Prefrontal cortex

Last updated
Prefrontal cortex
Gray726-Brodman-prefrontal.svg
Brodmann areas, 8, 9, 10, 11, 12, 13, 14, 24, 25, 32, 44, 45, 46, and 47 are all in the prefrontal cortex [1]
Details
Part of Frontal lobe
Parts Superior frontal gyrus
Middle frontal gyrus
Inferior frontal gyrus
Artery Anterior cerebral
Middle cerebral
Vein Superior sagittal sinus
Identifiers
Latin cortex praefrontalis
MeSH D017397
NeuroNames 2429
NeuroLex ID nlx_anat_090801, ilx_0109209
FMA 224850
Anatomical terms of neuroanatomy

In mammalian brain anatomy, the prefrontal cortex (PFC) covers the front part of the frontal lobe of the cerebral cortex. It is the association cortex in the frontal lobe. [2] The PFC contains the Brodmann areas BA8, BA9, BA10, BA11, BA12, BA13, BA14, BA24, BA25, BA32, BA44, BA45, BA46, and BA47. [1]

Contents

This brain region is involved in a wide range of higher-order cognitive functions, including speech formation (Broca's area), gaze (frontal eye fields), working memory (dorsolateral prefrontal cortex), and risk processing (e.g. ventromedial prefrontal cortex). The basic activity of this brain region is considered to be orchestration of thoughts and actions in accordance with internal goals. [3] Many authors have indicated an integral link between a person's will to live, personality, and the functions of the prefrontal cortex. [4]

This brain region has been implicated in executive functions, such as planning, decision making, working memory, personality expression, moderating social behavior and controlling certain aspects of speech and language. [5] [6] [7] [8] Executive function relates to abilities to differentiate among conflicting thoughts, determine good and bad, better and best, same and different, future consequences of current activities, working toward a defined goal, prediction of outcomes, expectation based on actions, and social "control" (the ability to suppress urges that, if not suppressed, could lead to socially unacceptable outcomes).

The frontal cortex supports concrete rule learning, with more anterior regions supporting rule learning at higher levels of abstraction. [9]

Structure

Definition

There are three possible ways to define the prefrontal cortex:

Granular frontal cortex

The prefrontal cortex has been defined based on cytoarchitectonics by the presence of a cortical granular layer IV. It is not entirely clear who first used this criterion. Many of the early cytoarchitectonic researchers restricted the use of the term prefrontal to a much smaller region of cortex including the gyrus rectus and the gyrus rostralis (Campbell, 1905; G. E. Smith, 1907; Brodmann, 1909; von Economo and Koskinas, 1925). In 1935, however, Jacobsen used the term prefrontal to distinguish granular prefrontal areas from agranular motor and premotor areas. [10] In terms of Brodmann areas, the prefrontal cortex traditionally includes areas 8, 9, 10, 11, 12, 13, 14, 24, 25, 32, 44, 45, 46, and 47, [1] however, not all of these areas are strictly granular – 44 is dysgranular, caudal 11 and orbital 47 are agranular. [11] The main problem with this definition is that it works well only in primates but not in nonprimates, as the latter lack a granular layer IV. [12]

Projection zone

To define the prefrontal cortex as the projection zone of the mediodorsal nucleus of the thalamus builds on the work of Rose and Woolsey, [13] who showed that this nucleus projects to anterior and ventral parts of the brain in nonprimates, however, Rose and Woolsey termed this projection zone "orbitofrontal." It seems to have been Akert, who, for the first time in 1964, explicitly suggested that this criterion could be used to define homologues of the prefrontal cortex in primates and nonprimates. [14] This allowed the establishment of homologies despite the lack of a granular frontal cortex in nonprimates.

The projection zone definition is still widely accepted today (e.g. Fuster [15] ), although its usefulness has been questioned. [11] [16] Modern tract tracing studies have shown that projections of the mediodorsal nucleus of the thalamus are not restricted to the granular frontal cortex in primates. As a result, it was suggested to define the prefrontal cortex as the region of cortex that has stronger reciprocal connections with the mediodorsal nucleus than with any other thalamic nucleus. [12] Uylings et al. [12] acknowledge, however, that even with the application of this criterion, it might be rather difficult to define the prefrontal cortex unequivocally.

Electrically silent area of frontal cortex

A third definition of the prefrontal cortex is the area of frontal cortex whose electrical stimulation does not lead to observable movements. For example, in 1890 David Ferrier [17] used the term in this sense. One complication with this definition is that the electrically "silent" frontal cortex includes both granular and non-granular areas. [11]

Subdivisions

Brodmann areas.jpg

According to Striedter, [18] the PFC of humans can be delineated into two functionally, morphologically, and evolutionarily different regions: the ventromedial PFC (vmPFC) consisting of:

  1. the ventral prefrontal cortex (VPFC)
  2. the medial prefrontal cortex present in all mammals (MPFC)

and the lateral prefrontal cortex (LPFC), consisting of:

  1. the dorsolateral prefrontal cortex (DLPFC)
  2. the ventrolateral prefrontal cortex (VLPFC) present only in primates.

The LPFC contains the Brodmann areas BA8, BA9, BA10, BA45, BA46, and BA47. Some researchers also include BA44. The vmPFC contains the Brodmann areas BA12, BA25, BA32, BA33, BA24, BA11, BA13, and BA14.

The table below shows different ways to subdivide parts of the human prefrontal cortex based upon Brodmann areas. [1]

8910464547441225323324111314
lateral ventromedial
dorsolateral ventrolateral medialventral

Interconnections

The prefrontal cortex is highly interconnected with much of the brain, including extensive connections with other cortical, subcortical and brain stem sites. [21] The dorsal prefrontal cortex is especially interconnected with brain regions involved with attention, cognition and action, [22] while the ventral prefrontal cortex interconnects with brain regions involved with emotion. [23] The prefrontal cortex also receives inputs from the brainstem arousal systems, and its function is particularly dependent on its neurochemical environment. [24] Thus, there is coordination between one's state of arousal and mental state. [25] The interplay between the prefrontal cortex and socioemotional system of the brain is relevant for adolescent development, as proposed by the Dual Systems Model.

The medial prefrontal cortex has been implicated in the generation of slow-wave sleep (SWS), and prefrontal atrophy has been linked to decreases in SWS. [26] Prefrontal atrophy occurs naturally as individuals age, and it has been demonstrated that older adults experience impairments in memory consolidation as their medial prefrontal cortices degrade. [26] In older adults, instead of being transferred and stored in the neocortex during SWS, memories start to remain in the hippocampus where they were encoded, as evidenced by increased hippocampal activation compared to younger adults during recall tasks, when subjects learned word associations, slept, and then were asked to recall the learned words. [26]

The ventrolateral prefrontal cortex (VLPFC) has been implicated in various aspects of speech production and language comprehension. The VLPFC is richly connected to various regions of the brain including the lateral and medial temporal lobe, the superior temporal cortex, the infertemporal cortex, the perirhinal cortex, and the parahippoccampal cortex. [27] These brain areas are implicated in memory retrieval and consolidation, language processing, and association of emotions. These connections allow the VLPFC to mediate explicit and implicit memory retrieval and integrate it with language stimulus to help plan coherent speech. [28] In other words, choosing the correct words and staying "on topic" during conversation come from the VLPFC.

Function

Executive function

The original studies of Fuster and of Goldman-Rakic emphasized the fundamental ability of the prefrontal cortex to represent information not currently in the environment, and the central role of this function in creating the "mental sketch pad". Goldman-Rakic spoke of how this representational knowledge was used to intelligently guide thought, action, and emotion, including the inhibition of inappropriate thoughts, distractions, actions, and feelings. [29] In this way, working memory can be seen as fundamental to attention and behavioral inhibition. Fuster speaks of how this prefrontal ability allows the wedding of past to future, allowing both cross-temporal and cross-modal associations in the creation of goal-directed, perception-action cycles. [30] This ability to represent underlies all other higher executive functions.

Shimamura proposed Dynamic Filtering Theory to describe the role of the prefrontal cortex in executive functions. The prefrontal cortex is presumed to act as a high-level gating or filtering mechanism that enhances goal-directed activations and inhibits irrelevant activations. This filtering mechanism enables executive control at various levels of processing, including selecting, maintaining, updating, and rerouting activations. It has also been used to explain emotional regulation. [31]

Miller and Cohen proposed an Integrative Theory of Prefrontal Cortex Function, that arises from the original work of Goldman-Rakic and Fuster. The two theorize that "cognitive control stems from the active maintenance of patterns of activity in the prefrontal cortex that represents goals and means to achieve them. They provide bias signals to other brain structures whose net effect is to guide the flow of activity along neural pathways that establish the proper mappings between inputs, internal states, and outputs needed to perform a given task". [32] In essence, the two theorize that the prefrontal cortex guides the inputs and connections, which allows for cognitive control of our actions.

The prefrontal cortex is of significant importance when top-down processing is needed. Top-down processing by definition is when behavior is guided by internal states or intentions. According to the two, "The PFC is critical in situations when the mappings between sensory inputs, thoughts, and actions either are weakly established relative to other existing ones or are rapidly changing". [32] An example of this can be portrayed in the Wisconsin Card Sorting Test (WCST). Subjects engaging in this task are instructed to sort cards according to the shape, color, or number of symbols appearing on them. The thought is that any given card can be associated with a number of actions and no single stimulus-response mapping will work. Human subjects with PFC damage are able to sort the card in the initial simple tasks, but unable to do so as the rules of classification change.

Miller and Cohen conclude that the implications of their theory can explain how much of a role the PFC has in guiding control of cognitive actions. In the researchers' own words, they claim that, "depending on their target of influence, representations in the PFC can function variously as attentional templates, rules, or goals by providing top-down bias signals to other parts of the brain that guide the flow of activity along the pathways needed to perform a task". [32]

Experimental data indicate a role for the prefrontal cortex in mediating normal sleep physiology, dreaming and sleep-deprivation phenomena. [33]

When analyzing and thinking about attributes of other individuals, the medial prefrontal cortex is activated, however, it is not activated when contemplating the characteristics of inanimate objects. [34]

Studies using fMRI have shown that the medial prefrontal cortex (mPFC), specifically the anterior medial prefrontal cortex (amPFC), may modulate mimicry behavior. Neuroscientists are suggesting that social priming influences activity and processing in the amPFC, and that this area of the prefrontal cortex modulates mimicry responses and behavior. [35]

As of recent, researchers have used neuroimaging techniques to find that along with the basal ganglia, the prefrontal cortex is involved with learning exemplars, which is part of the exemplar theory , one of the three main ways our mind categorizes things. The exemplar theory states that we categorize judgements by comparing it to a similar past experience within our stored memories. [36]

A 2014 meta-analysis by Professor Nicole P.Yuan from the University of Arizona found that larger prefrontal cortex volume and greater PFC cortical thickness were associated with better executive performance. [37]

Attention and memory

Lebedev et al. experiment that dissociated representation of spatial attention from representation of spatial memory in prefrontal cortex Attention vs memory.gif
Lebedev et al. experiment that dissociated representation of spatial attention from representation of spatial memory in prefrontal cortex

A widely accepted theory regarding the function of the brain's prefrontal cortex is that it serves as a store of short-term memory. This idea was first formulated by Jacobsen, who reported in 1936 that damage to the primate prefrontal cortex caused short-term memory deficits. [39] Karl Pribram and colleagues (1952) identified the part of the prefrontal cortex responsible for this deficit as area 46, also known as the dorsolateral prefrontal cortex (dlPFC). [40] More recently, Goldman-Rakic and colleagues (1993) evoked short-term memory loss in localized regions of space by temporary inactivation of portions of the dlPFC. [41] Once the concept of working memory (see also Baddeley's model of working memory) was established in contemporary neuroscience by Alan Baddeley (1986), these neuropsychological findings contributed to the theory that the prefrontal cortex implements working memory and, in some extreme formulations, only working memory. [42] In the 1990s this theory developed a wide following, and it became the predominant theory of PF function, especially for nonhuman primates. The concept of working memory used by proponents of this theory focused mostly on the short-term maintenance of information, and rather less on the manipulation or monitoring of such information or on the use of that information for decisions. Consistent with the idea that the prefrontal cortex functions predominantly in maintenance memory, delay-period activity in the PF has often been interpreted as a memory trace. (The phrase "delay-period activity" applies to neuronal activity that follows the transient presentation of an instruction cue and persists until a subsequent "go" or "trigger" signal.)

To explore alternative interpretations of delay-period activity in the prefrontal cortex, Lebedev et al. (2004) investigated the discharge rates of single prefrontal neurons as monkeys attended to a stimulus marking one location while remembering a different, unmarked location. [38] Both locations served as potential targets of a saccadic eye movement. Although the task made intensive demands on short-term memory, the largest proportion of prefrontal neurons represented attended locations, not remembered ones. These findings showed that short-term memory functions cannot account for all, or even most, delay-period activity in the part of the prefrontal cortex explored. The authors suggested that prefrontal activity during the delay-period contributes more to the process of attentional selection (and selective attention) than to memory storage. [8] [43]

Speech production and language

Various areas of the prefrontal cortex have been implicated in a multitude of critical functions regarding speech production, language comprehension, and response planning before speaking. [7] Cognitive neuroscience has shown that the left ventrolateral prefrontal cortex is vital in the processing of words and sentences.

The right prefrontal cortex has been found to be responsible for coordinating the retrieval of explicit memory for use in speech, whereas the deactivation of the left is responsible for mediating implicit memory retrieval to be used in verb generation. [7] Recollection of nouns (explicit memory) is impaired in some amnesic patients with damaged right prefrontal cortices, but verb generation remains intact because of its reliance on left prefrontal deactivation. [28]

Many researchers now include BA45 in the prefrontal cortex because together with BA44 it makes up an area of the frontal lobe called Broca's area. [20] Broca's Area is widely considered the output area of the language production pathway in the brain (as opposed to Wernicke's area in the medial temporal lobe, which is seen as the language input area). BA45 has been shown to be implicated for the retrieval of relevant semantic knowledge to be used in conversation/speech. [7] The right lateral prefrontal cortex (RLPFC) is implicated in the planning of complex behavior, and together with bilateral BA45, they act to maintain focus and coherence during speech production. [28] However, left BA45 has been shown to be activated significantly while maintaining speech coherence in young people. Older people have been shown to recruit the right BA45 more so than their younger counterparts. [28] This aligns with the evidence of decreased lateralization in other brain systems during aging.

In addition, this increase in BA45 and RLPFC activity in combination of BA47 in older patients has been shown to contribute to "off-topic utterances." The BA47 area in the prefrontal cortex is implicated in "stimulus-driven" retrieval of less-salient knowledge than is required to contribute to a conversation. [28] In other words, elevated activation of the BA47 together with altered activity in BA45 and the broader RLPFC has been shown to contribute to the inclusion of less relevant information and irrelevant tangential conversational speech patterns in older subjects.

Clinical significance

In the last few decades, brain imaging systems have been used to determine brain region volumes and nerve linkages. Several studies have indicated that reduced volume and interconnections of the frontal lobes with other brain regions is observed in patients diagnosed with mental disorders; those subjected to repeated stressors; [44] those who excessively consume sexually explicit materials; [45] suicides; [46] ; criminals; sociopaths; those affected by lead poisoning; [47] It is believed that at least some of the human abilities to feel guilt or remorse, and to interpret reality, are dependent on a well-functioning prefrontal cortex. [48] The advanced neurocircuitry and self-regulatory function of the human prefrontal cortex is also associated with the higher sentience in humans, [49] as the prefrontal cortex in humans occupies a far larger percentage of the brain than in any other animal. It is theorized that, as the brain has tripled in size over five million years of human evolution, [50] the prefrontal cortex has increased in size sixfold. [51]

A review on executive functions in healthy exercising individuals noted that the left and right halves of the prefrontal cortex, which is divided by the medial longitudinal fissure, appears to become more interconnected in response to consistent aerobic exercise. [52] Two reviews of structural neuroimaging research indicate that marked improvements in prefrontal and hippocampal gray matter volume occur in healthy adults that engage in medium intensity exercise for several months. [53] [54]

Chronic intake of alcohol leads to persistent alterations in brain function including altered decision-making ability. The prefrontal cortex of chronic alcoholics has been shown to be vulnerable to oxidative DNA damage and neuronal cell death. [55]

History

Perhaps the seminal case in prefrontal cortex function is that of Phineas Gage, whose left frontal lobe was destroyed when a large iron rod was driven through his head in an 1848 accident. The standard presentation is that, although Gage retained normal memory, speech and motor skills, his personality changed radically: He became irritable, quick-tempered, and impatient—characteristics he did not previously display — so that friends described him as "no longer Gage"; and, whereas he had previously been a capable and efficient worker, afterward he was unable to complete. [56] However, careful analysis of primary evidence shows that descriptions of Gage's psychological changes are usually exaggerated when held against the description given by Gage's doctor, the most striking feature being that changes described years after Gage's death are far more dramatic than anything reported while he was alive. [57] [58]

Subsequent studies on patients with prefrontal injuries have shown that the patients verbalized what the most appropriate social responses would be under certain circumstances. Yet, when actually performing, they instead pursued behavior aimed at immediate gratification, despite knowing the longer-term results would be self-defeating.

The interpretation of this data indicates that not only are skills of comparison and understanding of eventual outcomes harbored in the prefrontal cortex but the prefrontal cortex (when functioning correctly) controls the mental option to delay immediate gratification for a better or more rewarding longer-term gratification result. This ability to wait for a reward is one of the key pieces that define optimal executive function of the human brain.[ citation needed ]

There is much current research devoted to understanding the role of the prefrontal cortex in neurological disorders. Clinical trials have begun on certain drugs that have been shown to improve prefrontal cortex function, including guanfacine, which acts through the alpha-2A adrenergic receptor. A downstream target of this drug, the HCN channel, is one of the most recent areas of exploration in prefrontal cortex pharmacology. [59]

Etymology

The term "prefrontal" as describing a part of the brain appears to have been introduced by Richard Owen in 1868. [10] For him, the prefrontal area was restricted to the anterior-most part of the frontal lobe (approximately corresponding to the frontal pole). It has been hypothesized that his choice of the term was based on the prefrontal bone present in most amphibians and reptiles. [10]

Additional images

See also

Related Research Articles

<span class="mw-page-title-main">Entorhinal cortex</span> Area of the temporal lobe of the brain

The entorhinal cortex (EC) is an area of the brain's allocortex, located in the medial temporal lobe, whose functions include being a widespread network hub for memory, navigation, and the perception of time. The EC is the main interface between the hippocampus and neocortex. The EC-hippocampus system plays an important role in declarative (autobiographical/episodic/semantic) memories and in particular spatial memories including memory formation, memory consolidation, and memory optimization in sleep. The EC is also responsible for the pre-processing (familiarity) of the input signals in the reflex nictitating membrane response of classical trace conditioning; the association of impulses from the eye and the ear occurs in the entorhinal cortex.

Working memory is a cognitive system with a limited capacity that can hold information temporarily. It is important for reasoning and the guidance of decision-making and behavior. Working memory is often used synonymously with short-term memory, but some theorists consider the two forms of memory distinct, assuming that working memory allows for the manipulation of stored information, whereas short-term memory only refers to the short-term storage of information. Working memory is a theoretical concept central to cognitive psychology, neuropsychology, and neuroscience.

<span class="mw-page-title-main">Broca's area</span> Speech production region in the dominant hemisphere of the hominid brain

Broca's area, or the Broca area, is a region in the frontal lobe of the dominant hemisphere, usually the left, of the brain with functions linked to speech production.

<span class="mw-page-title-main">Cingulate cortex</span> Part of the brain within the cerebral cortex

The cingulate cortex is a part of the brain situated in the medial aspect of the cerebral cortex. The cingulate cortex includes the entire cingulate gyrus, which lies immediately above the corpus callosum, and the continuation of this in the cingulate sulcus. The cingulate cortex is usually considered part of the limbic lobe.

<span class="mw-page-title-main">Brodmann area</span> Region of the brain

A Brodmann area is a region of the cerebral cortex, in the human or other primate brain, defined by its cytoarchitecture, or histological structure and organization of cells. The concept was first introduced by the German anatomist Korbinian Brodmann in the early 20th century. Brodmann mapped the human brain based on the varied cellular structure across the cortex and identified 52 distinct regions, which he numbered 1 to 52. These regions, or Brodmann areas, correspond with diverse functions including sensation, motor control, and cognition.

<span class="mw-page-title-main">Frontal lobe</span> Part of the brain

The frontal lobe is the largest of the four major lobes of the brain in mammals, and is located at the front of each cerebral hemisphere. It is parted from the parietal lobe by a groove between tissues called the central sulcus and from the temporal lobe by a deeper groove called the lateral sulcus. The most anterior rounded part of the frontal lobe is known as the frontal pole, one of the three poles of the cerebrum.

<span class="mw-page-title-main">Brodmann area 9</span> Part of the frontal cortex in the brain of humans and other primates

Brodmann area 9, or BA9, refers to a cytoarchitecturally defined portion of the frontal cortex in the brain of humans and other primates. Its cytoarchitecture is referred to as granular due to the concentration of granule cells in layer IV. It contributes to the dorsolateral and medial prefrontal cortex.

<span class="mw-page-title-main">Brodmann area 10</span> Brain area

Brodmann area 10 is the anterior-most portion of the prefrontal cortex in the human brain. BA10 was originally defined broadly in terms of its cytoarchitectonic traits as they were observed in the brains of cadavers, but because modern functional imaging cannot precisely identify these boundaries, the terms anterior prefrontal cortex, rostral prefrontal cortex and frontopolar prefrontal cortex are used to refer to the area in the most anterior part of the frontal cortex that approximately covers BA10—simply to emphasize the fact that BA10 does not include all parts of the prefrontal cortex.

<span class="mw-page-title-main">Brodmann area 45</span> Brain area

Brodmann area 45 (BA45), is part of the frontal cortex in the human brain. It is situated on the lateral surface, inferior to BA9 and adjacent to BA46.

<span class="mw-page-title-main">Brodmann area 46</span> Brain area

Brodmann area 46, or BA46, is part of the frontal cortex in the human brain. It is between BA10 and BA45.

<span class="mw-page-title-main">Brodmann area 11</span> Brain area

Brodmann area 11 is one of Brodmann's cytologically defined regions of the brain. It is in the orbitofrontal cortex which is above the eye sockets (orbitae). It is involved in decision making, processing rewards, and encoding new information into long-term memory.

<span class="mw-page-title-main">Inferior frontal gyrus</span> Part of the brains prefrontal cortex

The inferior frontal gyrus (IFG),, is the lowest positioned gyrus of the frontal gyri, of the frontal lobe, and is part of the prefrontal cortex.

<span class="mw-page-title-main">Executive functions</span> Cognitive processes necessary for control of behavior

In cognitive science and neuropsychology, executive functions are a set of cognitive processes that are necessary for the cognitive control of behavior: selecting and successfully monitoring behaviors that facilitate the attainment of chosen goals. Executive functions include basic cognitive processes such as attentional control, cognitive inhibition, inhibitory control, working memory, and cognitive flexibility. Higher-order executive functions require the simultaneous use of multiple basic executive functions and include planning and fluid intelligence.

<span class="mw-page-title-main">Orbitofrontal cortex</span> Region of the prefrontal cortex of the brain

The orbitofrontal cortex (OFC) is a prefrontal cortex region in the frontal lobes of the brain which is involved in the cognitive process of decision-making. In non-human primates it consists of the association cortex areas Brodmann area 11, 12 and 13; in humans it consists of Brodmann area 10, 11 and 47.

<span class="mw-page-title-main">Frontal lobe disorder</span> Brain disorder

Frontal lobe disorder, also frontal lobe syndrome, is an impairment of the frontal lobe of the brain due to disease or frontal lobe injury. The frontal lobe plays a key role in executive functions such as motivation, planning, social behaviour, and speech production. Frontal lobe syndrome can be caused by a range of conditions including head trauma, tumours, neurodegenerative diseases, neurodevelopmental disorders, neurosurgery and cerebrovascular disease. Frontal lobe impairment can be detected by recognition of typical signs and symptoms, use of simple screening tests, and specialist neurological testing.

<span class="mw-page-title-main">Superior longitudinal fasciculus</span> Association fiber tract of the brain

The superior longitudinal fasciculus (SLF) is an association tract in the brain that is composed of three separate components. It is present in both hemispheres and can be found lateral to the centrum semiovale and connects the frontal, occipital, parietal, and temporal lobes. This bundle of tracts (fasciculus) passes from the frontal lobe through the operculum to the posterior end of the lateral sulcus where they either radiate to and synapse on neurons in the occipital lobe, or turn downward and forward around the putamen and then radiate to and synapse on neurons in anterior portions of the temporal lobe.

<span class="mw-page-title-main">Ventromedial prefrontal cortex</span> Body part

The ventromedial prefrontal cortex (vmPFC) is a part of the prefrontal cortex in the mammalian brain. The ventral medial prefrontal is located in the frontal lobe at the bottom of the cerebral hemispheres and is implicated in the processing of risk and fear, as it is critical in the regulation of amygdala activity in humans. It also plays a role in the inhibition of emotional responses, and in the process of decision-making and self-control. It is also involved in the cognitive evaluation of morality.

<span class="mw-page-title-main">Dorsolateral prefrontal cortex</span> Area of the prefrontal cortex of primates

The dorsolateral prefrontal cortex is an area in the prefrontal cortex of the primate brain. It is one of the most recently derived parts of the human brain. It undergoes a prolonged period of maturation which lasts into adulthood. The DLPFC is not an anatomical structure, but rather a functional one. It lies in the middle frontal gyrus of humans. In macaque monkeys, it is around the principal sulcus. Other sources consider that DLPFC is attributed anatomically to BA 9 and 46 and BA 8, 9 and 10.

<span class="mw-page-title-main">Ventrolateral prefrontal cortex</span> Part of the prefrontal cortex of the brain

The ventrolateral prefrontal cortex (VLPFC) is a section of the prefrontal cortex located on the inferior frontal gyrus, bounded superiorly by the inferior frontal sulcus and inferiorly by the lateral sulcus. It is attributed to the anatomical structures of Brodmann's area (BA) 47, 45 and 44.

Social cognitive neuroscience is the scientific study of the biological processes underpinning social cognition. Specifically, it uses the tools of neuroscience to study "the mental mechanisms that create, frame, regulate, and respond to our experience of the social world". Social cognitive neuroscience uses the epistemological foundations of cognitive neuroscience, and is closely related to social neuroscience. Social cognitive neuroscience employs human neuroimaging, typically using functional magnetic resonance imaging (fMRI). Human brain stimulation techniques such as transcranial magnetic stimulation and transcranial direct-current stimulation are also used. In nonhuman animals, direct electrophysiological recordings and electrical stimulation of single cells and neuronal populations are utilized for investigating lower-level social cognitive processes.

References

  1. 1 2 3 4 5 6 7 8 9 Murray E, Wise S, Grahatle K (2016). "Chapter 1: The History of Memory Systems". The Evolution of Memory Systems: Ancestors, Anatomy, and Adaptations (1st ed.). Oxford University Press. pp. 22–24. ISBN   978-0-19-150995-7 . Retrieved 12 March 2017.
  2. "Prefrontal Cortex - an overview | ScienceDirect Topics". www.sciencedirect.com. Retrieved 2024-02-08.
  3. Miller EK, Freedman DJ, Wallis JD (August 2002). "The prefrontal cortex: categories, concepts and cognition". Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 357 (1424): 1123–1136. doi:10.1098/rstb.2002.1099. PMC   1693009 . PMID   12217179.
  4. DeYoung CG, Hirsh JB, Shane MS, Papademetris X, Rajeevan N, Gray JR (June 2010). "Testing predictions from personality neuroscience. Brain structure and the big five". Psychological Science. 21 (6): 820–828. doi:10.1177/0956797610370159. PMC   3049165 . PMID   20435951.
  5. João, Rafael Batista; Filgueiras, Raquel Mattos (2018-10-03), Starcevic, Ana; Filipovic, Branislav (eds.), "Frontal Lobe: Functional Neuroanatomy of Its Circuitry and Related Disconnection Syndromes", Prefrontal Cortex, InTech, doi: 10.5772/intechopen.79571 , ISBN   978-1-78923-903-4 , retrieved 2023-06-24
  6. Yang Y, Raine A (November 2009). "Prefrontal structural and functional brain imaging findings in antisocial, violent, and psychopathic individuals: a meta-analysis". Psychiatry Research. 174 (2): 81–88. doi:10.1016/j.pscychresns.2009.03.012. PMC   2784035 . PMID   19833485.
  7. 1 2 3 4 Gabrieli JD, Poldrack RA, Desmond JE (February 1998). "The role of left prefrontal cortex in language and memory". Proceedings of the National Academy of Sciences of the United States of America. 95 (3): 906–913. Bibcode:1998PNAS...95..906G. doi: 10.1073/pnas.95.3.906 . PMC   33815 . PMID   9448258.
  8. 1 2 Baldauf D, Desimone R (April 2014). "Neural mechanisms of object-based attention". Science. 344 (6182): 424–427. Bibcode:2014Sci...344..424B. doi: 10.1126/science.1247003 . PMID   24763592. S2CID   34728448.
  9. Badre D, Kayser AS, D'Esposito M (April 2010). "Frontal cortex and the discovery of abstract action rules". Neuron. 66 (2): 315–326. doi:10.1016/j.neuron.2010.03.025. PMC   2990347 . PMID   20435006.
  10. 1 2 3 Finger S (1994). Origins of neuroscience: a history of explorations into brain function. Oxford [Oxfordshire]: Oxford University Press. ISBN   978-0-19-514694-3.[ page needed ]
  11. 1 2 3 Preuss TM (1995). "Do rats have prefrontal cortex? The rose-woolsey-akert program reconsidered". Journal of Cognitive Neuroscience. 7 (1): 1–24. doi:10.1162/jocn.1995.7.1.1. PMID   23961750. S2CID   2856619.
  12. 1 2 3 Uylings HB, Groenewegen HJ, Kolb B (November 2003). "Do rats have a prefrontal cortex?". Behavioural Brain Research. 146 (1–2): 3–17. doi:10.1016/j.bbr.2003.09.028. PMID   14643455. S2CID   32136463.
  13. Rose JE, Woolsey CN (1948). "The orbitofrontal cortex and its connections with the mediodorsal nucleus in rabbit, sheep and cat". Research Publications – Association for Research in Nervous and Mental Disease. 27: 210–232. PMID   18106857.
  14. Preuss TM, Goldman-Rakic PS (August 1991). "Myelo- and cytoarchitecture of the granular frontal cortex and surrounding regions in the strepsirhine primate Galago and the anthropoid primate Macaca". The Journal of Comparative Neurology. 310 (4): 429–474. doi:10.1002/cne.903100402. PMID   1939732. S2CID   34575725.
  15. Fuster JM (2008). The Prefrontal Cortex (4th ed.). Boston: Academic Press. ISBN   978-0-12-373644-4.[ page needed ]
  16. Markowitsch HJ, Pritzel M (1979). "The prefrontal cortex: Projection area of the thalamic mediodorsal nucleus?". Physiological Psychology. 7 (1): 1–6. doi: 10.3758/bf03326611 .
  17. Ferrier D (June 1890). "The Croonian Lectures on Cerebral Localisation". British Medical Journal. 1 (1537): 1349–1355. doi:10.1136/bmj.1.1537.1349. PMC   2207859 . PMID   20753055.
  18. Striedter GF (2005). Principles of brain evolution. Sinauer Associates. ISBN   978-0-87893-820-9.
  19. Sheline, Yvette; Yan, Shizi (2010). "Resting-state functional MRI in depression unmasks increased connectivity between networks via the dorsal nexus". PNAS. 107 (24): 11020–11025. Bibcode:2010PNAS..10711020S. doi: 10.1073/pnas.1000446107 . PMC   2890754 . PMID   20534464.
  20. 1 2 "Broca area | anatomy". Encyclopedia Britannica. Retrieved 2019-12-12.
  21. Alvarez JA, Emory E (March 2006). "Executive function and the frontal lobes: a meta-analytic review". Neuropsychology Review. 16 (1): 17–42. doi:10.1007/s11065-006-9002-x. PMID   16794878. S2CID   207222975.
  22. Goldman-Rakic PS (1988). "Topography of cognition: parallel distributed networks in primate association cortex". Annual Review of Neuroscience. 11: 137–156. doi:10.1146/annurev.ne.11.030188.001033. PMID   3284439.
  23. Price JL (June 1999). "Prefrontal cortical networks related to visceral function and mood". Annals of the New York Academy of Sciences. 877 (1): 383–396. Bibcode:1999NYASA.877..383P. doi:10.1111/j.1749-6632.1999.tb09278.x. PMID   10415660. S2CID   37564764.
  24. Robbins TW, Arnsten AF (2009). "The neuropsychopharmacology of fronto-executive function: monoaminergic modulation". Annual Review of Neuroscience. 32: 267–287. doi:10.1146/annurev.neuro.051508.135535. PMC   2863127 . PMID   19555290.
  25. Arnsten AF, Paspalas CD, Gamo NJ, Yang Y, Wang M (August 2010). "Dynamic Network Connectivity: A new form of neuroplasticity". Trends in Cognitive Sciences. 14 (8): 365–375. doi:10.1016/j.tics.2010.05.003. PMC   2914830 . PMID   20554470.
  26. 1 2 3 Mander BA, Rao V, Lu B, Saletin JM, Lindquist JR, Ancoli-Israel S, et al. (March 2013). "Prefrontal atrophy, disrupted NREM slow waves and impaired hippocampal-dependent memory in aging". Nature Neuroscience. 16 (3): 357–364. doi:10.1038/nn.3324. PMC   4286370 . PMID   23354332.
  27. Kuhl BA, Wagner AD (2009-01-01). "Strategic Control of Memory". In Squire LR (ed.). Encyclopedia of Neuroscience. Academic Press. pp. 437–444. doi:10.1016/b978-008045046-9.00424-1. ISBN   978-0-08-045046-9.
  28. 1 2 3 4 5 Hoffman P (January 2019). "Reductions in prefrontal activation predict off-topic utterances during speech production". Nature Communications. 10 (1): 515. Bibcode:2019NatCo..10..515H. doi:10.1038/s41467-019-08519-0. PMC   6355898 . PMID   30705284.
  29. Goldman-Rakic PS (October 1996). "The prefrontal landscape: implications of functional architecture for understanding human mentation and the central executive". Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 351 (1346): 1445–1453. doi:10.1098/rstb.1996.0129. JSTOR   3069191. PMID   8941956.
  30. Fuster JM, Bodner M, Kroger JK (May 2000). "Cross-modal and cross-temporal association in neurons of frontal cortex". Nature. 405 (6784): 347–351. Bibcode:2000Natur.405..347F. doi:10.1038/35012613. PMID   10830963. S2CID   4421762.
  31. Shimamura AP (2000). "The role of the prefrontal cortex in dynamic filtering". Psychobiology. 28 (2): 207–218. doi: 10.3758/BF03331979 . S2CID   140274181.
  32. 1 2 3 Miller EK, Cohen JD (2001). "An integrative theory of prefrontal cortex function". Annual Review of Neuroscience. 24: 167–202. doi:10.1146/annurev.neuro.24.1.167. PMID   11283309. S2CID   7301474.
  33. Muzur A, Pace-Schott EF, Hobson JA (November 2002). "The prefrontal cortex in sleep". Trends in Cognitive Sciences. 6 (11): 475–481. doi:10.1016/S1364-6613(02)01992-7. PMID   12457899. S2CID   5530174.
  34. Mitchell JP, Heatherton TF, Macrae CN (November 2002). "Distinct neural systems subserve person and object knowledge". Proceedings of the National Academy of Sciences of the United States of America. 99 (23): 15238–15243. Bibcode:2002PNAS...9915238M. doi: 10.1073/pnas.232395699 . PMC   137574 . PMID   12417766.
  35. Wang Y, Hamilton AF (April 2015). "Anterior medial prefrontal cortex implements social priming of mimicry". Social Cognitive and Affective Neuroscience. 10 (4): 486–493. doi:10.1093/scan/nsu076. PMC   4381231 . PMID   25009194.
  36. Schacter, Daniel L., Daniel Todd Gilbert, and Daniel M. Wegner. Psychology. 2nd ed, pages 364–366 New York, NY: Worth Publishers, 2011. Print.
  37. Yuan P, Raz N (May 2014). "Prefrontal cortex and executive functions in healthy adults: a meta-analysis of structural neuroimaging studies". Neuroscience and Biobehavioral Reviews. 42: 180–192. doi:10.1016/j.neubiorev.2014.02.005. PMC   4011981 . PMID   24568942.
  38. 1 2 Lebedev MA, Messinger A, Kralik JD, Wise SP (November 2004). "Representation of attended versus remembered locations in prefrontal cortex". PLOS Biology. 2 (11): e365. doi: 10.1371/journal.pbio.0020365 . PMC   524249 . PMID   15510225. Open Access logo PLoS transparent.svg
  39. Jacobsen C.F. (1936) Studies of cerebral function in primates. I. The functions of the frontal associations areas in monkeys. Comp Psychol Monogr 13: 3–60.
  40. Pribram KH, Mishkin M, Rosvold HE, Kaplan SJ (December 1952). "Effects on delayed-response performance of lesions of dorsolateral and ventromedial frontal cortex of baboons". Journal of Comparative and Physiological Psychology. 45 (6): 565–575. doi:10.1037/h0061240. PMID   13000029.
  41. Funahashi S, Bruce CJ, Goldman-Rakic PS (April 1993). "Dorsolateral prefrontal lesions and oculomotor delayed-response performance: evidence for mnemonic "scotomas"". The Journal of Neuroscience. 13 (4): 1479–1497. doi:10.1523/JNEUROSCI.13-04-01479.1993. PMC   6576716 . PMID   8463830.
  42. Baddeley A. (1986) Working memory. Oxford: Oxford University Press. p.289
  43. Bedini M, Baldauf D (August 2021). "Structure, function and connectivity fingerprints of the frontal eye field versus the inferior frontal junction: A comprehensive comparison". The European Journal of Neuroscience. 54 (4): 5462–5506. doi:10.1111/ejn.15393. PMC   9291791 . PMID   34273134. S2CID   235999643.
  44. Liston C, Miller MM, Goldwater DS, Radley JJ, Rocher AB, Hof PR, et al. (July 2006). "Stress-induced alterations in prefrontal cortical dendritic morphology predict selective impairments in perceptual attentional set-shifting". The Journal of Neuroscience. 26 (30): 7870–7874. doi:10.1523/JNEUROSCI.1184-06.2006. PMC   6674229 . PMID   16870732.
  45. "Viewers of pornography have a smaller reward system". MAX-PLANCK-GESELLSCHAFT. 2 June 2014. Retrieved 2 July 2018.
  46. Rajkowska G (December 1997). "Morphometric methods for studying the prefrontal cortex in suicide victims and psychiatric patients". Annals of the New York Academy of Sciences. 836 (1): 253–268. Bibcode:1997NYASA.836..253R. doi:10.1111/j.1749-6632.1997.tb52364.x. PMID   9616803. S2CID   32947726.
  47. Cecil KM, Brubaker CJ, Adler CM, Dietrich KN, Altaye M, Egelhoff JC, et al. (May 2008). Balmes J (ed.). "Decreased brain volume in adults with childhood lead exposure". PLOS Medicine. 5 (5): e112. doi: 10.1371/journal.pmed.0050112 . PMC   2689675 . PMID   18507499. Open Access logo PLoS transparent.svg
  48. Anderson SW, Bechara A, Damasio H, Tranel D, Damasio AR (November 1999). "Impairment of social and moral behavior related to early damage in human prefrontal cortex". Nature Neuroscience. 2 (11): 1032–1037. doi:10.1038/14833. PMID   10526345. S2CID   204990285.
  49. Fariba K, Gokarakonda SB (2021). "Impulse Control Disorders". StatPearls. Treasure Island (FL): StatPearls Publishing. PMID   32965950 . Retrieved 2021-05-04.
  50. Schoenemann PT, Budinger TF, Sarich VM, Wang WS (April 2000). "Brain size does not predict general cognitive ability within families". Proceedings of the National Academy of Sciences of the United States of America. 97 (9): 4932–4937. Bibcode:2000PNAS...97.4932S. doi: 10.1073/pnas.97.9.4932 . PMC   18335 . PMID   10781101.
  51. Cascio T. "House & Psychology, Episode 14". Psychology Today. Archived from the original on 2013-04-08. Retrieved 2011-11-15.
  52. Guiney H, Machado L (February 2013). "Benefits of regular aerobic exercise for executive functioning in healthy populations". Psychonomic Bulletin & Review. 20 (1): 73–86. doi: 10.3758/s13423-012-0345-4 . PMID   23229442. S2CID   24190840.
  53. Erickson KI, Leckie RL, Weinstein AM (September 2014). "Physical activity, fitness, and gray matter volume". Neurobiology of Aging. 35 (Suppl 2): S20–S28. doi:10.1016/j.neurobiolaging.2014.03.034. PMC   4094356 . PMID   24952993.
  54. Valkanova V, Eguia Rodriguez R, Ebmeier KP (June 2014). "Mind over matter--what do we know about neuroplasticity in adults?". International Psychogeriatrics. 26 (6): 891–909. doi:10.1017/S1041610213002482. PMID   24382194. S2CID   20765865.
  55. Fowler AK, Thompson J, Chen L, Dagda M, Dertien J, Dossou KS, et al. (2014). "Differential sensitivity of prefrontal cortex and hippocampus to alcohol-induced toxicity". PLOS ONE. 9 (9): e106945. Bibcode:2014PLoSO...9j6945F. doi: 10.1371/journal.pone.0106945 . PMC   4154772 . PMID   25188266.
  56. Antonio Damasio, Descartes' Error . Penguin Putman Pub., 1994[ page needed ]
  57. Malcolm Macmillan, An Odd Kind of Fame: Stories of Phineas Gage (MIT Press, 2000), pp.116–119, 307–333, esp. pp.11,333.
  58. Macmillan M (2008). "Phineas Gage – Unravelling the myth". The Psychologist. 21 (9). British Psychological Society: 828–831. Archived from the original on 2010-09-03. Retrieved 2014-06-21.
  59. Wang M, Ramos BP, Paspalas CD, Shu Y, Simen A, Duque A, et al. (April 2007). "Alpha2A-adrenoceptors strengthen working memory networks by inhibiting cAMP-HCN channel signaling in prefrontal cortex". Cell. 129 (2): 397–410. doi: 10.1016/j.cell.2007.03.015 . PMID   17448997. S2CID   741677.