Friday, July 8, 2016

The Whiteness of Whiteness or White Supremacy as White lunacy or why you think you sane when you crazy (denial)


When I announced the title of my recovery manual How to Recover from the Addiction to White Supremacy, people told me not to entitle the book as such but call it How to Recover from White Lunacy because lunacy is deeper than supremacy. Supposedly, supremacy suggests a certain level of sanity but according to my advisers and the elders white people are simply insane and supremacy is only a part of their insanity. Elijah Muhammad asked us why don't white people want us to have social equality with them? Answer: because if we have social equality with them we will discover how nasty and filthy they are. Was Elijah right? Yes! But Jesus told us even more about them. He said they are liars and murderers who abide not in the truth. He said if God were your Father you would love me but you seek to kill me because I tell you the truth. Has not America killed her truth tellers, e.g., JFK, RFK, MLK, MX, et al.?

But let's get to the whiteness of whiteness! What do I mean? I mean these people are so absorbed in their mythology of whiteness they don't have a clue how damaging their whiteness is to themselves and others, therefore they are qualified for admission to the mental institution. They condemn any and everything Black people express in our attempt to communicate with them our feelings of humanity. We say Black Lives Matter and they know Black lives don't matter to them for they treat their dogs better than they treat us, so BLM means nothing to them thus they ridicule it and wish we would dispense the idea that our lives matter, for all lives matter in their sick minds while the news shows them nightly how little Black lives matter when pigs kill our men in front of our women and children, and often kill our children and women. And then they are bewildered why we cry Black Lives Matter. Of course, Baldwin said they live in an airless room and furthermore the idea of white supremacy has led them to rationalizations so fantastic it approaches the pathological!

They kill us at the drop of a hat, and then when we kill them they are utterly astounded that we would have the nerve to harm precious white flesh, and yet they will flip the subject to why we kill each other as if they haven't taught us to hate each other in the same manner they hate us. Every institution in American society suggests we must hate ourselves and love them. Every image of a woman is that of a white woman or a almost white Black woman. One need only look at the Rap videos to see how often a Black skinned woman is projected as the object of beauty, even though they know and we know the Black woman's body is the true standard of beauty since the beginning of time.

So our women are brainwashed to hate their black skin from Africa to Jamaica--bleaching cream is imported into Africa by the tons so women can emulate the European standard of beauty, and the same is true in India and China. Yes, White lunacy is global and the addiction to white lunacy is global. As is taught in drug and alcohol recovery, addiction is cunning and vile. One can have a relapse at the drop of a hat, the slip of a tongue can reveal the residue of white lunacy even while the addict claims recovery.

The solution to recovery from the addiction to the whiteness of whiteness or white lunacy is long term recovery to engender neuroplasticity that will allow the brain cells to change due to a new environment.

The more we try to make them part of the human family, the more they reveal themselves as part of the animal family of beasts and predators of the worse jungle variety. They want us to stop killing each other but who taught us to kill? They did and as we speak they are killing around the world for nothing. They cannot tell us why they are still killing in Iraq, Afghanistan, Syria, Yemen, Somalia, etc.

For that matter, can they tell us why they kill a human being for selling single cigarettes, CDs, DVDs,
a defective tail light, signal light, for going to the store for a soda, for being mentally ill? Why would you kill the mentally ill? It is because you are addicted to murder under the color of law. We cannot have a trillion dollar military budget to kill around the world, including a president who checks off a murder list weekly, then expect no blow back? What did James Baldwin tell you, "The murder of my child will not make your child safe!"
 --Marvin X, How to Recover from the Addiction to White Supremacy, Black Bird Press, Berkeley CA.

 

 Neuroplasticity

NEUROPLASTICITY

Information in the brain is transmitted from neuron to neuron through specialized connections called synapses. A synapse between two neurons is made up of presynaptic and postsynaptic terminals, which are separated by a synaptic cleft. The presynaptic terminal is filled with small vesicles containing chemical neurotransmitters, and the postsynaptic terminal consists of receptors specific for these neurochemicals. Neurons carry information in the form of an electrical impulse called an action potential that is initiated at the cell body and travels down the axon. At the synapse, an action potential causes the voltage-dependent release of neurotransmitter-filled vesicles, thereby converting an electrical impulse into a chemical signal. Neurotransmitters diffuse across the synaptic cleft, where they bind to receptors and generate an electrical signal in the postsynaptic neuron. The postsynaptic cell will then, in turn, fire an action potential if the sum of all its synapses reaches an electrical threshold for firing. Since a neuron can receive synapses from many different presynaptic cells, each cell is able to integrate information from varied sources before passing along the information in the form of an electrical code. The ability of neurons to modify the strength of existing synapses, as well as form new synaptic connections, is called neuroplasticity. It is believed that neuroplasticity may be the underlying cellular mechanism for the brain's ability to encode information during learning. In theory, this is how information is stored as memory.
Defined in this way, neuroplasticity includes changes in strength of mature synaptic connections, as well as the formation and elimination of synapses in adult and developing brains. This encompasses a vast field of research, and similar processes may also occur at peripheral synapses, where much of the pioneering studies on synaptic transmission first took place. In addition, neuroplasticity includes the regrowth (or sprouting) of new synaptic connections following central nervous system injury; following stroke, for example.
The notion that the brain can store information by modifying synaptic connections is not a new one. In fact, Santiago Ramon y Cajal (a founder of modern neuroscience) expressed this theory in 1894, three years before Charles Sherrington coined the term synapse to describe the connections made between neurons. In the late 1940s the neuroplasticity model was advanced by Jerzy Konorski, who used the word plasticity to describe "permanent functional transformations," and Donald Hebb, who ascribed testable physiologic characteristics to synaptic plasticity. However, experimental evidence that synapses are capable of long-lasting changes in synaptic strength did not come until the early 1970s, when Timothy Bliss and Terry Lomo described an increase in the synaptic strength of neurons in the mammalian hippocampus (a region of the brain critical for some forms of memory) following electrical stimulation. They termed this increase long-lasting potentiation, now referred to as long-term potentiation (LTP).
Changes in synaptic strength proved to be bidirectionally modifiable (they increase and decrease in strength) as Serena Dudek and Mark Bear first demonstrated in 1992 by recording activity-driven, long-term depression (LTD) in the hippocampus. The evidence that learning and memory are based on these long-lasting changes in synaptic strength is substantial, but still incomplete. However, defining the molecular constituents in the mechanistic pathway leading from synaptic activity to plasticity continues to strengthen the evidence linking neuroplasticity with learning and memory. In addition, resolving the molecular mechanisms underlying synaptic modification should lead to targets for clinical intervention in eliminating age-related memory loss or synaptic loss following brain damage by enhancing new synaptic connections.

Mechanisms of plasticity

Synaptic plasticity can occur at either the presynaptic or postsynaptic terminal. Modifications to the presynaptic terminal affect the release of neurotransmitters. As the action potential invades the presynaptic terminal, it activates voltage-gated calcium channels that conduct calcium ions into the presynaptic terminal. This rise in intracellular calcium triggers the exocytosis of vesicles (fusion with the plasma membrane) and thus the release of neurotransmitters. Each presynaptic terminal contains between 200 and 500 vesicles, though only a small proportion of these are ready to be released at any time. Vesicles in the presynaptic terminal move through a specific release cycle, including vesicle storage, priming for release, release, vesicle reformation, and reloading with neurotransmitter.
Factors that alter the presynapse resulting in either modification of the calcium channel conductance or modification of the vesicle cycle will yield changes in synaptic strength. One such factor is the cyclic nucleotide cAMP. An increase in cAMP presynaptically can enhance transmitter release by activating protein kinase A (PKA). PKA activation induces a decrease in a specific potassium channel conductance called a delayed rectifier current. Decreased delayed rectifier conductance will increase the calcium entry into the presynaptic terminal by increasing the duration of the action potential. In addition, a rise in cAMP can activate vesicular release from presynaptic terminals that were previously dormant. Such terminals are present, but do not release neurotransmitters in response to an action potential prior to a rise in cAMP. A morphologically distinct synapse that is physiologically dormant has been termed a silent synapse and can be the result of deficient presynaptic release, or a deficiency of transmitter receptors expressed postsynaptically.

The postsynaptic terminal can also be modified to produce changes in synaptic efficacy. Signaling molecules in the postsynaptic compartment such as protein kinase A (PKA) and the alpha subunit of calcium/calmodulin-dependent kinase II (α-CaMKII) are thought to play major roles in synaptic plasticity. For example, when a mouse is genetically altered to express a version of α-CaMKII incapable of activation, LTP and learning are disrupted. While α-CaMKII can directly phosphorylate neurotransmitter receptors leading to an increase in conductance, it is likely to play additional roles in synaptic plasticity as well. Neurotransmitter receptors can cycle in and out of the postsynaptic membrane (in a process not unlike the presynaptic vesicles), and α-CaMKII phosphorylation of an as yet unidentified substrate could lead to the rapid insertion of more receptors. This would result in LTP of an active synapse and the unsilencing of a synapse that was not previously expressing these receptors in its membrane. As stated above, there is substantial evidence implicating long-lasting changes in synaptic strength with the formation of memory. It should be noted that synapses do not act in isolation. The neural circuits to which they belong are a result of the many thousands of synapses contained therein. Although the cellular coding of information may be encoded at synapses, memory itself is likely dependent upon the circuit(s) in which they are contained.

Plasticity, memory, and aging

As humans age, an impairment of memory occurs that is not associated with neurological damage or disease. The age of onset for this decline varies, but it is clear that this is a selective deficit and not a generalized decrease in cognitive skills. Moreover, the deficit is also apparent in animal models of aging and is manifest as a greater number of trials required to memorize a task and a decrease in memory retention that begins approximately twenty-four hours post-training. Interestingly, LTP also changes with age, typically requiring a more robust stimulus to induce and yielding a synaptic potentiation that decays more rapidly. Since aging animals and humans both maintain the ability to store memory, the fundamental mechanisms that underlie information storage may remain essentially intact. The deficit may not be a lack of ability, but rather a decline in the efficiency of storageor an inability to maintain the neural plasticity induced during learning. Since the formation of memory is dependent on new protein synthesis, one way to address the decreased stability of memory is to identify proteins made during learning. Consistent with this, synaptic plasticity has at least two temporally distinct components: transient changes that do not require new protein synthesis, and enduring modifications (e.g., LTP and LTD) that require the production of new proteins. Identification of newly formed proteins, their site of action, and the molecular basis for their role in neural plasticity may provide insights into the maintenance of memory, and thus indicate clinical targets for the amelioration of age-related memory decline.
David G. Wells
See also Brain; Learning; Memory; Neurochemistry; Neurodegenative Diseases.

BIBLIOGRAPHY

Bliss, T. V. P., and Lomo, T. "Long-Lasting Potentiation of Synaptic Transmission in the Dentate Area of the Anaesthetized Rabbit Following Stimulation of the Perforant Path." Journal of Physiology (London) 232 (1973): 331356.
Cowen, W. M., and Kandel, E. R. "A Brief History of Synapses and Synaptic Transmission." In Synapses. Edited by W. M. Cowen, T. C. Sudhof and C. F. Stevens, Baltimore, Md.: The Johns Hopkins University Press, 2001. Pages 188.
Davis, H. P., and Squire, L. R. "Protein Synthesis and Memory: A Review." Psychology Bulletin 96 (1984): 518559.
Dudek, S. M., and Bear, M. F. "Homosynaptic Long-Term Depression in Area CA1 of Hipocampus and Effects of N-methyl-D-aspartate Receptor Blockade." Proceedings of the National Academy of Science 89 (1992): 43634367.
Foster, T. C. "Involvement of Hippocampal Synaptic Plasticity in Age-Related Memory Decline." Brain Research Review 30 (1999): 236249.
Giese, K. P.; Fedorov, N. B.; Filipkowski, R. K.; and Silva, A. J. "Autophosphorylation at Thr286 of the Alpha Calcium-Calmodulin Kinase II in LTP and Learning." Science 279 (1998): 870873.
Hayashi, Y.; Shi, S.-H.; Esteban, J. A.; Piccini, A.; Poncer, J. C.; and Malinow, R. "Driving AMPA Receptors into Synapses by LTP and CaMKII: Requirements for GluR1 and PDZ Domain Interactions." Science 287 (2000): 22622267.
Ma, L.; Zablow, L.; Kandel, E. R.; and Siegelbaum, S. A. "Cyclic AMP Induces Functional Presynaptic Boutons in Hippocampal CA3-CA1 Neuronal Cultures." National Neuroscience 2 (1999): 2430.
Tong, G.; Malenka, R. C.; and Nicoll, R. A. "Long-Term Potentiation in Cultures of Single Hippocampal Granule Cells: A Presynaptic Form of Plasticity." Neuron 16 (1996): 11471157.



Language: English | Spanish | French

Neural plasticity: consequences of stress and actions of antidepressant treatment

Abstract

Neural plasticity is emerging as a fundamental and critical mechanism of neuronal function, which allows the brain to receive information and make the appropriate adaptive responses to subsequent related stimuli. Elucidation of the molecular and cellular mechanisms underlying neural plasticity is a major goal of neuroscience research, and significant advances have been made in recent years. These mechanisms include regulation of signal transduction and gene expression, and also structural alterations of neuronal spines and processes, and even the birth of new neurons in the adult brain. Altered plasticity could thereby contribute to psychiatric and neurological disorders. This article revievi/s the literature demonstrating altered plasticity in response to stress, and evidence that chronic antidepressant treatment can reverse or block the effects, and even induce neural piasiicity-iike responses. Continued elucidation of the mechanisms underlying neural plasticity will lead to novel drug targets that could prove to be effective and rapidly acting therapeutic interventions.
Keywords: signal transduction, gene expression, neurotrophic factor, neurogenesis, neuronal atrophy
Neural plasticity is a fundamental process that allows the brain to receive information and form appropriate adaptive responses to the same or similar stimuli. The molecular and cellular adaptations underlying learning and memory are the best-characterized and moststudied examples of neural plasticity. However, many different stimuli can activate neural plasticity processes in different brain structures, including environmental, social, behavioral, and pharmacological stimuli. In fact, it could be argued that neural plasticity is one of the most essential and important processes that the brain performs as it relates to many types of central nervous system functions.
Thus, disrupted or abnormal plasticity could lead to maladaptive neuronal responses and abnormal behavior. This could occur in response to genetic abnormalities of the cellular machinery required for plasticity, and abnormal or inappropriate stimuli. For example, exposure to inappropriate or prolonged stress has been reported to alter molecular and cellular markers of neural plasticity, and could contribute to stress-related mood disorders. This review will discuss the literature demonstrating altered neural plasticity in response to stress, and clinical evidence indicating that altered plasticity occurs in depressed patients. The second part of the review will present evidence that antidepressant treatment blocks the effects of stress or produces plasticity -like responses.

General mechanisms of neural plasticity

Neural plasticity encompasses many different types of molecular and cellular responses that occur when cells in the brain are induced to respond to inputs from other cells or circulating factors. The systems that have been most extensively studied are cellular and behavioral models of learning and memory, including long-term potentiation (LTP), in slices of brain and rodent models of behavior. The mechanisms identified for learning and memory most likely also subserve plasticity occurring in other regions and for other adaptive functions of the brain. This section will briefly discuss some general mechanisms and concepts of plasticity.

Mechanisms of acute neural plasticity: synaptic transmission and protein kinases

The effects underlying the rapid responses to neuronal activation are mediated by activation of the excitatory neurotransmitter glutamate and regulation of intracellular signaling cascades (for a review of acute mechanisms underlying LTP, see reference 1). Glutamate causes neuronal depolarization via activation of postsynaptic ionotropic receptors that increase intracellular Na+. This leads to the subsequent activation of /V-mcthyl-D-aspartatc (NMDA) receptors and the resulting influx of Ca2+. Ca2+ is a major intracellular signaling molecule that activates a signaling cascade, including activation of Ca2+/ calmodulin-dependent protein kinase. Within minutes to hours, activation of glutamate and Ca2+-dependent pathways can result in structural alterations at the level of dendritic spines. Spines mark the location of glutamate synapses and have been the subject of intensive investigation for understanding synaptic plasticity.2 Changes in the shape and even number of spines can occur very rapidly (minutes to hours) after glutamate stimulation. These alterations are made permanent or long-term when they arc stabilized or consolidated, a process that requires gene expression and protein synthesis.

Mechanisms of long-term plasticity: gene expression and protein synthesis

The Ca2+/cyclic adenosine monophosphate (cAMP) response element (CaRE) binding protein (CREB) is one of the major transcription factors that mediate the actions of Ca2+, as well as cAMP signaling. CREB has been reported to play a role in both cellular and behavioral models of learning and memory.3 There are a number of gene targets that are influenced by Ca2+, cAMP, and CREB, and the pattern of gene regulation is dependent on the cell type, the length of stimulation, as well as the magnitude of stimulation. Gene targets that have been implicated in learning and memory, and are relevant to the effects of stress and antidepressant treatment, are the neurotrophic factors. Of particular interest is brain-derived neurotrophic factor (BDNF), one of the most abundant neurotrophic factors in the brain.

Altered neural plasticity in response to stress

Recent reports have demonstrated altered molecular and cellular responses to stress and have contributed to the hypothesis that altered neural plasticity contributes to stress-related psychiatric illnesses. Some examples of stress responses are discussed in this section.

Stress alters learning and memory

Stress is known to significantly influence learning and memory, and the effects are dependent on the type, duration, and intensity of the stressor. Emotional arousal can enhance learning and memory via synaptic plasticity of amygdala-dependent pathways, and this is thought to be the basis for intense, long-term memories of traumatic events and posttraumatic stress disorder.4,5 However, stress can also impair subsequent learning and memory and can even lead to amnesia.6 The influence of stress on hippocampal-dependent learning is complex and dependent on the type of learning task.
In studies of LTP, a consistent suppression of neural plasticity is observed after exposure to stress or adrenal glucocorticoids.6,7 In one of these studies, the suppression of LTP was observed after exposure to an uncontrollable stressor and correlated with behavioral performance in a learning and memory task. Giving the animals control over the stress (ie, the stress could be terminated) did not lead to reduced LTP or decreased learning and memory.8 A role for BDNF in the actions of stress on LTP has also been suggested.9 For additional references and discussion of the effects of stress on learning and memory, see the reviews in references 4 to 7.

Stress causes atrophy of hippocainpal neurons

One of the best-characterized examples of altered structural plasticity in response to stress is the atrophy of hippocampal neurons, which was first described by McEwen and colleagues (Figure 1.).10 They found that repeated restraint stress results in atrophy of the dendrites of CA3 pyramidal neurons in the hippocampus, measured as a decrease in the number and length of apical dendrites.11 The reduction in dendritic arborization was found to be dependent on long-term, repeated exposure to restraint stress (3 weeks) and to be reversible when the animals are removed from stress. The atrophy of CA3 pyramidal cells appears to result from the elevation of adrenal glucocorticoids that occurs during stress because chronic administration of corticosteronc, the active form in rodent, results in a similar decrease in number and length of dendrites.12 The actions of stress and glucocorticoids are blocked by administration of an NMDA receptor antagonist, indicating that this glutamate receptor is required for atrophy of CA3 neurons.10 Atrophy of CA3 pyramidal neurons occurs after 2 to 3 weeks of exposure to restraint stress or more long-term social stress, and has been observed in rodents and tree shrews.11-13 In contrast to the atrophy of hippocampus, recent studies demonstrate that chronic stress causes hypertrophy of neurons in the amygdala.14 This study found chronic immobilization stress increased the dendritic arborization of pyramidal neurons in the basolateral amygdala, but decreased dendrite length and branching in the CA3 pyramidal neurons of the hippocampus. Hypertrophy of the amygdala could underlie increased learning and memory as a result of stressinduced emotional arousal, and may be relevant to the pathophysiology of stress-related disorders, including anxiety, posttraumatic stress, and depression. Increased arborization of neurons in the amygdala could thereby enhance emotional states or disrupt normal processing of emotional responses.
Figure 1.
Model of hippocampal plasticity showing structural alterations in response to stress: atropy of CA3 pyramidal neurons and decreased neurogenesis of dentate gyrus granule cells. Stress results in powerful effects on the hippocampus, partly because of the ...

Stress decreases neurogenesis in the adult hippocampus

In addition to regulation of the morphology of neurons in the hippocampus, stress influences the number of newborn neurons or neurogenesis in the adult hippocampus15,16 (Figures 1 and 2.) The hippocampus is one of two brain regions where neurogenesis continues to occur in adult organism (the other region is in the subventricular zone). In the hippocampus, neural progenitor cells are found in the subgranular zone, between the granule cell layer and the hilus. These cells give rise to newborn cells that migrate into the granule cell layer and mature into neurons with the morphological and physiological characteristics of adult granule cells.17 Interestingly, the process of neurogenesis is highly regulated by a variety of stimuli and can be considered a form of neural plasticity. For example, enriched environment, exercise, and learning increase neurogenesis, while aging and exposure to drugs of abuse decrease neurogenesis.15,16,18
Figure 2.
Model demonstrating the regulation of adult neurogenesis in the hippocampus. Neural progenitor cells are restricted to the subgranular zone (SGZ) that is located between the granule cell layer (GCL) and hilus. These progenitor cells give rise to newborn ...
In addition to these factors, stress also results in a dramatic downrcgulation of neurogenesis in the hippocampus.10,18 Exposure to just a single stressor is sufficient to significantly decrease neurogenesis in the adult hippocampus. Adult neurogenesis is decreased by different types of stress, including subordination stress,19 predator odor,20 maternal separation,21 and footshock.22 In addition, exposure to inescapable stress in the learned helplessness model of depression decreases adult neurogenesis and this effect correlates with behavioral despair in this model.22 Moreover, the reduction in neurogenesis and the behavioral despair is reversed by antidepressant treatment.

Regulation of CREB and decreased expression of BDNF in response to stress

Stress results in a wide range of effects that influence many different neurotransmitter and neuropeptide systems, signal transduction pathways, and altered gene expression. The hallmark of the stress response is activation of the hypothalamic-pituitary-adrcnal (HPA) axis, which includes increased circulating levels of adrenal glucocorticoids. The hippocampus contains veryhigh levels of glucocorticoid receptors and is therefore significantly impacted by stress. As mentioned above, studies by McEwen and colleagues have demonstrated that glucocorticoids contribute to the atrophy and decreased neurogenesis of hippocampal neurons resulting from exposure to stress.10
In addition, stress is reported to influence CREB and BDNF in the hippocampus and other brain regions. The transcriptional activity of CREB is regulated by phosphorylation and levels of phospho-CREB are used as an indirect measure of CREB activation and function (Figure 3.) The, regulation of phospho-CREB is complex and is dependent on the brain region and whether the stress is acute or chronic.23-26 Acute stress increases levels of phospho-CREB in many limbic regions associated with mood disorders and this may represent a normal or appropriate adaptive responsiveness.24 In contrast, chronic stress leads to decreased levels of phosphoCREB in many limbic brain regions, which could lead to decreased plasticity and function.26
Figure 3.
Model demonstrating the upregulation of the cyclic adenosine monophosphate (cAMP)-cAMP response element binding protein (CREB) cascade and expression of brain-derived neurotrophic factor (BDNF) by antidepressant treatment. Chronic, but not acute, antidepressant ...
Stress has profound effects on the expression of BDNF in the hippocampus. Levels of BDNF expression in hippocampus are dramatically downregulated by both acute and chronic stress, and this effect could contribute to the atrophy and decreased neurogenesis caused by stress (Figure l).27-29 The role of other factors that could underlie the actions of stress on adult neurogenesis is a subject of interest and could lead to novel targets for drug development.

Atrophy of limbic brain structures in depressed patients

Evidence from basic research studies provide strong support for the hypothesis that stress-related illnesses such as depression could include alterations in brain structure and neural plasticity. Indeed, direct evidence to support this hypothesis has been provided by brain imaging and postmortem studies of depressed patients.

Evidence from brain imaging studies

Magnetic resonance imaging studies have demonstrated that the size of certain brain structures is decreased in mood disorder patients. In particular, these studies demonstrate that the volume of the hippocampus is decreased in patients with depression.30,31 Reduced hip pocampal volume is also observed in patients with posttraumatic stress disorder (PTSD).32 The reduction in hippocampal volume is directly related to the length of illness.33,34 In addition to hippocampus, atrophy of prefrontal cortex and amygdala - brain regions that control cognition, mood, and anxiety - has also been reported in patients with depression or bipolar disorder.35

Evidence from postmortem studies

Atrophy of hippocampus or other brain regions could result from loss of cells (neurons or glia) or decreased size of the cell body or neuronal processes. The most extensive studies have been conducted on prefrontal and cingulatc cortex and demonstrate that the neuronal body size and number of glia is decreased in depressed patients.36-38 There is much less known about the hippocampus and additional studies will be required to determine what accounts for the atrophy of hippocampus observed in depressed patients.
Postmortem analysis of CREB and BDNF has also provided evidence consistent with a loss of neural plasticity in depression. Levels of CREB arc decreased in the cerebral cortex of depressed patients or suicide victims.39,40 Levels of BDNF are also decreased in prefrontal cortex and hippocampus of depressed patients.41 Reduced levels of CREB and BDNF“, two molecular markers of neural plasticity, indicate that the ability of limbic brain structures to mount adaptive responses is compromised in depressed patients.

Antidepressant treatment increases neural plasticity

In contrast to the effects of stress, antidepressant treatment results in molecular and cellular responses that demonstrate an increase in neural plasticity. Moreover, these studies have paved the way for additional studies that demonstrate that antidepressant treatment results in structural remodeling. In many cases, the effects of antidepressant treatment oppose or reverse the effects of stress. Taken together, these findings provide additional support for the hypothesis that neural plasticity plays a significant role in the treatment, as well as the pathophysiology of mood disorders. The evidence for regulation of neural plasticity at the level of neurogenesis, signal transduction, and gene expression is discussed in the second half of this review.

Antidepressant treatment increases adult neurogenesis

Neurogenesis is increased by chronic antidepressant administration

One of the most surprising discoveries of recent times in the field of depression is that antidepressant treatment regulates neurogenesis in the adult hippocampus (Figures 1 and 2). In contrast to the actions of stress, chronic antidepressant treatment increases the number of newborn neurons in the adult hippocampus of rodents or tree shrews.42,43 The upregulation of neurogenesis is dependent on chronic antidepressant treatment, consistent with the time course for the therapeutic action of antidepressants.43 In addition, different classes of antidepressants, including serotonin (5-hydroxytryptamine [5-HT]) and noradrenaline reuptake inhibitors, and electroconvulsive seizures are reported to increase adult neurogenesis.43-45 Antidepressant treatment influences two important aspects of neurogenesis, the rate of cell proliferation (ie, the number of newborn neurons) and the survival of newborn neurons.46 An increase in the number of newborn neurons could contribute to the reversal of hippocampal atrophy observed in depressed patients.

Antidepressant treatment blocks the downregulation of neurogenesis caused by stress

The influence of antidepressant treatment in the context of stress has also been examined. These studies demonstrate that chronic antidepressant treatment can block or reverse the downregulation of neurogenesis that results from exposure to stress. Several different types of stress have been tested, including blockade of intruder stress,42 maternal separation,47 and learned helplessness.22 In addition, different types of antidepressants have been tested, including an atypical antidepressant, tianeptine,42 a selective serotonin reuptake inhibitor (SSRI),22,47 and a neurokinin-1 receptor antagonist.48.
The influence of antidepressant treatment on the atrophy of CA3 pyramidal neurons resulting from chronic exposure to stress has been examined. These studies demonstrate that chronic administration of tianeptine blocks the atrophy of CA3 apical dendrites that is caused by stress.12 Chronic administration of an SSRI antidepressant did not block the atrophy of CA3 neurons in this study Analysis of dendrite branch number and length is tedious and labor intensive, but additional studies of other antidepressants are necessary to determine the relevance of this effect in the actions of antidepressant treatment.

A functional role for neurogenesis in the action of antidepressant treatment

A major issue in the field of adult neurogenesis is how to test the function of newborn neurons. A recent study has addressed this question by using a combination of irradiation and mutant mouse approaches.49 This study demonstrates that focused irradiation of hippocampus in the mouse completely blocks neurogenesis and there was a corresponding blockade of the behavioral actions of antidepressant treatment in two behavioral models, novelty suppressed feeding and chronic mild stress. In addition, Santarelli et al49 studied the effects of antidepressants in mice with a null mutation of the 5-HT1A receptor, a subtype that has been implicated in the actions of antidepressant treatment. They found that upregulation of neurogenesis by chronic administration of an SSRI was completely blocked in 5-HT1A null mutant mice, and that the behavioral effects of SSRI treatment were similarly blocked. These results are the first evidence that increased neurogenesis is necessary for an antidepressant response in behavioral models. rFh ere arc a few limitations to this study. First, although novelty-suppressed feeding is responsive to chronic antidepressant treatment - and this is why it was chosen - this paradigm is a better model of anxiety than depression. Second, although the effects of antidepressant treatment were blocked, irradiation and 5-HT1A null mutation alone, in the absence of antidepressant administration, did not produce a depressive phenotype. This is consistent with another report demonstrating that decreased neurogenesis is not correlated with behavior in the learned helplessness model of depression.50 Together these studies indicate that neurogenesis is not required for baseline response. However, it is possible that intact neurons are sufficient to sustain baseline response and that more long-term inhibition of neurogenesis would be required to influence activity.

The cAMP-CREB cascade and depression

Neural plasticity upon antidepressant treatment is likely to involve adaptations of multiple intracellular signaling cascades and even interactions of these pathways. One of the pathways that is regulated by antidepressant treatment and has been demonstrated to contribute to the actions of chronic antidepressant responses is the cAMP-CREB cascade, the subject of this section. However, it is likely that other signaling pathways are also regulated by - and play a role in - the actions of antidepressants. For reviews covering other signal transduction pathways, see reference 51 and 52.

Antidepressant treatment upregulates the cAMP CREB cascade

Several studies have investigated the influence of antidepressant treatment on the cAM'P-CREB pathway (Figure 3).53,54 This work demonstrates that chronic antidepressant treatment upregulates the cAMP second-messenger cascade at several different levels. This includes increased coupling of the stimulatory G protein to adenylyl cyclase, increased levels of cAMP-dependent protein kinase (PKA), and increased levels of CREB as well as phospho-CREB.55-57 Upregulation of these components of the cAMP-CREB signaling pathway is dependent, on chronic antidepressant treatment, consistent with the time course for the therapeutic action of antidepressants. In addition, upregulation of the cAMP-CREB cascade is observed in response to chronic administration of different classes of antidepressants, indicating that this is a common target of antidepressant treatment.
In addition to phosphorylation by PKA, CREB is also phosphorylated by Ca2+-dependent kinases, such as Ca2+/calmodulin-dependent protein kinase, and by mitogen-activated protein kinase pathways (Figure 3). In this way, CREB can serve as a target for multiple signal transduction pathways and neurotransmitter receptors that activate these cascades.

Activation of the cAMP-CREB cascade produces an antidepressant response

Direct, evidence for cAMP-CREB signaling in the action of antidepressant treatment has been tested by pharmacological, viral vector, and mutant mouse approaches. First, drugs that block the breakdown of cAMP produce an antidepressant response in behavioral models of depression.54 The primary target for inhibition of cAMP breakdown is cAMP-specific phosphodiesterase type IV (PDE4), and rolipram was one of the first selective PDE4 inhibitors. In addition, we have found that chronic rolipram administration increases neurogenesis in adult hippocampus.46,58
Second, viral expression of CREB in the hippocampus of rat produces an antidepressant response in the forced swim and learned helplessness models of depression.59 However, further studies demonstrated that the effects of CREB are dependent on the brain region where it is expressed. For example, expression of CREB in the nucleus accumbens produces a prodepressant effect, while expression of a dominant, negative mutant of CREB results in an antidepressant response in the forced swim test.60 Transgenic expression of dominant negative CREB in the nucleus accumbens is consistent with this effect.61 The different behavioral effects of CREB can be explained by different target genes in the hippocampus (ic, BDNF) versus the nucleus accumbens (ie,prodynorphin).

Regulation of neurotrophic factors and depression

The regulation of CREB by antidepressant treatment indicates that regulation of gene expression also plays a role in the actions of antidepressants. There have been many gene targets identified for antidepressants,51,52 but BDNF is one that has gained attention and is relevant to neural plasticity responses to antidepressant medications. Studies to identify additional gene targets and gene profiles using gene microarray analysis are currently being conducted.

Antidepressant treatment upregulates BDNF

Neurotrophic factors were originally identified and studied for their role in development, and neuronal survival. However, it is now clear that these factors are expressed in the adult brain, arc dynamically regulated by neuronal activity, and are critical for the survival and function of adult neurons. On the basis of these considerations, it is clear why decreased expression of BDNF could have serious consequences for the function of limbic brain structures that control mood and cognition. In contrast, antidepressant treatment results in significant upregulation of BDNF in the hippocampus and cerebral cortex of rodents.28,53,54 Increased expression of BDNF is dependent on chronic treatment, and is observed with different classes of antidepressants, but not other psychotropic drugs. The induction of BDNF would be expected to protect neurons from damage resulting from stress, elevated glucocorticoids, or other types of neuronal insult.

BDNF has antidepressant effects in behavioral models of depression

The possibility that BDNF contributes to the actions of antidepressant treatment is supported by behavioral studies of recombinant BDNF and transgenic mouse models. Microinfusions of BDNF into the hippocampus produce an antidepressant-like response in the learned helplessness and forced swim models of depression.62 The antidepressant, effect of BDNF is observed after a single infusion, compared with repeated administration of a. chemical antidepressant, and is relatively long-lasting (up to 10 days after infusion). Transgenic overexpression of a dominant negative mutant of the BDNF receptor, trkB, in the hippocampus and other forebrain structures is also reported to block the effect, of antidepressant treatment, demonstrating that BDNF signaling is necessary for an antidepressant response.63
Microinfusions of BDNF into the dorsal raphe, a midbrain region where 5-HT cell bodies are localized, also produces an antidepressant response in the learned helplessness model.64 Together, these studies indicate that BDNF could contribute to antidepressant responses in both forebrain and brain stem structures by affecting different populations of neurons. Alternatively, it is possible that, microinfusions of BDNF into the hippocampus influence 5-HT neuronal function by acting at presynaptic sites, and could therefore enhance 5-HT signaling as observed after brain stem infusions of BDNF.64

A neurotrophic hypothesis of depression

Basic research and clinical studies of BDNF have resulted in a. neurotrophic hypothesis of depression and antidepressant action.53,54 This hypothesis is based in part. on studies demonstrating that stress decreases BDNF, reduces neurogenesis, and causes atrophy or CA3 pyramidal neurons. Brain imaging and postmortem studies provide additional support, demonstrating atrophy and cell loss of limbic structures, including the hippocampus, prefrontal cortex, and amygdala. In contrast, antidepressant treatment, opposes these effects of stress and depression, increasing levels of BDNF, increasing neurogenesis, and reversing or blocking the atrophy and cell loss caused by stress and depression. Additional brain imaging and postmortem studies, as well as basic research approaches will be required to further test this hypothesis. In any case, the studies to date provide compelling evidence that, neural plasticity is a. critical factor in the pathophysiology and treatment of depression.

Antidepressants influence other neurotrophic factor systems

Because of the preclinical and clinical evidence implicating neurotrophic factors in the pathophysiology and treatment of depression, studies have been conducted to examine other neurotrophic factor systems. One of the most robust effects identified to date is that antidepressant treatment increases the expression of fibroblast. growth factor-2 (FGF-2).65 FGF-2 is known to have a potent influence on neurogenesis during development and in the adult brain, and could contribute to antide pressant regulation of neurogenesis. Studies are under way to examine the role of FGF-2 in antidepressant regulation of neurogenesis and regulation of behavior in models of depression. Several other growth factors have been identified by microarray analysis and gene expression profiling, including vascular endothelial growth factor, neuritin, and VGF.66 Studies are currently under way to determine the functional significance of these growth factors in models of depression.

Clinical evidence of relevance of neural plasticity to antidepressant treatment

Basic research studies clearly demonstrate that antidepressant treatment regulates signal transduction, gene expression, and the cellular responses that, represent neural plasticity. This issue is more difficult, to address in clinical studies, but evidence is slowly accumulating. Brain imaging studies have been conducted to examine the influence of antidepressants on the volume of limbic brain regions. One study demonstrates that hippocampal atrophy is inversely proportional to the length of time a patient receives antidepressant medication.67 A longitudinal study of PTSD patients before and after antidepressant treatment has found that there is a. partial reversal of hippocampal atrophy in patients receiving medication.68 The latter study demonstrated a corresponding increase in verbal declarative memory in response to antidepressant treatment.
Evidence at the molecular level is also provided by postmortem studies. Levels of CREB immunoreactivity are increased in patients receiving antidepressant treatment at the time of death relative to unmedicated patients.39 In addition, levels of BDNF are increased in patients taking an antidepressant at the time of death.59 Although these effects must be replicated and extended (for example, to the regulation of neurogenesis) in additional banks of postmortem tissue, the results are consistent with the hypothesis that neural plasticity is upregulatcd in patients receiving antidepressant medication.

Novel targets for the treatment of depression

The hypothesis that antidepressant treatment increases neural plasticity provides a number of novel targets for drug development. However, as with any fundamentally important mechanism, care must be taken that the drugs developed for such targets do not interfere with the normal function of the brain. Nevertheless, regulation of neural plasticity is an exciting area of research for design of new drugs for a variety of indications, including learning, memory, cognition, mood, and neurodegenerative disorders. This section discusses a few of these targets in the context of the pathways regulated by antidepressants and stress.

Targets for antidepressant regulation of neurogenesis

Identification of the signal transduction and gene expression pathways that are responsible for the actions of antidepressant regulation of neurogenesis is a subject, of intense investigation. Activation of the cAMP-CREB signaling cascade using either pharmacological or transgenic approaches is reported to increase both proliferation and survival of newborn neurons in the hippocampus,46,58 supporting the possibility that antidepressants increase neurogenesis via regulation of this intracellular pathway. Gene targets of CREB, as well as other neurotrophic/growfh factors that, have been shown to regulate adult neurogenesis, include BDNF, FGF-2, and insulin-like growth factor-1 , to name but. a few.18 Because antidepressant treatment increases the expression of both BDNF and FGF-2, these two factors are currently being investigated. This is just a partial listing of the signal transduction cascades and factors that could contribute to antidepressant regulation of adult neurogenesis.

Targets for regulation of the cAMP-CREB cascade

There are several different sites within the cAMP pathway that could be targeted for drug development. One that has already proven to be effective for antidepressant treatment is blockade of PDE4 and the breakdown of cAMP. Rolipram is a PDF'4-selective inhibitor that has been demonstrated to have antidepressant efficacy in early clinical trials and behavioral models of depression.69,70 However, the clinical use of rolipram has been limited by its side effects, primarily nausea.
The identification of four different. PDE4 isozymes that are equally inhibited by rolipram raises the possibility that one of the isozymes underlies the antidepressant actions of rolipram, while another mediates its side effects. Studies are currently under way to characterize the regional distribution and function of the three PDE4 isozymes expressed in brain (PDE4A, PDE4B, and PDE4D) and the role of these isozymes in the actions of antidepressant treatment.71 Studies of mutant mice demonstrate that null mutation of PDE4D produces an antidcpressant-like phenotype indicating a role for this isozyme,72 and similar studies are currently under way for PDE4A and PDE4B.

BDNF as a target for drug development

The use of BDNF and other neurotrophic factors for the treatment of neurological disorders has been a subject of interest, for several years, although problems with delivery, efficacy, and side effects have hampered these efforts. To more directly replicate the in vivo situation, it may be possible to stimulate the expression of endogenous BDNF expression by stimulating signaling pathways known to regulate this neurotrophic factor. First, activation of the cAMP-CREB cascade by inhibition of PDE4 increases the expression of BDNF.56
Small molecular agonists for neurotransmitter receptors have also exhibited some promise. Activation of ionotropic glutamate receptors increases BDNF expression and could be targeted for the treatment of depression.73 One drug that modulates glutamate transmission and increases BDNF expression is memantine.74 Riluzole, a. sodium channel blocker, also increases BDNF expression, as well as neurogenesis in adult hippocampus.75 Specific 5-HT and norepinephrine receptor subtypes that activate cAMP (eg, β-adrenergic, 5-HT7), Ca2+, or mitogen-activated protein kinase (α1-adrenergic, 5-HT1A) pathways could also be targets for development. Characterization of the antidepressant actions of these compounds will be needed, as well as identification of additional neurotransmitter and signal transduction systems that regulate BDNF

Conclusions

Studies of the molecular and cellular mechanisms underlying neural plasticity responses in learning and memory, as well as fear, anxiety, depression, and drug abuse to name but a few, are some of the most exciting and rapidly advancing areas of research in neuroscience. Progress in our understanding of neural plasticity has profound implications for the treatment of a number of psychiatric and neurodegenerative disorders, and for enhancing performance in what are considered normal subjects. One of the promising aspects of neural plasticity is that it implies that the alterations that occur are reversible, even neuronal atrophy and cell loss. Reversibility of structural as well as functional plasticity has already been demonstrated in response to pharmacological treatments or even behavioral therapy. As the fundamental mechanisms of neural plasticity are further elucidated, new targets and paradigms for enhancing plasticity will be revealed and will lead to more effective and faster-acting therapeutic interventions.

Selected abbrewiations and acronyms

BDNFbrain-derived neurotrophic factor
cAMPcyclic adenosine monophosphate
CaREcAMP response element
CREBcAMP response element binding protein
FGF-2fibroblast growth factor-2
5-HT5 -hydroxy tryptamine (serotonin)
LTPlong-term potentiation
NMDAN-methyl-D-aspartate
PDE4phosphodiesterase type IV
PKAprotein kinase
SSRIselective serotonin reuptake inhibitor

Notes

This work is supported by USPHS grants MH45481 and 2 P01 MH25642, a Veterans Administration National Center Grant for posttraumatic stress disorder, and by the Connecticut Mental Health Center.

REFERENCES

1. Malenka R., Nicoll RA. Long-term potentiation - a decade of progress?. Science. 1999;285:1870–1874. [PubMed]
2. Lamprecht R., LeDoux J. Structural plasticity and memory. Nat Rev Neurosci. 2004;5:45–54. [PubMed]
3. Silva A., Kogan JH., Frankland PW., Kida S. CREB and memory. Ann Rev Neurosci. 1998;21:127–148. [PubMed]
4. Cahill L., McGaugh JL. Mechanisms of emotional arousal and lasting declarative memory. Trends Neurosci. 1998;21:294–299. [PubMed]
5. LeDoux J. Emotion circuits in the brain. Ann Rev Neurosci. 2000;23:155–184. [PubMed]
6. Kim J., Diamond DM. The stressed hippocampus, synaptic plasticity and lost memories. Nat Rev Neurosci. 2002;3:453–462. [PubMed]
7. Pavlides C., Nivon LG., McEwen BS. Effects of chronic stress on hippocampal long-term potentiation. Hippocampus. 2002;12:245–257. [PubMed]
8. Shors T., Seib TB., Levine S., Thompson RF. Inescapable versus escapable shock modulates long-term potentiation in the rat hippocampus. Science. 1989;244:224–226. [PubMed]
9. Zhou J., Zhang F., Zhang Y. Corticosterone inhibits generation of longterm potentiation in rat hippocampus slice: involvement of brain-derived neurotrophic factor. Brain Res. 2000;885:182–191. [PubMed]
10. McEwen B. Stress and hippocampal plasticity. Curr Opin Neurobiol. 1999;5:205–216. [PubMed]
11. Wooley CS., Gould E., McEwen BS. Exposure to excess glucocorticoids alters dendritic morphology of adult hippocampal pyramidal neurons. Brain Res. 1990;531:225–231. [PubMed]
12. Watanabe Y., Gould E., Daniels DC., Cameron H., McEwen BS. Tianeptine attenuates stress-induced morphological changes in the hippocampus. Eur J Pharmacol. 1992;222:157–162. [PubMed]
13. Margarinos A., McEwen BS., Flugge G., Fuchs E. Chronic psychosocial stress causes apical dendritic atrophy of hippocampal CA3 pyramidal neurons in subordinate tree shrews. J Neurosci. 1996;16:3534–3540. [PubMed]
14. Vyas A., Mitra R., Shankaranarayana Rao BS., Chattarji S. Chronic stress induces contrasting patterns of dendritic remodeling in hippocampal and amygdaloid neurons. J Neurosci. 2002;22:6810–6818. [PubMed]
15. Gage F. Mammalian neural stem cells. Science. 2000;287:1433–1438. [PubMed]
16. Gould E., Beylin A., Tanapat P., Reeves A., Shors TJ. Learning enhances adult neurogenesis in the hippocampal formation. Nat Neurosci. 1999;2:260–265. [PubMed]
17. van Praag H., Schlinder AF., Christie BR., Toni N., Palmer TD., Gage FH. Functional neurogenesis in the adult mouse dentate gyrus. Nature. 2002;415:1030–1034. [PubMed]
18. Duman R., Malberg J., Nakagawa S. Regulation of adult neurogenesis by psychotropic drugs and stress. J Pharmacol Exp Ther. 2001;299:401–407. [PubMed]
19. Gould E., McEwen BS., Tanapat P., Galea LAM., Fuchs E. Neurogenesis in the dentate gyrus of the adult tree shrew is regulated by psychosocial stress and NMDA receptor activation. J Neurosci. 1997;17:2492–2498. [PubMed]
20. Tanapat P., Hastings NB., Rydel TA., Galea LAM., Gould E. Exposure to fox odor inhibits cell proliferation in the hippocampus of adult rats via an adrenal hormone-dependent mechanism. J Cornp Neurol. 2001;437:496–504. [PubMed]
21. Lee K., Lynch KR., Nguyen T., et al. Cloning and charactization of additional members of the G protein-coupled receptor family. Biochim Biophys Acta. 2000;1490:311–323. [PubMed]
22. Malberg J., Duman RS. Cell proliferation in adult hippocmpus is decreased by inescapable stress: reversal by fluoxetine treatment. Neuropsychopharmacology. 2003;28:1562–1571. [PubMed]
23. Barrot M., Olivier JD., Perrotti LI., et al. CREB activity in the nucleus accumbens shell controls gating of behavioral responses to emotional stimuli. Proc Natl Acad Sci USA. 2002;99:11435–11440. [PMC free article] [PubMed]
24. Bilang-Bleuel A., Rech J., De Carli S., Holsboer F., Reul JMHM. Forced swimming evokes a biphasic response in CREB phosphorylation in extrahypothalamic limbic and neocortical brain structures in the rat. Eur J Neurosci. 2002;15:1048–1060. [PubMed]
25. Bruijnzeel A., Stam R., Compaan JC., Wiegant VM. Stress-induced sensitization of CRH-ir but not P-CREB-ir responsivity in the rat central nervous system. Brain Res. 2001;908:187–196. [PubMed]
26. Trentani A., Kuipers SD., Ter Horst GJ., Den Boer JA. Selective chronic stress-induced in vivo ERK1/2 hyperphosphorylation in medial prefrontocortical dendrites: implications for stress-related cortical pathology?. Eur] Neurosci. 2002;15:1681–1691. [PubMed]
27. Duman R. Role of neurotrophic factors in the etiology and treatment of mood disorders. Neuromol Med. 2004;5:11–26. [PubMed]
28. Nibuya M., Morinobu S., Duman RS. Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J Neurosci. 1995;15:7539–7547. [PubMed]
29. Smith MA., Makino S., Kvetnansky R., Post RM. Stress alters the express of brain-derived neurotrophic factor and neurotrophin-3 mRNAs in the hippocampus. J Neurosci. 1995;15:1768–1777. [PubMed]
30. Bremner J., Narayan M., Anderson ER., Staib LH., Miller H., Charney DS. Smaller hippocampal volume in major depression. Am J Psychiatry. 2000;157:115–117. [PubMed]
31. Sheline Y., Wany P., Gado MH., Csernansky JG., Vannier MW. Hippocampal atrophy in recurrent major depression. Proc Natl Acad Sci USA. 1996;93:3908–3913. [PMC free article] [PubMed]
32. Bremner JD., Randall P., Scott TM., et al. MRI-based measurement of hippocampal volume in patients with combat-related posttraumatic stress disorder. Am J Psychiatry. 1995;152:973–981 . [PMC free article] [PubMed]
33. MacQueen G., Campbell S., McEwen BS., et al. Course of illness, hippocampal function, and hippocampal volume in major depression. Proc Natl Acad Sci USA. 2003;100:1387–1392. [PMC free article] [PubMed]
34. Sheline Y., Sanghavi M., Mintun MA., Gado MH. Depression duration but not age predicts hippocampal volume loss in medically healthy wormen with recurrent major depression. J Neurosci. 1999;19:5034–5043. [PubMed]
35. Manji H., Duman RS. Impairments of neuroplasticity and cellular resilience in severe mood disorders: implications for the development of novel therapeutics. Psychopharmacol Bull. 2001;35:5–49. [PubMed]
36. Cotter D., Mackay D., Landau S., Kerwin R., Everall I. Reduced glial cell density and neuronal size in the anterior cingulate cortex in major depressive disorder. Arch Gen Psychiatry. 2001;58:545–553. [PubMed]
37. Ongur D., Drevets WC., Price JL. Glial reduction in the subgenual prefrontal cortex in mood disorders. Proc Natl Acad Sci U S A. 1998;95:13290–13295. [PMC free article] [PubMed]
38. Rajkowska G., Miguel-Hidalgo JJ., Wei J., et al. Morphometric evidence for neuronal and glial prefrontal cell pathology in major depression. Biol Psychiatry. 1999;45:1085–1098. [PubMed]
39. Dowlatshahi D., MacQueen GM., Wang JF., Young LT. Increased temporal cortex CREB concentrations and antidepressant treatment in major depression. Lancet. 1998;352:1754–1755. [PubMed]
40. Dwivedi Y., Rizavi HS., Conley RR., Tamminga CA., Pandey GN. Altered gene expression of brain-derived neurotrophic factor and receptor tyrosine kinase B in postmortem brain of suicide subjects. Arch Gen Psychiatry. 2003;60:804–815. [PubMed]
41. Dwivedi Y., Rizavi HS., Roberts RC., Conley RC., Tamminga CA., Pandey GN. Reduced activation and expression of ERK1/2 MAP kinase in the postmortem brain of depressed suicide subjects. J Neurochem. 2001;77:916–928. [PubMed]
42. Czeh B., Michaelis T., Watanabe T., et al. Stress-induced changes in cerebral metabolites, hippocampal volume, and cell proliferation are prevented by antidepressant treatment with tianeptine. Proc Natl Acad Sci USA. 2001;98:12796–12801. [PMC free article] [PubMed]
43. Malberg J., Eisch AJ., Nestler EJ., Duman RS. Chronic antidepressant treatment increases neurogenesis in adult hippocampus. J Neurosci. 2000;20:9104–9110. [PubMed]
44. Madsen T., Treschow A., Bengzon J., Bolwig TG., Lindvall O., Tingstrôm A. Increased neurogenesis in a model of electroconvulsive therapy. Biol Psychiatry. 2000;47:1043–1049. [PubMed]
45. Manev H., Uz T., Smalheiser NR., Manev R. Antidepressants alter cell proliferation in the adult brain in vivo and in neural cultures in vitro. Eur J Pharmacol. 2001;411:67–70. [PubMed]
46. Nakagawa S., Kim JE., Lee R., et al. Regulation of neurogenesis in adult mouse hippocampus by cAMP and cAMP response element-binding protein. J Neurosci. 2002;22:9868–9876. [PubMed]
47. Lee H., Kim JW., Yim SV., et al. Fluoxetine enhances cell proliferation and prevents apoptosis in dentate gyrus of maternally separated rats. Mot Psychiatry. 2001;6:725–728. [PubMed]
48. van der Hart M., Czeh B., de Biurrun G., et al. Substance P receptor antagonist and clomipramine prevent stress-induced alterations in cerebral metabolites, cytogenesis in the dentate gyrus and hippocampal volume. Mol Psychiatry. 2002;7:933–941. [PubMed]
49. Santarelli L., Saxe M., Gross C., et al. Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science. 2003;301:805–809. [PubMed]
50. Vollmayr B., Simonis C., Weber S., Gass P., Henn F. Reduced cell proliferation in the dentate gyrus is not correlated with the development of learned helplessness. Biol Psychiatry. 2003;54:1035–1040. [PubMed]
51. Manji H., Drevets WC., Charney DS. The cellular neurobiology of depression. Nat Med. 2001;7:541–547. [PubMed]
52. Nestler E., Barrot M., DiLeone RJ., Eisch AJ., Gold SJ. Monteggia LM. Neurobiology of depression. Neuron. 2002;34:13–25. [PubMed]
53. Duman R., Heninger GR., Nestler EJ. A molecular and cellular theory of depression. Arch Gen Psychiatry. 1997;54:597–606. [PubMed]
54. Duman R., Malberg J., Nakagawa S. C D'Sa. Neuronal plasticity and survival in mood disorders. Biol Psychiatry. 2000;48:732–739. [PubMed]
55. Nestler E., Terwilliger RZ., Duman RS. Chronic antidepressant administration alters the subcellular distribution of cAMP-dependent protein kinase in rat frontal cortex. J Neurochem. 1989;53:1644–1647. [PubMed]
56. Nibuya M., Nestler EJ., Duman RS. Chronic antidepressant administration increases the expression of cAMP response element binding protein (CREB) in rat hippocampus. J Neurosci. 1996;16:2365–2372. [PubMed]
57. Thome J., Sakai N., Shin KH., et al. cAMP response element-mediated gene transcription is upregulated by chronic antidepressant treatment. J Neurosci. 2000;20:4030–4036. [PubMed]
58. Nakagawa S., Kim JE., Lee R., Chen J., Fujioka T., Malberg J. Localization of phosphorylated cAMP response element-binding protein in immature neurons of adult hippocampus. J Neurosci. 2002;22:9868–9876. [PubMed]
59. Chen A-H., Shirayama Y., Shin KH., Neve RL., Duman RS. Expression of the cAMP response element binding protein (CREB) in hippocampus produces antidepressant effect. Biol Psychiatry. 2001;49:753–762. [PubMed]
60. Pliakas A., Carlson RR., Neve RL., Konradi C., Nestler EJ., Carlezon WA. Altered responsiveness to cocaine and increased immobility in the forced swim test associated with elevated CREB expression in the nucleus accumbens. J Neurosci. 2001;21:7397–7403. [PMC free article] [PubMed]
61. Newton S., Thome J., Wallace TL., et al. Inhibition of cAMP response element-binding protein or dynorphin in the nucleus accumbens produces an antidepressant-like effect. J Neurosci. 2002;24:10883–10890. [PubMed]
62. Shirayama Y., Chen AC., Nakagawa S., Russell RS., Duman RS. Brain-derived neurotrophic factor produces antidepressant effects in behavioral models of depression. J Neurosci. 2002;22:3251–3261. [PubMed]
63. Saarelainen T., Hendolin P., Lucas G., et al. Activation of the trkB neurotrophin receptor is induced by antidepressant drugs and is required for antidepressant-induced behavioral. effects. J Neurosci. 2003;23:349–357. [PubMed]
64. Siuciak JA., Lewis DR., Wiegand SJ., Lindsay R. Antidepressant-like effect of brain-derived neurotrophic factor (BDNF). Pharmacol Biochem Behav. 1997;56:131–137. [PubMed]
65. Mallei A., Shi B., Mocchetti I. Antidepressant treatments induce the expression of basic fibroblast growth factor in cortical and hippocampal neurons. 2002;61:1017–1024. [PubMed]
66. Newton S., Collier E., Hunsberger J., Adams D., Salvanayagam E., Duman RS. Gene profile of electroconvulsive seizures: induction of neurogenic and angiogenic factors. J Neurosci. 2003;23:10841–10851. [PubMed]
67. Sheline Y., Gado MH., Kraemer HC. Untreated depression and hippocampal volume loss. Am J Psychiatry. 2003;160:1–3. [PubMed]
68. Vermetten E., Vythilingam M., Southwick SM., Charney DS., Bremner JD. Long-term treatment with paroxetine increases verbal declarative memory and hippocampal volume in posttraumatic stress disorder. Biol Psychiatry. 2003;54:693–702. [PMC free article] [PubMed]
69. Horowski R., Sastre-Y-Hernandez M. Clinical effects of the neurotrophic selective cAMP phosphodiesterase inhibitor rolipram in depressed patients: global evaluation of the preliminary reports. CurrTher Res. 1985;38:23–29.
70. Wachtel H. Potential antidepressant activity of rolipram and other selective cyclic adenosine 3',5“-monophosphate phosphodiesterase inhibitors. Neuropharmacology. 1983;22:267–272. [PubMed]
71. Takahashi M., Terwilliger R., Lane S., Mezes PS., Conti M., Duman RS. Chronic antidepressant administration increases the expression of cAMP phosphodiesterase 4A and 4B isoforms. J Neurosci. 1999;19:610–618. [PubMed]
72. Zhang H-T., Huang Y., Jin SJC., et al. Antidepressant-like profile and reduced sensitivity to rolipram in mice deficient in the PDE4D phosphodiesterase enzyme. Neuropsychopharmacology . 2002;27:587–595. [PubMed]
73. Li X., Tizzano JP., Griffey K., Clay M., Lindstron T., Skolnick P. Antidepressant-like actions of an AMPA receptor potentiator (LY392098). Neuropharmacology. 2001;40:1028–1033. [PubMed]
74. Marvanova M., Lakso M., Pirhonen J., Nawa H., Wong G., Castren E. The neuroprotective agent memantine induces brain-derived neurotrophic factor and trkB receptor expression in rat brain. Mol Cell Neurosci. 2001;18:247–258. [PubMed]
75. Katoh-Semba R., Asano T., Ueda H., et al. Riluzole enhances expression of brain-derived neurotrophic factor with consequent proliferation of granule precursor cells in the rat hippocampus. FASEBJ. 2001;16:1328–1330. [PubMed]

No comments:

Post a Comment