202 visitantes