Skip to main content

Advertisement

Log in

Timing in cognition and EEG brain dynamics: discreteness versus continuity

  • Review
  • Published:
Cognitive Processing Aims and scope Submit manuscript

Abstract

This article provides an overview of recent developments in solving the timing problem (discreteness vs. continuity) in cognitive neuroscience. Both theoretical and empirical studies have been considered, with an emphasis on the framework of operational architectonics (OA) of brain functioning (Fingelkurts and Fingelkurts in Brain Mind 2:291–29, 2001; Neurosci Biobehav Rev 28:827–836, 2005). This framework explores the temporal structure of information flow and interarea interactions within the network of functional neuronal populations by examining topographic sharp transition processes in the scalp EEG, on the millisecond scale. We conclude, based on the OA framework, that brain functioning is best conceptualized in terms of continuity–discreteness unity which is also the characteristic property of cognition. At the end we emphasize where one might productively proceed for the future research.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. This means that there exists a particular minimal inter-stimulus interval for which two successive events are consistently perceived as simultaneous.

  2. Usually it is considered that things which are analog to be continuous and things which are digital to be discrete (Suber 1988; see also Eliasmith 2000). However, see Blachowitz (1997) in support of the ubiquity of this synonymizing and for an alternative to it.

  3. Consider a typical transistor. The processes in it are indeed continuous: That is its transfer function traces out a (nonlinear, continuous) S-shaped curve—that is why they are used as both switches and amplifiers (Eliasmith 2002). Nevertheless, we treat them as if they are only ever in one of the two possible states. Consistently with this analogy, van Gelder (1995) claims that even though neurons and neuronal assemblies may be considered as discrete in state, this does not mean that they are discrete in time.

  4. Figure 1 also illustrates the continuity and discreteness for the spatial dimension (see B). For the case of continuity, the change of neuronal activity appeared in a gradual fashion, whereas for the discrete condition, the change of neuronal activity appeared abruptly.

  5. To illustrate this point, consider only one example, which we borrow form Llinas et al. (1998, p 1841): “We know full well that if we are tired we can fall asleep extraordinarily quickly and that if we are asleep and a strong stimulus is given to us (e.g. the havoc played by an alarm clock) we can awaken also extraordinarily fast. It is so fast, indeed, that the only substrate capable of supporting the speed of these two events must be electrical in nature given the large number of elements involved...” EEG is mainly the “product” of the cortex. With this respect, there is interesting finding: Sahraie et al. (1997) compared brain activity in a single blindsight subject (G.Y.) generated by stimuli which give rise to awareness with activity generated by stimuli (permitting similar levels of discrimination) without awareness. They found that the shift between “aware” and “unaware” modes was associated with a shift in the pattern of activity from cortical to subcortical levels. Nunez (2000) also stated that subcortical activity is only weakly correlated with cognition and behavior. However, one should acknowledge the importance of such factors as the effects of some definite neurotransmitters controlled by thalamus (Newman 1995) and the establishment of reentrant thalamocortical loops for oscillatory EEG synchronization in order to bind specific features for consciously cognitive representation of the objects that are integrated from these features (Bachmann 1984; Llinas et al. 2002).

  6. The OA framework has its origin in the work of Kaplan and coworkers (Kaplan 1995, 1998, 1999; Kaplan et al. 1997; Kaplan and Shishkin 2000). We thoroughly endorse what they state, since our own particular perspective on the problem of brain-mind functioning does not differ substantially, although our choice of emphasis is very different in places.

  7. This approach goes back to Hebb (1949); however, the classical neural assemblies are too slow and may be not suitable for cognitive operations (Kaplan and Borisov 2003). Modern understanding of neural assemblies stresses its functional nature, which is at scales both coarser and finer than that of the classical ones (von der Malsburg 1999). The idea is that large neuronal populations can quickly become associated or disassociated, thus giving rise to transient assemblies (Frison 2000; Triesch and von der Malsburg 2001), which thought to execute the basic operations of informational processing (Averbeck and Lee 2004). For definition of “brain operation,” see Fingelkurts and Fingelkurts (2003, 2005). It is important to note here that the cell assembly’s concept is difficult to falsify (see Appendix 2 for details).

  8. If the data is stationary, its dynamics does not change significantly during the acquisition period, thus been stable. Therefore, quasi-stationary means almost (or near) stable.

  9. The attributes are the following: (1) Average amplitude within each segment (μV)—as generally agreed, indicates mainly the volume or size of neuronal population: indeed, the more neurons recruited into assembly through local synchronization of their activity, the higher will be the amplitude of corresponding to this assembly oscillations in the EEG (Nunez 2000); (2) Average length of segments (ms)—illustrates the functional life-span of neuronal population or the duration of operations produced by this population: since the transient NA functions during a particular time interval, this period is reflected in EEG as a stabilized interval of quasi-stationary activity (Fell et al. 2000; Kaplan and Shishkin 2000); (3) Coefficient of amplitude variability within segments (%)—shows the stability of local neuronal synchronization within neuronal population or assembly (Truccolo et al. 2002); (4) Average amplitude relation among adjacent segments (%)—indicates the NA behavior—growth (recruiting of new neurons) or distraction (functional elimination of neurons) (Kaplan and Borisov 2003); (5) Average steepness among adjacent segments (estimated in the close area of RTP) (%)—reflects the speed of neuronal population growth or distraction (Kaplan and Borisov 2003).

  10. Although the concept of metastability has been around in physics for a long time, the specific interpretation of metastability in the context of a theoretical model of the coordination dynamics in the brain has been developed by Kelso (1991). Metastability is a theory of how global integrative and local segregative tendencies in the brain coexist (Kelso 1995; Friston 1997; Kaplan 1998). In the metastable regime of brain functioning, the individual parts of the brain exhibit tendencies to function autonomously at the same time as they exhibit tendencies for coordinated activity (Bressler and Kelso 2001; see also Fingelkurts and Fingelkurts 2001, 2004). The synchronized operations of distributed neuronal assemblies are metastable spatial-temporal patterns because intrinsic differences in activity between the neuronal assemblies are sufficiently large that they do their own job, while still retaining a tendency to be coordinated together.

  11. OM means that the set of the neuronal assemblies synchronously participated in the same cognitive act during the analyzed period. The criterion for defining an OM is a sequence of the same synchro-complexes (SC). Whereby, SC is a set of EEG channels in which each channel forms a paired combination (with high values of index of structural synchrony) with all other EEG channels in the same set (Fig. 3a); meaning that all pairs of channels in an SC have to have significant index of structural synchrony (Fingelkurts et al. 2004b). For the properties of OM see Fingelkurts and Fingelkurts (2005).

  12. The index of structural synchrony (ISS) is estimated through synchronization of rapid transition processes (RTP)—boundaries between quasi-stationary segments—between different EEG channels. This procedure reveals the functional (operational) interrelationships between cortical sites as distinct from those measured using correlation, coherence and phase analysis (Kaplan et al. 2005).

  13. These values coincide precisely with the mean microstate duration of entire neocortex (82 ± 4 ms) obtained for healthy young adults using Lehmann approach (Koenig et al. 2002), which is essentially different from our method. Thus, these data cross-validate each other.

  14. It has been demonstrated that if two areas of cortex are operationally synchronized, then they tend to be also synchronized with some other areas (Fingelkurts 1998). Calculations showed that the power-law statistics governs the probability that a number of cortical areas are recruited into an OM. This ubiquitous dependency is characterized by a fractal relation between different levels of resolution of the data, a property also called self-organized criticality (Bak et al. 1987).

  15. Isomorphism is generally defined as a mapping of one entity into another having the same elemental structure, whereby the behaviors of the two entities are identically describable (Warfield 1977). A functional isomorphism on the other hand requires the functional connectivity between its component entities (Lehar 2003). It is an extension to Müller's psychophysical postulate (Müller 1896), and Chalmers' principle of structural coherence (Chalmers 1995).

  16. Indeed, experimental evidence suggests that the behavioral or cognitive continuum is a succession of discrete behavioral/cognitive acts performed by an individual (Alexandrov 1999; Madison 2001). Each separate act is the integration of a certain number of operations, which are important and appropriate for the realization of this act. The change from one behavioral/cognitive act to another is embedded in a rapid “transitional process” (Alexandrov 1999). The same is true for the phenomenological structure of human consciousness which consists of stable nuclei (or thoughts) and transitive fringes (or periods)—as it is described by James’ metaphor of “Stream of Thoughts” (James 1890). It seems that metastability provides a mechanism of the functional isomorphism realization (Fingelkurts and Fingelkurts 2004).

  17. In this effect normal listeners report hearing audio-visually fusion syllables as some combination of the auditory and visual syllables (e.g., auditory /ba/ + visual /ga/ are perceived as /va/) or as a syllable dominated by the visual syllable (e.g., auditory /ba/ + visual /va/ are perceived as /va/). The vast majority of people (but not all) experience the McGurk illusion. It was also shown that the McGurk illusion exists between other sensory modalities.

  18. Here there are no restrictions for the relations between frequency bands, because the method we used for assessing the OMs is not associated with the phase relation as the usual techniques estimating synchrony (Kaplan et al. 2005).

  19. This point has been emphasized many times during history of psychophysiology science as a “limited capacity of conscious state” (James 1890; Kahneman 1978; Posner 1987; Baars 1988; von der Malsburg 1997).

  20. Complexity hierarchy enables the system to build complex representations from primitive ones so that the semantic value of the complex representation is determined by, and dependent on, the semantic values of the primitives (Fingelkurts and Fingelkurts 2003).

  21. It should be stressed that the concepts “state” and “contents” of consciousness should be differentiated from each other. The “contents” of consciousness refer to the patterns of subjective experience at the phenomenal level: percepts, emotions, sensations, mental images, etc. (Block 1995), while the term “state” of consciousness refers to the underlying context in the brain in which the phenomenal contents of consciousness are realized. Thus, the “state” does not refer to the subjective experiences themselves (Kallio and Revonsuo 2003).

  22. The use of a posthypnotic suggestion would minimize the need for suggestions of relaxation, drowsiness, etc. which are typically used in a hypnotic induction (Kallio and Revonsuo 2003).

References

  • Albertazzi L (1998) Perceptual saliences and nuclei of meaning. In: Poli R (ed) The Brentano puzzle. Aldershot, Ashgate, pp 113–138

    Google Scholar 

  • Alexandrov YuI (1999) Psychophysiological regularities of the dynamics of individual experience and the “stream of consciousness.” In: Taddei-Feretti C, Musio C (eds) Series on biophysics and biocybernetics. Neural basis and psychological aspects of consciousness, Vol. 8—Biocybernetic. World Scientific, Singapore, pp 201–219

  • Allport DA (1968) Phenomenal simultaneity and perceptual moment hypothesis. Br J Psychol 59:395–406

    PubMed  CAS  Google Scholar 

  • Alpern M (1952) Metacontrast. Am J Optomol 29:631–646

    CAS  Google Scholar 

  • Andrews TJ, White LE, Binder D, Purves D (1996) Temporal events in cyclopean vision. Proc Natl Acad Sci USA 93:3689–3692

    PubMed  CAS  Google Scholar 

  • Arbib MA (2001) Co-evolution of human consciousness and language. In: Marijuan EP (ed) Cajal and consciousness: scientific approaches to consciousness on the Centential of Ramon y Cajal’s Textura. Vol. 929. Annals of the NYAS, New York, pp 195–220

  • Averbeck BB, Lee D (2004) Coding and transmission of information by neural ensembles. Trends Neurosci 27:225–230

    PubMed  CAS  Google Scholar 

  • Baars BJ (1988) A cognitive theory of consciousness. Cambridge University Press, New York

    Google Scholar 

  • Baars BJ (1997) In the theatre of consciousness: global workspace theory, a rigorous scientific theory of consciousness. J Conscious Stud 4:292–309

    Google Scholar 

  • Bachmann T (1984) The process of perceptual retouch: nonspecific afferent activation dynamics in explaining visual masking. Percept Psychophys 35:69–84

    PubMed  CAS  Google Scholar 

  • Bachmann T (1994) Psychophysiology of visual masking. Nova Science, Commack, New York

    Google Scholar 

  • Bachmann T (1999) Twelve spatiotemporal phenomena, and one explanation. In: Aschersleben G, Bachmann T, Musseler J (eds) Cognitive contributions to the perception of spatial and temporal events. Elsevier, Amsterdam, pp 173–206

    Google Scholar 

  • Bachmann T, Luiga I, Poder E, Kalev K (2003) Perceptual acceleration of objects in stream: evidence from ?ash-lag displays. Conscious Cogn 12:279–297

    PubMed  Google Scholar 

  • Bair W (1999) Spike timing in the mammalian visual system. Curr Opin Neurobiol 9:447–453

    PubMed  CAS  Google Scholar 

  • Bak P, Tang C, Wiesenfeld K (1987) Self-organized criticality: an explanation of 1/f noise. Phys Rev Lett 59:364–374

    Google Scholar 

  • Baker SN, Spinks R, Jackson A, Lemon RN (2001) Synchronization in monkey motor cortex during a precision grip task. I. Task-dependent modulation in single-unit synchrony. J Neurophysiol 85:869–885

    PubMed  CAS  Google Scholar 

  • Barrie JM, Freeman WJ, Lenhart MD (1996) Spatiotemporal analysis of prepyriform, visual, auditory, and somesthetic surface EEGs in trained rabbits. J Neurophysiol 76:520–539

    PubMed  CAS  Google Scholar 

  • Barsalou LW (1999) Perceptual symbol systems. Behav Brain Sci 22:577–609

    PubMed  CAS  Google Scholar 

  • Basar E (2004) Macrodynamics of electrical activity in the whole brain. Int J Bifurcat Chaos 14:363–381

    Google Scholar 

  • Basar E (2005) Memory as the “whole brain work.” A large-scale model based on “oscillations in super-synergy.” Int J Psychophysiol 58:199–226

    PubMed  Google Scholar 

  • Basar E, Basar-Eroglu C, Karakas S, Schurmann M (2001) Gamma, alpha, delta, and theta oscillations govern cognitive processes. Int J Psychophysiol 39:241–248

    PubMed  CAS  Google Scholar 

  • Bayne T, Chalmers DJ (2002) What is the unity of consciousness? In: Cleeremans A (ed) The unity of consciousness: binding, integration, dissociation. Oxford University Press, Oxford URL = http://www.u.arizona.edu/∼chalmers/papers/unity.html

  • Bennett NVL, Zukin RS (2004) Electrical coupling and neuronal synchronization in the mammalian brain. Neuron 41:495–511

    PubMed  CAS  Google Scholar 

  • Bickle J, Worley C, Bernstein M (2000) Vector subtraction implemented neurally: a neurocomputational model of some sequential cognitive and conscious processes. Conscious Cogn 9:117–144

    PubMed  CAS  Google Scholar 

  • Birch T (2002) Introduction to mental images. Homepage, URL = http://www.gis.net/∼tbirch/hp5.html

  • Blachowicz J (1997) Analog representation beyond mental imagery. J Philos 94:55–84

    Google Scholar 

  • Block RA (1990) Models of psychological time. In: Block RA (ed) Cognitive models of psychological time. Erlbaum, Hillsdale, pp 1–35

    Google Scholar 

  • Block N (1995) On a confusion about a function of consciousness. Behav Brain Sci 18:227–287

    Google Scholar 

  • Bolton TL (1894) Rhythm. Am J Psychol 6:145–238

    Google Scholar 

  • Borisov SV (2002) Studying of a phasic structure of the alpha activity of human EEG. PhD dissertation, Moscow State University, Moscow, Russian Federation, 213 pp (in Russian)

  • Braitenberg V (1980) Alcune considerazione sui meccanismi cerebrali del linguaggio. In: Braga G, Braitenberg V, Cipolli C, Coseriu E, Crespi-Reghizzi S, Mehler J, Titone R (eds) L’accostamento interdisciplinare allo studio del linguaggio. Franco Angeli Editore, Braitenberg

  • Braitenberg V, Pulvermüller F (1992) Entwurf einer neurologischen Theorie der Sprache. Naturwissenschaften 79:103–117

    PubMed  CAS  Google Scholar 

  • Bressler SL (1995) Large-scale cortical networks and cognition. Brain Res Brain Res Rev 20:288–304

    PubMed  CAS  Google Scholar 

  • Bressler SL (2003) Cortical coordination dynamics and the disorganization syndrome in schizophrenia. Neuropsychopharmacology 28:S35–S39

    PubMed  Google Scholar 

  • Bressler SL, Kelso JAS (2001) Cortical coordination dynamics and cognition. Trends Cogn Sci 5:26–36

    PubMed  Google Scholar 

  • Brodsky BE, Darkhovsky BS (1993) Nonparametric methods in change-point problems. Kluwer, Dordrecht

    Google Scholar 

  • Brodsky BE, Darkhovsky BS, Kaplan AYA, Shishkin SL (1999) A nonparametric method for the segmentation of the EEG. Comput Methods Programs Biomed 60:93–106

    PubMed  CAS  Google Scholar 

  • Brown JW (1998) Fundamentals of process neuropsychology. Brain Cogn 38:234–245

    PubMed  CAS  Google Scholar 

  • Bullock TH (1997) Signals and signs in the nervous system: The dynamic anatomy of electrical activity. Proc Natl Acad Sci USA 94:1–6

    PubMed  CAS  Google Scholar 

  • Buzsáki G (2004) Large-scale recording of neuronal ensembles. Nat Neurosci 7:446–451

    PubMed  Google Scholar 

  • Buzsáki G, Draguhn A (2004) Neuronal oscillations in cortical networks. Science 304:1926–1929

    PubMed  Google Scholar 

  • Callaway E, Layne RS (1964) Interaction between the visual evoked response and two spontaneous biological rhythms: the EEG alpha cycle and the cardiac arousal cycle. Ann NY Acad Sci 112:421–431

    PubMed  Google Scholar 

  • Cariani P (1997) Emergence of new signal-primitives in neural systems. Intellectica 2:95–143

    Google Scholar 

  • Chafe WL (1994) Discourse, consciousness, and time: the flow and displacement of conscious experience in speaking and writing. University of Chicago Press, Chicago

    Google Scholar 

  • Chalmers DJ (1995) Facing up to the problems of consciousness. J Conscious Stud 2:200–219

    Google Scholar 

  • Chalmers DJ (2002) Consciousness and its place in nature. In: Chalmers D (ed) Philosophy of mind: classical and contemporary readings. Oxford, URL = http://www.jamaica.u.arizona.edu/∼chalmers/papers/nature.html

  • Chomsky N (1957) Syntactic structures. The Hague, Mouton

    Google Scholar 

  • Churchland PS, Sejnowski T (1992) The computational brain. MIT, Cambridge

    Google Scholar 

  • Cooper LA (1975) Mental transformation of random two-dimensional shapes. Cognit Psychol 7:20–43

    Google Scholar 

  • Crick F, Koch C (2003) A framework for consciousness. Nat Neurosci 6:119–126

    PubMed  CAS  Google Scholar 

  • Crone NE, Miglioretti DL, Gordon B, Lesser RP (1998) Functional mapping of human sensorimotor cortex with electrocorticographic spectral analysis. II. Event-related synchronization in the gamma band. Brain 121:2301–2315

    PubMed  Google Scholar 

  • Cuesta MJ, Peralta V (2001) Integrating psychopathological dimensions in functional psychoses: a hierarchical approach. Schizophr Res 52:215–229

    PubMed  CAS  Google Scholar 

  • Damasio AR (1994) Descartes’ error; emotion, reason and the human brain. Picador, New York

    Google Scholar 

  • Damasio AR (2000) The feeling of what happens. Body, emotion and the making of consciousness. Vintage, London

    Google Scholar 

  • Dennett DC (1991) Consciousness explained. Little Brown, Boston

    Google Scholar 

  • Dennett DC, Kinsbourne M (1995) Time and the observer: The where and when of consciousness in the brain. Behav Brain Sci 15:183–247

    Google Scholar 

  • Dietze G (1885) Untersuchungen über den Umfang des Bewusstseins bei regelmässig auf einander folgenden Schalleindrücken. Philos Stud 2:362–393

    Google Scholar 

  • DiLollo V, Enns JT, Rensink RA (2000) Competition for consciousness among visual events: the psychophysics of reentrant visual processes. J Exp Psychol Gen 129:481–507

    CAS  Google Scholar 

  • Dinse H (1990) A temporal structure of cortical information processing. Concepts Neurosci 1:199–238

    Google Scholar 

  • Dinse H (1994) A time-based approach towards cortical functions: neural mechanisms underlying dynamic aspects of information processing before and after postontogenetic plastic processes. Physica D 75:129–150

    Google Scholar 

  • Donald M (1993) Precis of origins of the modern mind: three stages in the evolution of culture and cognition. Behav Brain Sci 16:737–791

    Google Scholar 

  • Eagleman DM (2001) Visual illusions and neurobiology. Nat Rev Neurosci 2:920–926

    PubMed  CAS  Google Scholar 

  • Edelman GM, Tononi G (2000) A Universe of consciousness: how matter becomes imagination. Basic Books, New York

    Google Scholar 

  • Effern A, Lehnertz K, Fernandez G, Grunwald T, David P, Elger CE (2000) Single trial analysis of event related potentials: non-linear de-noising with wavelets. Clin Neurophysiol 111:2255–2263

    PubMed  CAS  Google Scholar 

  • Efron E (1970) The minimum duration of a perception. Neuropsychologia 8:57–63

    PubMed  CAS  Google Scholar 

  • Eliasmith C (2000) Is the brain analog or digital? The solution and its consequences for cognitive science. Cogn Sci Q 1:147–170

    Google Scholar 

  • Eliasmith C (2001) Attractive and in-discrete: a critique of two putative virtues of the dynamicist theory of mind. Minds Mach 11:417–426

    Google Scholar 

  • Eliasmith C (2002) Discreteness and relevance: a reply to Roman Poznanski. Minds Mach 12: 437–438

    Google Scholar 

  • Elul R (1972a) The genesis of the EEG. In: Pfeiffer CC, Smythes JR (eds) International review of neurobiology, Vol. 15. Academic, New York, pp 227–272

  • Elul R (1972b) Randomness and synchrony in the generation of electroencephalogram. In: Petsche H, Brazier MAB (eds) Synchronization of EEG activity in epilepsies. Springer, Vienna, pp 59–77

    Google Scholar 

  • Fell J, Kaplan A, Darkhovsky B, Röschke J (2000) EEG analysis with nonlinear deterministic and stochastic methods: a combined strategy. Acta Neurobiol Exp 60:87–108

    CAS  Google Scholar 

  • Fingelkurts AnA (1998) Time-spatial organization of human EEG segment’s structure. PhD Dissertation, Moscow State University, Moscow, Russian Federation 415 pp (in Russian)

  • Fingelkurts AnA, Fingelkurts AlA (2001) Operational architectonics of the human brain biopotential field: towards solving the mind-brain problem. Brain Mind 2:261–296, URL = http://www.bm-science.com/team/art18.pdf

  • Fingelkurts AnA, Fingelkurts AlA (2003) Operational architectonics of perception and cognition: a principle of self-organized metastable brain states. In: VI Parmenides workshop—Perception and thinking, Institute of Medical Psychology. April 5–10, Elba/Italy (invited full-text contribution) URL = http://www.bm-science.com/team/art24.pdf

  • Fingelkurts AnA, Fingelkurts AlA (2004) Making complexity simpler: multivariability and metastability in the Brain. Int J Neurosci 114:843–862

    PubMed  Google Scholar 

  • Fingelkurts AnA, Fingelkurts AlA (2005) Mapping of the brain operational architectonics. Chapter 2. In: Chen FJ (ed) Focus on brain mapping research. Nova Science Publishers, Inc., pp 59–98. URL = http://www.bm-science.com/team/chapt3.pdf

  • Fingelkurts AlA, Fingelkurts AnA, Kaplan AYa (2003) The regularities of the discrete nature of multi-variability of EEG spectral patterns. Int J Psychophysiol 47:23–41

    PubMed  Google Scholar 

  • Fingelkurts AnA, Fingelkurts AnA, Kähkönen SA (2005) Functional connectivity in the brain—is it an elusive concept? Neurosci Biobehav Rev 28:827–836

    PubMed  Google Scholar 

  • Fingelkurts AlA, Fingelkurts AnA, Krause CM, Sams M (2002) Probability interrelations between pre-/post-stimulus intervals and ERD/ERS during a memory task. Clin Neurophysiol 113:826–843

    PubMed  Google Scholar 

  • Fingelkurts AnA, Fingelkurts AlA, Krause CM, Möttönen R, Sams M (2003a) Cortical operational synchrony during audio-visual speech integration. Brain Lang 85:297–312

    Google Scholar 

  • Fingelkurts AnA, Fingelkurts AlA, Krause CM, Kaplan AYa, Borisov SV, Sams M (2003b) Structural (operational) synchrony of EEG alpha activity during an auditory memory task. NeuroImage 20:529–542

    Google Scholar 

  • Fingelkurts AlA, Fingelkurts AnA, Krause CM, Kaplan AYa (2003c) Systematic rules underlying spectral pattern variability: Experimental results and a review of the evidence. Int J Neurosci 113:1447–1473

    Google Scholar 

  • Fingelkurts AnA, Fingelkurts AlA, Fingelkurts AnA, Kivisaari R, Pekkonen E, Ilmoniemi RJ, Kähkönen SA (2004a) Enhancement of GABA-related signalling is associated with increase of functional connectivity in human cortex. Hum Brain Mapp 22:27–39

    Google Scholar 

  • Fingelkurts AnA, Fingelkurts AlA, Fingelkurts AnA, Kivisaari R, Pekkonen E, Ilmoniemi RJ, Kähkönen SA (2004b) Local and remote functional connectivity of neocortex under the inhibition influence. Neuroimage 22:1390–1406

    Google Scholar 

  • Finke RA, Kurtzman HS (1981) Mapping the visual field in mental imagery. J Exp Psychol Gen 110:501–517

    PubMed  CAS  Google Scholar 

  • Flohr H (1995) Sensations and brain processes. Behav Brain Res 71:157–161

    PubMed  CAS  Google Scholar 

  • Fodor J, Pylyshyn Z (1988) Connectionism and cognitive architecture: a critical analysis. Cognition 28:3–71

    PubMed  CAS  Google Scholar 

  • Fraisse P (1978) Time and rhythm perception. In: Carterette EC, Friedman MP (eds) Handbook of perception, Vol. 8. Academic, New York, pp 203–254

  • Fraisse P (1984) Perception and estimation of time. Annu Rev Psychol 35:1–36

    PubMed  CAS  Google Scholar 

  • Freeman WJ (1972) Waves, pulses and the theory of neural masses. Prog Theor Biol 2:87–165

    Google Scholar 

  • Freeman WJ (1974) A model for mutual excitation in a neuron population in olfactory bulb. IEEE Trans Biomed Eng 21:350–358

    PubMed  CAS  Google Scholar 

  • Freeman WJ (2003) Evidence from human scalp electroencephalograms of global chaotic itinerancy. Chaos 13:1067–1077

    PubMed  Google Scholar 

  • Freeman WJ, Barrie JM (1993) Chaotic oscillations and the genesis of meaning in cerebral cortex. In: The IPSEN foundation symposium on temporal coding in the brain, Paris, 11 October 1993

  • Freeman WJ, Vitiello G (2005) Nonlinear brain dynamics and many-body field dynamics. Electromagn Biol Med 24:233–241

    Article  Google Scholar 

  • Freeman WJ, Rogers LJ, Holmes MD, Silbergeld DL (2000) Spatial spectral analysis of human electrocorticograms including the alpha and gamma bands. J Neurosci Methods 95:111–121

    PubMed  CAS  Google Scholar 

  • Friston KJ (1997) Transients, metastability and neural dynamics. NeuroImage 5:164–171

    PubMed  CAS  Google Scholar 

  • Friston K (2000) The labile brain. I. Neuronal transients and nonlinear coupling. Philos Trans R Soc Lond B Biol Sci 355:215–236

    PubMed  CAS  Google Scholar 

  • Galambos R, Makeig S, Talmachoff PJ (1981) A 40-Hz auditory potential recorded from the human scalp. Proc Natl Acad Sci USA 78:2643–2647

    PubMed  CAS  Google Scholar 

  • Galin D (1994) The structure of awareness: contemporary applications of William James’ forgotten concept of “the fringe.” J Mind Behav 15:375–402

    Google Scholar 

  • Galin D (2000) Comments on Epstein’s neurocognitive interpretation of William James’s model of consciousness. Conscious Cogn 9: 576–583

    PubMed  CAS  Google Scholar 

  • Geissler H-G (1987) The temporal architecture of central information processing: evidence for a tentative time-quantum model. Psychol Res 49:99–106

    Google Scholar 

  • Geissler H-G (1997) Is there a way from behavior to non-linear brain dynamics? On quantal periods in cognition and the place of alpha in brain resonances. Int J Psychophysiol 26:381–393

    PubMed  CAS  Google Scholar 

  • Geissler H-G, Schebera F-U, Kompass R (1999) Ultra-precise quantal timing: evidence from simultaneity thresholds in long-range apparent movement. Percept Psychophys 61:707–726

    PubMed  CAS  Google Scholar 

  • Gho M, Varela FJ (1988) A quantitative assessment of the dependency of the visual temporal frame upon the cortical rhythm. Am J Physiol 83:95–101

    Google Scholar 

  • Giaquinta A, Argentina M, Velarde MG (2000) A simple generalized excitability model mimicking salient features of neuron dynamics. J Stat Phys 101:665–678

    Google Scholar 

  • Glicksohn J (2001) Temporal cognition and the phenomenology of time: a multiplicative function for apparent duration. Conscious Cogn 10:1–25

    PubMed  CAS  Google Scholar 

  • Gobet F, Lane PCR, Croker S, Cheng PC-H, Oliver I, Pine JM (2001) Chanking mechanisms in human learning. Trends Cogn Sci 5:236–243

    PubMed  Google Scholar 

  • Gomes G (2002) Problems in the timing of conscious experience. Conscious Cogn 11:191–197

    PubMed  Google Scholar 

  • Gray CM, Singer W (1989) Stimulus-specific neuronal oscillations in orientation columns of cat visual cortex. Proc Natl Acad Sci USA 86:1698–1702

    PubMed  CAS  Google Scholar 

  • Gray CM, Maldonado PE, Wilson M, McNaughton B (1995) Tetrodes markedly improve the reliability and yield of multiple single-unit isolation from multi-unit recordings in cat striate cortex. J Neurosci Methods 63:43–54

    PubMed  CAS  Google Scholar 

  • Grinvald A, Arieli A, Tsodyks M, Kenet T (2003) Neuronal assemblies: single cortical neurons are obedient members of a huge orchestra. Biopolymers 68:422–436

    PubMed  CAS  Google Scholar 

  • Habel Ch (1994) Discreteness, finiteness, and the structure of topological spaces. In: Eschenbach C, Habel Ch, Smith B (eds) Topological foundations of cognitive science. Report 37. Graduiertenkolleg Kognitionswissenschaft Hamburg, Hamburg, pp 81–90

  • Haig AR, Gordon E, De Pascalis V, Meares RA, Bahramali H, Harris A (2000) Gamma activity in schizophrenia: evidence of impaired network binding? Clin Neurophysiol 111:1461–1468

    PubMed  CAS  Google Scholar 

  • Haken H (1996) Principles of brain functioning: a synergetic approach to brain activity, behavior and cognition. Springer, Berlin Heidelberg New York

    Google Scholar 

  • Harter MR (1967) Excitability cycles and cortical scanning: a review of two hypothesis of cortical intermittency in perception. Psychol Bull 68:47–55

    PubMed  CAS  Google Scholar 

  • Hasty J, Collins JJ, Wiesenfeld K, Grigg P (2001) Wavelets of excitability in sensory neurons. J Neurophysiol 86:2097–2101

    PubMed  CAS  Google Scholar 

  • Hebb DO (1949) The organization of behavior. Wiley, New York

    Google Scholar 

  • Hill C (1991) Sensations: a defense of type materialism. Cambridge University Press, New York

  • Hirsh IJ, Sherrick CEJ (1961) Perceived order in different sense modalities. J Exp Psychol 62:423–432

    PubMed  CAS  Google Scholar 

  • Hobson JA (1992) A new model of brain-mind state: activation level, input source, and mode of processing (AIM). In: Antrobus J, Bertini M (eds) The neuropsychology of dreaming sleep. Lawrence Erlbaum Associates, Hillsdale

  • Hobson JA, Pace-Schott EF, Stickgold R (2000) Dreaming and the brain: toward a cognitive neuroscience of conscious states. Behav Brain Sci 23:793–842

    PubMed  CAS  Google Scholar 

  • Hwa RC, Ferree T (2002) Scaling properties of fluctuations in the human electroencephalogram. Phys Rev E Stat Nonlin Soft Matter Phys 66:021901

    Google Scholar 

  • Ivancich JE, Huyck CR, Kaplana S (1999) Cell assemblies as building blocks of larger cognitive structures. Behav Brain Sci 22:292–293

    Google Scholar 

  • Izhikevich EM (1999) Class 1 neural excitability, conventional synapses, weakly connected networks, and mathematical foundations of pulse-coupled models. IEEE Trans Neural Netw 10:499–507

    PubMed  CAS  Google Scholar 

  • James W (1890) The principles of psychology. Vol. I. Dover, New York

  • Jansen BH, Cheng WK (1988) Structural EEG analysis: an explorative study. Int J Biomed Comput 23:221–237

    PubMed  CAS  Google Scholar 

  • Jeannerod M (1994) The representing brain: neural correlates of motor intention and imagery. Behav Brain Sci 17:187–245

    Google Scholar 

  • John ER (1990) Machinery of the mind. Birkhauser, Boston

    Google Scholar 

  • John ER (2001) A field theory of consciousness. Conscious Cogn 9–10:184–213

    Google Scholar 

  • John ER (2002) The neurophysics of consciousness. Brain Res Brain Res Rev 39:1–28

    PubMed  Google Scholar 

  • Kahneman D (1978) Attention and effort. Prentice-Hall, Englewood Cliffs

    Google Scholar 

  • Kaplan AYa (1995) On the frame architecture of central information processing: EEG analysis. In: The fourth IBRO World congress of neuroscience, p 438

  • Kaplan AYa (1998) Nonstationary EEG: methodological and experimental analysis. Usp Fiziol Nauk (Success in Physiological Sciences) 29:35–55 (in Russian)

  • Kaplan AYa (1999) The problem of segmental description of human electroencephalogram (translated from Physiol Cheloveka). Human Physiol 25:107–114

    Google Scholar 

  • Kaplan AYa, Shishkin SL (2000) Application of the change-point analysis to the investigation of the brain’s electrical activity. Chapter 7. In: Brodsky BE, Darkhovsky BS (eds) Nonparametric statistical diagnosis: problems and methods. Kluwer, Dordrecht, pp 333–388

  • Kaplan AYa, Borisov SV (2002) The differences in structural synchrony of the brain electrical field in alpha range between normal control and schizophrenic adolescents. Human Brain Mapping Meeting (Sendai, Japan, 2002). Poster No.: 10472. NeuroImage No 329

  • Kaplan AYa, Borisov SV (2003) Dynamic properties of segmental characteristics of EEG alpha activity in rest conditions and during cognitive load (in Russian). Zh Vyssh Nerv Deiat Im IP Pavlova (IP Pavlov J High Nerv Act) 53:22–32

    Google Scholar 

  • Kaplan AYA, Fingelkurts ALA, Fingelkurts ANA, Darkhovsky BS (1997) Topological mapping of sharp reorganization synchrony in multichannel EEG. Am J Electroneurodiagnostic Technol 37:265–275

    Google Scholar 

  • Kaplan AYA, Fingelkurts ANA, Fingelkurts ALA, Borisov SV, Darkhovsky BS (2005) Nonstationary nature of the brain activity as revealed by EEG/MEG: methodological, practical and conceptual challenges. Signal Process 85:2190–2212

    Google Scholar 

  • Kallio S, Revonsuo A (2003) Hypnotic phenomena and altered states of consciousness: a multilevel framework of description and explanation. Contemp Hypn 20:111–164

    Google Scholar 

  • Kelso JAS (1991) Behavioral and neural pattern generation: the concept of neurobehavioral dynamical system (NBDS). In: Koepchen HP (ed) Cardiorespiratory and motor coordination. Springer, Berlin Heidelberg New York

  • Kelso JAS (1995) Review of dynamic patterns: the self-organization of brain and behavior. MIT, Cambridge

    Google Scholar 

  • Kinoshita T, Strik WK, Michel CM, Yagyu T, Saito M, Lehmann D (1995) Microstate segmentation of spontaneous multichannel EEG map series under diazepam and sulpiride. Pharmacopsychiatry 28:51–55

    Article  PubMed  CAS  Google Scholar 

  • Kirillov AB, Makarenko VI (1991) Metastability and phase transition in neural networks: statistical approach. In: Holden AV, Kryukov VI (eds). Neurocomputers and attention, Vol. 2. Manchester University Press, Manchester, pp 825–922

  • Koenig T, Lehmann D (1996) Microstates in language-related brain potentials show noun-verb differences. Brain Lang 53:169–182

    PubMed  CAS  Google Scholar 

  • Koenig T, Prichep L, Lehmann D, Sosa PV, Braeker E, Kleinlogel H, Isenhart R, John ER (2002) Millisecond by millisecond, year by year: normative EEG microstates and developmental stages. NeuroImage 16:41–48

    PubMed  Google Scholar 

  • Köhler W (1940) Dynamics in psychology. Grove Press, New York

    Google Scholar 

  • Korb KB (1993) Stage effects in the Cartesian theater: a review of Daniel Dennett’s consciousness explained. PSYCHE 1(4), December 1993, URL = http://www.psyche.cs.monash.edu.au/v1/psyche-1–04-korb.html

  • Kosslyn SM (1975) Information representation in visual images. Cognit Psychol 7:341–370

    Google Scholar 

  • Kristofferson AB (1967) Successiveness discrimination as a two-state, quantal process. Science 158:1337–1339

    PubMed  CAS  Google Scholar 

  • Landa P, Gribkov D, Kaplan A (2000) Oscillatory processes in biological systems. In: Malik SK, Chandrashekaran MK, Pradhan N (eds) Nonlinear phenomena in biological and physical sciences. Indian National Science Academy, New Deli, pp 123–152

    Google Scholar 

  • Laskaris NA, Ioannides AA (2001) Exploratory data analysis of evoked response single trials based on minimal spanning tree. Clin Neurophysiol 112:698–712

    PubMed  CAS  Google Scholar 

  • Latour PL (1967) Evidence of internal clocks in the human operator. Acta Psychol 27:341–348

    CAS  Google Scholar 

  • Lehar S (2003) Gestalt isomorphism and the primacy of subjective conscious experience: a gestalt bubble model. Behav Brain Sci 26:375–408

    PubMed  Google Scholar 

  • Lehmann D (1971) Multichannel topography of human alpha EEG fields. Electroencephalogr Clin Neurophysiol 31:439–449

    PubMed  CAS  Google Scholar 

  • Lehmann D, Koenig T (1997) Spatio-temporal dynamics of alpha brain electric fields, and cognitive modes. Int J Psychophysiol 26:99–112

    PubMed  CAS  Google Scholar 

  • Lehmann D, Ozaki H, Pal I (1987) EEG alpha map series: brain micro-states by space oriented adaptive segmentation. Electroencephalogr Clin Neurophysiol 67:271–288

    PubMed  CAS  Google Scholar 

  • Lehmann D, Wackermann J, Michel CM, Koenig T (1993) Space-oriented EEG segmentation reveals changes in brain electric field maps under the influence of a nootropic drug. Psychiatry Res 50:275–282

    PubMed  CAS  Google Scholar 

  • Lehmann D, Strik WK, Henggeler B, Koenig T, Koukkou M (1998) Brain electrical micro-states and momentary conscious mind states as building blocks of spontaneous thinking. I. Visual imagery and abstract thoughts. Int J Psychophysiol 29:1–11

    PubMed  CAS  Google Scholar 

  • Lennie P (1998) Single units and visual cortical organization. Perception 27:889–935

    PubMed  CAS  Google Scholar 

  • Leznik E, Makarenko V, Llinas R (2002) Electrotonically mediated oscillatory patterns in neuronal ensembles: an in vitro voltage-dependent dye-imaging study in the inferior olive. J Neurosci 22:2804–2815

    PubMed  Google Scholar 

  • Libet B, Wright EW, Feinstein B, Pearl DK (1979) Subjective referral of the timing for a conscious sensory experience. Brain 102:193–224

    PubMed  CAS  Google Scholar 

  • Libet B, Gleason CA, Wright EW, Pearl DK (1983) Time of conscious intention to act in relation to onset of cerebral activity (readiness potential): the unconscious initiation of a freely voluntary act. Brain 106:623–642

    PubMed  Google Scholar 

  • Linkenkaer-Hansen K, Nikouline VM, Palva JM, Iimoniemi RJ (2001) Long-range temporal correlations and scaling behavior in human brain oscillations. J Neurosci 15:1370–1377

    Google Scholar 

  • Llinas R (1988) The intrinsic electrophysiological properties of mammalian neurons: insights into central nervous system function. Science 242:1654–1664

    PubMed  CAS  Google Scholar 

  • Llinas R, Ribary U (1998) Temporal conjunction. In: Thalamocortical transactions. Consciousness: at the frontiers of neuroscience (Advances in Neurology), Vol. 77. Lippincott-Raven, Philadelphia, pp 213–217

  • Llinas R, Ribary U, Contreras D, Pedroarena C (1998) The neuronal basis for consciousness. Philos Trans R Soc Lond B Biol Sci 353:1841–1849

    PubMed  CAS  Google Scholar 

  • Llinas R, Leznik E, Urbano FJ (2002) Temporal binding via cortical coincidence detection of specific and nonspecific thalamocortical inputs: A voltage-dependent dyeimaging study in mouse brain slices. Proc Natl Acad Sci USA 99:449–454

    PubMed  CAS  Google Scholar 

  • Luria AR (1980) Higher cortical functions in man. Kluwer, Dordrecht

    Google Scholar 

  • Lutz A, Lachaux J-P, Martinerie J, Varela JF (2002) Guiding the study of brain dynamics by using first-person data: synchrony patterns correlate with ongoing conscious states during a simple visual task. Proc Natl Acad Sci USA 99:1586–1591

    PubMed  CAS  Google Scholar 

  • Makarenko V, Llinas R (1998) Experimentally determined chaotic phase synchronization in a neuronal system. Proc Natl Acad Sci USA 95:15747–15752

    PubMed  CAS  Google Scholar 

  • Makarenko VI, Welsh JP, Lang EJ, Llinas R (1997) A new approach to the analysis of multidementional neural activity: Markov random fields. Neural Netw 10:785–789

    PubMed  Google Scholar 

  • Mangan BB (1993a) Taking phenomenology seriously: the “fringe” and its implications for cognitive research. Conscious Cogn 2:89–108

    Google Scholar 

  • Mangan BB (1993b) Some philosophical and empirical implications of the fringe. Conscious Cogn 2:142–154

    Google Scholar 

  • McGurk H, MacDonald JW (1976) Hearing lips and seeing voices. Nature 264:746–748

    PubMed  CAS  Google Scholar 

  • McIntosh AR (1999) Mapping cognition to the brain through neural interactions. Memroy 7:523–548

    CAS  Google Scholar 

  • McIntosh AR, Fitzpatrick SM, Friston KJ (2001) On the marriage of cognition and neuroscience. NeuroImage 14:1231–1237

    PubMed  CAS  Google Scholar 

  • Medison G (2001) Functional modeling of the human timing mechanism. PhD Dissertation, Acta Universitatis Upsaliensis, Uppsala, Sweden, 77 pp

  • Meyer DE, Yantis S, Osman AM, Smith JEK (1985) Temporal properties of human information processing: tests of discrete versus continuous models. Cognit Psychol 17:445–518

    PubMed  CAS  Google Scholar 

  • Metzinger T (1995) Faster than thought. Holism, homogeneity and temporal coding. In: Metzinger T (ed) Conscious experience. Imprint Academic and Paderborn, Thorverton, UK, URL = http://www.imprint.co.uk/online/Metz1.html

  • Michon JA (1985) The complete time experiencer. In: Michon JA (ed) Time, mind, and behavior. Springer, Berlin Heidelberg New York, pp 20–52

    Google Scholar 

  • Müller GE (1896) Zur psychophysik der gesichtsempfindungen. Z Psychol 10:1–82

    Google Scholar 

  • Naish P (2001) Hypnotic time perception: Busy beaver or tardy timekeeper. Contemp Hypn 18:87–99

    Google Scholar 

  • Newman J (1995) Thalamic contributions to attention and consciousness. Conscious Cogn 4:172–193

    PubMed  CAS  Google Scholar 

  • Noë A, Thompson E (2004) Are there neural correlates of consciousness? J Conscious Stud 11:3–28

    Google Scholar 

  • Nunez PL (1995) Neocortical dynamics and human EEG rhythms. Oxford University Press, New York

    Google Scholar 

  • Nunez PL (2000) Toward a quantitative description of large-scale neocortical dynamic function and EEG. Behav Brain Sci 23:371–398

    PubMed  CAS  Google Scholar 

  • O’Brien G, Opie J (1999) A connectionist theory of phenomenal experience. Behav Brain Sci 22:127–148

    PubMed  CAS  Google Scholar 

  • Pascual-Marqui R, Michel C, Lehmann D (1995) Segmentation of brain electrical activity into microstates. IEEE Trans Biomed Eng 42:658–665

    PubMed  CAS  Google Scholar 

  • Pelliomisz A, Llinas R (1985) Tenzor network theory of the metaorganization of fanctional geometries in the central nervous system. Neuroscience 16:245–273

    Google Scholar 

  • Pöppel E (1988) Mindworks: time and conscious experience. Harcourt Brace Jovanovich, Boston

    Google Scholar 

  • Pöppel E (1996) Reconstruction of subjective time on the basis of hierarchically organized processing system. In: Pastor MA, Arteida J (eds) Time, internal clocks and movement. Elsevier, New York, pp 165–185

    Google Scholar 

  • Pöppel E (1997) A hierarchical model of temporal perception. Trends Cogn Sci 1:56–61

    Google Scholar 

  • Pouget A, Dayan P, Zemel R (2000) Information processing with population codes. Nat Rev Neurosci 1:125–132

    PubMed  CAS  Google Scholar 

  • Posner MI (1987) Chronometric exploration of mind. Erlbaum, Hillsdale

    Google Scholar 

  • Poznanski RR (2002) Dendritic integration in a recurrent network. J Integr Neurosci 1:69–99

    PubMed  Google Scholar 

  • Pylyshyn ZW (2002) Mental imagery: in search of a theory. Behav Brain Sci 25:157–238

    PubMed  Google Scholar 

  • Pulvermüller F (1999) Words in the brain’s language. Behav Brain Sci 22:253–336

    PubMed  Google Scholar 

  • Purpura DP (1972) Functional studies of thalamic internuclear interactions. Brain Behav 6:203–234

    CAS  Google Scholar 

  • Purushothaman G, Patel SS, Bedell HE, Ögmen H (1998) Moving ahead through differential visual latency. Nature 396:424

    PubMed  CAS  Google Scholar 

  • Rensing L, Meyer-Grahle U, Ruoff P (2001) Biological timing and the clock metaphor: oscillatory and hourglass mechanisms. Chronobiol Int 18:329–369

    PubMed  CAS  Google Scholar 

  • Revonsuo A (1993) Dennett and dissociations of consciousness. Psycoloquy: 4(59), Split Brain (4), URL = http://www.psycprints.ecs.soton.ac.uk/archive/00000353/

  • Revonsuo A (2000) Prospects for a scientific research program on consciousness. In: Metzinger T (ed) Neural correlates of consciousness. MIT, Cambridge, pp 57–75

    Google Scholar 

  • Revonsuo A (2001) Can functional brain imaging discover consciousness in the brain? J Conscious Stud 8:3–23

    Google Scholar 

  • Reynolds JH, Desimone R (1999) The role of neural mechanisms of attention in solving the binding problem. Neuron 24:19–29, 111–125

    Google Scholar 

  • Robins C, Shepard RN (1977) Spatio-temporal probing of apparent rotational movement. Percept Psychophys 22:12–18

    Google Scholar 

  • Rodriguez E, George N, Lachaux JP, Martinerie J, Renault B, Varela FJ (1999) Perception’s shadow: long-distance synchronization of human brain activity. Nature 397:430–433

    PubMed  CAS  Google Scholar 

  • Sahraie A, Weiskrantz L, Barbur IL, Simmone A, Williams SC, Brammer MJ (1997) Pattern of neocortical activity associated with conscious and unconscious processing of visual signals. Proc Natl Acad Sci USA 94:9406–9411

    PubMed  CAS  Google Scholar 

  • Seidemann E, Meilijson I, Abeles M, Bergman H, Vaadia E (1996) Simultaneously recorded single units in the frontal cortex go through sequences of discrete and stable states in monkeys performing a delayed localization task. J Neurosci 16:752–768

    PubMed  CAS  Google Scholar 

  • Shallice T (1964) The detection of change and the perceptual moment hypothesis. Br J Stat Psychol 17:113–135

    Google Scholar 

  • Shannon C (1948/1949) A mathematical theory of communication. In: Shannon C, Weaver W (eds). The mathematical theory of communication. University of Illinois Press, Urbana, IL, pp 623–656

  • Shevelev IA, Kostelianetz NB, Kamenkovich VM, Sharaev VA (1991) EEG alpha-wave in the visual cortex: check of the hypothesis of the scanning process. Int J Psychophysiol 11:195–201

    PubMed  CAS  Google Scholar 

  • Shevelev IA, Kamenkovich VM, Bark ED, Verkhlutov VM, Sharaev VA, Mikhailova ES (2000) Visual illusion and traveling alpha waves produced by flicker at alpha frequencies. Int J Psychophysiol 39:9–20

    PubMed  CAS  Google Scholar 

  • Shishkin SL, Darkhovsky BS, Fingelkurts AlA, Fingelkurts AnA, Kaplan AYa (1998) Interhemispheric synchrony of short-term variations in human EEG alpha power correlates with self-estimates of functional state. In: Proceedings of ninth world congress of psychophysiology, Tvaormin, Sicily/Italy, pp 133

  • Shepard RN, Metzler J (1971) Mental rotation of three-dimensional objects. Science 191:701–703

    Google Scholar 

  • Singer W (2001) Consciousness and the binding problem. Ann N Y Acad Sci 929:123–146

    Article  PubMed  CAS  Google Scholar 

  • Singer W, Engel AK, Kreiter AK, Munk MHJ, Neuenschwander S, Roelfsema PR (1997) Neural assemblies: necessity, signature and detectability. Trends Cogn Sci 1:252–261

    Google Scholar 

  • Skinner JE, Molnar M (2000) “Response cooperativity”: a sign of a nonlinear neocortical mechanism for stimulus-binding during classical conditioning in the act. In: Malik SK, Chandrashekaran MK, Pradhan N (eds) Nonlinear phenomena in biological and physical sciences. Indian National Science Academy, New Deli, pp 223–248

    Google Scholar 

  • Sternberg A (1969) The discovery of processing stages: extansions of Donders’ method. Acta Psychol 30:276–315

    Google Scholar 

  • Strik WK, Lehmann D (1993) Data-determined window size and space-oriented segmentation of spontaneous EEG map series. Electroencephalogr Clin Neurophysiol 87:169–174

    PubMed  CAS  Google Scholar 

  • Stroud JM (1955) The fine structure of psychological time. In: Quastler H (ed) Information theory in psychology: problems and methods. The Free Press, Glencoe, Ill, pp 174–205

    Google Scholar 

  • Suber P (1988) What is software? J Speculative Philos 2:89–119

    Google Scholar 

  • Surwillo WW (1966) Time perception and the “internal clock”: some observations on the role of the electroencephalogram. Brain Res 2:390–392

    PubMed  CAS  Google Scholar 

  • Taylor K (2001) Applying continuous modelling to consciousness. J Conscious Stud 8:45–60

    Google Scholar 

  • Tononi G, Edelman GM (1998) Consciousness and complexity. Science 282:1846–1851

    PubMed  CAS  Google Scholar 

  • Tonnelier A (2005) Categorization of neural excitability using threshold models. Neural Comput 17:1447–1455

    PubMed  CAS  Google Scholar 

  • Treisman A, Gelade G (1980) A feature-integration theory of attention. Cognit Psychol 12:97–136

    PubMed  CAS  Google Scholar 

  • Treisman M (1963) Temporal discrimination and the indifference interval: implications for a model of the “internal clock.” Psychol Monogr 77:1–31

    PubMed  CAS  Google Scholar 

  • Treisman M (1984) Temporal rhythms and cerebral rhythms. Ann N Y Acad Sci 423:542–565

    PubMed  CAS  Google Scholar 

  • Treisman M, Faulker A, Naish PL, Brogan D (1990) The internal clock: evidence for a temporal oscillator underlying time perception with some estimates of its characteristic frequency. Perception 19:705–743

    PubMed  CAS  Google Scholar 

  • Treisman M, Cook N, Naish PL, MacCrone JK (1994) The internal clock: electroencephalographic evidence for oscillatory processes underlying time perception. Q J Exp Psychol A 47:241–289

    PubMed  CAS  Google Scholar 

  • Triesch J, von der Malsburg C (2001) Democratic integration: self-organized integration of adaptive cues. Neural Comput 13:2049–2074

    PubMed  CAS  Google Scholar 

  • Truccolo WA, Ding M, Knuth KH, Nakamura R, Bressler S (2002) Trial-to-trial variability of cortical evoked responses: implications for analysis of functional connectivity. Clin Neurophysiol 113:206–226

    PubMed  Google Scholar 

  • Tsuda I (2001) Toward an interpretation of dynamic neural activity in terms of chaotic dynamical systems. Behav Brain Sci 24:793–810

    Article  PubMed  CAS  Google Scholar 

  • Uhr L (1994) Digital and analog microcircuit and sub-net structures for connectionist networks. In: Honavar V, Uhr L (eds) Artificial intelligence and neural networks: steps toward principled integration. Academic, Boston, pp 341–370

    Google Scholar 

  • Vaadia E, Haalman I, Abeles M, Bergman H, Prut Y, Slovin H, Aertsen A (1995) Dynamics of neuronal interactions in monkey cortex in relation to behavioural events. Nature 373:515–518

    PubMed  CAS  Google Scholar 

  • Vanagas V (1994) Active, hierarchical visual system organization and attentional information processing. In: 39. Internationales Wissenschaftliches Kolloquium. Tehnische Universitat Ilmenau, Thüringen, Band 2, pp 91–94

  • van Gelder T (1995) What might cognition be, if not computation? J Philos XCI:345–381

    Google Scholar 

  • VanRullen R, Koch C (2003) Is perception discrete or continuous? Trends Cogn Sci 7:207–213

    PubMed  Google Scholar 

  • Varela FJ (2000) Neural synchrony and consciousness: are we going somewhere? Conscious Cogn 9:S26–S27

    Google Scholar 

  • Varela FJ, Toro A, John ER, Schwartz EL (1981) Perceptual framing and cortical alpha rhythm. Neuropsychologia 19:675–686

    PubMed  CAS  Google Scholar 

  • Varela FJ (1995) Resonant cell assemblies: a new approach to cognitive functions and neuronal synchrony. Biol Res 28:81–95

    PubMed  CAS  Google Scholar 

  • Varela F, Lachaux J-P, Rodriguez E, Martinerie J (2001) The brainweb: phase synchronization and large-scale integration. Nat Rev Neurosci 2:229–239

    PubMed  CAS  Google Scholar 

  • Vartanyan GA, Pirogov AA, Konstantinov KV (1989) Changes produced in neuronal excitability by subthreshold depolarization as a possible mechanism of interval selective relationships within the central nervous system. Neurophysiol 21:201–208

    Google Scholar 

  • Velmans M (2002) How could conscious experience affect brains? J Conscious Stud 9:3–29

    Google Scholar 

  • Venables PH (1960) Periodicity in reaction time. Br J Psychol 51:37–43

    PubMed  CAS  Google Scholar 

  • von Baer KE (1864) Welche Auffasung der lebendigen Natur ist die richtige? Und wie ist diese Auffasung auf die Entomologie auzuwenden? In: Schmitzdorff H (ed) Reden gehalten in wiss. Versammlungen und kleine Aufsätze vermischten Inhalts. Verlag der kaiserl, Hofbuchhandlung, St. Petersburg, pp 237–287

    Google Scholar 

  • von der Malsburg C (1981) The correlation theory of brain function. Max-Planck-Institut für Biophysikalische Chemie, Postfach 2841, D-3400 Göttingen, FRG

  • von der Malsburg C (1997) The coherence definition of consciousness. In: Ito M, Miyashita Y, Rolls ET (eds) Cognition, computation and consciousness. Oxford University Press, Oxford, pp 193–204

    Google Scholar 

  • von der Malsburg C (1999) The what and why of binding: the modeler’s perspective. Neuron 24:95–104

    PubMed  Google Scholar 

  • von Rospatt A (1995) The buddhist doctrine of momentariness: a survey of the origins and early phase of this doctrine up to Vasubandhu. Franz Steiner Verlag, Stuttgart

  • Warfield JN (1977) Crossing theory and hierarchy mapping. IEEE Trans Syst Man Cybern 7:505–523

    Article  Google Scholar 

  • Watson C (2003) MRI cytoarchitectonics: the next level? J Neurol Sci 211:1–3

    PubMed  Google Scholar 

  • White C, Harter MR (1969) Intermittency in reaction time and perception, and evoked response correlates of image quality. Acta Psychol 30:368–377

    CAS  Google Scholar 

  • Wiener N (1961) Cybernetics: or control and communication in the animal and the machine, 2nd edn. MIT, MA

    Google Scholar 

  • Wright JJ, Robinson PA, Rennie CJ, Gordon E, Bourke PD, Chapman CL, Hawthorn N, Lees GJ, Alexander D (2001) Toward an integrated continuum model of cerebral dynamics: the cerebral rhythms, synchronous oscillation and cortical stability. Biosystems 63:71–88

    PubMed  CAS  Google Scholar 

  • Zeki S (2003) The disunity of consciousness. Trends Cogn Sci 7:214–218

    PubMed  Google Scholar 

  • Zhou YD, Fuster JM (2000) Visuo-tactile cross-modal associations in cortical somatosensory cells. Proc Natl Acad Sci USA 97:9777–9782

    PubMed  CAS  Google Scholar 

Download references

Acknowledgments

The authors are grateful for stimulating discussions on related questions to Prof. Alexander Kaplan, Prof. Walter Freeman, Prof. Erol Basar, Prof. Hermann Haken, Prof. Steve Bressler, and Mr. Carlos Neves (Computer Science specialist). Conversations with Prof. Antti Revonsuo about neurophysiology of consciousness have had a significant influence on the ideas in this paper. We wish also to thank Prof. William Banks and Prof. Max Velmans for their very useful comments on the earlier version of this paper. The writing of this paper has been supported by the BM-SCIENCE. Special thanks to Prof. Richard Lippa for skilful text editing.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andrew A. Fingelkurts.

Appendices

Appendix 1: OA and other related theoretical frameworks

The OA theoretical framework has a marked resemblance to other theoretical accounts currently dominating the field of research. This is not the place for a detailed discussion of the similarities and differences between various theoretical frameworks, since the scope of the present paper is limited to the “continuity–discreteness” problem. Therefore, we will touch this subject very briefly. The first theory which we want to address is the “Global Workspace (GW)” framework (Baars 1988). According to this theory, the brain seems to show a distributed style of functioning, in which the detailed work is done by millions of specialized neural groupings without specific instructions from some command centre. Mostly these are unconscious processes; however, consciousness creates widespread access (free from interference) to complex and unconscious systems. Using “theater metaphor,” Baars argued that as theatre combines very limited events taking place on stage with a vast audience, consciousness involves limited information that creates access to a vast number of unconscious sources of knowledge (Baars 1997).

Generally, the OA framework is also consistent with the “Framework for Consciousness” suggested by Crick and Koch (2003). The main lines of correspondence are the following: a preamble on the cerebral cortex; the fact, that most cortical areas is sensitive to correlations among correlations being expressed by other cortical areas; the importance of neuronal assemblies; the claim that conscious awareness is a series of discrete snapshots and that the durations of successive snapshots are not constant (Crick and Koch 2003).

Further, our theoretical framework is also compatible with Chalmers’ “Philosophy of Mind” (2002) in the part that any distinction in experience should be mirrored by a distinction in neural activity, and the pattern of experience should be matched by the pattern of awareness (see “functional isomorphism” in Fingelkurts and Fingelkurts 2001, 2003, 2004).

It is worth to note that OA framework is in vein with Revonsuo’ “Neuroconsciousness” conception (2001). According to this framework, consciousness is a real biological phenomenon that is physically located within the brain—it is the phenomenal level of brain organization (Revonsuo 2000). Further, it is suggested that physiologically neuroconsciousness consists of large-scale electrophysiological (or bioelectrical) activity spatio-temporal patterns; and synchrony in these patterns may be the mechanism by which the conscious state and its contents are realized in the brain. Further, it has been proposed that “neural synchrony seems to be capable of supporting higher-level electrophysiological entities that resemble the content of the phenomenal level of organization” (Revonsuo 2001, p 6).

Besides just mentioned theories, there are two related theoretical frameworks which have the closest similarities with the OA conceptualization. These are (a) the “resonant cell assemblies” framework developed by Varela (1995) and (b) the “dynamic core” (DC) theory proposed by Tononi and Edelman (1998). Because of limited space, we summarize the similarities and differences between these theoretical frameworks and our OA theory in Table 1. One can notice that besides similarities between all three theories, the OA framework has several notable differences. We will concentrate here only on the most relevant ones.

Table 1 Comparison of OA framework with resonant cell assemblies and dynamical core frameworks
  1. 1.

    Neither the NA, nor the DC is flexible enough to allow for a representation of complex objects or for the execution of complex combinatorial cognitive operations, which are also the gist of their internal structure. This is so because NA and DC do not have internal hierarchical organizations. Here, it is essentially necessary to allow for hierarchical organization with the structured integration of subcomponents (von der Malsburg 1999). The components in question are often necessarily activated under the same overall conditions; hence without defined internal structure a NA/DC could not distinguish between the two (or more) types of events. In contrast, OMs, which are the main constituents of the OA framework, do have such internal organization (see Fingelkurts and Fingelkurts 2003, 2005): one OM may be a member of another more complex one, or it may be decomposed until simple neuronal assemblies, each of which would be responsible for simple brain/cognitive operations. Therefore, the recombination of subsets of neuronal assemblies into OMs, and of different OMs into larger structured OMs yields a vast number of potential combinations needed to represent the multivariability of cognition and eventually consciousness (Fingelkurts and Fingelkurts 2004). Such complex structure of OM is also important for the semantic representations of words with similar meanings, for example, for hyponyms and hyperonyms (Pulvermüller 1999). For instance, it can be shown that between-assembly connections and activity dynamics are a possible basis of semantic associations and/or grammatical phenomena (see Ivancich et al. 1999; Pulvermüller 1999; Fingelkurts and Fingelkurts 2001, 2003 for further discussion).

  2. 2.

    In contrast to NA/DC models, the OA framework supposes large number of coexisting OMs. Considering the composite polyphonic character of the electrical brain field (EEG), this field may be presented as a mixture of many time-scale processes (individual frequency components) (Nunez 2000; Basar et al. 2001; Basar 2004). Consequently, a large amount of functionally distinct OMs can co-exist simultaneously at different time-scales and even between them (Kaplan and Shishkin 2000; Fingelkurts and Fingelkurts 2001; here there are no restrictions for the relations between frequency bands, because the method we used for assessing the OMs is not associated with the phase relation as the usual techniques estimating synchrony). Simultaneous existence of these OMs subserves the numerous operational acts on the functioning of the brain/organism and on the interaction of the organism with its environment (Arbib 2001). Only subset of these OMs constitutes mental states, some of which are of conscious nature (see Fingelkurts and Fingelkurts 2001, 2003, 2005 for further discussion).

  3. 3.

    Even though all three theoretical frameworks stress the importance of functional connections, the concepts they used to define the values of functional connectivity differ significantly between them. This subject is discussed in a grate detail in our previous publications; therefore we address interested reader to them (Fingelkurts et al. 2005; Fingelkurts and Fingelkurts 2005). Here we should only mention that OA framework supposes “true” functional synchrony which not necessarily requires any anatomical connections. It is the stimuli (either external or internal), the task, and other functions which cause the synchronization; therefore, it is a function-based synchronization (Fingelkurts and Fingelkurts 2005).

  4. 4.

    The OMs in contrast to NA/DC are characterized by the metastable nature. Attention, we speak here not about dynamics of OM/NA/DC which is also metastable, but about the functional entity (OM/NA/DC) pre se. As we have already mentioned in the main body of the text, the OMs are inherently metastable since they constructed by separate neuronal assemblies which process and represent different types of information from relatively independent brain functional systems; however, at the same time they exhibit tendency for the coordinated activity (Fingelkurts and Fingelkurts 2004, 2005). Such simultaneous existence of autonomous and coordinated tendencies is the essence of the metastable regime of system (brain) functioning (Kelso 1991, 1995).

  5. 5.

    NA/DC are lacking of the time-dimensional information in each cortical area separately, while the OM is based on the detailed and known time-dimensional information in each cortical area.

Appendix 2: The methodological problems with cell assembly model

The claim held by many researches that cell assembly framework cannot be easily falsified may be the basis for a premature rejection of many cell assembly based theories (as the OA theory) that offer a neurobiologically plausible framework within which so much can actually be explained. The existence of cell assemblies could, in principle, be tested by recording, in parallel, multiple neurons whose activity is correlated with different cognitive functions or conscious experience. Multielectrode recordings have already indicated that rapid changes in the functional connectivity among distributed populations of neurons can occur independently of firing rate (Vaadia et al. 1995). Furthermore, recent studies in monkey frontal cortex show abrupt and simultaneous shifts among stationary activity states involving many, but not all recorded neurons (Seidemann et al. 1996), thus clearly indicating the functional clustering of neurons into neuronal assemblies.

However, such studies have several methodological limitations. For example, understanding of how the cooperated activity of neurons gives rise to collective assembly behavior requires improved methods for simultaneous recording with minimal damage to the neuronal tissue (Buzsáki 2004). Another important step in multielectrode recording analysis is the isolation of single neurons on the basis of extracellular features. Several methods exist, however they based on the assumptions which are difficult to justify in most cases (Llinas 1988; Gray et al. 1995). Yet another difficulty is that no independent criteria are available for the assessment of unit isolation (Buzsáki 2004) and therefore, inter-laboratory comparison is difficult and makes interpretation of the results controversial. Additionally, it is often not known what a given cell assembly is coding for: there is considerable evidence that the firing pattern of neurons even in primary sensory cortices reflects not just the physical nature of a stimulus, but also internal factors (Zhou and Fuster 2000).

Additionally, cell assemblies widely distributed over distant cortical regions are obviously difficult to observe through electrophysiological recordings from local neuron clusters or small areas. If large-scale neuronal theories of cognitive functions are correct, then fast, large-scale recording techniques are necessary to visualize activity changes in distributed assemblies. Fortunately, several such new techniques have been already emerged (see for example, Grinvald et al. 2003; Buzsáki 2004; Bennett and Zukin 2004) and therefore there is a hope that the question of whether there are specific cell assemblies for different cognitive operations will be soon answered experimentally (and therefore will give bases for possible falsification of cell assembly based theories).

We should stress also, that among already existed large-scale techniques, EEG and MEG approaches can be extremely useful when new methods of their analysis (as described in the present paper) are used for investigating the cortical topographies of neuronal assemblies. A coherent picture can be drawn already on the basis of a number of studies (see Fingelkurts and Fingelkurts 2004, 2005 for the resent review of research).

It is relevant to point out here that the proposed concept of cell assemblies is necessarily fuzzy. We agree with Pulvermüller (1999) that this is not a problem: “It is essential to see that fuzziness is intrinsic to the assembly concept and that this is only problematic in the way it is a problem to determine the boundaries of the sun or the Milky Way” (Pulvermüller 1999, p 310). What is important—it is functional discreteness of cell assemblies (Braitenberg 1980; Braitenberg and Pulvermüller 1992). Exactly this main property of neuronal assemblies is used in the OA framework (see Fingelkurts and Fingelkurts 2005).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Fingelkurts, A.A., Fingelkurts, A.A. Timing in cognition and EEG brain dynamics: discreteness versus continuity. Cogn Process 7, 135–162 (2006). https://doi.org/10.1007/s10339-006-0035-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10339-006-0035-0

Keywords

Navigation