6
Spontaneous cortical activity in awake monkeys composed of neuronal avalanches Thomas Petermann a , Tara C. Thiagarajan a , Mikhail A. Lebedev b , Miguel A. L. Nicolelis b , Dante R. Chialvo c , and Dietmar Plenz a,1 a Section on Critical Brain Dynamics, National Institute of Mental Health, Bethesda, MD 20892; b Department of Neurobiology, Center for Neuroengineering, Duke University, Durham, NC 27710; and c Department of Physiology, Northwestern University, Chicago, IL 60611 Edited by Eve Marder, Brandeis University, Waltham, MA, and approved July 16, 2009 (received for review April 16, 2009) Spontaneous neuronal activity is an important property of the cerebral cortex but its spatiotemporal organization and dynami- cal framework remain poorly understood. Studies in reduced systems—tissue cultures, acute slices, and anesthetized rats— show that spontaneous activity forms characteristic clusters in space and time, called neuronal avalanches. Modeling studies suggest that networks with this property are poised at a critical state that optimizes input processing, information storage, and transfer, but the relevance of avalanches for fully functional cerebral systems has been controversial. Here we show that ongoing cortical synchronization in awake rhesus monkeys carries the signature of neuronal avalanches. Negative LFP deflections (nLFPs) correlate with neuronal spiking and increase in amplitude with increases in local population spike rate and synchrony. These nLFPs form neuronal avalanches that are scale-invariant in space and time and with respect to the threshold of nLFP detection. This dimension, threshold invariance, describes a fractal organization: smaller nLFPs are embedded in clusters of larger ones without destroying the spatial and temporal scale-invariance of the dy- namics. These findings suggest an organization of ongoing cortical synchronization that is scale-invariant in its three fundamental dimensions—time, space, and local neuronal group size. Such scale-invariance has ontogenetic and phylogenetic implications because it allows large increases in network capacity without a fundamental reorganization of the system. neuronal synchronization resting state rhesus monkey spontaneous activity ongoing activity T he cerebral cortex displays spontaneous activity, also known as ‘ongoing’ or ‘resting state’ activity, which persists in the absence of sensory stimuli or motor outputs. The ongoing activity is a robust feature of cortical dynamics as it is only modulated to a small extent by stimulus presentation (1), but contributes significantly to the large variability observed in stimulus responses (2–5). In fact, ongoing activity has been found to reflect multiple aspects of neuronal processing. The activity is similar to that observed during stimulus presentation (1, 6–8), incorporates previously acquired information (9), and carries information about the underlying neuronal network [for review see (10)]. Indeed, correlations during resting state activity are altered in disease states such as schizo- phrenia or chronic pain (11, 12), which raises the question whether there is a general framework that describes the statistics in the spatiotemporal organization of this dynamics. Recently, we found that spontaneous cortical activity in slice cultures, acute slices, and in the anesthetized rat in vivo has a scale-invariant dynamics called neuronal avalanches (13–15). These spontaneous bursts of synchronized activity occur in clusters of sizes s (where s is the number of active sites in an electrode array) that distribute according to a power law with exponent : Ps s [1] where usually lies between -1 and -2. Thus, given a burst of size s, a burst double that size will occur 2 times less often, independent of s. More generally, given synchronized activity bursts of sizes s and k s (where k is a constant), there will always be a fixed ratio k between the corresponding probabilities of occurrence. This scale- invariant dynamics has several key properties: (i) it requires bal- anced excitatory and inhibitory synaptic transmission (13, 15, 16); (ii) it arises in vitro and in vivo when superficial layers form during development (15); and (iii) it is maintained over many weeks in the isolated cortex in the absence of input (17), suggesting that the dynamics is an intrinsic property of the cortex. Importantly, neu- ronal avalanches resemble the dynamics observed in other systems at the critical point between an ordered and disordered dynamics, commonly referred to as the critical state (18 –20). Simulations of neuronal networks show that in the critical state the dynamic range of inputs is optimized, along with the amount of information that can be processed (21), stored (22), and transferred (13). To date, neuronal avalanches have been described only in reduced preparations from rat brains. These limitations have led to doubts about their relevance to intact and fully functional nervous systems. Accordingly, the present study examined the generality of avalanche dynamics in awake rhesus monkeys. We demonstrate that ongoing activity is composed of neuronal avalanches by demon- strating its invariance to various scale transformations in space, time, and a third dimension, the nLFP amplitude threshold. This scale invariance is in line with predictions from a critical state dynamics (19, 20); it implies that the ongoing functional linking of cortical sites occurs across all distances and on all time scales with a fractal ordering in the size of participating neuronal groups. Results We measured ongoing LFP and unit activity continuously for approximately 40 min using microelectrode arrays in the cerebral cortex of two awake rhesus monkeys, which were seated in a monkey chair. External sounds were minimized. The monkeys were not required to do any behavioral task or respond to a particular stimulus, while they maintained their body posture and general vigilance. The array configurations covered a range of scales, spanning 64 and 36 mm 2 of the primary motor and premotor areas in two monkeys (Fig. 1A; one array in monkey A, four arrays in monkey B). The simultaneously recorded LFP traces revealed irregular deflections typical of wakefulness (Fig. 1B). Spike- triggered LFP averages demonstrated that nLFP peak times cor- related with unit activity (Fig. S1; 92% of units M1 left monkey A; 88 4% of units monkey B) and nLFP peak amplitudes increased with instantaneous local firing rate (number of spikes from all units Author contributions: T.P. and D.P. designed research; M.A.L. and M.A.L.T. performed research; D.R.C. contributed new reagents/analytic tools; T.P. and D.P. analyzed data; and T.P., T.C.T., and D.P. wrote the paper . Conflict of interest statement: National Institutes of Health international patent filing PCT/US2006/031884 ‘‘Neuronal Avalanche Assay’’ This article is a PNAS Direct Submission. Freely available online through the PNAS open access option. 1 To whom correspondence should be addressed. E-mail: [email protected]. This article contains supporting information online at www.pnas.org/cgi/content/full/ 0904089106/DCSupplemental. www.pnas.orgcgidoi10.1073pnas.0904089106 PNAS September 15, 2009 vol. 106 no. 37 15921–15926 NEUROSCIENCE

Spontaneous cortical activity in awake monkeys … · Spontaneous cortical activity in awake monkeys composed of neuronal avalanches Thomas Petermanna, Tara C. Thiagarajana, Mikhail

Embed Size (px)

Citation preview

Spontaneous cortical activity in awake monkeyscomposed of neuronal avalanchesThomas Petermanna, Tara C. Thiagarajana, Mikhail A. Lebedevb, Miguel A. L. Nicolelisb, Dante R. Chialvoc,and Dietmar Plenza,1

aSection on Critical Brain Dynamics, National Institute of Mental Health, Bethesda, MD 20892; bDepartment of Neurobiology, Center for Neuroengineering,Duke University, Durham, NC 27710; and cDepartment of Physiology, Northwestern University, Chicago, IL 60611

Edited by Eve Marder, Brandeis University, Waltham, MA, and approved July 16, 2009 (received for review April 16, 2009)

Spontaneous neuronal activity is an important property of thecerebral cortex but its spatiotemporal organization and dynami-cal framework remain poorly understood. Studies in reducedsystems—tissue cultures, acute slices, and anesthetized rats—show that spontaneous activity forms characteristic clusters inspace and time, called neuronal avalanches. Modeling studiessuggest that networks with this property are poised at a criticalstate that optimizes input processing, information storage, andtransfer, but the relevance of avalanches for fully functionalcerebral systems has been controversial. Here we show thatongoing cortical synchronization in awake rhesus monkeys carriesthe signature of neuronal avalanches. Negative LFP deflections(nLFPs) correlate with neuronal spiking and increase in amplitudewith increases in local population spike rate and synchrony. ThesenLFPs form neuronal avalanches that are scale-invariant in spaceand time and with respect to the threshold of nLFP detection. Thisdimension, threshold invariance, describes a fractal organization:smaller nLFPs are embedded in clusters of larger ones withoutdestroying the spatial and temporal scale-invariance of the dy-namics. These findings suggest an organization of ongoing corticalsynchronization that is scale-invariant in its three fundamentaldimensions—time, space, and local neuronal group size. Suchscale-invariance has ontogenetic and phylogenetic implicationsbecause it allows large increases in network capacity without afundamental reorganization of the system.

neuronal synchronization � resting state � rhesus monkey �spontaneous activity � ongoing activity

The cerebral cortex displays spontaneous activity, also known as‘ongoing’ or ‘resting state’ activity, which persists in the absence

of sensory stimuli or motor outputs. The ongoing activity is a robustfeature of cortical dynamics as it is only modulated to a small extentby stimulus presentation (1), but contributes significantly to thelarge variability observed in stimulus responses (2–5). In fact,ongoing activity has been found to reflect multiple aspects ofneuronal processing. The activity is similar to that observed duringstimulus presentation (1, 6–8), incorporates previously acquiredinformation (9), and carries information about the underlyingneuronal network [for review see (10)]. Indeed, correlations duringresting state activity are altered in disease states such as schizo-phrenia or chronic pain (11, 12), which raises the question whetherthere is a general framework that describes the statistics in thespatiotemporal organization of this dynamics.

Recently, we found that spontaneous cortical activity in slicecultures, acute slices, and in the anesthetized rat in vivo has ascale-invariant dynamics called neuronal avalanches (13–15). Thesespontaneous bursts of synchronized activity occur in clusters of sizess (where s is the number of active sites in an electrode array) thatdistribute according to a power law with exponent �:

P�s� � s� [1]

where � usually lies between -1 and -2. Thus, given a burst of sizes, a burst double that size will occur 2� times less often, independent

of s. More generally, given synchronized activity bursts of sizes s andk � s (where k is a constant), there will always be a fixed ratio k�

between the corresponding probabilities of occurrence. This scale-invariant dynamics has several key properties: (i) it requires bal-anced excitatory and inhibitory synaptic transmission (13, 15, 16);(ii) it arises in vitro and in vivo when superficial layers form duringdevelopment (15); and (iii) it is maintained over many weeks in theisolated cortex in the absence of input (17), suggesting that thedynamics is an intrinsic property of the cortex. Importantly, neu-ronal avalanches resemble the dynamics observed in other systemsat the critical point between an ordered and disordered dynamics,commonly referred to as the critical state (18–20). Simulations ofneuronal networks show that in the critical state the dynamic rangeof inputs is optimized, along with the amount of information thatcan be processed (21), stored (22), and transferred (13).

To date, neuronal avalanches have been described only inreduced preparations from rat brains. These limitations have led todoubts about their relevance to intact and fully functional nervoussystems. Accordingly, the present study examined the generality ofavalanche dynamics in awake rhesus monkeys. We demonstrate thatongoing activity is composed of neuronal avalanches by demon-strating its invariance to various scale transformations in space,time, and a third dimension, the nLFP amplitude threshold. Thisscale invariance is in line with predictions from a critical statedynamics (19, 20); it implies that the ongoing functional linking ofcortical sites occurs across all distances and on all time scales witha fractal ordering in the size of participating neuronal groups.

ResultsWe measured ongoing LFP and unit activity continuously forapproximately 40 min using microelectrode arrays in the cerebralcortex of two awake rhesus monkeys, which were seated in amonkey chair. External sounds were minimized. The monkeys werenot required to do any behavioral task or respond to a particularstimulus, while they maintained their body posture and generalvigilance. The array configurations covered a range of scales,spanning 64 and 36 mm2 of the primary motor and premotor areasin two monkeys (Fig. 1A; one array in monkey A, four arrays inmonkey B). The simultaneously recorded LFP traces revealedirregular deflections typical of wakefulness (Fig. 1B). Spike-triggered LFP averages demonstrated that nLFP peak times cor-related with unit activity (Fig. S1; 92% of units M1left monkey A;88 � 4% of units monkey B) and nLFP peak amplitudes increasedwith instantaneous local firing rate (number of spikes from all units

Author contributions: T.P. and D.P. designed research; M.A.L. and M.A.L.T. performedresearch; D.R.C. contributed new reagents/analytic tools; T.P. and D.P. analyzed data; andT.P., T.C.T., and D.P. wrote the paper .

Conflict of interest statement: National Institutes of Health international patent filingPCT/US2006/031884 ‘‘Neuronal Avalanche Assay’’

This article is a PNAS Direct Submission.

Freely available online through the PNAS open access option.

1To whom correspondence should be addressed. E-mail: [email protected].

This article contains supporting information online at www.pnas.org/cgi/content/full/0904089106/DCSupplemental.

www.pnas.org�cgi�doi�10.1073�pnas.0904089106 PNAS � September 15, 2009 � vol. 106 � no. 37 � 15921–15926

NEU

ROSC

IEN

CE

at each electrode within 10 ms) [Fig. 1C; monkey A: F (3, 73) � 26,P � 10�11 M1left; monkey B: F (3, 53) � 4.0, P � 0.01 M1left; F (3,26) � 6.0, P � 0.003 M1right; F (3, 38) � 10.0 PMdleft, P � 5�10�5;F (3, 37) � 23, P � 10�8 PMdright; one-way ANOVA; �0.27 SD formonkey A; �0.16 � 0.05 SD for monkey B per 100 Hz; R �0.55�0.7; P � 0.0001 all arrays; linear regression]. nLFP peakamplitude also increased with local synchronization of unit activity(Fig. 1D; number of units with at least 1 spike/10 ms; see Fig. S2;2 vs. 3 units: 0.6 SD monkey A; 0.3 � 0.1 SD monkey B; t test; P �0.0005). These locally synchronized spikes were significantly differ-ent from chance occurrence and were correlated across corticalsites over many seconds in time (Fig. S2). Thus, the peak nLFPamplitudes during ongoing activity provides a good approximationof the instantaneous firing rate and degree of spike synchrony of thelocal neuronal population.

nLFPs in Vivo Organize into Neuronal Avalanches. We then groupednLFPs into spatiotemporal clusters based on their occurrence insuccessive time bins, regardless of the electrode on which theyoccurred (Fig. 2A). Such temporal grouping of nLFPs is supportedby the strong crosscorrelation between nLFPs at different elec-trodes (Fig. 2B). The beginning of a cluster was defined by the

occurrence of a time bin (chosen at �t � 4 ms) with at least onenLFP and the end by the next encounter of an empty time bin.Clusters had size s equal to their number of nLFPs and a lifetime

PMdright

M1right

PMdleft

M1leftM1left

Monkey A Monkey B

10 mm

A

10SD200 ms

B

C

0

1.4

Local firing rate (Hz) Local firing rate (Hz)

0

0.8

Local rate -> nLFP

100 100300 300400 400200 200

nLFP

(SD

)

100

300200

100

1

300

3

200

2

PMdPMd

M1M1

rightleft

rightleft

M1left

Monkey A

Monkey B

D

0

1.4

units/10 ms (n) units/10 ms (n)

0

0.8Local synchrony -> nLFP

11

1

33

3

22

2

nLFP

(SD

)

Fig. 1. nLFP peak amplitude correlates with local rate and synchrony of unitactivity. (A) Position of microelectrode arrays in primary motor (M1) anddorsolateral premotor cortex (PMd) of two rhesus monkeys (monkey A: M1left;monkey B: M1left,right and PMdleft, right). Circles: electrode positions with LFPand unit activity. (B) Segment of simultaneous LFP recordings scaled in SD(PMdright). (C) nLFP peak amplitude increases with local firing rate (number ofspikes per 10 ms at each electrode). Average (� SE) for all electrodes (left:monkey A; right: four arrays monkey B). Insets: Example average LFP triggeredby 10 ms periods with one, two, or three spikes at a single electrode. Left:M1left. Right: PMdright. (Scale bars, 0.4 SD left; 0.2 SD right; 100 ms.) (D) nLFPamplitude increases with number of units that contribute at least one spikeper 10 ms, left inset) that is, local spike synchrony (see Fig. S2D for significanceof two vs. three units). Right inset: Example average LFP triggered on one, two,or three active units (PMdright, Scale bar, 0.2 SD; 100 ms).

Fig. 2. nLFP peaks cluster as neuronal avalanches that are invariant totemporal scale. (A) Raster of nLFP peak occurrences (dots) extracted at �2SD(M1left; monkey A) with a period shown at higher temporal resolution (bot-tom). nLFP peaks were grouped into a cluster (gray) when they occurred in thesame or contiguous bins of duration �t. Clusters ended when an empty bin wasencountered. Spatiotemporal size (i.e., number of nLFPs or electrode activa-tions) and lifetime (in multiples of �t) for the three identified clusters areindicated at the bottom. (B) Mean cross-correlation function for nLFPs be-tween electrodes (�t � 2 ms). Top: Peak at around time 0 reveals strongcorrelation between simultaneous and successive nLFPs. Bottom: Significantcross-correlation persists over many seconds (Right side from top plotted inloglog-coordinates). Broken lines: correlation obtained from rasters withrate-matched, but uniformly randomized inter nLFP times. (C) Cluster sizesdistribute according to a power law (straight line in log-log coordinates) up toa cut-off determined by the total number of electrodes (empty circles, monkeyA; M1left, cut-off � 32; see F). Slope � varies systematically with �t. Brokenlines, squares: Exponentially decaying size distributions from correspondingtime-shuffled rasters (R0.999). (D) Relationship between slope � and �tplotted in log-log coordinates for each array (�2 SD; solid lines: linear regres-sion). In vitro: replotted from (13). (E) Collapse in cluster size distributionsobtained for different �t using the scaling relationship (�t/4ms)��. � � 0.15was taken from D (monkey A; M1left). Scaling exponent 0.22 minimizes devi-ations in P(s) for visualization purposes. (F) The cluster sizes for M1left distrib-ute according to a power law up to a cut-off determined by the number ofelectrodes of the array (monkey A; M1left, empty circles). The cut-off shiftsfrom 32 to 16 electrodes if only half of the array is considered.

15922 � www.pnas.org�cgi�doi�10.1073�pnas.0904089106 Petermann et al.

T, that is, duration, equal to the number of bins times �t. We foundthat P(s) distributed according to the power law in Eq. 1. The powerlaw is readily identifiable as a linear relationship in log-log coor-dinates, where the slope approximates the power law exponent �(Fig. 2C, open circles). For both monkeys and all arrays, the powerlaw ranged from the lowest cluster size of 1 electrode, i.e., s � 1, tothe maximal number of electrodes in the array (s � 32, monkey A;s � 16, monkey B).

The power law identifies a scaling relationship in the spontaneousformation of long-range spatial correlations in the system. This canbe seen by the greater probability of occurrence of spatiotemporallylarge clusters relative to random or independent nLFP groupings.Indeed, after random shuffling of nLFPs in time, nLFP clustersdistribute exponentially in sizes (R 0.999, all cases; Fig. 2C,broken lines) indicating a rapid decrease in the probability of largerclusters, that is, long-range spatial correlations, and a loss of thescaling relationship.

While the power law was preserved for different choices of �t, theabsolute value of its corresponding exponent � decreased when �tincreased (Fig. 2 C and D). This relationship was previously foundin vitro as

� � �t�, [2]

with � � -0.16 � 0.01 (13). A similar relationship between � and�t was found in vivo with a range of � between �0.04 and �0.16(Fig. 2D, monkey A; M1left � � �0.15 � 0.01, R � �0.98, P � 0.001;monkey B: M1right � � �0.16 � 0.1, M1left � � �0.11 � 0.02,PMdright � � �0.09 � 0.02, PMdleft � � �0.04 � 0.01, R � �0.94 �0.04, P � 0.01 all arrays; linear regression). Formulae 1 and 2 allowfor the size distributions P(s) obtained at different �t to be collapsedusing a general scaling ansatz (see Materials and Methods) demon-strated for M1left of monkey A in Fig. 2E. The collapse experi-mentally demonstrates a temporal scale-invariance of the power lawin cluster sizes for temporal resolutions of 2–32 ms.

As reported previously in vitro (13), the maximal number ofelectrodes on the array marked the cut-off of the power law becausecortical sites rarely participated in a cluster more than once. Indeed,this cut-off shifted systematically as electrodes were removed fromthe analysis (Fig. 2F) demonstrating finite-size scaling typical forcritical state dynamics (19).

The power law distribution of nLFP clusters is specific to nLFPfluctuations, is not predicted by a 1/f�-decay in LFP frequencypower (23) or volume conduction, and is cortical depth specific.First, the power law is not observed for other signal characteristicsof the LFP, such as all baseline crossings (Fig. S3A). We note thatnLFPs are often followed by positive deflections, which thereforecannot serve as an independent control. Second, phase-shuffling ofthe LFP, which changes the temporal relationship between fre-quencies at and between sites without affecting the 1/f� scaling,destroys the power law in nLFP cluster sizes (Fig. S3 B and C).Third, nLFP clusters are most often composed of spatially non-contiguous nLFPs irrespective of the temporal resolution chosen, incontrast to what is predicted for either volume conduction oroverlapping signal detection (Fig. S3 D–F). Fourth, the relationshipbetween nLFP and unit activity as shown in Fig. 1C as well as thepower law in nLFP cluster sizes is only present in more superficiallayers of the cortex demonstrating layer specificity of the organi-zation (Fig. S4), in line with the localization of neuronal avalanchesto superficial layers in the acute slice (16), slice culture (17), andanesthetized rat (15).

The Avalanche Organization Is Invariant to nLFP Detection Threshold.A remaining and previously unexplored free parameter in thedescription of nLFP clusters is the threshold z to detect nLFPs.Because nLFP peak amplitudes increase with local firing rate andsynchrony (Fig. 1C and D), a threshold operation removes asignificant number of suprathreshold, active sites from the analysis.

In fact, when z increased from �1.5 to �4.5 SD the number ofextracted nLFPs decayed by almost two orders of magnitude (Fig.3A; exponential decay, slope:-2.4 and �3.5 � 0.4; monkey A and Brespectively; R � 0.999). Importantly, the power law in cluster sizedistribution remained unchanged [Fig. 3B; �t � 4 ms; P 0.05;Kolmogorov-Smirnov (KS)-test] including � which remained con-stant as z increased (Fig. 3C; average slope: 0.07 � 0.06/SD; P 0.1all arrays) and accordingly the collapse of P(s) for different z isachieved with � � 0 (Fig. 3B; Eq. 2). Thus, large nLFPs on thearray are parsed into spatiotemporal clusters in exactly the sameway after smaller nLFPs are removed. This demonstrates that theneuronal avalanche dynamics is independent of the chosen thresh-old. In fact, an avalanche identified at a low threshold fragmentsinto multiple avalanches as the threshold increases, but the averagesize stays close to its original value demonstrating similarity of nLFPamplitudes within avalanches beyond chance (see Fig. S5). As acontrol, we demonstrate that when nLFPs were extracted at z ��1.5 SD and then randomly removed from this raster to arrive attotal nLFP numbers that corresponded to a given threshold z (Fig.3A), the power law in cluster size distribution readily collapsed withan increase in z (P � 0.05; KS-test; Fig. 3D; see also Fig. S6 formonkey B PMdright).

We visualized the self-similar, hierarchical organization of nLFPamplitudes by replotting nLFP rasters at different temporal andamplitude scales. At z � �4.5 SD, the rate of nLFPs was low, butviewing the data at a low temporal scale revealed the spatiotem-poral clustering of nLFPs (‘columns’, Fig. 3E). On the other hand,clusters that appeared homogenous at high threshold and lowtemporal resolution, ‘spawned’ new clusters at successively lowerthresholds and increased temporal resolutions (Fig. 3E; A 3 A,B 3 B).

These results clearly demonstrate that proper subsampling of thescale-invariant activity, that is, all events above a given threshold,differs from randomized subsampling of synchronized activity,which underestimates the probability of large clusters (24). Forexample, when estimating local synchrony using the relatively fewextracellular units discriminated successfully at individual elec-trodes, numerous local, synchronized events as reflected in thenLFP, will not be captured, establishing a condition of partial,random subsampling. Indeed, the resulting distributions of spatio-temporal clusters based on local unit synchrony where heavy-tailedand, while significantly different from an exponential distribution(Fig. S7A), deviated from a power law. The deviation found is in linewith the finding of Fig. 3D and Fig. S6, where a scale-invariantorganization of nLFPs was subjected to random event removals (cp.transition from the original organization at �1.5 SD to an nLFPraster with equivalent nLFP numbers as found at �2 SD and �2.5SD by random removal of nLFPs).

Avalanche Lifetimes Are Scale Invariant in Time and nLFP Threshold.For dynamical systems in the critical state, scale-invariance inavalanche sizes is accompanied by scale-invariance in avalancheduration, that is, the lifetime T (18–20). In vivo, T varied greatlyeven for avalanches of any given size, although large avalanchestended to have longer lifetimes (Fig. 4A). As reported previously invitro (13, 16), avalanche lifetimes in vivo were significantly longerthan when nLFPs were randomized in time (Fig. 4B; broken lines;P � 0.05; KS-test; inter nLFP time shuffling) and lifetime distri-butions collapsed to a similar distribution when T was scaled inmultiples of �t (P 0.05; KS-test). Futhermore, similar to ava-lanche sizes, the lifetime distributions were also scale-invariant withrespect to z (Fig. 4C; P 0.05; KS-test).

DiscussionThe present data show that ongoing cortical activity in awakemonkeys is composed of neuronal avalanches, indicating the uni-versality of these dynamics in the context of cortical activity. nLFPscorrelate in time and amplitude with local synchronized spike

Petermann et al. PNAS � September 15, 2009 � vol. 106 � no. 37 � 15923

NEU

ROSC

IEN

CE

activity and organize into spatiotemporal clusters with a power-lawdistribution in sizes. Not only occur these clusters irrespective ofspatial and temporal scale, as shown previously for reduced prep-arations (13–15), but they are also scale-invariant with respect todetection threshold.

Scale Invariance in Time, nLFP Amplitude, and Finite-Size Scaling IsConsistent with a Critical State. Power laws in cluster-size distribu-tions can result from numerous mechanisms (14). However, theself-similarity at different scales suggests critical state dynamics asthe most likely one. Self-similar hierarchies of events across time,space, and amplitude are universally found in systems that are neara critical point of a phase transition between order and disorder

at z � �1.5SD (original data, monkey A). Open circles: nLFPs are randomlyremoved from the original �1.5SD raster to arrive at rasters with equivalentnumber of nLFPs as found for z � �2, �2.5, . . ., �4.5SD (see A). Correspondingcluster size distributions rapidly deviate from a power law. (E) Avalanches‘spawn’ new avalanches when lowering threshold z (�t � constant). Top: nLFPraster period at z � �4.5 SD (nLFP clusters appear ‘columnar’). Middle: Rasterof expanded time period from top (broken lines) at z � �3 SD and rescaled fornLFP amplitude (color bar). Note segment A (gray) contains more nLFPs someof which form a new cluster A’ (ellipsoid). Bottom: Raster of expanded timeperiod from middle (broken line) for z � �1.5 SD rescaled in amplitude. Spawnof new cluster in B (middle, gray) indicated by B’ (ellipsoid).

Fig. 3. The avalanche organization is scale-invariant to nLFP amplitudethreshold. (A) Decay in total number of nLFPs with increase in threshold z. (B)Power law distributions in sizes are scale-invariant with respect to nLFPamplitude. Size distributions P(s) plotted for different z and two arrays (�t �4 ms). (C) The slope � does not depend on z (five arrays; color code in A). (D)Random removal of nLFPs differs from threshold-dependent nLFP removaland does not maintain scale- invariance. Filled circles: cluster size distribution

Fig. 4. The scale-invariance of avalanche lifetimes. (A) Avalanche lifetime Tvaries widely for a given size s, although longer avalanches tend to have largersizes. P(T,s): Probability density. (B) Lifetime distributions (open circles) col-lapse when scaled in multiples of �t. This scale-invariance is lost for time-shuffled rasters (broken lines, open squares). (C) Lifetime distributions ob-tained for z � �1.5SD to �4.5SD are similar demonstrating scale-invariance innLFP amplitude.

15924 � www.pnas.org�cgi�doi�10.1073�pnas.0904089106 Petermann et al.

(18–20). Moreover, such a critical state appears to be advantageousfor information processing (13, 21, 22).

Alternative dynamic scenarios, where nLFP clusters of small sizesarise by virtue of different mechanisms compared to those respon-sible for medium- or large-sized clusters, are difficult to translateinto different scales. For example, the removal of 90% of nLFPsbelow a given amplitude threshold (Fig. 3B) demonstrates that theremaining, much fewer, but large nLFPs organize like smallernLFPs. This property suggests that the generation of large clustershas similar mechanisms as small clusters. Similarly, an equivalentstatistical structure after removal of full sections of electrodesimplies a common mechanism in space. An analogous argumentwould apply for different temporal scales.

A critical state, as implied by the neuronal avalanche dynamics,does not arise arbitrarily in cortical networks. Instead, this propertyof spontaneous activity emerges when superficial layers of cortexfirst differentiate from the cortical plate during development (15).Given the dependence of neuronal avalanches on the balance of fastsynaptic excitation and inhibition and the neuromodulator dopa-mine (13, 15–17), it seems clear that these properties arise from thespecific organization of cortical circuitry.

Implications for Large-Scale Cortical Networks. The scale invariancein the spontaneous formation of clusters has important implicationsfor large-scale cortical networks. A large network can support alarge diversity of cluster sizes, including those that percolate themaximal extent of the system. This indicates that the maximalcluster size, that is, the long-range correlations established by sucha cluster, is constrained by the system size and not system dynamics.Consistent with this, in our experiments, the maximally measuredcluster size or area occupied by a cluster was constrained only by thenumber of electrodes. In these recordings from monkeys, whereinter-electrode distances were up to 2 mm, the scale invariance wasdemonstrated over 64 mm2, the area covered by the largest array.This greatly extends the range of scale invariance shown previously,where the arrays and correspondingly the maximal cluster, spanneda maximum of 2 mm2 (13, 15). Furthermore, the large number ofregions and conditions where neuronal avalanche dynamics havenow been observed—motor, premotor, and somatosensory cortexin awake monkeys (present data), somatosensory cortex of anes-thetized rats (15), and both rat somatosensory and prefrontal cortexin vitro in acute slices and slice cultures (13, 17)—indicates that thisdynamics is likely to be a general property of the cerebral cortex.This also suggests that the dynamics may extend across multipleregions rather than being restricted to specific regions.

As has been shown in well-identified networks, a target dynamicscan be achieved by a plethora of different network realizations (25).Indeed, the finding of scale invariance does not exclude long-rangeor short-range heterogeneous corticocortical projections. Numer-ical simulations have shown that neuronal avalanches can arise atthe critical state in models with scale-free (26), fully connected (27),random (28), and nearest-neighbor (18, 26) topologies, although ineach case the conditions to reach criticality can be different. Incortical cultures, the neuronal avalanches establish a functionalsmall-world architecture with specific and highly diverse point-to-point connectivity between cortical sites that is, the networkrepresenting these site activations is densely interconnected, glo-bally as well as locally (29). The superficial layers where thesedynamics arise in cultures are also where long-range corticocorticalconnections are most prominent and therefore where informationis likely to be integrated across multiple sensory modalities. Asdiscussed above, a hallmark of criticality is the fact that the lengthof correlations (in space and time) is only limited by the size of thesystem (18, 19, 30), for instance in the present case the extent of thecorrelations goes as the size of the largest avalanche, which divergeswith system size. Thus an expanding critical system increases thenumber, the diversity and the maximal cluster size without thenecessity to rewire existing connections. This property has signif-

icant implications from both a developmental as well as an evolu-tionary perspective. The ontogenetic and phylogenetic increases incortical volume would naturally allow for greater functional capac-ity without requiring a fundamental and sophisticated reorganiza-tion of the system (31). Importantly, the superficial layers of cortex,where this dynamics has been identified in vitro and in the anes-thetized rat in vivo (15–17), expand most dramatically during bothindividual development and during the evolution of the cerebralcortex in advanced mammals (32).

Functional Significance. Our results demonstrate that the neuronalavalanche dynamics in awake animals is remarkably similar to thatarising in vivo early during development (15) and in explant corticalnetworks that mature in the absence of external inputs (13, 17).Thus, the phenomenon described previously for reduced systemsapplies to intact ones, as well. The present results also generalizeavalanche dynamics across species (including to a primate) andindividuals, as well as the broad diversity of cortical areas listedabove. Taken together, these findings strongly suggest that critical-ity is a generic property of cortical network activity. We note thatthe resting condition of the monkeys in the current experiments isan approximation of the idle state, which would be the propercondition for measuring spontaneous activity as found in in vitropreparations and it remains to be compared explicitly to highlyactive states as for example found during precisely controlledbehavioral paradigms.

An intuitive functional understanding of the critical state can beobtained by analyzing the internal organization of avalanches, thatis, spatiotemporal cascades, within the context of a branchingprocess. We previously demonstrated that during an avalanche, onaverage, one nLFP (the ancestor) is followed by one nLFP in thenear future (the descendant) in line with predictions from theoryfor a critical system (13). The critical branching ratio �, which is thenumber of descendants divided by the number of ancestors is thusprecisely 1. For subcritical processes (��1), cascades terminateprematurely, whereas for supercritical processes (�1), they willquickly engage the whole system in a non-selective manner (similarto an uncontrolled nuclear chain reaction). If systems try tomaximize the likelihood of linking distant sites through cascades,while avoiding massive, non-selective activation, criticality is theoptimal scenario, and it is from this perspective that the functionalsignificance of the critical state expressed as neuronal avalanchesneeds to be appreciated. Specifically, modeling studies have shownthat critical branching maximizes the mutual information betweengroups of neurons (13) as well as the ability of a system to respondwith diverse internal states to a wide range of inputs (21).

The cerebral cortex in all mammalian species follows a commonarchitectural blueprint and cortical areas must perform, at somelevel of analysis, an alike function. This function must be moregeneral than the motor, sensory, and associational roles commonlyascribed to particular cortical areas. Thus, independently of the typeof inputs that each region of cortex receives, the critical state wouldendow each spatiotemporal column or patch of cortex with amaximal dynamic range of processing, along with maximizinginformation transfer to other cortical areas. In addition, sincecriticality exhibits a mixture of ordered and disordered excitationpatterns, neuronal networks at the critical state can generate thelargest repertoire of dynamical configurations in a flexible manner.It would be reasonable to think that in some instances thosenetworks with largest repertoires were selected over othersthroughout evolution. As such, the principles of synchronous,spontaneous activity described here could underlie all corticalfunctions. The present results open the way to testing this theoryand for applying the knowledge gained from reduced systems toawake, behaving animals.

Petermann et al. PNAS � September 15, 2009 � vol. 106 � no. 37 � 15925

NEU

ROSC

IEN

CE

Materials and MethodsMicroelectrode Array Positioning. The ACUC of Duke University approved allprocedures. Arrays of monopolar tungsten electrodes (30 �m in diameter, 1 M�impedance; 1-mm spacing) were chronically implanted into the cortex of twoadult rhesus monkeys (Macaca mulatta) (33). Monkey A was implanted with a64-electrode array in the left motor cortex (M1left; leg representation). Monkey Bwas implanted bilaterally with four 32-electrode arrays in the M1 arm represen-tation (M1left, M1right) and in dorsolateral premotor cortex (PMdleft, PMdright).Areas were identified by cortical landmarks, evoked LFP and unit responsesduring surgery. Electrodes were placed �1 mm deep for PMd and �1.5 mm forM1. Those electrode tips were aligned in a single plane of the array. To studylaminar properties, a staggered array was inserted 1–1.5 mm deep in the somato-sensory cortex of monkey A (S1left; Fig. S4; electrode tip separation: �300–500�m; inter-electrode distance: 1 mm). No postmortem histology was conducted tofurther specify the depth of electrode locations.

Data Acquisition. Simultaneous recordings were carried out in M1left and S1left

(monkey A) and the four arrays in M1 and PMd (monkey B). LFP and unit activitywassampledsimultaneously fromeveryotherelectroderesulting in32electrodesfor M1left and S1left for monkey A and 16 electrodes per array in monkey B usinga dual amplifier Plexon system (Fig. 1A and Fig. S4; filled circles). An epiduralstainless steel T-bolt, at least 20 mm from all recording areas, served as a commonground.

LFPs were sampled at 500 Hz and band-pass filtered (1–100Hz). Using anotch-filter, 60 Hz noise was removed. Extracellular spiking activity was sampledat 40 kHz and band-pass filtered with a two-pole low-cut and a four-pole high cutfilter at 0.4–8 kHz. Off-line unit discrimination was based on principal compo-nentanalysisandspike-templatematching(33).The68units resolved inM1left (32electrodes)firedonaverage11,300�9,700times(4.3�3.7Hz).Forthestaggeredarray in S1left of this monkey, 37 units were resolved on short- shank electrodesand 40 units on long-shank electrodes with 13,300 � 14,500 and 12,100 � 12,500discharges, respectively (4.3 � 4.8Hz and 5.2 � 5.6Hz). For the four arrays inmonkey B, on average 47 � 13 units were resolved (16 electrodes/array), whichfired 16,200 � 5,200 times during the recording (5.7 � 1.8 Hz). Unit firing rateswere not significantly different between monkeys A and B (Student t test,P � 0.3).

Removal of Slow-Wave Activity. To exclude sleep states, we first identified highamplitude, slow-wave activity characteristic of sleep-spindles (90% of power at1–10 Hz (34); 2-s sliding window at interval steps of 1 s). Periods with 50%channels showing this activity were removed (12.0% of the recorded data inmonkey A; 6.8 � 1.7% in monkey B).

nLFP Detection. LFP activity at each electrode was first normalized by subtractingits mean and dividing by its standard deviation (SD). The detection threshold for

negative LFP peaks at each electrode (nLFP) was determined as �nSD (where n �1.5, 2.0, . . . 4.5). Lower temporal resolutions for nLFPs were obtained by down-sampling the original LFP thereby taking every nth value at n 2 ms (n � 1, 2, 4,8, 16) before nLFP detection. nLFPs from each electrode were combined into nLFPrasters (Fig. 2A). Because such a down sampling can be done in n different ways(e.g., for n � 4, one can start with the 1st, 2nd, 3rd, or 4th data point), the resultingdistributions of nLFP cluster sizes and lifetimes (see Fig. 2A) were obtained as theaverages over the n nLFP rasters.

For controls, up to 80 time-shuffled nLFP rasters were created for each tem-poral resolution by uniformly redistributing nLFPs for each electrode randomly intime excluding slow-wave periods. This procedure de-correlates nLFPs in spaceand time, while maintaining the nLFP rate at each electrode.

For nLFPs, the cross-correlation functions between cortical sites were calcu-lated from nLFP rasters (temporal resolution 2 ms) for all pair-wise electrodecombinations and averaged for each array over a window of � 10 s. As controls,uniformly randomly timed nLFP rasters were used.

Scaling of Size Distributions. From 10,000–100,000 clusters contributed to eachdistribution, thus allowing for robust heavy-tail analysis. Distributions obtainedwith logarithmic binning were calculated for two parameters, temporal binwidth �t and negative threshold z (multiples of SD). The slope � of the power lawwas obtained for each value of �t or z using linear regression of log-transformeddistributions. The relationship between � and �t at a given value of z alsorevealed a power law relationship with an exponent � [Eq. 2, (13, 14)]. Accord-ingly, the argument s of a size distribution at a given temporal resolution �t canbe scaled by (�t/�t0)�� with �t0 constant [�t0 � 4 ms in our experiments; (19, 20)].This results in size distributions with a single slope � (�t0) for any �t. For graphs,thecollapseofdistributions intovirtually identicaloverplots isachievedbyscalingP(s) by �t� where � is estimated by minimizing deviations in P(s) for the variousdistributions.

Statistical Analysis. Distributions were compared using the two-sided Kolmog-orov-Smirnov (KS) test [Matlab (v7)]. ANOVA and linear regression were appliedto normal or log-transformed data (Statview v5.0; Origin v7). Significance wasestablishedatP less thanorequal to0.05andvaluesaregivenasmean�SDunlessstated otherwise.

For additional information see SI Text.

ACKNOWLEDGMENTS. We thank Drs. S.P. Wise and D. Leopold for comments onearlier manuscript versions, Dr. S. Pajevic for help with the scaling analysis. T.P.,T.C.T., and D.P. were supported by Division of Intramural Research Programs/National Institute of Mental Health. D.R.C. is supported by National Institute ofNeurological Disorders and Stroke. T.P. acknowledges the Swiss National ScienceFoundation (Grant No. PBEL2–110211). M.L. and M.N. provided the data sets andwere funded by Defense Advanced Research Projects Agency.

1. Fiser J, Chiu C, Weliky M (2004) Small modulation of ongoing cortical dynamics bysensory input during natural vision. Nature 431:573–578.

2. Arieli A, Sterkin A, Grinvald A, Aertsen A (1996) Dynamics of ongoing activity: Expla-nation of the large variability in evoked cortical responses. Science 273:1868–1871.

3. Kisley MA, Gerstein GL (1999) Trial-to-trial variability and state-dependent modulationof auditory-evoked responses in cortex. J Neurosci 19:10451–10460.

4. Azouz R, Gray CM (1999) Cellular mechanisms contributing to response variability ofcortical neurons in vivo. J Neurosci 19:2209–2223.

5. Fox MD, Snyder AZ, Zacks JM, Raichle ME (2006) Coherent spontaneous activity accountsfor trial-to-trial variability in human evoked brain responses. Nat Neurosci 9:23–25.

6. Leopold DA, Murayama Y, Logothetis NK (2003) Very slow activity fluctuations in monkeyvisual cortex: Implications for functional brain imaging. Cereb Cortex 13:422–433.

7. Tsodyks M, Kenet T, Grinvald A, Arieli A (1999) Linking spontaneous activity of singlecortical neurons and the underlying functional architecture. Science 286:1943–1946.

8. Kenet T, Bibitchkov D, Tsodyks M, Grinvald A, Arieli A (2003) Spontaneously emergingcortical representations of visual attributes. Nature 425:954–956.

9. Ohl FW, Scheich H, Freeman WJ (2001) Change in pattern of ongoing cortical activitywith auditory category learning. Nature 412:733–736.

10. Fox MD, Raichle ME (2007) Spontaneous fluctuations in brain activity observed withfunctional magnetic resonance imaging. Nat Rev Neurosci 8:700–711.

11. Baliki M, Geha PY, Apkarian AV, Chialvo DR (2008) Beyond feeling: Chronic pain hurtsthe brain disrupting the default-mode network dynamics. J Neurosci 28:1398–1403.

12. Liang M, et al. (2006) Widespread functional disconnectivity in schizophrenia withresting-state functional magnetic resonance imaging. NeuroReport 17:209–213.

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

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

15. Gireesh ED, Plenz D (2008) Neuronal avalanches organize as nested theta- and beta/gamma-oscillations during development of cortical layer 2/3. Proc Natl Acad Sci USA105:7576–7581.

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

17. Stewart CV, Plenz D (2007) Homeostasis of neuronal avalanches during postnatalcortex development in vitro. J Neurosci Meth 169:405–416.

18. Bak P, Paczuski M (1995) Complexity, contingency, and criticality. Proc Natl Acad SciUSA 92:6689–6696.

19. Jensen HJ (1998) in Self-Organized Criticality (Cambridge Univ Press).20. Stanley HE (1971) in Introduction to Phase Transitions and Critical Phenomena (Oxford

Univ Press, New York).21. Kinouchi O, Copelli M (2006) Optimal dynamical range of excitable networks at

criticality. Nat Phys 2:348–351.22. Legenstein R, Maass W (2007) Edge of chaos and prediction of computational perfor-

mance for neural circuit models. Neural Netw 20:323–334.23. De Los Rios P, Zhang Y-C (1999) Universal 1/f noise from dissipative self-organized

criticality models. Phys Rev Lett 82:472–475.24. Priesemann V, Munk MH, Wibral M (2009) Subsampling effects in neuronal avalanche

distributions recorded in vivo. BMC Neurosci 10:40.25. Prinz AA, Bucher D, Marder E (2004) Similar network activity from disparate circuit

parameters Nat Neurosci.26. De Arcangelis L, Perrone-Capano C, Herrmann HJ (2006) Self-organized criticality

model for brain plasticity. Phys Rev Lett 96:028107.27. Levina A, Hermann JM, Geisel T (2007) Dynamical synapses causing self-organized

criticality in neural networks Nat Phys 857–860.28. Teramae JN, Fukai T (2007) Local cortical circuit model inferred from power-law

distributed neuronal avalanches. J Comput Neurosci 22:301–312.29. Pajevic S, Plenz D (2008) Efficient network reconstruction from dynamical cascades

identifies small-world topology from neuronal avalanches PLoS Comp. Biol. 5,e1000271.

30. Bak P (1996) in How Nature Works: The Science of Self-Organized Criticality (Coper-nicus Books, New York).

31. Marder E, Goaillard JM (2006) Variability, compensation and homeostasis in neuronand network function. Nat Rev Neurosci 7:563–574.

32. Hutsler JJ, Lee DG, Porter KK (2005) Comparative analysis of cortical layering andsupragranular layer enlargement in rodent carnivore and primate species. Brain Res1052:71–81.

33. Nicolelis MA, et al. (2003) Chronic, multisite, multielectrode recordings in macaquemonkeys. Proc Natl Acad Sci USA 100:11041–11046.

34. Gervasoni D, Lin SC, Ribeiro S, Soares ES, Pantoja J, Nicolelis MA (2004) Global forebraindynamics predict rat behavioral states and their transitions. J Neurosci 24:11137–11147.

15926 � www.pnas.org�cgi�doi�10.1073�pnas.0904089106 Petermann et al.