Affective neuroscience

Last updated

Affective neuroscience is the study of how the brain processes emotions. This field combines neuroscience with the psychological study of personality, emotion, and mood. [1] The basis of emotions and what emotions are remains an issue of debate within the field of affective neuroscience. [2]

Contents

The term "affective neuroscience" was coined by neuroscientist Jaak Panksepp, at a time when cognitive neuroscience focused on parts of psychology that did not include emotion, such as attention or memory. [3]

Affective neuroscience

Emotions are thought to be related to activity in brain areas that direct our attention, motivate our behavior, and help us make decisions about our environment. Early stages of research on emotions and the brain was conducted by Paul Broca, [4] James Papez, [5] and Paul D. MacLean. [6] Their work suggests that emotion is related to a group of structures in the center of the brain called the limbic system. The limbic system is made up of the following brain structures:

Limbic system

Research has shown the limbic system is directly related to emotion, but there are other brain areas and structures that are important for producing and processing emotion. [22]

Other brain structures

Right hemisphere

Many theories about the role of the right hemisphere in emotion has resulted in several models of emotional functioning.After observing decreased emotional processing after right hemisphere injuries, C.K. Mills hypothesized emotions are directly related to the right hemisphere. [42] [43] In 1992, researchers found that emotional expression and understanding may be controlled by smaller brain structures in the right hemisphere. [44] These findings were the basis for the right hemisphere hypothesis and the valence hypothesis.

Right hemisphere hypothesis

It is believed that the right hemisphere is more specialized in processing emotions than the left hemisphere. [45] The right hemisphere is associated with nonverbal, synthetic, integrative, holistic and gestaltic mental strategies. [44] As demonstrated by patients who have increased spatial neglect when damage affects the right brain rather than the left brain, the right hemisphere is more connected to subcortical systems of autonomic arousal and attention. [46] Right hemisphere disorders have been associated with abnormal patterns of autonomic nervous system responses. [47] These findings suggest the right hemisphere and subcortical brain areas are closely related.

Valence hypothesis

According to the valence hypothesis, although the right hemisphere is involved in emotion, it is primarily involved in the processing of negative emotions, while the left hemisphere is involved in processing positive emotions. In one explanation, negative emotions are processed by the right brain, while positive emotions are processed by the left. [48] An alternative explanation is that the right hemisphere is dominant when it comes to feeling both positive and negative emotions. [49] [50] Recent studies indicate that the frontal lobes of both hemispheres play an active role in emotions, while the parietal and temporal lobes process them. [51] Depression has been associated with decreased right parietal lobe activity, while anxiety has been associated with increased right parietal lobe activity. [52] Based on the original valence model, increasingly complex models have been developed as a result of the increasing understanding of the different hemispheres. [53]

Cognitive neuroscience

While emotions are integral to thought processes, cognition has been investigated without emotion until the late 1990s, focusing instead on non-emotional processes such as memory, attention, perception, problem solving, and mental imagery. [54] Cognitive neuroscience and affective neuroscience have emerged as separate fields for studying the neural basis of non-emotional and emotional processes. Despite the fact that fields are classified according to how the brain processes cognition and emotion, the neural and mental mechanisms behind emotional and non-emotional processes often overlap. [55]

Cognitive neuroscience tasks in affective neuroscience research

Emotion go/no-go

Emotion go/no-go tasks are used to study behavioral inhibition, especially how it is influenced by emotion. [56] A "go" cue tells the participant to respond rapidly, but a "no-go" cue tells them to withhold a response. Because the "go" cue occurs more frequently, it can be used to measure how well a subject suppresses a response under different emotional conditions. [57]

This task is often used in combination with neuroimaging in healthy individuals and patients with affective disorders to identify relevant brain functions associated with emotional regulation. [56] [58] [59] Several studies, including go/no-go studies, suggest that sections of the prefrontal cortex are involved in controlling emotional responses to stimuli during inhibition. [60]

Emotional Stroop

Adapted from the Stroop, the emotional Stroop test measures how much attention you pay to emotional stimuli. [61] [62] In this task, participants are instructed to name the ink color of words while ignoring their meanings. [63] Generally, people have trouble detaching their attention from words with an affective meaning compared with neutral words. [64] [65] It has been demonstrated in several studies that naming the color of neutral words results in a quicker response. [62]

Selective attention to negative or threatening stimuli, which are often related to psychological disorders, is commonly tested with this task. [66] Different mental disorders have been associated with specific attentional biases. [66] [67] Participants with spider phobia, for example, tend to be more inclined to use spider-related words than negatively charged words. [68] Similar findings have been found for threat words related to other anxiety disorders. [66] Even so, other studies have questioned these conclusions. When the words are matched for emotionality, anxious participants in some studies show the Stroop interference effect for both negative and positive words. [69] [70] In other words, the specificity effects of words for various disorders may be primarily due to their conceptual relation to the disorder's concerns rather than their emotionality. [66]

Ekman 60 faces task

The Ekman faces task is used to measure emotion recognition of six basic emotions. [71] [72] Black and white photographs of 10 actors (6 male, 4 female) are presented, with each actor displaying each emotion. Participants are usually asked to respond quickly with the name of the displayed emotion. The task is a common tool to study deficits in emotion regulation in patients with dementia, Parkinson's, and other cognitively degenerative disorders. [73] The task has been used to analyze recognition errors in disorders such as borderline personality disorder, schizophrenia, and bipolar disorder. [74] [75] [76]

Dot probe (emotion)

The emotional dot-probe paradigm is a task used to assess selective visual attention to and failure to detach attention from affective stimuli. [77] [78] The paradigm begins with a fixation cross at the center of a screen. An emotional stimulus and a neutral stimulus appear side by side, after which a dot appears behind either the neutral stimulus (incongruent condition) or the affective stimulus (congruent condition). Participants are asked to indicate when they see this dot, and response latency is measured. Dots that appear on the same side of the screen as the image the participant was looking at will be identified more quickly. Thus, it is possible to discern which object the participant was attending to by subtracting the reaction time to respond to congruent versus incongruent trials. [77]

The best documented research with the dot probe paradigm involves attention to threat related stimuli, such as fearful faces, in individuals with anxiety disorders. Anxious individuals tend to respond more quickly to congruent trials, which may indicate vigilance to threat and/or failure to detach attention from threatening stimuli. [77] [79] A specificity effect of attention has also been noted, with individuals attending selectively to threats related to their particular disorder. For example, those with social phobia selectively attend to social threats but not physical threats. [80] However, this specificity may be even more nuanced. Participants with obsessive-compulsive disorder symptoms initially show attentional bias to compulsive threat, but this bias is attenuated in later trials due to habituation to the threat stimuli. [81]

Fear potentiated startle

Fear-potentiated startle (FPS) has been utilized as a psychophysiological index of fear reaction in both animals and humans. [82] FPS is most often assessed through the magnitude of the eyeblink startle reflex, which can be measured by electromyography. [83] This eyeblink reflex is an automatic defensive reaction to an abrupt elicitor, making it an objective indicator of fear. [84] Typical FPS paradigms involve bursts of noise or abrupt flashes of light transmitted while an individual attends to a set of stimuli. [84] Startle reflexes have been shown to be modulated by emotion. For example, healthy participants tend to show enhanced startle responses while viewing negatively valenced images and attenuated startle while viewing positively valenced images, as compared with neutral images. [85] [86]

The startle response to a particular stimulus is greater under conditions of threat. [87] A common example given to indicate this phenomenon is that one's startle response to a flash of light will be greater when walking in a dangerous neighborhood at night than it would under safer conditions. In laboratory studies, the threat of receiving shock is enough to potentiate startle, even without any actual shock. [88]

Fear potentiated startle paradigms are often used to study fear learning and extinction in individuals with post-traumatic stress disorder (PTSD) and other anxiety disorders. [89] [90] [91] In fear conditioning studies, an initially neutral stimulus is repeatedly paired with an aversive one, borrowing from classical conditioning. [92] FPS studies have demonstrated that PTSD patients have enhanced startle responses during both danger cues and neutral/safety cues as compared with healthy participants. [92] [93]

Learning

Affect plays many roles during learning. Deep, emotional attachment to a subject area allows a deeper understanding of the material and therefore, learning occurs and lasts. [94] The emotions evoked when reading in comparison to the emotions portrayed in the content affects comprehension. Someone who is feeling sad understands a sad passage better than someone feeling happy. [95] Therefore, a student's emotion plays an important role during the learning process.

Emotion can be embodied or perceived from words read on a page or in a facial expression. Neuroimaging studies using fMRI have demonstrated that the same area of the brain that is activated when feeling disgust is activated when observing another's disgust. [96] In a traditional learning environment, the teacher's facial expression can play a critical role in language acquisition. Showing a fearful facial expression when reading passages that contain fearful tones facilitates students learning of the meaning of certain vocabulary words and comprehension of the passage. [97]

Models

The neurobiological basis of emotion is still disputed. [98] The existence of basic emotions and their defining attributes represents a long lasting and yet unsettled issue in psychology. [98] The available research suggests that the neurobiological existence of basic emotions is still tenable and heuristically seminal, pending some reformulation. [98]

Basic emotions

These approaches hypothesize that emotion categories (including happiness, sadness, fear, anger, and disgust) are biologically basic. [99] [100] In this view, emotions are inherited, biologically based modules that cannot be separated into more basic psychological components. [99] [100] [101] Models following this approach hypothesize that all mental states belonging to a single emotional category can be consistently and specifically localized to either a single brain region or a defined network of brain regions. [100] [102] Each basic emotion category also shares other universal characteristics: distinct facial behavior, physiology, subjective experience and accompanying thoughts and memories. [99]

Psychological constructionist approaches

This approach to emotion hypothesizes that emotions like happiness, sadness, fear, anger and disgust (and many others) are constructed mental states that occur when brain systems work together. [103] In this view, networks of brain regions underlie psychological operations (e.g., language, attention, etc.) that interact to produce emotion, perception, and cognition. [104] One psychological operation critical for emotion is the network of brain regions that underlie valence (feeling pleasant/unpleasant) and arousal (feeling activated and energized). [103] Emotions emerge when neural systems underlying different psychological operations interact (not just those involved in valence and arousal), producing distributed patterns of activation across the brain. Because emotions emerge from more basic components, heterogeneity affects each emotion category; for example, a person can experience many different kinds of fear, which feel differently, and which correspond to different neural patterns in the brain. [105]

Aging

People typically associate aging with a decline in the functioning of all mental processing abilities; however, this is not the case for emotion regulation. Older adults typically have a stronger drive to maintain and improve on their emotional well being. [106] Thus providing them to utilize emotion regulation skills that provide a higher satisfaction in life.

Role of the vmPFC in emotion regulation of older adults

The ventromedial prefrontal cortex (vmPFC) has a significant influence on emotion regulation, especially regarding high emotional arousing stimuli. [107] Compared to other areas of the prefrontal cortex (PFC), the vmPFC loses volume at a much lower rate . Due to this, an older persons emotion regulation abilities are not very heavily impacted by brain changes associated with aging. Additionally, the anterior cingulate cortex (ACC) is an important area of the brain that is used for emotion regulation. The ACC has proven to be a key player in emotion regulation in not just young adults, but also in older adults. [106] In older adults the ACC is important to create connections with from the vmPFC in order to regulate emotions. This connection was the most salient when negative emotions were reappraised. This demonstrates that older adults use the vmPFC to regulate their emotions in a more positive manner. Despite other areas of the brain decreasing in functionality as humans age, the connection between the vmPFC and ACC remains strong to reappraise negative emotions into more positive emotions. This is different from younger adults who rely on a more on other areas of the PFC.

Neuropsychology behind older adults emotion regulation differences

As people age, most cognitive functions decline. This is not the case when it comes to emotion regulation. A study conducted by Carstensen and colleagues (2000) found that as people increase in age so does their ability to regulate their emotions. [108] It is important to note that just because older adults had better emotion regulation skills, does not mean they live more stable daily lives. In fact, they tend to have more unstable negative emotions especially in comparison to the stability of their positive emotions. [109] The major difference observed in how older adults and younger adults regulate their emotions when negative emotional stimuli are present can be explained by numerous theories.

Theories of emotion regulation in aging

Passive method of emotion regulation

How older adults handle emotionally salient events or stimuli are often vastly different from younger adults, and even middle aged adults. There does not appear to be many differences in ways that younger, middle, and older adults handle social situations; however, when a social situation becomes emotionally charged differences emerge. [110] When intense emotions in a social situation were evoked for older adults, they tended to the situation in a more passive manner in comparison to middle aged adults. They also tend to rely more on their previous problem solving skills than both younger and older adults. This is because as people age, there tends to be a shift in preferences to maintain a more positive emotional affect. In fact, there seems to be a decrease in negative emotions felt by older adults once until they reach the age of 60, in which this decrease stops. [108] It is important to note that while the frequency of negative emotion decreases with age, the intensity of the emotions experienced does not change. Additionally, emotional satisfaction is not lower just because they experience less frequent negative emotions.

Socioemotional selectivity theory

Carstensen (2003) hypothesized that the reason that older adults tended to have better emotion regulation skills than younger adults is due to the socioemotional selectivity theory. [111] This theory highlights the role of social interactions in the ability to regulate emotions. Social interactions, while are often positive, can sometimes lead to negative emotional arousal. Since older adults have been alive longer, they have more dense social networks. This creates a drastic increase in social interaction that cause positive emotional arousal. On the chance they experience a negative emotional reaction from a social event, they are likely to be able to pair it with something that is more positively emotionally salient. This causes the negative emotion to be less potent, and therefore increase their hedonic perspective on life.

Meta-analyses

A meta-analysis is a statistical approach to synthesizing results across multiple studies. Included studies investigated healthy, unmedicated adults and that used subtraction analysis to examine brain areas that were more active during emotional processing than during a neutral (control) condition.

Phan et al. 2002

In the first neuroimaging meta-analysis of emotion, Phan et al. (2002) analyzed the results of 55 peer reviewed studies between January 1990 and December 2000 to determine if the emotions of fear, sadness, disgust, anger, and happiness were consistently associated with activity in specific brain regions. All studies used fMRI or PET techniques to investigate higher-order mental processing of emotion (studies of low-order sensory or motor processes were excluded). The authors' tabulated the number of studies that reported activation in specific brain regions. For each brain region, statistical chi-squared analysis was conducted. Two regions showed a statistically significant association. In the amygdala, 66% of studies inducing fear reported activity in this region, as compared to ~20% of studies inducing happiness, ~15% of studies inducing sadness (with no reported activations for anger or disgust). In the subcallosal cingulate, 46% of studies inducing sadness reported activity in this region, as compared to ~20% inducing happiness and ~20% inducing anger. This pattern of clear discriminability between emotion categories was in fact rare, with other patterns occurring in limbic regions, paralimbic regions, and uni/heteromodal regions. Brain regions implicated across discrete emotion included the basal ganglia (~60% of studies inducing happiness and ~60% of studies inducing disgust reported activity in this region) and medial prefrontal cortex (happiness ~60%, anger ~55%, sadness ~40%, disgust ~40%, and fear ~30%). [112]

Murphy et al. 2003

Murphy, et al. 2003 analyzed 106 peer reviewed studies published between January 1994 and December 2001 to examine the evidence for regional specialization of discrete emotions (fear, disgust, anger, happiness and sadness) across a larger set of studies. Studies included in the meta-analysis measured activity in the whole brain and regions of interest (activity in individual regions of particular interest to the study). 3-D Kolmogorov-Smirnov (KS3) statistics were used to compare rough spatial distributions of 3-D activation patterns to determine if statistically significant activations were specific to particular brain regions for all emotional categories. This pattern of consistently activated, regionally specific activations was identified in four brain regions: amygdala with fear (~40% of studies), insula with disgust (~70%), globus pallidus with disgust (~70%), and lateral orbitofrontal cortex with anger (80%). Other regions showed different patterns of activation across categories. For example, both the dorsal medial prefrontal cortex and the rostral anterior cingulate cortex showed consistent activity across emotions (happiness ~50%, sadness ~50%, anger ~ 40%, fear ~30%, and disgust ~ 20%). [113]

Barrett et al. 2006

Barrett, et al. 2006 examined 161 studies published between 1990 and 2001. The authors compared the consistency and specificity of prior meta-analytic findings specific to each notional basic emotion. Consistent neural patterns were defined by brain regions showing increased activity for a specific emotion (relative to a neutral control condition), regardless of the method of induction used (for example, visual vs. auditory cue). Specific neural patterns were defined as separate circuits for one emotion vs. the other emotions (for example, the fear circuit must be discriminable from the anger circuit, although both may include common brain regions). In general, the results supported Phan et al. and Murphy et al., but not specificity. Consistency was determined through the comparison of chi-squared analyses that revealed whether the proportion of studies reporting activation during one emotion was significantly higher than the proportion of studies reporting activation during the other emotions. Specificity was determined through the comparison of emotion-category brain-localizations by contrasting activations in key regions that were specific to particular emotions. Increased amygdala activation during fear was the most consistently reported across induction methods (but not specific). Both meta-analyses associated the anterior cingulate cortex with sadness, although this finding was less consistent (across induction methods) and was not specific. Both meta-analyses found that disgust was associated with the basal ganglia, but these findings were neither consistent nor specific. Neither consistent nor specific activity was observed across the meta-analyses for anger or happiness. This meta-analysis introduced the concept of the basic, irreducible elements of emotional life as dimensions such as approach and avoidance. [103]

Kober et al. 2008

Kober reviewed 162 neuroimaging studies published between 1990 and 2005 in order to determine if specific brain regions were activated when experiencing an emotion directly and (indirectly) through the experience of someone else. [114] According to the study, six different functional groups showed similar activation patterns. The psychological functions of each group were discussed in more basic terms. These regions may also play a role in processing visual information and paying attention to emotional signals. [115]

Groups
GroupRegionsNotes
Core limbicleft amygdala, hypothalamus, periaqueductal gray/thalamus regions, and amygdala/ventral striatum/ventral globus pallidus/thalamus regionsA functional emotional center responsible for evaluating affective significance.
Lateral Paralimbicventral anterior insula/frontal operculum/right temporal pole/ posterior orbitofrontal cortex, the anterior insula/ posterior orbitofrontal cortex, the ventral anterior insula/ temporal cortex/ orbitofrontal cortex junction, the midinsula/ dorsal putamen, and the ventral striatum /mid insula/ left hippocampusContributes to motivation, especially in reward, by making stimuli more valuable.
Medial Prefrontal Cortexdorsal medial prefrontal cortex, pregenual anterior cingulate cortex, and rostral dorsal anterior cingulate cortexPlays a role in both the generation and regulation of emotion.
Cognitive/ Motor Networkright frontal operculum, the right interior frontal gyrus, and the pre-supplementray motor area/ left interior frontal gyrus, regionsPlays a general role in information processing and cognitive control.
Occipital/ Visual AssociationV8 and V4 areas of the primary visual cortex, the medial temporal lobe, and the lateral occipital cortex
Medial Posteriorposterior cingulate cortex and area V1 of the primary visual cortex

Vytal et al. 2010

Vytal, et al. 2010 examined 83 neuroimaging studies published between 1993–2008 to examine whether neuroimaging evidence supports biologically discrete, basic emotions (i.e. fear, anger, disgust, happiness, and sadness). Consistency analyses identified brain regions associated with individual emotions. Discriminability analyses identified brain regions that were differentially active under contrasting pairs of emotions. This meta-analysis examined PET or fMRI studies that reported whole brain analyses identifying significant activations for at least one of the five emotions relative to a neutral or control condition. The authors used activation likelihood estimation (ALE) to perform spatially sensitive, voxel-wise (sensitive to the spatial properties of voxels) statistical comparisons across studies. This technique allows for direct statistical comparison between activation maps associated with each discrete emotion. Thus, discriminability between the five discrete emotion categories was assessed on a more precise spatial scale than in prior meta-analyses.

Consistency was first assessed by comparing the cross-study ALE map for each emotion to ALE maps generated by random permutations. Discriminability was assessed by pair-wise contrasts of emotion maps. Consistent and discriminable activation patterns were observed for the five categories.

Associations [116]
EmotionPeakRegions
Happinessright superior temporal gyrus, left rostral anterior cingulate cortex9 regional brain clusters
Sadnessleft medial frontal gyrus35 clusters - especially, left medial frontal gyrus, right middle temporal gyrus, and right inferior frontal gyrus
Angerleft inferior frontal gyrus13 clusters - bilateral inferior frontal gyrus, and in right parahippocampal gyrus
Fearleft amygdala11 clusters - left amygdala and left putamen
Disgustright insula/ right inferior frontal gyrus16 clusters - right putamen and the left insula.

Lindquist et al. 2012

Lindquist, et al. reviewed 91 PET and fMRI studies published between January 1990 and December 2007. Induction methods were used to elicit fear, sadness, disgust, anger, and happiness. The goal was to compare basic emotions approaches with psychological constructionist approaches. [114]

It was found that many brain regions activated consistently or selectively for one emotion category when experienced or perceived. As predicted by constructionist models, no region demonstrated functional specificity for fear, disgust, happiness, sadness, or anger.

The authors suggest that certain brain areas traditionally assigned to certain emotions are incorrect and instead correspond to different emotion categories. There is some evidence that the amygdala, anterior insula, and orbitofrontal cortex all contribute to "core affect", which are feelings of pleasure or discomfort.

Core affects
RegionRole
AmygdalaIdentifying whether external sensory information is motivationally relevant, new, or evokes uncertainty
Anterior insulaDescribes the core affective feelings, mostly driven by body sensations, across all emotion categories
Orbitofrontal cortexGuides behavior by combining sensory information from the body and our environment

The anterior cingulate and the dorsolateral prefrontal cortex play a key role in attention, which is closely related to core affect. By using sensory information, the anterior cingulate directs attention and motor responses. According to psychological constructionist theory, emotions are conceptualizations connecting the world and the body, and the dorsolateral prefrontal cortex facilitates executive attention. As well as playing an active role in conceptualizing, the prefrontal cortex and hippocampus also simulate previous experiences. In several studies, the ventrolateral prefrontal cortex, which supports language, was consistently active during emotion perception and experience. [104]

See also

Related Research Articles

<span class="mw-page-title-main">Amygdala</span> Each of two small structures deep within the temporal lobe of complex vertebrates

The amygdala is one of two almond-shaped clusters of nuclei located deep and medially within the temporal lobes of the brain's cerebrum in complex vertebrates, including humans. Shown to perform a primary role in the processing of memory, decision making, and emotional responses, the amygdalae are considered part of the limbic system. The term "amygdala" was first introduced by Karl Friedrich Burdach in 1822.

<span class="mw-page-title-main">Anterior cingulate cortex</span> Brain region

In the human brain, the anterior cingulate cortex (ACC) is the frontal part of the cingulate cortex that resembles a "collar" surrounding the frontal part of the corpus callosum. It consists of Brodmann areas 24, 32, and 33.

<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. It contributes to the dorsolateral and medial prefrontal cortex.

In animals, including humans, the startle response is a largely unconscious defensive response to sudden or threatening stimuli, such as sudden noise or sharp movement, and is associated with negative affect. Usually the onset of the startle response is a startle reflex reaction. The startle reflex is a brainstem reflectory reaction (reflex) that serves to protect vulnerable parts, such as the back of the neck and the eyes (eyeblink) and facilitates escape from sudden stimuli. It is found across many different species, throughout all stages of life. A variety of responses may occur depending on the affected individual's emotional state, body posture, preparation for execution of a motor task, or other activities. The startle response is implicated in the formation of specific phobias.

Reduced affect display, sometimes referred to as emotional blunting or emotional numbing, is a condition of reduced emotional reactivity in an individual. It manifests as a failure to express feelings either verbally or nonverbally, especially when talking about issues that would normally be expected to engage emotions. In this condition, expressive gestures are rare and there is little animation in facial expression or vocal inflection. Additionally, reduced affect can be symptomatic of autism, schizophrenia, depression, post-traumatic stress disorder, depersonalization disorder, schizoid personality disorder or brain damage. It may also be a side effect of certain medications.

<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">Posterior cingulate cortex</span> Caudal part of the cingulate cortex of the brain

The posterior cingulate cortex (PCC) is the caudal part of the cingulate cortex, located posterior to the anterior cingulate cortex. This is the upper part of the "limbic lobe". The cingulate cortex is made up of an area around the midline of the brain. Surrounding areas include the retrosplenial cortex and the precuneus.

<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.

Memory and trauma is the deleterious effects that physical or psychological trauma has on memory.

Scientific studies have found that different brain areas show altered activity in humans with major depressive disorder (MDD), and this has encouraged advocates of various theories that seek to identify a biochemical origin of the disease, as opposed to theories that emphasize psychological or situational causes. Factors spanning these causative groups include nutritional deficiencies in magnesium, vitamin D, and tryptophan with situational origin but biological impact. Several theories concerning the biologically based cause of depression have been suggested over the years, including theories revolving around monoamine neurotransmitters, neuroplasticity, neurogenesis, inflammation and the circadian rhythm. Physical illnesses, including hypothyroidism and mitochondrial disease, can also trigger depressive symptoms.

The biology of obsessive–compulsive disorder (OCD) refers biologically based theories about the mechanism of OCD. Cognitive models generally fall into the category of executive dysfunction or modulatory control. Neuroanatomically, functional and structural neuroimaging studies implicate the prefrontal cortex (PFC), basal ganglia (BG), insula, and posterior cingulate cortex (PCC). Genetic and neurochemical studies implicate glutamate and monoamine neurotransmitters, especially serotonin and dopamine.

<span class="mw-page-title-main">Brain activity and meditation</span>

Meditation and its effect on brain activity and the central nervous system became a focus of collaborative research in neuroscience, psychology and neurobiology during the latter half of the 20th century. Research on meditation sought to define and characterize various practices. The effects of meditation on the brain can be broken up into two categories: state changes and trait changes, respectively alterations in brain activities during the act of meditating and changes that are the outcome of long-term practice.

<span class="mw-page-title-main">Parental brain</span>

Parental experience, as well as changing hormone levels during pregnancy and postpartum, cause changes in the parental brain. Displaying maternal sensitivity towards infant cues, processing those cues and being motivated to engage socially with her infant and attend to the infant's needs in any context could be described as mothering behavior and is regulated by many systems in the maternal brain. Research has shown that hormones such as oxytocin, prolactin, estradiol and progesterone are essential for the onset and the maintenance of maternal behavior in rats, and other mammals as well. Mothering behavior has also been classified within the basic drives.

Pain empathy is a specific variety of empathy that involves recognizing and understanding another person's pain.

Emotion perception refers to the capacities and abilities of recognizing and identifying emotions in others, in addition to biological and physiological processes involved. Emotions are typically viewed as having three components: subjective experience, physical changes, and cognitive appraisal; emotion perception is the ability to make accurate decisions about another's subjective experience by interpreting their physical changes through sensory systems responsible for converting these observed changes into mental representations. The ability to perceive emotion is believed to be both innate and subject to environmental influence and is also a critical component in social interactions. How emotion is experienced and interpreted depends on how it is perceived. Likewise, how emotion is perceived is dependent on past experiences and interpretations. Emotion can be accurately perceived in humans. Emotions can be perceived visually, audibly, through smell and also through bodily sensations and this process is believed to be different from the perception of non-emotional material.

<span class="mw-page-title-main">Mechanisms of mindfulness meditation</span>

Mindfulness has been defined in modern psychological terms as "paying attention to relevant aspects of experience in a nonjudgmental manner", and maintaining attention on present moment experience with an attitude of openness and acceptance. Meditation is a platform used to achieve mindfulness. Both practices, mindfulness and meditation, have been "directly inspired from the Buddhist tradition" and have been widely promoted by Jon Kabat-Zinn. Mindfulness meditation has been shown to have a positive impact on several psychiatric problems such as depression and therefore has formed the basis of mindfulness programs such as mindfulness-based cognitive therapy, mindfulness-based stress reduction and mindfulness-based pain management. The applications of mindfulness meditation are well established, however the mechanisms that underlie this practice are yet to be fully understood. Many tests and studies on soldiers with PTSD have shown tremendous positive results in decreasing stress levels and being able to cope with problems of the past, paving the way for more tests and studies to normalize and accept mindful based meditation and research, not only for soldiers with PTSD, but numerous mental inabilities or disabilities.

Neuromorality is an emerging field of neuroscience that studies the connection between morality and neuronal function. Scientists use fMRI and psychological assessment together to investigate the neural basis of moral cognition and behavior. Evidence shows that the central hub of morality is the prefrontal cortex guiding activity to other nodes of the neuromoral network. A spectrum of functional characteristics within this network to give rise to both altruistic and psychopathological behavior. Evidence from the investigation of neuromorality has applications in both clinical neuropsychiatry and forensic neuropsychiatry.

<span class="mw-page-title-main">Biology of bipolar disorder</span> Biological Study Of Bipolar Disorder

Bipolar disorder is an affective disorder characterized by periods of elevated and depressed mood. The cause and mechanism of bipolar disorder is not yet known, and the study of its biological origins is ongoing. Although no single gene causes the disorder, a number of genes are linked to increase risk of the disorder, and various gene environment interactions may play a role in predisposing individuals to developing bipolar disorder. Neuroimaging and postmortem studies have found abnormalities in a variety of brain regions, and most commonly implicated regions include the ventral prefrontal cortex and amygdala. Dysfunction in emotional circuits located in these regions have been hypothesized as a mechanism for bipolar disorder. A number of lines of evidence suggests abnormalities in neurotransmission, intracellular signalling, and cellular functioning as possibly playing a role in bipolar disorder.

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.

Affect labeling is an implicit emotional regulation strategy that can be simply described as "putting feelings into words". Specifically, it refers to the idea that explicitly labeling one's, typically negative, emotional state results in a reduction of the conscious experience, physiological response, and/or behavior resulting from that emotional state. For example, writing about a negative experience in one's journal may improve one's mood. Some other examples of affect labeling include discussing one's feelings with a therapist, complaining to friends about a negative experience, posting one's feelings on social media or acknowledging the scary aspects of a situation.

References

  1. Panksepp, Jaak (1990), "A Role for Affective Neuroscience in Understanding Stress: The Case of Separation Distress Circuitry", Psychobiology of Stress, Dordrecht: Springer Netherlands, pp. 41–57, doi:10.1007/978-94-009-1990-7_4, ISBN   978-94-010-7390-5 , retrieved 2022-11-27
  2. Celeghin, Alessia; Diano, Matteo; Bagnis, Arianna; Viola, Marco; Tamietto, Marco (2017-08-24). "Basic Emotions in Human Neuroscience: Neuroimaging and Beyond". Frontiers in Psychology. 8: 1432. doi: 10.3389/fpsyg.2017.01432 . ISSN   1664-1078. PMC   5573709 . PMID   28883803.
  3. Panksepp, Jaak (2004). Affective Neuroscience: The Foundations of Human and Animal Emotions (1st ed.). Oxford University Press. ISBN   978-0195178050.
  4. Broca, Paul (1878). "Anatomie comparée des circonvolutions cérébrales: le grand lobe limbique". Rev. Anthropology. 1: 385–498.
  5. PAPEZ, JAMES W. (1937-10-01). "A Proposed Mechanism of Emotion". Archives of Neurology and Psychiatry. 38 (4): 725. doi:10.1001/archneurpsyc.1937.02260220069003. ISSN   0096-6754.
  6. MacLean, Paul D. (1952). "Some psychiatric implications of physiological studies on frontotemporal portion of limbic system (Visceral brain)". Electroencephalography and Clinical Neurophysiology. 4 (4): 407–418. doi:10.1016/0013-4694(52)90073-4. ISSN   0013-4694. PMID   12998590.
  7. LeDoux, Joseph E. (1995). "Emotion: Clues from the Brain". Annual Review of Psychology. 46 (1): 209–235. doi:10.1146/annurev.ps.46.020195.001233. ISSN   0066-4308. PMID   7872730.
  8. Breiter, Hans C; Etcoff, Nancy L; Whalen, Paul J; Kennedy, William A; Rauch, Scott L; Buckner, Randy L; Strauss, Monica M; Hyman, Steven E; Rosen, Bruce R (1996). "Response and Habituation of the Human Amygdala during Visual Processing of Facial Expression". Neuron. 17 (5): 875–887. doi: 10.1016/S0896-6273(00)80219-6 . PMID   8938120. S2CID   17284478.
  9. Sherman, S. (2006). "Thalamus". Scholarpedia. 1 (9): 1583. Bibcode:2006SchpJ...1.1583S. doi: 10.4249/scholarpedia.1583 . ISSN   1941-6016.
  10. Steriade, M; Llinás, R R (1988). "The functional states of the thalamus and the associated neuronal interplay". Physiological Reviews. 68 (3): 649–742. doi:10.1152/physrev.1988.68.3.649. ISSN   0031-9333. PMID   2839857.
  11. The Cambridge handbook of human affective neuroscience. Jorge Armony, Patrik Vuilleumier. Cambridge: Cambridge University Press. 2013. ISBN   978-1-107-31401-6. OCLC   825767970.{{cite book}}: CS1 maint: others (link)
  12. Rolls, Edmund T. (2022). "The hippocampus, ventromedial prefrontal cortex, and episodic and semantic memory". Progress in Neurobiology. 217: 102334. doi:10.1016/j.pneurobio.2022.102334. PMID   35870682. S2CID   250722212.
  13. Fischer, Håkan; Wright, Christopher I; Whalen, Paul J; McInerney, Sean C; Shin, Lisa M; Rauch, Scott L (2003). "Brain habituation during repeated exposure to fearful and neutral faces: A functional MRI study". Brain Research Bulletin. 59 (5): 387–392. doi:10.1016/S0361-9230(02)00940-1. PMID   12507690. S2CID   43705711.
  14. Modi, Shilpi; Trivedi, Richa; Singh, Kavita; Kumar, Pawan; Rathore, Ram K.S.; Tripathi, Rajendra P.; Khushu, Subash (2013). "Individual differences in trait anxiety are associated with white matter tract integrity in fornix and uncinate fasciculus: Preliminary evidence from a DTI based tractography study". Behavioural Brain Research. 238: 188–192. doi:10.1016/j.bbr.2012.10.007. PMID   23085341. S2CID   1741104.
  15. Vann, Seralynne D. (2010). "Re-evaluating the role of the mammillary bodies in memory". Neuropsychologia. 48 (8): 2316–2327. doi:10.1016/j.neuropsychologia.2009.10.019. PMID   19879886. S2CID   2424758.
  16. Imai, Takeshi (2014). "Construction of functional neuronal circuitry in the olfactory bulb". Seminars in Cell & Developmental Biology. 35: 180–188. doi: 10.1016/j.semcdb.2014.07.012 . ISSN   1084-9521. PMID   25084319.
  17. Tong, Michelle T.; Peace, Shane T.; Cleland, Thomas A. (2014-07-07). "Properties and mechanisms of olfactory learning and memory". Frontiers in Behavioral Neuroscience. 8: 238. doi: 10.3389/fnbeh.2014.00238 . ISSN   1662-5153. PMC   4083347 . PMID   25071492.
  18. Rolls, Edmund T. (2019), "The cingulate cortex and limbic systems for action, emotion, and memory", Cingulate Cortex, Handbook of Clinical Neurology, vol. 166, Elsevier, pp. 23–37, doi:10.1016/b978-0-444-64196-0.00002-9, ISBN   978-0-444-64196-0, PMID   31731913, S2CID   208060755 , retrieved 2022-11-27
  19. Weissman, D.H. (2004-07-06). "Dorsal Anterior Cingulate Cortex Resolves Conflict from Distracting Stimuli by Boosting Attention toward Relevant Events". Cerebral Cortex. 15 (2): 229–237. doi: 10.1093/cercor/bhh125 . ISSN   1460-2199. PMID   15238434.
  20. Medford, Nick; Critchley, Hugo D. (2010). "Conjoint activity of anterior insular and anterior cingulate cortex: awareness and response". Brain Structure and Function. 214 (5–6): 535–549. doi:10.1007/s00429-010-0265-x. ISSN   1863-2653. PMC   2886906 . PMID   20512367.
  21. Drevets, Wayne C.; Savitz, Jonathan; Trimble, Michael (2008). "The Subgenual Anterior Cingulate Cortex in Mood Disorders". CNS Spectrums. 13 (8): 663–681. doi:10.1017/S1092852900013754. ISSN   1092-8529. PMC   2729429 . PMID   18704022.
  22. Dalgleish, Tim (2004). "The emotional brain". Nature Reviews Neuroscience. 5 (7): 583–589. doi:10.1038/nrn1432. ISSN   1471-003X. PMID   15208700. S2CID   148864726.
  23. Da Cunha, Claudio; Gomez-A, Alexander; Blaha, Charles D. (2012). "The role of the basal ganglia in motivated behavior". Reviews in the Neurosciences. 23 (5–6): 747–767. doi:10.1515/revneuro-2012-0063. ISSN   0334-1763. PMID   23079510. S2CID   30904211.
  24. Mikhael, John G.; Bogacz, Rafal (2016-09-02). Blackwell, Kim T. (ed.). "Learning Reward Uncertainty in the Basal Ganglia". PLOS Computational Biology. 12 (9): e1005062. Bibcode:2016PLSCB..12E5062M. doi: 10.1371/journal.pcbi.1005062 . ISSN   1553-7358. PMC   5010205 . PMID   27589489.
  25. Bechara, A. (2000-03-01). "Emotion, Decision Making and the Orbitofrontal Cortex". Cerebral Cortex. 10 (3): 295–307. doi: 10.1093/cercor/10.3.295 . PMID   10731224.
  26. Davidson, Richard J; Sutton, Steven K (1995). "Affective neuroscience: the emergence of a discipline". Current Opinion in Neurobiology. 5 (2): 217–224. doi:10.1016/0959-4388(95)80029-8. PMID   7620310. S2CID   14639993.
  27. Suzuki, Yukihiro; Tanaka, Saori C. (2021-09-14). "Functions of the ventromedial prefrontal cortex in emotion regulation under stress". Scientific Reports. 11 (1): 18225. Bibcode:2021NatSR..1118225S. doi:10.1038/s41598-021-97751-0. ISSN   2045-2322. PMC   8440524 . PMID   34521947.
  28. Kringelbach, Morten L.; Berridge, Kent C. (2010). "The functional neuroanatomy of pleasure and happiness". Discovery Medicine. 9 (49): 579–587. ISSN   1944-7930. PMC   3008353 . PMID   20587348.
  29. Rolls, Edmund T. (2016-12-01). "Functions of the anterior insula in taste, autonomic, and related functions". Brain and Cognition. Food for thought: The functional and neural mechanisms of food perception and choice. 110: 4–19. doi:10.1016/j.bandc.2015.07.002. ISSN   0278-2626. PMID   26277487. S2CID   5961714.
  30. Gu, Xiaosi; Hof, Patrick R.; Friston, Karl J.; Fan, Jin (2013). "Anterior insular cortex and emotional awareness: Anterior Insular Cortex and Emotional Awareness". Journal of Comparative Neurology. 521 (15): 3371–3388. doi:10.1002/cne.23368. PMC   3999437 . PMID   23749500.
  31. Adamaszek, Michael; Manto, Mario; Schutter, Dennis J. L. G. (2022), Adamaszek, Michael; Manto, Mario; Schutter, Dennis J. L. G. (eds.), "Current and Future Perspectives of the Cerebellum in Affective Neuroscience", The Emotional Cerebellum, Cham: Springer International Publishing, vol. 1378, pp. 303–313, doi:10.1007/978-3-030-99550-8_19, ISBN   978-3-030-99549-2, PMID   35902479 , retrieved 2023-11-16
  32. Turner, Beth M.; Paradiso, Sergio; Marvel, Cherie L.; Pierson, Ronald; Boles Ponto, Laura L.; Hichwa, Richard D.; Robinson, Robert G. (2007-03-25). "The cerebellum and emotional experience". Neuropsychologia. 45 (6): 1331–1341. doi:10.1016/j.neuropsychologia.2006.09.023. ISSN   0028-3932. PMC   1868674 . PMID   17123557.
  33. Martin-Sölch, C.; Magyar, S.; Künig, G.; Missimer, J.; Schultz, W.; Leenders, K. L. (2001). "Changes in brain activation associated with reward processing in smokers and nonsmokers. A positron emission tomography study". Experimental Brain Research. 139 (3): 278–286. doi:10.1007/s002210100751. ISSN   0014-4819. PMID   11545466. S2CID   29989526.
  34. Sell, L. A.; Morris, J.; Bearn, J.; Frackowiak, R. S.; Friston, K. J.; Dolan, R. J. (1999). "Activation of reward circuitry in human opiate addicts". The European Journal of Neuroscience. 11 (3): 1042–1048. doi:10.1046/j.1460-9568.1999.00522.x. hdl: 21.11116/0000-0001-A04C-5 . ISSN   0953-816X. PMID   10103096. S2CID   12004748.
  35. Zhou, Huan-Xiang (2008-11-05). "The debut of PMC Biophysics". PMC Biophysics. 1 (1): 1. doi: 10.1186/1757-5036-1-1 . ISSN   1757-5036. PMC   2605105 . PMID   19351423.
  36. Parvizi, J. (2001-09-01). "Pathological laughter and crying: A link to the cerebellum". Brain. 124 (9): 1708–1719. doi:10.1093/brain/124.9.1708. PMID   11522574.
  37. Barbas, Helen; García-Cabezas, Miguel Ángel (2017), "Prefrontal Cortex Integration of Emotion and Cognition", The Prefrontal Cortex as an Executive, Emotional, and Social Brain, Tokyo: Springer Japan, pp. 51–76, doi:10.1007/978-4-431-56508-6_4, ISBN   978-4-431-56506-2 , retrieved 2022-11-27
  38. Kropf, Erika; Syan, Sabrina K.; Minuzzi, Luciano; Frey, Benicio N. (2019). "From anatomy to function: the role of the somatosensory cortex in emotional regulation". Brazilian Journal of Psychiatry. 41 (3): 261–269. doi:10.1590/1516-4446-2018-0183. ISSN   1809-452X. PMC   6794131 . PMID   30540029.
  39. Goghari, V. M.; MacDonald, A. W.; Sponheim, S. R. (2010-05-19). "Temporal Lobe Structures and Facial Emotion Recognition in Schizophrenia Patients and Nonpsychotic Relatives". Schizophrenia Bulletin. 37 (6): 1281–1294. doi:10.1093/schbul/sbq046. ISSN   0586-7614. PMC   3196942 . PMID   20484523.
  40. Kumfor, Fiona; Irish, Muireann; Hodges, John R.; Piguet, Olivier (2014-06-24). "Frontal and temporal lobe contributions to emotional enhancement of memory in behavioral-variant frontotemporal dementia and Alzheimer's disease". Frontiers in Behavioral Neuroscience. 8: 225. doi: 10.3389/fnbeh.2014.00225 . ISSN   1662-5153. PMC   4067999 . PMID   25009480.
  41. Venkatraman, Anand; Edlow, Brian L.; Immordino-Yang, Mary Helen (2017-03-09). "The Brainstem in Emotion: A Review". Frontiers in Neuroanatomy. 11: 15. doi: 10.3389/fnana.2017.00015 . ISSN   1662-5129. PMC   5343067 . PMID   28337130.
  42. Mills, C. K. (1912a). "The cerebral mechanism of emotional expression". Transactions of the College of Physicians of Philadelphia. 34: 381–390.
  43. Mills, C. K. (1912b). "The cortical representation of emotion, with a discussion of some points in the general nervous mechanism of expression in its relation to organic nervous disease and insanity". Proceedings of the American Medico-Psychological Association. 19: 297–300.
  44. 1 2 Borod, J. C. (1992). "Intel-hemispheric and intrahemispheric control of emotion: A focus on unilateral brain damage". Journal of Consulting and Clinical Psychology. 60 (3): 339–348. doi:10.1037/0022-006x.60.3.339. PMID   1619088.
  45. Borod, J. C., Koff, E., & Caron, H. S. (1983). Right hemispheric specialization for the expression and appreciation of emotion: A focus on the face. In E. Perecman (Ed.), Cognitive functions in the right hemisphere (pp. 83-110). New York: Academic Press.
  46. Heilman, K. M. (February 1982). Discussant comments. In J. C. Borod & R. Buck (Chairs), Asymmetries in facial expression: Method and meaning. Symposium conducted at the International Neuropsychological Society, Pittsburgh, PA.
  47. Yokoyama, K.; Jennings, R.; Ackles, P.; Hood, B. S.; Boiler, F. (1987). "Lack of heart rate changes during an attention-demanding task after right hemisphere lesions". Neurology. 37 (4): 624–630. doi:10.1212/wnl.37.4.624. PMID   3561774. S2CID   38096461.
  48. Silberman, E. K.; Weingartner, H. (1986). "Hemispheric lateralization of functions related to emotion". Brain and Cognition. 5 (3): 322–353. doi:10.1016/0278-2626(86)90035-7. PMID   3530287. S2CID   2636348.
  49. Fox, N. A. (1991). "If it's not left, it's right". American Psychologist. 46 (8): 863–872. doi:10.1037/0003-066x.46.8.863. PMID   1928939.
  50. Davidson, R.; Ekman, P.; Saron, C. D.; Senulis, J. A.; Friesen, W V (1990). "Approach-withdrawal and cerebral asymmetry: Emotional expression and brain physiology I". Journal of Personality and Social Psychology. 58 (2): 330–341. doi:10.1037/0022-3514.58.2.330. PMID   2319445.
  51. Borod, J.C.; Cicero, B.A.; Obler, L.K.; Welkowitz, J.; Erhan, H.M.; Santschi, C.; et al. (1998). "Right hemisphere emotional perception: Evidence across multiple channels". Neuropsychology. 12 (3): 446–458. doi:10.1037/0894-4105.12.3.446. PMID   9673999.
  52. Engels, AS; Heller, W; Mohanty, A; Herrington, JD; Banich, MT; Webb, AG; Miller, GA (2007). "Specificity of regional brain activity in anxiety types during emotion processing". Psychophysiology. 44 (3): 352–363. doi:10.1111/j.1469-8986.2007.00518.x. PMID   17433094.
  53. Spielberg, J.; Stewart, J; Levin, R.; Miller, G.; Heller, W. (2008). "Prefrontal Cortex, Emotion, and Approach/Withdrawal Motivation". Soc Personal Psychol Compass. 2 (1): 135–153. doi:10.1111/j.1751-9004.2007.00064.x. PMC   2889703 . PMID   20574551.
  54. Cacioppo, J.T.; Gardner, W.L. (1999). "Emotion". Annual Review of Psychology. 50: 191–214. doi:10.1146/annurev.psych.50.1.191. PMID   10074678.
  55. Davidson, R.J. (2000). "Cognitive neuroscience needs affective neuroscience (and vice versa)". Brain and Cognition. 42 (1): 89–92. CiteSeerX   10.1.1.487.9015 . doi:10.1006/brcg.1999.1170. PMID   10739607. S2CID   22183667.
  56. 1 2 Drevets, W.C.; Raichle, M.E. (1998). "Reciprocal suppression of regional cerebral blood flow during emotional versus higher cognitive processes: Implications for interactions between emotion and cognition". Cognition and Emotion. 12 (3): 353–385. doi:10.1080/026999398379646.
  57. Schulz, K.P.; Fan, J.; Magidina, O.; Marks, D.J.; Hahn, B.; Halperin, J.M. (2007). "Does the emotional go/no-go task really measure behavioral inhibition? Convergence with measures on a non-emotional analog". Archives of Clinical Neuropsychology. 22 (2): 151–160. doi:10.1016/j.acn.2006.12.001. PMC   2562664 . PMID   17207962.
  58. Elliott, R.; Ogilvie, A.; Rubinsztein, J.S.; Calderon, G.; Dolan, R.J.; Sahakian, B.J. (2004). "Abnormal ventral frontal response during performance of an affective go/no go task in patients with mania". Biological Psychiatry. 55 (12): 1163–1170. doi:10.1016/j.biopsych.2004.03.007. hdl: 21.11116/0000-0001-A247-8 . PMID   15184035. S2CID   6511699.
  59. Elliott, R.; Rubinsztein, J.S.; Sahakian, B.J.; Dolan, R.J. (2000). "Selective attention to emotional stimuli in a verbal go/no-go task: An fMRI study". Brain Imaging. 11 (8): 1739–1744. doi:10.1097/00001756-200006050-00028. PMID   10852235. S2CID   16872952.
  60. Hare, T.A.; Tottenham, N.; Davidson, M.C.; Glover, G.H.; Casey, B.J. (2005). "Contributions of amygdala and striatal activity in emotion regulation". Biological Psychiatry. 57 (6): 624–632. doi:10.1016/j.biopsych.2004.12.038. PMID   15780849. S2CID   17690328.
  61. Stroop, J.R. (1935). "Studies of interference in serial verbal reactions". Journal of Experimental Psychology. 18 (6): 643–662. doi:10.1037/h0054651. hdl: 11858/00-001M-0000-002C-5ADB-7 .
  62. 1 2 Epp, A.M.; Dobson, K.S.; Dozois, D.J.A.; Frewen, P.A. (2012). "A systematic meta-analysis of the stroop task in depression". Clinical Psychology Review. 32 (4): 316–328. doi:10.1016/j.cpr.2012.02.005. PMID   22459792.
  63. Pratto, F.; John, O. P. (1991). "Automatic vigilance: The attention grabbing power of negative social information". Journal of Personality & Social Psychology. 61 (3): 380–391. doi:10.1037/0022-3514.61.3.380. PMID   1941510.
  64. Wentura, D.; Rothermund, K.; Bak, P. (2000). "Automatic vigilance: The attention-grabbing power of approach and avoidance-related social information". Journal of Personality and Social Psychology. 78 (6): 1024–1037. doi:10.1037/0022-3514.78.6.1024. PMID   10870906.
  65. Williams, J.M.; Broadbent, K. (1986). "Distraction by emotional stimuli: use of a Stroop task with suicide attempters". British Journal of Clinical Psychology. 25 (2): 101–110. doi:10.1111/j.2044-8260.1986.tb00678.x. PMID   3730646.
  66. 1 2 3 4 Williams, M.G.; Matthews, A.; MacLeod, C. (1996). "The emotional stroop task and psychopathology". Psychological Bulletin. 120 (1): 3–24. doi:10.1037/0033-2909.120.1.3. PMID   8711015.
  67. Gotlib, I.H., Roberts, J.E., & Gilboa, E. (1996). Cognitive interference in depression. in I.G. Sarason, G.R. Pierce, & B.R. Sarason (Eds.), Cognitive interference: Theories, methods, and findings, Lawrence Erlbaum Associates Inc., Hillsdale, NJ, pp. 347–377.
  68. Watts, E N.; McKenna, E P.; Sharrock, R.; Trezise, L. (1986). "Colour naming of phobia-related words". British Journal of Psychology. 77: 97–108. doi:10.1111/j.2044-8295.1986.tb01985.x. PMID   2869817.
  69. Martin, M.; Williams, R. M.; Clark, D. M. (1991). "Does anxiety lead to selective processing of threat-related information?". Behaviour Research and Therapy. 29 (2): 147–160. doi:10.1016/0005-7967(91)90043-3. PMID   2021377.
  70. Mogg, K.; Mathews, A. M.; Weinman, J. (1989). "Selective processing of threat cues in anxiety states: A replication". Behaviour Research and Therapy. 27 (4): 317–323. doi:10.1016/0005-7967(89)90001-6. PMID   2775141.
  71. Diehl-Schmid, J.; Pohl, C.; Ruprecht, C.; Wagenpfeil, S.; Foerstl, H.; Kurz, A. (2007). "The Ekman 60 faces test as a diagnostic instrument in frontotemporal dementia". Archives of Clinical Neuropsychology. 22 (4): 459–464. doi: 10.1016/j.acn.2007.01.024 . PMID   17360152.
  72. Ekman, P., & Friesen, W. (1976). Pictures of facial affect. Consulting Psychologists Press: Palo Alto, CA.
  73. Diehl-Schmida, J.; Pohla, C.; Ruprechta, C.; Wagenpfeilb, S.; Foerstla, H.; Kurza, A. (2007). "The Ekman 60 faces test as a diagnostic instrument in frontotemporal dementia". Archives of Clinical Neuropsychology. 22 (4): 459–464. doi: 10.1016/j.acn.2007.01.024 . PMID   17360152.
  74. Ibarretxe-Bilbao, N.; Junque, C.; Tolosa, E.; Marti, M.; Valldeoriola, F.; Bargalloand, N.; Zareio, M. (2009). "Neuroanatomical correlates of impaired decision-making and facial emotion recognition in early Parkinson's disease". European Journal of Neuroscience. 30 (6): 1162–1171. doi:10.1111/j.1460-9568.2009.06892.x. PMID   19735293. S2CID   413809.
  75. Soeiro-de-Souza, M.G.; Bio, D.S.; David, D.P.; Santos, D.R.; Kerr, D.S.; Moreno, R.A.; Machado-Vieira, Rodrigo; Moreno, Ricardo Albeto (2012). "COMT Met (158) modulates facial emotion recognition in bipolar I disorder mood episodes". Journal of Affective Disorders. 136 (3): 370–376. doi:10.1016/j.jad.2011.11.021. PMID   22222175.
  76. Ebert, A.; Haussleiter, I.S.; Juckel, G.; Brune, M.; Roser, P. (2012). "Impaired facial emotion recognition in a ketamine model of psychosis". Psychiatry Research. 200 (2–3): 724–727. doi:10.1016/j.psychres.2012.06.034. PMID   22776754. S2CID   43744551.
  77. 1 2 3 Koster, E.H.W.; Crombez, G.; Verschuere, B.; De Houwer, J. (2004). "Selective attention to threat in the dot probe paradigm: Differentiating vigilance and difficulty to disengage". Behaviour Research and Therapy. 42 (10): 1183–1192. doi:10.1016/j.brat.2003.08.001. PMID   15350857.
  78. MacLeod, C.; Mathews, A.; Tata, P. (1986). "Attentional bias in emotional disorders". Journal of Abnormal Psychology. 95 (1): 15–20. doi:10.1037/0021-843x.95.1.15. PMID   3700842.
  79. Mogg, K.; Bradley, B. P. (1998). "A cognitive–motivational analysis of anxiety". Behaviour Research and Therapy. 36 (9): 809–848. doi:10.1016/s0005-7967(98)00063-1. PMID   9701859.
  80. Asmundson, G.J.G.; Stein, M.B. (1994). "Selective processing of social threat in patients with generalized social phobia: Evaluation using a dot-probe paradigm". Journal of Anxiety Disorders. 8 (2): 107–117. doi:10.1016/0887-6185(94)90009-4.
  81. Amir, N.; Naimi, S.; Morrison, A.S. (2009). "Attenuation of attention bias in obsessive-compulsive disorder". Behaviour Research and Therapy. 47 (2): 153–157. doi:10.1016/j.brat.2008.10.020. PMC   2662360 . PMID   19046576.
  82. Davis, M. (1986). "Pharmacological and anatomical analysis of fear conditioning using the fear-potentiated startle paradigm". Behavioral Neuroscience. 100 (6): 814–824. doi:10.1037/0735-7044.100.6.814. PMID   3545257.
  83. Baskin-Sommers, A.R.; Vitale, J.E.; MacCoon, D.; Newman, J.P. (2012). "Assessing emotion sensitivity in female offenders with borderline personality symptoms: Results from a fear-potentiated startle paradigm". Journal of Abnormal Psychology. 121 (2): 477–483. doi:10.1037/a0026753. PMC   3358451 . PMID   22250659.
  84. 1 2 Vaidyanathan, U.; Patrick, C.J.; Cuthbert, B.N. (2009). "Linking dimensional models of internalizing psychopathology to neurobiological systems: Affect-modulated startle as an indicator of fear and distress disorders and affiliated traits". Psychological Bulletin. 135 (6): 909–942. doi:10.1037/a0017222. PMC   2776729 . PMID   19883142.
  85. Lang, P. J.; Bradley, M. M.; Cuthbert, B. N. (1990). "Emotion, attention, and the startle reflex". Psychological Review. 97 (3): 377–395. doi:10.1037/0033-295x.97.3.377. PMID   2200076.
  86. Vrana, S. R.; Spence, E. L.; Lang, P. J. (1988). "The startle probe response: A new measure of emotion?". Journal of Abnormal Psychology. 97 (4): 487–491. doi:10.1037/0021-843x.97.4.487. PMID   3204235.
  87. Norrholm, S.D.; Anderson, K.M.; Olin, I.W.; Jovanovic, T.; Kwon, C.; Warren, Victor T.; Bradley, B.; Bosshardt, Lauren; Sabree, Justin; Duncan, Erica J.; Rothbaum, Barbara O.; Bradley, Bekh (2011). "Versatility of fear-potentiated startle paradigms for assessing human conditioned fear extinction and return of fear". Frontiers in Behavioral Neuroscience. 5: 77. doi: 10.3389/fnbeh.2011.00077 . PMC   3221285 . PMID   22125516.
  88. Grillon, C.; Ameli, R.; Woods, S. W.; Merikangas, K.; Davis, M. (1991). "Fear-potentiated startle in humans: Effects of anticipatory anxiety on the acoustic blink reflex". Psychophysiology. 28 (5): 588–595. doi:10.1111/j.1469-8986.1991.tb01999.x. PMID   1758934.
  89. Norrholm, S.; Jovanovic, Tanja (2011). "Translational fear inhibition models as indices of trauma-related psychopathology". Current Psychiatry Reviews. 7 (3): 194–204. doi:10.2174/157340011797183193.
  90. Norrholm, S. D.; Jovanovic, T.; Olin, I. W.; Sands, L.; Karapanou, I.; Bradley, B.; Ressler, K. J. (2010). "Fear extinction in traumatized civilians with posttraumatic stress disorder: relation to symptom severity". Biological Psychiatry. 69 (6): 556–563. doi:10.1016/j.biopsych.2010.09.013. PMC   3052965 . PMID   21035787.
  91. Lang, P. J.; McTeague, L. M. (2009). "The anxiety disorder spectrum: Fear imagery, physiological reactivity, and differential diagnosis". Anxiety, Stress & Coping. 22 (1): 5–25. doi:10.1080/10615800802478247. PMC   2766521 . PMID   19096959.
  92. 1 2 Cousens, G.A.; Kearns, A.; Laterza, F.; Tundidor, J. (2012). "Excitotoxic lesions of the medial amygdala attenuate olfactory fear-potentiated startle and conditioned freezing behavior". Behavioural Brain Research. 229 (2): 427–432. doi:10.1016/j.bbr.2012.01.011. PMID   22249137. S2CID   25116098.
  93. Jovanovic, T; Norrholm, Seth D.; Fennell, Jennifer E.; Keyes, Megan; Fiallos, Ana M.; Myers, Karyn M.; Davis, Michael; Duncan, Erica J. (15 May 2009). "Posttraumatic stress disorder may be associated with impaired fear inhibition: relation to symptom severity". Psychiatry Research. 167 (1–2): 151–160. doi:10.1016/j.psychres.2007.12.014. PMC   2713500 . PMID   19345420.
  94. Picard, R. W.; Papert, S; Bender, W; Blumberg, B; Breazeal, C; Cavallo, D; Machover, T; Resnick, M; Roy, D; Strohecker, C (2004). "Affective Learning – a manifesto". BT Technology Journal. 22 (4): 253–269. CiteSeerX   10.1.1.110.3308 . doi:10.1023/b:bttj.0000047603.37042.33. S2CID   518899. Closed Access logo transparent.svg
  95. Havas, D.A.; Glenberg, A.M.; Rinck, M. (2007). "Emotion simulation during language comprehension". Psychonomic Bulletin & Review. 14 (3): 436–441. doi: 10.3758/bf03194085 . PMID   17874584.
  96. Wicker, B.; Keysers, Christian; Plailly, Jane; Royet, Jean-Pierre; Gallese, Vittorio; Rizzolatti, Giacomo (2003). "Both of Us Disgusted in My Insula: The Common Neural Basis of Seeing and Feeling Disgust". Neuron. 40 (3): 655–664. doi: 10.1016/s0896-6273(03)00679-2 . PMID   14642287. S2CID   766157.
  97. Niedenthal, P. M. (2007). "Embodying emotion". Science. 316 (5827): 1002–1005. Bibcode:2007Sci...316.1002N. doi:10.1126/science.1136930. PMID   17510358. S2CID   14537829.
  98. 1 2 3 Celeghin, Alessia; Diano, Matteo; Bagnis, Arianna; Viola, Marco; Tamietto, Marco (2017). "Basic Emotions in Human Neuroscience: Neuroimaging and Beyond". Frontiers in Psychology. 8: 1432. doi: 10.3389/fpsyg.2017.01432 . ISSN   1664-1078. PMC   5573709 . PMID   28883803. Creative Commons by small.svg  This article incorporates text by Alessia Celeghin, Matteo Diano, Arianna Bagnis, Marco Viola, and Marco Tamietto available under the CC BY 4.0 license.
  99. 1 2 3 Ekman, P.; Cordaro, D. (2011). "What is meant by calling emotions basic". Emotion Review. 3 (4): 364–370. doi:10.1177/1754073911410740. S2CID   52833124.
  100. 1 2 3 Panksepp, J.; Watt, D. (2011). "What is basic about basic emotions? Lasting lessons from affective neuroscience". Emotion Review. 3 (4): 387–396. doi:10.1177/1754073911410741. S2CID   144198152.
  101. Izard, C.E. (2011). "Forms and functions in emotions: Matters of emotion-cognition interactions". Emotion Review. 3 (4): 371–378. doi:10.1177/1754073911410737. S2CID   144037625.
  102. Panksepp, J. (1998). Affective Neuroscience: The foundations of human and animal emotions. New York: Oxford University Press.
  103. 1 2 3 Barrett, L.F.; Wager, T. (2006). "The structure of emotion: Evidence from the neuroimaging of emotion". Current Directions in Psychological Science. 15 (2): 79–85. CiteSeerX   10.1.1.470.7762 . doi:10.1111/j.0963-7214.2006.00411.x. S2CID   14489624.
  104. 1 2 Lindquist, K.; Wager, T.; Kober, Hedy; Bliss-Moreau, Eliza; Barrett, Lisa Feldman (2012). "The brain basis of emotion: A meta-analytic review". Behavioral and Brain Sciences. 35 (3): 121–143. doi:10.1017/s0140525x11000446. PMC   4329228 . PMID   22617651.
  105. Fox, Andrew S.; Shackman, Alexander J. (February 2019). "The central extended amygdala in fear and anxiety: Closing the gap between mechanistic and neuroimaging research". Neuroscience Letters. 693: 58–67. doi:10.1016/j.neulet.2017.11.056. ISSN   0304-3940. PMC   5976525 . PMID   29195911.
  106. 1 2 Allard, Eric; Kensinger, Elizabeth (November 10, 2014). "Age-Related Differences in Functional Connectivity During Cognitive Emotion Regulation". The Journals of Gerontology. Retrieved 2023-10-25.
  107. Suzuki, Yukihiro; Tanaka, Saori C. (2021-09-14). "Functions of the ventromedial prefrontal cortex in emotion regulation under stress". Scientific Reports. 11 (1): 18225. Bibcode:2021NatSR..1118225S. doi: 10.1038/s41598-021-97751-0 . ISSN   2045-2322. PMC   8440524 . PMID   34521947.
  108. 1 2 Carstensen, Laura; Pasupathi, Monisha; Ulrich, Mayr; Nesselroade, John (2000). "Emotional experience in everyday life across the adult life span". Journal of Personality and Social Psychology. 79 (4): 644–655. doi:10.1037/0022-3514.79.4.644. PMID   11045744 . Retrieved 2023-10-26.
  109. Urry, Heather L.; Gross, James J. (2010). "Emotion Regulation in Older Age". Current Directions in Psychological Science. 19 (6): 352–357. doi:10.1177/0963721410388395. ISSN   0963-7214. S2CID   1400335.
  110. Blanchard-Fields, Fredda; Stein, Renee; Watson, Tonya (2004). "Age Differences in Emotion-Regulation Strategies in Handling Everyday Problems". The Journals of Gerontology. Series B, Psychological Sciences and Social Sciences. The Journals of Gerontology. 59 (6): P261-9. doi:10.1093/geronb/59.6.p261. PMID   15576853 . Retrieved 2023-10-26.
  111. Carstensen, Laura L.; Fung, Helene H.; Charles, Susan T. (2003-06-01). "Socioemotional Selectivity Theory and the Regulation of Emotion in the Second Half of Life". Motivation and Emotion. 27 (2): 103–123. doi:10.1023/A:1024569803230. ISSN   1573-6644. S2CID   143149171.
  112. Phan, K.L.; Wager, T.D.; Taylor, S.F.; Liberzon, I. (2002). "Functional neuroanatomy of emotion: A meta-analysis of emotion activation studies in PET and fMRI". NeuroImage. 16 (2): 331–348. doi:10.1006/nimg.2002.1087. PMID   12030820. S2CID   7150871.
  113. Murphy, F.C.; Nimmo-Smith, I.; Lawrence, A.D. (2003). "Functional Neuroanatomy: A meta-analysis". Cognitive, Affective, & Behavioral Neuroscience. 3 (3): 207–233. doi: 10.3758/cabn.3.3.207 . PMID   14672157.
  114. 1 2 Wager, T.D.; Lindquist, M.; Kaplan, L. (2007). "Meta-analysis of functional neuroimaging data: Current and future directions". Social Cognitive and Affective Neuroscience. 2 (2): 150–158. doi:10.1093/scan/nsm015. PMC   2555451 . PMID   18985131.
  115. Kober, H.; Barrett, L.F.; Joseph, J.; Bliss-Moreau, E.; Lindquist, K.; Wager, T.D. (2008). "Functional grouping and cortical-subcortical interactions in emotion: A meta-analysis of neuroimaging studies". NeuroImage. 42 (2): 998–1031. doi:10.1016/j.neuroimage.2008.03.059. PMC   2752702 . PMID   18579414.
  116. Vytal, K.; Hamann, S. (2010). "Neuroimaging support for discrete neural correlates of basic emotions: A voxel-based meta-analysis". Journal of Cognitive Neuroscience. 22 (12): 2864–2885. doi:10.1162/jocn.2009.21366. PMID   19929758. S2CID   18151469.

Further reading