Week 1-3 Readings
Week 1-3 Readings
Week 1-3 Readings
http://www.scienticamerican.com/article/gut-second-brain/?p...
ADVERTISEMENT
As Olympians go for the gold in Vancouver, even the steeliest are likely to experience
that familiar feeling of "butterflies" in the stomach. Underlying this sensation is an
often-overlooked network of neurons lining our guts that is so extensive some
scientists have nicknamed it our "second brain".
A deeper understanding of this mass of neural tissue, filled with important
neurotransmitters, is revealing that it does much more than merely handle digestion
or inflict the occasional nervous pang. The little brain in our innards, in connection
with the big one in our skulls, partly determines our mental state and plays key roles
in certain diseases throughout the body.
Although its influence is far-reaching, the second brain is not the seat of any
conscious thoughts or decision-making.
"The second brain doesn't help with the great thought processesreligion,
philosophy and poetry is left to the brain in the head," says Michael Gershon,
ISTOCKPHOTO/ERAXION
ADVERTISEMENT
1 of 3
20/07/2015 2:32 pm
http://www.scienticamerican.com/article/gut-second-brain/?p...
physiology, psychiatry and biobehavioral sciences at the David Geffen School of Medicine at the University of California, Los Angeles
(U.C.L.A.). For example, scientists were shocked to learn that about 90 percent of the fibers in the primary visceral nerve, the vagus,
carry information from the gut to the brain and not the other way around. "Some of that info is decidedly unpleasant," Gershon says.
The second brain informs our state of mind in other more obscure ways, as well. "A big part of our emotions are probably influenced by
the nerves in our gut," Mayer says. Butterflies in the stomachsignaling in the gut as part of our physiological stress response, Gershon
saysis but one example. Although gastrointestinal (GI) turmoil can sour one's moods, everyday emotional well-being may rely on
messages from the brain below to the brain above. For example, electrical stimulation of the vagus nervea useful treatment for
depressionmay mimic these signals, Gershon says.
Given the two brains' commonalities, other depression treatments that target the mind can unintentionally impact the gut. The enteric
nervous system uses more than 30 neurotransmitters, just like the brain, and in fact 95 percent of the body's serotonin is found in the
bowels. Because antidepressant medications called selective serotonin reuptake inhibitors (SSRIs) increase serotonin levels, it's little
wonder that meds meant to cause chemical changes in the mind often provoke GI issues as a side effect. Irritable bowel syndrome
which afflicts more than two million Americansalso arises in part from too much serotonin in our entrails, and could perhaps be
regarded as a "mental illness" of the second brain.
Scientists are learning that the serotonin made by the enteric nervous system might also play a role in more surprising diseases: In a
new Nature Medicine study published online February 7, a drug that inhibited the release of serotonin from the gut counteracted the
bone-deteriorating disease osteoporosis in postmenopausal rodents. (Scientific American is part of Nature Publishing Group.) "It was
totally unexpected that the gut would regulate bone mass to the extent that one could use this regulation to cureat least in rodents
osteoporosis," says Gerard Karsenty, lead author of the study and chair of the Department of Genetics and Development at Columbia
University Medical Center.
Serotonin seeping from the second brain might even play some part in autism, the developmental disorder often first noticed in early
childhood. Gershon has discovered that the same genes involved in synapse formation between neurons in the brain are involved in the
alimentary synapse formation. "If these genes are affected in autism," he says, "it could explain why so many kids with autism have GI
motor abnormalities" in addition to elevated levels of gut-produced serotonin in their blood.
Down the road, the blossoming field of neurogastroenterology will likely offer some new insight into the workings of the second
brainand its impact on the body and mind. "We have never systematically looked at [the enteric nervous system] in relating lesions in
it to diseases like they have for the" central nervous system, Gershon says. One day, perhaps there will be well-known connections
between diseases and lesions in the gut's nervous system as some in the brain and spinal cord today indicate multiple sclerosis.
Cutting-edge research is currently investigating how the second brain mediates the body's immune response; after all, at least 70
percent of our immune system is aimed at the gut to expel and kill foreign invaders.
U.C.L.A.'s Mayer is doing work on how the trillions of bacteria in the gut "communicate" with enteric nervous system cells (which they
greatly outnumber). His work with the gut's nervous system has led him to think that in coming years psychiatry will need to expand to
treat the second brain in addition to the one atop the shoulders.
So for those physically skilled and mentally strong enough to compete in the Olympic Gamesas well as those watching at homeit
may well behoove us all to pay more heed to our so-called "gut feelings" in the future.
SEE ALSO:
Energy & Sustainability: 5 Steps to Feed the World and Sustain the Planet | Evolution: New Fossil Reveals Velociraptor Sported
Feathers | Mind & Brain: Nail Biting May Arise from Perfectionism | Space: Europa's "Brown Gunk" Suggests a Briny
Sea | Technology: Timeline: The Amazing Multimillion-Year History of Processed Food | More Science: The Flavor Connection
2 of 3
20/07/2015 2:32 pm
http://www.scienticamerican.com/article/gut-second-brain/?p...
The Problem with Female Superheroes 3 weeks ago scienticamerican.com ScienticAmerican.com Mind & Brain
2.
Metacognition Is the Forgotten Secret to Success 11 months ago scienticamerican.com ScienticAmerican.com Features
What Kind of Introvert Are You? 9 months ago blogs.scienticamerican.com ScienticAmerican.com Jennifer Odessa Grimes
Subscribe Now
3 of 3
20/07/2015 2:32 pm
EDITORIAL
761
762
that sickness behaviour is not merely an automatic, inflexible reaction to illness, but rather the
expression of a central motivational state. That is, animals will actively learn responses to be able
to engage in these behaviours when they are infected and will choose conditions under which they
can occur (Dantzer et al. 1996 ; Maier & Watkins, 1999).
ACUTE SICKNESS BEHAVIOUR IS IMMUNOLOGICALLY MEDIATED
The production of the pro-inflammatory cytokines such as tumour necrosis factor (TNF),
interleukin-1 (IL-1) and IL-6 by activated immune cells (monocyte}macrophages, lymphocytes) is
an integral part of the host response to infection. These cytokines act as messenger molecules and
play a pivotal role in the orchestration of the acute phase response (Dinarello, 1997 ; Papanicolaou,
1998). In recent years, it has also become clear that the necessary synchrony between metabolic,
physiological and behavioural aspects of the individuals response to infection depends on the
activities of these same cytokines. It has long been known that fever is not caused directly by
invading pathogens but rather by the action of soluble immunological products. These endogenous
pyrogens were subsequently identified as being pro-inflammatory cytokines, with IL-1 the most
potent pyrogenic agent (Dinarello et al. 1977). The first indication of a role for cytokines in the
induction of acute sickness behaviour came from clinical trials with purified or recombinant
cytokines. Administration of cytokines in the treatment of cancer and chronic viral infections such
as hepatitis B and C produced a syndrome of adverse effects, including fever, fatigue, malaise,
headaches, anorexia, depression and, at high doses, delirium (Renault & Hoofnagel, 1989 ;
Dinarello, 1997). However, as these clinical observations were made predominantly on patients with
significant medical illnesses, generalization from these findings to argue for a role for cytokines
in normal sickness behaviour is problematical.
Animal experiments have confirmed that most aspects of sickness behaviour can be induced in
a dose-dependent fashion by systemic or intracerebral injections of pro-inflammatory cytokines,
particularly IL-1 (Dantzer et al. 1996). Systemic injections with lipopolysaccharide (LPS), a strong
inducer of the synthesis and release of most pro-inflammatory cytokines, also produces the full
spectrum of sickness behaviour. Conversely, administration of specific antagonists (e.g. IL-1
receptor antagonist (IL-1Ra)) inhibits many of the central effects of this cytokine (Rothwell &
Hopkins, 1995). Given the pleiotropism and redundancy of the cytokine network, it has been
difficult to determine the precise contribution of individual cytokines to sickness behaviour.
However, there is evidence to suggests that TNF- and IL-1-, which are synthesized very early in
the immune response, are more potent than those that are induced later (e.g. IL-6) and act in
synergy to produce sickness behaviour (Dantzer et al. 1996). IL-6 appears to require the presence
of other pro-inflammatory cytokines (e.g. IL-1) to produce behavioural symptoms of sickness
(Bluthe et al. 1998). The role of IL-6 in sickness behaviour is likely to be complex, as recent evidence
suggests that IL-6 has both pro- and anti-inflammatory functions, the latter including the induction
of natural antagonists to TNF and IL-1 and stimulation of the hypothalamicpituitaryadrenal
(HPA) axis (Papanicolaou, 1998).
LINKS BETWEEN ACUTE SICKNESS BEHAVIOUR AND THE STRESS RESPONSE
Experimental administration of pro-inflammatory cytokines or LPS to animals has revealed that,
in addition to acute sickness behaviour, a classic stress response is produced. This is characterized
by activation of the sympathetic nervous system and release of plasma catecholamines as well as
activation of the HPA system leading to the release of adrenocorticotrophic hormone (ACTH) and
glucocorticoids (Dunn, 1995 ; Maier & Watkins, 1999). In humans, IL-6 appears to be a particularly
potent stimulator of the HPA axis. For example, daily administration of recombinant IL-6 over a
week was found to produce a remarkable activation and enlargement of the adrenal glands, similar
to that seen in patients with Cushings disease (Mastorakos et al. 1993). Because glucocorticoids,
in turn, exert negative feedback on the secretion of IL-6, it has been argued that this cytokine
displays the traditional characteristics of a hormone (Papanicolaou, 1998).
763
It is well established that exposure to stressors may result in the suppression of cell-mediated
immune responses (Kusnekov & Rabin, 1994). However, new evidence suggests that the sequelae
of exposure to acute stressors are considerably more complex than previously thought and include
aspects of immune activation not unlike the host response to acute infections (Maier & Watkins,
1998 ; Maes, 1999 ; Tringali et al. 2000, for reviews). Briefly, in both humans and animals, diverse
physical and psychosocial stressors were found to produce leucocytosis, a shift in liver metabolism
toward production of acute phase proteins, secretion of pro-inflammatory cytokines and fever.
Moreover, behavioural changes including anorexia, decreases in activation, exploration, social
interaction and aggression, depressed affect and cognitive impairment, bear a strong resemblance
to cytokine-induced sickness behaviour and appear more adaptive in the context of sickness then
as a component of the frightfight response.
The similarities between aspects of the stress response and acute sickness behaviour point to a
common mediator. Indeed, a direct role for pro-inflammatory cytokines in the stress response was
established by experiments blocking specific cytokine receptors in the brain during exposure to a
stressor. For example, injections of the IL-1 receptor antagonist (IL-1Ra) abolished the learned
helplessness response and exaggerated fear typically induced by inescapable shock (Maier &
Watkins, 1995), as well as blunting the pituitaryadrenal response and inhibiting the release of
hypothalamic monoamines normally associated with immobilization stress (Shintani et al. 1995).
There is also evidence that the stress-induced secretion of cytokines in hypothalamic structures and
the periphery are mediated by catecholaminergic mechanisms such as the activation of -adrenergic
receptors (Takaki et al. 1994 ; Papanicolaou, 1998 ; Tringali et al. 2000).
There remains little doubt that the systems mediating the organisms defence against infection
and injury on the one hand, and against psychosocial and environmental stressors on the other, are
closely connected. Indeed, it has been proposed that environmental (i.e. external ) stressors
essentially activate the same neuroendorineimmune circuitry as immunogenic agents (i.e.
internal stressors), although they may enter the circuit at different sites (Maier & Watkins, 1998 ;
Maes, 1999). The end product of this activation (in terms of neuroendocrine, behavioural, cognitive,
emotional, or immunological consequences which are clearly different for different trigger stimuli)
presumably depends on the specific characteristics of the trigger (Maier &Watkins, 1998 ; Tringali
et al. 2000).
From an evolutionary perspective, the similarities in the response to stress and acute illness may
be understood by studying the systems subserving adaptation and defence in primitive organisms.
Even in the most primitive organisms such as molluscs and sponges, which are incapable of
elaborate defence against distal threats, some form of host response against infection and injury can
be identified (Beck et al. 1994). Moreover, there is evidence that as early in evolution as the molluscs,
defence against infection and injury involved pro-inflammatory cytokines in bi-directional
communication between immunological and neural structures, as well as the release peptides
traditionally viewed as stress hormones (Clatworthy, 1996 ; Maier & Watkins, 1998, for reviews).
When the frightflight response evolved later in more complex organisms, existing mechanisms such
as cytokine-based communication networks appear to have been incorporated in this new
adaptation (Maier & Watkins, 1998). This account, although speculative, offers an explanation for
the mobilization of inflammatory processes during acute stress, and provides a basis for the
manifestation of seemingly maladaptive behavioural changes (i.e. sickness behaviour) after
exposure to a significant stressor.
POSSIBLE MECHANISMS UNDERLYING THE PRODUCTION OF ACUTE SICKNESS
BEHAVIOUR
In view of their documented involvement in diverse centrally-mediated phenomena, there is little
doubt that cytokines are capable of providing signals to the brain to alter neural activity. The
mechanism through which these peripherally produced molecules might act on the brain, however,
has been the subject of much debate. Cytokines are not thought to cross the bloodbrain barrier,
764
owing to their large molecular weight and hydrophilic properties. A number of specialized
mechanisms have been identified, however, that would allow blood-borne cytokines to signal the
brain. These include an active transport system across the barrier (Banks et al. 1991), and cytokine
entry at regions of the brain (e.g. circumventricular organs) where the barrier is weak or absent to
trigger the production of second messengers (e.g. prostaglandins) to neural targets (e.g.
hypothalamic regions) or to induce local cytokine production (Saper & Breder, 1994). As an
alternative to the bloodborne route, several authors have pursued the possibility of direct neural
signalling via the vagus nerve (Dantzer et al. 1996 ; Maier & Watkins, 1999). Whatever the route
used to signal the brain, most authors agree that local production of cytokines by glial cells plays
an important role in the induction of sickness behaviour. This offers the best explanation of the welldocumented increase in brain levels of cytokines after peripheral administration, as well as
accounting for the presence of cytokine-producing cells (glial cells) and specific binding sites
throughout the brain (Vollmer-Conna et al. 1998 ; Maier & Watkins, 1999). The exact mechanism
responsible to translate the immune signal into a neural signal is still unclear, but appears to involve
alterations in a variety of neuropeptide (e.g. corticotropin-releasing hormone (CRH), substance P,
opioids) and neurotransmitter systems (nonadrenaline, serotonin, gamma-aminobutyric acid
(GABA)) (Rothwell & Hopkins, 1995).
HUMAN STUDIES OF ACUTE SICKNESS BEHAVIOUR
To date, very few studies have systematically examined the potential relationships between proinflammatory cytokines and mental and behavioural symptoms in sick humans. Although animal
studies have contributed much to our understanding of acute sickness behaviour, they are poorly
suited to examine complex behaviours and}or more subtle changes in brain function (e.g. cognitive
deficits, mood alterations) that are nonetheless an integral part of sickness behaviour. Smith and
colleagues (1987, 1988) studied cognitive performance in healthy volunteers with experimentally
induced common cold (rhinovirus) or influenza, or an infusion of interferon-. Both influenza
infections and interferon- administration produced impairments in stimulus detection tasks,
whereas common colds were associated with impaired handeye coordination. Interestingly, these
cognitive performance deficits were also demonstrated in subjects with subclinical infections.
Interpretation of these findings is somewhat limited by the artificial context in which they were
obtained, and the use of a cytokine not generally considered one of the prototypic inflammatory
cytokines. They do suggest, however, that even minor infections with agents that do not directly
infect the brain, may be associated with significant impairment in cognitive performance. Moreover,
such impairment appears to be related to the action of cytokines.
Significant neurocognitive deficits have also been demonstrated in a clinical sample of patients
with acute infections (both EpsteinBarr virus (EBV), and influenza-like infections ; Vollmer-Conna
et al. 1997). In comparison to well-matched healthy controls, these patients reported a significant
increase in negative affect and fatigue, and showed deficits in selective attention, response speed,
pursuit tracking, and performance accuracy. A similar study recently demonstrated that patients
with influenza-like illnesses were impaired on aspects of everyday memory (Capuron et al. 1999).
These findings are suggestive only, as immunological correlates were not assessed and not all
infections were serologically documented. What is clearly needed is the unambiguous identification
of cognitive deficits associated with specific acute infections. In addition, the relationship between
such deficits and the action of cytokines must be examined.
CLINICAL AND PRACTICAL IMPLICATIONS OF THE BEHAVIOURAL AND MENTAL
EFFECTS OF CYTOKINES
An accumulation of evidence suggests that behavioural}mental changes typically associated with
acute infections (i.e. acute sickness behaviours) are immunologically mediated and form part of the
host defence. The efficacy of these behavioural disease-fighting strategies is demonstrated by the fact
765
that both animals and man have survived infections and injury through evolutionary history. A
better understanding of acute sickness behaviour and its underlying mechanism(s) will provide new
insights into how sickness and recovery processes are organized in the brain. This may lead to the
development of management practices that complement the innate disease-coping strategies, and
may be particularly beneficial when dealing with viruses or drug resistant bacteria (Hart, 1988).
Several reports suggest that acute infective illness is associated with significant cognitive
impairment (Smith et al. 1987 ; Volmer-Conna et al. 1997 ; Capuron et al. 1999). Yet, many patients
continue with their usual daily routine throughout an illness such as influenza, glandular fever or
Q fever. There is some indication from an ongoing large, prospective study of infective cohorts
(Bennett et al. 1998) that accidents in the workplace are more prevalent in the context of acute Q
fever (Lloyd, A. unpublished data). Elucidation of the full extent of behavioural and mental changes
during acute infections may thus have important implications for road and work place safety.
POTENTIALLY ABNORMAL MANIFESTATIONS OF SICKNESS BEHAVIOUR
Sickness behaviour, similar to the frightflight response, is viewed as a highly organized strategy
critical to survival. However, in the same way as pathological fear and anxiety are debilitating,
excessive sickness behaviour can be detrimental. Over the past decade there has been much
speculation on a possible role for cytokines in the pathogenesis of neuropsychiatric syndromes,
notably post-infective fatigue syndromes (e.g. Chao et al. 1991 ; Hickie & Lloyd, 1995 ; VollmerConna et al. 1998 ;) and, more recently, major depression (e.g. Maes et al. 1995 a ; Maes 1999 ;
Charlton, 2000).
Clinical and scientific interest in exploring the possibility of an immunological basis for
neuropsychiatric disorders was initially fuelled by the striking similarity between acute sickness
behaviours and key symptoms reported in depression and fatigue syndromes (i.e. loss of appetite,
malaise, psychomotor slowing, altered sleep, fatigue, anhedonia, depressed affect and cognitive
impairment). In addition, the discovery that psychosocial stressors can activate the inflammatory
response system has lent more credence to a proposed role for pro-inflammatory cytokines in the
pathogenesis of stress-related disorders such as depression (Maes, 1999).
Research examining immunological correlates of depression has produced evidence consistent
with the notion of immune activation, including elevated leukocyte counts and activation markers,
increased production of pro-inflammatory cytokines and acute phase proteins. Moreover, tricyclic
antidepressants and serotonin reuptake inhibitors appear to suppress these inflammatory responses
(Maes et al. 1995 a, for review). On the other hand, there is substantial evidence documenting
immunosuppression and increased susceptibility to disease in patients with major depression
(Herbert & Cohen, 1993). It has been suggested that the simultaneous signs of immune activation
and suppression in depression may be reconciled by reference to the action of multiple feedback
systems generated to contain the immune response or T cell exhaustion (Maes et al. 1995 b).
The emergence of immunological hypotheses for neuropsychiatric disorders, such as depression,
challenges traditional pathophysiological views, and provides a fresh perspective on symptomatology (e.g. depression as a variant of sickness behaviour). However, the data to date are not
sufficient to determine whether altered levels of pro-inflammatory cytokines play an aetiological
role, are the consequence of depression, or merely reflect an epiphenomenon. Interpretation of the
available evidence is limited by a variety of methodological problems including the predominant
reliance on cross-sectional, correlational designs, one-off sampling and the inevitable heterogeneity
of subject samples. An additional complication inherent in this type of research lies in the attempt
to establish associations between centrally-mediated symptoms and measurements obtained from
peripheral blood samples. Although such endeavours are justified, in principle, by the bi-directional
nature of the connection between the brain and the immune system, it is unlikely that clear and
definitive answers about the role of specific circulating cytokines in the production of specific
psychopathological symptoms can be established in this way.
Understanding potentially abnormal variants of a natural phenomenon clearly requires an in-
766
depth knowledge of the phenomenon in question. Therefore, a systematic study of normal sickness
behaviour in humans (examining the full spectrum, development and immunological mediators of
symptoms during acute infections) is needed to construct a sound knowledge-base enabling
identification and evaluation of inappropriate or excessive manifestations of this phenomenon.
Moreover, such knowledge is essential in light of the fact that the manipulation of cytokine systems
is rapidly developing into a new area of therapeutics (Dinarello, 1997), and which may have
unanticipated neurobehavioural consequences.
! -
I wish to thank Professor Gordon Parker for invaluable advice and encouragement throughout the preparation
of this paper.
REFERENCES
Ader, R., Felten, D. L. & Cohen, N. (1991). Psychoneuroimmunology,
2nd edn. Academic Press : San Diego, CA.
Banks, W. A., Oritz, L., Plotkin, S. R. & Kastin, A. J. (1991).
Human interleukin (IL) 1 alpha, murine IL-1 alpha, and murine
IL-1 beta are transported from blood to brain in the mouse by a
shared saturable mechanism. Journal of Pharmacology and Experimental Therapeutics 259, 988996.
Beck, G., Cooper, E. L., Hobicht, G. S. & Marchalonis, J. J. (1994).
Primordial immunity : foundations for the vertebrate immune
system. Annals of the New York Academy of Sciences 712,
206212.
Bennett, B. K., Hickie, I. B., Vollmer-Conna, U. S., Quigley, B.,
Brennan, C. M., Wakefield, D., Douglas, M. P., Hansen, G. R.,
Tahmindjis, A. J. & Lloyd, A. R. (1998). The relationship between
fatigue, psychological and immunological variables in acute
infectious illness. Australian and New Zealand Journal of Psychiatry
32, 180186.
Bluthe, R. M., Michaud, B., Poli, V., Bernay, F., Parnet, P. &
Dantzer, R. (1998). Interleukin-6 is active only in presence of other
proinflammatory cytokines to induce sickness behaviour. Neuroimmunomodulation 5, 7.
Capuron, L., Lamarque, D., Dantzer, R. & Goodall, G. (1999).
Attentional and mnemonic deficits associated with infectious
diseases in humans. Psychological Medicine 29, 291297.
Chao, C. C., Janoff, E. N., Hu, S., Thomas, K., Gallagher, M.,
Tsang, M. & Peterson, P. K. (1991). Altered cytokine release in
peripheral blood mononuclear cell cultures from patients with
chronic fatigue syndrome. Cytokine 3, 292298.
Charlton, B. G. (2000). The malaise theory of depression : major
depressive disorder is sickness behaviour and antidepressants are
analgesics. Medical Hypotheses 54, 126130.
Clatworthy, A. L. (1996). A simple system approach to neuralimmune communication. Comparative Biochemistry and Physiology
115A, 110.
Dantzer, R., Bluthe, R.-M., Aubert, A., Goodall, G., Bret-Dibat, JL., Kent, S., Goujon, E., Laye, S., Parnet, P. & Kelley, K. W.
(1996). Cytokine actions on behavior. In Cytokines in the Nervous
System (ed. N. J. Rothwell), pp. 117144. R. G. Landes Co.: New
York.
Dinarello, C. A. (1997). Proinflammatory and anti-inflammatory
cytokines as mediators in the pathogenesis of septic shock. Chest
112, 312S329S.
Dinarello, C. A., Renfer, L. & Wolff, S. M. (1977). Human leukocytic
pyrogen : purification and development of a radioummunoassay.
Proceedings of the National Academy of Science USA 74, 4624.
Dunn, A. J. (1995). Interactions between the nervous system and the
immune system : implications for psychopharmacology. In Psychopharmacology : The fourth Generation of Progress (ed. F. E. Bloom
and D. J. Kupfer), pp. 719733. Raven Press : New York.
Hart, B. L. (1988). Biological basis of the behavior of sick animals.
Neuroscience and Biobehavioral Reviews 12, 123137.
Herbert, I. & Cohen, S. (1993). Depression and immunity : a metaanalytic review. Psychological Bulletin 113, 472486.
767
NEUROSCIENCE
Grandmother
Each concepteach person or thing in our
everyday experiencemay have a set of
corresponding neurons assigned to it
By Rodrigo Quian Quiroga,
Itzhak Fried and Christof Koch
IN BRIEF
around the world, including Kochs group at the California Institute of Technology and Quian Quirogas laboratory at the University of Leicester in England. This technique furnishes an extraordinary opportunity to record directly from single neurons for
days at a time in awake patients and provides the ability to study
the iring of neurons during various tasksmonitoring the incessant chattering that occurs while patients look at images on a
laptop, recall memories or perform other tasks. That is how we
discovered the Jennifer Aniston neurons and unwittingly revived the debate ignited by Lettvins parable.
GRANDMOTHER CELLS REVISITED
are nerve cells such as the Jennifer Aniston neuron the longdebated grandmother cells? To answer that question, we have to
be more precise about what we mean by grandmother cells. One
extreme way of thinking about the grandmother cell hypothesis
is that only one neuron responds to one concept. But if we could
ind one single neuron that ired to Jennifer Aniston, it strongly
suggests that there must be morethe chance of inding the one
and only one among billions is minuscule. Moreover, if only a
single neuron would be responsible for a persons entire concept
of Jennifer Aniston, and it were damaged or destroyed by disease or accident, all trace of Jennifer Aniston would disappear
from memory, an extremely unlikely prospect.
A less extreme deinition of
grandmother cells postulates
that many more than a solitary
neuron respond to any one concept. This hypothesis is plausible but very diicult, if not impossible, to prove. We cannot
try every possible concept to
prove that the neuron ires only
to Jennifer Aniston. In fact, the
opposite is often the case: we
often ind neurons that respond
to more than one concept. Thus,
if a neuron ires only to one person during an experiment, we
cannot rule out that it could have also ired to some other stimuli that we did not happen to show.
For example, the day after inding the Jennifer Aniston neuron we repeated the experiment, now using many more pictures
related to her, and found that the neuron also ired to Lisa Kudrow, a costar in the TV series Friends that catapulted both to
fame. The neuron that responded to Luke Skywalker also ired to
Yoda, another Jedi from Star Wars; another neuron ired to two
basketball players; another to one of the authors (Quian Quiroga) of this article and other colleagues who interacted with the
patient at U.C.L.A., and so on. Even then, one can still argue that
these neurons are grandmother cells that are responding to
broader concepts, namely, the two blond women from Friends,
the Jedis from Star Wars, the basketball players, or the scientists
doing experiments with the patient. This expanded deinition
turns the discussion of whether these neurons should be considered grandmother cells into a semantic issue.
Let us leave semantics aside for now and focus instead on a
few critical aspects of these so-called Jennifer Aniston neurons.
First, we found that the responses of each cell are quite selec-
A single neuron
that responded to
Luke Skywalker
and his written
and spoken name
also ired to the
image of Yoda.
to understand the way a small number of cells become attached to a particular concept such as Jennifer Aniston, it helps
to know something about the brains complex processes for
capturing and storing images of the myriad of objects and people encountered in the world around us. The information taken
in by the eyes irst goesvia the optic nerve leaving the eyeballto the primary visual cortex at the back of the head. Neurons there ire in response to a tiny portion of the minute details that compose an image, as if each were lighting up like a
pixel in a digital image or as if they were the colored dots in a
pointillist painting by Georges Seurat.
One neuron does not suice to tell whether the detail is part
of a face, a cup of tea or the Eifel Tower. Each cell forms part of
an ensemble, a combination that generates a composite image
presented, say, as A Sunday Afternoon on the Island of La
Grande Jatte. If the picture changes slightly, some of the details
will vary, and the iring of the corresponding set of neurons will
change as well.
The brain needs to process sensory information so that it
captures more than a photographit must recognize an object
and integrate it with what is already known. From the primary
visual cortex, the neuronal activation triggered by an image
moves through a series of cortical regions toward more frontal
areas. Individual neurons in these higher visual areas respond
to entire faces or whole objects and not to local details. Just one
of these high-level neurons can tell us that the image is a face
and not the Eifel Tower. If we slightly vary the picture, move it
about or change the lighting illuminating it, it will change some
features, but these neurons do not care much about small differences in detail, and their iring will remain more or less the
samea property known as visual invariance.
Neurons in high-level visual areas send their information to
the medial temporal lobethe hippocampus and surrounding
cortexwhich is involved in memory functions and is where we
found the Jennifer Aniston neurons. The responses of neurons in
the hippocampus are much more speciic than in the higher visual
cortex. Each of these neurons responds to a particular person or,
more precisely, to the concept of that person: not only to the face
and other facets of appearance but also to closely associated attributes such as the persons name.
In our research, we have tried to explore how many individual neurons ire to represent a given concept. We had to ask
cause we use pictures of things that are very familiar to the patients in our researchwhich tend to trigger more responsesthis
number should be taken strictly as an upper bound; the number
of cells representing a concept may be 10 or 100 times as small,
perhaps close to Lettvins guess of 18,000 neurons per concept.
Contrary to this argument, one reason to think that the brain
does not code concepts sparsely, but rather distributes them
across very large neuronal populations, is that we may not have
enough neurons to represent all possible
concepts and their variations. Do we, for
CONCEPT CELLS
instance, have a big enough store of brain
cells to picture Grandma smiling, weaving, drinking tea or waiting at the bus
stop, as well as the Queen of England
Neuroscientists ardently debate two alternative theories of how memories are encoded
greeting the crowds, Luke Skywalker as a
in the brain. One theory contends that the representation of a single memorythe
child on Tatooine or ighting Darth Vader,
image of Luke Skywalker, for instanceis stored as bits and pieces distributed across
and so on?
millions or perhaps billions of neurons. The alternative view, which has gained more
To answer this question, we should
scientic credibility in recent years, holds that a relatively few neurons, numbering in the
irst consider that, in fact, a typical person
thousands or perhaps even less, constitute a sparse representation of an image. Each of
remembers no more than 10,000 conthose neurons will switch on to the image of Luke, whether from a distance or close-up.
cepts. And this is not a lot in comparison
Some but not all of the same group of neurons will also re to the related image of Yoda.
to the billion nerve cells that make up the
Similarly, a separate set of specic neurons activates when perceiving Jennifer Aniston.
medial temporal lobe. Furthermore, we
have good reason to think that concepts
may be coded and stored very efficiently in
Sparse
Distributed
a sparse way. Neurons in the medial temImage of
poral lobe just do not care about different
Luke Skywalker
instances of the same conceptthey do
not care if Luke is sitting or standing; they
only care if a stimulus has something to do
with Luke. They re to the concept itself
no matter how it is presented. Making the
concept more abstractring to all inMedial temporal lobe
stances of Lukereduces the information
that a neuron needs to encode and allows
Dierent image of
Luke Skywalker
it to become highly selective, responding
to Luke but not to Jennifer.
Simulation studies by Waydo underscore this view even further. Drawing on
a detailed model of visual processing,
Waydo built a software-based neural network that learned to recognize many unlabeled pictures of airplanes, cars, motorImage of
bikes and human faces. The software did
Yoda
so without supervision from a teacher. It
was not told this is a plane and that a
car. It had to gure this out by itself, using the assumption that the immense variety of possible images is in reality based
on a small number of people or things
and that each is represented by a small
Image of
subset of neurons, just as we found in the
Jennifer Aniston
medial temporal lobe. By incorporating
this sparse representation in the software
simulation, the network learned to distinguish the same persons or objects even
when shown in myriad different ways, a
nding similar to our observations from
human brain recordings.
To Code a Memory
Sparse but Not Grandmother-Cell Coding in the Medial Temporal Lobe. R. Quian
Quiroga, G. Kreiman, C. Koch and I. Fried in Trends in Cognitive Sciences, Vol. 12, No. 3, pages
8791; March 2008.
Percepts to Recollections: Insights from Single Neuron Recordings in the Human
Brain. Nanthia Suthana and Itzhak Fried in Trends in Cognitive Sciences, Vol. 16, No. 8, pages
427436; July 16, 2012.
Concept Cells: The Building Blocks of Declarative Memory Functions. Rodrigo Quian
Quiroga in Nature Reviews Neuroscience, Vol. 13, pages 587597; August 2012.
SCIENTIFIC AMERICAN ONLINE
Read an excerpt of Quian Quirogas book on memory at
ScientiicAmerican.com/feb2013/brain-cells
Neuron
Review
Experimental and Theoretical Approaches
to Conscious Processing
Stanislas Dehaene1,2,3,4,* and Jean-Pierre Changeux4,5,*
1INSERM,
Recent experimental studies and theoretical models have begun to address the challenge of establishing
a causal link between subjective conscious experience and measurable neuronal activity. The present
review focuses on the well-delimited issue of how an external or internal piece of information goes
beyond nonconscious processing and gains access to conscious processing, a transition characterized
by the existence of a reportable subjective experience. Converging neuroimaging and neurophysiological
data, acquired during minimal experimental contrasts between conscious and nonconscious processing,
point to objective neural measures of conscious access: late amplification of relevant sensory activity,
long-distance cortico-cortical synchronization at beta and gamma frequencies, and ignition of
a large-scale prefronto-parietal network. We compare these findings to current theoretical models of
conscious processing, including the Global Neuronal Workspace (GNW) model according to which
conscious access occurs when incoming information is made globally available to multiple brain systems
through a network of neurons with long-range axons densely distributed in prefrontal, parieto-temporal,
and cingulate cortices. The clinical implications of these results for general anesthesia, coma, vegetative
state, and schizophrenia are discussed.
Introduction
Understanding the neuronal architectures that give rise to
conscious experience is one of the central unsolved problems
of todays neuroscience, despite its major clinical implications
for general anesthesia, coma, vegetative-state, or minimally
conscious patients. The difficulties are numerous. Notably, the
term consciousness has multiple meanings, most of which
are difficult to precisely define in a manner amenable to experimentation. In this review, we outline recent advances made in
understanding the delimited issue of conscious access: how
does an external or internal piece of information gain access
to conscious processing, defined as a reportable subjective
experience?
We start with a brief overview of the relevant vocabulary and
theoretical concepts. We then examine the experimental studies
that have attempted to delineate the objective physiological
mechanisms of conscious sensory perception by contrasting it
with minimally different, yet nonconscious processing conditions, using a variety of methods: behavior, neuroimaging,
time-resolved electro- and magneto-encephalography, and
finally single-cell electrophysiology and pharmacology. We critically examine how the present evidence fits or argues against
existing models of conscious processing, including the Global
Neuronal Workspace (GNW) model. We end by examining
possible consequences of these advances for pathological brain
states, including general anesthesia, coma, and vegetative
states.
Neuron
Review
on a given trial, is not consciously perceived due to temporary
distraction or inattention.
Subliminal presentation is often achieved by masking,
a method whereby the subjective visibility of a stimulus is
reduced or eliminated by the presentation, in close spatial and
temporal contiguity, of other stimuli acting as masks (Breitmeyer, 2006). For instance, a word flashed for 33 ms is visible
when presented in isolation but becomes fully invisible when
preceded and followed by geometrical shapes. Masked stimuli
are frequently used to induce subliminal priming, the facilitation
of the processing of a visible target by the prior presentation of
an identical or related subliminal prime (for review, see Kouider
and Dehaene, 2007). Subliminal presentation can also be
achieved with threshold stimuli, where the contrast or energy
of a stimulus is progressively reduced until its presence is unnoticeable. Binocular rivalry is another common paradigm whereby
the image in one eye becomes subliminal by competition with
a rivaling image presented in the other eye. Participants typically
report temporal alternations in the image that is consciously
perceived. However, a variant of binocular rivalry, the continuous
flash suppression paradigm allows an image to be made permanently invisible by presenting continuously flashing shapes in the
other eye (Tsuchiya and Koch, 2005).
An equally large range of techniques allows for preconscious
presentation. In inattentional blindness, a potentially visible but
unexpected stimulus remains unreported when the participants
attention is focused on another task (Mack and Rock, 1998;
Simons and Ambinder, 2005). The attentional blink (AB) is
a short-term variant of this effect where a brief distraction by
a first stimulus T1 prevents the conscious perception of a second
stimulus T2 briefly presented within a few hundreds of milliseconds of T1 (Raymond et al., 1992). In the related psychological
refractory period (PRP) effect (Pashler, 1994; Welford, 1952),
T2 is unmasked and is therefore eventually perceived and processed, but only after a delay during which it remains nonconscious (Corallo et al., 2008; Marti et al., 2010). The distracting
event T1 can be a surprise event that merely captures attention
(Asplund et al., 2010). The minimum requirement, in order to
induce AB, appears to be that T1 is consciously perceived
(Nieuwenstein et al., 2009). Thus, PRP and AB are closely related
phenomena that point to a serial limit or bottleneck in conscous access (Jolicoeur, 1999; Marti et al., 2010; Wong, 2002)
and can be used to contrast the neural fate of two identical
stimuli, only one of which is consciously perceived (Sergent
et al., 2005).
Objective versus Subjective Criteria for Conscious
Access
How can an experimenter decide whether his experimental
subject was or was not conscious of a stimulus? According to
a long psychophysical tradition, grounded in signal-detection
theory, a stimulus should be accepted as nonconscious only if
subjects are unable to perform above chance on some direct
task of stimulus detection or classification. This strict objective
criterion raises problems, however (Persaud et al., 2007;
Schurger and Sher, 2008). First, it tends to overestimate
conscious perception: there are many conditions in which
subjects perform better than chance, yet still deny perceiving
the stimulus. Second, performance can be at chance level for
some tasks, but not others, raising the issue of which tasks count
as evidence of conscious perception or merely of subliminal processing. Third, the approach requires accepting the null hypothesis of chance-level performance, yet performance never really
falls down to zero, and whether it is significant or not often
depends on arbitrary choices such as the number of trials dedicated to its measurement.
For these reasons, recent alternative approaches emphasize
either pure subjective reports, such as ratings of stimulus visibility
(Sergent and Dehaene, 2004), or second-order commentaries
such as postdecision wagering (e.g., would you bet that your
response was correct?; Persaud et al., 2007). The wagering
method and related confidence judgements provide a high motivation to respond truthfully and in an unbiased manner (Schurger
and Sher, 2008). Furthemore, they can be adapted to nonhuman
subjects (Kiani and Shadlen, 2009; Terrace and Son, 2009).
However, they can sometimes exceed chance level even when
subjects deny seeing the stimulus (Kanai et al., 2010).
Conversely, subjective report is arguably the primary data of
interest in consciousness research. Furthermore, reports of stimulus visibility can be finely quantified, leading to the discovery that
conscious perception can be all-or-none in some paradigms
(Del Cul et al., 2007; Del Cul et al., 2006; Sergent and Dehaene,
2004). Subjective reports also present the advantage of assessing conscious access immediately and on every trial, thus permitting postexperiment sorting of conscious versus nonconscious
trials with identical stimuli (e.g., Del Cul et al., 2007; Lamy et al.,
2009; Pins and Ffytche, 2003; Sergent et al., 2005; Wyart and
Tallon-Baudry, 2008).
Although the debate about optimal measures of conscious
perception continues, it is important to acknowledge that objective assessments, wagering indices and subjective reports are
generally in excellent agreement (Del Cul et al., 2006; Del Cul
et al., 2009; Persaud et al., 2007). For instance, in visual masking,
the conscious perception thresholds derived from objective and
subjective data are essentially identical across subjects (r2 =
0.96, slope z 1) (Del Cul et al., 2006). Those data suggest that
conscious access causes a major change in the global availability of information, whether queried by objective or by subjective means, whose mechanism is the focus of the present review.
Selective Attention versus Conscious Access
Conscious access must be distinguished from the related
concept of attention. William James (1890) provided a wellknown definition of attention as the taking possession by the
mind, in clear and vivid form, of one out of what seem several
simultaneously possible objects or trains of thought. The
problem with this definition is that it conflates two processes
that are now clearly separated in cognitive psychology and cognitive neuroscience (e.g., Huang, 2010; Posner and Dehaene,
1994): selection and access. Selection, also called selective
attention, refers to the separation of relevant versus irrelevant
information, isolation of an object or spatial location, based on
its saliency or relevance to current goals, and amplification of
its sensory attributes. Access refers to its conscious taking
possession of the mindthe subject of the present review.
Empirical evidence indicates that selection can occur without
conscious processing (Koch and Tsuchiya, 2007). For instance,
selective spatial attention can be attracted to the location of
Neuron
Review
a target stimulus that remains invisible (Bressan and Pizzighello,
2008; McCormick, 1997; Robitaille and Jolicoeur, 2006;
Woodman and Luck, 2003). Selective attention can also amplify
the processing of stimuli that remain nonconscious (Kentridge
et al., 2008; Kiefer and Brendel, 2006; Naccache et al., 2002).
Finally, in simple displays with a single target, conscious access
can occur independently of selection (Wyart and Tallon-Baudry,
2008). In cluttered displays, however, selection appears to be
a prerequisite of conscious access: when faced with several
competing stimuli, we need attentional selection in order to
gain conscious access to just one of them (Dehaene and Naccache, 2001; Mack and Rock, 1998). These findings indicate that
selective attention and conscious access are related but dissociable concepts that should be carefully separated, attention
frequently serving as a gateway that regulates which information reaches conscious processing.
II. Experimental Studies of the Brain Mechanisms
of Conscious Access
With this vocabulary at hand, we turn to empirical studies of
conscious access. The simplest experiments consist in presenting a brief sensory stimulus that is sometimes consciously accessible, sometimes not, and using behavior, neuroimaging, and
neurophysiological recording to monitor the depth of its processing and how it differs as a function of conscious reportability.
Experiments Contrasting Visible and Invisible Stimuli
Behavioral evidence. A visual stimulus that is masked and
remains invisible can nevertheless affect behavior and brain
activity at multiple levels (for review, see Kouider and Dehaene,
2007; Van den Bussche et al., 2009b). Subliminal priming has
now been convincingly demonstrated at visual, semantic, and
even motor levels. For instance, when a visible target image is
preceded by a subliminal presentation of the same image, simple
decisions, such as judging whether it refers to an object or
animal, are accelerated compared to when the image is not
repeated. Crucially, this repetition effect resists major changes
in the physical stimulus, such as presenting the same word in
upper case versus lower case (Dehaene et al., 2001) or presenting the same face in two different orientations (Kouider et al.,
2009), suggesting that invariant visual recognition can be
achieved without awareness. At the semantic level, subliminal
extraction of the meaning of words has now been demonstrated
for a variety of word categories (e.g., Gaillard et al., 2006; Naccache and Dehaene, 2001; Van den Bussche et al., 2009a). At
even more advanced levels, a subliminal stimulus can bias motor
responses (Dehaene et al., 1998b; Leuthold and Kopp, 1998).
Subliminal monetary incentives enhance subjects motivation
in a demanding force task, indicating that motivation is modulated by nonconscious signals (Pessiglione et al., 2007). So is
task setting: masked shapes can act as cues for task switching
and lead to detectable changes in task set (Lau and Passingham,
2007). Even inhibitory control can be partially launched nonconsciously, as when a nonconscious stop signal slows down or
interrupts motor responses (van Gaal et al., 2008) (see Figure 1).
The above list suggests that entire chains of specialized
processors can be subject to nonconscious influences. Nevertheless, three potential limits to subliminal processing have
been identified (Dehaene and Naccache, 2001). First, subliminal
Neuron
Review
A
Visible word
Invisible word
0.3
Visible words
percent signal change
0.2
0.1
Masked
words
0.0
-5
10
15
time (s)
-0.1
Detected sound
Non-detected sound
Heard
Not heard
Inhibitory control
by visible cue
Inhibitory control
by invisible cue
Visible go/nogo signals
Neuron
Review
recordings and ERP-fMRI correlation to involve a highly distributed set of nearly simultaneous active areas including hippocampus and temporal, parietal, and frontal association cortices
(Halgren et al., 1998; Mantini et al., 2009). The P3b has been
reproducibly observed as strongly correlated with subjective
reports, both when varying stimulus parameters (e.g., Del Cul
et al., 2007) and when comparing identical trials with or without
conscious perception (e.g., Babiloni et al., 2006; Del Cul et al.,
2007; Fernandez-Duque et al., 2003; Koivisto et al., 2008;
Lamy et al., 2009; Niedeggen et al., 2001; Pins and Ffytche,
2003; Sergent et al., 2005) (however, this effect may disappear
when the subject already has a conscious working memory
representation of the target: Melloni et al., 2011). The effect is
not easily imputable to increased postperceptual processing or
other task confounds, as many studies equated attention and
response requirements on conscious and nonconscious trials
(e.g., Del Cul et al., 2007; Gaillard et al., 2009; Lamy et al.,
2009; Sergent et al., 2005). For instance, Lamy et al. (2009)
Neuron
Review
A Late event-related
B Late gamma-band
C Beta phase
D Long-distance
potentials
power
synchrony
causality
1.4
100
Freq. (Hz)
invisible
80
0.7
60
0
40
-0.7
20
-1.4
100
1.4
Freq. (Hz)
visible
80
0.7
60
0
40
-0.7
20
0.5
-1.4
80
300
200
visible
100
100
400
Freq. (Hz)
Voltage Power
0.01
100
300
invisible
500
700 ms
0.03
0.02
60
visible
0.003
0.01
40
-0.003
20
0
0.04
-0.01
0
-0.01
-500
invisible
0
500
ms
Neuron
Review
Figure 4. Human Single-Cell Recordings
during Conscious Access
Single cells were recorded from the human medial
temporal lobe and hippocampus during presentation of masked pictures, with a variable targetmask delay (Quiroga et al., 2008). The example at
left shows a single cell that fired specifically to
pictures of the World Trade Center, and did so only
on trials when the patient recognized the picture
(dark blue raster plots), not on trials when recognition failed (red raster plots). Graphs at right show
the average firing rate across all neurons. Although
a small transient firing could be seen on unrecognized trials, conscious perception was characterized by a massive and durable amplification of
activity (for complementary results using electrocorticography (ECoG) in human occipito-temporal
areas, see also Fisch et al., 2009).
in the same time window (300500 ms) and suggested that they
might constitute different measures of the same state of distributed ignition of a large cortical network including prefrontal
cortex. Indeed, seen stimuli had a global impact on late evoked
activity virtually anywhere in the cortex: 68.8% of electrode sites,
although selected for clinical purposes, were modulated by the
presence of conscious words (as opposed to 24.4% of sites
for nonconscious words).
Neuronal recordings. A pioneering research program was conducted by Logothetis and collaborators using monkeys trained
to report their perception during binocular rivalry (Leopold and
Logothetis, 1996; Sheinberg and Logothetis, 1997; Wilke et al.,
2006). By recording from V1, V2, V4, MT, MST, IT, and STS
neurons and presenting two rivaling images, only one of which
led to high neural firing, they identified a fraction of cells whose
firing rate increased when their preferred stimuli was perceived,
thus participating in a conscious neuronal assembly. The proportion of such cells increased from about 20% in V1/V2 to 40% in
V4, MT, or MST to as high as 90% in IT and STS. This finding
supports the hypothesis that subjective perception is associated
with distributed cell assemblies whose neurons are denser in
higher associative cortices than in primary and secondary visual
cortices. Surprisingly, fMRI signals correlated quite strongly with
conscious perception during rivalry in area V1 (Haynes and Rees,
2005; Polonsky et al., 2000) and even in the lateral geniculate
nucleus of the thalamus (Haynes et al., 2005a; Wunderlich
et al., 2005). The discrepancy between fMRI and single-cell
recordings was addressed in a recent electrophysiological study
(Maier et al., 2008; see also Wilke et al., 2006): within area V1 of
the same monkeys, fMRI signals and low-frequency (530 Hz)
local field potentials (LFPs) correlated with subjective visibility
while high-frequency (3090 Hz) LFPs and single-cell firing rate
did not. One interpretation of this finding is that V1 neurons
receive additional top-down synaptic signals during conscious
perception compared to nonconscious perception, although
these signals need not be translated into changes in average
firing rate (Maier et al., 2008).
Neuron
Review
perception (Rodriguez et al., 1999; Varela et al., 2001). Within
a single area such as V4, the degree to which single neurons
synchronize with the ongoing fluctuations in local-field potential
is a predictor of stimulus detection (Womelsdorf et al., 2006).
Across distant areas such as FEF and V4 (Gregoriou et al.,
2009) or PFC and LIP (Buschman and Miller, 2007), synchrony
is enhanced when the stimulus in the receptive field is attended
and is thus presumably accessed consciously. Consistent with
human MEG and intracranial studies (e.g., Gaillard et al., 2009;
Gross et al., 2004), synchronization involves both gamma and
beta bands, the latter being particularly enhanced during topdown attention (Buschman and Miller, 2007). During the late
phase of attention-driven activity, causal relations between
distant areas are durably enhanced in both directions, but
more strongly so in the bottom-up direction from V4 to FEF (Gregoriou et al., 2009), again similar to human findings (Gaillard
et al., 2009) and compatible with the idea that sensory information needs to be propagated anteriorily, particularly to PFC,
before becoming consciously reportable.
Experiments with Perceived and Unperceived Stimuli
outside the Visual Modality
Although vision remains the dominant paradigm, remarkably
similar signatures of conscious access have been obtained in
other sensory or motor modalities (see Figure 1).
In the tactile modality, threshold-level stimuli were studied
both in humans with fMRI and magneto-encephalography
(Boly et al., 2007; Jones et al., 2007) and in awake monkeys
with single-cell electrophysiology (de Lafuente and Romo,
2005, 2006). In the monkey, the early activity of neurons in the
primary somatosensory area S1 was identical on detected and
undetected trials, but within 180 ms the activation expanded
into parietal and medial frontal cortices (MFC) where it showed
a large difference predictive of behavioral reports (high activation
on detected trials and low activity on undetected trials, even for
constant stimuli). In humans, a similar two-phase pattern was
identified within area S1 (Jones et al., 2007). According to the
authors, modeling of these S1 potentials required the postulation
of a late top-down input from unknown distant areas to supragranular and granular layers, specific to detected stimuli. Thus,
as in the visual modality (Del Cul et al., 2007; Supe`r et al.,
2001), tactile cortices may be mobilized into a conscious
assembly only during a later phase of top-down amplification,
synchronous to the activation of higher association cortices.
In the auditory modality, similarly, stimuli that are not
consciously detected still trigger considerable sensory processing, including 40 Hz steady-state responses (Gutschalk et al.,
2008) and mismatch negativities (MMN), i.e., electrophysiological
responses that arise primarily from the temporal lobe in response
to rare, deviant, or otherwise unpredictable auditory stimuli (Allen
et al., 2000; Bekinschtein et al., 2009a; Diekhof et al., 2009; Naatanen, 1990). Once again, conscious and nonconscious stimuli
differ in a late (>200 ms) and global P3 wave arising from bilateral
prefronto-parietal generators, with joint enhancement of
temporal auditory cortices (Bekinschtein et al., 2009a; Diekhof
et al., 2009). These localizations are confirmed by an fMRI study
that contrasted detected versus undetected near-threshold noise
bursts (Sadaghiani et al., 2009) (Figure 1). Similarly, an fMRI study
of speech listening at different levels of sedation showed partially
preserved responses in temporal cortices but the total disappearance of activation in the left inferior frontal gyrus during deep
sedation (Davis et al., 2007). A study by Hasson et al. (2007)
further suggests that the content of what we consciously hear
does not depend on early modality-specific responses in auditory
cortex, but rather on late fronto-parietal cross-modal computations. Using the McGurk illusion (perception of a syllable ta
when simultaneously hearing pa and seeing a face saying
ka), they dissociated the objective auditory and visual stimuli
from the subjective percept. Using fMRI repetition suppression,
they then showed that early auditory cortices coded solely for
the objective auditory stimulus, while the perceived subjective
conscious content was reflected in the activation of the left posterior inferior frontal gyrus and anterior inferior parietal lobule. In this
instance, at least, PFC activation could not be attributed to
a generic process of attention, detection, or memory but demonstrably encoded the specific syllable perceived.
Turning to the action domain, several studies have demonstrated that the awareness of ones action, surprisingly, is not
associated with primary or premotor cortices but arises from
a higher-level representation of intentions and their expected
sensory consequences; this representation involves prefrontal
and parietal cortices, notably the angular gyrus (AG) (Desmurget
et al., 2009; Farrer et al., 2008). Using direct cortical stimulation,
Desmurget et al. (2009) observed a double dissociation: premotor stimulation often led to overt movements that the subject was
not aware of performing, while angular gyrus stimulation led to
a subjective perception of movement intention and performance
even in the absence of any detectable muscle activation. In
normal subjects, disrupted sensori-motor feedback has also
been used to define a minimal contrast between subliminal
versus conscious gestures. For instance, when a temporal delay
or a spatial bias was introduced in the visual feedback provided
to participants about their own hand movements, they continuously adjusted their behavior, but these motor adjustments
were only perceived consciously when the disruption exceeded
a certain threshold (Farrer et al., 2008; Slachevsky et al., 2001).
fMRI revealed that this nonlinearity related to a bilateral distributed network involving AG and PFC cortices (Farrer et al., 2008).
Perhaps the clearest evidence for a two-stage process in
action awareness comes from studies of error awareness (Nieuwenhuis et al., 2001). In an antisaccade paradigm, participants
were instructed to move their eyes in the direction opposite to
a visual target. This instruction generated frequent errors, where
the eyes first moved toward the stimulus and then away from it.
Many of these erroneous eye movements remained undetected.
Remarkably, immediately after such undetected errors, a strong
and early (80 ms) ERP component called the error-related
negativity arose from midline frontal cortices (anterior cingulate
or pre-SMA). Only when the error was consciously detected
was this early waveform amplified and followed by a massive
P3-like waveform, which fMRI associated with the expansion
of activation into a broader network including left inferior
frontal/anterior insula activity (Klein et al., 2007).
Convergence with Studies of Inattention and Dual Tasks
The experiments reviewed so far considered primarily subliminal
paradigms where access to conscious reportability was modulated by reducing the incoming sensory information. However,
Neuron
Review
Figure 5. Recruitment of Global FrontoParietal Networks in Effortful Serial Tasks
(A) Simulations of the original global neuronal
workspace proposal before, during, and after
learning of an effortful Stroop-like task (adapted
from Dehaene et al., 1998a). The figure shows the
activity of various processor and workspace units
as a function of time. Workspace units show
strong activation (a) during the search for a taskappropriate configuration of workspace units; (b)
during the effortful execution of a novel task (but
not after its routinization); and (c) after errors, or
whenever higher control is needed.
(BD) Example of corresponding global frontoparietal activations as seen with fMRI. (B) Strong
activation of a distributed network involving PFC
during effortful search for the solution of
a master-mind type problem, with a sudden
collapse as soon as a routine solution is found
(adapted from Landmann et al., 2007). (C) Activation of inferior PFC during dual-task performance which diminishes with training (adapted
from Dux et al., 2009). (D) Activation of a distribution parieto-prefrontal-cingulate network on error
and conflict trials (adapted from the meta-analysis
by Klein et al., 2007).
similar findings arise from preconscious paradigms where withdrawal of attentional selection is used to modulate conscious
access (Dehaene et al., 2006), resulting in either failed (attentional
blink, AB) or delayed (psychological refractory period or PRP)
conscious access. In such states, initial visual processing, indexed by P1 and N1 waves, can be largely or even entirely unaffected (Sergent et al., 2005; Sigman and Dehaene, 2008; Vogel
et al., 1998). However, only perceived stimuli exhibit an amplification of activation in task-related sensory areas (e.g., parahippocampal place area for pictures of places) as well as the unique
emergence of lateral and midline prefrontal and parietal areas
(see also Asplund et al., 2010; Marois et al., 2004; Slagter et al.,
2010; Williams et al., 2008). Temporally resolved fMRI studies
indicate that, during the dual-task bottleneck, PFC activity
evoked by the second task is delayed (Dux et al., 2006; Sigman
and Dehaene, 2008). With electrophysiology, the P3b waveform
again appears as a major correlate of conscious processing
that is both delayed during the PRP (Dellacqua et al., 2005;
Sigman and Dehaene, 2008) and absent during AB (Kranczioch
et al., 2007; Sergent et al., 2005). Seen versus blinked trials are
also distinguished by another marker, the synchronization of
distant frontoparietal areas in the beta band (Gross et al., 2004).
William James (1890) noted how conscious attention and
effort are required for the controlled execution of novel nonroutine sequential tasks but is no longer needed or even detrimental
once routine sets in. Thus, the comparison of effortful versus
automatic tasks provides another contrast that, although not
quite as minimal as the previous ones, should at least provide
signatures of conscious-level processing consistent with other
paradigms. Indeed, a broad network including inferior and
dorsolateral prefrontal, anterior cingulated, and lateral parietal
Neuron
Review
Evaluative
Systems
(VALUE)
Attentional
Systems
(FOCUSING)
Long-Term
Memory
(PAST)
Global
Workspace
Perceptual
systems
(PRESENT)
Motor
systems
(FUTURE)
frontal
sensory
II
III
II
III
In conclusion, human neuroimaging methods and electrophysiological recordings during conscious access, under
a broad variety of paradigms, consistently reveal a late amplification of relevant sensory activity, long-distance cortico-cortical
synchronization at beta and gamma frequencies, and ignition
of a large-scale prefronto-parietal network.
III. Theoretical Modeling of Conscious Access
The above experiments provide a convergent database of
observations. In the present section, we examine which theoretical principles may account for these findings. We briefly survey
the major theories of conscious processing, with the goal to try to
isolate a core set of principles that are common to most theories
and begin to make sense of existing observations. We then
describe in more detail a specific theory, the Global Neuronal
Workspace (GNW), whose simulations coarsely capture the contrasting physiological states underlying nonconscious versus
conscious processing.
Neuron
Review
semantic, and motor processors can unfold without our awareness, as reviewed in the previous section, but conscious perception seems needed for the flexible control of their execution, such
as their onset, termination, inhibition, repetition, or serial chaining.
A serial processing system. Descartes (1648) first observed
that ideas impede each other. Broadbent (1958) theorized
conscious perception as involving access to a limited-capacity
channel where processing is serial, one object at a time. The
attentional blink and psychological refractory period effects
indeed confirm that conscious processing of a first stimulus
renders us temporarily unable to consciously perceive other
stimuli presently shortly thereafter. Several psychological
models now incorporate the idea that initial perceptual processing is parallel and nonconscious and that conscious access is
serial and occurs at the level of a later central bottleneck (Pashler, 1994) or second processing stage of working memory
consolidation (Chun and Potter, 1995).
A coherent assembly formed by re-entrant or top-down loops.
In the context of the maintenance of invariant representations of
the body/world through reafference (von Holst and Mittelstaedt,
1950), Edelman (1987) proposed re-entry as an essential component of the creation of a unified percept: the bidirectional
exchange of signals across parallel cortical maps coding for
different aspects of the same object. More recently, the dynamic
core hypothesis (Tononi and Edelman, 1998) proposes that information encoded by a group of neurons is conscious only if it
achieves not only differentiation (i.e., the isolation of one specific
content out of a vast repertoire of potential internal representations) but also integration (i.e., the formation of a single,
coherent, and unified representation, where the whole carries
more information than each part alone). A notable feature of
the dynamic core hypothesis is the proposal of a quantitative
mathematical measure of information integration called F, high
values of which are achieved only through a hierarchical recurrent connectivity and would be necessary and sufficient to
sustain conscious experience: consciousness is integrated
information (Tononi, 2008). This measure has been shown to
be operative for some conscious/nonconscious distinctions
such as anesthesia (e.g., Lee et al., 2009b; Schrouff et al.,
2011), but it is computationally complicated and, as a result,
has not yet been broadly applied to most of the minimal empirical
contrasts reviewed above.
In related proposals, Crick and Koch (1995, 2003, 2005) suggested that conscious access involves forming a stable global
neural coalition. They initially introduced reverberating gammaband oscillations around 40 Hz as a crucial component, then
proposed an essential role of connections to prefrontal cortex.
Lamme and colleagues (Lamme and Roelfsema, 2000; Supe`r
et al., 2001) produced data strongly suggesting that feedforward
or bottom-up processing alone is not sufficient for conscious
access and that top-down or feedback signals forming recurrent
loops are essential to conscious visual perception. Llinas and
colleagues (Llinas et al., 1998; Llinas and Pare, 1991) have also
argued that consciousness is fundamentally a thalamocortical
closed-loop property in which the ability of cells to be intrinsically
active plays a central role.
A global workspace for information sharing. The theater metaphor (Taine, 1870) compares consciousness to a narrow scene
that allows a single actor to diffuse his message. This view has
been criticized because, at face value, it implies a conscious
homunculus watching the scene, thus leading to infinite regress
(Dennett, 1991). However, capitalizing on the earlier concept of
a blackboard system in artificial intelligence (a common data
structure shared and updated by many specialized modules),
Baars (1989) proposed a homunculus-free psychological model
where the current conscious content is represented within
a distinct mental space called global workspace, with the
capacity to broadcast this information to a set of other processors (Figure 6). Anatomically, Baars speculated that the neural
bases of his global workspace might comprise the ascending
reticular formation of the brain stem and midbrain, the outer shell
of the thalamus and the set of neurons projecting upward
diffusely from the thalamus to the cerebral cortex.
We introduced the Global Neuronal Workspace (GNW) model
as an alternative cortical mechanism capable of integrating the
supervision, limited-capacity, and re-entry properties (Changeux
and Dehaene, 2008; Dehaene and Changeux, 2005; Dehaene
et al., 1998a, 2003b, 2006; Dehaene and Naccache, 2001).
Our proposal is that a subset of cortical pyramidal cells with
long-range excitatory axons, particularly dense in prefrontal,
cingulate, and parietal regions, together with the relevant thalamocortical loops, form a horizontal neuronal workspace interconnecting the multiple specialized, automatic, and nonconscious processors (Figure 6). A conscious content is assumed
to be encoded by the sustained activity of a fraction of GNW
neurons, the rest being inhibited. Through their numerous reciprocal connections, GNW neurons amplify and maintain a specific
neural representation. The long-distance axons of GNW neurons
then broadcast it to many other processors brain-wide. Global
broadcasting allows information to be more efficiently processed
(because it is no longer confined to a subset of nonconscious
circuits but can be flexibly shared by many cortical processors)
and to be verbally reported (because these processors include
those involved in formulating verbal messages). Nonconscious
stimuli can be quickly and efficiently processed along automatized or preinstructed processing routes before quickly decaying
within a few seconds. By contrast, conscious stimuli would be
distinguished by their lack of encapsulation in specialized
processes and their flexible circulation to various processes of
verbal report, evaluation, memory, planning, and intentional
action, many seconds after their disappearance (Baars, 1989;
Dehaene and Naccache, 2001). Dehaene and Naccache (2001)
postulate that this global availability of information (.) is what
we subjectively experience as a conscious state.
Explicit Simulations of Conscious Ignition
The GNW has been implemented as explicit computer simulations of neural networks (Dehaene and Changeux, 2005; Dehaene et al., 1998a, 2003b; see also Zylberberg et al., 2009).
These simulations incorporate spiking neurons and synapses
with detailed membrane, ion channel, and receptor properties,
organized into distinct cortical supragranular, granular, infragranular, and thalamic sectors with reasonable connectivity and
temporal delays. Although the full GNW architecture was not
simulated, four areas were selected and hierarchically interconnected (Figure 7). Bottom-up feed-forward connections linked
each area to the next, while long-distance top-down
Neuron
Review
A
Feed-forward propagation
(subliminal processing)
Feedforward connections
(AMPA)
Feedback connections
(NMDA)
Area D
Area C
Area B1
Area A1
T1
Supra
granular
T2
layer IV
Infra
granular
Thalamocortical
column
Thalamus
neuromodulation
Simulated areas
Time
Neuron
Review
associative cortices, with simultaneous increases in highfrequency power and synchrony (e.g., de Lafuente and Romo,
2006; Del Cul et al., 2007; Gaillard et al., 2009).
In GNW simulations, ignition manifests itself, at the cortical
level, as a depolarization of layer II/III apical dendrites of pyramidal dendrites in a subset of activated GNW neurons defining
the conscious contents, the rest being inhibited. In a geometrically accurate model of the pyramidal cell, the summed postsynaptic potentials evoked by long-distance signaling among these
distributed sets of active cells would create slow intracellular
currents traveling from the apical dendrites toward the cells
soma, summing up on the cortical surface as negative slow
cortical potentials (SCPs) over regions coding for the conscious
stimulus (see He and Raichle, 2009). Simultaneously, many other
GNW neurons are strongly suppressed by lateral inhibition via
GABAergic interneurons and define what the current conscious
content is not. As already noted by Rockstroh et al. (1992, p.
175), assuming that many more neurons are inhibited than activated, The surface positivity corresponding to these inhibited
networks would then dominate over the relatively smaller spots
of negativity caused by the reverberating excitation. Thus, the
model can explain why, during conscious access, the resulting
event-related potential is dominated by a positive waveform,
the P3b. This view also predicts that scalp negativities should
appear specifically over areas dense in neurons coding for the
current conscious content. Indeed, in a spatial working memory
task, all stimuli evoke a broad P3b, but when subtracting ERPs
ipsilateral and controlateral to the side of the memorized items,
negative potentials appeared over parietal cortex contralateral
to the memorized locations (Vogel and Machizawa, 2004).
Further GNW simulations showed that ignition could fail to be
triggered under specific conditions, thus leading to simulated
nonconscious states. For very brief or low-amplitude stimuli,
a feedforward wave was seen in the initial thalamic and cortical
stages of the simulation, but it died out without triggering the
late global activation, because it was not able to gather sufficient
self-sustaining reverberant activation (Dehaene and Changeux,
2005). Even at higher stimulus amplitudes, the second global
phase could also be disrupted if another incoming stimulus
had been simultaneously accessed (Dehaene et al., 2003b).
Such a disruption occurs because during ignition, the GNW is
mobilized as a whole, some GNW neurons being active while
the rest is actively inhibited, thus preventing multiple simultaneous ignitions. A strict seriality of conscious access and processing is therefore predicted and has been simulated (Dehaene
and Changeux, 2005; Dehaene et al., 2003b; Zylberberg et al.,
2010). Overall, these simulations capture the two main types of
experimental conditions known to lead to nonconscious processing: subliminal states due to stimulus degradation (e.g.,
masking), and preconscious states due to distraction by a simultaneous task (e.g., attentional blink).
The transition to the ignited state can be described, in theoretical physics terms, as a stochastic phase transitiona sudden
change in neuronal dynamics whose occurrence depends in
part on stimulus characteristics and in part on spontaneous fluctuations in activity (Dehaene and Changeux, 2005; Dehaene
et al., 2003b). In GNW simulations, prestimulus fluctuations in
neural discharges only have a small effect on the early sensory
Neuron
Review
In summary, we propose that a core set of theoretical concepts
lie at the confluence of the diverse theories that have been
proposed to account for conscious access: high-level supervision; serial processing; coherent stability through re-entrant
loops; and global information availability. Furthermore, once implemented in the specific neuronal architecture of the GNW
model, these concepts begin to provide a schematic account of
the neurophysiological signatures that, empirically, distinguish
conscious access from nonconscious processing. In particular,
simulations of the GNW architecture can explain the close similarity of the brain activations seen during (1) conscious access to
a single external stimulus; (2) effortful serial processing; and (3)
spontaneous fluctuations in the absence of any stimulus or task.
IV. Present Experimental and Theoretical Challenges
The existing empirical data on conscious access still present
many challenges for theorizing. Indeed, the above theoretical
synthesis may still be refuted if some of its key neural components were found to be implausible or altogether absent in
primate cerebral architecture, or if its predicted patterns of
activity (the late ignition) were found to be unnecessary, artifactual, noncoding, or noncausally related to conscious states.
We consider each of these potential challenges in turn.
Connectivity and Architecture of Long-Distance Cortical
Networks
Pyramidal neurons with long-distance axons. The main anatomical premise of the GNW model is that it consists of a distributed
set of cortical neurons characterized by their ability to receive
from and send back to homologous neurons in other cortical
areas horizontal projections through long-range excitatory axons
mostly originating from the pyramidal cells of layers II and III
(Dehaene et al., 1998a) and more densely distributed in
prefrontal and inferior parietal cortices. Do these units actually
exist? The special morphology of the pyramidal cells from
the cerebral cortex was already noted by Cajal (18991904),
who mentioned their long axons with multiple collaterals and
their very numerous and complex dendrites. Von Economo
(1929) further noted that these large pyramidal cells in layers III
and V are especially abundant in areas spread over the anterior
two-thirds of the frontal lobe, (.) the superior parietal lobule
and the cingulate cortex, among other cortical areas. Recent
investigations have confirmed that long-distance corticocortical and callosal fibers primarily (though not exclusively) arise
from layer II-III pyramids. Furthermore, quantitative analyses of
the dendritic field morphology of layer III pyramidal neurons revealed a continuous increase of complexity of the basal
dendrites from the occipital up to the prefrontal cortex within
a given species (DeFelipe and Farinas, 1992; Elston and Rosa,
1997, 1998) and from lower species (owl monkey, marmoset)
up to humans (Elston, 2003). Layer IV PFC pyramidal neurons
have as many as 16 times more spines in PFC than in V1 and,
as a result, the highly spinous cells in prefrontal areas may integrate many more inputs than cells in areas such as V1, TE, and
7a (Elston, 2000). These observations confirm that PFC cells
exhibit the morphological adaptations needed for massive
long-distance communication, information integration, and
broadcasting postulated in the GNW model and suggest that
this architecture is particularly developed in the human species.
Neuron
Review
A
Spatial neglect
Multiple sclerosis
Inferior longitudinal
fasciculus
30
70
90 ms
Neglect
50
30
50
70
90 ms
Neuron
Review
phase, mimicking experimental observations. It can be criticized
as both receptor types are known to be present in variable
proportions at glutamatergic synapses (for pioneering data on
human receptor distribution, see Amunts et al., 2010). However,
in agreement with the model, physiological recordings suggest
that NMDA antagonists do not interfere with early bottom-up
sensory activity, but only affect later integrative events such as
the mismatch negativity in auditory cortex (Javitt et al., 1996).
Thus, although GNW simulations adopted a highly simplified
anatomical assumption of radically distinct distributions of
NMDA and AMPA, which may have to be qualified in more realistic models, the notion that NMDA receptors contribute
primarily to late, slow, and top-down integrative processes is
plausible (for a related argument, see Wong and Wang, 2006).
Is Conscious Perception Slow and Late?
A strong statement of the proposed theoretical synthesis is that
early bottom-up sensory events, prior to global ignition (<200
300 ms), contribute solely to nonconscious percept construction
and do not systematically distinguish consciously seen from
unseen stimuli. In apparent contradiction with this view, certain
experiments, using both visual (Pins and Ffytche, 2003) or tactile
stimuli (Palva et al., 2005), have observed that the early incoming
wave of sensory-evoked activity (e.g., P1 component) is already
enhanced on conscious compared to nonconscious trials.
Lamme and collaborators (Fahrenfort et al., 2007) found amplification in visual cortex, just posterior to the P1 wave (110
140 ms). More frequently, at around 200300 ms, surrounding
the P2 ERP component, more negative voltages are reported
over posterior cortices on visible compared to invisible trials
(Del Cul et al., 2007; Fahrenfort et al., 2007; Koivisto et al.,
2008, 2009; Railo and Koivisto, 2009; Sergent et al., 2005). Koivisto and collaborators have called this event the visual awareness negativity (VAN).
Several arguments, however, mitigate the possibility that
these early or midlatency differences already reflect conscious
perception. First, they may not be necessary and sufficient, as
they are absent from several experiments (e.g., Lamy et al.,
2009; van Aalderen-Smeets et al., 2006) (although one cannot
exclude that they failed to be detected). Second, and most
crucially, their profile of variation with stimulus variables such
as target-mask delay does not always track the variations in
subjects conscious reports (Del Cul et al., 2007; van AalderenSmeets et al., 2006). Third, they typically consist only in small
modulations that ride on top of early sensory activations that
are still strongly present on nonconscious trials (Del Cul et al.,
2007; Fahrenfort et al., 2007; Sergent et al., 2005). Fourth, in
this respect they resemble the small electrophysiological modulations that have been found to partially predict later perception
even prior to the stimulus (e.g., Boly et al., 2007; Palva et al.,
2005; Sadaghiani et al., 2009; Supe`r et al., 2003; Wyart and Tallon-Baudry, 2009). The timing of these events makes it logically
impossible that they already participate in the neural mechanism
of conscious access. Similar, early differences in sensory activation between conscious and nonconscious trials may reflect fluctuations in prestimulus priors and in sensory evidence that
contribute to subsequent conscious access, rather than be
constitutive of a conscious state per se (Dehaene and Changeux,
2005; Wyart and Tallon-Baudry, 2009).
Neuron
Review
preSMA and the junction of the bilateral anterior insula with the
inferior frontal gyrus. Only conscious no-go signals triggered
a broad and more anterior activation expanding into anterior
cingulate, inferior, and middle frontal gyrus, dorsolateral
prefrontal cortex, and inferior parietal cortexa network fully
compatible with the GNW model (see Figure 1).
Identifying the limits of nonconscious processing remains an
active area of research, as new techniques for presentation of
nonconscious stimuli are constantly appearing (e.g., Arnold
et al., 2008; Wilke et al., 2003). A recent masking study observed
that subliminal task-switching cues evoked detectable activations in premotor, prefrontal, and temporal cortices (Lau and
Passingham, 2007), but with a much reduced amplitude
compared to conscious cues. Another more challenging study
(Diaz and McCarthy, 2007) reported a large network of cortical
perisylvian regions (inferior frontal, inferior temporal, and angular
gyrus) activated by subliminal words relative to subliminal pseudowords, and surprisingly more extended than in previous
reports (e.g., Dehaene et al., 2001). Attentional blink studies
also suggest that unseen words may cause surprisingly longlasting ERP components (N400) (see also Gaillard et al., 2007;
Vogel et al., 1998). A crucial question for future research is
whether these activations remain confined to specialized subcircuits, for instance in the left temporal lobe (Sergent et al., 2005),
or whether they constitute true instances of global cortical processing without consciousness.
Do Prefrontal and Parietal Networks Play a Causal Role
in Conscious Access?
Brain imaging is only correlational in nature, and leaves open the
possibility that distributed ignition involving PFC is a mere
epiphenomenon or a consequence of conscious access, rather
than being one of its necessary causes. Causality is a demanding
concept that can only be assessed by systematic lesion or interference methods, which are of very limited applicability in human
subjects. Nevertheless, one prediction of the GNW model is testable: lesioning or interfering with prefrontal or parietal cortex
activity, at sites quite distant from visual areas, should disrupt
conscious vision. This prediction was initially judged as so counterintuitive as to be immediately refuted by clinical observations,
because frontal lobe patients do not appear to be unconscious
(Pollen, 1999). However, recent evidence actually supports the
GNW account. In normal subjects, transcranial magnetic stimulation (TMS) over either parietal or prefrontal cortex can prevent
conscious perception and even trigger a sudden subjective
disappearance of visual stimulis during prolonged fixation (Kanai
et al., 2008), change blindness (Beck et al., 2006), binocularly
rivalry (Carmel et al., 2010), inattentional blindness (Babiloni
et al., 2007), and attentional blink paradigms (Kihara et al.,
2011). Over prefrontal cortex, bilateral theta-burst TMS leads
to a reduction of subjective visibility with preserved objective
sensori-motor performance (Rounis et al., 2010). We recently
made similar observations in patients with focal prefrontal
lesions (Del Cul et al., 2009): their masking threshold was significantly elevated, in tight correlation with the degree of expansion
of the lesions into left anterior prefrontal cortex, while subliminal
performance on not-seen trials did not differ from normal. In
more severe and diffuse cases, following traumatic brain injury,
bilateral lesions of fronto-parietal cortices or, characteristically,
Neuron
Review
A
Figure 9. Cortical Measures of Loss of Consciousness in Sleep, Anesthesia, and Vegetative State
(A) Massive drops in cortical metabolism observed with PET rCBF measurements in slow-wave sleep (Maquet et al., 1997), anesthesia (Kaisti et al., 2002), and
vegetative state (Laureys et al., 2004).
(B) Reduced activity in a resting-state distributed cortical network in three vegetative state patients, as measured by independent component analysis of fMRI
data (adapted from Cauda et al., 2009).
(C) Sudden change in dimensional activation, a nonlinear dynamics measure of EEG complexity, at the precise point of loss of consciousness during anesthesia
(adapted from Velly et al., 2007). Signals were measured from the scalp as well as from the thalamus using depth electrodes (left). Only the scalp (cortical) EEG
showed a dramatic and discontinuous change accompanying loss of consciousness (right).
Neuron
Review
and propofol in human patients with Parkinson disease, cortical
EEG complexity decreased dramatically at the precise time where
consciousness was lost, while for several minutes there was little
change in subcortical signals, and eventually a slow decline
(Figure 9). These data suggest that in humans, the early stage of
anesthesia correlates with cortical disruption, and that the effects
on the thalamus are indirectly driven by cortical feedback (Alkire
et al., 2008). Indeed, in the course of anesthesia induction, there
is a decrease in EEG coherence in the 20 to 80 Hz frequency range
between right and left frontal cortices and between frontal and
occipital territories (John and Prichep, 2005). Quantitative analysis
of EEG under propofol induction further indicates a reduction of
mean information integration, as measured by Tononis Phi
measure, around the g-band (40 Hz) and a breakdown of the
spatiotemporal organization of this particular band (Lee et al.,
2009b). In agreement with experiments carried out with rats
(Imas et al., 2005; Imas et al., 2006), quantitative EEG analysis in
humans under propofol anesthesia induction noted a decrease
of directed feedback connectivity with loss of consciousness
and a return with responsiveness to verbal command (Lee et al.,
2009a). Also, during anesthesia induced by the benzodiazepine
midazolam, an externally induced transcranial pulse evoked reliable initial activity monitored by ERPs in humans, but the subsequent late phase of propagation to distributed areas was abolished (Ferrarelli et al., 2010). These observations are consistent
with the postulated role of top-down frontal-posterior amplification in conscious access (see also Supe`r et al., 2001).
Coma and vegetative state. The clinical distinctions between
coma, vegetative state (Laureys, 2005), and minimal consciousness (Giacino, 2005) remain poorly defined, and even fully
conscious but paralyzed patients with locked-in syndrome can
remain undetected. It is therefore of interest to see whether
objective neural measures and GNW theory can help discriminate
them. In coma and vegetative state, as with general anesthesia,
global metabolic activity typically decreases to 50% of normal
levels (Laureys, 2005). This decrease is not homogeneous,
however, but particularly pronounced in GNW areas including
lateral and mesial prefrontal and inferior parietal cortices
(Figure 9). Spontaneous recovery from VS is accompanied by
a functional restoration of this broad frontoparietal network (Laureys et al., 1999) and some of its cortico-thalamo-cortical
connections (Laureys et al., 2000; see also Voss et al., 2006).
Anatomically, prediction of recovery from coma relies on the
comprehensive assessment of all structures involved in arousal
and awareness functions, namely, the ascending reticular activating system located in the postero-superior part of the brainstem and structures encompassing thalamus, basal forebrain,
and fronto-parietal association cortices (Tshibanda et al.,
2009). Lesion or inhibition of part of this system suffices to cause
immediate coma (e.g., Parvizi and Damasio, 2003). Studies on
traumatic coma patients with conventional MRI showed that
lesions of the pons, midbrain, and basal ganglia were predictive
of poor outcome especially when they were bilateral (Tshibanda
et al., 2009). In relation with the GNW model, it is noteworthy that
prediction of nonrecovery after 1 year could be calculated with
up to 86% sensitivity and 97% specificity when taking into
account both diffusion tensor and spectroscopic measures of
cortical white matter integrity (Tshibanda et al., 2009).
Neuron
Review
received a sufficient empirical and neurophysiological definition to
figure in this review. Following Crick and Koch (1990), we focused
solely here on the simpler and well-studied question of what
neurophysiological mechanisms differentiate conscious access
to some information from nonconscious processing of the same
information. Additional work will be needed to explore, in the
future, these important aspects of higher-order consciousness.
In the present state of investigations, experimental measures
of conscious access identified in this review include: (1) sudden,
all-or-none ignition of prefronto-parietal networks; (2) concomitant all-or-none amplification of sensory activation; (3) a late
global P3b wave in event-related potentials; (4) late amplification
of broad-band power in the gamma range; (5) enhanced longdistance phase synchronization, particularly in the beta range;
and (6) enhanced causal relations between distant areas,
including a significant top-down component. Many of these
measures are also found during complex serial computations
and in spontaneous thought. There is evidence that they rely
on an anatomical network of long-distance connections that is
particularly developed in the human brain. Finally, pathologies
of these networks or their long-distance connections are associated with impairments of conscious access.
In the future, as argued by Haynes (2009), the mapping of
conscious experiences onto neural states will ultimately require
not only a neural distinction between seen and not-seen trials,
but also a proof that the proposed conscious neural state actually encodes all the details of the participants current subjective
experience. Criteria for a genuine one-to-one mapping should
include verifying that the proposed neural state has the same
perceptual stability (for instance over successive eye movements) and suffers from the same occasional illusions as the
subjects own report. Multivariate decoding techniques provide
pertinent tools to address this question and have already been
used to infer conscious mental images from early visual areas
(Haynes and Rees, 2005; Thirion et al., 2006) and from inferotemporal cortex (Schurger et al., 2010; Sterzer et al., 2008).
However, decoding the more intermingled neural patterns expected from PFC and other associative cortices is clearly a challenge for future research (though see Fuentemilla et al., 2010).
Another important question concerns the genetic mechanisms
that, in the coure of biological evolution, have led to the development of the GNW architecture, particularly the relative expansion
of PFC, higher associative cortices, and their underlying longdistance white matter tracts in the course of hominization (see
Avants et al., 2006; Schoenemann et al., 2005; Semendeferi
et al., 2002). Finally, now that measures of conscious processing
have been identified in human adults, it should become possible
to ask how they transpose to lower animal species (Changeux,
2006, 2010) and to human infants and fetuses (Dehaene-Lambertz et al., 2002; Gelskov and Kouider, 2010; Lagercrantz and
Changeux, 2009), in whom genuine but immature long-distance
networks have been described (Fair et al., 2009; Fransson et al.,
2007).
ACKNOWLEDGMENTS
We gratefully acknowledge extensive discussions with Lionel Naccache,
Sid Kouider, Jerome Sackur, Bechir Jarraya, and Pierre-Marie Lledo as
Neuron
Review
Boly, M., Tshibanda, L., Vanhaudenhuyse, A., Noirhomme, Q., Schnakers, C.,
Ledoux, D., Boveroux, P., Garweg, C., Lambermont, B., Phillips, C., et al.
(2009). Functional connectivity in the default network during resting state is
preserved in a vegetative but not in a brain dead patient. Hum. Brain Mapp.
30, 23932400.
Brazdil, M., Rektor, I., Dufek, M., Jurak, P., and Daniel, P. (1998). Effect of
subthreshold target stimuli on event-related potentials. Electroencephalogr.
Clin. Neurophysiol. 107, 6468.
Brazdil, M., Rektor, I., Daniel, P., Dufek, M., and Jurak, P. (2001). Intracerebral
event-related potentials to subthreshold target stimuli. Clin. Neurophysiol.
112, 650661.
Breitmeyer, B. (2006). Visual Masking: Time Slices through Conscious and
Unconscious Vision (New York: Oxford University Press).
Bressan, P., and Pizzighello, S. (2008). The attentional cost of inattentional
blindness. Cognition 106, 370383.
Bridgeman, B. (1975). Correlates of metacontrast in single cells of the cat
visual system. Vision Res. 15, 9199.
Bridgeman, B. (1988). Visual evoked potentials: Concomitants of metacontrast
in late components. Percept. Psychophys. 43, 401403.
Britz, J., Van De Ville, D., and Michel, C.M. (2010). BOLD correlates of EEG
topography reveal rapid resting-state network dynamics. Neuroimage 52,
11621170.
Broadbent, D.E. (1958). Perception and Communication (London: Pergamon).
Bullmore, E.T., Frangou, S., and Murray, R.M. (1997). The dysplastic net
hypothesis: An integration of developmental and dysconnectivity theories of
schizophrenia. Schizophr. Res. 28, 143156.
Buschman, T.J., and Miller, E.K. (2007). Top-down versus bottom-up control
of attention in the prefrontal and posterior parietal cortices. Science 315,
18601862.
Cajal, S.R. (18991904). Cajal on the Cerebral Cortex: An Annotated Translation of the Complete Writings, J. DeFelipe and E.G. Jones, trans. and eds.
(New York: Oxford University Press, 1988).
Christoff, K., Gordon, A.M., Smallwood, J., Smith, R., and Schooler, J.W.
(2009). Experience sampling during fMRI reveals default network and executive system contributions to mind wandering. Proc. Natl. Acad. Sci. USA
106, 87198724.
Chun, M.M., and Potter, M.C. (1995). A two-stage model for multiple target
detection in rapid serial visual presentation. J. Exp. Psychol. Hum. Percept.
Perform. 21, 109127.
Clark, R.E., Manns, J.R., and Squire, L.R. (2002). Classical conditioning,
awareness, and brain systems. Trends Cogn. Sci. (Regul. Ed.) 6, 524531.
Cleeremans, A., Timmermans, B., and Pasquali, A. (2007). Consciousness and
metarepresentation: A computational sketch. Neural Netw. 20, 10321039.
Corallo, G., Sackur, J., Dehaene, S., and Sigman, M. (2008). Limits on introspection: Distorted subjective time during the dual-task bottleneck. Psychol.
Sci. 19, 11101117.
Crick, F., and Koch, C. (1990). Some reflections on visual awareness. Cold
Spring Harb. Symp. Quant. Biol. 55, 953962.
Crick, F., and Koch, C. (1995). Are we aware of neural activity in primary visual
cortex? Nature 375, 121123.
Crick, F., and Koch, C. (2003). A framework for consciousness. Nat. Neurosci.
6, 119126.
Crick, F.C., and Koch, C. (2005). What is the function of the claustrum? Philos.
Trans. R. Soc. Lond. B Biol. Sci. 360, 12711279.
Damasio, A. (1999). The feeling of what happens (New York: Harcourt
Brace & Co.).
Damasio, A., and Meyer, D.E. (2009). Consciousness: An overview of the
phenomenon and of its possible neural basis. In The neurology of consciousness, S. Laureys and G. Tononi, eds. (Amsterdam: Elsevier), pp. 314.
Davis, M.H., Coleman, M.R., Absalom, A.R., Rodd, J.M., Johnsrude, I.S.,
Matta, B.F., Owen, A.M., and Menon, D.K. (2007). Dissociating speech
perception and comprehension at reduced levels of awareness. Proc. Natl.
Acad. Sci. USA 104, 1603216037.
Carmel, D., Walsh, V., Lavie, N., and Rees, G. (2010). Right parietal TMS
shortens dominance durations in binocular rivalry. Curr. Biol. 20, R799R800.
Cauda, F., Micon, B.M., Sacco, K., Duca, S., DAgata, F., Geminiani, G., and
Canavero, S. (2009). Disrupted intrinsic functional connectivity in the vegetative state. J. Neurol. Neurosurg. Psychiatry 80, 429431.
Cavada, C., Company, T., Tejedor, J., Cruz-Rizzolo, R.J., and Reinoso-Suarez,
F. (2000). The anatomical connections of the macaque monkey orbitofrontal
cortex. A review. Cereb. Cortex 10, 220242.
Cavanna, A.E., and Trimble, M.R. (2006). The precuneus: A review of its functional anatomy and behavioural correlates. Brain 129, 564583.
Changeux, J.P. (2006). The Ferrier Lecture 1998. The molecular biology of
consciousness investigated with genetically modified mice. Philos. Trans. R.
Soc. Lond. B Biol. Sci. 361, 22392259.
Changeux, J.P. (2010). Nicotine addiction and nicotinic receptors: Lessons
from genetically modified mice. Nat. Rev. Neurosci. 11, 389401.
Changeux, J.P., and Danchin, A. (1976). Selective stabilisation of developing
synapses as a mechanism for the specification of neuronal networks. Nature
264, 705712.
Changeux, J.P., and Dehaene, S. (2008). The neuronal workspace model:
Conscious processing and learning. In Learning Theory and Behavior.
Volume 1 of Learning and Memory: A Comprehensive Reference, J. Byrne
and R. Menzel, eds. (Oxford: Elsevier), pp. 729758.
Changeux, J.P., and Michel, C.M. (2004). Mechanisms of neural integration at
the brain-scale level. The neuronal workspace and microstate models. In
Microcircuits: The Interface between Neurons and Global Brain Function, S.
Grillner and A.M. Graybiel, eds. (Cambridge, MA: MIT Press), pp. 347370.
DeFelipe, J., and Farinas, I. (1992). The pyramidal neuron of the cerebral
cortex: Morphological and chemical characteristics of the synaptic inputs.
Prog. Neurobiol. 39, 563607.
Dehaene, S. (2008). Conscious and nonconscious processes: Distinct forms of
evidence accumulation? In Better Than Conscious? Decision Making, the
Human Mind, and Implications for Institutions. Strungmann Forum Report,
C. Engel and W. Singer, eds. (Cambridge: MIT Press).
Dehaene, S., and Changeux, J.P. (1989). A simple model of prefrontal cortex
function in delayed-response tasks. J. Cogn. Neurosci. 1, 244261.
Dehaene, S., and Changeux, J.P. (1991). The Wisconsin Card Sorting Test:
Theoretical analysis and modeling in a neuronal network. Cereb. Cortex 1,
6279.
Dehaene, S., and Changeux, J.P. (1997). A hierarchical neuronal network for
planning behavior. Proc. Natl. Acad. Sci. USA 94, 1329313298.
Dehaene, S., and Changeux, J.P. (2005). Ongoing spontaneous activity
controls access to consciousness: A neuronal model for inattentional blindness. PLoS Biol. 3, e141.
Dehaene, S., and Naccache, L. (2001). Towards a cognitive neuroscience of
consciousness: Basic evidence and a workspace framework. Cognition 79,
137.
Dehaene, S., Kerszberg, M., and Changeux, J.P. (1998a). A neuronal model of
a global workspace in effortful cognitive tasks. Proc. Natl. Acad. Sci. USA 95,
1452914534.
Neuron
Review
Dehaene, S., Naccache, L., Le ClecH, G., Koechlin, E., Mueller, M., DehaeneLambertz, G., van de Moortele, P.F., and Le Bihan, D. (1998b). Imaging unconscious semantic priming. Nature 395, 597600.
Dehaene, S., Naccache, L., Cohen, L., Bihan, D.L., Mangin, J.F., Poline, J.B.,
and Rivie`re, D. (2001). Cerebral mechanisms of word masking and unconscious repetition priming. Nat. Neurosci. 4, 752758.
Dehaene, S., Artiges, E., Naccache, L., Martelli, C., Viard, A., Schurhoff, F., Recasens, C., Martinot, M.L., Leboyer, M., and Martinot, J.L. (2003a). Conscious
and subliminal conflicts in normal subjects and patients with schizophrenia:
The role of the anterior cingulate. Proc. Natl. Acad. Sci. USA 100, 13722
13727.
Dehaene, S., Sergent, C., and Changeux, J.P. (2003b). A neuronal network
model linking subjective reports and objective physiological data during
conscious perception. Proc. Natl. Acad. Sci. USA 100, 85208525.
Dehaene, S., Changeux, J.P., Naccache, L., Sackur, J., and Sergent, C. (2006).
Conscious, preconscious, and subliminal processing: A testable taxonomy.
Trends Cogn. Sci. (Regul. Ed.) 10, 204211.
Dehaene-Lambertz, G., Dehaene, S., and Hertz-Pannier, L. (2002). Functional
neuroimaging of speech perception in infants. Science 298, 20132015.
Dejerine, J. (1895). Anatomie des Centres Nerveux, Volume 1 (Paris: Rueff
et Cie).
Del Cul, A., Dehaene, S., and Leboyer, M. (2006). Preserved subliminal processing and impaired conscious access in schizophrenia. Arch. Gen. Psychiatry 63, 13131323.
Del Cul, A., Baillet, S., and Dehaene, S. (2007). Brain dynamics underlying the
nonlinear threshold for access to consciousness. PLoS Biol. 5, e260.
Del Cul, A., Dehaene, S., Reyes, P., Bravo, E., and Slachevsky, A. (2009).
Causal role of prefrontal cortex in the threshold for access to consciousness.
Brain 132, 25312540.
Dellacqua, R., Jolicoeur, P., Vespignani, F., and Toffanin, P. (2005). Central
processing overlap modulates P3 latency. Exp. Brain Res. 165, 5468.
Demiralp, T., Ucok, A., Devrim, M., Isoglu-Alkac, U., Tecer, A., and Polich, J.
(2002). N2 and P3 components of event-related potential in first-episode
schizophrenic patients: Scalp topography, medication, and latency effects.
Psychiatry Res. 111, 167179.
Edelman, G.M. (1989). The Remembered Present (New York: Basic Books).
Elston, G.N. (2000). Pyramidal cells of the frontal lobe: All the more spinous to
think with. J. Neurosci. 20, RC95.
Elston, G.N. (2003). Cortex, cognition and the cell: New insights into the pyramidal neuron and prefrontal function. Cereb. Cortex 13, 11241138.
Elston, G.N., and Rosa, M.G. (1997). The occipitoparietal pathway of the
macaque monkey: Comparison of pyramidal cell morphology in layer III of
functionally related cortical visual areas. Cereb. Cortex 7, 432452.
Elston, G.N., and Rosa, M.G. (1998). Morphological variation of layer III pyramidal neurones in the occipitotemporal pathway of the macaque monkey
visual cortex. Cereb. Cortex 8, 278294.
Fahrenfort, J.J., Scholte, H.S., and Lamme, V.A. (2007). Masking disrupts
reentrant processing in human visual cortex. J. Cogn. Neurosci. 19, 1488
1497.
Fair, D.A., Cohen, A.L., Power, J.D., Dosenbach, N.U., Church, J.A., Miezin,
F.M., Schlaggar, B.L., and Petersen, S.E. (2009). Functional brain networks
develop from a local to distributed organization. PLoS Comput. Biol. 5,
e1000381.
Farrer, C., Frey, S.H., Van Horn, J.D., Tunik, E., Turk, D., Inati, S., and Grafton,
S.T. (2008). The angular gyrus computes action awareness representations.
Cereb. Cortex 18, 254261.
Faugeras, F., Rohaut, B., Weiss, N., Bekinschtein, T., Galanaud, D., Puybasset, L., Bolgert, F., Sergent, C., Cohen, L., Dehaene, S., and Naccache, L.
(2011). Probing consciousness in clinically defined vegetative patients with
event-related potentials. Neurology, in press.
Fernandez-Duque, D., Grossi, G., Thornton, I.M., and Neville, H.J. (2003).
Representation of change: Separate electrophysiological markers of attention,
awareness, and implicit processing. J. Cogn. Neurosci. 15, 491507.
Ferrarelli, F., Massimini, M., Sarasso, S., Casali, A., Riedner, B.A., Angelini, G.,
Tononi, G., and Pearce, R.A. (2010). Breakdown in cortical effective connectivity during midazolam-induced loss of consciousness. Proc. Natl. Acad.
Sci. USA 107, 26812686.
Fisch, L., Privman, E., Ramot, M., Harel, M., Nir, Y., Kipervasser, S., Andelman,
F., Neufeld, M.Y., Kramer, U., Fried, I., and Malach, R. (2009). Neural ignition:
Enhanced activation linked to perceptual awareness in human ventral stream
visual cortex. Neuron 64, 562574.
Fischer, C., Luaute, J., Adeleine, P., and Morlet, D. (2004). Predictive value of
sensory and cognitive evoked potentials for awakening from coma. Neurology
63, 669673.
Desmurget, M., Reilly, K.T., Richard, N., Szathmari, A., Mottolese, C., and
Sirigu, A. (2009). Movement intention after parietal cortex stimulation in humans. Science 324, 811813.
Fleming, S.M., Weil, R.S., Nagy, Z., Dolan, R.J., and Rees, G. (2010). Relating
introspective accuracy to individual differences in brain structure. Science
329, 15411543.
Forget, J., Buiatti, M., and Dehaene, S. (2010). Temporal integration in visual
word recognition. J. Cogn. Neurosci. 22, 10541068.
Diekhof, E.K., Biedermann, F., Ruebsamen, R., and Gruber, O. (2009). Topdown and bottom-up modulation of brain structures involved in auditory
discrimination. Brain Res. 1297, 118123.
Forman, S.A., and Miller, K.W. (2011). Anesthetic sites and allosteric mechanisms of action on Cys-loop ligand-gated ion channels. Can. J. Anaesth. 58,
191205.
Doesburg, S.M., Green, J.J., McDonald, J.J., and Ward, L.M. (2009). Rhythms
of consciousness: Binocular rivalry reveals large-scale oscillatory network
dynamics mediating visual perception. PLoS ONE 4, e6142.
Fox, M.D., Corbetta, M., Snyder, A.Z., Vincent, J.L., and Raichle, M.E. (2006).
Spontaneous neuronal activity distinguishes human dorsal and ventral attention systems. Proc. Natl. Acad. Sci. USA 103, 1004610051.
Dupoux, E., de Gardelle, V., and Kouider, S. (2008). Subliminal speech perception and auditory streaming. Cognition 109, 267273.
Dux, P.E., Ivanoff, J., Asplund, C.L., and Marois, R. (2006). Isolation of a central
bottleneck of information processing with time-resolved FMRI. Neuron 52,
11091120.
Fransson, P., Skiold, B., Horsch, S., Nordell, A., Blennow, M., Lagercrantz, H.,
and Aden, U. (2007). Resting-state networks in the infant brain. Proc. Natl.
Acad. Sci. USA 104, 1553115536.
Dux, P.E., Tombu, M.N., Harrison, S., Rogers, B.P., Tong, F., and Marois, R.
(2009). Training improves multitasking performance by increasing the speed
of information processing in human prefrontal cortex. Neuron 63, 127138.
Eccles, J.C. (1994). How the Self Controls Its Brain (New York: Springer
Verlag).
Edelman, G. (1987). Neural Darwinism (New York: Basic Books).
Frith, C. (2007). Making up the Mind. How the Brain Creates Our Mental World
(London: Blackwell).
Fuentemilla, L., Penny, W.D., Cashdollar, N., Bunzeck, N., and Duzel, E. (2010).
Theta-coupled periodic replay in working memory. Curr. Biol. 20, 606612.
Neuron
Review
Fuster, J.M. (2008). The Prefrontal Cortex, Fourth Edition (London: Academic
Press).
Gaillard, R., Del Cul, A., Naccache, L., Vinckier, F., Cohen, L., and Dehaene, S.
(2006). Nonconscious semantic processing of emotional words modulates
conscious access. Proc. Natl. Acad. Sci. USA 103, 75247529.
Gaillard, R., Cohen, L., Adam, C., Clemenceau, S., Hasboun, D., Baulac, M.,
Willer, J.C., Dehaene, S., and Naccache, L. (2007). Subliminal words durably
affect neuronal activity. Neuroreport 18, 15271531.
Gaillard, R., Dehaene, S., Adam, C., Clemenceau, S., Hasboun, D., Baulac, M.,
Cohen, L., and Naccache, L. (2009). Converging intracranial markers of
conscious access. PLoS Biol. 7, e61.
Gazzaniga, M.S., LeDoux, J.E., and Wilson, D.H. (1977). Language, praxis, and
the right hemisphere: Clues to some mechanisms of consciousness.
Neurology 27, 11441147.
Geldard, F.A., and Sherrick, C.E. (1972). The cutaneous rabbit: A perceptual
illusion. Science 178, 178179.
Gelskov, S.V., and Kouider, S. (2010). Psychophysical thresholds of face visibility during infancy. Cognition 114, 285292.
Giacino, J.T. (2005). The minimally conscious state: Defining the borders of
consciousness. Prog. Brain Res. 150, 381395.
Goldman-Rakic, P.S. (1988). Topography of cognition: Parallel distributed
networks in primate association cortex. Annu. Rev. Neurosci. 11, 137156.
Goldman-Rakic, P.S. (1999). The psychic neuron of the cerebral cortex. Ann.
N Y Acad. Sci. 868, 1326.
Greenwald, A.G., Draine, S.C., and Abrams, R.L. (1996). Three cognitive
markers of unconscious semantic activation. Science 273, 16991702.
Gregoriou, G.G., Gotts, S.J., Zhou, H., and Desimone, R. (2009). Highfrequency, long-range coupling between prefrontal and visual cortex during
attention. Science 324, 12071210.
Greicius, M.D., Krasnow, B., Reiss, A.L., and Menon, V. (2003). Functional
connectivity in the resting brain: A network analysis of the default mode
hypothesis. Proc. Natl. Acad. Sci. USA 100, 253258.
Greicius, M.D., Kiviniemi, V., Tervonen, O., Vainionpaa, V., Alahuhta, S., Reiss,
A.L., and Menon, V. (2008). Persistent default-mode network connectivity
during light sedation. Hum. Brain Mapp. 29, 839847.
Haynes, J.D., Driver, J., and Rees, G. (2005b). Visibility reflects dynamic
changes of effective connectivity between V1 and fusiform cortex. Neuron
46, 811821.
He, B.J., and Raichle, M.E. (2009). The fMRI signal, slow cortical potential and
consciousness. Trends Cogn. Sci. (Regul. Ed.) 13, 302309.
He, B.J., Snyder, A.Z., Vincent, J.L., Epstein, A., Shulman, G.L., and Corbetta,
M. (2007). Breakdown of functional connectivity in frontoparietal networks
underlies behavioral deficits in spatial neglect. Neuron 53, 905918.
He, B.J., Snyder, A.Z., Zempel, J.M., Smyth, M.D., and Raichle, M.E. (2008).
Electrophysiological correlates of the brains intrinsic large-scale functional
architecture. Proc. Natl. Acad. Sci. USA 105, 1603916044.
He, Y., Dagher, A., Chen, Z., Charil, A., Zijdenbos, A., Worsley, K., and Evans,
A. (2009). Impaired small-world efficiency in structural cortical networks in
multiple sclerosis associated with white matter lesion load. Brain 132, 3366
3379.
Heinemann, A., Kunde, W., and Kiesel, A. (2009). Context-specific primecongruency effects: On the role of conscious stimulus representations for
cognitive control. Conscious. Cogn. 18, 966976.
Hipp, J.F., Engel, A.K., and Siegel, M. (2011). Oscillatory synchronization in
large-scale cortical networks predicts perception. Neuron 69, 387396.
Holland, P.C., and Gallagher, M. (2004). Amygdala-frontal interactions and
reward expectancy. Curr. Opin. Neurobiol. 14, 148155.
Huang, L. (2010). What is the unit of visual attention? Object for selection, but
Boolean map for access. J. Exp. Psychol. Gen. 139, 162179.
Husain, M., and Kennard, C. (1996). Visual neglect associated with frontal lobe
infarction. J. Neurol. 243, 652657.
Imas, O.A., Ropella, K.M., Ward, B.D., Wood, J.D., and Hudetz, A.G. (2005).
Volatile anesthetics enhance flash-induced gamma oscillations in rat visual
cortex. Anesthesiology 102, 937947.
Imas, O.A., Ropella, K.M., Wood, J.D., and Hudetz, A.G. (2006). Isoflurane
disrupts anterio-posterior phase synchronization of flash-induced field potentials in the rat. Neurosci. Lett. 402, 216221.
Grill-Spector, K., Kushnir, T., Hendler, T., and Malach, R. (2000). The dynamics
of object-selective activation correlate with recognition performance in
humans. Nat. Neurosci. 3, 837843.
Groom, M.J., Bates, A.T., Jackson, G.M., Calton, T.G., Liddle, P.F., and Hollis,
C. (2008). Event-related potentials in adolescents with schizophrenia and their
siblings: A comparison with attention-deficit/hyperactivity disorder. Biol.
Psychiatry 63, 784792.
Gross, J., Schmitz, F., Schnitzler, I., Kessler, K., Shapiro, K., Hommel, B., and
Schnitzler, A. (2004). Modulation of long-range neural synchrony reflects
temporal limitations of visual attention in humans. Proc. Natl. Acad. Sci. USA
101, 1305013055.
Gutschalk, A., Micheyl, C., and Oxenham, A.J. (2008). Neural correlates of
auditory perceptual awareness under informational masking. PLoS Biol.
6, e138.
Hagmann, P., Cammoun, L., Gigandet, X., Meuli, R., Honey, C.J., Wedeen,
V.J., and Sporns, O. (2008). Mapping the structural core of human cerebral
cortex. PLoS Biol. 6, e159.
Halgren, E., Marinkovic, K., and Chauvel, P. (1998). Generators of the late
cognitive potentials in auditory and visual oddball tasks. Electroencephalogr.
Clin. Neurophysiol. 106, 156164.
Javitt, D.C., Steinschneider, M., Schroeder, C.E., and Arezzo, J.C. (1996). Role
of cortical N-methyl-D-aspartate receptors in auditory sensory memory and
mismatch negativity generation: Implications for schizophrenia. Proc. Natl.
Acad. Sci. USA 93, 1196211967.
Jaynes, J. (1976). The Origin of Consciousness in the Breakdown of the Bicameral Mind (New York: Houghton Mifflin Company).
John, E.R., and Prichep, L.S. (2005). The anesthetic cascade: A theory of how
anesthesia suppresses consciousness. Anesthesiology 102, 447471.
Jolicoeur, P. (1999). Concurrent response-selection demands modulate the
attentional blink. J. Exp. Psychol. Hum. Percept. Perform. 25, 10971113.
Jones, S.R., Pritchett, D.L., Stufflebeam, S.M., Hamalainen, M., and Moore,
C.I. (2007). Neural correlates of tactile detection: A combined magnetoencephalography and biophysically based computational modeling study. J. Neurosci. 27, 1075110764.
Hasson, U., Skipper, J.I., Nusbaum, H.C., and Small, S.L. (2007). Abstract
coding of audiovisual speech: Beyond sensory representation. Neuron 56,
11161126.
Kaisti, K.K., Metsahonkala, L., Teras, M., Oikonen, V., Aalto, S., Jaaskelainen,
S., Hinkka, S., and Scheinin, H. (2002). Effects of surgical levels of propofol and
sevoflurane anesthesia on cerebral blood flow in healthy subjects studied with
positron emission tomography. Anesthesiology 96, 13581370.
Kanai, R., Muggleton, N.G., and Walsh, V. (2008). TMS over the intraparietal
sulcus induces perceptual fading. J. Neurophysiol. 100, 33433350.
Neuron
Review
Kanai, R., Walsh, V., and Tseng, C.H. (2010). Subjective discriminability of
invisibility: A framework for distinguishing perceptual and attentional failures
of awareness. Conscious. Cogn. 19, 10451057.
Kringelbach, M.L., and Rolls, E.T. (2004). The functional neuroanatomy of the
human orbitofrontal cortex: Evidence from neuroimaging and neuropsychology. Prog. Neurobiol. 72, 341372.
Karlsgodt, K.H., Sun, D., Jimenez, A.M., Lutkenhoff, E.S., Willhite, R., van Erp,
T.G., and Cannon, T.D. (2008). Developmental disruptions in neural connectivity in the pathophysiology of schizophrenia. Dev. Psychopathol. 20, 1297
1327.
Kayser, J., Tenke, C.E., Gates, N.A., Kroppmann, C.J., Gil, R.B., and Bruder,
G.E. (2006). ERP/CSD indices of impaired verbal working memory subprocesses in schizophrenia. Psychophysiology 43, 237252.
Kentridge, R.W., Nijboer, T.C., and Heywood, C.A. (2008). Attended but
unseen: Visual attention is not sufficient for visual awareness. Neuropsychologia 46, 864869.
Kiani, R., and Shadlen, M.N. (2009). Representation of confidence associated
with a decision by neurons in the parietal cortex. Science 324, 759764.
Kiebel, S.J., Daunizeau, J., and Friston, K.J. (2008). A hierarchy of time-scales
and the brain. PLoS Comput. Biol. 4, e1000209.
Kiefer, M., and Brendel, D. (2006). Attentional modulation of unconscious
automatic processes: Evidence from event-related potentials in a masked
priming paradigm. J. Cogn. Neurosci. 18, 184198.
Kunde, W. (2003). Sequential modulations of stimulus-response correspondence effects depend on awareness of response conflict. Psychon. Bull.
Rev. 10, 198205.
Lagercrantz, H., and Changeux, J.P. (2009). The emergence of human
consciousness: From fetal to neonatal life. Pediatr. Res. 65, 255260.
Lamme, V.A., and Roelfsema, P.R. (2000). The distinct modes of vision offered
by feedforward and recurrent processing. Trends Neurosci. 23, 571579.
Lamme, V.A., Zipser, K., and Spekreijse, H. (1998). Figure-ground activity in
primary visual cortex is suppressed by anesthesia. Proc. Natl. Acad. Sci.
USA 95, 32633268.
Lamme, V.A., Zipser, K., and Spekreijse, H. (2002). Masking interrupts figureground signals in V1. J. Cogn. Neurosci. 14, 10441053.
Lamy, D., Salti, M., and Bar-Haim, Y. (2009). Neural correlates of subjective
awareness and unconscious processing: An ERP study. J. Cogn. Neurosci.
21, 14351446.
Kihara, K., Ikeda, T., Matsuyoshi, D., Hirose, N., Mima, T., Fukuyama, H., and
Osaka, N. (2011). Differential contributions of the intraparietal sulcus and the
inferior parietal lobe to attentional blink: Evidence from transcranial magnetic
stimulation. J. Cogn. Neurosci. 23, 247256.
Landmann, C., Dehaene, S., Pappata, S., Jobert, A., Bottlaender, M., Roumenov, D., and Le Bihan, D. (2007). Dynamics of prefrontal and cingulate activity
during a reward-based logical deduction task. Cereb. Cortex 17, 749759.
Kim, C.Y., and Blake, R. (2005). Psychophysical magic: Rendering the visible
invisible. Trends Cogn. Sci. (Regul. Ed.) 9, 381388.
Kinoshita, S., Forster, K.I., and Mozer, M.C. (2008). Unconscious cognition
isnt that smart: Modulation of masked repetition priming effect in the word
naming task. Cognition 107, 623649.
Klein, T.A., Endrass, T., Kathmann, N., Neumann, J., von Cramon, D.Y., and
Ullsperger, M. (2007). Neural correlates of error awareness. Neuroimage 34,
17741781.
Koch, C., and Tsuchiya, N. (2007). Attention and consciousness: Two distinct
brain processes. Trends Cogn. Sci. (Regul. Ed.) 11, 1622.
Koechlin, E., Ody, C., and Kouneiher, F. (2003). The architecture of cognitive
control in the human prefrontal cortex. Science 302, 11811185.
Koivisto, M., Revonsuo, A., and Lehtonen, M. (2006). Independence of visual
awareness from the scope of attention: An electrophysiological study. Cereb.
Cortex 16, 415424.
Koivisto, M., Lahteenmaki, M., Srensen, T.A., Vangkilde, S., Overgaard, M.,
and Revonsuo, A. (2008). The earliest electrophysiological correlate of visual
awareness? Brain Cogn. 66, 91103.
Koivisto, M., Kainulainen, P., and Revonsuo, A. (2009). The relationship
between awareness and attention: Evidence from ERP responses. Neuropsychologia 47, 28912899.
Kouider, S., and Dehaene, S. (2007). Levels of processing during nonconscious perception: A critical review of visual masking. Philos. Trans. R.
Soc. Lond. B Biol. Sci. 362, 857875.
Kouider, S., Dehaene, S., Jobert, A., and Le Bihan, D. (2007). Cerebral bases of
subliminal and supraliminal priming during reading. Cereb. Cortex 17, 2019
2029.
Kouider, S., Eger, E., Dolan, R., and Henson, R.N. (2009). Activity in faceresponsive brain regions is modulated by invisible, attended faces: Evidence
from masked priming. Cereb. Cortex 19, 1323.
Lau, H.C., and Passingham, R.E. (2007). Unconscious activation of the cognitive control system in the human prefrontal cortex. J. Neurosci. 27, 58055811.
Laureys, S. (2005). The neural correlate of (un)awareness: Lessons from the
vegetative state. Trends Cogn. Sci. (Regul. Ed.) 9, 556559.
Laureys, S., Lemaire, C., Maquet, P., Phillips, C., and Franck, G. (1999). Cerebral metabolism during vegetative state and after recovery to consciousness.
J. Neurol. Neurosurg. Psychiatry 67, 121.
Laureys, S., Faymonville, M.E., Luxen, A., Lamy, M., Franck, G., and Maquet,
P. (2000). Restoration of thalamocortical connectivity after recovery from
persistent vegetative state. Lancet 355, 17901791.
Laureys, S., Owen, A.M., and Schiff, N.D. (2004). Brain function in coma, vegetative state, and related disorders. Lancet Neurol. 3, 537546.
Lee, U., Kim, S., Noh, G.J., Choi, B.M., Hwang, E., and Mashour, G.A. (2009a).
The directionality and functional organization of frontoparietal connectivity
during consciousness and anesthesia in humans. Conscious. Cogn. 18,
10691078.
Lee, U., Mashour, G.A., Kim, S., Noh, G.J., and Choi, B.M. (2009b). Propofol
induction reduces the capacity for neural information integration: Implications
for the mechanism of consciousness and general anesthesia. Conscious.
Cogn. 18, 5664.
Lehmann, D., and Koenig, T. (1997). Spatio-temporal dynamics of alpha brain
electric fields, and cognitive modes. Int. J. Psychophysiol. 26, 99112.
Lehmann, D., Strik, W.K., Henggeler, B., Koenig, T., and Koukkou, M. (1998).
Brain electric microstates and momentary conscious mind states as building
blocks of spontaneous thinking: I. Visual imagery and abstract thoughts. Int.
J. Psychophysiol. 29, 111.
Kovacs, G., Vogels, R., and Orban, G.A. (1995). Cortical correlate of pattern
backward masking. Proc. Natl. Acad. Sci. USA 92, 55875591.
Lehmann, D., Pascual-Marqui, R.D., Strik, W.K., and Koenig, T. (2010). Core
networks for visual-concrete and abstract thought content: A brain electric
microstate analysis. Neuroimage 49, 10731079.
Kranczioch, C., Debener, S., Maye, A., and Engel, A.K. (2007). Temporal
dynamics of access to consciousness in the attentional blink. Neuroimage
37, 947955.
Leopold, D.A., and Logothetis, N.K. (1996). Activity changes in early visual
cortex reflect monkeys percepts during binocular rivalry. Nature 379,
549553.
Neuron
Review
Leuthold, H., and Kopp, B. (1998). Mechanisms of priming by masked stimuli:
Inferences from event-related potentials. Psychol. Sci. 9, 263269.
Lhermitte, F. (1983). Utilization behaviour and its relation to lesions of the
frontal lobes. Brain 106, 237255.
Li, G.D., Chiara, D.C., Cohen, J.B., and Olsen, R.W. (2010). Numerous classes
of general anesthetics inhibit etomidate binding to gamma-aminobutyric acid
type A (GABAA) receptors. J. Biol. Chem. 285, 86158620.
Libet, B., Gleason, C.A., Wright, E.W., and Pearl, D.K. (1983). Time of
conscious intention to act in relation to onset of cerebral activity (readinesspotential). The unconscious initiation of a freely voluntary act. Brain 106,
623642.
Llinas, R.R., and Pare, D. (1991). Of dreaming and wakefulness. Neuroscience
44, 521535.
Llinas, R., Ribary, U., Contreras, D., and Pedroarena, C. (1998). The neuronal
basis for consciousness. Philos. Trans. R. Soc. Lond. B Biol. Sci. 353, 1841
1849.
Logan, G.D., and Crump, M.J. (2010). Cognitive illusions of authorship reveal
hierarchical error detection in skilled typists. Science 330, 683686.
Luck, S.J., Fuller, R.L., Braun, E.L., Robinson, B., Summerfelt, A., and Gold,
J.M. (2006). The speed of visual attention in schizophrenia: Electrophysiological and behavioral evidence. Schizophr. Res. 85, 174195.
Lynall, M.E., Bassett, D.S., Kerwin, R., McKenna, P.J., Kitzbichler, M., Muller,
U., and Bullmore, E. (2010). Functional connectivity and brain networks in
schizophrenia. J. Neurosci. 30, 94779487.
Mack, A., and Rock, I. (1998). Inattentional Blindness (Cambridge, Mass: MIT
Press).
Macknik, S.L., and Haglund, M.M. (1999). Optical images of visible and invisible percepts in the primary visual cortex of primates. Proc. Natl. Acad. Sci.
USA 96, 1520815210.
Macknik, S.L., and Livingstone, M.S. (1998). Neuronal correlates of visibility
and invisibility in the primate visual system. Nat. Neurosci. 1, 144149.
Maier, A., Wilke, M., Aura, C., Zhu, C., Ye, F.Q., and Leopold, D.A. (2008).
Divergence of fMRI and neural signals in V1 during perceptual suppression
in the awake monkey. Nat. Neurosci. 11, 11931200.
Mantini, D., Perrucci, M.G., Del Gratta, C., Romani, G.L., and Corbetta, M.
(2007). Electrophysiological signatures of resting state networks in the human
brain. Proc. Natl. Acad. Sci. USA 104, 1317013175.
Mantini, D., Corbetta, M., Perrucci, M.G., Romani, G.L., and Del Gratta, C.
(2009). Large-scale brain networks account for sustained and transient activity
during target detection. Neuroimage 44, 265274.
Maquet, P., Degueldre, C., Delfiore, G., Aerts, J., Peters, J.M., Luxen, A., and
Franck, G. (1997). Functional neuroanatomy of human slow wave sleep. J.
Neurosci. 17, 28072812.
Marois, R., and Ivanoff, J. (2005). Capacity limits of information processing in
the brain. Trends Cogn. Sci. (Regul. Ed.) 9, 296305.
Marois, R., Yi, D.J., and Chun, M.M. (2004). The neural fate of consciously
perceived and missed events in the attentional blink. Neuron 41, 465472.
Marti, S., Sackur, J., Sigman, M., and Dehaene, S. (2010). Mapping introspections blind spot: Reconstruction of dual-task phenomenology using quantified
introspection. Cognition 115, 303313.
Mason, M.F., Norton, M.I., Van Horn, J.D., Wegner, D.M., Grafton, S.T., and
Macrae, C.N. (2007). Wandering minds: The default network and stimulusindependent thought. Science 315, 393395.
Mattler, U. (2005). Inhibition and decay of motor and nonmotor priming.
Percept. Psychophys. 67, 285300.
Melchitzky, D.S., Sesack, S.R., Pucak, M.L., and Lewis, D.A. (1998). Synaptic
targets of pyramidal neurons providing intrinsic horizontal connections in
monkey prefrontal cortex. J. Comp. Neurol. 390, 211224.
Melchitzky, D.S., Gonzalez-Burgos, G., Barrionuevo, G., and Lewis, D.A.
(2001). Synaptic targets of the intrinsic axon collaterals of supragranular pyramidal neurons in monkey prefrontal cortex. J. Comp. Neurol. 430, 209221.
Melloni, L., Molina, C., Pena, M., Torres, D., Singer, W., and Rodriguez, E.
(2007). Synchronization of neural activity across cortical areas correlates
with conscious perception. J. Neurosci. 27, 28582865.
Melloni, L., Schwiedrzik, C.M., Muller, N., Rodriguez, E., and Singer, W. (2011).
Expectations change the signatures and timing of electrophysiological correlates of perceptual awareness. J. Neurosci. 31, 13861396.
Merikle, P.M., and Joordens, S. (1997). Parallels between perception without
attention and perception without awareness. Conscious. Cogn. 6, 219236.
Meyer, K., and Damasio, A. (2009). Convergence and divergence in a neural
architecture for recognition and memory. Trends Neurosci. 32, 376382.
Monti, M.M., Vanhaudenhuyse, A., Coleman, M.R., Boly, M., Pickard, J.D.,
Tshibanda, L., Owen, A.M., and Laureys, S. (2010). Willful modulation of brain
activity in disorders of consciousness. N. Engl. J. Med. 362, 579589.
Muller-Gass, A., Macdonald, M., Schroger, E., Sculthorpe, L., and Campbell,
K. (2007). Evidence for the auditory P3a reflecting an automatic process: Elicitation during highly-focused continuous visual attention. Brain Res. 1170,
7178.
Naatanen, R. (1990). The role of attention in auditory information processing as
revealed by event-related potentials and other brain measures of cognitive
function. Behav. Brain Sci. 13, 201288.
Naccache, L., and Dehaene, S. (2001). Unconscious semantic priming extends
to novel unseen stimuli. Cognition 80, 215229.
Naccache, L., Blandin, E., and Dehaene, S. (2002). Unconscious masked
priming depends on temporal attention. Psychol. Sci. 13, 416424.
Niedeggen, M., Wichmann, P., and Stoerig, P. (2001). Change blindness and
time to consciousness. Eur. J. Neurosci. 14, 17191726.
Nieuwenhuis, S., Ridderinkhof, K.R., Blom, J., Band, G.P., and Kok, A. (2001).
Error-related brain potentials are differentially related to awareness of
response errors: Evidence from an antisaccade task. Psychophysiology 38,
752760.
Nieuwenstein, M., Van der Burg, E., Theeuwes, J., Wyble, B., and Potter, M.
(2009). Temporal constraints on conscious vision: On the ubiquitous nature
of the attentional blink. J. Vis. 9, 18.118.14.
Niswender, C.M., and Conn, P.J. (2010). Metabotropic glutamate receptors:
Physiology, pharmacology, and disease. Annu. Rev. Pharmacol. Toxicol. 50,
295322.
Norman, D.A., and Shallice, T. (1980). Attention to action: Willed and automatic
control of behavior. In Consciousness and Self-Regulation, R.J. Davidson,
G.E. Schwartz, and D. Shapiro, eds. (New York: Plenum Press), pp. 118.
Nury, H., Van Renterghem, C., Weng, Y., Tran, A., Baaden, M., Dufresne, V.,
Changeux, J.P., Sonner, J.M., Delarue, M., and Corringer, P.J. (2011). X-ray
structures of general anaesthetics bound to a pentameric ligand-gated ion
channel. Nature 469, 428431.
Oh, J.S., Kubicki, M., Rosenberger, G., Bouix, S., Levitt, J.J., McCarley, R.W.,
Westin, C.F., and Shenton, M.E. (2009). Thalamo-frontal white matter alterations in chronic schizophrenia: A quantitative diffusion tractography study.
Hum. Brain Mapp. 30, 38123825.
Owen, A.M., Coleman, M.R., Boly, M., Davis, M.H., Laureys, S., and Pickard,
J.D. (2006). Detecting awareness in the vegetative state. Science 313, 1402.
Palva, S., Linkenkaer-Hansen, K., Naatanen, R., and Palva, J.M. (2005). Early
neural correlates of conscious somatosensory perception. J. Neurosci. 25,
52485258.
Pandya, D.N., and Yeterian, E.H. (1990). Prefrontal cortex in relation to other
cortical areas in rhesus monkey: Architecture and connections. Prog. Brain
Res. 85, 6394.
Neuron
Review
Parvizi, J., and Damasio, A.R. (2003). Neuroanatomical correlates of brainstem
coma. Brain 126, 15241536.
Parvizi, J., Van Hoesen, G.W., Buckwalter, J., and Damasio, A. (2006). Neural
connections of the posteromedial cortex in the macaque. Proc. Natl. Acad.
Sci. USA 103, 15631568.
Pashler, H. (1994). Dual-task interference in simple tasks: Data and theory.
Psychol. Bull. 116, 220244.
Passingham, R. (1993). The Frontal Lobes and Voluntary Action, Volume 21
(New York: Oxford University Press).
Penrose, R. (1990). The Emperors New Mind. Concerning Computers, Minds,
and the Laws of Physics (London: Vintage books).
Roopun, A.K., Cunningham, M.O., Racca, C., Alter, K., Traub, R.D., and Whittington, M.A. (2008). Region-specific changes in gamma and beta2 rhythms in
NMDA receptor dysfunction models of schizophrenia. Schizophr. Bull. 34,
962973.
Pessiglione, M., Schmidt, L., Draganski, B., Kalisch, R., Lau, H., Dolan, R.J.,
and Frith, C.D. (2007). How the brain translates money into force: A neuroimaging study of subliminal motivation. Science 316, 904906.
Rosenthal, D.M. (2004). Varieties of higher-order theory. In Higher-Order Theories of Consciousness, R.J. Gennaro, ed. (Philadelphia: John Benjamins),
pp. 1944.
Petrides, M., and Pandya, D.N. (2009). Distinct parietal and temporal pathways
to the homologues of Brocas area in the monkey. PLoS Biol. 7, e1000170.
Rougier, N.P., Noelle, D.C., Braver, T.S., Cohen, J.D., and OReilly, R.C.
(2005). Prefrontal cortex and flexible cognitive control: Rules without symbols.
Proc. Natl. Acad. Sci. USA 102, 73387343.
Pins, D., and Ffytche, D. (2003). The neural correlates of conscious vision.
Cereb. Cortex 13, 461474.
Pollen, D.A. (1999). On the neural correlates of visual perception. Cereb.
Cortex 9, 419.
Polonsky, A., Blake, R., Braun, J., and Heeger, D.J. (2000). Neuronal activity in
human primary visual cortex correlates with perception during binocular
rivalry. Nat. Neurosci. 3, 11531159.
Posner, M.I., and Dehaene, S. (1994). Attentional networks. Trends Neurosci.
17, 7579.
Posner, M.I., and Rothbart, M.K. (1998). Attention, self-regulation and
consciousness. Philos. Trans. R. Soc. Lond. B Biol. Sci. 353, 19151927.
Posner, N.I., and Snyder, C.R.R. (1975). Attention and cognitive control. In
Information Processing and Cognition: The Loyola Symposium, R.L. Solso,
ed. (Hillsdale: L. Erlbaum), pp. 205223.
Rounis, E., Maniscalco, B., Rothwell, J.C., Passingham, R., and Lau, H. (2010).
Theta-burst transcranial magnetic stimulation to the prefrontal cortex impairs
metacognitive visual awareness. Cognitive Neuroscience 1, 165175.
Sackur, J., and Dehaene, S. (2009). The cognitive architecture for chaining of
two mental operations. Cognition 111, 187211.
Sadaghiani, S., Hesselmann, G., and Kleinschmidt, A. (2009). Distributed and
antagonistic contributions of ongoing activity fluctuations to auditory stimulus
detection. J. Neurosci. 29, 1341013417.
Sadaghiani, S., Scheeringa, R., Lehongre, K., Morillon, B., Giraud, A.L., and
Kleinschmidt, A. (2010). Intrinsic connectivity networks, alpha oscillations,
and tonic alertness: A simultaneous electroencephalography/functional
magnetic resonance imaging study. J. Neurosci. 30, 1024310250.
Salisbury, D., Squires, N.K., Ibel, S., and Maloney, T. (1992). Auditory eventrelated potentials during stage 2 NREM sleep in humans. J. Sleep Res. 1,
251257.
Schiller, P.H., and Chorover, S.L. (1966). Metacontrast: Its relation to evoked
potentials. Science 153, 13981400.
Procyk, E., Tanaka, Y.L., and Joseph, J.P. (2000). Anterior cingulate activity
during routine and non-routine sequential behaviors in macaques. Nat. Neurosci.
3, 502508.
Schoenemann, P.T., Sheehan, M.J., and Glotzer, L.D. (2005). Prefrontal white
matter volume is disproportionately larger in humans than in other primates.
Nat. Neurosci. 8, 242252.
Pucak, M.L., Levitt, J.B., Lund, J.S., and Lewis, D.A. (1996). Patterns of intrinsic
and associational circuitry in monkey prefrontal cortex. J. Comp. Neurol. 376,
614630.
Schrouff, J., Perlbarg, V., Boly, M., Marrelec, G., Boveroux, P., Vanhaudenhuyse, A., Bruno, M.A., Laureys, S., Phillips, C., Pelegrini-Issac, M., Maquet,
P., and Benali, H. (2011). Brain functional integration decreases during
propofol-induced loss of consciousness. NeuroImage, in press. 10.1016/j.
neuroimage.2011.04.020.
Quiroga, R.Q., Mukamel, R., Isham, E.A., Malach, R., and Fried, I. (2008).
Human single-neuron responses at the threshold of conscious recognition.
Proc. Natl. Acad. Sci. USA 105, 35993604.
Railo, H., and Koivisto, M. (2009). The electrophysiological correlates of stimulus visibility and metacontrast masking. Conscious. Cogn. 18, 794803.
Ray, S., and Maunsell, J.H. (2010). Differences in gamma frequencies across
visual cortex restrict their possible use in computation. Neuron 67, 885896.
Raymond, J.E., Shapiro, K.L., and Arnell, K.M. (1992). Temporary suppression
of visual processing in an RSVP task: An attentional blink? J. Exp. Psychol.
Hum. Percept. Perform. 18, 849860.
Reuter, F., Del Cul, A., Audoin, B., Malikova, I., Naccache, L., Ranjeva, J.P.,
Lyon-Caen, O., Ali Cherif, A., Cohen, L., Dehaene, S., and Pelletier, J.
(2007). Intact subliminal processing and delayed conscious access in multiple
sclerosis. Neuropsychologia 45, 26832691.
Reuter, F., Del Cul, A., Malikova, I., Naccache, L., Confort-Gouny, S., Cohen,
L., Cherif, A.A., Cozzone, P.J., Pelletier, J., Ranjeva, J.P., et al. (2009). White
matter damage impairs access to consciousness in multiple sclerosis. Neuroimage 44, 590599.
Schurger, A., and Sher, S. (2008). Awareness, loss aversion, and post-decision
wagering. Trends Cogn. Sci. (Regul. Ed.) 12, 209210, author reply 210.
Schurger, A., Cowey, A., and Tallon-Baudry, C. (2006). Induced gamma-band
oscillations correlate with awareness in hemianopic patient GY. Neuropsychologia 44, 17961803.
Schurger, A., Pereira, F., Treisman, A., and Cohen, J.D. (2010). Reproducibility
distinguishes conscious from nonconscious neural representations. Science
327, 9799.
Semendeferi, K., Lu, A., Schenker, N., and Damasio, H. (2002). Humans and
great apes share a large frontal cortex. Nat. Neurosci. 5, 272276.
Sergent, C., and Dehaene, S. (2004). Is consciousness a gradual phenomenon? Evidence for an all-or-none bifurcation during the attentional blink.
Psychol. Sci. 15, 720728.
Sergent, C., Baillet, S., and Dehaene, S. (2005). Timing of the brain events
underlying access to consciousness during the attentional blink. Nat. Neurosci.
8, 13911400.
Neuron
Review
Seth, A.K. (2007). Models of consciousness. Scholarpedia 2, 1328.
Shallice, T. (1972). Dual functions of consciousness. Psychol. Rev. 79,
383393.
confined to areas in the occipital cortex beyond human V1/V2. Proc. Natl.
Acad. Sci. USA 102, 1717817183.
Tshibanda, L., Vanhaudenhuyse, A., Galanaud, D., Boly, M., Laureys, S., and
Puybasset, L. (2009). Magnetic resonance spectroscopy and diffusion tensor
imaging in coma survivors: Promises and pitfalls. Prog. Brain Res. 177,
215229.
Sheinberg, D.L., and Logothetis, N.K. (1997). The role of temporal cortical
areas in perceptual organization. Proc. Natl. Acad. Sci. USA 94, 34083413.
Sigman, M., and Dehaene, S. (2008). Brain mechanisms of serial and parallel
processing during dual-task performance. J. Neurosci. 28, 75857598.
Sigman, M., Pan, H., Yang, Y., Stern, E., Silbersweig, D., and Gilbert, C.D.
(2005). Top-down reorganization of activity in the visual pathway after learning
a shape identification task. Neuron 46, 823835.
Simons, D.J., and Ambinder, M.S. (2005). Change blindness: Theory and
consequences. Curr. Dir. Psychol. Sci. 14, 4448.
Slachevsky, A., Pillon, B., Fourneret, P., Pradat-Diehl, P., Jeannerod, M., and
Dubois, B. (2001). Preserved adjustment but impaired awareness in a sensorymotor conflict following prefrontal lesions. J. Cogn. Neurosci. 13, 332340.
Slagter, H.A., Johnstone, T., Beets, I.A., and Davidson, R.J. (2010). Neural
competition for conscious representation across time: An fMRI study. PLoS
ONE 5, e10556.
Smallwood, J., Beach, E., Schooler, J.W., and Handy, T.C. (2008). Going
AWOL in the brain: Mind wandering reduces cortical analysis of external
events. J. Cogn. Neurosci. 20, 458469.
Stephan, K.E., Friston, K.J., and Frith, C.D. (2009). Dysconnection in schizophrenia: From abnormal synaptic plasticity to failures of self-monitoring.
Schizophr. Bull. 35, 509527.
Sterzer, P., Haynes, J.D., and Rees, G. (2008). Fine-scale activity patterns in
high-level visual areas encode the category of invisible objects. J. Vis. 8,
10.110.12.
Sukhotinsky, I., Zalkind, V., Lu, J., Hopkins, D.A., Saper, C.B., and Devor, M.
(2007). Neural pathways associated with loss of consciousness caused by
intracerebral microinjection of GABA A-active anesthetics. Eur. J. Neurosci.
25, 14171436.
Supe`r, H., Spekreijse, H., and Lamme, V.A. (2001). Two distinct modes of
sensory processing observed in monkey primary visual cortex (V1). Nat.
Neurosci. 4, 304310.
Supe`r, H., van der Togt, C., Spekreijse, H., and Lamme, V.A. (2003). Internal
state of monkey primary visual cortex (V1) predicts figure-ground perception.
J. Neurosci. 23, 34073414.
Taine, H. (1870). De lintelligence (Paris: Hachette).
Terrace, H.S., and Son, L.K. (2009). Comparative metacognition. Curr. Opin.
Neurobiol. 19, 6774.
Thiebaut de Schotten, M., Urbanski, M., Duffau, H., Volle, E., Levy, R., Dubois,
B., and Bartolomeo, P. (2005). Direct evidence for a parietal-frontal pathway
subserving spatial awareness in humans. Science 309, 22262228.
Thirion, B., Duchesnay, E., Hubbard, E., Dubois, J., Poline, J.B., Lebihan, D.,
and Dehaene, S. (2006). Inverse retinotopy: Inferring the visual content of
images from brain activation patterns. Neuroimage 33, 11041116.
Uhlhaas, P.J., Linden, D.E., Singer, W., Haenschel, C., Lindner, M., Maurer, K.,
and Rodriguez, E. (2006). Dysfunctional long-range coordination of neural
activity during Gestalt perception in schizophrenia. J. Neurosci. 26, 8168
8175.
Urbanski, M., Thiebaut de Schotten, M., Rodrigo, S., Catani, M., Oppenheim,
C., Touze, E., Chokron, S., Meder, J.F., Levy, R., Dubois, B., and Bartolomeo,
P. (2008). Brain networks of spatial awareness: Evidence from diffusion tensor
imaging tractography. J. Neurol. Neurosurg. Psychiatry 79, 598601.
van Aalderen-Smeets, S.I., Oostenveld, R., and Schwarzbach, J. (2006). Investigating neurophysiological correlates of metacontrast masking with magnetoencephalography. Adv. Cogn. Psychol. 2, 2135.
Van de Ville, D., Britz, J., and Michel, C.M. (2010). EEG microstate sequences
in healthy humans at rest reveal scale-free dynamics. Proc. Natl. Acad. Sci.
USA 107, 1817918184.
Van den Bussche, E., Segers, G., and Reynvoet, B. (2008). Conscious and
unconscious proportion effects in masked priming. Conscious. Cogn. 17,
13451358.
Van den Bussche, E., Notebaert, K., and Reynvoet, B. (2009a). Masked primes
can be genuinely semantically processed: A picture prime study. Exp. Psychol.
56, 295300.
Van den Bussche, E., Van den Noortgate, W., and Reynvoet, B. (2009b). Mechanisms of masked priming: A meta-analysis. Psychol. Bull. 135, 452477.
van der Stelt, O., Frye, J., Lieberman, J.A., and Belger, A. (2004). Impaired P3
generation reflects high-level and progressive neurocognitive dysfunction in
schizophrenia. Arch. Gen. Psychiatry 61, 237248.
van Gaal, S., Ridderinkhof, K.R., Fahrenfort, J.J., Scholte, H.S., and Lamme,
V.A. (2008). Frontal cortex mediates unconsciously triggered inhibitory control.
J. Neurosci. 28, 80538062.
van Gaal, S., Lamme, V.A., and Ridderinkhof, K.R. (2010). Unconsciously triggered conflict adaptation. PLoS ONE 5, e11508.
van Gaal, S., Lamme, V.A., Fahrenfort, J.J., and Ridderinkhof, K.R. (2011).
Dissociable brain mechanisms underlying the conscious and unconscious
control of behavior. J. Cogn. Neurosci. 23, 91105.
Varela, F., Lachaux, J.P., Rodriguez, E., and Martinerie, J. (2001). The brainweb: Phase synchronization and large-scale integration. Nat. Rev. Neurosci.
2, 229239.
Thompson, K.G., and Schall, J.D. (1999). The detection of visual signals by
macaque frontal eye field during masking. Nat. Neurosci. 2, 283288.
Velly, L.J., Rey, M.F., Bruder, N.J., Gouvitsos, F.A., Witjas, T., Regis, J.M.,
Peragut, J.C., and Gouin, F.M. (2007). Differential dynamic of action on cortical
and subcortical structures of anesthetic agents during induction of anesthesia.
Anesthesiology 107, 202212.
Thompson, K.G., and Schall, J.D. (2000). Antecedents and correlates of visual
detection and awareness in macaque prefrontal cortex. Vision Res. 40, 1523
1538.
Veselis, R.A., Feshchenko, V.A., Reinsel, R.A., Dnistrian, A.M., Beattie, B., and
Akhurst, T.J. (2004). Thiopental and propofol affect different regions of the
brain at similar pharmacologic effects. Anesth. Analg. 99, 399408.
Vincent, J.L., Patel, G.H., Fox, M.D., Snyder, A.Z., Baker, J.T., Van Essen,
D.C., Zempel, J.M., Snyder, L.H., Corbetta, M., and Raichle, M.E. (2007).
Intrinsic functional architecture in the anaesthetized monkey brain. Nature
447, 8386.
Vincent, J.L., Kahn, I., Snyder, A.Z., Raichle, M.E., and Buckner, R.L. (2008).
Evidence for a frontoparietal control system revealed by intrinsic functional
connectivity. J. Neurophysiol. 100, 33283342.
Neuron
Review
Vogel, E.K., and Machizawa, M.G. (2004). Neural activity predicts individual
differences in visual working memory capacity. Nature 428, 748751.
Vogel, E.K., Luck, S.J., and Shapiro, K.L. (1998). Electrophysiological evidence
for a postperceptual locus of suppression during the attentional blink. J. Exp.
Psychol. Hum. Percept. Perform. 24, 16561674.
Vogt, B.A., and Laureys, S. (2005). Posterior cingulate, precuneal and retrosplenial cortices: Cytology and components of the neural network correlates
of consciousness. Prog. Brain Res. 150, 205217.
Von Economo, C. (1929). The Cytoarchitectonics of the Human Cerebral
Cortex (London: Oxford University Press).
von Holst, E., and Mittelstaedt, H. (1950). Das Reafferenzprinzip. Naturwissenschaften 37, 464476.
Voss, H.U., Uluc, A.M., Dyke, J.P., Watts, R., Kobylarz, E.J., McCandliss, B.D.,
Heier, L.A., Beattie, B.J., Hamacher, K.A., Vallabhajosula, S., et al. (2006).
Possible axonal regrowth in late recovery from the minimally conscious state.
J. Clin. Invest. 116, 20052011.
Voytek, B., and Knight, R.T. (2010). Prefrontal cortex and basal ganglia contributions to visual working memory. Proc. Natl. Acad. Sci. USA 107, 18167
18172.
Vul, E., and MacLeod, D.I. (2006). Contingent aftereffects distinguish
conscious and preconscious color processing. Nat. Neurosci. 9, 873874.
Wegner, D.M. (2003). The Illusion of Conscious Will (Cambridge: MIT Press).
Welford, A.T. (1952). The psychological refractory period and the timing of
high speed performanceA review and a theory. Br. J. Psychol. 43, 219.
Williams, M.A., Visser, T.A., Cunnington, R., and Mattingley, J.B. (2008). Attenuation of neural responses in primary visual cortex during the attentional blink.
J. Neurosci. 28, 98909894.
Womelsdorf, T., Fries, P., Mitra, P.P., and Desimone, R. (2006). Gamma-band
synchronization in visual cortex predicts speed of change detection. Nature
439, 733736.
Wong, K.F.E. (2002). The Relationship Between Attentional Blink and Psychological Refractory Period. J. Exp. Psychol. Hum. Percept. Perform. 28, 5471.
Wong, K.F., and Wang, X.J. (2006). A recurrent network mechanism of time
integration in perceptual decisions. J. Neurosci. 26, 13141328.
Woodman, G.F., and Luck, S.J. (2003). Dissociations among attention,
perception, and awareness during object-substitution masking. Psychol.
Sci. 14, 605611.
Wunderlich, K., Schneider, K.A., and Kastner, S. (2005). Neural correlates of
binocular rivalry in the human lateral geniculate nucleus. Nat. Neurosci. 8,
15951602.
Wyart, V., and Tallon-Baudry, C. (2008). Neural dissociation between visual
awareness and spatial attention. J. Neurosci. 28, 26672679.
Wyart, V., and Tallon-Baudry, C. (2009). How ongoing fluctuations in human
visual cortex predict perceptual awareness: Baseline shift versus decision
bias. J. Neurosci. 29, 87158725.
Wilke, M., Logothetis, N.K., and Leopold, D.A. (2003). Generalized flash
suppression of salient visual targets. Neuron 39, 10431052.
Zylberberg, A., Dehaene, S., Mindlin, G.B., and Sigman, M. (2009). Neurophysiological bases of exponential sensory decay and top-down memory retrieval:
A model. Front Comput Neurosci 3, 4.
Wilke, M., Logothetis, N.K., and Leopold, D.A. (2006). Local field potential
reflects perceptual suppression in monkey visual cortex. Proc. Natl. Acad.
Sci. USA 103, 1750717512.
Zylberberg, A., Fernandez Slezak, D., Roelfsema, P.R., Dehaene, S., and
Sigman, M. (2010). The brains router: A cortical network model of serial processing in the primate brain. PLoS Comput. Biol. 6, e1000765.
http://www.scienticamerican.com/article/consciousness-migh...
ADVERTISEMENT
ADVERTISEMENT
experiences. [Full disclosure: I have worked with Tononi on this theory.] In contrast,
the Global Workspace Model of consciousness moves in the opposite direction. Its starting point is behavioral experiments that
manipulate conscious experience of people in a very controlled setting. It then seeks to identify the areas of the brain that underlie these
experiences.
Stanislas Dehaene, the French cognitive neuroscientist at the Collge de France in Paris who has devoted much of his career to studying
the psychology of consciousness, has just published a compelling book on his investigations into how the Global Workspace Model
maps onto the brain.
The model derives from the realization that whenever we become conscious of somethingwhether a familiar face in a crowd or the
voice of a strangerwe can retain what we perceive in our mind for a brief period. This perception can remain in this short-term
memory storage, a kind of mental scratch pad, even after the face has disappeared or the voice has died away. Cognitive scientist
Bernard Baars of the Neurosciences Institute in La Jolla, Calif., who came up with the Global Workspace Model, took his central insight
from the early days of artificial intelligence, in which specialized programs accessed a shared repository of information, the blackboard.
According to Baars, it is the act of broadcasting data from the blackboard throughout a computational system, whether cybernetic or
biological, that makes it conscious. Consciousness is just brain-wide sharing of information that is in the memory buffer of the
blackboard.
1 of 4
29/07/2015 11:05 am
http://www.scienticamerican.com/article/consciousness-migh...
This neural buffer does more than process recent sensory inputs. It can also call up a memory from long ago and move it into the buffer.
Once information is loaded into this workspace, a host of powerful cognitive processes can make use of it. The data can be sent off to a
particular brain area that processes languagea language modulewhere this knowledge can be readied for sharing with other people
by formulating a spoken explanation: Guess who I just saw over there. It can also be forwarded to a planning module to be reasoned
about, and it can be stored in long-term memory. The act of transmitting these data from the brain's memory buffers to its various
functional modules is what gives rise to consciousness.
Unfortunately, this workspace has extremely limited capacity. At any one time, we can be conscious of only one or a few items or events,
although we can quickly shift things into and out of consciousness. New information competes with the old and may ultimately
overwrite it. This limitation probably is an unavoidable design characteristic of any information-processing system that is overwhelmed
by inflowing data streams and has to concentrate its most precious resources on dealing with a couple of critical items as fast as
possible.
The brain compensates for the dearth of neural bandwidth by calling on a host of unconscious processes that either totally bypass this
central scratch pad or interact with it below the level of awareness. The vast subliminal onslaught of data thereby turns sounds into
meaningful words and photons into objects and identifiable people. These processes evaluate and weigh evidence, pass judgment and
synchronize the movements initiated by the musculoskeletal system so that an organism can survive in a constantly and rapidly
changing world. They are sophisticated and act quickly but do not share information with one another, nor do they transfer it into the
common workspace. As with an intelligence agency, information is shared only on a need-to-know basis.
Yet these myriad agents of the unconscious shape our daily routines. Because we have, by definition, no access to these subliminal
events, we consistently underestimate their importance. Yet occasionally they manifest themselves quite dramatically. Japanese novelist
Haruki Murakami put it well in a striking interview: We have rooms in ourselves. Most of them we have not visited yet. Forgotten
rooms. From time to time we can find the passage. We find strange things ... old phonographs, pictures, books ... they belong to us, but
it is the first time we have found them.
Dehaene probes these unconscious lairs using a technique called masking. A picture, say, of a face or a word is briefly flashed onto a
monitor, preceded and followed by images of a bunch of randomly drawn lines or a cloud of X's. These masks prevent the displayed
face or word from becoming consciousa subject reports seeing only a mask. Combining versions of this technique with recordings
from electrodes implanted deep into the brain of patients monitored for epilepsy seizures, Dehaene and his colleagues demonstrated
that the unconscious can process the meaning of word combinationsthe brain responds differently to happy war than to happy
loveimplying that it has noticed the incongruence of having a word with a positive emotional meaning followed by a word with a
negative one.
Dehaene and the distinguished molecular biologist Jean-Pierre Changeux have gone beyond this rather abstract model and are
searching for the specific brain areas and populations of neurons that correspond to the global workspace. Their ongoing research using
functional brain imaging and electroencephalographic electrodes placed on the skull has uncovered distinct neural signatures in these
regions that appear to represent the theorized mental buffer.
SEE ALSO:
Energy & Sustainability: 5 Steps to Feed the World and Sustain the Planet | Evolution: Clues to How Homo sapiens Conquered the Earth
Emerge from Digs in South Africa [Slide Show] | Health: The Conflicted History of Alcohol in Western Civilization | Space: Europa's
"Brown Gunk" Suggests a Briny Sea | Technology: Timeline: The Amazing Multimillion-Year History of Processed Food | More
Science: The Flavor Connection
In one classic experiment, Dehaene and his colleagues had volunteers lie inside a magnetic resonance imaging scanner while they
watched a stream of words on a computer screen, each one displayed for 29 milliseconds. Some of the words were masked, which
triggered only a slight brain response. But when the words were legible, an avalanche of neural activity occurred.
The activated regions make up a dense tapestry of interlocking brain cellsspecifically pyramidal neuronsthat tie together the
prefrontal cortex, the inferior parietal lobe, the middle and anterior temporal lobes and other brain regions. Axons, the wirelike
extensions from a neuron's cell body, fan out from the brain's fissured surface, the cerebral cortex, to bind together vast reaches of
neural topography. This network is where Dehaene and his colleagues have started to look both for the brain's scratch pad and for how
signals streaming through this web of connections are communicated to the rest of the brain.
Whenever a stimulus is consciously perceived, its neuronal footprinta particular type of brain activityshows up in many parts of the
cerebral cortex. Take, for instance, the intense electrical activity triggered by an image that passes into the primary visual cortex at the
back of the head and from there to many cortical regions. As it reaches anterior regions of cortex, the signals increase in amplitude,
prompting Dehaene to call it a neuronal avalanche.
2 of 4
29/07/2015 11:05 am
http://www.scienticamerican.com/article/consciousness-migh...
The intense neuronal firing can be caught in the act with EEG electrodes by measuring the P300 wave, a brain wave that, in
experiments, starts about 300 milliseconds after an image is projected onto a computer screen. As Dehaene's experiments demonstrate,
becoming conscious of a sight or sound by having it broadcast throughout the brain from areas postulated to make up the global
workspace often goes hand in hand with the presence of a P300 wave in the prefrontal cortex, a brain area associated with higher
mental processes. Conversely, without the signature P300 wave, electrical activity dies out, and the image displayed is not consciously
perceived. The information fails to enter the global workspace and so remains subliminal.
First Glimmers
Dehaene and his colleagues used this electrophysiological marker of conscious perception to map when consciousness first arises in
five- to 15-month-old infants [see The Conscious Infant; Scientific American Mind, September/October 2013] and to devise a clever
test for consciousness in severely brain-injured patients with whom no reliable communication using speech, eyes or gestures is
possible. The tests depend on the ability of a conscious individual to detect a novel stimulusimagine reading a book when your cell
phone abruptly rings. This unexpected event can trigger a massive P300 wave that is readily noticeable. Yet when you do not pick up the
phone and it rings again and again, you come to expect it, and the P300 becomes fainter until it cannot be detected.
In the laboratory, the researchers play a sequence of five simple tones: beep beep beep beep boop. The last odd-man-out tone generates
a strong P300. When the entire sequence of five tones is repeated three times, the brain adapts to the deviant sound, and the
consciousness marker disappears.
Then, along comes a beep beep beep beep beep sequence. As an attentive subject becomes conscious of the lack of a deviating sound in
the fourth sequence, her brain responds with a P300 to the final beep because it was conditioned to expect a boop.
Preliminary trials using this test with brain-injured patients are intriguing. Patients in whom behavioral evidence indicates a minimal
level of consciousness show this pattern of P300 activity on their EEGs, whereas those in a coma, thought to be without any sensation
whatsoever, do not. Ongoing experiments seek to exploit the same odd-man-out paradigm in monkeys and in mice.
Proposing that what we consciously experience can be defined as the brain's ability to distribute information from the global workspace
to the rest of the brain brings up several questions. Why and how, for instance, does broadcasting information from the global
workspace give rise to consciousness? What message is being broadcast? Blood-borne hormones and chemicals that regulate neural
activity also relay information throughout the body and brain. Yet we are not aware of them. Why not? And can data transmitted over
the Internet or information coursing through the nervous system of a roundworm represent conscious activity? For now the Global
Workspace Model avoids such thorny questions.
When the molecular-biologist-turned-neuroscientist Francis Crick and I started our joint work in the late 1980s on trying to understand
the brain activity underlying vision and other mental processes, scant experimental work was dedicated to empirical studies of the
hallmarks of consciousness.
As the work by Dehaene, Changeux and their colleagues makes abundantly clear, this sorry situation has changed radically. Their
research program is beginning to untangle how the firing of networks of brain cells translates into this most mysterious of all
phenomena.
Buy this digital issue or subscribe to access other articles from the May 2014 publication.
Already have an account? Sign In
3 of 4
Digital Issue
$5.99
Digital Subscription
$19.99
Add To Cart
Subscribe
29/07/2015 11:05 am
http://www.scienticamerican.com/article/consciousness-migh...
The Problem with Female Superheroes a month ago scienticamerican.com ScienticAmerican.com Mind & Brain
2.
Your Facial Bone Structure Has a Big Inuence on How People See You a month ago scienticamerican.com ScienticAmerican.com
Everyday Science
What Kind of Introvert Are You? 9 months ago blogs.scienticamerican.com ScienticAmerican.com Jennifer Odessa Grimes
Subscribe Now
4 of 4
29/07/2015 11:05 am
REVIEWS
Consciousness and Anesthesia
Michael T. Alkire,1 Anthony G. Hudetz,2 Giulio Tononi3*
When we are anesthetized, we expect consciousness to vanish. But does it always? Although
anesthesia undoubtedly induces unresponsiveness and amnesia, the extent to which it causes
unconsciousness is harder to establish. For instance, certain anesthetics act on areas of the brains
cortex near the midline and abolish behavioral responsiveness, but not necessarily consciousness.
Unconsciousness is likely to ensue when a complex of brain regions in the posterior parietal area is
inactivated. Consciousness vanishes when anesthetics produce functional disconnection in this
posterior complex, interrupting cortical communication and causing a loss of integration; or when
they lead to bistable, stereotypic responses, causing a loss of information capacity. Thus, anesthetics
seem to cause unconsciousness when they block the brains ability to integrate information.
ow consciousness arises in the brain remains unknown. Yet, for nearly two centuries our ignorance has not hampered
the use of general anesthesia for routinely extinguishing consciousness during surgery. Unfortunately, once in every 1000 to 2000 operations
a patient may temporarily regain consciousness or
even remain conscious during surgery (1). Such
intraoperative awareness arises in part because our
ability to evaluate levels of consciousness remains
limited. Nevertheless, progress is being made in
identifying general principles that underlie how
anesthetics bring about unconsciousness (26) and
how, occasionally, they may fail to do so.
convenient, has drawbacks. For instance, unresponsiveness can occur without unconsciousness.
When we dream, we have vivid conscious experiences, but are unresponsive because inhibition
by the brainstem induces muscle paralysis (13).
Similarly, paralyzing agents used to prevent unwanted movements during anesthesia do not remove consciousness (14).
Certain anesthetics may impair a persons
willfulness to respond by affecting brain regions
where executive decisions are made. This is not
an issue for anesthetics that globally deactivate
the brain, but it may be problematic for dissociative anesthetics like ketamine. Low doses of
ketamine cause depersonalization, out-of-body experiences, forgetfulness, and loss of motivation to
follow commands (15). At higher doses, ketamine
causes a characteristic state in which the eyes are
open and the face takes on a disconnected blank
stare. Neuroimaging data show a complex pattern
of regional metabolic changes (16), including a
deactivation of executive circuits in anterior
cingulate cortex and basal ganglia (Fig. 1) (17).
A similar open-eyed unresponsiveness is seen in
akinetic mutism after bilateral lesions around the
anterior cingulate cortex (18). In at least some
of these cases, patients understand questions, but
may fail to respond. Indeed, a woman with large
frontal lesions who was clinically unresponsive
was asked to imagine playing tennis or to navigate her room, and she showed cortical activation
patterns indistinguishable from those of healthy
subjects (19). Thus, clinical unresponsiveness is not
necessarily synonymous with unconsciousness.
At doses near the unconsciousness threshold,
some anesthetics block working memory (20).
Thus, patients may fail to respond because they
immediately forget what to do. At much lower
doses, anesthetics cause profound amnesia. Studies
with the isolated forearm technique, in which a
tourniquet is applied to the arm before paralysis is
Potassium channels
Nicotinic Muscarinic
Two Inwardly Voltage
Serotonin AMPA Kainate
GABAA NMDA pore rectifying gated Glycine Ach
Ach
Intravenous
anesthestics
Barbiturates
Propofol
Etomidate
Ketamine
Inhalational
anesthestics
Nitrous oxide
Isoflurane
Sevoflurane
Department of Anesthesiology and the Center for the Neurobiology of Learning and Memory, University of California,
Irvine, CA 92868, USA. 2Department of Anesthesiology, Medical College of Wisconsin, Milwaukee, WI 53226, USA.
3
Department of Psychiatry, University of Wisconsin, Madison,
WI 53719, USA.
*To whom correspondence should be addressed. E-mail:
[email protected]
876
Desflurane
Major potentiation
Minor potentiation
Major inhibition
Minor inhibition
Biphasic
No effect
Table 1. Ionic mechanisms and targets of current clinical anesthetics (6, 8). Abbreviations: Ach,
acetylcholine; AMPA, a-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; GABAA, g-aminobutyric
acid, type A; NMDA, N-methyl-D-aspartate.
7 NOVEMBER 2008
VOL 322
SCIENCE
www.sciencemag.org
REVIEWS
induced (to allow the hand to
Anterior
move while the rest of the body
cingulate
is paralyzed), show that patients
cortex
under general anesthesia can
sometimes carry on a conversation using hand signals, but postoperatively deny ever being awake
(21). Thus, retrospective oblivion
is no proof of unconsciousness.
Nevertheless, at some level of
anesthesia between behavioral unresponsiveness and the induction
of a flat EEG [indicating the cessation of the brains electrical activity, one of the criteria for brain
Fig. 1. Brain
death (22)], consciousness must
vanish. Therefore, the use of brain-function monitors could improve consciousness assessment
during anesthesia (23). For instance, bispectral
index monitors record the EEG signal over the
forehead and reduce the complex signal into a
single number that tracks a patients depth of
anesthesia over time (12). Such devices help guide
anesthetic delivery and may reduce cases of intraoperative awareness (24), but they remain limited at directly indicating the presence or absence
of consciousness, especially around the transition
point. The isolated forearm technique has shown
that individual patients can be aware and responsive during surgery even though their bispectral
index value suggests they are not (25). Either the
EEG is not sensitive enough to the neural processes underlying consciousness, or we still do
not yet fully understand what to look for.
The ThalamusSwitch or Readout?
The most consistent regional effect produced by
anesthetics at (or near) loss of consciousness is a
reduction of thalamic metabolism and blood flow
(Fig. 1), suggesting that the thalamus may serve as
a consciousness switch (2). Indeed, switchlike effects
have been found with a number of thalamic manipulations. For example, g-aminobutyric acid (GABA)
agonists (mimicking anesthetic action) injected into
the intralaminar nuclei cause rats to rapidly fall
asleep, with a corresponding slowing of the EEG (26).
Conversely, rats under anesthetic concentrations of
sevoflurane can be awakened by a minute injection
of nicotine into the intralaminar thalamus (27). In
humans, midline thalamic damage can result in a
vegetative state (18). Conversely, recovery from the
vegetative state is heralded by the restoration of functional connectivity between thalamus and cingulate
cortex (28). Also, deep brain electrical stimulation of
the central thalamus improved behavioral responsiveness in a patient who was minimally conscious (29).
Nevertheless, thalamic activity does not decrease with all anesthetics. Ketamine increases
global metabolism, especially in the thalamus
(16). Other anesthetics can substantially reduce
thalamic activity at doses that cause sedation, not
unconsciousness. For instance, sevoflurane sedation causes a 23% reduction of relative thalamic
metabolism when subjects are still awake and
responsive (30). Indeed, anesthetic effects on the
www.sciencemag.org
Mesial parietal
cortex, precuneus
SCIENCE
VOL 322
7 NOVEMBER 2008
877
REVIEWS
(56). Also, anesthesia suppresses the late component (>100 ms) of visual responses, possibly by
inhibiting feedback connections (57), but not the
early feedforward components. Moreover, anesthesia abolishes contextual and attentional modulation of firing, presumably mediated by feedback
connections (58). The corticothalamic system may
be especially vulnerable to anesthetics due to its
small-world organization. Small-world networks
have mostly local connectivity with comparatively
few long-range connections. Augmented with hubs,
such networks maximize interactions while minimizing wiring. By the same token, anesthetics need
only disrupt a few long-range connections to produce a set of disconnected components. Indeed,
computer simulations demonstrate a rapid state
transition at a critical anesthetic dose (59), consistent with a breakdown in network integration.
Integrated information
activity patterns. When the repertoire of discriminable firing patterns available to the corticothalamic
system shrinks, neural activity becomes less informative, even though it may be globally integrated
(52). As described above, at high enough doses
several anesthetics produce a burst-suppression
pattern in which a near-flat EEG is interrupted
every few seconds by brief, quasi-periodic bursts
of global activationa stereotypic, global on-off
pattern. Such stereotypic burst-suppression can also
be elicited by visual, auditory, and mechanical
stimuli (Fig. 3B) (60, 61). Thus, during deep anesthetic unconsciousness, the corticothalamic system can still be activein fact, hyperexcitable
and can produce global responses. However, the
repertoire of responses has shrunk to a stereotypic burst-suppression pattern, with a corresponding
loss of information, essentially creating a system having only two possible states (on or off).
Generalized convulsive seizures provide another
example in which consciousness can be lost even
though neural activity remains high and highly
Human
Premotor cortex
Awake
12 V
TMS
50 ms
Rat
A/mm2
0.01
Anesthetized
Awake
1.1% Isoflurane
Fr
Fr
15 ms
Oc
Par
250 ms
0.002
Occipital (Oc)
12 V
Parietal (Par)
50 ms
0.20
TMS
0.15
0.01
0.10
A/mm2
150 ms
0.25
0.05
0.00
-200
-100
100
200
300 -200
-100
100
200
300
Time (ms)
15 ms
878
100 ms
Asleep
Par
Frontal (Fr)
50 ms
Oc
7 NOVEMBER 2008
VOL 322
50 ms
100 ms
150 ms
250 ms
0.002
anesthesia (56). When the rat is awake, each flash evokes a sustained gfrequency (20 to 60 Hz) response in visual occipital cortex (blue) and a later
response in parietal association cortex (red). During anesthesia, the occipital
response is preserved, although it is shorter (blue), and the parietal response
is attenuated, indicating that anesthesia reduces cortical interactions and thus
reduces integration. (C) Sleeping reduces cortical integration in humans. EEG
voltages and current densities are shown from a representative subject in
which the premotor cortex was stimulated with transcranial magnetic stimulation (TMS) (black arrow). During waking (top), stimulation evokes EEG responses first near the stimulation site (black circle; the white cross is the site of
maximum evoked current) and then in sequence at other cortical locations.
During deep sleep (bottom), the stimulus-evoked response remains local,
indicating a loss of cortical integration.
SCIENCE
www.sciencemag.org
REVIEWS
(TMS) applied to premotor cortex and other cortical areas induces a sustained response (300 ms)
involving the sequential activation of specific brain
areas, the identity of which depends upon the precise site of stimulation (63, 64). During early
NREM sleep, possibly due to the induction of a
local down-state, TMS pulses produce instead a
short (<150 ms) local response (64), suggesting
a loss of integration. Intriguingly, TMS pulses to
mesial parietal regions, overlying the main hub in
the cortical connectional core (49), trigger a stereotypic, high-amplitude slow wave closely resembling spontaneous slow waves (63). This stereotypic
response, presumably due to the simultaneous activation of the cortical connectional core and to the
induction of a global down-state, reflects a limited
repertoire of activity patterns and thus a loss of
information.
synchronized: A large portion of the corticothalamic complex is engaged in strong, hypersynchronous activity, but this activity is stereotypic
(60, 61).
A Bit Like Sleep
Sleep is the only time when healthy humans regularly lose consciousness. Subjects awakened during slow wave sleep early in the night may report
short, thoughtlike fragments of experience, or often
nothing at all (13). Although anesthesia is not the
same as natural sleep, brain-arousal systems are
similarly deactivated (6, 62). Also, as under anesthesia, during slow wave sleep, cortical and
thalamic neurons become bistable and undergo
slow oscillations (1 Hz or less) between up- and
down-states. Like animal studies during anesthesia (Fig. 2B and 3B), human studies during slow
wave sleep suggest that the bistability of cortical
neurons has consequences for the brains capacity
to integrate information (Figs. 2C and 3C). During wakefulness, transcranial magnetic stimulation
C Human
Loss of
information capacity
Integrated information
Awake
20 V
50 ms
TMS
B Rat
A/mm2
0.01
Anesthetized
Awake
1.8% Isoflurane
5 V
Fr
0.002
Fr
Par
15 ms
Par
Oc
50 ms
100 ms
150 ms
250 ms
Asleep
20 V
Oc
25s
Occipital (Oc)
50 ms
TMS
Parietal (Par)
0.15
0.01
0.10
A/mm2
Frontal (Fr)
0.20
0.05
0.00
-200
200
400
-200
200
400
0.002
Time (ms)
15 ms
SCIENCE
50 ms
100 ms
150 ms
250 ms
7 NOVEMBER 2008
879
REVIEWS
experience the left half of the visual field independently of the right half, or visual shapes independently of their color. In other words, the die
of experience is a single onethrowing multiple
dice and combining the numbers will not do.
Less metaphorically, the theory claims that
the level of consciousness of a physical system is
related to the repertoire of different states (information) that can be discriminated by the system
as a whole (integration). A measure of integrated
information, called phi (F), can be used to quantify the information generated when a system enters one particular state of its repertoire, above
and beyond the information generated independently by its parts (52, 65). In practice, F can
only be measured rigorously for small, simulated
systems. However, empirical measures could be
devised to evaluate integrated information on the
basis of EEG data, resting functional connectivity,
or TMS-evoked responses. This approach could
allow the development of consciousness monitors
that evaluate both loss of integration, as revealed
by reduced functional or effective connectivity, and
loss of information, as evidenced by stereotypic
responses.
This theory has some interesting implications
for anesthesia. For example, it explains why a corticothalamic complex is essential for consciousness and is thus the proper target for anesthesia:
By conjoining functional specialization (each cortical area and neuronal group within each area is
exquisitely specialized) with functional integration
(thanks to extensive corticocortical and corticothalamocortical connectivity), a corticothalamic
complex is well suited to behave as a single dynamic entity endowed with a large number of discriminable states. By contrast, parts of the brain
made up of small, quasi-independent modules,
such as the cerebellum, and parallel loops through
the basal ganglia, are not sufficiently integrated,
which is perhaps why they can be lesioned without
loss of consciousness (18, 52). The theory suggests
that one should not interpret individual motor responses, or localized activations, as signs of consciousness, and conversely, should not interpret
the absence of motor responses as a sure sign of
unconsciousness. Finally, from this theoretical
perspective, consciousness is not an all-or-none
property, but it is graded: Specifically, it increases
in proportion to a systems repertoire of discriminable states. The shrinking or dimming of the
field of consciousness during sedation is consistent with this idea. On the other hand, the
abrupt loss of consciousness at a critical concentration of anesthetics suggests that the integrated
repertoire of neural states underlying consciousness may collapse nonlinearly.
Conclusions
Despite different mechanisms and sites of action,
most anesthetic agents appear to cause unconsciousness by targeting, directly or indirectly, a
posterior lateral corticothalamic complex centered
around the inferior parietal lobe, and perhaps a
medial cortical core. Whether the medial or lateral
880
7 NOVEMBER 2008
VOL 322
SCIENCE
www.sciencemag.org
Coma Science Group, Cyclotron Research Center, University of Li`ege, Li`ege, Belgium
Spontaneous brain activity has recently received increasing interest in the neuroimaging community. However, the value of resting-state studies to a better understanding of brainbehavior
relationships has been challenged. That altered states of consciousness are a privileged way to
study the relationships between spontaneous brain activity and behavior is proposed, and common resting-state brain activity features observed in various states of altered consciousness are
reviewed. Early positron emission tomography studies showed that states of extremely low or high
brain activity are often associated with unconsciousness. However, this relationship is not absolute,
and the precise link between global brain metabolism and awareness remains yet difficult to assert.
In contrast, voxel-based analyses identified a systematic impairment of associative frontoparieto
cingulate areas in altered states of consciousness, such as sleep, anesthesia, coma, vegetative
state, epileptic loss of consciousness, and somnambulism. In parallel, recent functional magnetic
resonance imaging studies have identified structured patterns of slow neuronal oscillations in the
resting human brain. Similar coherent blood oxygen leveldependent (BOLD) systemwide patterns
can also be found, in particular in the default-mode network, in several states of unconsciousness,
such as coma, anesthesia, and slow-wave sleep. The latter results suggest that slow coherent spontaneous BOLD fluctuations cannot be exclusively a reflection of conscious mental activity, but may
reflect default brain connectivity shaping brain areas of most likely interactions in a way that
transcends levels of consciousness, and whose functional significance remains largely in the dark.
Key words: functional neuroimaging; resting state; disorders of consciousness; vegetative state
Introduction
In recent years, there has been a growing interest
from the neuroscientific community concerning spontaneous brain activity and its relation to cognition and
behavior. The concept of a default mode of brain
function arose from the need to explain consistent
brain-activity decreases in a set of areas during cognitive processing as compared to a passive resting baseline.1 These areas, encompassing the posterior cingulate cortex/precuneus, the medial prefrontal cortex,
and bilateral temporoparietal junctions, began to be
120
neuroscience. In contrast, the aims of cognitive neuroscience would be best served by the study of specific
task manipulations, rather than of rest.3
In response to this criticism, Raichle and Snyder5
argued that there is likely much more to brain function than that revealed by experiments manipulating
momentary demands of the environment. In their
view, a first argument in this direction is the cost of intrinsic brain activity, which far exceeds that of evoked
activity.6 Indeed, relative to the high rate of ongoing
or basal brain metabolism,6 the amount dedicated
to task-evoked regional imaging signals is remarkably
small (estimated to be less than 5%). The brain continuously expends a considerable amount of energy, even
in the absence of a particular task (i.e., when a subject is
awake and at rest). A significant fraction of the energy
consumed by the brain (quite possibly the majority) has
been shown to be a result of functionally significant
spontaneous neuronal activity.7 From this cost-based
analysis of brain functional activity, it seems reasonable
to conclude that intrinsic activity may be as significant,
if not more so, than evoked activity in terms of overall brain function.6 Another argument for the interest
of studying spontaneous brain activity is the striking
degree of functional organization exhibited by this intrinsic activity.5 The first clue of this organization is the
consistent activity decreases in default network during
cognitive tasks.1 Even more striking data recently arose
from fMRI blood oxygen leveldependent (BOLD)
connectivity studies in awake resting subjects, which
will be discussed later in this chapter. Maps of spontaneous network correlations have also been proposed
to provide tools for functional localization, for the understanding of clinical conditions such as Alzheimers
disease8,9 and autism,10 or for the study of comparative anatomy between primate species.11 However, the functional significance of the observed patterns of intrinsic brain activity remains actually poorly
understood.
We here propose that disorders of consciousness
are a privileged way to investigate the links between spontaneous brain activity and behavior. These
states are indeed mainly characterized by the alteration of intrinsic brain activity, which induces dramatic changes in the contents of awareness and
responses to environmental stimuli and demands.
We will illustrate our view, reviewing common features of spontaneous brain-activity patterns in altered states of consciousness, as shown by metabolic
positron emission tomography (PET) data as well
as recent BOLD fMRI studies. We will also discuss
methodological issues of resting-state neuroimaging
experiments.
Consciousness as a Multidimensional
Concept
Consciousness has two major components: awareness (i.e., the content of consciousness) and arousal (i.e.,
the level of consciousness).12 Arousal and awareness
are usually positively correlated: when your arousal
decreases, so does your awareness [rapid eye movement (REM) sleep being a notable exception]. Awareness can also be divided into two components: selfawareness and external awareness. Self- and external
awareness usually behave in an anti-correlated manner. When you are engaged in self-related processes,
you are less receptive to environmental demands, and
vice versa.13,14 A number of studies have compared
brain activation in circumstances that do or do not give
rise to consciousness in either of its two main senses of
awareness and arousal. Very few groups, however, have
studied situations in which wakefulness and arousal are
dissociated.
The vegetative state (VS) is a classic example of a
dissociated state of unconsciousness. VS patients are
fully aroused, but are unaware of themselves and their
environment. They can show automatic reactions like
moving their eyes, head, and limbs in a meaningless manner, and may even grimace, cry, or smile (albeit never contingently upon specific external stimuli).
Some patients might evolve toward full recovery or remain in the minimally conscious state,15 where some
nonreflexive or nonmeaningful behaviors are shown,
but patients are still unable to communicate. In addition to their clinical and ethical importance, the study
of vegetative and minimally conscious states offers a
still widely unexploited means of studying human consciousness.12 In contrast to other unconscious states,
such as general anesthesia and deep sleep, where impairment in arousal cannot be disentangled from impairment in awareness, these states represent a unique
lesional approach enabling us to identify the neural
correlates of (un)awareness.
121
perpolarized state of the system, is maximized in conditions of intermediate neural activity, and decreases
in states where the neural activity is extremely high
and near-synchronous.27 It has been proposed that a
proper balance between excitatory and inhibitory activity would be necessary to allow neurons to respond
appropriately to correlational changes in their input,
and to establish the functional connectivity as required
for a particular cognitive task or behavior.29 In the
same line, Raichle and Gusnard suggested that a large
part of the brains default activity could be devoted to
ongoing synaptic processes associated with the maintenance of this balance.30 In this view, global brain
metabolism would have a potentially decisive role by
allowing the presence of conscious perception or behavior.
The equivocal link between global brain activity and
consciousness is, however, further challenged by the
fact that in some patients who subsequently recovered
from a VS to normal consciousness, global metabolic
rates for glucose metabolism did not show substantial
changes.31 Moreover, some awake healthy volunteers
have global brain metabolism values comparable to
those observed in some patients in a VS.12 Inversely,
some well-documented vegetative patients have shown
close to normal global cortical metabolism.32 These
data led us to focus rather on regional metabolism
in our quest for a better understanding of the links
between consciousness and resting-brain activity.
Voxel-based statistical analyses have sought to identify regions showing metabolic dysfunction in VS patients as compared with the conscious resting state
in healthy controls. These studies have identified a
systematic metabolic dysfunction, not in one brain
region but in a wide frontoparietal network encompassing the polymodal associative cortices in
VS: lateral and medial frontal regions bilaterally, parietotemporal and posterior parietal areas bilaterally, posterior cingulated, and precuneal
cortices,12,21 known to be the most active by
default in resting nonstimulated conditions.33 In contrast, arousal structures (encompassing the pedunculopontine reticular formation, the hypothalamus, and
the basal forebrain) are relatively preserved in these
patients.19 The same frontoparietal functional impairment is found in various other states of unconsciousness, that is, in sleep,34 coma,35 general
anesthesia,36 generalized seizures,37 or in other dissociated unconscious states like absence seizures,38,39
complex partial seizures,40 or somnambulism.41
FIGURE 2 illustrates the involvement of the frontoparietal cortical network in awareness, while arousal rather
relies on subcortical structures.
122
FIGURE 2. (Left ) Consciousness has two main components: arousal, or the level of consciousness, and awareness,
corresponding to the contents of consciousness per se. Arousal and awareness are usually positively correlated. However,
they involve different brain structures. Arousal involves the activity of subcortical structures encompassing brain-stem reticular
formation, hypothalamus, and basal forebrain. Awareness is related to the activity of a widespread set of frontoparietal
associative areas, both on the convexity and on the midline. (Right ) Awareness can in turn be divided into two main
components: self and external awareness. In healthy volunteers, self- and external awareness are usually negatively
correlated. Similarly, the frontoparietal awareness network can in turn be divided into two sub-systems, involved in selfand external awareness. Self-awareness networks encompass the posterior cingulate/precuneal cortices, medial frontal
cortex, and bilateral temporoparietal junctions. The external awareness network encompasses lateral frontal and parietal
cortices. In healthy volunteers, self- and external awareness networks usually show an anticorrelated pattern of activity.
These findings emphasize the importance of frontoparietal association areas in consciousness, and are
in line with the global workspace theory as introduced
by Baars.42,43 This theory views the brain as a massive parallel set of specialized processors. Consciousness might be a gateway to brain integration, enabling
access between otherwise separate neuronal functions.
In such a system, coordination and control may take
place by way of a central information exchange, allowing some processorssuch as sensory systems in
the brainto distribute information to the system as
a whole. According to Baars,35 frontoparietal association areas would be an ideal candidate for being the
global workspace processor.
As reported below, awareness can in turn be divided
in two main components: self- and external awareness. In the same line, FIGURE 2 illustrates that frontoparietal network can be subdivided in areas involved
in external awareness, and in self-awareness. External
awareness network activity is crucial for conscious external stimuli perception, as documented in healthy
awake volunteers.13,44 Self-awareness network encompasses the so-called default network and has been
involved in various aspects of self-related processes.45
In awake healthy volunteers, self- and external awareness networks usually show an anticorrelated pattern
of activity. Anticorrelations between self- and external
awareness networks have indeed been observed during cognitive tasks,46 sensory perception,13 as well as
in studies of resting-state brain activity.45 In contrast, in
most states of altered consciousness, both the activity
of subnetworks is similarly impaired.
123
124
components correspond to noise or neural systems. Solutions to these practical problems have been proposed,
for example, the probabilistic ICA.59
So far, we have only considered fMRI recordings,
but the electroencephalogram (EEG) is more and more
routinely recorded alongside fMRI to study spontaneous brain activity.65 Importantly, EEG data provide
access to very useful information regarding the timing of spontaneous brain activity. Features can be detected in the EEG signal and used to build an activation regressor for fMRI. For example, during sleep
studies, typical waves (e.g., slow waves or spindles; or
epileptic spikes) are easily detected on the EEG trace
and a spontaneous hemodynamic event is associated
with each occurrence of such wave. The analysis of the
fMRI data can then proceed as usual in stimulus induced tasks.6668 EEG data can also be processed to
yield a continuous regressor associated with spontaneous brain dynamics. For example, correlation between the BOLD signal of each voxel and the EEG
power in a specific frequency band, convoluted with
the standard hemodynamic response function, provides a correlation map, similar to what is done with the
seed-region activity. Typically, the spectrogram, that is,
the power spectrum evolving over time, of some or all
EEG channels is calculated and resampled at the fMRI
acquisition frequency. Then the time course of power
within frequency bands of interest is used to build a
correlations map.69,70
125
FIGURE 3. Spontaneous anticorrelations between self- and external awareness networks in the conscious resting state, as observed in an individual volunteer. (Left ) Areas correlated (above) and anticorrelated (below ) with the blood oxygen leveldependent (BOLD) time course of a seed voxel located in
the posterior cingulate/precuneus. (Right ) Plot of the BOLD time courses of posterior cingulate/precuneus
(PCC, red/gray line) and of middle frontal gyrus (MFG, blue/dark line) in the same volunteer. As previously reported, anticorrelations between these area time courses occur in slow frequencies with a period
below 0.1 Hz. (In color in Annals online.)
Peltier et al.92 assessed the effect of sevoflurane anesthesia on the temporal BOLD correlations in activity
in the motor cortices of healthy humans. Across all
volunteers, they found that the number of significant
voxels in the functional connectivity maps was reduced
by 78% for light anesthesia and by 98% for deep anesthesia, compared with the awake state. Additionally,
significant correlations in the connectivity maps were
bilateral in the awake state, but unilateral in the light
anesthesia state. Interestingly, this loss of interhemispheric connectivity was also found in an independent
resting-state fMRI study on a minimally conscious patient compared to healthy volunteers.85
In contrast to these findings, recent data from several independent BOLD fMRIs suggest that lowfrequency systemwide BOLD coherent spontaneous
activity can be preserved in various states of unconsciousness. First, low-frequency BOLD fluctuations
have recently been investigated using ICA during light
sleep in humans.93,94 In this work (collapsing nonREM sleep stages 1 and 2), significant increases in the
fluctuation level of the BOLD signal were observed
in several cortical areas, among which visual cortex
was the most significant.93 Furthermore, correlations
among brain regions involved in the default network
(encompassing posterior cingulate/precuneus, medial
frontal cortex, and bilateral temporoparietal junctions) persisted during light non-REM sleep.94 Vincent
et al.91 demonstrated in deeply isoflurane-anesthetised
126
FIGURE 4. Preserved coherent blood oxygen leveldependent (BOLD) oscillations in the default
network persist in three documented states of unawareness. Brain areas showing correlations with a seed
voxel in the posterior cingulate cortex, after correction for spurious variance as described in Reference
91. From the left to the right, results of 12 volunteers random-effect analysis, from an individual sleeping
volunteer scanned during sleep stage 2 (from Ref. 68), from a patient in coma due to a nontraumatic
origin, and anaesthetized monkey data (reproduced by permission from Vincent et al.91 ). Sleep and
coma patients were masked inclusively with healthy volunteers results to check for spatial consistency of
the resting-state connectivity patterns.
monkeys preserved and coherent resting-state spontaneous fluctuations within three well-known neuroanatomical systems (oculomotor, somatomotor, and
visual) and within a network very close to the human
default system (see FIG. 4), a set of brain regions
thought by some to support uniquely human capabilities. These results demonstrate that cortical systems
previously associated with performance in sensory, motor, and/or cognitive tasks are manifest in the correlation structure of spontaneous BOLD fluctuations
observed in the absence of normal perception or behavior. Finally, using a method similar to that used in
Vincent et al.,91 we could identify persisting coherent
BOLD oscillations within the default-mode network
in coma, and during stage 2 slow-wave sleep (FIG. 4,
unpublished results).
All these results indicate that coherent systemwide
fluctuations probably reflect an aspect of brain
functional organization that transcends levels of consciousness.91 Thus, coherent spontaneous BOLD
fluctuations cannot be exclusively a reflection of
conscious mental activity,3 but may reflect a more
fundamental or intrinsic property of functional brain
organization. They should be considered as certainly
necessary, but not sufficient to support consciousness.
One could argue that the temporal dynamics of our
ongoing stream of consciousness (classically considered around 500 ms95 ) is much faster than the slow
fMRI BOLD oscillations occurring at around the 10-s
time period (0.1 Hz) observed here.
The physiological origin and functional significance
of low-frequency spontaneous brain activity fluctu-
Conclusion
Even if states of extremely low or high brain activity
are often associated with unconsciousness, the precise
link between global brain metabolism and awareness
remains difficult to assert. On the contrary, regional
brain activity in a widespread frontoparietal associative
network has been shown to be systematically altered in
all documented states of unconsciousness. In line with
studies in awake volunteers, these data emphasize the
potential role of frontoparietal association cortices in
the genesis of awareness.
Recent functional MRI studies have identified
coherent low-frequency fluctuations among welldocumented neuroanatomical networks. We, however,
9.
10.
11.
Acknowledgments
This work was supported by grants from the Belgian Fonds National de la Recherche Scientifique
(FNRS) and from the Centre Hospitalier Universitaire Sart Tilman, the University of Li`ege, the Mind
Science Foundation, the European Comission, the
French Speaking Community Concerted Research Action, and the Fondation Medicale Reine Elisabeth.
M.B. and T.D.V., C.P., S.L., and P.M. are, respectively, Research Fellows, Research Associate, Senior
Research Associate, and Research Director at FNRS.
M.S. was supported by an Austrian Science Fund
Erwin-Schrodinger Fellowship J2470-B02 (to M.S.).
We thank Dimitri Haye, Jacques Trantsieaux, and
Charlemagne Noukoua for their assistance in acquiring the fMRI data.
12.
13.
14.
15.
16.
17.
18.
Competing Interest
References
1. SHULMAN, G.L. et al. 1997. Common blood flow changes
across visual tasks: II. Decreases in cerebral cortex. J.
Cogn. Neurosci. 9: 648663.
2. RAICHLE, M.E. et al. 2001. A default mode of brain function.
Proc. Natl. Acad. Sci. USA 98: 676682.
3. MORCOM, A.M. & P.C. FLETCHER. 2007. Does the brain
have a baseline? Why we should be resisting a rest. NeuroImage 37: 10731082.
4. MORCOM, A.M. & P.C. FLETCHER. 2007. Cognitive neuroscience: the case for design rather than default. NeuroImage 37: 10971099.
5. RAICHLE, M.E. & A.Z. SNYDER. 2007. A default mode of
brain function: a brief history of an evolving idea. NeuroImage 37: 10831090.
6. RAICHLE, M.E. & M.A. MINTUN. 2006. Brain work and
brain imaging. Annu. Rev. Neurosci. 29: 449476.
7. FOX, M.D. & M.E. RAICHLE. 2007. Spontaneous fluctuations in brain activity observed with functional magnetic
resonance imaging. Nat. Rev. Neurosci. 8: 700711.
8. GREICIUS, M.D. et al. 2004. Default-mode network activity
distinguishes Alzheimers disease from healthy aging: evi-
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
127
128
29. SALINAS, E. & T.J. SEJNOWSKI. 2001. Correlated neuronal
activity and the flow of neural information. Nat. Rev. Neurosci. 2: 539550.
30. RAICHLE, M.E. & D.A. GUSNARD. 2002. Appraising the
brains energy budget. Proc. Natl. Acad. Sci. USA 99:
1023710239.
31. LAUREYS, S. et al. 1999. Cerebral metabolism during vegetative state and after recovery to consciousness. J. Neurol.
Neurosurg. Psychiatry 67: 121.
32. SCHIFF, N.D. et al. 2002. Residual cerebral activity and behavioural fragments can remain in the persistently vegetative brain. Brain 125: 12101234.
33. GUSNARD, D.A. & M.E. RAICHLE. 2001. Searching for a
baseline: functional imaging and the resting human brain.
Nat. Rev. Neurosci. 2: 685694.
34. MAQUET, P. 2000. Functional neuroimaging of normal human sleep by positron emission tomography. J. Sleep Res.
9: 207231.
35. BAARS, B.J., T.Z. RAMSOY & S. LAUREYS. 2003. Brain, conscious experience and the observing self. Trends Neurosci.
26: 671675.
36. KAISTI, K.K. et al. 2002. Effects of surgical levels of propofol and sevoflurane anesthesia on cerebral blood flow in
healthy subjects studied with positron emission tomography. Anesthesiology 96: 13581370.
37. BLUMENFELD, H. et al. 2003. Selective frontal, parietal, and
temporal networks in generalized seizures. NeuroImage
19: 15561566.
38. SALEK-HADDADI, A. et al. 2003. Functional magnetic resonance imaging of human absence seizures. Ann. Neurol.
53: 663667.
39. LAUFS, H. et al. 2006. Linking generalized spike-andwave discharges and resting state brain activity by using
EEG/fMRI in a patient with absence seizures. Epilepsia
47: 444448.
40. BLUMENFELD, H. et al. 2004. Positive and negative network
correlations in temporal lobe epilepsy. Cereb. Cortex. 14:
892902.
41. BASSETTI, C. et al. 2000. SPECT during sleepwalking. Lancet
356: 484485.
42. BAARS, B.J. 1988. A Cognitive Theory of Consciousness.
Cambridge University Press. Cambridge, UK.
43. BAARS, B.J. 2002. The conscious access hypothesis: origins
and recent evidence. Trends Cogn. Sci. 6: 4752.
44. DEHAENE, S. et al. 2001. Cerebral mechanisms of word masking and unconscious repetition priming. Nat. Neurosci. 4:
752758.
45. FOX, M.D. et al. 2005. The human brain is intrinsically organized into dynamic, anticorrelated functional networks.
Proc. Natl. Acad. Sci. USA 102: 96739678.
46. GUSNARD, D.A., M.E. RAICHLE & M.E. RAICHLE. 2001.
Searching for a baseline: functional imaging and the resting human brain. Nat. Rev. Neurosci. 2: 685694.
47. LAUREYS, S. et al. 2000. Restoration of thalamocortical connectivity after recovery from persistent vegetative state.
Lancet 355: 17901791.
48. WHITE, N.S. & M.T. ALKIRE. 2003. Impaired thalamocortical connectivity in humans during general-anestheticinduced unconsciousness. NeuroImage 19: 402
411.
129
82. HAMPSON, M. et al. 2002. Detection of functional connectivity using temporal correlations in MR images. Hum.
Brain Mapp. 15: 247262.
83. FOX, M.D. et al. 2006. Spontaneous neuronal activity distinguishes human dorsal and ventral attention systems. Proc.
Natl. Acad. Sci. USA 103: 1004610051.
84. CORDES, D. et al. 2002. Hierarchical clustering to measure
connectivity in fMRI resting-state data. Magn. Reson.
Imaging 20: 305317.
85. SALVADOR, R. et al. 2005. Neurophysiological architecture
of functional magnetic resonance images of human brain.
Cereb. Cortex. 15: 13321342.
86. BARTELS, A. & S. ZEKI. 2005. The chronoarchitecture of the
cerebral cortex. Philos. Trans. R Soc. Lond. B Biol. Sci.
360: 733750.
87. DAMOISEAUX, J.S. et al. 2006. Consistent resting-state networks across healthy subjects. Proc. Natl. Acad. Sci. USA
103: 1384813853.
88. TIAN, L. et al. 2007. The relationship within and between
the extrinsic and intrinsic systems indicated by resting state
correlational patterns of sensory cortices. NeuroImage 36:
684690.
89. MANTINI, D. et al. 2007. Electrophysiological signatures of
resting state networks in the human brain. Proc. Natl.
Acad. Sci. USA 104: 1317013175.
90. MASON, M.F. et al. 2007. Wandering minds: the default
network and stimulus-independent thought. Science 315:
393395.
91. VINCENT, J.L. et al. 2007. Intrinsic functional architecture
in the anaesthetized monkey brain. Nature 447: 83
86.
92. PELTIER, S.J. et al. 2005. Functional connectivity changes
with concentration of sevoflurane anesthesia. Neuroreport
16: 285288.
93. FUKUNAGA, M. et al. 2006. Large-amplitude, spatially correlated fluctuations in BOLD fMRI signals during extended
rest and early sleep stages. Magn. Reson. Imaging 24:
979992.
94. HOROVITZ, S.G. et al. 2007. Low frequency BOLD fluctuations during resting wakefulness and light sleep: a simultaneous EEG-fMRI study. Hum. Brain Mapp. (In press.)
95. LIBET, B. 2006. Reflections on the interaction of the mind
and brain. Prog. Neurobiol. 78: 322326.
REVIEWS
Abstract | Research over the past two decades broadly supports the claim that mindfulness
meditation practiced widely for the reduction of stress and promotion of health
exertsbeneficial effects on physical and mental health, and cognitive performance. Recent
neuroimaging studies have begun to uncover the brain areas and networks that mediate
these positive effects. However, the underlying neural mechanisms remain unclear, and it is
apparent that more methodologically rigorous studies are required if we are to gain a full
understanding of the neuronal and molecular bases of the changes in the brain that
accompany mindfulness meditation.
Longitudinal studies
Study designs that compare
data from one or more groups
at several time points and that
ideally include a (preferably
active) control condition and
random assignment to
conditions.
Cross-sectional studies
Study designs that compare
data from an experimental
group with those from a control
group at one point in time.
1
Department of Psychological
Sciences, Texas Tech
University, Lubbock, Texas
79409, USA.
2
Department of Psychology,
University of Oregon, Eugene,
Oregon 97403, USA.
3
Department of
Neuroradiology, Technical
University of Munich, 81675
Munich, Germany.
4
Massachusetts General
Hospital, Charlestown,
Massachusetts 02129, USA.
*These authors contributed
equally to this work.
Correspondence to Y.Y.T.
email: [email protected]
doi:10.1038/nrn3916
Published online
18 March 2015
REVIEWS
Box 1 | Mindfulness meditation
Different styles and forms of meditation are found in almost all cultures and religions. Mindfulness meditation originally
stems from Buddhist meditation traditions3. Since the 1990s, mindfulness meditation has been applied to multiple mental
and physical health conditions, and has received much attention in psychological research2,47. In current clinical and
research contexts, mindfulness meditation is typically described as non-judgemental attention to experiences in the
present moment38. This definition encompasses the Buddhist concepts of mindfulness and equanimity158 and describes
practices that require both the regulation of attention (in order to maintain the focus on immediate experiences, such as
thoughts, emotions, body posture and sensations) and the ability to approach ones experiences with openness and
acceptance210,53,158161. Mindfulness meditation can be subdivided into methods involving focused attention and those
involving open monitoring of present-moment experience9.
The mindfulness practices that have been the subject of neuroscientific research comprise a broad range of methods
and techniques, including Buddhist meditation traditions, such as Vipassana meditation, Dzogchen and Zen, as well as
mindfulness-based approaches such as integrative bodymind training (IBMT), mindfulness-based stress reduction
(MBSR) and clinical interventions based on MBSR46. Both
MBSR and IBMT have adopted mindfulness practices from a
Mindfulness meditation
the Buddhist traditions and aim to develop
moment-to-moment, non-judgemental awareness through
various techniques8-11,53,73. IBMT has been categorized in
Attention
Emotion
Selfthe literature as open-monitoring mindfulness
control
regulation
awareness
9,10,161
meditation
, whereas MBSR includes both focused
attention and open-monitoring practices8.
It has been suggested that mindfulness meditation
Self-regulation
includes at least three components that interact closely to
constitute a process of enhanced self-regulation:
enhanced attention control, improved emotion regulation b
and altered self-awareness (diminished self-referential
Early stage
Middle stage
Advanced stage
processing and enhanced body awareness)10 (see the
figure, part a). Mindfulness meditation can be roughly
Eort to
Eortful
Eortless
divided into three different stages of practice early,
reduce mind
doing
being
middle (intermediate) and advanced that involve
wandering
different amounts of effort11 (see the figure, part b).
Correlational studies
Studies that assess the
co-variation between two
variables: for example,
co-variation of functional or
structural properties of the
brain and a behavioural
variable, such as reported
stress.
groups of non-meditators matched on various dimensions9,18. The rationale was that any effects of meditation
would be most easily detectable in highly experienced
practitioners.
A number of cross-sectional studies revealed differences in brain structure and function associated with
meditation (see below). Although these differences may
constitute training-induced effects, a cross-sectional
study design precludes causal attribution: it is possible
that there are pre-existing differences in the brains of
meditators, which might be linked to their interest in
meditation, personality or temperament 2,19. Although
correlational studies have attempted to discover whether
more meditation experience is related to larger changes
in brain structure or function, such correlations still
cannot prove that meditation practice has caused the
changes because it is possible that individuals with these
particular brain characteristics may be drawn to longer
meditation practice.
More recent research has used longitudinal designs,
which compare data from one or more groups at several time points and ideally include a (preferably active)
control condition and random assignment to conditions1114,2025. In meditation research, longitudinal studies are still relatively rare. Among those studies, some
have investigated the effects of mindfulness training
over just a few days, whereas others have investigated
programmes of 1 to 3months. Some of these studies
have revealed changes in behaviour, brain structure and
function1114,2025. A lack of similar changes in the control
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
REVIEWS
Blood-oxygen-leveldependent contrasts
(BOLD contrasts). Signals that
can be extracted with
functional MRI and that reflect
the change in the amount of
deoxyhaemoglobin that is
induced by changes in the
activity of neurons and their
synapses in a region of the
brain. The signals thus reflect
the activity in a local brain
region.
Brain state
The reliable patterns of brain
activity that involve the
activation and/or connectivity
of multiple large-scale brain
networks.
Fractional anisotropy
A parameter in diffusion tensor
imaging, which images brain
structures by measuring the
diffusion properties of water
molecules. It provides
information about the
microstructural integrity of
white matter.
REVIEWS
Table 1 | Structural changes in the brain associated with mindfulness meditation
Meditation
tradition*
Control
Sample size of
meditation (M) and
control (C) groups
Type of measurement
Refs
Non-meditators
M: 20, C: 15
Cortical thickness
32
Zen
Non-meditators
M: 13, C: 13
Grey-matter volume
34
Insight
Non-meditators
M: 20, C: 20
Grey-matter density
31
Tibetan
Dzogchen
Non-meditators
M: 10, C: 10
Grey-matter density
33
Zen
Non-meditators
M: 17, C: 18
Cortical thickness
51
MBSR
Non-meditators
M: 20, C: 16
Grey-matter volume
52
Zen
Non-meditators
M: 10, C: 10
37
Active control:
relaxation training
M: 22, C: 23
38
MBSR
Individuals on a
waiting list
M: 16, C: 17
Grey-matter density
40
IBMT
(2weeks)
Active control:
relaxation training
M: 34, C: 34
39
Usual care
(patients with
Parkinson disease)
M: 14, C: 13
Grey-matter density
42
MBSR
Waiting list
(patients with
mild cognitive
impairment)
M: 8, C: 5
Hippocampal volume
(region of interest analysis)
41
DTI, diffusion tensor imaging; FA, fractional anisotropy; IBMT, integrative bodymind training; MBI, mindfulness-based intervention; MBSR, mindfulness-based
stress reduction. *Studies that include meditators from traditions other than mindfulness or studies only investigating correlations with other variables are not
listed. Meditators show increased values, unless otherwise noted.
Activation likelihood
estimation meta-analysis
A technique for
coordinate-based
meta-analysis of neuroimaging
data. It determines the
convergence of foci reported
from different experiments,
weighted by the number of
participants in each study.
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
REVIEWS
Posterior cingulate
cortex/precuneus
Anterior cingulate
cortex
Multiple
prefrontal
regions
Medial
prefrontal
cortex
Striatum
Amygdala
Insula
Medial view
Lateral view
REVIEWS
of attentional performance along with functional MRI
(fMRI) are needed. If supported by more rigorous
future research, the evidence of improved attention regulation and strengthened brain activity in the regions
underlying attentional control following mindfulness
meditation might be promising for the treatment of
psychiatric disorders in which there are deficiencies in
these functions74,84,85.
is important, with beginners showing a different pattern from expert meditators. However, although several
studies have pointed to the involvement of fronto-limbic
regions, very few studies have begun to relate changes
in these regions to changes in measures of behaviour or
well-being 10.
A frequently reported finding is that mindfulness
practice leads to (or is associated with) a diminished activation of the amygdala in response to emotional stimuli
during mindful states83,94,95 as well as in a resting state93,
suggesting a decrease in emotional arousal. However,
although such results have been reported for meditation
beginners, they have less consistently been detected in
experienced meditators95 (but see REF.18).
Prefrontal activations are often enhanced as an effect
of mindfulness meditation in novice meditators (but see
REF.29): for example, greater dorsolateral PFC responses
were found during executive processing within an emotional Stroop task in healthy individuals after 6weeks
of mindfulness training 82. Enhanced dorsomedial and
dorsolateral PFC activation was also detected when participants expected to see negative images while engaging
in a mindful state94. Moreover, after an MBSR course, an
enhanced activation in the ventrolateral PFC in people
suffering from anxiety was found when they labelled the
affect of emotional images97. By contrast, experienced
meditators have been found to show diminished activation in medial PFC regions95. This finding could be
interpreted as indicating reduced control (disengagement
of elaboration and appraisal) and greater acceptance of
affectivestates.
Neuroimaging studies of ameliorated pain processing through mindfulness meditation have also pointed
to expertise-related differences in the extent of cognitive control over sensory experience. Meditation beginners showed increased activity in areas involved in the
cognitive regulation of nociceptive processing (the ACC
and anterior insula) and areas involved in reframing the
evaluation of stimuli (the orbitofrontal cortex), along with
reduced activation in the primary somatosensory cortex
in a 4-day longitudinal study with no control group30,
whereas meditation experts were characterized by
decreased activation in dorsolateral and ventrolateral PFC
regions and enhancements in primary pain processing
regions (the insula, somatosensory cortex and thalamus)
compared with controls in two cross-sectional studies35,81.
These findings are in line with the assumption that
the process of mindfulness meditation is characterized
as an active cognitive regulation in meditation beginners, who need to overcome habitual ways of internally
reacting to ones emotions and might therefore show
greater prefrontal activation. Expert meditators might
not use this prefrontal control. Rather, they might have
automated an accepting stance towards their experience
and thus no longer engage in top-down control efforts
but instead show enhanced bottom-up processing 100.
In the early stages of meditation training, achieving
the meditation state seems to involve the use of attentional control and mental effort; thus, areas of the lateral prefrontal and parietal cortex are more active than
before training 11,16,100,101. This may reflect the higher level
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
REVIEWS
of effort often found when participants struggle to obtain
the meditation state in the early stages11,73,98,102. However,
in the advanced stages, prefrontalparietal activity is
often reduced or eliminated, but ACC, striatum and
insula activity remains9,10,53,73,76,101103. Whether effort has
a key role in PFC and ACC activation during or following
meditation needs further investigation.
Analysis of functional connectivity between regions
of the fronto-limbic network could help to further elucidate the regulatory function of executive control regions.
Only a few studies have included such analyses. One
cross-sectional study on pain processing in meditators
demonstrated decreased connectivity of executive and
pain-related brain regions35, and one study of mindfulness-naive smokers demonstrated reduced connectivity
between craving-related brain regions during a mindfulness condition compared to passive viewing of smokingrelated images during cigarette craving 96, suggesting
a functional decoupling of involved regions. Another
longitudinal, randomized study reported that people
suffering from anxiety showed a change from a negative
correlation between the activity of frontal regions and
that of the amygdala before intervention (that is, negative
connectivity) to a positive correlation between the activity
of these regions (positive connectivity) after a mindfulness intervention97. Because such a negative correlation
will occur when prefrontal regions downregulate limbic
activation104,105, it was speculated that the positive coupling
between the activity of the two regions after mindfulness
intervention might indicate that meditation involves
monitoring of arousal rather than a downregulation or
suppression of emotional responses, and that it might
be a unique signature of mindful emotion regulation.
REVIEWS
Although there is much debate about its exact function,
a widespread view holds that the default mode network
(DMN)119,120 is involved in self-referential processing. This
network includes midline structures of the brain, such as
areas of the medial PFC, posterior cingulate cortex (PCC),
anterior precuneus and inferior parietal lobule121,122. These
structures show high activity during rest, mind wandering
and conditions of stimulus-independent thought 121 and
have been suggested to support diverse mechanisms by
which an individual can project themselves into another
perspective123. fMRI studies have investigated activity
in the DMN in association with mindfulness practice.
Regions of the DMN (the medial PFC and PCC) showed
relatively little activity in meditators compared to controls across different types of meditation, which has
been interpreted as indicating diminished self-referential
processing 117. Functional connectivity analysis revealed
stronger coupling in experienced meditators between the
PCC, dorsal ACC and dorsolateral PFC, both at baseline
and during meditation, which was interpreted as indicating increased cognitive control over the function of
the DMN117. Increased functional connectivity was also
found between DMN regions and the ventromedial PFC
in participants with more compared to less meditation
experience118. It has been speculated that this increased
connectivity with ventromedial PFC regions supports
greater access of the default circuitry to information about
internal states because this region is highly interconnected
with limbic regions118.
Awareness of present-moment experiences. Evaluative
self-referential processing is assumed to decrease as an
effect of mindfulness meditation, whereas awareness of
present-moment experiences is thought to be enhanced.
Mindfulness practitioners often report that the practice of
attending to present-moment body sensations results in
an enhanced awareness of bodily states and greater perceptual clarity of subtle interoception. Empirical findings
to support this claim are mixed. Although studies that
assessed performance on a heartbeat detection task a
standard measure of interoceptive awareness found
no evidence that meditators had superior performance to
non-meditators124,125, other studies found that meditators
showed greater coherence between objective physiological
data and their subjective experience in regard to an emotional experience126 and the sensitivity of body regions127.
Multiple studies have shown the insula to be implicated in mindfulness meditation: it shows stronger activation during compassion meditation128 and following
mindfulness training 23,52,98, and has greater cortical thickness in experienced meditators32. Given its known role in
awareness129, it is conceivable that enhanced insula activity in meditators might represent the amplified awareness
of present-moment experience.
Similarly, a study reported an uncoupling of the right
insula and medial PFC and increased connectivity of the
right insula with dorsolateral PFC regions in individuals
after mindfulness training 98. The authors interpret their
findings as a shift in self-referential processing towards
a more self-detached and objective analysis of interoceptive and exteroceptive sensory events, rather than their
Future questions
Mechanisms of mindfulness-induced changes. A number of studies seem to suggest that mindfulness meditation induces changes in brain structure and function,
raising the question of which underlying mechanisms
support these processes. It is possible that engaging the
brain in mindfulness affects brain structure by inducing
dendritic branching, synaptogenesis, myelinogenesis or
even adult neurogenesis. Alternatively, it is possible that
mindfulness positively affects autonomic regulation and
immune activity, which may result in neuronal preservation, restoration and/or inhibition of apoptosis14,23,131.
It is well known that mindfulness-based techniques are
highly effective in stress reduction, and it is possible
that such stress reduction may mediate changes in brain
function14,48,132137 (BOX4). A combination of all of these
mechanisms may evenoccur.
It is also important to realize that the direction of
the observed effects of mindfulness meditation has
not been consistent across all studies. Although larger
values in meditators compared to controls are predominantly reported, a cross-sectional study also revealed
smaller fractional anisotropy and cortical thickness
values in meditators in some brain regions, including
the medial PFC, postcentral and inferior parietal cortices, PCC and medial occipital cortex 138. Along these
lines, mindfulness-induced increases are predominantly
observed in longitudinal studies. However, it was also
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
REVIEWS
Table 2 | Evidence for changes in the core brain regions after mindfulness meditation
Brain region
Study design
Findings*
ACC
(self-regulation of
attention and emotion)
76
23
82
97
157
117
118
23
52
23
Striatum
(regulation of attention
and emotion)
23
Amygdala
(emotional processing)
Longitudinal, mindful attention training (N = 12), Decreased activation in right amygdala in response to emotional
compassion training (N = 12) and active control pictures in a non-meditative state
(N = 12)
93
83
95
PFC
Longitudinal, mindfulness training (N = 30)
(attention and emotion) versus active control (N = 31)
PCC
(self-awareness)
Insula
(awareness and
emotional processing)
Refs
128
106
Exemplary studies for each region support its involvement in mindfulness (the list is not comprehensive). Future research will need to test the hypothesized
functions by relating behavioural and neuroimaging findings. ACC, anterior cingulate cortex; IBMT, integrative bodymind training; MBSR, mindfulness-based
stress reduction; PCC, posterior cingulate cortex; PFC, prefrontal cortex. *Meditators show increased values, unless otherwise noted.
reported, for example, that as a consequence of meditation, larger decreases in perceived stress were associated with larger decreases in grey-matter density in the
amygdala48. Thus, the underlying mechanisms seem to
be more complex than currently assumed, and further
research is necessary.
Although neuroimaging has advanced our understanding of the individual brain regions involved in
mindfulness meditation, most evidence supports the
idea that the brain processes information through the
dynamic interactions of distributed areas operating in
large-scale networks139,140. Because the complex mental
state of mindfulness is probably supported by alterations
in large-scale brain networks, future work should consider the inclusion of complex network analyses, rather
than restricting analyses to comparisons of the strength
of activations in single brain areas. Recent studies have
explored functional network architecture during the
resting state using these new tools141,142.
REVIEWS
Box 4 | Mindfulness meditation and stress
Stress reduction might be a potential mediator of the effects of mindfulness practice on
neural function. Mindfulness meditation has been shown to reduce stress14,132137; this is
most consistently documented in self-reported data132,133. A review of
mindfulness-based stress reduction (MBSR) studies showed a non-specific effect on
stress reduction, which is similar to that of standard relaxation training134. However,
findings in studies that have examined biomarkers of stress, such as cortisol levels, are
less consistent: changes in cortisol levels have been found in association with
mindfulness training in some studies14,136 but not in others132,135.
The brain is a target for stress and stress-related hormones. It undergoes functional
and structural remodelling in response to stress in a manner that is adaptive under
normal circumstances but can lead to damage when stress is excessive172. Evidence
suggests that vulnerability to stress-induced brain plasticity is prominent in the
prefrontal cortex (PFC), hippocampus, amygdala and other areas associated with
fear-related memories and self-regulatory behaviours172,173. The interactions between
these brain regions determine whether life experiences lead to successful adaptation
or maladaptation and impaired mental and physical health173. A study has shown that
chronic stress induces less flexibility in attention shifting in the rodent and human
adult174. This was paralleled by a reduction in apical dendritic arborization in rodent
medial PFC (specifically, in the anterior cingulate cortex) and fewer feedforward PFC
connections in humans under stress, effects that recovered when the stressor was
removed174. This suggests that the effects of chronic psychosocial stress on PFC
function and connectivity are plastic and can change quickly as a function of mental
state174. Studies have also shown that moderate to severe stress seems to increase the
volume of the amygdala but reduce the volume of the PFC and hippocampus175.
Mindfulness training, however, has been shown to enhance grey-matter density in the
hippocampus40. Furthermore, after mindfulness training, reductions in perceived stress
correlate with reductions in amygdala grey-matter density48. These findings suggest
that mindfulness meditation might be a potential intervention and prevention
strategy176. Thus, it is possible that mindfulness meditation reduces stress by improving
self-regulation, which enhances neuroplasticity and leads to health benefits. It should be
noted that mindfulness meditation might also directly modulate stress processing via a
bottom-up pathway, through which it alters the sympatheticadrenalmedullary and
hypothalamicpituitaryadrenal axes by increasing activity in the parasympathetic
nervous system; thus, mindfulness meditation could prevent sympathetic nervous
system fight-or-flight stress responses177,178. Indeed, some research has suggested that
mindfulness leads to increased activity of the parasympathetic nervous system23,179.
Brain-derived neurotrophic factor (BDNF) has been linked to numerous aspects of
plasticity in the brain. Stress-induced remodelling of the PFC, hippocampus and
amygdala coincides with changes in the levels of BDNF, supporting its role as a trophic
factor modulating neuronal survival and regulating synaptic plasticity131. However,
glucocorticoids and other molecules have been shown to act in conjunction with BDNF
to facilitate both morphological and molecular changes. Because some forms of
mindfulness meditation training have been found to reduce stress-induced cortisol
secretion, this could potentially have neuroprotective effects by increasing levels of
BDNF, and future research should explore this possible causal relationship136,149,180.
Multivariate pattern
analysis
A method of analysing
functional MRI data that is
capable of detecting and
characterizing information
represented in patterns of
activity distributed within and
across multiple regions of the
brain. Unlike univariate
approaches, which only
identify magnitudes of activity
in localized parts of the brain,
this approach can monitor
multiple areas at once.
subjects effectively and allow their mental states at different stages of mindfulness training to be decoded from
their brain activity 144146, possibly by applying techniques
such as multivariate pattern analysis147.
Interpretations of study outcomes remain tentative until they are clearly linked to subjective reports or
behavioural findings. Future studies should therefore
increasingly draw connections between behavioural
outcomes and neuroimaging data using the advanced
multi-level analyses mentionedabove.
Investigating individual differences. People respond to
mindfulness meditation differently. These differences
may derive from temperamental, personality or genetic
differences. Studies in other fields have suggested that
genetic polymorphisms may interact with experience to
Conclusions
Interest in the psychological and neuroscientific investigation of mindfulness meditation has increased markedly
over the past two decades. As is relatively common in a
new field of research, studies suffer from low methodological quality and present with speculative post-hoc interpretations. Knowledge of the mechanisms that underlie
the effects of meditation is therefore still in its infancy.
However, there is emerging evidence that mindfulness
meditation might cause neuroplastic changes in the structure and function of brain regions involved in regulation
of attention, emotion and self-awareness. Further research
needs to use longitudinal, randomized and actively controlled research designs and larger sample sizes to advance
the understanding of the mechanisms of mindfulness
meditation in regard to the interactions of complex brain
networks, and needs to connect neuroscientific findings
with behavioural data. If supported by rigorous research
studies, the practice of mindfulness meditation might be
promising for the treatment of clinical disorders and might
facilitate the cultivation of a healthy mind and increased
well-being.
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
REVIEWS
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
REVIEWS
72. Pagnoni,G. & Cekic,M. Age effects on grey matter
volume and attentional performance in Zen
meditation. Neurobiol. Aging 28, 16231627 (2007).
73. Tang,Y.Y. & Posner,M.I. Training brain networks and
states. Trends Cogn. Sci. 18, 345350 (2014).
74. Tang,Y.Y., Tang,R. & Posner,M.I. Brief meditation
training induces smoking reduction. Proc. Natl Acad.
Sci. USA 110, 1397113975 (2013).
75. Cahn,B.R. & Polich,J. Meditation states and traits:
EEG, ERP, and neuroimaging studies. Psychol. Bull.
132, 180211 (2006).
76. Hlzel,B.K. etal. Differential engagement of anterior
cingulate and adjacent medial frontal cortex in adept
meditators and non-meditators. Neurosci. Lett. 421,
1621 (2007).
77. Van Veen,V. & Carter,C.S. The anterior cingulate as a
conflict monitor: fMRI and ERP studies. Physiol.
Behav. 77, 477482 (2002).
78. Posner,M.I., Sheese,B., Rothbart,M. & Tang,Y.Y.
The anterior cingulate gyrus and the mechanism of
self-regulation. Cogn. Affect. Behav. Neurosci. 7,
391395 (2007).
79. Tang,Y.Y. & Tang,R. Ventral-subgenual anterior
cingulate cortex and self-transcendence. Front.
Psychol. 4, 1000(2014).
80. Sridharan,D., Levitin,D.J. & Menon,V. A critical role
for the right fronto-insular cortex in switching between
central-executive and default-mode networks. Proc.
Natl Acad. Sci. USA 105, 1256912574 (2008).
81. Gard,T. etal. Pain attenuation through mindfulness is
associated with decreased cognitive control and
increased sensory processing in the brain. Cereb.
Cortex 22, 26922702 (2012).
82. Allen,M. etal. Cognitive-affective neural plasticity
following active-controlled mindfulness intervention.
J.Neurosci. 32, 1560115610 (2012).
One of the first studies to document the effects of
mindfulness using active controls.
83. Goldin,P.R. & Gross,J.J. Effects of mindfulnessbased stress reduction (MBSR) on emotion regulation
in social anxiety disorder. Emotion 10, 8391
(2010).
84. Deckersbach,T., Hlzel,B.K., Eisner,L.R.,
Lazar,S.W. & Nierenberg,A.A. MindfulnessBased
Cognitive Therapy for Bipolar Disorder (Guildford
Press, 2014).
85. Passarotti,A.M., Sweeney,J.A. & Pavuluri,M.N.
Emotion processing influences working memory
circuits in pediatric bipolar disorder and attentiondeficit/hyperactivity disorder. J.Am. Acad. Child
Adolesc. Psychiatry 49, 10641080 (2010).
86. Gross,J.J. in Handbook of Emotion Regulation 2nd
edn (ed. Gross, J.J.) 320 (Guildford Press, 2014).
87. Ortner,C.N.M., Kilner,S.J. & Zelazo,P.D.
Mindfulness meditation and reduced emotional
interference on a cognitive task. Motiv. Emot. 31,
271283 (2007).
88. Goleman,D.J. & Schwartz,G.E. Meditation as an
intervention in stress reactivity. J.Consult. Clin.
Psychol. 44, 456466 (1976).
89. Robins,C.J., Keng,S.-L., Ekblad,A.G. &
Brantley,J.G. Effects of mindfulness-based stress
reduction on emotional experience and expression: a
randomized controlled trial. J.Clin. Psychol. 68,
117131 (2012).
90. Chambers,R., Lo,B.C.Y. & Allen,N.B. The impact of
intensive mindfulness training on attentional control,
cognitive style, and affect. Cogn. Ther. Res. 32,
303322 (2008).
91. Ding,X., Tang,Y.Y., Tang,R. & Posner,M.I. Improving
creativity performance by short-term meditation.
Behav. Brain Funct. 10, 9 (2014).
92. Jain,S. etal. A randomized controlled trial of
mindfulness meditation versus relaxation training:
effects on distress, positive states of mind,
rumination, and distraction. Ann. Behav. Med. 33,
1121 (2007).
93. Desbordes,G. etal. Effects of mindful-attention and
compassion meditation training on amygdala response
to emotional stimuli in an ordinary, non-meditative
state. Front. Hum. Neurosci. 6, 292 (2012).
94. Lutz,J. etal. Mindfulness and emotion regulation
an fMRI study. Soc. Cogn. Affect. Neurosci. 9,
776785 (2014).
95. Taylor,V.A. etal. Impact of mindfulness on the neural
responses to emotional pictures in experienced and
beginner meditators. Neuroimage 57, 15241533
(2011).
96. Westbrook,C. etal. Mindful attention reduces neural
and self-reported cue-induced craving in smokers. Soc.
Cogn. Affect. Neurosci. 8, 7384 (2013).
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
REVIEWS
148. van I.Jendoorn,M.H. etal. Gene-by-environment
experiments: a new approach to finding the missing
heritability. Nature Rev. Genet. 12, 881 (2011).
149. Jung,Y.H. etal. Influence of brain-derived
neurotrophic factor and catechol O-methyl transferase
polymorphisms on effects of meditation on plasma
catecholamines and stress. Stress 15, 97104 (2012).
150. Ding,X., Tang,Y.Y., Deng,Y., Tang,R. & Posner,M.I.
Mood and personality predict improvement in
creativity due to meditation training. Learn. Individ.
Differ. 37, 217221 (2014).
151. Rothbart,M.K. Becoming Who We Are (Guilford
Press, 2011).
152. Takahashi,T. etal. Changes in EEG and autonomic
nervous activity during meditation and their
association with personality traits. Int.
J.Psychophysiol. 55, 199207 (2005).
153. Moffitt,T.E. etal. A gradient of childhood self-control
predicts health, wealth, and public safety. Proc. Natl
Acad. Sci. USA 108, 26932698 (2011).
154. Hofmann,S.G., Sawyer,A.T., Witt,A. A. & Oh,D. The
effect of mindfulness-based therapy on anxiety and
depression: a meta-analytic review. J.Consult. Clin.
Psychol. 78, 169183 (2010).
A review of the effect of mindfulness-based therapy
on anxiety and mood symptoms.
155. Bowen,S. etal. Relative efficacy of mindfulness-based
relapse prevention, standard relapse prevention, and
treatment as usual for substance use disorders: a
randomized clinical trial. JAMA Psychiatry 71,
547556 (2014).
One of the first longitudinal studies to document the
effects of mindfulness on drug use and heavy drinking.
156. Schoenberg,P.L.A. etal. Effects of mindfulness-based
cognitive therapy on neurophysiological correlates of
performance monitoring in adult attention-deficit/
hyperactivity disorder. Clin. Neurophysiol. 125,
14071416 (2014).
157. Zeidan,F., Martucci,K.T., Kraft,R.A., McHaffie,J.G.
& Coghill,R.C. Neural correlates of mindfulness
meditation-related anxiety relief. Soc. Cogn. Affect.
Neurosci. 9, 751759 (2014).
158. Desbordes,G. etal. Moving beyond mindfulness:
defining equanimity as an outcome measure in
meditation and contemplative research. Mindfulness
http://dx.doi.org/10.1007/s12671-013-0269-8 (2014).
Acknowledgements
This work was supported by the US Office of Naval Research.
We thank E.Luders for her contributions to an earlier version
of this manuscript. We benefited from discussions with
R.Davidson and A.Chiesa. We thank four anonymous reviewers for their constructive comments and R.Tang for manuscript preparation.
CORRESPONDENCE
L I N K T O O R I G I N A L A RT I C L E
1.
2.
3.
4.
5.
6.
7.
8.
www.nature.com/reviews/neuro
2015 Macmillan Publishers Limited. All rights reserved
PeRSPecTiveS
OpiniOn
Abstract | After a pause of nearly 40 years in research into the effects of psychedelic
drugs, recent advances in our understanding of the neurobiology of psychedelics,
such as lysergic acid diethylamide (LSD), psilocybin and ketamine have led to
renewed interest in the clinical potential of psychedelics in the treatment of various
psychiatric disorders. Recent behavioural and neuroimaging data show that
psychedelics modulate neural circuits that have been implicated in mood and
affective disorders, and can reduce the clinical symptoms of these disorders. These
findings raise the possibility that research into psychedelics might identify novel
therapeutic mechanisms and approaches that are based on glutamate-driven
neuroplasticity.
Psychedelic drugs have long held a special
fascination for mankind because they produce an altered state of consciousness that is
characterized by distortions of perception,
hallucinations or visions, ecstasy, dissolution of self boundaries and the experience
of union with the world. As plant-derived
materials, they have been used traditionally
by many indigenous cultures in medical
and religious practices for centuries, if
not millennia1.
However, research into psychedelics
did not begin until the 1950s after the
breakthrough discovery of the classical
hallucinogen lysergic acid diethylamide
(LSD) by Albert Hofmann2 (TIMELINE). The
classical hallucinogens include indoleamines, such as psilocybin and LSD, and
phenethylamines, such as mescaline and
2,5-dimethoxy-4-iodo-amphetamine
(DOI). Research into psychedelics was
advanced in the mid 1960s by the finding
that dissociative anaesthetics such as ketamine and phencyclidine (PCP) also produce psychedelic-like effects3 (BOX 1). Given
their overlapping psychological effects,
both classes of drugs are included here
as psychedelics.
www.nature.com/reviews/neuro
2010 Macmillan Publishers Limited. All rights reserved
PersPectives
Current therapeutic studies
Several preclinical studies in the 1990s
revealed an important role for the NMDA
glutamate receptor in the mechanism of
action of antidepressants. These findings
consequently gave rise to the hypothesis that
the NMDA-antagonist ketamine might have
potential as an antidepressant24. This hypothesis was validated in an initial double-blind
placebo-controlled clinical study in seven
medication-free patients with major depression. Specifically, a significant reduction in
depression scores on the Hamilton depression
rating scale (HDRS) was observed 3 hours
after a single infusion of ketamine (0.5 mg
per kg), and this effect was sustained for at
least 72 hours25. Several studies have since
replicated this rapid antidepressant effect of
ketamine using larger sample sizes and treatment-resistant patients with depression2630.
Given that 71% of the patients met response
criteria (defined as a 50% reduction in HDRS
scores from baseline) within 24 hours26, this
rapid effect has a high therapeutic value. In
particular, patients with depression who are
suicidal might benefit from such a rapid and
marked effect as their acute mortality risk is
not considerably diminished with conventional antidepressants owing to their long
delay in onset of action (usually 23 weeks).
Indeed, suicidal ideations were reduced
24 hours after a single ketamine infusion28.
However, despite these impressive and
rapid effects, all but 2 of the patients relapsed
within 2 weeks after a single dose of ketamine26. Previous relapse prevention strategies,
such as the administration of either five
additional ketamine infusions29 or riluzole
(Rilutek; Sanofi-aventis) on a daily basis30,
yielded success only in some patients and
1897
Synthesis
of PcP
1919
Synthesis of
mescaline
by e. Spth
1926
Discovery of
psychoactive
effects of LSD
by A. Hofmann
1938
Synthesis
of LSD by
A. Hofmann
1943
1947
First LSD
study in
humans by
W. Stoll
1952
isolation and
synthesis of
psilocin and
psilocybin by
A. Hofmann
1953
1958
LSD appears on
the streets
1962
Synthesis
of ketamine
1963
Sandoz recalls
samples of
LSD and
ceases
supplying it
1965
introduction
of the term
dissociative
anaesthetic
by e. Domino
1966
Demonstration
of antagonistic
action of PcP at
NMDA receptors
by N. Anis
1970
LSD, psilocin
and mescaline
are placed in
Schedule i in
the US
1983
1988
First neuroimaging
study on psilocybin
and ketamine
1990
Demonstration of
agonistic action of
LSD at 5-HT2A
receptors; first
neuroimaging
study on mescaline
1999
Ketamine is
placed in
schedule iii
in the US
LSD, lysergic acid diethylamide; NMDA, N-methyl-d-aspartate; PcP, phencyclidine. Discoveries relating to classical hallucinogens and to dissociative anaesthetics are
shown by black and red boxes, respectively.
PersPectives
Box 1 | Assessing altered states of consciousness
Elementary
visual
alterations
Audiovisual
synesthaesia
Disembodiment
Impaired control
and cognition
Vivid imagery
Elementary
visual
alterations
Disembodiment
Audiovisual
synesthaesia
Impaired control
and cognition
Vivid imagery
10 20 30 40 50 60 70 Anxiety
Changed meaning
of percepts
20
30 40
50
60
Anxiety
Changed meaning
of percepts
Blissful state
Blissful state
Insightfulness
Insightfulness
Religious
experience
Experience
of unity
Experience
of unity
Religious
experience
Classical hallucinogens. The classical hallucinogens are comprised of three main chemical classes: the plant-derived tryptamines
(for example, psilocybin) and phenethylamines (for example, mescaline), and the
semisynthetic ergolines (for example, LSD)48.
Although all classical hallucinogens display
high affinity for 5-HT2 receptors, they also
interact to some degree with 5-HT1, 5-HT4,
5-HT5, 5-HT6 and 5-HT7 receptors12. In contrast to the tryptamines, the ergolines also
show high intrinsic activity at dopamine D2
receptors and at -adrenergic receptors49.
Converging evidence from pharmacological50, electrophysiological51,52 and behavioural studies in animals53,54 suggests that
classical hallucinogens produce their effects
in animals and possibly in humans primarily
through agonistic actions at cortical 5-HT2A
receptors (FIG. 1a). Consistent with this view,
selectively restoring 5-HT2A receptors in
www.nature.com/reviews/neuro
PersPectives
cortical pyramidal neurons is sufficient to
rescue hallucinogen-induced head shaking
in transgenic mice that lack 5-HT2A receptors53,55. Importantly, administration of the
5-HT2A receptor antagonist ketanserin abolishes virtually all of the psilocybin-induced
subjective effects in humans56. Recent studies have demonstrated that hallucinogenic
and non-hallucinogenic 5-HT2A agonists
differentially regulate intracellular signalling
pathways in cortical pyramidal neurons and
that this results in a differential expression
of downstream signalling proteins, such as
early growth response protein 1 (EGR1),
EGR2 and -arrestin 255,57. This suggests that
further elucidation of hallucinogen-specific
signalling pathways may aid the development of functionally selective ligands with
specific therapeutic properties for example, ligands that have antidepressant effects
but no hallucinogenic effects.
Several studies have demonstrated that
activation of 5-HT2A receptors by classical
hallucinogens or by serotonin leads to a
robust, glutamate-dependent increase in the
activity of pyramidal neurons, preferentially
Activation of 5-HT2A and 5-HT1A receptors in the medial PFC (mPFC) also has
downstream effects on serotonergic and
dopaminergic activity through descending projections to the dorsal raphe and the
ventral tegmental area (VTA). For example,
activation of 5-HT2A receptors in the mPFC
increases the firing rate of 5-HT neurons in
the dorsal raphe and of dopamine neurons
in the VTA, resulting in an increased release
of 5-HT in the mPFC58,66 and of dopamine in
mesocortical areas67 in animals. In a study
in humans, the hallucinogenic 5-HT2A agonist psilocybin increased striatal dopamine
concentrations, and this increase correlated
with euphoria and depersonalization
phenomena68. blocking dopamine D2
receptors by haloperidol, however, reduced
these effects by only about 30%. This
suggests that the dopaminergic system contributes only moderately to the broad spectrum of psilocybin-induced psychological
alterations56.
Interestingly, 5-HT2A receptor activation
not only seems to underlie the preponderance of the acute psychedelic effects of hallucinogens but may also lead to neuroplastic
adaptations in an extended prefrontallimbic
network. For example, in rats a single dose
of the hallucinogen DOI transiently
increased the dendritic spine size in cortical neurons69 and repeated doses of LSD
downregulated cortical 5-HT2A but not
5-HT1A receptors; effects that were the most
pronounced in the frontomedial cortex and
ACC70,71. It is possible that such adaptations
and specifically a downregulation of prefrontal
5-HT2A receptors might underlie some of
the therapeutic effects of hallucinogens in the
treatment of depression, anxiety and chronic
pain. In favour of this hypothesis, 5-HT2A
receptor density was found to be increased
in the PFC in post-mortem samples72 and
in vivo73,74 in patients with major depression,
and to be reduced after chronic treatment with
various antidepressants the reduction coinciding with the onset of clinical efficacy7577. In
addition, chronic, antisense-mediated downregulation of 5-HT2A receptors in rats78 and in
5-HT2A knockout mice79 reduced anxiety-like
behaviour, and selective restoration of 5-HT2A
receptors in the PFC normalized anxiety-like
behaviour in these 5-HT2A knockout mice.
These findings suggest that prefrontal 5-HT2A
receptors might modulate the activity of subcortical structures, such as the amygdala79.
Anxiety and depression are interrelated with
stress80, which also affects the serotonin system81. Stress elevates corticotropin-releasing
factor (CRF)82, and administration of CRF
into the mPFC of mice enhanced anxiety-like
VOLUME 11 | SEPTEMbER 2010 | 645
PersPectives
a
Cortical layer V
Glutamate
release
NMDAR
Brainstem
5-HT neuron
5-HT2A
+
AMPAR
Psilocin/
LSD/DMT
5-HT2A
BDNF
Psilocin/
LSD/DMT
b
Cortex
Subcortical areas
Glutamate
release
Ketamine
NMDAR
AMPAR
Interneuron
+
BDNF
NMDAR
GABA
Ketamine
Figure 1 | Activation of the prefrontal network and glutamate release by psychedelics. a | The
figure shows a model in which hallucinogens, such as psilocin, lysergic acid diethylamide (LSD) and
dimethyltryptamine (DMT), increase extracellular glutamate levels in the prefrontal cortex through
stimulation of postsynaptic serotonin (5-hydroxytryptamine) 2A (5-HT2A) receptors that are located
on large glutamatergic pyramidal cells in deep cortical layers (v and vi) projecting to layer v pyramidal
neurons. This glutamate release leads to an activation of AMPA (-amino-3-hydroxy-5-methyl-4isoxazole propionic acid) and NMDA (N-methyl-d-aspartate) receptors on cortical pyramidal neurons. in
addition, hallucinogens directly activate 5-HT2A receptors located on cortical pyramidal neurons. This
activation is thought to ultimately lead to increased expression of brain-derived neurotrophic factor
(BDNF). b | The figure shows a model in which dissociative NMDA antagonists, such as ketamine, block
inhibitory GABA (-aminobutyric acid)-ergic interneurons in cortical and subcortical brain areas, leading to enhanced firing of glutamatergic projection neurons and increased extracellular glutamate
levels in the prefrontal cortex. As ketamine also blocks NMDA receptors on cortical pyramidal neurons,
the increased glutamate release in the cortex is thought to stimulate cortical AMPA more than NMDA
receptors. The increased AMPA-receptor-mediated throughput relative to NMDA-receptor-mediated
throughput is thought ultimately to lead to increased expression of BDNF.
www.nature.com/reviews/neuro
2010 Macmillan Publishers Limited. All rights reserved
PersPectives
symptoms98,99, but in another study systemic
administration of the dopamine D2 receptor antagonist haloperidol did not attenuate
ketamine-induced psychotic symptoms
in healthy volunteers100. Although 5-HT2A
receptor antagonists reverse the disruptive
effects of NMDA antagonists on sensorimotor gating 101 and on object recognition102 in
animals, no comparable studies of the role
of serotonin in the mechanism of action of
NMDA antagonists have been conducted
in humans.
The enhanced glutamate release that
results from NMDA receptor blockade
by ketamine leads to an increased activation of AMPA receptors relative to NMDA
receptors95. The antidepressant-like effects
of ketamine and the selective NR2b antagonist CP-101,606 in animals can be blocked
by administration of the AMPA receptor
antagonist 2,3-dihydroxy-6-nitro-7-sulphamoyl-benzo[f]quinoxaline-2,3-dione
(NbQX)95, suggesting that enhanced AMPA
activation in cortical circuits is crucial for
the therapeutic effect of NMDA receptor
antagonists34,95.
A common mechanism? There is accumulating evidence that, despite their different primary modes of action, classical hallucinogens
and dissociative anaesthetics both modulate
glutamatergic neurotransmission in the prefrontallimbic circuitry that is implicated in
the pathophysiology of mood disorders. This
modulation is evidenced by the observation
in rats that hallucinogens103,104 and dissociative anaesthetics88,89 have a similar effect in
enhancing extracellular glutamate release
in the PFC, leading to increased activation
of pyramidal cells63,65,105,106. Furthermore,
and congruent with these findings, human
neuroimaging studies have shown that both
psilocybin and ketamine markedly activate
prefrontal cortical areas, including the ACC
and insula and, to a lesser extent, temporal and
parieto-occipital regions107111 (FIG. 2).
According to current models of emotion
regulation the PFC, including the ACC, exerts
cognitive, top-down control over emotion
and stress responses through its connections to the amygdala and dorsal raphe47,85.
Reduced prefrontal glutamate levels that are
associated with attenuated PFC activation
b
s-Ketamine
Psilocybin
PersPectives
treatment strategies that could improve the
well-being of patients and the associated
economic burden on patients and society.
Accumulating evidence shows a crucial
role for the glutamate system in the regulation of neuronal plasticity, and indicates that
abnormalities in neuroplasticity contribute
to the pathophysiology of mood disorders.
Thus, drugs that target neuronal plasticity
may offer a novel approach to their treatment. This Perspective proposes that classical
Glossary
Cluster period
A period of time during which cluster headache attacks
occur regularly.
Enantiomers
Two stereoisomeric molecules that are mirror images of
each other and are not superimposable.
Neurosis
A former term for a category of mental disorders
characterized by anxiety and a sense of distress. This
category includes disorders now classified as mood
disorders, anxiety disorders, dissociative disorders,
sexual disorders and somatoform disorders.
Regression
In Freudian psychoanalytic theory this term describes a
psychological strategy to cope with reality by means of
a temporary reversion of the ego to an earlier stage of
development.
Riluzole
A drug used to treat amyotrophic lateral sclerosis and that
has NMDA (N-methyl-D-aspartate) receptor blocking
properties similar to those of ketamine.
Schedule 1
A legislative category containing controlled drugs that have
a high potential for abuse, a lack of accepted safety and no
currently accepted medical use in treatments.
Self-actualization
The motivation to realize all of ones potential.
Structureactivity relationship
(Often abbreviated to SAR.) This is the relationship between
the chemical structure of a molecule and its biological activity.
Transference
A phenomenon in psychoanalysis characterized by
unconscious redirection of feelings or desires from one
person to another.
www.nature.com/reviews/neuro
2010 Macmillan Publishers Limited. All rights reserved
PersPectives
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
PersPectives
72. Shelton, R. C., Sanders-Bush, E., Manier, D. H. &
Lewis, D. A. Elevated 5-HT 2A receptors in
postmortem prefrontal cortex in major depression is
associated with reduced activity of protein kinase, A.
Neuroscience 158, 14061415 (2008).
73. Bhagwagar, Z. et al. Increased 5-HT2A receptor binding
in euthymic, medication-free patients recovered from
depression: a positron emission study with [11C]MDL
100,907. Am. J. Psychiatry 163, 15801587
(2006).
74. Meyer, J. H. et al. Dysfunctional attitudes and 5-HT2
receptors during depression and self-harm. Am.
J. Psychiatry 160, 9099 (2003).
75. Sibille, E. et al. Antisense inhibition of
5-hydroxytryptamine2a receptor induces an
antidepressant-like effect in mice. Mol. Pharmacol.
52, 10561063 (1997).
76. Yamauchi, M., Miyara, T., Matsushima, T. &
Imanishi, T. Desensitization of 5-HT2A receptor
function by chronic administration of selective
serotonin reuptake inhibitors. Brain Res. 1067,
164169 (2006).
77. Gomez-Gil, E. et al. Decrease of the platelet 5-HT2A
receptor function by long-term imipramine treatment
in endogenous depression. Hum. Psychopharmacol.
19, 251258 (2004).
78. Cohen, H. Anxiolytic effect and memory improvement
in rats by antisense oligodeoxynucleotide to
5-hydroxytryptamine-2A precursor protein. Depress.
Anxiety. 22, 8493 (2005).
79. Weisstaub, N. V. et al. Cortical 5-HT2A receptor
signaling modulates anxiety-like behaviors in mice.
Science 313, 536540 (2006).
80. Anisman, H., Merali, Z. & Stead, J. D. Experiential and
genetic contributions to depressive- and anxiety-like
disorders: clinical and experimental studies. Neurosci.
Biobehav. Rev. 32, 11851206 (2008).
81. Lukkes, J., Vuong, S., Scholl, J., Oliver, H. & Forster, G.
Corticotropin-releasing factor receptor antagonism
within the dorsal raphe nucleus reduces social anxietylike behavior after early-life social isolation.
J. Neurosci. 29, 99559960 (2009).
82. Reul, J. M. & Holsboer, F. Corticotropin-releasing
factor receptors 1 and 2 in anxiety and depression.
Curr. Opin. Pharmacol. 2, 2333 (2002).
83. Magalhaes, A. C. et al. CRF receptor 1 regulates
anxiety behavior via sensitization of 5-HT2 receptor
signaling. Nature Neurosci. 13, 622629 (2010).
84. Frokjaer, V. G. et al. Frontolimbic serotonin 2A
receptor binding in healthy subjects is associated with
personality risk factors for affective disorder. Biol.
Psychiatry 63, 569576 (2008).
85. Amat, J. et al. Medial prefrontal cortex determines
how stressor controllability affects behavior and
dorsal raphe nucleus. Nature Neurosci. 8, 365371
(2005).
86. Kupers, R. et al. A PET [18F]altanserin study of
5-HT12A receptor binding in the human brain and
responses to painful heat stimulation. Neuroimage
44, 10011007 (2009).
87. Oye, I., Paulsen, O. & Maurset, A. Effects of ketamine
on sensory perception: Evidence for a role of
N-methyl-D-aspartate receptors. J. Pharmac. Exp.
Ther. 260, 12091213 (1992).
88. Moghaddam, B., Adams, B., Verma, A. & Daly, D.
Activation of glutamatergic neurotransmission by
ketamine: a novel step in the pathway from NMDA
receptor blockade to dopaminergic and cognitive
disruptions associated with the prefrontal cortex.
J. Neurosci. 17, 29212927 (1997).
89. Lopez-Gil, X. et al. Clozapine and haloperidol
differently suppress the MK-801-increased
glutamatergic and serotonergic transmission in the
medial prefrontal cortex of the rat.
Neuropsychopharmacology 32, 20872097 (2007).
90. Jackson, M. E., Homayoun, H. & Moghaddam, B.
NMDA receptor hypofunction produces concomitant
firing rate potentiation and burst activity reduction in
the prefrontal cortex. Proc. Natl Acad. Sci. USA 101,
84678472 (2004).
91. Homayoun, H. & Moghaddam, B. NMDA receptor
hypofunction produces opposite effects on prefrontal
cortex interneurons and pyramidal neurons.
J. Neurosci. 27, 1149611500 (2007).
92. Jodo, E. et al. Activation of medial prefrontal cortex by
phencyclidine is mediated via a hippocampo-prefrontal
pathway. Cereb. Cortex 15, 663669 (2005).
93. Moghaddam, B. & Adams, B. W. Reversal of
phencyclidine effects by a group II metabotropic
glutamate receptor agonist in rats. Science 281,
13491352 (1998).
www.nature.com/reviews/neuro
2010 Macmillan Publishers Limited. All rights reserved
PersPectives
136. Osmond, H. A review of the clinical effects of
psychotomimetic agents. Ann. NY Acad. Sci. 66,
418434 (1957).
137. Kurland, A. A. LSD in the supportive care of the
terminally ill cancer patient. J. Psychoactive Drugs
17, 279290 (1985).
138. Abramson, H. A. The Use of LSD in Psychotherapy
and Alcoholism (Bobbs-Merrill, Indianapolis, 1967).
139. Hollister, L. E., Shelton, J. & Krieger, G. A controlled
comparison of lysergic acid diethylamide (LSD) and
dextroamphetmine in alcoholics. Am. J. Psychiatry
125, 13521357 (1969).
140. Savage, C. & McCabe, O. L. Residential psychedelic
(LSD) therapy for the narcotic addict. A controlled
study. Arch. Gen. Psychiatry 28, 808814 (1973).
141. Grof, S., Goodman, L. E., Richards, W. A. & Kurland,
A. A. LSD-assisted psychotherapy in patients with
terminal cancer. Int. Pharmacopsychiatry 8,
129144 (1973).
142. Pahnke, W. N. Psychedelic drugs and mystical
experience. Int. Psychiatry Clin. 5, 149162
(1969).
143. Grinspoon, L. & Bakalar, J. B. Psychedelic Drugs
Reconsidered (Basic Books., New York, 1979).
144. Crocket, R., Sandison, R. A. & Walk, A. in Proc.
R. MedPsychol. Assoc. (Lewis & Co., London,
1963).
145. Leuner H. in Ethnopsychotherapie (eds Dittrich, A. &
Scharfetter, C.) 151161 (Enke, Stuttgard, 1987)
146. Geert-Jorgensen, E. Further observations regarding
hallucinogenic treatment. Acta Psychiatr. Scand. 203
(Suppl.), 195200 (1968).
Acknowledgements
The authors would like to acknowledge the financial support
of the Swiss Neuromatrix Foundation (to F.X.V. and M.K.),
and of the Heffter Research Institute (to F.X.V.). The authors
thank D. Nichols for critical comments on the manuscript.
DATABASES
clinicaltrials.gov: http://clinicaltrials.gov
NcT00302744 | NcT00465595 | NcT00920387 |
NcT00947791 | NcT00957359
UniProtKB: http://www.uniprot.org
b-arrestin 2 | eGR1 | eGR2 | mGluR2 | mGluR3
FURTHER inFORMATiOn
University of Zurich Neuropsychopharmacology and Brain
imaging Groups homepage: http://www.dcp.uzh.ch/
research/groups/neuropsychopharmacology.html
All liNks Are Active iN the oNliNe pdf
that are influenced by SES (BOX 1). In addition, neuroscience research in animals
and in humans has provided candidate
mechanisms for the causeeffect relationships between SES and neural development.
This research has also demonstrated that at
least some of these effects are reversible. Such
a mechanistic understanding will enable the
design of more specific and powerful interventions to prevent and remediate the effects
of low childhood SES79.
Other recent reviews have discussed
research on SES-related differences in
neurocognitive development 79. In this
Perspective, we focus on the candidate
mechanisms by which SES influences brain
development, drawing from research in
humans and in animal models. We first
describe studies in humans that show that
SES influences cognitive and affective function in children, adolescents and young
adults. We then discuss studies in human
populations that have identified possible
mediators of the effects of SES, and review
research in animals in which these factors
were directly manipulated to assess their
effect on offspring outcomes.
SES effects on mental health and cognition
SES is a complex construct that is based
on household income, material resources,
education and occupation, as well as related
neighbourhood and family characteristics,
such as exposure to violence and toxins,
parental care and provision of a cognitively
stimulating environment 2,5,10,11 (for controversies regarding the measurement and
defining levels of SES see REFS 1,10,11). Not
only the lowest stratum but all levels of SES
affect emotional and cognitive development
to varying degrees1,1214. This implies that the
effects of SES that are reviewed here are
relevant to the entire population, although
it should be noted that the strongest effects
are often seen in people with the lowest
levels of SES.
Compared with children and adolescents
from higher-SES backgrounds, children
and adolescents from low-SES backgrounds
show higher rates of depression, anxiety,
attention problems and conduct disorders12,1518, and a higher prevalence of internalizing (that is, depression- or anxiety-like)
and externalizing (that is, aggressive and
impulsive) behaviours6,1921, all of which
increase with the duration of impoverishment 12,21. In addition, childhood SES influences cognitive development; it is positively
correlated with intelligence and academic
achievement from early childhood and
through adolescence2,3,6,14,19,22,23.
David A.
Rosenbaum
need not imply that intention formation and intention enactment are unrelated. Knowledge of how one can perform
affects what one intends to do, just as what one intends to
do affects how one acts. Still, it has proven useful to take
for granted that when an actor performs some voluntary
action, he or she has some goal in mind. Insofar as there are
different means of achieving that goal, the question of
interest for students of motor control is how the performed
movements are selected and controlled.
Researchers interested in the translation of intentions
into physical actions have largely focused on anticipatory
phenomena. The logic of the approach is straightforward. If
the activity of the nervous system prior to the performance
of some motor act differs from the activity of the nervous
system prior to the performance of some other motor act, it
is reasonable to suppose that the state of the nervous system
played a role in differentiating the two acts. Said another
way, the state of the nervous system is a necessary, if not
sufficient, condition for performing any particular act. By
this way of thinking, changes in the nervous system prior to
performance of a motor act reflect the history of the acts
genesis (e.g., Jeannerod, 1988).
The analysis of the precursors of voluntary motor acts
is not restricted to studies of neural activity. Behaviors, too,
betray their histories. Errors in performance provide clues
into the nature of plans for forthcoming actions, whether
for speech (e.g., Dell, 1986; Fromkin, 1973, 1980; Lashley,
1951), typewriting (Cooper, 1983; Rosenbaum, 1991, chap.
8), or other kinds of performance (Norman, 1981; Reason,
1990). Reaction times to begin production of motor sequences also provide information about the processes underlying movement generation (Henry & Rogers, 1960;
Klapp, 1977; Rosenbaum, 1987; S. Sternberg, Monsell,
Knoll, & Wright, 1978).
MayJune 2005 American Psychologist
Studies of motor control have also focused on learning. Through learning, motor acts are performed more
quickly, automatically, and consistently (Schmidt & Lee,
1999). An insight from the study of motor learning is that,
as practice continues, actors achieve greater flexibility in
the way they perform. One way they do so is by unlocking
biomechanical degrees of freedom. Thus, novice pistol
shooters tend to keep their elbows and wrists locked, but
with practice they allow these joints to counterrotate so
extension of one joint compensates for flexion of the other
(Arutyunyan, Gurfinkel, & Mirsky, 1969). Learners can
also acquire the ability to decouple joint motions. Thus,
skilled pianists can achieve greater independence of the
two hands than can novice piano players (Shaffer, 1976).
Similarly, experienced percussionists can generate more
complex polyrhythms than can new drummers (Pressing,
Summers, & Magill, 1996).
Studies of motor control have also focused on the
connection between perception and performance. What one
perceives affects how one acts, just as how one acts affects
what one perceives. Thus, ones capacity to tune ones
actions to the perceptual environment depends on ones
opportunity to actively explore the relations between ones
actions and the perceptions those actions afford (Held,
1965). Such exploration permits prediction of the perceptual consequences of behavior, and such predictions help
one determine whether perceptual changes originate from
changes in the external environment or from ones activity
in the environment. As a result, if the image of the visual
world shifts across the retina and the eye has been commanded to move, the retinal shift can be ascribed to motion
of the eye rather than to motion of the external environment
(von Helmholtz, 1909/1911).
Predicting perceptual consequences of motor acts
plays a key role in motor planning. As James (1890) wrote,
If I will to write Peter rather than Paul, it is the thought
of certain digital sensations, of certain alphabetic sounds,
of certain appearances on the paper, and of no others,
which immediately precedes the motion of my pen
(p. 500).
The idea that plans for motor action include perceptual
goals has received a great deal of support in recent years
(for reviews, see Hommel, Musseler, Aschersleben, &
Prinz, 2001). That plans for motor activity are partly perceptual makes sense from the perspective of feedback
processing. To respond adaptively to movement-related
feedback, one needs to have a goal against which the
feedback can be compared.
Table 1
Citations of Selected Topics in the Social Science
Citation Index From 1986 to 2004
Topic
Attention
Cognitive
Decision or judgment or reasoning
Language
Memory
Motor
Perception or pattern recognition
Occurrence
51,946
65,039
54,367
42,205
48,867
17,424
34,328
Note. All topics except for Motor are referred to in Ashcrafts (2002) Cognitive Psychology (3rd ed.; see Appendix B).
310
The methodological difficulties that Evarts (1973) described were later overcome, thanks in part to his own
refinement of microelectrode technology. A great deal of
research has subsequently been done on the brain activity
of awake, moving animals. Still, the range of movements
that is possible while brain activity is being recorded remains small compared with what is possible in the everyday environment because the technology used for recording
brain activity in awake subjects (e.g., functional magnetic
resonance imagery) requires the use of scanners that significantly limit mobility.
In addition to the technical difficulties of studying
motor control, the processes underlying movement planning and generation are relatively immune to conscious
inspection. As James (1890) wrote, For many actions, we
are aware of nothing between the conception and the execution. All sorts of neuromuscular processes come between
. . . but we know absolutely nothing of them. We think the
act, and it is done (p. 790).
The fact that psychologists seem to have no sense of
motor innervationa topic discussed at length by James
(1890)may have put the study of motor control at a
disadvantage, compared, say, with the study of visual perception, where visual images can be formed in ones mind.
There are problems with the too-hard-to-study-hypothesis, however. One is that much of what psychologists
know about motor control was discovered with techniques
that are no more complex than the techniques used to study
other psychological phenomena. Psychologists have
learned about the planning and control of movements by
recording mistakes that people make, by recording how
long it takes to initiate and perform predetermined motor
sequences, and by using other methods that are common in
experimental psychology. For example, using a simple
311
land, university students majoring in psychology are invited to take courses in motor control. Such classes are rare
in America.
These points notwithstanding, a predilection of Americas foremost behaviorist had the surprising effect of turning psychologists away from movement. When B. F. Skinner promoted operant conditioning, he downplayed the
importance of body movements per se, stressing instead the
instrumental effects of muscle activity. Thus, whether a
pigeon pecked a key or kicked the key did not much matter
to Skinner, though it mattered much to the pigeon. Paradoxically, Skinners pooh-poohing of movements appears
to have struck a chord with psychologists. With few exceptions, the analysis of body movements has received
little attention in psychology except as a means of addressing other questions. Thus, facial expressions have been
used to study emotion (Ekman & Oster, 1979), eye movements have been used to study reading (Rayner, 1983),
eyeblinks have been used to study memory (McCormick &
Thompson, 1984), and hand gestures have been used to
study language (Goldin-Meadow, 1999). The movements
people make to control external devices such as buttons or
joysticks have generally been less studied than the participants disposition to use the devices, as measured by
response probabilities or reaction times. Generally, the
question of how movements are controlled has been
ignored.
The Neuroscientists-Have-It-Covered
Hypothesis
The final viable hypothesis about the cause of psychologys
neglect of motor control is that motor control has long been
a forte of neuroscience. Why study a topic when another
group of researchers handles it well? To evaluate the neuroscientists-have-it-covered hypothesis, I used Web of Science to access the Science Citation Index in order to count
articles on the same topics as I had checked earlier in the
Social Science Citation Index. To limit the journals to ones
that were relevant, I restricted the search to journals that
had the word brain or the letters neur in their title. The
results appear in Table 2, where it is seen that the topic that
yielded the most citations was motor. Figure 1 shows the
relation between number of citations for the same set of
topics in Social Science Citation Index and Science Citation Index. Overall, there was a negative relation between
the number of citations in the two sources, with motor
being the topic with the greatest disparity in citation counts.
Why has motor-control research prospered in neuroscience? Apart from the fact that motor disorders cry out
for medical research, motor neurophysiologists can precisely stimulate different parts of the nervous system and
record the ensuing motor effects. Thus, motor neurophysiologists can escape the problem Broadbent (1993, p. 863)
identified: [I]t is much harder to control what a person
does than what stimulates them.
Does the success of motor-control research in neuroscience account for psychologys neglect of this topic? I
doubt it is the only source of the neglect, but I think it has
been an important one. Psychologists, like professionals in
MayJune 2005 American Psychologist
Table 2
Citations of Selected Topics in the Science Citation
Index for Journals Containing Brain or Neur in
Their Titles From 1986 to 2004
Topic
Attention
Cognitive
Decision or judgment or reasoning
Language
Memory
Motor
Perception or pattern recognition
Occurrence
3,747
6,049
1,177
1,833
8,537
10,913
2,791
Note. All topics except for Motor are referred to in Ashcrafts (2002) Cognitive Psychology (3rd ed.; see Appendix B).
any field, are less prone to pursue topics that are well
covered in other disciplines, particularly when they feel
they may have nothing special to offer. Psychologists do in
fact have something special to offer the study of motor
control: They can analyze macroscopic as well as microscopic aspects of behavior, and they can exploit their
knowledge of experimental design to reveal functional
principles that might otherwise go unnoticed. Still, if psychologists come from a tradition that is epistemologically
rather than action-based (the think-before-you-act hypothesis), if their tradition has made motor behavior a pariah
rather than an attractive research target (the baby-with-thebathwater hypothesis), and if grant money and other
sources of recognition are more liberally doled out to
scientists in other fields (the neuroscientists-have-it-covered hypothesis), there is not much incentive for psychologists to get on the move.
The Future
In the story of Cinderella, a modest chamber maid, abandoned by her wicked stepmother and stepsisters, is rescued
by a handsome prince. My aim in likening motor control to
Cinderella has not been to equate research domains to
wicked relatives, nor to equate myself with a prince, handsome or otherwise. Instead, my aim has been to point out
that motor control, which one may argue lies at the heart of
the science of mental life and behavior because it joins the
two, has had a surprisingly modest presence in psychology.
The reasons, I have suggested, are intellectual and economic. Intellectually, psychology grew out of philosophy,
where questions of knowing were taken to be quintessential
to epistemology. Only recently have psychologists come to
appreciate that acting and knowing are inseparable (Carlson, 1997), and only recently have psychologists come to
appreciate that purposeful movement helps initiate or sustain perceptionaction cycles rather than just being a response to input (for a particularly eloquent, early statement
of this position, see Weimer, 1977). Economically, psychologists have been inclined to work on problems for
313
Figure 1
Number of Articles in the Social Science Citation Index (Abscissa) and Science Citation Index (Ordinate)
Pertaining to Each Topic Listed in the Graph
Note. Values on the abscissa are 1/10,000 the number of reported values. Values on the ordinate are 1/1,000 the number of reported values. For the Science
Citation Index, the only journals included had the word brain or the letters neur in their titles.
315
316
Thelen, E. (1995). Motor development: A new synthesis. American Psychologist, 50, 79 95.
Tipper, S. P., Lortie, C., & Baylis, G. C. (1992). Selective reaching:
Evidence for action-centered attention. Journal of Experimental Psychology: Human Perception and Performance, 18, 891905.
Tolman, C. E. (1948). Cognitive maps in rats and man. Psychological
Review, 55, 189 208.
Turvey, M. T. (1990). Coordination. American Psychologist, 45,
938 953.
Vereijken, B., Whiting, H. T. A., & Beek, W. J. (1992). A dynamicsystems approach to skill acquisition. Quarterly Journal of Experimental Psychology: Human Experimental Psychology, 45(A), 323344.
von Helmholtz, H. (1911). Treatise on physiological optics (J. P. Southall,
Ed. & Trans.; 3rd ed., Vol. 3). Rochester, NY: Optical Society of
America. (Original work published 1909)
Weimer, W. B. (1977). A conceptual framework for cognitive psychology: Motor theories of the mind. In R. Shaw & J. Bransford (Eds.),
Perceiving, acting, and knowing: Toward an ecological psychology
(pp. 267311). Hillsdale, NJ: Erlbaum.
Wiesendanger, M. (1997). Paths of discovery in human motor control. In
M. C. Hepp-Reymond & G. Marini (Eds.), Perspectives of motor
behavior and its neural basis (pp. 103134). Basel, Switzerland:
Karger.
Willingham, D. T. (2004). Cognition: The thinking animal (2nd ed.).
Upper Saddle River, NJ: Prentice Hall.
Woodworth, R. S. (1899). The accuracy of voluntary movement. Psychological Review Monograph Supplements, 3(3, Suppl. 13), 1119.
Woodworth, R. S. (1938). Experimental psychology. New York: Holt.
Woodworth, R. S., & Schlosberg, H. (1954). Experimental psychology
(2nd ed.). New York: Holt.
Zanone, P. G., & Kelso, J. A. S. (1997). Coordination dynamics of
learning and transfer: Collective and component levels. Journal of
Experimental Psychology: Human Perception and Performance, 23,
1454 1480.
Appendix A
Brief Contents of Neissers (1967) Cognitive Psychology
1. The Cognitive Approach
2. Iconic Storage and Verbal Coding
3. Pattern Recognition
4. Focal Attention and Figural Synthesis
5. Words as Visual Patterns
6. Visual Memory
7. Speech Perception
8. Echoic Memory and Auditory Attention
9. Active Verbal Memory
10. Sentences
11. A Cognitive Approach to Memory and Thought
Appendix B
Brief Contents of Ashcrafts (2002) Cognitive Psychology
(3rd ed.)
1. Cognitive Psychology: An Introduction
2. The Cognitive Science Approach
3. Perception and Pattern Recognition
4. Attention
5. Short-Term Working Memory
6. Episodic Long-Term Memory
7. Semantic Long-Term Memory
8. Interactions in Long-Term Memory
9. Language
10. Comprehension: Written and Spoken Language
11. Decisions, Judgments, and Reasoning
12. Problem Solving
317