13
Neuronal long-range temporal correlations and avalanche dynamics are correlated with behavioral scaling laws J. Matias Palva a,1,2 , Alexander Zhigalov a,1 , Jonni Hirvonen a , Onerva Korhonen a , Klaus Linkenkaer-Hansen b,3 , and Satu Palva a,c,3 a Neuroscience Center, University of Helsinki, FIN-00014 Helsinki, Finland; b Department of Integrative Neurophysiology, Center for Neurogenomics and Cognitive Research, Neuroscience Campus Amsterdam, VU University Amsterdam, 1081 HV Amsterdam, The Netherlands; and c BioMag Laboratory, HUS Medical Imaging Center, Helsinki University Central Hospital, FIN-00029, Hus, Finland Edited by Marcus E. Raichle, Washington University in St. Louis, St. Louis, MO, and approved January 17, 2013 (received for review October 1, 2012) Scale-free uctuations are ubiquitous in behavioral performance and neuronal activity. In time scales from seconds to hundreds of seconds, psychophysical dynamics and the amplitude uctuations of neuronal oscillations are governed by power-law-form long- range temporal correlations (LRTCs). In millisecond time scales, neuronal activity comprises cascade-like neuronal avalanches that exhibit power-law size and lifetime distributions. However, it remains unknown whether these neuronal scaling laws are corre- lated with those characterizing behavioral performance or whether neuronal LRTCs and avalanches are related. Here, we show that the neuronal scaling laws are strongly correlated both with each other and with behavioral scaling laws. We used source recon- structed magneto- and electroencephalographic recordings to char- acterize the dynamics of ongoing cortical activity. We found robust power-law scaling in neuronal LRTCs and avalanches in resting-state data and during the performance of audiovisual threshold stimulus detection tasks. The LRTC scaling exponents of the behavioral per- formance uctuations were correlated with those of concurrent neuronal avalanches and LRTCs in anatomically identied brain systems. The behavioral exponents also were correlated with neuronal scaling laws derived from a resting-state condition and with a similar anatomical topography. Finally, despite the differ- ence in time scales, the scaling exponents of neuronal LRTCs and avalanches were strongly correlated during both rest and task performance. Thus, long and short time-scale neuronal dynamics are related and functionally signicant at the behavioral level. These data suggest that the temporal structures of human cog- nitive uctuations and behavioral variability stem from the scaling laws of individual and intrinsic brain dynamics. spontaneous activity | threshold detection | criticality H uman cognitive and behavioral performance is highly vari- able and exhibits slow uctuations that are salient in con- tinuous performance tasks (CPTs) (1). Psychophysical time series have been known since the early 1950s to be nonrandomly clus- tered (2), and later studies have shown that hit-rate and/or re- action-time uctuations in CPT data are fractal and power-law autocorrelated across hundreds of seconds (39). The biological origins and relevance of these dynamic, however, remain unclear (10, 11). Similar to those in behavioral performance, the uctuations of collective neuronal activity at many levels of the nervous system are scale-free and governed by power-law scaling laws. On long time scales (10 0 10 3 s), scale-free uctuations and long-range temporal correlations (LRTCs) are salient in the amplitude enve- lopes of spontaneous neuronal oscillations in data recorded with magneto- and electroencephalography (M/EEG) (12). It appears that these oscillation amplitude uctuations reect the un- derlying dynamic architecture of spontaneous brain activity dis- covered with functional MRI (fMRI) and dened by correlated slow uctuations in blood oxygenation leveldependent (BOLD) signals among well-delineated functional brain systems (1315). The oscillation amplitudes are directly correlated with these BOLD uctuations (1620) and exhibit interareal correlations that closely match those of BOLD signals (17, 2123). Moreover, BOLD signals also exhibit scale-free temporal (2426) and spatiotem- poral correlations (2729). The scaling laws of LRTCs thus are a unifying fundamental characteristic of spontaneous brain activity (1, 30, 31). On short time scales (10 3 10 1 s), negative deections in local eld potentials form spatiotemporal cascades of activity, neu- ronal avalanches(3234), the size and lifetime distributions of which are power laws akin to those of a critical branching process (33). Neuronal avalanches characterize spontaneous neuronal network activity in organotypic cultures (32), brain slices in vitro (35), and monkey (34) and human cortex (36) in vivo. In monkey cortex, the avalanches are delimited by cycles of ongoing neuro- nal oscillations (34) showing that in addition to LRTCs (12), neuronal avalanches also coexist with neuronal oscillations. The power-law scaling behavior and fractal properties of neu- ronal LRTCs and avalanches strongly suggest that the brain operates near a critical state (12, 30, 32, 33, 37). Computational modeling predicts that LRTCs and neuronal avalanches are cou- pled (38) and suggests that they coemerge from neuronal inter- actions in a critical regime (30). So far, however, neither the relationship between LRTCs and avalanche dynamics nor their signicance at the behavioral level has been elucidated. We hy- pothesize here that the scaling laws of LRTCs and neuronal ava- lanches are related and are correlated with the interindividual variability in behavioral scaling laws. We test the hypothesis with source reconstructed task- and resting-state M/EEG recordings and show that behavioral scaling laws, LTRCs, and neuronal avalanches are strongly correlated. Results Paradigm for Mapping Individual Behavioral and Neuronal Scaling Laws. As an experimental task that yields psychometric data for probing the scaling laws of cognitive dynamics (7), we used a factorial audiovisual threshold-stimulus detection task (TSDT) in which the subjects were presented independent and continuous auditory and/or visual stimulus streams. The subjects attended Author contributions: J.M.P. and S.P. designed research; A.Z. and J.H. performed research; J.M.P. and O.K. contributed new reagents/analytic tools; J.M.P., A.Z., J.H., and S.P. ana- lyzed data; and J.M.P., A.Z., K.L.-H., and S.P. wrote the paper. The authors declare no conict of interest. This article is a PNAS Direct Submission. 1 J.M.P. and A.Z. contributed equally to this work. 2 To whom correspondence should be addressed. E-mail: matias.palva@helsinki.. 3 K.L.-H. and S.P. contributed equally to this work. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1216855110/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1216855110 PNAS | February 26, 2013 | vol. 110 | no. 9 | 35853590 NEUROSCIENCE

PNAS-2013-Palva-3585-90 (1)

Embed Size (px)

Citation preview

Page 1: PNAS-2013-Palva-3585-90 (1)

Neuronal long-range temporal correlations andavalanche dynamics are correlated withbehavioral scaling lawsJ. Matias Palvaa,1,2, Alexander Zhigalova,1, Jonni Hirvonena, Onerva Korhonena, Klaus Linkenkaer-Hansenb,3,and Satu Palvaa,c,3

aNeuroscience Center, University of Helsinki, FIN-00014 Helsinki, Finland; bDepartment of Integrative Neurophysiology, Center for Neurogenomics andCognitive Research, Neuroscience Campus Amsterdam, VU University Amsterdam, 1081 HV Amsterdam, The Netherlands; and cBioMag Laboratory, HUSMedical Imaging Center, Helsinki University Central Hospital, FIN-00029, Hus, Finland

Edited by Marcus E. Raichle, Washington University in St. Louis, St. Louis, MO, and approved January 17, 2013 (received for review October 1, 2012)

Scale-free fluctuations are ubiquitous in behavioral performanceand neuronal activity. In time scales from seconds to hundreds ofseconds, psychophysical dynamics and the amplitude fluctuationsof neuronal oscillations are governed by power-law-form long-range temporal correlations (LRTCs). In millisecond time scales,neuronal activity comprises cascade-like neuronal avalanches thatexhibit power-law size and lifetime distributions. However, itremains unknown whether these neuronal scaling laws are corre-lated with those characterizing behavioral performance or whetherneuronal LRTCs and avalanches are related. Here, we show thatthe neuronal scaling laws are strongly correlated both with eachother and with behavioral scaling laws. We used source recon-structed magneto- and electroencephalographic recordings to char-acterize the dynamics of ongoing cortical activity. We found robustpower-law scaling in neuronal LRTCs and avalanches in resting-statedata and during the performance of audiovisual threshold stimulusdetection tasks. The LRTC scaling exponents of the behavioral per-formance fluctuations were correlated with those of concurrentneuronal avalanches and LRTCs in anatomically identified brainsystems. The behavioral exponents also were correlated withneuronal scaling laws derived from a resting-state condition andwith a similar anatomical topography. Finally, despite the differ-ence in time scales, the scaling exponents of neuronal LRTCs andavalanches were strongly correlated during both rest and taskperformance. Thus, long and short time-scale neuronal dynamicsare related and functionally significant at the behavioral level.These data suggest that the temporal structures of human cog-nitive fluctuations and behavioral variability stem from the scalinglaws of individual and intrinsic brain dynamics.

spontaneous activity | threshold detection | criticality

Human cognitive and behavioral performance is highly vari-able and exhibits slow fluctuations that are salient in con-

tinuous performance tasks (CPTs) (1). Psychophysical time serieshave been known since the early 1950s to be nonrandomly clus-tered (2), and later studies have shown that hit-rate and/or re-action-time fluctuations in CPT data are fractal and power-lawautocorrelated across hundreds of seconds (3–9). The biologicalorigins and relevance of these dynamic, however, remain unclear(10, 11).Similar to those in behavioral performance, the fluctuations of

collective neuronal activity at many levels of the nervous systemare scale-free and governed by power-law scaling laws. On longtime scales (100−103 s), scale-free fluctuations and long-rangetemporal correlations (LRTCs) are salient in the amplitude enve-lopes of spontaneous neuronal oscillations in data recorded withmagneto- and electroencephalography (M/EEG) (12). It appearsthat these oscillation amplitude fluctuations reflect the un-derlying dynamic architecture of spontaneous brain activity dis-covered with functional MRI (fMRI) and defined by correlatedslow fluctuations in blood oxygenation level–dependent (BOLD)

signals among well-delineated functional brain systems (13–15).The oscillation amplitudes are directly correlated with these BOLDfluctuations (16–20) and exhibit interareal correlations that closelymatch those of BOLD signals (17, 21–23). Moreover, BOLDsignals also exhibit scale-free temporal (24–26) and spatiotem-poral correlations (27–29). The scaling laws of LRTCs thus are aunifying fundamental characteristic of spontaneous brain activity(1, 30, 31).On short time scales (10−3−10−1 s), negative deflections in local

field potentials form spatiotemporal cascades of activity, “neu-ronal avalanches” (32–34), the size and lifetime distributions ofwhich are power laws akin to those of a critical branching process(33). Neuronal avalanches characterize spontaneous neuronalnetwork activity in organotypic cultures (32), brain slices in vitro(35), and monkey (34) and human cortex (36) in vivo. In monkeycortex, the avalanches are delimited by cycles of ongoing neuro-nal oscillations (34) showing that in addition to LRTCs (12),neuronal avalanches also coexist with neuronal oscillations.The power-law scaling behavior and fractal properties of neu-

ronal LRTCs and avalanches strongly suggest that the brainoperates near a critical state (12, 30, 32, 33, 37). Computationalmodeling predicts that LRTCs and neuronal avalanches are cou-pled (38) and suggests that they coemerge from neuronal inter-actions in a critical regime (30). So far, however, neither therelationship between LRTCs and avalanche dynamics nor theirsignificance at the behavioral level has been elucidated. We hy-pothesize here that the scaling laws of LRTCs and neuronal ava-lanches are related and are correlated with the interindividualvariability in behavioral scaling laws. We test the hypothesis withsource reconstructed task- and resting-state M/EEG recordingsand show that behavioral scaling laws, LTRCs, and neuronalavalanches are strongly correlated.

ResultsParadigm for Mapping Individual Behavioral and Neuronal ScalingLaws. As an experimental task that yields psychometric data forprobing the scaling laws of cognitive dynamics (7), we used afactorial audiovisual threshold-stimulus detection task (TSDT)in which the subjects were presented independent and continuousauditory and/or visual stimulus streams. The subjects attended

Author contributions: J.M.P. and S.P. designed research; A.Z. and J.H. performed research;J.M.P. and O.K. contributed new reagents/analytic tools; J.M.P., A.Z., J.H., and S.P. ana-lyzed data; and J.M.P., A.Z., K.L.-H., and S.P. wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.1J.M.P. and A.Z. contributed equally to this work.2To whom correspondence should be addressed. E-mail: [email protected]. and S.P. contributed equally to this work.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1216855110/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1216855110 PNAS | February 26, 2013 | vol. 110 | no. 9 | 3585–3590

NEU

ROSC

IENCE

Page 2: PNAS-2013-Palva-3585-90 (1)

either the visual or the auditory stream, or both streamsconcurrently, and reported each perceived stimulus in the atten-ded modality. In TSDTs, the stimulus intensity is adjusted beforethe experiment so that approximately half the stimuli are per-ceived and then kept constant (Fig. 1A and Fig. S1A) (7, 39).Auditory, visual, and audiovisual tasks revealed complex tem-poral structures in individual behavioral time series with intermit-tent sequences of detected (Hits) and undetected stimuli (Misses)(Fig. 1B). We used detrended fluctuation analysis (DFA) to eval-uate the underlying scaling laws and found that the Hit–Miss serieswere characterized by robust LRTCs (Fig. 1C and Fig. S2A; r2 =0.978 ± 0.003, mean ± SD for 56 datasets). At the group level, thebehavioral scaling exponents were significantly modulated by boththe stimulus modality (auditory/visual) and the task (uni-/bimodal)(Fig. S2A). Nevertheless, although the task effects accounted for11% of total variance in behavioral scaling exponents, as much as44% of total variance was attributable to interindividual variability.To test whether the scaling laws of neuronal activity underlie

the interindividual variability in the behavioral scaling laws, weused M/EEG for noninvasive measurements of ongoing brainactivity and source reconstruction methods for transforming fil-tered M/EEG data into time series of 400 patches fully covering

the cortical surface (Materials and Methods). We extracted am-plitude fluctuations of ongoing cortical oscillations with narrow-(Fig. 1D) and broad-band filters. The scaling exponents of LRTCsin neuronal amplitude fluctuations then were estimated withDFA (Fig. 1E). These source-level data revealed fractal ampli-tude fluctuations (Fig. 1D) and hence corroborate earlier M/EEGsensor-level (12) and electrocorticographic (40) investigations inrevealing highly robust LRTCs in all studied frequency bands(r2 = 0.993 ± 0.0006; Fig. S2B). The mean LRTC scaling expo-nents of 10-Hz amplitudes were significantly different among thethree task conditions (Fig. S2B).To obtain a complementary view of the scaling laws of human

brain dynamics, we searched for neuronal avalanches—uninter-rupted cascades of large-magnitude events across the corticalsurface—by using broad-band filtered data and peak detection(34). These events were salient in source-reconstructed M/EEGdata (Fig. 1F and Fig. S3A) and were characterized by power-lawsize and lifetime distributions with exponents close to those ofa critical branching process (−1.5 and −2, respectively (33) (Fig. 1G and H). To confirm the power-law nature of these distributions,we reproduced all analyses with surrogate data. Although the sizeand lifetime distributions of surrogate data were fit by exponentials

Fig. 1. Paradigm for mapping individual behavioral and neuronal scaling laws with TSDTs and source-reconstructed M/EEG recordings. (A) Examples of noise-embedded visual and auditory stimuli whose SNRs are tuned before the experiment to yield an ∼50% hit rate and then kept constant (Fig. S1). (B) Behavioralperformance time series of detected (upward ticks) and undetected (downward ticks) display rich dynamics in a bimodal audiovisual TSDT (visual, red; auditory, blue;time series are for thefirst 10minof a 30-min sessionof a representative subject). (C) Visual and auditory detection time series exhibit LRTCs thatmaybe characterizedfor each subject byDFAexponents, βV and βA. (D) Amplitudefluctuations ofneuronal oscillations in local cortical patches (here, 10Hz in the inferior parietal gyrus) arefractally self-similar and (E) show robust LRTC. (F) Avalanche dynamics are salient in source-reconstructed broad-band data. The time series of cortical patches in theexample avalanche (see also Fig. S3A) are color codedby the peak latency. These colors correspond to those displayedonpial andflattened cortical surfaces and showthe progression of this activity cascade from posterior parietal to temporal and postcentral loci. (Bottom) The avalanche time series (black lines) show the number ofcortical patches in which a peak was found, with zeros indicating interavalanche periods. (G) The sizes and lifetimes of cortical avalanches are approximately power-law distributed with exponents, α, close to those of a critical branching process (−1.5 and −2, respectively). (H) In line with this notion, the kappa index, κ, for the sizedistribution is close to 1. All data in this figure are from the same 30-min session of a subject representative in having β closest to population mean.

3586 | www.pnas.org/cgi/doi/10.1073/pnas.1216855110 Palva et al.

Page 3: PNAS-2013-Palva-3585-90 (1)

better than by power laws (P < 10−4, T test), the original data weremuch better represented by a power-law than by an exponential(P < 10−10, T test; for further corroboration, see Fig. S3 B–D). Likethose of LRTCs, the scaling exponents of neuronal avalancheswere modulated by the tasks (Fig. S3 B andC). To confirm that theLRTCs and neuronal avalanches in these data were not attribut-able to artificial sources, we reproduced these analyses on inverse-modeled empty-room magnetoencephalography (MEG) data aswell as on forward- and inverse-modeled simulated uncorrelatedbrain activity. These control experiments rule out the possibilitythat environmental or instrument noise, the mixing of neuronalsignals at the scalp level, or the source reconstruction methodsmight contribute to the power-law scaling behavior of LRTCs andavalanches found here to characterize brain activity (Fig. S4).

Neuronal and Behavioral Scaling Laws Are Correlated. To test ourhypothesis of a relation between power-law scaling of brain ac-tivity and behavior, we first averaged the DFA exponents acrossbrain regions and task conditions. Individual variation in the DFAexponents of the behavioral time series exhibited a remarkabledependence on the LRTCs of 10-Hz oscillations, whether estimatedfrom task data or from separate resting-state data (Fig. 2A). A sim-ilar relationship was found throughout frequency bands from 5 to

30 Hz, for broad-band amplitudes, and for the DFA exponents ofthe avalanche time series (Fig. 2B). Interestingly, the short–timescale avalanche dynamics, as quantified by the power-law expo-nents of size or lifetime distributions, also were correlated with thebehavioral scaling exponents in both task- and resting-state data(Fig. 2C and Fig. S5 A and B). Hence, a large fraction of interin-dividual variability in behavioral scaling laws is explained linearlyby corresponding variability in the neuronal scaling laws.To test whether arousal fluctuations driven by the autonomic

nervous system played a role in this correlation, we characterizedthe scaling behavior of heart-rate fluctuations and evaluated thecorrelation of these scaling exponents with those of behavioraland neuronal (10-Hz) LRTCs during task performance (Fig. S6A–D). The scaling exponents of heart-rate variability indeed werecorrelated with both neuronal and behavioral exponents, but apartial correlation analysis revealed that the correlation withbehavior was indirect and that the neuronal LRTCs were a me-diating variable (Fig. S6E). The same result was found whencomparing heart-rate and neuronal LRTCs measured during restwith task-state behavioral LRTCs (Fig. S6 F–H).We then tested our second hypothesis about whether the

LRTC and avalanche dynamics were related. The LRTCs of 10-

Fig. 2. Scale-free neuronal dynamics are correlated with interindividual variability in behavioral scaling laws. (A) Mean local LRTCs in the 10-Hz band (β) bothduring the TSDT task performance and in a separate resting-state session are correlated with the mean behavioral scaling exponents (βbehav.). (B) This cor-relation was significant in frequency bands from 5 to 30 Hz, in broad-band data, and for the avalanche DFA (*P < 0.05, **P < 0.01; ***P < 0.005). (C) Scalingexponents of the size (purple) and lifetime (black) distributions of neuronal avalanches in task- and resting-state data are correlated with the behavioralscaling exponents. (D) The LRTC scaling exponents of neuronal amplitude fluctuations in the 10-Hz and (E) all other studied frequency bands are stronglycorrelated with the scaling exponents of neuronal avalanches. (F) Partial correlation analysis shows that neuronal LRTCs also are correlated with behavioralLRTC when the contribution of neuronal avalanches is factored out and that the correlation between avalanches (α) and behavioral LRTC (β*) is mediatedthrough the correlation between neuronal LRTCs (β) and avalanches (see also D and E).

Palva et al. PNAS | February 26, 2013 | vol. 110 | no. 9 | 3587

NEU

ROSC

IENCE

Page 4: PNAS-2013-Palva-3585-90 (1)

Hz oscillations were strongly correlated with the scaling expo-nents of neuronal avalanches (Fig. 2D and Fig. S5B). Surpris-ingly, the correlation between LRTCs and avalanches waseven more pronounced in the other frequency bands (Fig. 2E).These data thus indicate that short (10−3−10−1 s) and long(100−103 s) time-scale neuronal dynamics are related and cor-related with behavioral scaling laws. The correlation betweenneuronal LRTCs and avalanches raises the question of whetherthey both are directly correlated with the behavioral scaling laws.We addressed this question with a partial correlation analysis andfound that, in fact, the correlation of neuronal avalanches withbehavior was fully explained by their correlation with neuronalLRTCs, although factoring out the contribution of neuronalavalanches did not alter the correlation between neuronal andbehavioral LRTCs (Fig. 2F).

Specific Cortical Regions Underlie the Correlation Between Neuronaland Behavioral Scaling Laws. The cortical structures underlyingbrain–behavior correlations were identified by correlating theDFA exponents of each cortical patch with the exponents of thebehavioral scaling laws. For clarity, we collapsed the data into θ/α(5, 7.5, and 10 Hz) and β/γ (15, 20, and 30 Hz) frequency bands.In both θ/α and β/γ bands, visual behavioral LRTCs were cor-related with neuronal LRTCs during task performance in theposterior parietal cortex and, in the β/γ band, also in the cuneusand inferotemporal visual regions. These cortical regions aretask relevant in supporting visual attentional and representa-tional functions, respectively. In addition, however, neuronal andbehavioral LRTCs in the visual task were correlated in assum-ingly task-irrelevant sensorimotor regions and in those belongingto the default mode system, such as the posterior cingulate,precuneus, medial prefrontal cortex, and inferior parietal cortex(Fig. 3A). These observations are well in line with the notion thatthe detection-probability fluctuations in TSDT experiments maybe determined largely by antagonistic (41) fluctuations ofmodality-specific attentional and default-mode systems (1, 19).When comparing the visual behavioral LRTCs with neuronalLRTCs in the resting state, we found a similar, albeit morewidespread, anatomical pattern of correlations (Fig. 3B). Thissuggests that individual endogenous brain dynamics largely arepreserved during the tasks. The resting-state data also show thatthe brain–behavior correlation of LRTCs in the sensorimotorsystem cannot be explained simply by motor response–relatedamplitude transients during task performance. Additional anal-yses of sensorimotor amplitude time series and motor responsesshowed that even during task performance, the motor responsesdo not bias the correlation between sensorimotor and behavioralLRTC exponents or introduce artificial correlations (Fig. S7).Similar to the visual modality, the comparison of neuronal andbehavioral scaling laws in the auditory task revealed significantcorrelations in both task-relevant and -irrelevant regions, al-though the neuronal correlates of the behavioral scaling laws invisual and auditory tasks clearly involved distinct cortical regions(compare Fig. 3 A and B with C and D). Task-relevant structuresincluded predominantly the inferior frontal and superior tem-poral gyri that support auditory attentional and sensory process-ing, respectively, as well as the anterior cingulate and insula (Fig.3C). Neuronal LRTCs in the conceivably task-irrelevant visualcortical areas in the occipital and inferotemporal cortices, how-ever, also were strongly correlated with behavioral LRTCs in theauditory task. Importantly, these regions together closely matchthose found to be correlated with auditory TSDT performance infMRI (42). A similar anatomical pattern also was observed in theresting state (Fig. 3D), corroborating the notion that endogenousneuronal dynamics are preserved during task performance.

DiscussionIn the present study, we found that the strength of autocorrela-tions in neuronal oscillations during task performance, as indexedby individual LRTC scaling exponents, was correlated with theLRTCs in behavioral time series. Importantly, these behavioralscaling exponents were correlated with neuronal exponents alsomeasured during the performance of another task or during rest,indicating that the relationship is not specific to the concurrenttask (Fig. S8) and suggests that this neuronal dynamics arises en-dogenously. The neuronal sources showing correlations betweenbrain and behavior were localized to task-positive systems, in-cluding the visual and auditory cortices, as well as to regionscontrolling attention. Interestingly, neuronal LRTCs in a subset ofthe default-mode network (DMN) also were correlated with be-havioral LRTCs. These observations are in line with fMRI studiessuggesting that out-of-phase slow fluctuations (41) in such task-relevant and task-irrelevant structures, e.g., in somatosensory (43)and auditory (42) TSDTs, predict trial-by-trial variation in be-havioral performance. Furthermore, a recent study shows thatDMN activity might underlie focused attentional states during

Fig. 3. Neuronal correlates of behavioral scaling laws. Pearson correlationcoefficients were computed between βbehav. and β in the six narrow-fre-quency bands for each cortical patch (Fig. 2B), and significant (P < 0.05, FDRcorrected) correlations were displayed on cortical surfaces collapsed into θ/α(5, 7.5, and 10 Hz) and β/γ (15, 20, and 30 Hz) frequency ranges. For eachcortical patch of the Destrieux parcellation, the color intensity indicates thefraction of significant correlations across the three bands (pale, 1/3; medium,2/3; full, 3/3). (A) Correlation of visual behavioral scaling exponents, βV, withthe β of neuronal LRTCs during visual task performance. (B) Correlation of βVwith β in separate resting-state data. (C) Correlation of auditory behavioralscaling exponents, βA, with the β of neuronal LRTCs during auditory taskperformance. (D) Correlation of βA with β in separate resting-state data. a,Anterior; C, central; CI, cingulate; CN, cuneus; F, frontal; G, gyrus; i, inferior;LIN, lingual; m, middle; O, occipital; P, parietal; p, posterior; pr, pre; s,superior; T, temporal. Colors: red, occipital; green, parietal; blue, frontal;yellow, temporal; purple, cingulate. iPG shows the angular part.

3588 | www.pnas.org/cgi/doi/10.1073/pnas.1216855110 Palva et al.

Page 5: PNAS-2013-Palva-3585-90 (1)

continuous performance tasks (44). These BOLD signals and theoscillation amplitude fluctuations, such as those measured in thisstudy, covary in intrinsic connectivity networks (16, 18, 19, 21, 22).Importantly, because similar albeit more widespread correla-

tions between neuronal and behavioral LRTCs also were observedduring rest (Fig. 3), the individual phenotypic variability in neu-ronal scaling laws may be a determinant of the dynamic nature ofvariability in task performance. Supporting this idea, LRTCs inoscillations are heritable (45), test–retest reliable (46), and cor-related with brain pathologies (40, 47–50). Considering that theexponents of autonomic nervous system fluctuations were corre-lated only indirectly with behavior (Fig. S6), the intrinsic organi-zation of neuronal networks as reflected in scaling laws of neuronalfluctuations is a likely physiological substrate of the scaling lawsgoverning the variability in psychophysical performance (1, 12).The notion of phenotypic and individual neuronal scaling laws

is in line with computational modeling showing that emergentfluctuations in resting-state neuronal activity with realistic struc-tural connectivity exhibit spatiotemporal power-law scaling (51).The individual variability in the scaling laws of neuronal ava-lanches, LRTCs, and behavior thus might arise mechanisticallyfrom the dynamics of neuronal activity in the individual brainarchitecture of structural connectivity (30, 38, 52, 53) togetherwith individual variability in the expression of cellular levelmechanisms (54–56) that regulate, for instance, the excitation–inhibition balance (38). Neuronal avalanches appear phenome-nologically distinct from LRTCs in that they involve a spatialdimension and time scales much shorter than those typically as-sociated with LRTCs (10−3−10−1 vs. 100−103 s, respectively).Nevertheless, as predicted theoretically (38), we found that thepower-law exponents of avalanche size and lifetime distributionswere strongly correlated with those of LRTCs in a range of fre-quency bands. Although the avalanche exponents also were cor-related with behavioral LRTCs, this relationship was indirect andmediated by neuronal LRTCs. We suggest that neuronal ava-lanches and oscillations exhibiting LRTCs coemerge in near–critical-state brain dynamics and reflect propagating neuronalactivity at widespread spatial and temporal scales.Criticality has gained widespread interest in neuroscience as a

framework for understanding the character and functional im-plications of variability in brain activity (12, 30, 32). The meta-stability of critical systems maximizes their dynamic range (57),storage capacity (58), and computational power. Fractal self-similarity, power-law scaling behavior, and “1/f noise” at thephenomenological level are typical for systems exhibiting “ava-lanche dynamics” and operating in a critical (37) or self-organizedcritical state (59). Nevertheless, the fact that numerous complexsystems exhibit similar dynamics raises the question of whetherfractal neuronal dynamics are an epiphenomenon without func-tional relevance. Our results contribute to this context in tworespects. First, the discovery that neuronal scaling only in well-delineated and task-specific brain systems is correlated with hu-man behavioral scaling suggests that neuronal criticality is notepiphenomenal. Second, the observations of LRTCs and neuro-nal avalanches with power-law size and lifetime distributions addto the growing body of data suggesting that the human brainoperates near a critical regime (12, 30, 32, 33), which may fun-damentally determine the dynamics of human perceptual, cog-nitive, and behavioral processes.

Materials and MethodsSubjects and Stimuli. Fourteen healthy subjects (seven females) participated inthe experiment, which comprised three continuous 30-min M/EEG task-staterecordings followed by a 10-min resting-state recording. In each of three 30-min task-state experiments, continuous, concurrent, and uncorrelated seriesof constant-intensity noise-embedded threshold-level auditory and visualstimuli were presented. The interstimulus interval ranged from 1.5 to 6 s, witha mean of 3.75 s, yielding a total of 480 auditory and visual stimuli per

experiment. Before each experiment, the subject was instructed to indicatethe perception of each stimulus in the attended sensory modality by pressinga button with the index finger (left/right hand pseudo-randomly counter-balanced across subjects). In the three randomly ordered task-state record-ings, the tasks were to attend the auditory (A), the visual (V), or both (B)stimuli. Before the onset of the experiment, the intensities of auditory andvisual stimuli with respect to the continuous background, as indexed bya signal-to-noise ratio (SNR), were calibrated separately to yield an initialdetection probability (hit rate, HR) of ∼50%. After calibration, these SNRvalues were kept constant throughout the experiment.

M/EEG Recordings and Source Reconstruction. We recorded 366-channel M/EEG data with 204 planar gradiometers, 102 magnetometers, and 60 EEGelectrodes (Elekta Neuromag Ltd.) at a 600-Hz sampling rate. The MaxFiltersoftware was used to suppress extracranial noise and to colocalize the signalspace data from different recording sessions and subjects. For cortical surfacereconstructions, we recorded T1-weighted (magnetization-prepared rapidgradient echo) anatomical magnetic resonance images at a ≤ 1 × 1 × 1-mmresolution with a 1.5-T MRI scanner (Siemens). This study was approved bythe Ethical Committee of Helsinki University Central Hospital and was per-formed according to the Declaration of Helsinki. Written informed consentwas obtained from each subject before the experiment. FreeSurfer softwarewas used for automatic volumetric segmentation of the MRI data, surfacereconstruction, flattening, cortical parcellation, and labeling with the Des-trieux atlas; the MNE software was used to create three-layer boundary el-ement conductivity and cortically constrained source models, to colocalizeM/EEG–MRI data, and to prepare the forward and inverse operators (SIMaterials and Methods). M/EEG time series were filtered into seven fre-quency bands, f. A bank of Morlet wavelets (f = 5, 7.5, 10, 15, 20, and 30 Hz)yielded narrow-band data and a finite impulse-response filter was used forbroad-band filtering from 0.1 to 45 Hz (pass-band from 1 to 30 Hz). Afterfiltering, M/EEG sensor data were inverse transformed and then collapsedinto time series of 400 cortical patches with individual fidelity-optimizedparcellations (FOPs; SI Materials and Methods). Statistics and visualizationwere performed in the 148-patch Destrieux parcellation. The scaling expo-nent of each Destrieux parcellation patch was calculated by averaging theexponents of corresponding subpatches in the FOPs.

Estimation of Neuronal and Behavioral Scaling Laws. The monofractal scalingexponents of neuronal and behavioral LRTCs were estimated with DFA (12) ofband-amplitude envelopes and Hit–Miss time series (SI Materials andMethods). A single neuronal avalanche was defined as a set of continuoussamples in which one or more peaks in broad-band filtered data were ob-served. The number of peaks in the avalanche was taken as its size and thetime spanned by the samples as its lifetime. The scaling exponents of the sizeand lifetime distributions then were estimated as in ref. 34 and the κ indicesas in ref. 58 (for details, see SI Materials and Methods).

Statistical Analysis. The correlations among neuronal and behavioral LRTCscaling exponents and those of avalanche dynamics were assessed withPearson correlation coefficients. In Fig. 2, the behavioral scaling exponentswere averaged across conditions (βbehav.; Fig. 2A), the neuronal scaling ex-ponents were averaged across patches and task conditions for each frequencyband (Fig. 2B), and the significance of the correlation coefficient was assessedwith t-statistics: t = r

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðn− 2Þ=ð1− r2Þ

p, where r is the correlation coefficient

and n is the sample size (number of subjects). To measure the degree of as-sociation between βbehav., β, and αlifetime/size, we used a partial correlationanalysis by computing Pearson correlation coefficients between each pair ofsets, regressing out the effect of the third set (Fig. 2F and Fig. S6).

In Fig. 3, the correlation between behavioral scaling exponents and thepatchwise neuronal scaling exponents was estimated with the Pearson corre-lation coefficient as above. To identify the most central cortical regions un-derlying behavioral scaling laws separately for the auditory and visual modali-ties, the fraction of significant coefficients (P < 0.05, corrected for multiplecomparisons) was assessed in the unimodal tasks (Fig. 3). False discovery rate(FDR) control was applied to correct the statistical significance for multiplecomparisons collectively across all cortical regions and the six frequency bands.

To estimate the task and modality effects on the behavioral scalingexponents, two-way ANOVA was applied with task (auditory/visual) andmodality (uni-/bimodal) as the independent variables and the scaling expo-nents were computed for each subject and condition as the dependentvariables. The task effect for the neuronal and avalanche scaling exponentswas estimated using one-way ANOVA with condition (either auditory/visual/audiovisual or task/rest) as the independent variables and scaling exponentsof each subject and condition as the dependent variable.

Palva et al. PNAS | February 26, 2013 | vol. 110 | no. 9 | 3589

NEU

ROSC

IENCE

Page 6: PNAS-2013-Palva-3585-90 (1)

ACKNOWLEDGMENTS. This work was funded by Academy of Finland Grants253130 and 256472 (to J.M.P.) and 1126967 (to S.P.), University of HelsinkiResearch Funds (S.P.), the Sigrid Juselius Foundation (S.P. and J.M.P.), Centre

for International Mobility CIMO (J.M.P. and A.Z.), and Netherlands Orga-nization for Scientific Research (NWO) Physical Sciences Grant 612.001.123(to K.L.-H.).

1. Palva JM, Palva S (2011) Roles of multiscale brain activity fluctuations in shaping thevariability and dynamics of psychophysical performance. Prog Brain Res 193:335–350.

2. VerplanckWS, Collier GH, Cotton JW (1952) Nonindependence of successive responsesin measurements of the visual threshold. J Exp Psychol 44(4):273–282.

3. Gilden DL, Thornton T, Mallon MW (1995) 1/f noise in human cognition. Science267(5205):1837–1839.

4. Gilden DL (2001) Cognitive emissions of 1/f noise. Psychol Rev 108(1):33–56.5. Helps SK, Broyd SJ, James CJ, Karl A, Sonuga-Barke EJS (2010) The attenuation of very

low frequency brain oscillations in transitions from a rest state to active attention.J Psychophysiol 23:191–198.

6. Ihlen EA, Vereijken B (2010) Interaction-dominant dynamics in human cognition:Beyond 1/f(alpha) fluctuation. J Exp Psychol Gen 139(3):436–463.

7. Monto S, Palva S, Voipio J, Palva JM (2008) Very slow EEG fluctuations predict thedynamics of stimulus detection and oscillation amplitudes in humans. J Neurosci28(33):8268–8272.

8. Thornton TL, Gilden DL (2005) Provenance of correlations in psychological data.Psychon Bull Rev 12(3):409–441.

9. Chen Y, Ding M, Kelso JA (2001) Origins of timing errors in human sensorimotorcoordination. J Mot Behav 33(1):3–8.

10. Kello CT, et al. (2010) Scaling laws in cognitive sciences. Trends Cogn Sci 14(5):223–232.

11. Farrell S, Wagenmakers EJ, Ratcliff R (2006) 1/f noise in human cognition: Is it ubiq-uitous, and what does it mean? Psychon Bull Rev 13(4):737–741.

12. Linkenkaer-Hansen K, Nikouline VV, Palva JM, Ilmoniemi RJ (2001) Long-range tem-poral correlations and scaling behavior in human brain oscillations. J Neurosci 21(4):1370–1377.

13. Biswal B, Yetkin FZ, Haughton VM, Hyde JS (1995) Functional connectivity in themotor cortex of resting human brain using echo-planar MRI. Magn Reson Med 34(4):537–541.

14. Damoiseaux JS, et al. (2006) Consistent resting-state networks across healthy subjects.Proc Natl Acad Sci USA 103(37):13848–13853.

15. Raichle ME (2011) The restless brain. Brain Connect 1(1):3–12.16. Goldman RI, Stern JM, Engel J, Jr., Cohen MS (2002) Simultaneous EEG and fMRI of

the alpha rhythm. Neuroreport 13(18):2487–2492.17. Leopold DA, Murayama Y, Logothetis NK (2003) Very slow activity fluctuations in

monkey visual cortex: Implications for functional brain imaging. Cereb Cortex 13(4):422–433.

18. Mantini D, Perrucci MG, Del Gratta C, Romani GL, Corbetta M (2007) Electrophysio-logical signatures of resting state networks in the human brain. Proc Natl Acad SciUSA 104(32):13170–13175.

19. Sadaghiani S, et al. (2010) Intrinsic connectivity networks, alpha oscillations, and tonicalertness: A simultaneous electroencephalography/functional magnetic resonanceimaging study. J Neurosci 30(30):10243–10250.

20. Schölvinck ML, Maier A, Ye FQ, Duyn JH, Leopold DA (2010) Neural basis of globalresting-state fMRI activity. Proc Natl Acad Sci USA 107(22):10238–10243.

21. Brookes MJ, et al. (2011) Investigating the electrophysiological basis of restingstate networks using magnetoencephalography. Proc Natl Acad Sci USA 108(40):16783–16788.

22. de Pasquale F, et al. (2010) Temporal dynamics of spontaneous MEG activity in brainnetworks. Proc Natl Acad Sci USA 107(13):6040–6045.

23. Nikouline VV, Linkenkaer-Hansen K, Huttunen J, Ilmoniemi RJ (2001) Interhemisphericphase synchrony and amplitude correlation of spontaneous beta oscillations in humansubjects: A magnetoencephalographic study. Neuroreport 12(11):2487–2491.

24. Suckling J, Wink AM, Bernard FA, Barnes A, Bullmore E (2008) Endogenous multi-fractal brain dynamics are modulated by age, cholinergic blockade and cognitiveperformance. J Neurosci Methods 174(2):292–300.

25. Wink AM, Bullmore E, Barnes A, Bernard F, Suckling J (2008) Monofractal and mul-tifractal dynamics of low frequency endogenous brain oscillations in functional MRI.Hum Brain Mapp 29(7):791–801.

26. He BJ (2011) Scale-free properties of the functional magnetic resonance imagingsignal during rest and task. J Neurosci 31(39):13786–13795.

27. Eguíluz VM, Chialvo DR, Cecchi GA, Baliki M, Apkarian AV (2005) Scale-free brainfunctional networks. Phys Rev Lett 94(1):018102.

28. Expert P, et al. (2011) Self-similar correlation function in brain resting-state functionalmagnetic resonance imaging. J R Soc Interface 8(57):472–479.

29. Tagliazucchi E, Balenzuela P, Fraiman D, Chialvo DR (2012) Criticality in large-scalebrain FMRI dynamics unveiled by a novel point process analysis. Front Physiol 3:15.

30. Chialvo DR (2010) Emergent complex neural dynamics. Nat Phys 6:744–750.31. He BJ, Zempel JM, Snyder AZ, Raichle ME (2010) The temporal structures and func-

tional significance of scale-free brain activity. Neuron 66(3):353–369.

32. Beggs JM, Plenz D (2003) Neuronal avalanches in neocortical circuits. J Neurosci23(35):11167–11177.

33. Plenz D, Thiagarajan TC (2007) The organizing principles of neuronal avalanches: Cellassemblies in the cortex? Trends Neurosci 30(3):101–110.

34. Petermann T, et al. (2009) Spontaneous cortical activity in awake monkeys composedof neuronal avalanches. Proc Natl Acad Sci USA 106(37):15921–15926.

35. Stewart CV, Plenz D (2006) Inverted-U profile of dopamine-NMDA-mediated spon-taneous avalanche recurrence in superficial layers of rat prefrontal cortex. J Neurosci26(31):8148–8159.

36. Solovey G, Miller KJ, Ojemann JG, Magnasco MO, Cecchi GA (2012) Self-regulateddynamical criticality in human ECoG. Front Integr Neurosci 6:44.

37. Werner G (2010) Fractals in the nervous system: Conceptual implications for theo-retical neuroscience. Front Physio 1:1–28.

38. Poil SS, Hardstone R, Mansvelder HD, Linkenkaer-Hansen K (2012) Critical-state dy-namics of avalanches and oscillations jointly emerge from balanced excitation/in-hibition in neuronal networks. J Neurosci 32(29):9817–9823.

39. Linkenkaer-Hansen K, Nikulin VV, Palva JM, Kaila K, Ilmoniemi RJ (2004) Stimulus-induced change in long-range temporal correlations and scaling behaviour of sen-sorimotor oscillations. Eur J Neurosci 19(1):203–211.

40. Monto S, Vanhatalo S, Holmes MD, Palva JM (2007) Epileptogenic neocortical net-works are revealed by abnormal temporal dynamics in seizure-free subdural EEG.Cereb Cortex 17(6):1386–1393.

41. Fox MD, et al. (2005) The human brain is intrinsically organized into dynamic, anti-correlated functional networks. Proc Natl Acad Sci USA 102(27):9673–9678.

42. Sadaghiani S, Hesselmann G, Kleinschmidt A (2009) Distributed and antagonisticcontributions of ongoing activity fluctuations to auditory stimulus detection. J Neu-rosci 29(42):13410–13417.

43. Boly M, et al. (2007) Baseline brain activity fluctuations predict somatosensory per-ception in humans. Proc Natl Acad Sci USA 104(29):12187–12192.

44. Esterman M, Noonan SK, Rosenberg M, Degutis J (2012) In the zone or zoning out?Tracking behavioral and neural fluctuations during sustained attention. Cereb Cortex,10.1093/cercor/bhs261.

45. Linkenkaer-Hansen K, et al. (2007) Genetic contributions to long-range temporalcorrelations in ongoing oscillations. J Neurosci 27(50):13882–13889.

46. Nikulin VV, Brismar T (2004) Long-range temporal correlations in alpha and betaoscillations: Effect of arousal level and test-retest reliability. Clin Neurophysiol 115(8):1896–1908.

47. Linkenkaer-Hansen K, et al. (2005) Breakdown of long-range temporal correlations intheta oscillations in patients with major depressive disorder. J Neurosci 25(44):10131–10137.

48. Montez T, et al. (2009) Altered temporal correlations in parietal alpha and prefrontaltheta oscillations in early-stage Alzheimer disease. Proc Natl Acad Sci USA 106(5):1614–1619.

49. Nikulin VV, Jönsson EG, Brismar T (2012) Attenuation of long-range temporal cor-relations in the amplitude dynamics of alpha and beta neuronal oscillations in pa-tients with schizophrenia. Neuroimage 61(1):162–169.

50. Lai MC, et al.; MRC AIMS Consortium (2010) A shift to randomness of brain oscillationsin people with autism. Biol Psychiatry 68(12):1092–1099.

51. Deco G, Jirsa VK (2012) Ongoing cortical activity at rest: Criticality, multistability, andghost attractors. J Neurosci 32(10):3366–3375.

52. Honey CJ, et al. (2009) Predicting human resting-state functional connectivity fromstructural connectivity. Proc Natl Acad Sci USA 106(6):2035–2040.

53. Deco G, Jirsa VK, McIntosh AR (2011) Emerging concepts for the dynamical organi-zation of resting-state activity in the brain. Nat Rev Neurosci 12(1):43–56.

54. Eagleson KL, Campbell DB, Thompson BL, Bergman MY, Levitt P (2011) The autism riskgenes MET and PLAUR differentially impact cortical development. Autism Res 4(1):68–83.

55. Bruening S, et al. (2006) The anxiety-like phenotype of 5-HT receptor null mice isassociated with genetic background-specific perturbations in the prefrontal cortexGABA-glutamate system. J Neurochem 99(3):892–899.

56. Won H, et al. (2011) GIT1 is associated with ADHD in humans and ADHD-like be-haviors in mice. Nat Med 17(5):566–572.

57. Kinouchi O, Copelli M (2006) Optimal dynamical range of excitable networks atcriticality. Nat Phys 2:348–351.

58. Shew WL, Yang H, Petermann T, Roy R, Plenz D (2009) Neuronal avalanches implymaximum dynamic range in cortical networks at criticality. J Neurosci 29(49):15595–15600.

59. Bak P, Tang C, Wiesenfeld K (1987) Self-organized criticality: An explanation of the 1/fnoise. Phys Rev Lett 59(4):381–384.

3590 | www.pnas.org/cgi/doi/10.1073/pnas.1216855110 Palva et al.

Page 7: PNAS-2013-Palva-3585-90 (1)

Supporting InformationPalva et al. 10.1073/pnas.1216855110SI Materials and MethodsMagneto-/Electroencephalography Recordings and Source Recon-struction. As earlier (1, 2), for each subject, the FreeSurfer soft-ware (http://surfer.nmr.mgh.harvard.edu/) was used for automaticvolumetric segmentation of the MRI data, reconstruction of white-gray matter and pial surfaces, flattening of complete cut surfaces,and cortical parcellation and labeling with the Destrieux atlas (3–5). MNE software (www.martinos.org/mne/) was used to createthree-layer boundary element conductivity models and corticallyconstrained source models with fixed-orientation dipoles at 7-mmspacing oriented orthogonal to the cortical surface. MNE soft-ware also was used for magneto-/electroencephalography (M/EEG)–MRI colocalization and for the preparation of the for-ward and inverse operators (6–8).M/EEG signals, Y(t), are related to the source activity by re-

lation Y(t) = ΓX(t) + N(t), where Γ is the lead field matrix (for-ward operator) that relates the source dipole strengths to thesensor level data, X(t) is the source dipole data, and N(t) is noise.To obtain X(t) from the measured Y(t), we used a minimum-normestimator, such that X(t) =MY(t) = RΓT(ΓRΓT + λ2χ), whereM isthe inverse operator, R the source covariance matrix, λ2 a regu-larization parameter, and χ the noise covariance matrix (6–8).After exclusion of environmental and physiological artifacts,

the recorded M/EEG data were filtered (Materials and Methods)and transformed to time series of ∼7,000 source dipoles withX(t) = MY(t) and then collapsed to time series in a corticalparcellation, i.e., a set of cortical patches, each of which is a setof source vertices (1, 2). Here, we used a parcellation strategyaiming to minimize spurious interactions between patch timeseries and, to this end, searched iteratively for each individuala parcellation that maximized the correlation between simulatedand inverse modeled time series for each given patch (fidelity, ξ)and minimized the correlation between the inverse modeledtime series of each patch with the simulated time series of otherpatches (infidelity, ψ). This fidelity-optimization approach wasapplied with a selection criterion of maximal 0.7*ξ +0.3*(1 − ψ)to an anatomical parcellation of 400 patches that were derivedfrom the Destrieux atlas by iteratively splitting the patches thathad the largest size in the subject population along the axis (an-terior-posterior, lateral-medial, ventral-dorsal) with largest meanvariance (1, 2). The iterative fidelity optimization was performed,in short, so that patch and source vertex time series were simu-lated with each source vertex sharing the time series of the patchowning it. The source vertex time series then were forward andinverse modeled, and for each patch, the vertices yielding themaximal value of selection criterion were selected. ξ and ψ werecalculated at each step of the iterative selection process betweenthe simulated patch time series and a weighted average of thetime series of the vertex constellation tested. The inverse modeledvertex data were collapsed into patch data by obtaining the timeseries for each patch as an average of the time series of vertices inthat patch, weighted by jRe(ξ)j.Neuronal long-range temporal correlation (LRTC) and ava-

lanche analyses in individual subjects were performed on timeseries in the 400-patch parcellation. Group statistics and visual-ization of brain–behavior correlations were performed in theDestrieux parcellation with the scaling exponent of each Des-trieux parcellation patch obtained by averaging the exponents ofthe corresponding subpatches.

Estimation of Scaling Laws for LRTCs. Detrended fluctuation anal-ysis (DFA) was applied to estimate the scaling laws for LRTCs.

DFA is a two-stage procedure: In the first stage, time series X(k)is normalized to zero mean and integrated, yðkÞ=Pk

i=1½XðiÞ−hXi�, then segmented into time windows of various sizes Δt. Inthe second stage, each segment of integrated data is locally fittedto a linear function yΔtðkÞ and the mean-squared residual F(Δt)is computed:

FðΔtÞ =ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1N

XN

k= 1

½ yðkÞ− yΔtðkÞ�2vuut ;

where N is the total number of data points.The scaling law exponent β is defined as the slope of linear

regression of the function F(Δt) in log–log coordinates, esti-mated using a least-squares algorithm.

Estimation of Behavioral and Neuronal DFA. To assess the LRTCs,DFA was applied to the behavioral sequence of Hits and Misses,neuronal oscillation amplitude envelopes, and the time series ofneuronal avalanches. Unlike the magnetoencephalography (MEG)data, the behavioral time series of Hits/Misses were sparsely andirregularly sampled because of the variability in interstimulusintervals that were distributed uniformly between 1.5 s and 6 s(mean 3.75 s). The behavioral time series hence were linearlyinterpolated into a time series of 0s and 1s with a 10-Hz samplingrate. To discard any possible effects of this interpolation on thescaling law exponents, only DFA time windows between 4 and400 s were used in the regression. To consolidate that the be-havioral LRTCs could not be explained by a random process, weshuffled the Hit–Miss time series and reproduced the resamplingand DFA estimation for 1,000 iterations. Of 56 behavioral timeseries, 44 yielded LRTC exponents greater than the mean onsurrogates (P < 10−5, binomial test), showing that the behavioralLRTCs were a robust phenomenon. The scaling exponents werenot correlated with hit rate (P > 0.5), even though for randomHit–Miss sequences, very low hit rates impose a positive bias. Toassess the scaling exponents of the neuronal LRTCs, DFA wasapplied to the amplitude envelope of the filtered signal for eachcortical patch and frequency band. For consistency with thebehavioral DFA analysis and to exclude the possibility that short-range correlations or filtering biased the neuronal LRTC ex-ponents, we used the same 4−400-s range for regression of theDFA slope. All neuronal LRTC exponents were greater than the99.99 percentile of surrogates estimated as in ref. 9 and, hence,highly significant.

Estimation of Scaling Laws for Neuronal Avalanches. Broad-bandfiltered neuronal activity time series of each cortical patch werenormalized by subtracting the mean and dividing by SD. Thenthe time series were down-sampled to 300 Hz to achieve a time binsize comparable with prior experiments with local field potential(LFP) recordings. Each patch time series was transformed toa binary point process by detecting positive and negative peaksabove a threshold of three SDs, and setting the samples corre-sponding to peak latencies to ones (“1”). These binary sequencesthen were summed across cortical patches to create the avalanchetime series (Fig. 1F). Zeros in this time series constitute theinteravalanche “waiting” periods and sequences of consecutivevalues above 0 an avalanche. The avalanches were described interms of their lifetimes, given by the durations in milliseconds,and sizes, given by the total number of peaks. To estimate theavalanche lifetime and size distributions (10), the histogram ap-

Palva et al. www.pnas.org/cgi/content/short/1216855110 1 of 7

Page 8: PNAS-2013-Palva-3585-90 (1)

proach was used. Sizes of the histogram bins were selected frompower-spaced series (1.8m, m = 1, . . ., 9). A complementarycharacteristic, the κ-index, for neuronal avalanches was in-troduced by Shew et al. (11). This nonparametric measurequantifies the difference between an experimental cumulativedensity function (CDF) of the avalanche size, F(bk), and thetheoretical reference CDF, FNA(bk), which is a power-law functionwith the theoretically expected exponent −3/2:

κ = 1+1m

Xm

k= 1

�FNAðbkÞ−FðbkÞ

�;

where bk are the power-spaced avalanche sizes with, in our im-plementation, the base of 1.8 (1.8m) and m = 1, . . ., 9. κ-indexvalues at around 1 are characteristic of systems in a critical state,whereas values below and above 1 suggest sub- and supercriticalstates, respectively.

Simulations and Validation. Forward-inverse modeling, which usesspatiotemporal information of the M/EEG data, potentially mayinduce the spatial and temporal correlation in neuronal timeseries. To verify that scaling law exponents are not affected bymodeling, we assessed the scaling exponents for simulated data.Random signals with uniform probability density function

(PDF) were generated for each cortical patch, then Morletwavelets with central frequency of 10 Hz and broad-band filterswere applied to patch time courses. Cortical patches weremapped to source space (∼7,000 sources), then forward andinverse operators were applied sequentially. Finally, the signalswere mapped back to cortical patch space, and scaling exponentswere computed. We used the same approach to assess the impactof noise generated by the empty scanner.

SI Note. It is important to note that the present data relate onlyindirectly to 1/f scaling in frequency spectra. The scaling in thelow-frequency end of the classical power spectra is biased by theconductive properties of brain tissue (12), and the signals in thisfrequency regime may have several incompletely understoodneurophysiological or hemodynamic sources (13). These con-founders influence the scaling laws of the power spectra ob-tained directly from real-valued M/EEG signals, but are notrelevant for the amplitude envelopes of narrow-band neuronaloscillations studied here. Our LRTC analyses focus on slow fluc-tuations in the amplitude envelopes of fast neuronal oscillationsthat reflect the dynamics of more specific neuronal processes andyield (amplitude) scaling laws that are not influenced by the pas-sage of electrical and magnetic fields in brain tissue (12).

1. Palva JM, Monto S, Kulashekhar S, Palva S (2010) Neuronal synchrony reveals workingmemory networks and predicts individual memory capacity. Proc Natl Acad Sci USA107(16):7580–7585.

2. Palva S, Kulashekhar S, Hämäläinen M, Palva JM (2011) Localization of cortical phaseand amplitude dynamics during visual working memory encoding and retention.J Neurosci 31(13):5013–5025.

3. Dale AM, Fischl B, Sereno MI (1999) Cortical surface-based analysis. I. Segmentationand surface reconstruction. Neuroimage 9(2):179–194.

4. Fischl B, Sereno MI, Dale AM (1999) Cortical surface-based analysis. II: Inflation,flattening, and a surface-based coordinate system. Neuroimage 9(2):195–207.

5. Fischl B, et al. (2002) Whole brain segmentation: automated labeling of neuroanatomical structures in the human brain. Neuron 33(3):341–355.

6. Hämäläinen MS, Sarvas J (1989) Realistic conductivity geometry model of the humanhead for interpretation of neuromagnetic data. IEEE Trans Biomed Eng 36(2):165–171.

7. Hämäläinen MS, Ilmoniemi RJ (1994) Interpreting magnetic fields of the brain:Minimum norm estimates. Med Biol Eng Comput 32(1):35–42.

8. Lin FH, Belliveau JW, Dale AM, Hämäläinen MS (2006) Distributed current estimatesusing cortical orientation constraints. Hum Brain Mapp 27(1):1–13.

9. Linkenkaer-Hansen K, Nikouline VV, Palva JM, Ilmoniemi RJ (2001) Long-rangetemporal correlations and scaling behavior in human brain oscillations. J Neurosci21(4):1370–1377.

10. Beggs JM, Plenz D (2003) Neuronal avalanches in neocortical circuits. J Neurosci23(35):11167–11177.

11. Shew WL, Yang H, Petermann T, Roy R, Plenz D (2009) Neuronal avalanches implymaximum dynamic range in cortical networks at criticality. J Neurosci 29(49):15595–15600.

12. Dehghani N, Bédard C, Cash SS, Halgren E, Destexhe A (2010) Comparative powerspectral analysis of simultaneous elecroencephalographic and magnetoencepha-lographic recordings in humans suggests non-resistive extracellular media. J ComputNeurosci 29(3):405–421.

13. Palva JM, Palva S (2012) Infra-slow fluctuations in electrophysiological recordings,blood-oxygenation-level-dependent signals, and psychophysical time series. Neuroimage62(4):2201–2211.

Fig. S1. Signal-to-noise ratio (SNR) of the visual and auditory stimuli was adjusted for each subject in a calibration session before the experiment. (A) Visualstimuli (circular pattern) lasted 50 ms and were superimposed on a continuous, very slowly moving Perlin noise background. The auditory stimuli were dual-frequency (115- and 185-Hz) Hanning-windowed sinusoidal signals lasting 50 ms and were superimposed on continuous Poisson noise (50–350 Hz). (B) In-tensities of visual and auditory stimuli were calibrated before the experiment to yield an initial hit rate of ∼50%. Auditory and visual stimuli were detectedwith equal probabilities (P > 0.34) in uni- and bimodal experiments. The mean final hit rates in the unimodal tasks (mean of auditory and visual hit rates: 0.44 ±0.04, mean ± SEM) were significantly (P < 0.05) higher than the final hit rates in the bimodal task (0.36 ± 0.04).

Palva et al. www.pnas.org/cgi/content/short/1216855110 2 of 7

Page 9: PNAS-2013-Palva-3585-90 (1)

Fig. S2. Behavioral data, as well as task- and resting-state neuronal data, reveal robust scaling laws and significant task effects. (A1) DFA of behavioral timeseries (averaged across subjects). Blue, unimodal auditory; red, unimodal visual; green ○, bimodal auditory; green △, bimodal visual. Lines indicate best-fitlinear regression in the time range from 4 to 400 s. (A2) Comparisons of the scaling exponents across subjects and conditions with a two-way ANOVA revealedan effect of both sensory modality (P < 0.03, F-test) and task (P < 0.02, F-test), but not an interaction between modality and task. Error bars indicate the SEM ofscaling exponents across subjects. Note that within-subject means were subtracted before ANOVA to remove the contribution of large interindividual vari-ability therein (Results). A, auditory; B, bimodal; V, visual. (A3) Mean ± SEM r2 values for the goodness of fit of the linear regression on behavioral DFAs. (B1)DFA of task- and resting-state 10-Hz neuronal oscillations (averaged across subjects). Blue, auditory; red, visual; green, audiovisual; orange, resting-state.Colored lines indicate best-fit linear regression in the time range from 4 to 400 s; gray line is the DFA of phase-randomized surrogate data with an exponentof ∼0.5; all individuals had exponents above this value. (B2) Comparisons of the neuronal scaling exponents between conditions with one-way ANOVA re-vealed an effect of task (P < 0.008, F-test; see note on A2), whereas the effect of task (all three conditions) vs. resting state was not significant (P > 0.11, F-test).(B3) Mean ± SEM r2 values for the goodness of fit of the linear regression on neuronal DFAs. Systematic variability of both the LRTC and avalanche scalingexponents among different task conditions (see also Fig. S3) yields a complementary view on the putative significance of scaling laws in neuronal dynamics.Comparable effects were observed earlier with MEG and functional MRI between task and rest conditions (1, 2). In these data, the scaling exponents were notsuppressed from resting-state levels during task performance, which likely is attributable to the threshold-stimulus detection task paradigm, in which thestimulus-evoked disruption of ongoing dynamics is minimal. The task effects show that rather than being a stationary property of individual nervous systems,the scaling exponents are malleable, dynamic, and dependent on brain states in time scales from minutes to tens of minutes. In other words, if the scalingexponents are taken as indicators of the brain operating near a critical state, then this operating point, even in the healthy brain, may flexibly fluctuate in theneighborhood of the critical regime.

1. Linkenkaer-Hansen K, Nikulin VV, Palva JM, Kaila K, Ilmoniemi RJ (2004) Stimulus-induced change in long-range temporal correlations and scaling behaviour of sensorimotoroscillations. Eur J Neurosci 19(1):203–211.

2. He BJ (2011) Scale-free properties of the functional magnetic resonance imaging signal during rest and task. J Neurosci 31(39):13786–13795.

Palva et al. www.pnas.org/cgi/content/short/1216855110 3 of 7

Page 10: PNAS-2013-Palva-3585-90 (1)

Fig. S3. Avalanche dynamics are prominent in source-reconstructed M/EEG data during both task- and resting-state conditions. (A) Example avalancheshowing a cascade propagating in the left hemispheric occipital and temporal regions (see Fig. 1F for color coding). (B1) Lifetime distribution of avalanchedynamics is significantly different from that of surrogates obtained by random rotation of cortical patch time series, which preserves the autocorrelationstructure but mixes temporal relationships between patches [gray lines, P < 0.005, Kolmogorov–Smirnov (KS) test; for colors, see B2; distributions averagedacross subjects]. (B2) Task effect on the lifetime scaling exponents (mean ± SEM across subjects) estimated with a one-way ANOVA was significant (P < 0.002,F-test), but the effect of task vs. rest, R, was not (P > 0.16, F-test). A, auditory; AV, audiovisual, V, visual. (B3) Avalanche lifetime distributions of original data ofindividual datasets were fit (r2 on the y axis) by power laws better than by exponentials (P < 10−10, t-test), whereas the surrogate data were represented betterby exponentials than by power laws (P < 10−4, t-test). This observation was fully reproduced when using the maximum likelihood approach (1) (maximumlikelihood ratio test for the original data being better described by a power law than by an exponential; P < 0.0001). (C1) Size distributions of avalanchesshowed similar characteristics as lifetime distributions (B) and were different from surrogate data (gray lines, P < 0.002, KS test). (C2) One-way ANOVA againrevealed a task effect (P < 0.03, F-test) but no difference between task and rest conditions (P > 0.22, F-test). (C3) The size distributions of original data werebetter fit by power laws than by exponentials (P < 10−11, t-test), and vice versa for surrogate data (P < 10−14, t-test); this finding also was corroborated with themaximum likelihood analysis (as in B; P < 0.0001) (D1 and D2) We used the κ-index as an additional measure of the size distribution scaling. The cumulativedistribution function (CDF) for observing an avalanche with a given size was well fit with the CDF of the power-law function G(s) = s−3/2 (dashed line), inagreement with previous studies on avalanche dynamics in nonhuman data (2).

1. Klaus A, Yu S, Plenz D (2011) Statistical analyses support power law distributions found in neuronal avalanches. PLoS ONE 6(5):e19779.2. Petermann T, et al. (2009) Spontaneous cortical activity in awake monkeys composed of neuronal avalanches. Proc Natl Acad Sci USA 106(37):15921–15926.

Palva et al. www.pnas.org/cgi/content/short/1216855110 4 of 7

Page 11: PNAS-2013-Palva-3585-90 (1)

Fig. S4. Artificial LRTC or power-law avalanche dynamics are not produced by residual environmental or instrument noise, the mixing of neuronal signals atthe scalp level, or the source reconstruction methods. We used forward modeling to simulate virtual M/EEG recordings of uncorrelated brain activity andidentically preprocessed empty-room MEG recordings to probe environmental and instrument noise. These M/EEG and MEG datasets then were inversemodeled with the inverse operators and methods used elsewhere in the data analyses as well. The 10-Hz LRTC scaling exponents of both (A1) forward-modeleddata (β = 0.52 ± 0.001) and (A2) those recorded by an empty scanner (β = 0.53 ± 0.002) were close to the theoretical DFA scaling exponent for white noise (β =0.5). The avalanche lifetime and size distributions are exponentials for both (B1) forward- and inverse-modeled white noise (r2lifetime = 0.991, r2size = 0.994) and(B2) for inverse-modeled scanner noise (r2lifetime = 0.991, r2size = 0.989).

Fig. S5. As in Fig. 2, the scaling-law exponents of behavioral and neuronal LRTCs also are correlated with avalanche κ-indices (A and B). Note the consistencyof the results within subjects in task (circles) and rest (triangles) conditions; unique colors/fills indicate individual subjects.

Palva et al. www.pnas.org/cgi/content/short/1216855110 5 of 7

Page 12: PNAS-2013-Palva-3585-90 (1)

Fig. S6. Dynamics of interheartbeat (RR) intervals are scale-free and their scaling exponents are correlated during both task and rest with behavioral andneuronal exponents so that the correlation with behavior is mediated by correlation with neuronal LRTCs. (A) Time series of RR intervals of 10-min recording ofa representative subject shows fractal-like dynamics. (B) DFA plot for the time series of RR intervals (A) reveals salient LRTCs (βRR, scaling exponent of RR timeseries; βref, reference value for random data). (C) The behavioral LRTC scaling exponents, βbehav., and (D) the LRTC scaling exponents of 10-Hz neuronal os-cillations, β, were correlated with those of RR intervals across subjects (P < 0.04 and P < 0.03, respectively; Pearson correlation test) (E) Whereas the correlationsamong behavioral, neuronal, and heartbeat scaling exponents were significant, partial correlation analysis suggested that correlation of heartbeat withbehavior was indirect and mediated by neuronal LRTCs. (F–H) RR, neuronal, and behavioral LRTC data as in D, E, and F, respectively. Red (not significant) andgreen (significant) correlation coefficients indicate the values of partial correlations and show that the correlation between neuronal and behavioral LRTCexponents remains significant when the heartbeat LRTC is factored out, but not vice versa. Gray numbers denote the pairwise correlation coefficients.

Fig. S7. Motor responses do not affect the LRTC scaling exponents of the sensorimotor cortex during task performance or their correlation with behavioralLRTCs. (A) Thirty-second 10-Hz oscillation amplitude fluctuations were averaged across the corresponding time series from each subarea of the bilateral sen-sorimotor cortex (SMC; comprising the pre- and postcentral gyri and the central sulcus of the Destrieux atlas) of a representative subject. Arrows indicate buttonpresses. (B) Motor response averaged in a 0–1.5-s peri–button press time window across all button presses of the representative subject composed of a transientamplitude suppression after the button press. (C) To investigate the contribution of motor responses on SMC amplitude dynamics and LRTCs at the group level,we isolated for each subject the spontaneous amplitude time series of the bilateral SMC (as in A) and estimated the mean amplitude dynamics associated withmotor responses (as in B). (C, a) Post hoc DFA analysis of the intact SMC amplitude time series reproduced the brain–behavior correlation of the LRTC scalingexponents (r = 0.70, P < 0.006). We then used three approaches to examine the contribution of the motor responses. (C, b) First, we removed and interpolatedthe amplitude time series in the time windows of the motor responses. This decreased the brain–behavior correlation only slightly (r = 0.68, P < 0.007) from thevalue observed for the original time series. (C, c) Second, we added mean motor responses to the original time series; if the motor responses biased the originaldata toward a higher correlation, it should be observed here. However, we found that this manipulation did not change the brain–behavior correlation at all (r =0.70, P < 0.006). (C, d) Third, we used frequency-domain phase randomization to shuffle the temporal structure of the original SMC amplitude time series whilemaintaining an identical power spectrum; then, we superimposed the individual mean response waveform into individual response latencies. If the motor re-sponses per se introduced an artificial brain–behavior correlation in the scaling-law exponents, it should be reproduced here. However, there was no significantcorrelation between the scaling exponents of these time series with those of behavior (r = 0.11, P > 0.72). These three tests thus convergently indicate that evenin SMC, the motor responses do not introduce artificial brain–behavior correlations among the scaling exponents. **P < 0.01.

Palva et al. www.pnas.org/cgi/content/short/1216855110 6 of 7

Page 13: PNAS-2013-Palva-3585-90 (1)

Fig. S8. Neuronal LRTCs measured during the performance of an auditory or visual task are correlated with behavioral LRTCs measured in a separate sessionand with the other (visual or auditory, respectively) modality. Together with the correlation of resting-state neuronal LRTCs with behavioral task-state LRTC(Fig. 2A), these data suggest that the scaling exponents reflect individual phenotypic brain dynamic characteristics and are relatively independent of thespecific task at hand. (A) Scaling exponents of behavioral LRTCs in the auditory task are correlated with the neuronal scaling exponents in the visual task, and(B) vice versa. These plots are comparable with those in Fig. 2A, in which the behavioral and neuronal scaling exponents averaged across conditions are highlycorrelated.

Palva et al. www.pnas.org/cgi/content/short/1216855110 7 of 7