Journal articles on the topic 'Brain's default mode'

To see the other types of publications on this topic, follow the link: Brain's default mode.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Brain's default mode.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Raichle, Marcus E. "The Brain's Default Mode Network." Annual Review of Neuroscience 38, no. 1 (July 8, 2015): 433–47. http://dx.doi.org/10.1146/annurev-neuro-071013-014030.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Davey, Christopher G., Jesus Pujol, and Ben J. Harrison. "Mapping the self in the brain's default mode network." NeuroImage 132 (May 2016): 390–97. http://dx.doi.org/10.1016/j.neuroimage.2016.02.022.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Horovitz, S. G., A. R. Braun, W. S. Carr, D. Picchioni, T. J. Balkin, M. Fukunaga, and J. H. Duyn. "Decoupling of the brain's default mode network during deep sleep." Proceedings of the National Academy of Sciences 106, no. 27 (June 19, 2009): 11376–81. http://dx.doi.org/10.1073/pnas.0901435106.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Shin, Jonghan, Vladimir Kepe, Gary W. Small, Michael E. Phelps, and Jorge R. Barrio. "Multimodal Imaging of Alzheimer Pathophysiology in the Brain's Default Mode Network." International Journal of Alzheimer's Disease 2011 (2011): 1–8. http://dx.doi.org/10.4061/2011/687945.

Full text
Abstract:
The spatial correlations between the brain's default mode network (DMN) and the brain regions known to develop pathophysiology in Alzheimer's disease (AD) have recently attracted much attention. In this paper, we compare results of different functional and structural imaging modalities, including MRI and PET, and highlight different patterns of anomalies observed within the DMN. Multitracer PET imaging in subjects with and without dementia has demonstrated that [C-11]PIB- and [F-18]FDDNP-binding patterns in patients with AD overlap within nodes of the brain's default network including the prefrontal, lateral parietal, lateral temporal, and posterior cingulate cortices, with the exception of the medial temporal cortex (especially, the hippocampus) where significant discrepancy between increased [F-18]FDDNP binding and negligible [C-11]PIB-binding was observed. [F-18]FDDNP binding in the medial temporal cortex—a key constituent of the DMN—coincides with both the presence of amyloid and tau pathology, and also with cortical areas with maximal atrophy as demonstrated by T1-weighted MR imaging of AD patients.
APA, Harvard, Vancouver, ISO, and other styles
5

Sämann, Philipp G., Renate Wehrle, David Hoehn, Victor I. Spoormaker, Henning Peters, Carolin Tully, Florian Holsboer, and Michael Czisch. "Development of the Brain's Default Mode Network from Wakefulness to Slow Wave Sleep." Cerebral Cortex 21, no. 9 (February 17, 2011): 2082–93. http://dx.doi.org/10.1093/cercor/bhq295.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Spunt, Robert P., Meghan L. Meyer, and Matthew D. Lieberman. "The Default Mode of Human Brain Function Primes the Intentional Stance." Journal of Cognitive Neuroscience 27, no. 6 (June 2015): 1116–24. http://dx.doi.org/10.1162/jocn_a_00785.

Full text
Abstract:
Humans readily adopt an intentional stance to other people, comprehending their behavior as guided by unobservable mental states such as belief, desire, and intention. We used fMRI in healthy adults to test the hypothesis that this stance is primed by the default mode of human brain function present when the mind is at rest. We report three findings that support this hypothesis. First, brain regions activated by actively adopting an intentional rather than nonintentional stance to a social stimulus were anatomically similar to those demonstrating default responses to fixation baseline in the same task. Second, moment-to-moment variation in default activity during fixation in the dorsomedial PFC was related to the ease with which participants applied an intentional—but not nonintentional—stance to a social stimulus presented moments later. Finally, individuals who showed stronger dorsomedial PFC activity at baseline in a separate task were generally more efficient when adopting the intentional stance and reported having greater social skills. These results identify a biological basis for the human tendency to adopt the intentional stance. More broadly, they suggest that the brain's default mode of function may have evolved, in part, as a response to life in a social world.
APA, Harvard, Vancouver, ISO, and other styles
7

Hui, Kathleen K. S., Ovidiu Marina, Joshua D. Claunch, Erika E. Nixon, Jiliang Fang, Jing Liu, Ming Li, et al. "Acupuncture mobilizes the brain's default mode and its anti-correlated network in healthy subjects." Brain Research 1287 (September 2009): 84–103. http://dx.doi.org/10.1016/j.brainres.2009.06.061.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Briggs, Robert G., Onur Tanglay, Nicholas B. Dadario, Isabella M. Young, R. Dineth Fonseka, Jorge Hormovas, Vukshitha Dhanaraj, et al. "The Unique Fiber Anatomy of Middle Temporal Gyrus Default Mode Connectivity." Operative Neurosurgery 21, no. 1 (April 30, 2021): E8—E14. http://dx.doi.org/10.1093/ons/opab109.

Full text
Abstract:
Abstract BACKGROUND The middle temporal gyrus (MTG) is understood to play a role in language-related tasks such as lexical comprehension and semantic cognition. However, a more specific understanding of its key white matter connections could promote the preservation of these functions during neurosurgery. OBJECTIVE To provide a detailed description of the underlying white matter tracts associated with the MTG to improve semantic preservation during neurosurgery. METHODS Tractography was performed using diffusion imaging obtained from 10 healthy adults from the Human Connectome Project. All tracts were mapped between cerebral hemispheres with a subsequent laterality index calculated based on resultant tract volumes. Ten postmortem dissections were performed for ex vivo validation of the tractography based on qualitative visual agreement. RESULTS We identified 2 major white matter bundles leaving the MTG: the inferior longitudinal fasciculus and superior longitudinal fasciculus. In addition to long association fibers, a unique linear sequence of U-shaped fibers was identified, possibly representing a form of visual semantic transfer down the temporal lobe. CONCLUSION We elucidate the underlying fiber-bundle anatomy of the MTG, an area highly involved in the brain's language network. Improved understanding of the unique, underlying white matter connections in and around this area may augment our overall understanding of language processing as well as the involvement of higher order cerebral networks like the default mode network in these functions.
APA, Harvard, Vancouver, ISO, and other styles
9

Davey, Christopher G., and Ben J. Harrison. "The brain's center of gravity: how the default mode network helps us to understand the self." World Psychiatry 17, no. 3 (September 7, 2018): 278–79. http://dx.doi.org/10.1002/wps.20553.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Sambataro, F., N. D. Wolf, M. Pennuto, N. Vasic, and R. C. Wolf. "Revisiting default mode network function in major depression: evidence for disrupted subsystem connectivity." Psychological Medicine 44, no. 10 (October 31, 2013): 2041–51. http://dx.doi.org/10.1017/s0033291713002596.

Full text
Abstract:
BackgroundMajor depressive disorder (MDD) is characterized by alterations in brain function that are identifiable also during the brain's ‘resting state’. One functional network that is disrupted in this disorder is the default mode network (DMN), a set of large-scale connected brain regions that oscillate with low-frequency fluctuations and are more active during rest relative to a goal-directed task. Recent studies support the idea that the DMN is not a unitary system, but rather is composed of smaller and distinct functional subsystems that interact with each other. The functional relevance of these subsystems in depression, however, is unclear.MethodHere, we investigated the functional connectivity of distinct DMN subsystems and their interplay in depression using resting-state functional magnetic resonance imaging.ResultsWe show that patients with MDD exhibit increased within-network connectivity in posterior, ventral and core DMN subsystems along with reduced interplay from the anterior to the ventral DMN subsystems.ConclusionsThese data suggest that MDD is characterized by alterations of subsystems within the DMN as well as of their interactions. Our findings highlight a critical role of DMN circuitry in the pathophysiology of MDD, thus suggesting these subsystems as potential therapeutic targets.
APA, Harvard, Vancouver, ISO, and other styles
11

Anderson, Daniel R., and Matthew C. Davidson. "Receptive versus interactive video screens: A role for the brain's default mode network in learning from media." Computers in Human Behavior 99 (October 2019): 168–80. http://dx.doi.org/10.1016/j.chb.2019.05.008.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Frankland, Steven M., and Joshua D. Greene. "Concepts and Compositionality: In Search of the Brain's Language of Thought." Annual Review of Psychology 71, no. 1 (January 4, 2020): 273–303. http://dx.doi.org/10.1146/annurev-psych-122216-011829.

Full text
Abstract:
Imagine Genghis Khan, Aretha Franklin, and the Cleveland Cavaliers performing an opera on Maui. This silly sentence makes a serious point: As humans, we can flexibly generate and comprehend an unbounded number of complex ideas. Little is known, however, about how our brains accomplish this. Here we assemble clues from disparate areas of cognitive neuroscience, integrating recent research on language, memory, episodic simulation, and computational models of high-level cognition. Our review is framed by Fodor's classic language of thought hypothesis, according to which our minds employ an amodal, language-like system for combining and recombining simple concepts to form more complex thoughts. Here, we highlight emerging work on combinatorial processes in the brain and consider this work's relation to the language of thought. We review evidence for distinct, but complementary, contributions of map-like representations in subregions of the default mode network and sentence-like representations of conceptual relations in regions of the temporal and prefrontal cortex.
APA, Harvard, Vancouver, ISO, and other styles
13

Otti, A., H. Guendel, L. Läer, A. M. Wohlschlaeger, R. D. Lane, J. Decety, C. Zimmer, P. Henningsen, and M. Noll-Hussong. "I know the pain you feel—how the human brain's default mode predicts our resonance to another's suffering." Neuroscience 169, no. 1 (August 2010): 143–48. http://dx.doi.org/10.1016/j.neuroscience.2010.04.072.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

McKinnon, Andrew C., Jim Lagopoulos, Zoe Terpening, Ron Grunstein, Ian B. Hickie, Jennifer Batchelor, Simon JG Lewis, Shantel L. Duffy, James M. Shine, and Sharon L. Naismith. "P1-281: Sleep Disturbance in Mild Cognitive Impairment is Associated With Alterations in The Brain's Default Mode Network." Alzheimer's & Dementia 12 (July 2016): P526—P527. http://dx.doi.org/10.1016/j.jalz.2016.06.1031.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Hansen, Naja Liv, Martin Lauritzen, Erik Lykke Mortensen, Merete Osler, Kirsten Avlund, Birgitte Fagerlund, and Egill Rostrup. "Subclinical cognitive decline in middle-age is associated with reduced task-induced deactivation of the brain's default mode network." Human Brain Mapping 35, no. 9 (February 27, 2014): 4488–98. http://dx.doi.org/10.1002/hbm.22489.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Bar, Moshe. "The proactive brain: memory for predictions." Philosophical Transactions of the Royal Society B: Biological Sciences 364, no. 1521 (May 12, 2009): 1235–43. http://dx.doi.org/10.1098/rstb.2008.0310.

Full text
Abstract:
It is proposed that the human brain is proactive in that it continuously generates predictions that anticipate the relevant future. In this proposal, analogies are derived from elementary information that is extracted rapidly from the input, to link that input with the representations that exist in memory. Finding an analogical link results in the generation of focused predictions via associative activation of representations that are relevant to this analogy, in the given context. Predictions in complex circumstances, such as social interactions, combine multiple analogies. Such predictions need not be created afresh in new situations, but rather rely on existing scripts in memory, which are the result of real as well as of previously imagined experiences. This cognitive neuroscience framework provides a new hypothesis with which to consider the purpose of memory, and can help explain a variety of phenomena, ranging from recognition to first impressions, and from the brain's ‘default mode’ to a host of mental disorders.
APA, Harvard, Vancouver, ISO, and other styles
17

Hui, Kathleen K. S., Ovidiu Marina, Joshua D. Claunch, Erika E. Nixon, Jiliang Fang, Jing Liu, Ming Li, et al. "Corrigendum to “Acupuncture mobilizes the brain's default mode and its anti-correlated network in healthy subjects” [Brain Res. 1287 (2009) 84–103]." Brain Research 1308 (January 2010): 185. http://dx.doi.org/10.1016/j.brainres.2009.10.017.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Turkheimer, Federico E., Robert Leech, Paul Expert, Louis-David Lord, and Anthony C. Vernon. "The brain's code and its canonical computational motifs. From sensory cortex to the default mode network: A multi-scale model of brain function in health and disease." Neuroscience & Biobehavioral Reviews 55 (August 2015): 211–22. http://dx.doi.org/10.1016/j.neubiorev.2015.04.014.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Hou, Z., W. Jiang, Y. Yue, Y. Yin, Y. Zhang, Y. Sui, and Y. Yuan. "Decreased interhemispheric functional coordination underlying the cognitive impairment in late-onset depression." European Psychiatry 33, S1 (March 2016): S197. http://dx.doi.org/10.1016/j.eurpsy.2016.01.462.

Full text
Abstract:
BackgroundThe intuitive association between cognitive dysfunction in late onset depression (LOD) and the aberrant functional activity in the brain's default-mode network (DMN) has prompted interest in exploring the role of the DMN in LOD. The altered pattern of resting state voxel-mirrored homotopic connectivity (VMHC) in cognitive processes is not yet well understood in LOD.MethodsThe study was designed to examine the implicit coupling between the alteration of interhemispheric functional coordination and cognitive impairment in LOD. Thirty-one LOD patients and 37 matched healthy controls (HC) underwent neuropsychological tests and functional magnetic resonance imaging (fMRI) in this study.ResultsCompared to HC group, attenuated VMHC in superior frontal gyrus, superior temporal gyrus, posterior cerebellar lobe, postcentral and precentral gyrus was observed in LOD. Neuro-behavioral relevancy approach revealed that the imbalanced interhemispheric functional coordination in bilateral cerebellum was positively correlated with the performance of trail making test in LOD (r = 0.367, P = 0.040).ConclusionAltered linkage pattern of intrinsic homotopic connectivity and cognition was firstly investigated in LOD, and it would provide a novel clue to reveal the neural substrates underlying the cognitive dysfunction in LOD.Disclosure of interestThe authors have not supplied their declaration of competing interest.
APA, Harvard, Vancouver, ISO, and other styles
20

Koelsch, Stefan, Jessica R. Andrews‐Hanna, and Stavros Skouras. "Tormenting thoughts: The posterior cingulate sulcus of the default mode network regulates valence of thoughts and activity in the brain's pain network during music listening." Human Brain Mapping 43, no. 2 (October 15, 2021): 773–86. http://dx.doi.org/10.1002/hbm.25686.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Quah-Smith, Im, Chao Suo, Mark A. Williams, and Perminder S. Sachdev. "The Antidepressant Effect of Laser Acupuncture: A Comparison of the Resting Brain's Default Mode Network in Healthy and Depressed Subjects During Functional Magnetic Resonance Imaging." Medical Acupuncture 25, no. 2 (April 2013): 124–33. http://dx.doi.org/10.1089/acu.2012.0901.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Lambert, Steve. "Emotional awareness amongst middle leadership." Journal of Work-Applied Management 12, no. 2 (May 26, 2020): 233–43. http://dx.doi.org/10.1108/jwam-02-2020-0009.

Full text
Abstract:
PurposeThe purpose of this viewpoint paper is to explore middle leaders' ability to recognise emotions in the context of workplace research, and to propose measures that might support them in their role.Design/methodology/approachThis paper combines a contemporary literature review with reflections from practice to develop more nuanced understandings of middle leadership. This paper applied the Geneva Emotional Recognition Test (GERT) to explore the level of emotional recognition of 86 individuals (teachers to headteachers (equivalent to school principals)).FindingsThe preliminary findings suggest that teachers and headteachers have higher levels of emotional recognition than middle and senior leaders. This paper subsequently argues that the task-orientated nature middle leadership compounds an individual's ability to engage effectively in relationship-orientated tasks. This explains why middle leaders scored lower on the GERT assessment. This is further inhibited by the anti-correlation in the brain's ability to deal with the task-positive network (TDM) and default mode network (DMN) processing functions where individuals operate in one neural mode for long periods.Research limitations/implicationsThe viewpoint paper proposes a number of implications for middle leaders and suggests that middle leaders should proactively seek out opportunities to be engaged in activities that support the DMN function of the brain and subsequently the relationship-orientated aspects of leadership, for example, coaching other staff members. However, it has to be recognised that the sample size is small and further work is needed before any generalisations can be made.Originality/valueThis paper offers a contemporary review of the role of middle leaders underpinned by a preliminary study into individuals' ability to recognise emotions.
APA, Harvard, Vancouver, ISO, and other styles
23

Fox, Michael D., Dongyang Zhang, Abraham Z. Snyder, and Marcus E. Raichle. "The Global Signal and Observed Anticorrelated Resting State Brain Networks." Journal of Neurophysiology 101, no. 6 (June 2009): 3270–83. http://dx.doi.org/10.1152/jn.90777.2008.

Full text
Abstract:
Resting state studies of spontaneous fluctuations in the functional MRI (fMRI) blood oxygen level dependent (BOLD) signal have shown great promise in mapping the brain's intrinsic, large-scale functional architecture. An important data preprocessing step used to enhance the quality of these observations has been removal of spontaneous BOLD fluctuations common to the whole brain (the so-called global signal). One reproducible consequence of global signal removal has been the finding that spontaneous BOLD fluctuations in the default mode network and an extended dorsal attention system are consistently anticorrelated, a relationship that these two systems exhibit during task performance. The dependence of these resting-state anticorrelations on global signal removal has raised important questions regarding the nature of the global signal, the validity of global signal removal, and the appropriate interpretation of observed anticorrelated brain networks. In this study, we investigate several properties of the global signal and find that it is, indeed, global, not residing preferentially in systems exhibiting anticorrelations. We detail the influence of global signal removal on resting state correlation maps both mathematically and empirically, showing an enhancement in detection of system-specific correlations and improvement in the correspondence between resting-state correlations and anatomy. Finally, we show that several characteristics of anticorrelated networks including their spatial distribution, cross-subject consistency, presence with modified whole brain masks, and existence before global regression are not attributable to global signal removal and therefore suggest a biological basis.
APA, Harvard, Vancouver, ISO, and other styles
24

Bhat, Dhananjaya I., B. Indira Devi, Komal Bharti, and Rajanikant Panda. "Cortical plasticity after brachial plexus injury and repair: a resting-state functional MRI study." Neurosurgical Focus 42, no. 3 (March 2017): E14. http://dx.doi.org/10.3171/2016.12.focus16430.

Full text
Abstract:
OBJECTIVE The authors aimed to understand the alterations of brain resting-state networks (RSNs) in patients with pan–brachial plexus injury (BPI) before and after surgery, which might provide insight into cortical plasticity after peripheral nerve injury and regeneration. METHODS Thirty-five patients with left pan-BPI before surgery, 30 patients after surgery, and 25 healthy controls underwent resting-state functional MRI (rs-fMRI). The 30 postoperative patients were subdivided into 2 groups: 14 patients with improvement in muscle power and 16 patients with no improvement in muscle power after surgery. RSNs were extracted using independent component analysis to evaluate connectivity at a significance level of p < 0.05 (familywise error corrected). RESULTS The patients with BPI had lower connectivity in their sensorimotor network (SMN) and salience network (SN) and greater connectivity in their default mode network (DMN) before surgery than the controls. Connectivity of the left supplementary motor cortex in the SMN and medial frontal gyrus and in the anterior cingulate cortex in the SN increased in patients whose muscle power had improved after surgery, whereas no significant changes were noted in the unimproved patients. There was a trend toward reduction in DMN connectivity in all the patients after surgery compared with that in the preoperative patients; however, this result was not statistically significant. CONCLUSIONS The results of this study highlight the fact that peripheral nerve injury, its management, and successful treatment cause dynamic changes within the brain's RSNs, which includes not only the obvious SMN but also the higher cognitive networks such as the SN and DMN, which indicates brain plasticity and compensatory mechanisms at work.
APA, Harvard, Vancouver, ISO, and other styles
25

Legostaeva, Liudmila A., Elena I. Kremneva, Dmitry O. Sinitsyn, Elizaveta G. Iazeva, Dmitry V. Sergeev, Alexandra G. Poydasheva, Ilya S. Bakulin, et al. "Features of residual brain activity in patients with chronic disorders of consciousness on resting-state functional MRI." Annals of Clinical and Experimental Neurology 16, no. 2 (June 30, 2022): 15–24. http://dx.doi.org/10.54101/acen.2022.2.2.

Full text
Abstract:
Introduction. Rapid advances in critical care medicine have led to an increased survival rate of patients with severe brain damage and, consequently, to an increased prevalence of chronic disorders of consciousness (CDC). The lack of or fluctuations in signs of consciousness, which accompany the restoration of alertness after recovery from coma, indicate whether the type of CDC is a vegetative state or minimally conscious state. Correct diagnosis determines not only the rehabilitation outcome but also the economic outlook for a particular patient. However, the subjective nature of signs of consciousness, which are identified during clinical examination using neurological scales, is a common cause of diagnostic errors. The study of spontaneous activity using resting-state functional magnetic resonance imaging (fMRI) has helped to identify resting state networks. The default mode network (DMN) is one of the most studied brain networks. Its signal can change or be absent in patients with various types of CDC. Purpose. To study the signal of residual spontaneous brain activity in patients with CDC at rest. Materials and methods. Twenty-two patients with permanent CDC underwent resting state fMRI as an additional tool in the differential diagnosis between vegetative state and minimally conscious state at the Research Centre of Neurology. Results. It was found that the nature of the signal coming from anatomical regions that are part of the DMN changes when signs of consciousness emerge. Conclusion. These changes confirm that resting state fMRI is an important additional tool for differential diagnosis of CDC types. Accumulating knowledge about the brain's functional state helps us to expand our overall understanding of the nature of consciousness.
APA, Harvard, Vancouver, ISO, and other styles
26

Immordino-Yang, Mary Helen. "Emotion, Sociality, and the Brain’s Default Mode Network." Policy Insights from the Behavioral and Brain Sciences 3, no. 2 (July 7, 2016): 211–19. http://dx.doi.org/10.1177/2372732216656869.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Lu, H., Q. Zou, H. Gu, M. E. Raichle, E. A. Stein, and Y. Yang. "Rat brains also have a default mode network." Proceedings of the National Academy of Sciences 109, no. 10 (February 21, 2012): 3979–84. http://dx.doi.org/10.1073/pnas.1200506109.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Lochtenberg, Jessica, Ari Kirshenbaum, and Matthew Johnson. "Adjunct pharmacotherapy for psychotherapy." BJPsych Open 7, S1 (June 2021): S166—S167. http://dx.doi.org/10.1192/bjo.2021.460.

Full text
Abstract:
AimsA variety of pharmacotherapies have been used to assist the psychotherapy process as “adjunctive therapies.” These drugs are used in an acute, targeted fashion, such that they are explicitly delivered in the context of psychotherapy for anxiety, mood and substance-dependence disorders (SUDs). Our narrative review highlights the potential of medically-assisted psychotherapy by outlining the current state of research on few of these medications and describing the basic science that supports their use.MethodFirstly, we researched an assortment of medications that have been used off-label to enhance psychotherapy, and selected a few that have received the most empirical attention in preclinical and clinical-trial settings. Our review of clinical trials focused on three of the most common psychiatric ailments. For all studies reviewed, we identify the strengths and weaknesses of the data supporting the use of the medications for the three aforementioned disorders.ResultD-cycloserine: accelerates the process of associative emotional learning, enhancing exposure therapy in the treatment of various anxiety disorders, including obsessive-compulsive disorder and posttraumatic stress disorder. Limited studies are available on efficacy in treating SUDs.Intranasal oxytocin: accelerates memory retrieval-extinction procedures used in posttraumatic stress disorder, and promotes prosocial cognition and behaviour, facilitating a therapeutic alliance. Sufficiently powered studies and safety studies are required before strong conclusions can be made.Propranolol: interrupts the reconsolidation of memories (leading to maladaptive learned responses) involved in posttraumatic stress disorder during memory-reactivation therapy sessions, but there is little evidence that this drug can be used for depression or SUDs.Psychedelics: may effect the brain's default mode network, engendering a transformative experience that is often followed by a reduction in psychiatric symptoms. 3,4-methylenedioxymethamphetamine may additionally modulate the amygdala response in a way that allows for reprocessing of traumatic memories, and improves the therapeutic alliance. Anxiety, mood, and SUDs appear to be positively influence by traditional and non-traditional (ketamine) psychedelics.ConclusionAlthough the efficacy of the medically-assisted psychotherapies reviewed is still under investigation, we propose that these novel treatment approaches may be preferred over traditional psychopharmacological treatments due to the presence of fewer chronic side effects, as well less toxicity and abuse potential. Furthermore, these adjunctive pharmacotherapies may help to reinforce the psychotherapeutic alliance and may ultimately yield better long-term treatment outcomes. If at least some of the adjunctive pharmacotherapies outlined in this review are found to be clinically efficacious and safe, patients will benefit from having more treatment options available to them in the future.
APA, Harvard, Vancouver, ISO, and other styles
29

Duarte, Antonio. "Musement: The activity of the brain’s default mode network." Semiotica 2020, no. 233 (March 26, 2020): 145–58. http://dx.doi.org/10.1515/sem-2018-0118.

Full text
Abstract:
AbstractThe main purpose of this article is to identify the inner human activity Peirce calls musement with the mental processes that arise through the workings of the brain’s default mode network. This network is a specific, recently anatomically defined brain system, which is most active when individuals are not focused on the external environment. In doing so, musement, which was defined by Peirce over a hundred years ago, will finally be situated within what today we understand as its neurobiological origin.
APA, Harvard, Vancouver, ISO, and other styles
30

Postema, Merel Charlotte, Matteo De Marco, Elisa Colato, and Annalena Venneri. "A study of within-subject reliability of the brain’s default-mode network." Magnetic Resonance Materials in Physics, Biology and Medicine 32, no. 3 (February 7, 2019): 391–405. http://dx.doi.org/10.1007/s10334-018-00732-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Aso, Toshihiko, and Hidenao Fukuyama. "Functional Heterogeneity in the Default Mode Network Edges." Brain Connectivity 5, no. 4 (May 2015): 203–13. http://dx.doi.org/10.1089/brain.2014.0256.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Vatansever, Deniz, Anne Manktelow, Barbara J. Sahakian, David K. Menon, and Emmanuel A. Stamatakis. "Default Mode Network Engagement Beyond Self-Referential Internal Mentation." Brain Connectivity 8, no. 4 (May 2018): 245–53. http://dx.doi.org/10.1089/brain.2017.0489.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Gorges, M., H. P. Müller, A. C. Ludolph, E. Pinkhardt, and J. Kassubek. "P19. The brain’s default mode is associated with executive oculomotor deficits in Parkinson’s disease." Clinical Neurophysiology 126, no. 8 (August 2015): e95-e96. http://dx.doi.org/10.1016/j.clinph.2015.04.138.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Anderson, Jeffrey S., Michael A. Ferguson, Melissa Lopez-Larson, and Deborah Yurgelun-Todd. "Connectivity Gradients Between the Default Mode and Attention Control Networks." Brain Connectivity 1, no. 2 (August 2011): 147–57. http://dx.doi.org/10.1089/brain.2011.0007.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Tylén, K., P. Christensen, A. Roepstorff, T. Lund, S. Østergaard, and M. Donald. "Brains striving for coherence: Long-term cumulative plot formation in the default mode network." NeuroImage 121 (November 2015): 106–14. http://dx.doi.org/10.1016/j.neuroimage.2015.07.047.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Gordon, Evan M., Timothy O. Laumann, Scott Marek, Ryan V. Raut, Caterina Gratton, Dillan J. Newbold, Deanna J. Greene, et al. "Default-mode network streams for coupling to language and control systems." Proceedings of the National Academy of Sciences 117, no. 29 (July 6, 2020): 17308–19. http://dx.doi.org/10.1073/pnas.2005238117.

Full text
Abstract:
The human brain is organized into large-scale networks identifiable using resting-state functional connectivity (RSFC). These functional networks correspond with broad cognitive domains; for example, the Default-mode network (DMN) is engaged during internally oriented cognition. However, functional networks may contain hierarchical substructures corresponding with more specific cognitive functions. Here, we used individual-specific precision RSFC to test whether network substructures could be identified in 10 healthy human brains. Across all subjects and networks, individualized network subdivisions were more valid—more internally homogeneous and better matching spatial patterns of task activation—than canonical networks. These measures of validity were maximized at a hierarchical scale that contained ∼83 subnetworks across the brain. At this scale, nine DMN subnetworks exhibited topographical similarity across subjects, suggesting that this approach identifies homologous neurobiological circuits across individuals. Some DMN subnetworks matched known features of brain organization corresponding with cognitive functions. Other subnetworks represented separate streams by which DMN couples with other canonical large-scale networks, including language and control networks. Together, this work provides a detailed organizational framework for studying the DMN in individual humans.
APA, Harvard, Vancouver, ISO, and other styles
37

Sharp, David J., Christian F. Beckmann, Richard Greenwood, Kirsi M. Kinnunen, Valerie Bonnelle, Xavier De Boissezon, Jane H. Powell, Serena J. Counsell, Maneesh C. Patel, and Robert Leech. "Default mode network functional and structural connectivity after traumatic brain injury." Brain 134, no. 8 (August 2011): 2233–47. http://dx.doi.org/10.1093/brain/awr175.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Bigler, Erin D. "Default mode network, connectivity, traumatic brain injury and post-traumatic amnesia." Brain 139, no. 12 (December 2016): 3054–57. http://dx.doi.org/10.1093/brain/aww277.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Chiong, Winston, Stephen M. Wilson, Mark D’Esposito, Andrew S. Kayser, Scott N. Grossman, Pardis Poorzand, William W. Seeley, Bruce L. Miller, and Katherine P. Rankin. "The salience network causally influences default mode network activity during moral reasoning." Brain 136, no. 6 (April 9, 2013): 1929–41. http://dx.doi.org/10.1093/brain/awt066.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Luo, Li, Lai Xu, Rex Jung, Godfrey Pearlson, Tülay Adali, and Vince D. Calhoun. "Constrained Source-Based Morphometry Identifies Structural Networks Associated with Default Mode Network." Brain Connectivity 2, no. 1 (February 2012): 33–43. http://dx.doi.org/10.1089/brain.2011.0026.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

McKinnon, Andrew C., Jim Lagopoulos, Zoe Terpening, Ron Grunstein, Ian B. Hickie, Jennifer Batchelor, Simon J. G. Lewis, Shantel Duffy, James M. Shine, and Sharon L. Naismith. "Sleep disturbance in mild cognitive impairment is associated with alterations in the brain’s default mode network." Behavioral Neuroscience 130, no. 3 (June 2016): 305–15. http://dx.doi.org/10.1037/bne0000137.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Lim, Tow Keang. "The predictive brain model in diagnostic reasoning." Asia Pacific Scholar 6, no. 2 (May 4, 2021): 1–8. http://dx.doi.org/10.29060/taps.2021-6-2/ra2370.

Full text
Abstract:
Introduction: Clinical diagnosis is a pivotal and highly valued skill in medical practice. Most current interventions for teaching and improving diagnostic reasoning are based on the dual process model of cognition. Recent studies which have applied the popular dual process model to improve diagnostic performance by “Cognitive De-biasing” in clinicians have yielded disappointing results. Thus, it may be appropriate to also consider alternative models of cognitive processing in the teaching and practice of clinical reasoning. Methods: This is critical-narrative review of the predictive brain model. Results: The theory of predictive brains is a general, unified and integrated model of cognitive processing based on recent advances in the neurosciences. The predictive brain is characterised as an adaptive, generative, energy-frugal, context-sensitive action-orientated, probabilistic, predictive engine. It responds only to predictive errors and learns by iterative predictive error management, processing and hierarchical neural coding. Conclusion: The default cognitive mode of predictive processing may account for the failure of de-biasing since it is not thermodynamically frugal and thus, may not be sustainable in routine practice. Exploiting predictive brains by employing language to optimise metacognition may be a way forward
APA, Harvard, Vancouver, ISO, and other styles
43

Geng, Xiujuan, Yuzheng Hu, Hong Gu, Betty Jo Salmeron, Bryon Adinoff, Elliot A. Stein, and Yihong Yang. "Salience and default mode network dysregulation in chronic cocaine users predict treatment outcome." Brain 140, no. 5 (February 20, 2017): 1513–24. http://dx.doi.org/10.1093/brain/awx036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Mak, Lauren E., Luciano Minuzzi, Glenda MacQueen, Geoffrey Hall, Sidney H. Kennedy, and Roumen Milev. "The Default Mode Network in Healthy Individuals: A Systematic Review and Meta-Analysis." Brain Connectivity 7, no. 1 (February 2017): 25–33. http://dx.doi.org/10.1089/brain.2016.0438.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Wang, Jia-Xi, Jin-Ying Zhuang, Lulu Fu, Qin Lei, and Weidong Zhang. "Association of ovarian hormones with mapping concept of self and others in the brain’s default mode network." NeuroReport 31, no. 10 (May 21, 2020): 717–23. http://dx.doi.org/10.1097/wnr.0000000000001477.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Carhart-Harris, R. L., and K. J. Friston. "The default-mode, ego-functions and free-energy: a neurobiological account of Freudian ideas." Brain 133, no. 4 (February 28, 2010): 1265–83. http://dx.doi.org/10.1093/brain/awq010.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

De Simoni, Sara, Patrick J. Grover, Peter O. Jenkins, Lesley Honeyfield, Rebecca A. Quest, Ewan Ross, Gregory Scott, et al. "Disconnection between the default mode network and medial temporal lobes in post-traumatic amnesia." Brain 139, no. 12 (October 22, 2016): 3137–50. http://dx.doi.org/10.1093/brain/aww241.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Weaver, Kurt E., Andrew Poliakov, Edward J. Novotny, Jared D. Olson, Thomas J. Grabowski, and Jeffrey G. Ojemann. "Electrocorticography and the early maturation of high-frequency suppression within the default mode network." Journal of Neurosurgery: Pediatrics 21, no. 2 (February 2018): 133–40. http://dx.doi.org/10.3171/2017.7.peds17269.

Full text
Abstract:
OBJECTIVEThe acquisition and refinement of cognitive and behavioral skills during development is associated with the maturation of various brain oscillatory activities. Most developmental investigations have identified distinct patterns of low-frequency electrophysiological activity that are characteristic of various behavioral milestones. In this investigation, the authors focused on the cross-sectional developmental properties of high-frequency spectral power from the brain’s default mode network (DMN) during goal-directed behavior.METHODSThe authors contrasted regionally specific, time-evolving high gamma power (HGP) in the lateral DMN cortex between 3 young children (age range 3–6 years) and 3 adults by use of electrocorticography (ECoG) recordings over the left perisylvian cortex during a picture-naming task.RESULTSAcross all participants, a nearly identical and consistent response suppression of HGP, which is a functional signature of the DMN, was observed during task performance recordings acquired from ECoG electrodes placed over the lateral DMN cortex. This finding provides evidence of relatively early maturation of the DMN. Furthermore, only HGP relative to evoked alpha and beta band power showed this level of consistency across all participants.CONCLUSIONSRegionally specific, task-evoked suppression of the high-frequency components of the cortical power spectrum is established early in brain development, and this response may reflect the early maturation of specific cognitive and/or computational mechanisms.
APA, Harvard, Vancouver, ISO, and other styles
49

Kiviniemi, Vesa, Tapani Vire, Jukka Remes, Ahmed Abou Elseoud, Tuomo Starck, Osmo Tervonen, and Juha Nikkinen. "A Sliding Time-Window ICA Reveals Spatial Variability of the Default Mode Network in Time." Brain Connectivity 1, no. 4 (October 2011): 339–47. http://dx.doi.org/10.1089/brain.2011.0036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Karten, Ariel, Spiro P. Pantazatos, David Khalil, Xian Zhang, and Joy Hirsch. "Dynamic Coupling Between the Lateral Occipital-Cortex, Default-Mode, and Frontoparietal Networks During Bistable Perception." Brain Connectivity 3, no. 3 (June 2013): 286–93. http://dx.doi.org/10.1089/brain.2012.0119.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography