18
Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom 1 , Matthew H Higgs 1,2 , William J Spain 1,2 & Adrienne L Fairhall 1 Neural systems adapt to changes in stimulus statistics. However, it is not known how stimuli with complex temporal dynamics drive the dynamics of adaptation and the resulting firing rate. For single neurons, it has often been assumed that adaptation has a single time scale. We found that single rat neocortical pyramidal neurons adapt with a time scale that depends on the time scale of changes in stimulus statistics. This multiple time scale adaptation is consistent with fractional order differentiation, such that the neuron’s firing rate is a fractional derivative of slowly varying stimulus parameters. Biophysically, even though neuronal fractional differentiation effectively yields adaptation with many time scales, we found that its implementation required only a few properly balanced known adaptive mechanisms. Fractional differentiation provides single neurons with a fundamental and general computation that can contribute to efficient information processing, stimulus anticipation and frequency-independent phase shifts of oscillatory neuronal firing. In response to the sudden change of a stimulus, the firing rate of many neurons undergoes a quick, large change followed by slower adaptation to steady state. First described over 75 years ago 1 , this adaptation causes the spike rate to accentuate changes in the input rather than constant values, and in this sense the neuronal response resembles a derivative of the input. In general, neural systems adapt to a change in stimulus statistics, which may serve to maintain maximal information transmis- sion 2–4 . Recently many neural systems have been shown to respond adaptively to changes not just in mean input, but also in input variance or contrast 3–12 , higher-order moments 13 and correlation structure 14 . In several systems, it has also been shown that the time scale of firing-rate adaptation is not static, but depends on the stimulus, for example, the duration of stimulus steps 4,15 (B. Wark, A.L.F. and F. Rieke, unpublished data). The extent to which such adaptation of neural systems results from a property of the neural circuit or whether it is a property of single neurons is often unclear. Neocortical neurons have adaptive processes that modify mean firing rate over time scales ranging from tens of milliseconds to tens of seconds 16–20 and have been implicated in single neuron adaptation to stimulus statistics 7,11 . But for single neurons, the time course of firing-rate adaptation is often taken to be fixed, or at least independent of the history of change in stimulus statistics. The mean and variance of fluctuating inputs to cortical neurons can be modulated by changes in input synchrony 21,22 , changes in network excitability resulting from changes in brain state 23,24 or by stimulus properties. Here, we found that the firing rate in single rat neocortical pyramidal neurons adapted to such changes with multiple time scales, as revealed by corresponding time scales of change in stimulus statistics. We found that these multiple time scales are a signature of a more general property; the neuron’s firing rate encodes slowly varying stimulus statistics through fractional order differentiation. This gentle form of differentiation emphasizes change in the stimulus while preserving low-frequency information and produces a power law response to sudden stimulus transitions. The fractional differentiation model captures the firing rate response to a stimulus with complex, time-varying statistics. This operation can be biophysically implemen- ted through a balance of only a few known adaptive mechanisms 17–20 . Fractional differentiation provides single neurons with a form of adaptation in which no single time scale is preferred; these dynamics may tune the effective time scale of adaptation to the time scales of the stimulus. This operation expands our understanding of the basic processing capabilities of single neurons to the encoding of changes in stimulus statistics over behaviorally relevant time scales in the range of seconds to tens of seconds. RESULTS We made recordings from 55 neocortical pyramidal neurons using sharp (n ¼ 49) and whole-cell patch (n ¼ 6) electrodes (see also Supplementary Data and Discussion online). All were layer 2-3 (L2-3) neurons of the rat sensorimotor cortex, except for three of the patch- recorded neurons, which were from layer 5; the results were not distinguishable from those from L2-3. We probed the dynamics of the firing rate with a range of stimulus dynamics. Multiple time scales of adaptation in single neurons The responses of L2-3 pyramidal neurons in acute neocortical slices were recorded using sharp intracellular electrodes. The neurons were driven with square wave current stimuli, which alternated periodically, © 2008 Nature Publishing Group http://www.nature.com/natureneuroscience Received 16 July; accepted 16 September; published online 19 October 2008; doi:10.1038/nn.2212 1 Department of Physiology and Biophysics, Box 357290, University of Washington, Seattle, Washington 98195, USA. 2 Veterans Affairs Puget Sound Health Care System, Department of Neurology, 1660 South Columbian Way, Seattle, Washington 98108, USA. Correspondence should be addressed to A.L.F. ([email protected]). NATURE NEUROSCIENCE ADVANCE ONLINE PUBLICATION 1 ARTICLES

Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

Fractional differentiation by neocorticalpyramidal neuronsBrian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2 & Adrienne L Fairhall1

Neural systems adapt to changes in stimulus statistics. However, it is not known how stimuli with complex temporal dynamicsdrive the dynamics of adaptation and the resulting firing rate. For single neurons, it has often been assumed that adaptation has asingle time scale. We found that single rat neocortical pyramidal neurons adapt with a time scale that depends on the time scaleof changes in stimulus statistics. This multiple time scale adaptation is consistent with fractional order differentiation, such thatthe neuron’s firing rate is a fractional derivative of slowly varying stimulus parameters. Biophysically, even though neuronalfractional differentiation effectively yields adaptation with many time scales, we found that its implementation required only afew properly balanced known adaptive mechanisms. Fractional differentiation provides single neurons with a fundamental andgeneral computation that can contribute to efficient information processing, stimulus anticipation and frequency-independentphase shifts of oscillatory neuronal firing.

In response to the sudden change of a stimulus, the firing rate of manyneurons undergoes a quick, large change followed by slower adaptationto steady state. First described over 75 years ago1, this adaptation causesthe spike rate to accentuate changes in the input rather than constantvalues, and in this sense the neuronal response resembles a derivative ofthe input. In general, neural systems adapt to a change in stimulusstatistics, which may serve to maintain maximal information transmis-sion2–4. Recently many neural systems have been shown to respondadaptively to changes not just in mean input, but also in input varianceor contrast3–12, higher-order moments13 and correlation structure14. Inseveral systems, it has also been shown that the time scale of firing-rateadaptation is not static, but depends on the stimulus, for example,the duration of stimulus steps4,15 (B. Wark, A.L.F. and F. Rieke,unpublished data).The extent to which such adaptation of neural systems results from a

property of the neural circuit or whether it is a property of singleneurons is often unclear. Neocortical neurons have adaptive processesthat modify mean firing rate over time scales ranging from tens ofmilliseconds to tens of seconds16–20 and have been implicated in singleneuron adaptation to stimulus statistics7,11. But for single neurons, thetime course of firing-rate adaptation is often taken to be fixed, or atleast independent of the history of change in stimulus statistics.The mean and variance of fluctuating inputs to cortical neurons can

be modulated by changes in input synchrony21,22, changes in networkexcitability resulting from changes in brain state23,24 or by stimulusproperties. Here, we found that the firing rate in single rat neocorticalpyramidal neurons adapted to such changes with multiple time scales,as revealed by corresponding time scales of change in stimulus statistics.We found that these multiple time scales are a signature of a more

general property; the neuron’s firing rate encodes slowly varyingstimulus statistics through fractional order differentiation. This gentleform of differentiation emphasizes change in the stimulus whilepreserving low-frequency information and produces a power lawresponse to sudden stimulus transitions. The fractional differentiationmodel captures the firing rate response to a stimulus with complex,time-varying statistics. This operation can be biophysically implemen-ted through a balance of only a few known adaptive mechanisms17–20.Fractional differentiation provides single neurons with a form of

adaptation in which no single time scale is preferred; these dynamicsmay tune the effective time scale of adaptation to the time scales of thestimulus. This operation expands our understanding of the basicprocessing capabilities of single neurons to the encoding of changesin stimulus statistics over behaviorally relevant time scales in the rangeof seconds to tens of seconds.

RESULTSWe made recordings from 55 neocortical pyramidal neurons usingsharp (n ¼ 49) and whole-cell patch (n ¼ 6) electrodes (see alsoSupplementary Data and Discussion online). All were layer 2-3 (L2-3)neurons of the rat sensorimotor cortex, except for three of the patch-recorded neurons, which were from layer 5; the results were notdistinguishable from those from L2-3. We probed the dynamics ofthe firing rate with a range of stimulus dynamics.

Multiple time scales of adaptation in single neuronsThe responses of L2-3 pyramidal neurons in acute neocortical sliceswere recorded using sharp intracellular electrodes. The neurons weredriven with square wave current stimuli, which alternated periodically,

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

Received 16 July; accepted 16 September; published online 19 October 2008; doi:10.1038/nn.2212

1Department of Physiology and Biophysics, Box 357290, University of Washington, Seattle, Washington 98195, USA. 2Veterans Affairs Puget Sound Health Care System,Department of Neurology, 1660 South Columbian Way, Seattle, Washington 98108, USA. Correspondence should be addressed to A.L.F. ([email protected]).

NATURE NEUROSCIENCE ADVANCE ONLINE PUBLICATION 1

ART ICLES

Page 2: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

with cycle period T, between two values (Fig. 1a), and with filteredGaussian noise whose s.d. alternated periodically between two values asthemean remained constant (Fig. 1b). Generally, the binned spike ratesadapted upward after a step decrease and downward after a stepincrease (Fig. 1c,d; see Methods). Although the time course ofadaptation was not necessarily exponential, we fitted the averagedfiring rates for each half cycle by a single exponential to determine aneffective adaptation time scale for both the low and high mean orvariance conditions. As T increased, the effective adaptation timeconstant t increased proportionally for both the noiseless step stimuli(Fig. 1e) and the noisy stimuli with a step change in variance (Fig. 1f),thus showing history-dependent adaptation to be a property of a singleneocortical neuron.The close-to-linear stimulus dependence of the adaptation time scale

suggests that the observed rate adaptation was generated by a scale-invariant (no preferred time scale), multiple time scale process ratherthan by a single exponential mechanism. This is consistent with power-law dynamics, as has been observed in some primary sensoryreceptors25,26. Power-law response to a step is one example of aninput-output relationship for a system implementing fractional orderdifferentiation. We tested whether the adaptive mechanisms of singleneurons allow them to function as fractional differentiators of thesomatic input current.Fractional differentiation27 (see also Supplementary Equations

online) is a linear operation that can be written in the time t orfrequency f domain as

r tð Þ ¼ kh tð Þ $ x tð Þ + r0 ,FR fð Þ ¼ kH fð ÞX fð Þ + r0d fð Þ ð1Þ

where r(t) is the firing rate response to a time-varying input x(t), h(t) isthe fractional differentiating filter in the time domain and H(f) is theFourier-transformed filter in the frequency domain (Fig. 2). Theconstants k and r0 account for the filter gain and the overall mean

firing rate; these constants can be fixed by the mean and s.d. of theobserved firing rate. In the frequency domain, the filterH(f) is (i2pf)a,where a is the order of fractional differentiation. If a neuron functionsas a fractional differentiator, the gain of the frequency response will beproportional to (2pf)a and each frequency component of the rateresponse will have a frequency-independent phase lead of ap/2 withrespect to the stimulus (see Methods and Supplementary Equations).For comparison, a single exponential filter leads to a frequency-dependent gain of 1

ð1+4p2f 2t2Þ and phase of arctan(%2pf t). If theseconditions hold, the order a of fractional differentiation can bedetermined from the gain-frequency relationship and phase leads.For example, when x(t) ¼ sin(2pft), then the fractional derivativewith order a is dax

dta ¼ ð2pf Þa sinð2pft + ap2 Þ. Finally, fractional differ-

entiation differs from integer order differentiation in that it is non-local. Although integer order differentiation depends instantaneouslyon the input function, the result of fractional differentiation dependson the history of the stimulus, as seen in the slow (power-law) decay ofh(t) with time (as in Fig. 2a).

Single neurons as fractional differentiatorsWe tested this functional model of the neuronal response throughlinear analysis. We recorded firing-rate responses of L2-3 neurons toinjected sine wave (noiseless) currents and sine-modulated noise(constant mean) (Fig. 3a,b). Although previous work has focused on

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

0 10 20 30

2

4

Period T (s)

Tim

e co

nsta

nt !

(s)

Firi

ng r

ate

(Hz)

High to lowLow to high

0 10 20 30

2

4

Period T (s)

0 8 16

10

20

Firi

ng r

ate

(Hz)

Time (s)

0 8 16

10

Time (s)

2 s

50 mV

0.1 nA

a b

e f

c

d

" = 0

" = 1

" = 0.25

" = 0.50

" = 0.75

a b

30 700

2

Frequency (Hz)

Mag

nitu

de

30 700

10

20

Frequency (Hz)

Pha

se (

deg)

0 0.2 0.4

0

2

Time (s)

h(t)

3 70

2

Period T (s)

Mag

nitu

de

Figure 2 The effect of the fractional differentiating filter H(f) ¼ (i2pf)a. (a) Inthe frequency domain, the absolute amplitude response is a power law andthe phase response is constant (top). The inset shows the magnitude as afunction of period rather than frequency. The filter h(t) is also shown in thetime domain (bottom, see Supplementary Equations). For these examples,a ¼ 0.15 with a sampling rate of 200 Hz. (b) When a ¼ 0, the responsewaveform is equivalent to that of the input, whereas when a ¼ 1, a fullderivative of the input is taken. Values of a between 0 and 1 are fractionalderivatives and contain similarities to both the waveforms resulting whena ¼ 0 and a ¼ 1.

Figure 1 The time constant of rate adaptationincreases with step duration or period T. (a,b) Inresponse to many repeats of square-wave current(a, top) or square-wave noise (b, top) with a periodT, injected into the soma, the neuron fires actionpotentials (bottom). (c,d) The neuron’s mean firingrate adapted both upward and downward (solidlines, 30 bins per period; see also SupplementaryFigs. 4 and 5 online) in response to square-wavecurrent (c) or square-wave noise (d). The upwardand downward adaptation time courses were eachleast-squares fit by an exponential of the formA expð%t

t Þ (dashed lines). The upward anddownward responses were separately fitted with atime constant t for each stimulus period T. (e,f) AsT increased, t similarly increased for both thenoiseless current (n ¼ 8, e) and constant mean,noisy (n ¼ 11, f) stimuli. Data in a and c are fromone example neuron, whereas data in b and d arefrom another.

2 ADVANCE ONLINE PUBLICATION NATURE NEUROSCIENCE

ART ICLES

Page 3: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

sudden stimulus changes and response characterization for high inputfrequencies (B1–1,000 Hz)28,29, we focused on low-frequencyresponses (B0.03–1 Hz), for which stimulus periods are large com-pared with typical interspike intervals.The average gain over all neurons in response to sine wave current

and sine wave–modulated noise decreased with increasingT in a powerlaw–like manner (Fig. 3c). The exponent a can be determined from theslope of the gain-period relationship on a log-log curve. In response toboth sine wave current and sine wave noise, the phase leads were closeto frequency independent (Fig. 3d) and the value of a extracted fromthe phase shift was consistent with that estimated from the gain. Theincreased phase leads in response to sine wave currents for T ¼ 1–2 smay reflect a voltage-dependent activation of a conductance with ashort time scale, as the mean membrane potential was relativelydepolarized relative to that observed during the sine wave noise.Our data indicate that mean firing-rate responses to stimuli that

change slowly relative to the interspike intervals are consistent with amodel of fractional differentiation. In these experiments, we found thatslowly varying, noiseless currents and modulated noise with constantmean are transformed into firing rate according to the fractionaldifferentiation model. In subsequent experiments, we focused on theresponses to sinusoidally modulated noise currents with constant mean.Because fractional differentiation, equation (1), is linear we tested

whether the neuronal response to single sine waves predicts the

response to signals constructed as a sum of sine waves. Neurons weregiven noisy stimuli modulated by an envelope consisting of a sum ofsine waves (Fig. 4a). These responses were compared with thoseobserved when each period of sine wave noise was presented indivi-dually (Fig. 4b). Neither the gain curves (n = 12, two-way ANOVA,F1,88 = 0.02, P = 0.902) nor the phase leads (F1,88 = 1.37, P = 0.245)were significantly different, suggesting that the firing rate of theseneurons depends linearly on the input in the tested ranges ofstimulus parameters.A further test of linearity is given by the responses to square waves

that we have already shown. Although these curves were fit with singleexponentials to extract a representative time scale for each curve(Fig. 1), the existence of multiple time scales of adaptation impliesthat this is not an optimal procedure; both faster and slower time scaleswere apparent during some individual stimulus steps. Therefore, spikeresponses to square waves were fitted with the multiple time-scaleresponse predicted by fractional differentiation, equation (1).Fractionally differentiating a step function yields a symmetric power-

law increase and power-law decrease that is governed by the parametera (see Methods and Supplementary Equations). We found theresponse of a neuron to square wave–current stimuli for three different

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

0.7

0.8

Cur

rent

(nA

)

0 8 16

7

15

Time (s)

Current inputFiring rate

0.2

0.4

0 2 4

0

15

Firi

ng r

ate

(Hz)

Time (s)

Noise envelopeFiring rate

0 16 32

20

40

60

Gai

n (H

z nA

–1)

Period T (s)

Sine-wave currentsSine-wave noise

100 101

101.4

101.8

" = 0.16

" = 0.15

Period T (s)

0 10 20 30

10

20

" = 0.14

Pha

se le

ad #

(de

g)

Period T (s)

a b

c

d

0 2 4

0

1

Time (s)

Cur

rent

(nA

)

a

b

c

0 16 32

5

15

Time (s)

Firi

ng r

ate

(Hz)

Noise envelopeFiring rate

0 0.1 0.2 0.3

1

2

Frequency (Hz)

Am

plitu

de (

Hz)

0 10 20 30

20

40

Gai

n (H

z nA

–1

)

Period T (s)

Sum of sines noiseSine-wave noise

101

101.4

101.5

" = 0.19

" = 0.17

Period T (s)

0 10 20 300

10

20

" = 0.13

Pha

se le

ad #

(de

g)

Period T (s)

Figure 4 Neuronal response to sine-wave noise envelopes is similar whenperiods T ¼ 4, 8, 16 and 32 s are presented simultaneously (with phasesf ¼ 0, 1, 2 and 3 rad) or individually. (a) The stimulus noise envelope is asum of four sine waves; the amplitude of the neuronal response can beobtained from the Fourier transform of spike responses. Data are from anexample neuron. (b,c) Gain curves were similar (n ¼ 12, two-way ANOVA,F1,88 ¼ 0.02, P ¼ 0.902, b), as were phase leads (F1,88 ¼ 1.37, P ¼ 0.245,c). For individually presented sine-wave noise, the mean phase lead was13.11 (a ¼ 0.14) and for the sum of sine-wave noise, the mean was 11.91(a ¼ 0.13). Error bars represent standard error.

Figure 3 The gain of the neuronal firing response can be described by apower law and the phase lead is frequency independent. (a,b) Sine–wavecurrent (a) or sinusoidally modulated noise (b) was injected and elicited asinusoidal response (60 bins per period) with a phase lead with respect to thestimulus (see also Supplementary Fig. 6 online). Data are from two exampleneurons. (c) The average gain of the response decreased with period T (n ¼ 8and 11 for sine-wave currents and sine-modulated noise, respectively),similar to a power law with exponent a, as seen from the log-log gain curves;the slope of the best-fit line is –a. (d) In response to sine-wave current, thephase lead was frequency independent for T ¼ 4–32 s (n ¼ 8, one-wayANOVA, F3,28 ¼ 1.63, P ¼ 0.204) with mean 11.91 (a ¼ 0.13) and wasfrequency independent in response to sinusoidally modulated noise forT ¼ 1–32 s (n ¼ 11, one-way ANOVA, F5,60 ¼ 0.51, P ¼ 0.766) withmean 12.31 (a ¼ 0.14). Error bars represent standard error.

NATURE NEUROSCIENCE ADVANCE ONLINE PUBLICATION 3

ART ICLES

Page 4: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

periods (Fig. 5a), which we overlaid with the predicted responsederived from the linear filter with amplitude and exponent a derivedfrom the sinusoidal experiments (as inFig. 3c).We also determined theresponse to a more complex stimulus in which the s.d. changedperiodically between three different values (Fig. 5b). This induced

upward adaptation when the stimulus stepped from the high to themedium variance and downward adaptation when the step was fromthe low to the medium variance. Given appropriate scale factors, thesedynamics were well captured by the fractional differentiator modelwithout fitting (see Methods).Thus, the dynamics of neuronal responses to complex stimuli were

well predicted by the differentiating filter. We then least-squares fit thesquare wave responses (Fig. 1c,d) with the response predicted byequation (1) using the parameter a. In this case, the best a was foundfor each neuron in response to square wave current or square noise. Forsquare wave current and noise, a was 0.163 (s.d. ¼ 0.034) and 0.137(s.d. ¼ 0.048), respectively (Fig. 5c). The difference in error betweenfitting each curve using an exponential form and fitting all with a singlefractional-differentiating filter was small. For square-wave current, thedifference in the mean of the absolute error was 0.18 Hz per bin,whereas the difference was 0.06 Hz per bin for square-wave noise,where errors were slightly smaller with exponential fitting. Differencesin squared errors were significant for square-wave noise (two-wayANOVA, F1, 3,938 ¼ 7.94, P ¼ 0.005), but not for square-wave current(two-way ANOVA, F1, 2,864 ¼ 2.92, P ¼ 0.087). Thus, scale-invariantfractional differentiation provides an accurate description of thedata, with fewer parameters, compared with fitting separate exponen-tial time scales.We then compared the value of a as estimated in eight ways (Fig. 5c).

The overall mean was 0.15 (s.d. ¼ 0.06). These data show that theneuronal firing-rate response to time-varying stimulus statistics iscompactly approximated as fractional differentiation of order 0.15.

Implementation of fractional differentiationIt remains unclear how power-law dynamics are implementedbiologically, although previous theoretical work suggests several possi-bilities25,26,30–32. For single neurons, several mechanisms are known tounderlie rate adaptation, including slow sodium-channel inactivation18

and after-hyperpolarization (AHP) currents16,17,19,20, which decay witha mixture of time scales33. Because slow AHP (sAHP) currents aredistinct from the spike-generating mechanism, in contrast with slowsodium inactivation, it is relatively straightforward to manipulate thesecurrents experimentally. We experimentally manipulated sAHP cur-rents, as we expected that these currents contribute to the adaptationtime scales as measured by the phase lead; increasing sAHP currentsshould increase phase leads, whereas decreasing sAHP currents shoulddecrease phase leads.

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

0 8 16

10

20

Time (s)

Firi

ng r

ate

(Hz)

0 8 16 24

5

15

Time (s)

Firing rate" = 0.05" = 0.15" = 0.25

0

0.1

0.2

Fra

ctio

nal o

rder

"

Boxes(filter)n = 8

Sines(phase)n = 8

Sines(gain)n = 8

Boxnoise(filter)n = 11

Sinenoise

(phase)n = 23

Sinenoise(gain)n = 23

Sum ofsinenoise

(phase)n = 12

Sum ofsine

noise(gain)n = 12

NoiseCurrent

" = 0.15

a

c

b

Figure 5 Responses can be predicted using fractional differentiation with theparameter a, which was estimated by several methods to beB0.15. (a,b) Theresponses of two neurons (solid line) were consistent with the outputpredicted by a fractional differentiator (dashed line), equation (1), with a asderived from other experiments (as in Fig. 3c,d). In a, the gain amplitude anda for the predicted response from the fractional differentiator was determinedusing individual sine wave data from the same neuron (as in Fig. 3c); themean of the predicted response was matched to the firing-rate response. In b,the predicted response was normalized to the mean and s.d. of the firing-rateresponse. Average firing rates are shown in response to square-wave currentwith T ¼ 4, 8 and 16 s (a) and square-wave noise with T ¼ 24 s and threelevels of s.d. ¼ 3, 2, 1 and 2 presented for 8, 4, 8 and 4 s (b), respectively.We also show the predicted response obtained for a ¼ 0.05 and 0.25. (c) Bargraph shows the mean (and s.e.) for a calculated in eight ways in response tosquare- and sine-wave current (three datasets) as well as to square- and sine-wave noise (five datasets), where the method of finding a is shown inparentheses. n is the number of a¢s that contribute to the mean and isequivalent to the number of contributing neurons. The overall mean wasa ¼ 0.15 (s.d. ¼ 0.06), where none of the obtained values were significantlydifferent (one-sample t test, P ¼ 0.05). Sine-wave current data from T ¼ 1and 2 were excluded, as these phases were frequency dependent (Fig. 3d).

a

b0 10 20 30

10

20

30

**

**

Pha

se le

ad #

(de

g) With artifical AHP ! = 3 s

Control

0 10 20 30

10

20

* * *

Pha

se le

ad #

(de

g)

Period T (s)

ControlWith α-methyl-5-HT

Figure 6 Fractional differentiation depends on a balance of multipleadaptation mechanisms with different time scales. Artificially increasing orpharmacologically blocking sAHP currents led to opposite changes in phaseleads. (a) sAHP currents were added to patch-clamped neurons (n ¼ 6) usingdynamic clamp (1-ms pulse, t ¼ 3 s, DG ¼ 0.05 nS per spike, –100 mVreversal potential). Without the artificial sAHP conductance, phase leads werefrequency independent (one-way ANOVA, F5,30 ¼ 1.60, P ¼ 0.190). With theartificial sAHP conductance, phase leads were strongly frequency dependent(F5,30 ¼ 7.48, P ¼ 0.000). The two conditions were significantly different(two-way ANOVA, F1,60 ¼ 27.9, P ¼ 0.000) and phase leads increased forT ¼ 4–16 s. (b) Application of a-methyl-5-HT reduced sAHP currents witht E 1 s by 63% (s.d. ¼ 13%), as measured 450–550 ms after a train of30 spikes at 50 Hz from a pre-train membrane potential set to –65 mV byinjection of positive holding current. Drug application (n ¼ 10) altered phaseleads (two-way ANOVA, F1,108 ¼ 9.39, P ¼ 0.003) by reducing them forT ¼ 4–16 s. Error bars indicate standard error and asterisks indicatesignificant differences (paired t tests, P o 0.05).

4 ADVANCE ONLINE PUBLICATION NATURE NEUROSCIENCE

ART ICLES

Page 5: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

First, we artificially added a sAHP conductance with a decay t ¼ 3 sto patch-clamped neurons (L2-3 and L5) using spike-triggereddynamic current clamp (Fig. 6a and Supplementary Fig. 1 online).Without the artificial conductance, phase leads were frequency inde-pendent, but with the artificial conductance, phase leads increased forT ¼ 4–16 s with a frequency dependence consistent with a previousstudy34. Thus, the addition of a sAHP conductance disturbed thebalance of underlying adaptive mechanisms.Next, again using sharp electrodes (L2-3 neurons), we pharmacolo-

gically partially blocked the early sAHP current (t B1 s) using the 5-HT2 receptor agonist a-methyl-5-HT (refs. 6,7), reducing it by 63%(s.d.¼ 13%) (see Fig. 6 and Supplementary Figs. 2 and 3 online). Thismanipulation decreased phase leads forT¼ 4–16 s (Fig. 6b). However,the magnitude of the pharmacologically induced reduction of thesAHP was only weakly correlated with the magnitude of reduction inphase leads (Pearson correlation coefficients were 0.51, 0.21 and 0.33with P values of 0.13, 0.55 and 0.36 for T values of 4, 8 and 16 s,respectively). On the basis of this and previous data showing thatother mechanisms also contribute to spike-frequency adaptation18,19,we do not believe that sAHP currents are solely responsible for theobserved phase leads. Rather, it is probable that a combination ofmechanisms gives rise to fractional differentiation in these neurons.Here, we have shown that manipulating the sAHP current affects phaseleads, as expected if this current at least partially underlies multipletime-scale adaptation.We examined whether a few AHP currents with different time

constants are sufficient to generate the observed multiple time-scaleadaptation. To examine this, we first added 2–3 AHP currents to astandard, single-compartment Hodgkin-Huxley model neuronwith noslow adaptive processes and then added two time scales of slow sodiuminactivation (Fig. 6a; see Supplementary Methods online). AHP

currents were added with time constants that roughly spanned therange of observed sAHP currents in neocortical pyramidal neurons20.The amplitudes of these currents were precisely balanced with respectto one another to approximate fractional differentiation. For slowsodium inactivation, we augmented sodium current dynamics withslow inactivation gates having two time scales and with kinetics asmeasured in neocortical neurons18.These models showed multiple time scales of adaptation (Fig. 7a)

and approximately frequency-independent phase leads (Fig. 7b), aswas observed in the real neurons (Figs. 1 and 3). The observation thatonly a few adaptive processes can underlie rate adaptation over a widerange of time scales is consistent with theoretical analysis showing thata single adaptive process affects the neuronal response over a widerange of frequencies34. Together, these results demonstrate that frac-tional differentiation can be implemented biophysically with knownadaptive mechanisms.

DISCUSSIONAlthough single neurons respond to current fluctuations on very shorttime scales35, the mean firing rate can be considered to encode slowlyvarying statistical properties of the stimulus4 or the stimulusenvelope36, acting as an additional channel of information37. Wefound that the dynamics underlying spike-rate adaptation to stimulussteps in single cortical neurons functionally approximate fractionaldifferentiation. This provides a general model for the firing-rateresponse to time-varying stimulus statistics or envelope encoding.Far from simply accommodating to a new state, these adaptivedynamics convey detailed information about stimulus components.Furthermore, this rate response is a linear function of the stimulusmean or envelope. Although extracting a stimulus envelope requires anonlinear operation36, here the dynamics of spike generation andadaptation provide this operation, while retaining the neuron’s abilityto represent high-frequency fluctuations. We considered responses tovariations in mean and variance separately. Further work is required toevaluate the encoding of simultaneous and potentially correlatedchange in both statistics.We have focused on the slower time scales that are relevant for time-

varying firing rates. One could combine this approachwith methods ofcapturing fast time-scale dynamics (for examples, see refs. 38,39) togive a model that captures both specific stimulus components causingsingle spikes and slower-varying components relevant for the firingrate. The simplest possibility is that the rate dynamics r(t) act as amultiplicative gain on the fast dynamics of spike generation. Consider,for example, the case of a slowly time-varying s.d. s(t). For a given s,the occurrence of single spikes is determined by a gain function gs,which acts on the input after convolution with a filter fs. Decomposingthe input into an envelope and a fast-varying component as s(t)Z(t),this multiplicative model for the spike probability is

P spike s tð ÞZ tð Þjð Þ ¼ r s tð Þð Þgs fs $ Z tð Þð Þ;

where the rate r is determined by the envelope s(t), as we have exploredhere. Such a description may hold true when there is a separation oftime scales between the short time-scale action of fs and gs and theslowly time-varying stimulus parameter s, and allows for changes inthe instantaneous coding of inputs resulting from statistical propertiesof the stimulus, such as gain normalization4.Previous work has shown that slow adaptation currents in neocor-

tical neurons have multiple time scales16,17,18,20. One study19 found thatin vitro data from rat L5 pyramidal neurons could be modeled by aleaky integrate-and-fire (LIF) model neuron modified to include two

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

0 50 100 150 200

0

10

20

Period T (s)

Pha

se le

ad #

(de

g)

AHP with ! = 0.3, 1 and 6 sAHP with ! = 0.3 and 6 sSlow Na inact. with ! = ~0.3 and 6 s

Without slow adaptation

0 20

2

4

Period T (s)

Tim

e co

nsta

nt !

(s)

AHP with ! = 0.3 and 6 s

High to lowLow to high

0 20

2

4

Period T (s)

Slow Na inact. with ! = ~0.3 and 6 sa

b

0 100 200

2

4

T (s)

! (s

)

0 100 200

2

4

T (s)

! (s

)

Figure 7 Single-compartment Hodgkin-Huxley model neurons with twoadaptation time scales can approximate a fractional differentiator over a30-fold range of stimulus period T. (a) A standard Hodgkin-Huxley neuronwith two sAHP currents or two time scales of slow sodium inactivationshowed multiple time scales of adaptation (similar to Fig. 1c,d); as periodTwas further increased, the time scales became constant (insets). (b) Addinga single, exponential adaptive mechanism increased phase leads over a widerange of time scales, and two such mechanisms yielded a roughly constantphase lead over an approximately 50-fold change in Twhen the amplitudes ofthe adaptation mechanisms were properly balanced. Additional currents wereable to lead to responses that varied less over T.

NATURE NEUROSCIENCE ADVANCE ONLINE PUBLICATION 5

ART ICLES

Page 6: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

spike-dependent adaptation currents and one facilitation current. Theaverage fitted time constants for these currents were 48 ms, 5.8 s and580 ms, respectively. As a result of the facilitating current, this modifiedLIF model19 does not generate results consistent with fractionaldifferentiation. However, a modified LIF model with an early sAHPcurrent (t ¼ B1 s)20 instead of a facilitating current can give resultssimilar to ours (Fig. 7). As the LIF and Hodgkin-Huxley models cangive similar results, it is likely that the fractional differentiation modeldoes not depend on the details of spike generation. Other mechanismsof adaptation, such as slow sodium inactivation, cannot be readilyimplemented in an LIF model.A fractional differentiator has a power-law response to stimulus

steps. Power law responses have been observed in a range of neuralsystems, from single ion channels15 to cognitive behavior32,40. Thepresence of power-law adaptation to a step and fractional differentia-tion are not synonymous; it is possible to have a power law responsewithout the frequency-independent phase property of the fractionaldifferentiator. Fractional order dynamics have been observed in thevestibular-ocular system30,41,42 and in the fly motion sensitive neuronH1 (ref. 43). However, these results were from neural systems ratherthan from single neurons. A range of mechanisms may contribute tofractional order dynamics and power laws, including circuit30 andsynaptic40 mechanisms, geometrical properties of cells25 and den-drites42, and the multiple inactivation states of sodium channels31.Our results show that this computation can be carried out by singlecortical neurons and can be implemented by known adaptationmechanisms in a spatially restricted region of the neuron.Unlike a full derivative, fractional differentiation enhances responses

to stimulus change, but does not entirely remove information aboutvery low-frequency stimulus fluctuations. Retaining information aboutboth the signal magnitude and the signal rate-of-change is a property offractional order derivatives that leads to dependence on stimulushistory. Unlike integer order derivatives, fractional orders are non-local. We found the fractional order (a) to be B0.15. This gentledifferentiation may reflect a requirement for the neural system toencode only a modest amount of rate-of-change information by asingle neuron. Increasing the effective value of a beyond 0.15 could beaccomplished by a neural circuit through the sequential adaptationeffects of multiple neurons and intervening synaptic dynamics30,44.We found that the fractional differentiation model for spike rate

changes holds true over time scales from B1–30 s, a range that isbehaviorally relevant for fluctuations in input statistics. It has beensuggested that fractional order differentiation may be an importantproperty of motor control systems that compensates for fractionalorder integration dynamics of muscle and tissue, resulting from theirviscoelastic properties42. Here, we found that single neurons in thesensory-motor cortex have this property. This kind of rate encodingmay also allow neurons to process information efficiently by matchingresponse time scales to input time scales45 or to temporally decorrelate,or whiten, the stimulus envelope46. The power-law form of the gainmay be related to the typical power-law spectra of natural stimuli47,48,which would facilitate stimulus whitening. The constant phase shiftallows phase to be modulated independently of frequency, which maybe important for the interaction of neocortical neurons with slowoscillatory frequencies in the brain49. Finally, representing an input’sderivative could allow neural circuits to predict future stimuli, similarto a Taylor approximation44.Our results suggest that fractional differentiation is a fundamental

and elementary computation of L2-3 neocortical pyramidal neuronsand provides a general model of the response of adapting neocorticalneurons to time-varying stimuli. These intrinsic dynamics provide a

substrate for a form of short-term memory, retaining stimulus infor-mation over the intermediate time scales of seconds to tens of seconds.Although we focused on L2-3 neurons distributed throughout thesensorimotor cortex, we found similar results among a small sample ofL5 neurons. Fractional differentiation may be a general property ofadapting neurons in the neocortex. A subset of L5 neurons showminimal slow firing-rate adaptation7 and these cells may not showfractional differentiation. Although it may seem surprising that thesedynamics have not been discovered previously, it is probable that theexperiments that would have revealed these dynamics were not carriedout at the slow time scales investigated here. The appearance of thesedynamics in single neurons and our ability to disrupt them suggest thatadaptation currents and slow sodium inactivation, among otheradaptive mechanisms, may be finely tuned to contribute these multipletime scales. This tuning may achieve a balance between providing rate-of-change information and preserving a continuous response in theface of negative feedback.

METHODSPreparation of cortical slices. We deeply anesthetized 5–12-week-old SpragueDawley rats in a chamber filled with 4% isoflurane in oxygen and quicklydecapitated them. The rostral, caudal and ventral portions of the brain wereremoved and the remaining block was attached to the stage of a Vibratometissue slicer (TPI) using cyanoacrylate glue (Loctite 404, Loctite) and immersedin ice-cold cutting solution containing 220 mM sucrose, 3 mM KCl, 1 mMCaCl2, 5 mM MgCl2, 26 mM NaHCO3, 1.25 mM NaH2PO4 and 10 mMD-glucose. The cutting solution was bubbled with 95% O2/5% CO2 to maintainpH at 7.4. Coronal slices (300 mm thick) were cut (B+1 mm to –1 mm relativeto bregma) and transferred to a holding chamber filled with artificial cere-brospinal fluid containing 125 mM NaCl, 3 mM KCl, 2 mM CaCl2, 2 mMMgCl2, 26 mM NaHCO3, 1.25 mM NaH2PO4 and 20 mM D-glucose, bubbledwith 95% O2/5% CO2. The temperature of the holding chamber was main-tained at 34 1C forB60 min and then allowed to cool to approximately 22 1C.

Recording. Slices were transferred to a recording chamber and perfused atB2 ml min–1 with warmed artificial cerebrospinal fluid (B33 ± 1 1C). Forpharmacological experiments, the control recording solution also contained6,7-dinitroquinoxaline-2,3(1 H,4 H)-dione (20 mM), ±3-(2-carboxypiperazin-4-yl)-propyl-1-phosphonic acid (5 mM) and picrotoxin (100 mM) to blockAMPA/kainate, NMDA and GABAA receptors, respectively. Recordings wereobtained from cells in L2-3 (sharp electrodes) and L2-3 and L5 (whole-cellpatch electrodes) in the dorsal neocortexB1-4 mm from the midline using anAxoclamp 2-A amplifier (Molecular Devices) in continuous bridge mode.Sharp electrodes were filled with 3 M KCl and had resistances of B60–100 MO. Patch pipettes were filled with solution containing 127 mMKCH3SO4, 10 mM myo-inositol, 2 mM MgCl2, 5 mM KCl, 10 mM HEPES,0.02 mM EGTA, 6 mM sodium phosphocreatine, 2 mM Na2ATP and 0.5 mMNa3GTP, pH 7.2–7.3. Access resistances during whole-cell recording wereB10–20 MO. Data were filtered at 10 kHz and sampled at 10–20 kHz usingITC-16 and ITC-18 data acquisition boards (Instrutech). Data acquisition wascontrolled by custom macros written in Igor Pro (WaveMetrics).

Current stimuli. Single neurons were stimulated with time-varying currentstimuli in current clamp. Square waves were generated to switch between valuesof 1 and 2 with a range of oscillation periods. Similarly, standard sine waveswere generated to vary between 1 and 2 with a range of periods. A positiveholding current was applied to depolarize each neuron to within 10% ofrheobase, near the threshold for firing. The basic periodic waveforms werescaled by a factor between 50 and 250 pA, chosen individually for each cell suchthat the neuron fired throughout the stimulus block. In trials using noisestimuli, the simple periodic waveforms were multiplied by exponentiallyfiltered (1-ms time constant) Gaussian white noise to give periodicallymodulated noise, with a s.d. given by the square or sine wave. To producecurrent stimuli, these noise sequences were again shifted using a positiveholding current to near rheobase and scaled as described above. During the

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

6 ADVANCE ONLINE PUBLICATION NATURE NEUROSCIENCE

ART ICLES

Page 7: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

noise stimuli, this scaling produced membrane-potential fluctuations with s.d.of B2–8 mV. These stimulus sets were presented in blocks with periods T ¼ 1,2, 4, 8, 16 and 32 s, where each period was presented for 96 s, giving a totalstimulus length of 576 s, following which periods were presented in reverseorder. On every individual cycle repetition, the particular noise sequencediffered. Finally, modulated noise stimuli were created where the modulationenvelope was generated as a sum of sinusoids. Four sine waves with periodsT ¼ 4, 8, 16 and 32 s and phase shifts f ¼ 0, 1, 2 and 3 rad were summed andthe result was used to multiply a Gaussian white-noise sequence as above. Inthis case, the block length was 192 s. Control blocks consisting of individualsine waves with T ¼ 4, 8, 16 and 32 s and block length 96 s were presented tothe same cells. Stimuli were scaled before injection, as described above.

Cell-acceptance criteria and firing rates. Neurons were considered to behealthy and stable if their resting membrane potential was o–70 mV, spikeswere465 mV in height, and the resting membrane potential and rheobase (thecurrent at which the neuron first began to spike) did not fluctuate by morethan 10% between stimulus blocks (except during pharmacological or dynamicclamp manipulations). Spike times were recorded when the membranepotential crossed –10 mV and reached its peak, and subsequent upwardcrossings within 2 ms were not considered. Time-varying firing rates werefound by taking a histogram of spike times modulo the cycle period.

Fitting exponentials to firing rates. Firing-rate time courses (as in Fig. 1a,b)were least-squares fit with the two-parameter exponential A expð%t

t Þ, where thesteady state was defined to be the last three bins of each half period and 30 binswere used for each period. Cells for the square-wave (n ¼ 8) and square-noise(n ¼ 11) experiments were included if firing was41 Hz throughout and firingrates did not show upward adaptation after a step increase. Upward adaptationin the high condition was observed inB10% of neurons and was usually seenin cells with unusually high firing rates, signs of poor cell health or unstablerecordings. Thus, experiments showing this type of response were excludedfrom the present analyses. Exponential t’s were not considered if the amplitudeof the exponential relaxation, A, was less than 10% of the steady-state firingrate. This criterion excluded 14 of the total 228 measurements of t.

Fitting sine waves to firing rates. The amplitude and phase lead of theresponse with respect to the stimulus were found by least-squares fitting themean firing-rate response of each neuron, determined, as done previously,by taking a histogram of spike times, with A sinð2pT +fÞ, with the twoparameters A and f. Identical results were obtained via Fourier transform ofthe firing-rate response, represented as a vector of zeros and ones designatingbins in which there was not or was a spike; the dominant frequency amplitudeand corresponding phase were easily identified. The gain for each Twas foundby dividing A by the stimulus scaling factor used for a given neuron, whichranged from 50–250 pA. For the sum of sines experiment, a vector of zeros andones to signify no spike or spike, respectively, was created with a sampling rateof 500 Hz; amplitudes and phases were found from the Fourier transform ofthis vector, with results nearly identical to those from the least-squares fitting ofthe mean firing- rate histogram, or from the Fourier transform of thehistogram. Neurons were included if at least four blocks (12.8 min) of datain response to the sum of sines stimulus were obtained, so that phases could bedetermined with sufficient accuracy. Neurons were excluded if they did not firethroughout T.

Fitting square-wave data with fractional differentiator responses. The shapesof the firing-rate curves were fit with the response that a fractional differ-entiator with order a has to a square-wave stimulus. The fractional-differentiating filter was found as follows. In the frequency domain, the filterwas defined as A(i2pf)a for frequencies –960/T to 960/T with Df ¼ 0.01/T. Inthe time domain, this filter was convolved with step functions of period Twitha sample rate of 1,920/T, using only the result computed without the zero-padded edges. Neighboring bins of this finely sampled response were averagedto match the bin number of the firing-rate responses. As we simply wished toevaluate the exponent a for the population, we fixed the amplitude A bynormalizing each response such that input and output had the same mean ands.d. For each experiment, which consisted of six periods, the single best a wasfound by least-squares fitting to the firing-rate response. Thus, fitting responses

from a single experiment required three parameters (a plus normalization ofmean and variance) instead of 36 parameters (six parameters for twoexponential fits per T), as was done previously.

Using dynamic clamp to add sAHP currents. When the membrane potentialcrossed –10 mV, a pulse generator injected a 1-ms negative current pulse into aresistor-capacitor circuit, the voltage, VRC, of which relaxed with a 3-s timeconstant. The resistor-capacitor circuit provided input to a custom-madeanalog dynamic-clamp circuit that produced an output GasAHP(Vmembrane –EK), where GasAHP ¼ kVRC ¼ 0.05 nS per spike and EK ¼ –100 mV. This wassummed with the stimulus current.

Biophysical modeling. The single-compartment, conductance-based Hodgkin-Huxley model neuron50 was used with standard parameters (SupplementaryMethods), except as noted. For each AHP current, an additional current, orterm, was added to the equation for dV/dt of the form –GAHPa(V – Ereversal),where a was incremented by one after each spike and decayed according toda/dt ¼ –a/t, with t ¼ 0.3, 1 and 6 s, and GAHP ¼ (0.05, 0.006 and 0.004)GLeak

with the time scales used as indicated. To implement slow sodium inactivationwith two time scales, two extra gating variables, s1 and s2, were added to thesodium current so that it became: GNam

3hs1s2(V – ENa). The kinetics for theseadditional gating variables were18 ds

dt ¼ kðasð1% sÞ % bssÞ with as ¼0:001 expð%85%V

30 Þ and bs ¼ 0:0034expð%17%V

10 Þ+1, where k was used to modify the time

constant ts ¼ 1ðas+bsÞ

of s. Specifically, given these kinetics, ts is voltagedependent with a peak of B2,300 ms at –50 mV, tapering off to B500 msby 0 mV. To approximate AHP currents with time constants of 0.3 s and 6 s (asin Fig. 7), k was set equal to 2/0.3 and 2/6, respectively. Equations were solvednumerically using fourth-order Runge-Kutta integration with a fixed time stepof 0.05 ms and spike times were identified as the upward crossing of the voltagetrace at –10 mV (resting potential ¼ –65 mV) separated by more than 2 ms.For these data, the mean of the exponentially filtered (t ¼ 1 ms) Gaussianwhite-noise stimulus was 5.5 mA cm–2 and the s.d. of the square and sine wavesvaried between 20 and 32 mA cm–2.

Magnitude and phase of a fractional differentiating filter. Because complexnumbers can be written in terms of amplitudes and phases, the response R(f) toa signal X(f) ¼ Cf exp(iyf) through a fractional differentiating filter may bewritten as

R fð Þ ¼ i2pfð ÞaX fð Þ

¼ 2pfð Þaexp iap2

! "Cf exp iyf

# $

¼ 2pfð ÞaCf exp iyf + iap2

! ";

where it can be seen that the magnitude of R(f ) is (2pf )aCf and the phase ofR(f) has a frequency-independent phase shift of ap/2. If the signal x(t) is a stepfunction, the resulting response after the step will be a power-law decay, as theFourier transform of a power law is a power law.

Note: Supplementary information is available on the Nature Neuroscience website.

ACKNOWLEDGMENTSWe are grateful to B. Aguera y Arcas, W. Bialek, F. Dunn, M. Famulare,M. Giugliano, S. Hong, M. Maravall, P. Murphy, F. Rieke, L. Sorensen andB. Wark for helpful discussions and/or comments on the manuscript. We thankS. Usher for excellent technical assistance. This work was supported by aBurroughs-Wellcome Careers at the Scientific Interface grant, and a McKnightScholar Award and a Sloan Research Fellowship to A.L.F. B.N.L. was supportedby grant number F30NS055650 from the National Institute of NeurologicalDisorders and Stroke, the University of Washington’s Medical Scientist TrainingProgram (supported by the National Institute of General Medical Sciences), andan Achievement Rewards for College Scientists (ARCS) fellowship. M.H.H. andW.J.S. were supported by a Veterans Affairs Merit Review to W.J.S.

AUTHOR CONTRIBUTIONSAll authors conceived of and designed the experiments. B.N.L. and M.H.H.performed the experiments. B.N.L. analyzed the data, performed the modelingand wrote the initial draft. All authors revised the paper.

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

NATURE NEUROSCIENCE ADVANCE ONLINE PUBLICATION 7

ART ICLES

Page 8: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

Published online at http://www.nature.com/natureneuroscience/Reprints and permissions information is available online at http://npg.nature.com/reprintsandpermissions/

1. Adrian, E.D. & Zotterman, Y. The impulses produced by sensory nerve endings: part 2.The response of a single end-organ. J. Physiol. (Lond.) 61, 151–171 (1926).

2. Barlow, H.B. Possible principles underlying the transformation of sensory messages.in Sensory Communication (ed. Rosenblith, W.) 217–234 (MIT Press, Cambridge,Massachusetts, 1961).

3. Brenner, N., Bialek, W. & de Ruyter van Steveninck, R. Adaptive rescaling maximizesinformation transmission. Neuron 26, 695–702 (2000).

4. Fairhall, A.L., Lewen, G.D. & Bialek, W. de Ruyter Van Steveninck, R.R. Efficiency andambiguity in an adaptive neural code. Nature 412, 787–792 (2001).

5. Dean, I., Harper, N.S. & McAlpine, D. Neural population coding of sound level adapts tostimulus statistics. Nat. Neurosci. 8, 1684–1689 (2005).

6. Diaz-Quesada, M. & Maravall, M. Intrinsic mechanisms for adaptive gain rescaling inbarrel cortex. J. Neurosci. 28, 696–710 (2008).

7. Higgs, M.H., Slee, S.J. & Spain, W.J. Diversity of gain modulation by noise in neocorticalneurons: regulation by the slow after-hyperpolarization conductance. J. Neurosci. 26,8787–8799 (2006).

8. Kim, K.J. & Rieke, F. Temporal contrast adaptation in the input and output signals ofsalamander retinal ganglion cells. J. Neurosci. 21, 287–299 (2001).

9. Maravall, M., Petersen, R.S., Fairhall, A.L., Arabzadeh, E. & Diamond, M.E. Shifts incoding properties and maintenance of information transmission during adaptation inbarrel cortex. PLoS Biol. 5, e19 (2007).

10.Nagel, K.I. & Doupe, A.J. Temporal processing and adaptation in the songbird auditoryforebrain. Neuron 51, 845–859 (2006).

11. Sanchez-Vives, M.V., Nowak, L.G. & McCormick, D.A. Cellular mechanisms of long-lasting adaptation in visual cortical neurons in vitro. J. Neurosci. 20, 4286–4299(2000).

12. Smirnakis, S.M., Berry, M.J., Warland, D.K., Bialek, W. & Meister, M. Adaptation ofretinal processing to image contrast and spatial scale. Nature 386, 69–73 (1997).

13.Kvale, M.N. & Schreiner, C.E. Short-term adaptation of auditory receptive fields todynamic stimuli. J. Neurophysiol. 91, 604–612 (2004).

14.Hosoya, T., Baccus, S.A. & Meister, M. Dynamic predictive coding by the retina.Nature436, 71–77 (2005).

15. Toib, A., Lyakhov, V. & Marom, S. Interaction between duration of activity and timecourse of recovery from slow inactivation inmammalian brain Na+ channels. J. Neurosci.18, 1893–1903 (1998).

16. Schwindt, P.C., Spain, W.J. & Crill, W.E. Long-lasting reduction of excitability by asodium-dependent potassium current in cat neocortical neurons. J. Neurophysiol. 61,233–244 (1989).

17. Abel, H.J., Lee, J.C., Callaway, J.C. & Foehring, R.C. Relationships between intracellularcalcium and after-hyperpolarizations in neocortical pyramidal neurons. J. Neurophysiol.91, 324–335 (2004).

18. Fleidervish, I.A., Friedman, A. & Gutnick, M.J. Slow inactivation of Na+ current and slowcumulative spike adaptation in mouse and guinea-pig neocortical neurones in slices.J. Physiol. (Lond.) 493, 83–97 (1996).

19. La Camera, G. et al. Multiple time scales of temporal response in pyramidal and fast-spiking cortical neurons. J. Neurophysiol. 96, 3448–3464 (2006).

20. Schwindt, P.C., Spain, W.J., Foehring, R.C., Chubb, M.C. & Crill, W.E. Slow conduc-tances in neurons from cat sensorimotor cortex in vitro and their role in slow excitabilitychanges. J. Neurophysiol. 59, 450–467 (1988).

21.Destexhe, A., Rudolph, M., Fellous, J.M. & Sejnowski, T.J. Fluctuating synapticconductances recreate in vivo–like activity in neocortical neurons. Neuroscience 107,13–24 (2001).

22.Richardson, M.J. Effects of synaptic conductance on the voltage distribution and firingrate of spiking neurons. Phys. Rev. E 69, 051918 (2004).

23. Crochet, S. & Petersen, C.C. Correlating whisker behavior with membrane potential inbarrel cortex of awake mice. Nat. Neurosci. 9, 608–610 (2006).

24.Hasenstaub, A., Sachdev, R.N. & McCormick, D.A. State changes rapidly modulatecortical neuronal responsiveness. J. Neurosci. 27, 9607–9622 (2007).

25. Thorson, J. & Biederman-Thorson, M. Distributed relaxation processes in sensoryadaptation. Science 183, 161–172 (1974).

26. French, A.S. & Torkkeli, P.H. The power law of sensory adaptation: simulation by amodelof excitability in spider mechanoreceptor neurons. Ann. Biomed. Eng. 36, 153–161(2008).

27.Kleinz, M. & Osler, T.J.A. A child’s garden of fractional derivatives. Coll. Math. J. 31,82–88 (2000).

28. Fourcaud-Trocme, N., Hansel, D., van Vreeswijk, C. & Brunel, N. How spike generationmechanisms determine the neuronal response to fluctuating inputs. J. Neurosci. 23,11628–11640 (2003).

29.Kondgen, H. et al. The dynamical response properties of neuocortical neuronsto temporally modulated noisy inputs in vitro. Cereb. Cortex 18, 2086–2097(2008).

30.Anastasio, T.J. Nonuniformity in the linear network model of the oculomotor integratorproduces approximately fractional-order dynamics and more realistic neuron behavior.Biol. Cybern. 79, 377–391 (1998).

31.Gilboa, G., Chen, R. & Brenner, N. History-dependent multiple time-scale dynamics in asingle-neuron model. J. Neurosci. 25, 6479–6489 (2005).

32.Drew, P.J. & Abbott, L.F. Models and properties of power-law adaptation in neuralsystems. J. Neurophysiol. 96, 826–833 (2006).

33.Powers, R.K., Sawczuk, A., Musick, J.R. & Binder, M.D. Multiple mechanisms ofspike-frequency adaptation in motoneurones. J. Physiol. (Paris) 93, 101–114(1999).

34.Benda, J. & Herz, A.V. A universalmodel for spike-frequency adaptation.Neural Comput.15, 2523–2564 (2003).

35.Mainen, Z.F. & Sejnowski, T.J. Reliability of spike timing in neocortical neurons.Science268, 1503–1506 (1995).

36.Middleton, J.W., Longtin, A., Benda, J. & Maler, L. The cellular basis for parallel neuraltransmission of a high-frequency stimulus and its low-frequency envelope. Proc. Natl.Acad. Sci. USA 103, 14596–14601 (2006).

37. Lundstrom,B.N. & Fairhall, A.L. Decoding stimulus variance froma distributional neuralcode of interspike intervals. J. Neurosci. 26, 9030–9037 (2006).

38. Jolivet, R., Rauch, A., Luscher, H.R. & Gerstner, W. Predicting spike timing ofneocortical pyramidal neurons by simple threshold models. J. Comput. Neurosci. 21,35–49 (2006).

39.Slee, S.J., Higgs, M.H., Fairhall, A.L. & Spain, W.J. Two-dimensional time coding in theauditory brainstem. J. Neurosci. 25, 9978–9988 (2005).

40. Fusi, S., Drew, P.J. & Abbott, L.F. Cascade models of synaptically stored memories.Neuron 45, 599–611 (2005).

41.Paulin, M.G., Hoffman, L.F. & Assad, C. Dynamics and the single spike. IEEE Trans.Neural Netw. 15, 987–994 (2004).

42.Anastasio, T.J. The fractional-order dynamics of brainstem vestibulo-oculomotorneurons. Biol. Cybern. 72, 69–79 (1994).

43. Fairhall, A.L., Lewen, G.D., Bialek, W. & de Ruyter van Steveninck, R. in Advances inNeural Information Processing Systems 13 (eds. Leen, T.K., Dietterich, T.G. & Tresp, V.)124–130 (MIT Press, Cambridge, Massachusetts, 2001).

44.Puccini, G.D., Sanchez-Vives, M.V. & Compte, A. Integratedmechanisms of anticipationand rate-of-change computations in cortical circuits. PLoS Comput. Biol. 3, e82(2007).

45.Wark, B., Lundstrom, B.N. & Fairhall, A. Sensory adaptation.Curr. Opin. Neurobiol. 17,423–429 (2007).

46.Wang, X.J., Liu, Y., Sanchez-Vives, M.V. & McCormick, D.A. Adaptation and temporaldecorrelation by single neurons in the primary visual cortex. J. Neurophysiol. 89,3279–3293 (2003).

47.Ruderman, D.L. & Bialek, W. Statistics of natural images: scaling in the woods. Phys.Rev. Lett. 73, 814–817 (1994).

48.Simoncelli, E.P. & Olshausen, B.A. Natural image statistics and neural representation.Annu. Rev. Neurosci. 24, 1193–1216 (2001).

49.Buzsaki, G. & Draguhn, A. Neuronal oscillations in cortical networks. Science 304,1926–1929 (2004).

50.Hodgkin, A.L. & Huxley, A.F. A quantitative description of membrane current and itsapplication to conduction and excitation in nerve. J. Physiol. (Lond.) 117, 500–544(1952).

©20

08 N

atur

e P

ublis

hing

Gro

up h

ttp:

//ww

w.n

atur

e.co

m/n

atur

eneuroscience

8 ADVANCE ONLINE PUBLICATION NATURE NEUROSCIENCE

ART ICLES

Page 9: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

3

Responses averaged across (a) 8 neurons in response to sine wave current and (b) 11

neurons in response to sine wave noise for a given period. In both cases a phase lead is observed, where the dotted black line is a sine wave with zero phase.

Phase leads increased during the dynamic clamp condition for T = [4 8 16] sec. These data are summarized in Figure 5a.

Reduction of the sAHP by !-methyl-5-HT. For each trace, the magnitude of the early sAHP

current (IsAHP) was approximated by taking the mean membrane voltage at 450-550 msec and subtracting the resting membrane voltage before the train of 30 action potentials at 50-Hz. For clarity, spikes have been removed.

! !"# $

!$

!

$

%&'()*(+&,-

.,+'/0&1(-2+/3(24567

5&8(29/:(2;<++(83

2

2

=(+&,-2>2$2?(;

@

A

B

$C

D@

! !"# $

!$

!

$

%&'()*(+&,-

5&8(29/:(28,&?(a b

10 20 30

10

20

30

Phase during control (deg)

T = 4 sec

Ph

ase

with

dyn

am

ic c

lam

p (

de

g)

10 20 30

10

20

30

Phase during control (deg)

T = 8 sec

10 20 30

10

20

30

Phase during control (deg)

T = 16 sec

L5, compared to 1st control

L2/3, 1st control

L5, 2nd control

L2/3, 2nd control

Figure S6:

Figure S1:

Figure S2:

Page 10: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

4

Phase leads decreased, on average, after application of !-methyl 5-HT. For each neuron (n

= 10), the fraction of the control phase lead ("AfterDrug/"BeforeDrug) for T = [4 8 16] sec remaining after drug application was plotted against the fraction of the sAHP remaining (DVAfterDrug/DVBeforeDrug). Supplementary Equations: Filtering with a fractional differentiator We provide here a more detailed outline of the mathematics underlying fractional differentiation (Miller and Ross, 1993; Podlubny, 1999; Kleinz and Osler, 2000; Sokolov et al., 2002; Oldham and Spanier, 2006), which has already been presented in the main text. We suggest that a single cortical neuron transforms mean current input and the envelope of modulated current noise into mean firing rate in a manner that has the general properties of fractional differentiation. We first clarify the specific filtering properties of fractional differentiation. It is convenient to express fractional differentiation in the frequency domain rather than in the time domain. In the frequency domain, a linear operator such as fractional differentiation is expressed as a filter, or transfer function. Recall

that any function in the time domain can be transformed into the frequency domain (# = 2$f) and back into the time domain using the following Fourier transforms:

X !( ) = e" i! t

x t( )dt"#

#

$

x t( ) =1

2%ei! tX !( )d!

"#

#

$,

so that

x t( )F

! "!# !! X $( ) .

The first derivative with respect to time in the time domain can be written in the frequency domain as

d

dtx t( )

F! "!# !! i$( )X $( ) .

Therefore, the nth derivative of x(t) can be written as

xn( )t( )

F! "!# !! i$( )

n

X $( ) . (1)

Mathematically, n is not restricted to be an integer, and indeed Eq. (1) can be used to define fractional differentiation in the frequency domain for any real n; when n is negative, then Eq. (1) describes

0.2 0.6 1 1.4

0.2

0.6

1

1.4

Fraction sAHP after drug

Fra

ctio

n p

ha

se

le

ad

aft

er

dru

g

T = 4 sec

8

16

Figure S3:

Page 11: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

2

Normalized average firing rate responses show a similar shape regardless of period. Firing

rates from 8 neurons averaged for a given period in response to the (a) low and (c) high condition of square-wave currents. Similar results for 11 neurons are seen in response to the (b) low and (d) high conditions of square-wave noise.

Using fixed bin widths yielded similar results as Figures 1c and 1d in response to (a)

square-wave current and (b) square-wave noise. Here, the inverse of each inter-spike-interval was used to find a firing rate estimate using 50 msec bins, and the resulting firing rate estimates were fit with exponentials as described previously.

! !"#$ !"$

!!"$

!

!"$

%&'()*&+,-(+&).-)+/0'*&)1'2&)30**&(.+

4-*5'678&9)*&+,-(+&

)

)

:&*7-9);)<)+&3

#

=

>

<?

@#

!"$ !"A$ <

!

!"$

<

B75&C,&*7-9

4-*5'678&9)*&+,-(+&

! !"#$ !"$

!!"$

!

!"$

%&'()*&+,-(+&).-)+/0'*&)1'2&)(-7+&

!"$ !"A$ <

!

!"$

<

B75&C,&*7-9

a b

c d

! "! #!

"

#

$

%&'()*+,+-.&/0

,(1&+/)2.3423+!+-.&/0

5674'&+849&+/7''&23:+;(<&*+=(2+8(*3>

+

+

?(@>!3)!A)8B)8!3)!>(@>

! "! #!

"

#

$

%&'()*+,+-.&/0

5674'&+849&+2)(.&:+;(<&*+=(2+8(*3>a b

Figure S4:

Figure S5:

Page 12: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

3

Responses averaged across (a) 8 neurons in response to sine wave current and (b) 11

neurons in response to sine wave noise for a given period. In both cases a phase lead is observed, where the dotted black line is a sine wave with zero phase.

Phase leads increased during the dynamic clamp condition for T = [4 8 16] sec. These data are summarized in Figure 5a.

Reduction of the sAHP by !-methyl-5-HT. For each trace, the magnitude of the early sAHP

current (IsAHP) was approximated by taking the mean membrane voltage at 450-550 msec and subtracting the resting membrane voltage before the train of 30 action potentials at 50-Hz. For clarity, spikes have been removed.

! !"# $

!$

!

$

%&'()*(+&,-

.,+'/0&1(-2+/3(24567

5&8(29/:(2;<++(83

2

2

=(+&,-2>2$2?(;

@

A

B

$C

D@

! !"# $

!$

!

$

%&'()*(+&,-

5&8(29/:(28,&?(a b

10 20 30

10

20

30

Phase during control (deg)

T = 4 sec

Ph

ase

with

dyn

am

ic c

lam

p (

de

g)

10 20 30

10

20

30

Phase during control (deg)

T = 8 sec

10 20 30

10

20

30

Phase during control (deg)

T = 16 sec

L5, compared to 1st control

L2/3, 1st control

L5, 2nd control

L2/3, 2nd control

Figure S6:

Figure S1:

Figure S2:

Page 13: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

Supporting Online Materials Methods The following equations comprise the standard HH model (Hodgkin and Huxley, 1952; Koch, 1999; Dayan and Abbott, 2001; Gerstner and Kistler, 2002):

CdV

dt= !G

Nam

3h V ! E

Na( ) ! G

Kn

4V ! E

K( ) ! G

LeakV ! E

Leak( ) + I

dn

dt= !

n(1" n) " #

nn,

dm

dt= !

m(1" m) " #

mm,

dh

dt= !

h(1" h) " #

hh,

!nV( ) =

0.01 V + 55( )

1" e"0.1 V +55( ) #

nV( ) = 0.125e

" V +65( )/80!,

!mV( ) =

0.1 V + 40( )

1" e"0.1 V +40( ) #

mV( ) = 4e

" V +65( )/18,

!hV( ) = 0.07e

" V +65( )/20 #

hV( ) =

1

1+ e"0.1 V +35( )

,

where steady-state gating values, such as n!, are equal to " / ("+#). Standard parameters for HH are:

GNa = 120, GK = 36, and GLeak = 0.3 mS/cm2; ENa = 50, EK = -77, and ELeak = -54.4 mV; and C = 1 µF/cm2. Data and Discussion Physiological properties of neurons recorded. For 45 of the sharp electrode recordings, the resting

membrane potential was –86 ± 5 mV, the rheobase was 0.47 ± 0.21 nA, and the input resistance, calculated using the steady-state current needed to depolarize the cell from rest to just below

rheobase, was 67 ± 25 M! (mean ± SD). For the patch recordings, three cells were in layer 2/3 and had resting membrane potentials of [-80, –87, –88] mV, rheobases of [0.28, 0.34, 0.51] nA, and chord

resistances of [20, 27, 32] M!; three cells were in layer 5 and had resting membrane potentials of [-72,

-75, -77] mV, rheobases of [0.35, 0.37, 0.70] nA, and chord resistances of [63, 59, 46] M!. Effect of data binning on adaptation measures for square-wave current and square-wave noise. For each T, the time-dependent mean firing rate of each neuron was determined from a histogram of spike times with 30 bins; because equal time was spent on each T and we were looking for a single exponential time constant to describe adaptation, we used a fixed number of bins rather than a fixed time interval for each bin. When the mean firing rates from each neuron were averaged and normalized, it was evident that the shape of the average response was similar over the tested range of T (Figure S1).

Although a bin width of 1/30 T masks exponentials with $ < ~ 0.01 T, in the present studies we

observed exponential relaxation of firing rate with $ ~ 0.1 T. To investigate whether our results were substantially biased by using 30 bins for each T, the firing rate for each neuron was also found using the inverse of each inter-spike interval (ISI), sampled at 500 Hz and down-sampled to a bin width of 50 msec. The results obtained (Figure S2) were similar to Figures 1c and 1d, despite inherent sampling and edge biases. For the modeling data (Figures 6a and 6b), when data sets were large, fixed bin width and fixed bin number gave nearly identical results.

Page 14: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

Phase leads decreased, on average, after application of !-methyl 5-HT. For each neuron (n

= 10), the fraction of the control phase lead ("AfterDrug/"BeforeDrug) for T = [4 8 16] sec remaining after drug application was plotted against the fraction of the sAHP remaining (DVAfterDrug/DVBeforeDrug). Supplementary Equations: Filtering with a fractional differentiator We provide here a more detailed outline of the mathematics underlying fractional differentiation (Miller and Ross, 1993; Podlubny, 1999; Kleinz and Osler, 2000; Sokolov et al., 2002; Oldham and Spanier, 2006), which has already been presented in the main text. We suggest that a single cortical neuron transforms mean current input and the envelope of modulated current noise into mean firing rate in a manner that has the general properties of fractional differentiation. We first clarify the specific filtering properties of fractional differentiation. It is convenient to express fractional differentiation in the frequency domain rather than in the time domain. In the frequency domain, a linear operator such as fractional differentiation is expressed as a filter, or transfer function. Recall

that any function in the time domain can be transformed into the frequency domain (# = 2$f) and back into the time domain using the following Fourier transforms:

X !( ) = e" i! t

x t( )dt"#

#

$

x t( ) =1

2%ei! tX !( )d!

"#

#

$,

so that

x t( )F

! "!# !! X $( ) .

The first derivative with respect to time in the time domain can be written in the frequency domain as

d

dtx t( )

F! "!# !! i$( )X $( ) .

Therefore, the nth derivative of x(t) can be written as

xn( )t( )

F! "!# !! i$( )

n

X $( ) . (1)

Mathematically, n is not restricted to be an integer, and indeed Eq. (1) can be used to define fractional differentiation in the frequency domain for any real n; when n is negative, then Eq. (1) describes

0.2 0.6 1 1.4

0.2

0.6

1

1.4

Fraction sAHP after drug

Fra

ctio

n p

ha

se

le

ad

aft

er

dru

g

T = 4 sec

8

16

Figure S3:

Page 15: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

fractional integration. To emphasize that n may be non-integer, from here onward we use ! instead of n to designate the order of the fractional derivative. In general, a linear, time-invariant system that filters an input x(t) with a filter h(t) and gives an output r(t) can be expressed via convolution as:

r t( ) = h t( )! x t( ) = h t " #( )x(# )d#"$

$

% . (2)

In the frequency domain, convolution becomes multiplication, and so Eq.(2) becomes:

R !( ) = H !( )X !( ) .

From Eq. (1), the filter for fractional differentiation is

H !( ) = i!( )"

.

Thus, if x(t) is some time-varying input variable (specifically, here we have considered the mean or SD of a time-varying current injected into the soma) and r(t) is the neuron’s response as measured by its mean firing rate, then the neuron’s computation can be modeled as:

r t( ) =d!

dt!x t( ) ,

where the parameter ! is the fractional order, or, in the frequency domain,

R !( ) = i!( )"

X !( ) . (3)

Because X(") is in general a complex number, it can be rewritten in its polar form in terms of its

magnitude C" and phase #":

X !( ) = X !( ) exp i!X !( )( ) ,

where

X !( ) = XR!( )

2

+ XI!( )

2

= C!

!X !( ) = arctanXI!( )

XR!( )

"

#$%

&'= (!

,

where XR(") and XI(") are the real and imaginary parts of X("), respectively, and C" and #" denote

the magnitude and phase of the component at frequency ". From Euler’s formula,

exp ix( ) = cos x + i sin x ,

and noticing that:

exp i!2

"#$

%&'= cos

!2+ i sin

!2= i ,

we can now write Eq. (3) as:

R !( ) = i!( )"X !( )

=!"exp i

"#2

$%&

'()C! exp i*!( )

=!"C! exp i*! + i

"#2

$%&

'()

.

(4)

Eq. (4) shows two properties: the magnitude of R(") is "!C" and R(") has a frequency-independent

phase shift of !$/2.

Page 16: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

Response characteristics of a fractional differentiating filter A particular output of this operation that has experimental relevance is the response to step stimuli. Here, we compute how a fractional differentiator should respond to this kind of stimulus. If the input

x(t) is the Heaviside step function, such that x(t)=0 for t<0 and x(t)=1 for t!0, then the input in the frequency domain is

X !( ) =1

i!+ "# !( ) .

Ignoring the constant " when #=0, which means that our result will need to be shifted by some average value, we then have:

H !( )X !( ) =i!( )

"

i!= i!( )

" #1

.

Given the definition of the gamma function as:

! k( ) " # k$1e$#d#

0

%

& ,

by setting $=i#t and d$=i# dt, one finds:

i!( )"k

=1

# k( )tk"1e" i! tdt

0

$

% .

Then, by letting % =1-k,

i!( )" #1

= H !( )X !( ) =1

$ 1#"( )t#"e# i! tdt

0

%

& ,

where the right hand side is proportional to the Fourier transform of t--% with 0<%<1 as the order of the fractional differentiator. Thus, a fractionally differentiating neuron will respond to a step function

with a power law t--%. In summary, here are three properties of the response of a fractional

differentiator with order %:

(1) Its magnitude in the frequency domain is #%C#.

(2) Its phase shift in the frequency domain is (%")/2.

(3) Its time-dependent response to a step increase is proportional to t--%. Form of the fractional differentiating filter in the time domain The form of the fractional differentiating filter in the time domain is

h t( ) = F

!1H "( )#$ %& =

1

2'i"( )

(

!)

)

* exp i"t( )d" .

Using the following approximation for the delta function with finite range of frequencies (Arfken and Weber, 1995),

!

nt( ) =

sin nt

"t=

1

2"exp iwt( )dw

#n

n

$ ,

the form of the filter is

h t( ) = a

Dt

! " t( )#$ %& ,

where fractional differentiation can be defined as (Podlubny, 1999; Oldham and Spanier, 2006):

a Dt

!f t( ) =

1

" 1#!( )

d

dtt # $( )

#!

a

t

% f $( )d$ , 0 & ! < 1( ) .

For discrete time from 0 to t,

h t( ) =1

! 1"#( )d

dtt "$( )

"#

0

t

%sin nt

&td$ , 0 '# < 1( ) .

Page 17: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

Supporting Online Materials Methods The following equations comprise the standard HH model (Hodgkin and Huxley, 1952; Koch, 1999; Dayan and Abbott, 2001; Gerstner and Kistler, 2002):

CdV

dt= !G

Nam

3h V ! E

Na( ) ! G

Kn

4V ! E

K( ) ! G

LeakV ! E

Leak( ) + I

dn

dt= !

n(1" n) " #

nn,

dm

dt= !

m(1" m) " #

mm,

dh

dt= !

h(1" h) " #

hh,

!nV( ) =

0.01 V + 55( )

1" e"0.1 V +55( ) #

nV( ) = 0.125e

" V +65( )/80!,

!mV( ) =

0.1 V + 40( )

1" e"0.1 V +40( ) #

mV( ) = 4e

" V +65( )/18,

!hV( ) = 0.07e

" V +65( )/20 #

hV( ) =

1

1+ e"0.1 V +35( )

,

where steady-state gating values, such as n!, are equal to " / ("+#). Standard parameters for HH are:

GNa = 120, GK = 36, and GLeak = 0.3 mS/cm2; ENa = 50, EK = -77, and ELeak = -54.4 mV; and C = 1 µF/cm2. Data and Discussion Physiological properties of neurons recorded. For 45 of the sharp electrode recordings, the resting

membrane potential was –86 ± 5 mV, the rheobase was 0.47 ± 0.21 nA, and the input resistance, calculated using the steady-state current needed to depolarize the cell from rest to just below

rheobase, was 67 ± 25 M! (mean ± SD). For the patch recordings, three cells were in layer 2/3 and had resting membrane potentials of [-80, –87, –88] mV, rheobases of [0.28, 0.34, 0.51] nA, and chord

resistances of [20, 27, 32] M!; three cells were in layer 5 and had resting membrane potentials of [-72,

-75, -77] mV, rheobases of [0.35, 0.37, 0.70] nA, and chord resistances of [63, 59, 46] M!. Effect of data binning on adaptation measures for square-wave current and square-wave noise. For each T, the time-dependent mean firing rate of each neuron was determined from a histogram of spike times with 30 bins; because equal time was spent on each T and we were looking for a single exponential time constant to describe adaptation, we used a fixed number of bins rather than a fixed time interval for each bin. When the mean firing rates from each neuron were averaged and normalized, it was evident that the shape of the average response was similar over the tested range of T (Figure S1).

Although a bin width of 1/30 T masks exponentials with $ < ~ 0.01 T, in the present studies we

observed exponential relaxation of firing rate with $ ~ 0.1 T. To investigate whether our results were substantially biased by using 30 bins for each T, the firing rate for each neuron was also found using the inverse of each inter-spike interval (ISI), sampled at 500 Hz and down-sampled to a bin width of 50 msec. The results obtained (Figure S2) were similar to Figures 1c and 1d, despite inherent sampling and edge biases. For the modeling data (Figures 6a and 6b), when data sets were large, fixed bin width and fixed bin number gave nearly identical results.

Page 18: Fractional differentiation by neocortical pyramidal neurons...Fractional differentiation by neocortical pyramidal neurons Brian N Lundstrom1, Matthew H Higgs1,2, William J Spain1,2

References

Arfken GB, Weber H-J (1995) Mathematical methods for physicists, 4th Edition. San Diego:

Academic Press.

Dayan P, Abbott LF (2001) Theoretical neuroscience: computational and mathematical modeling of

neural systems. Cambridge, Mass.: Massachusetts Institute of Technology Press.

Gerstner W, Kistler WM (2002) Spiking neuron models: single neurons, populations, plasticity.

Cambridge, U.K. ; New York: Cambridge University Press.

Hodgkin AL, Huxley AF (1952) A quantitative description of membrane current and its application to

conduction and excitation in nerve. Journal of Physiology 117:500-544.

Kleinz M, Osler TJ (2000) A Child's Garden of Fractional Derivatives. The College Mathematics

Journal 31:82-88.

Koch C (1999) Biophysics of computation: information processing in single neurons. New York:

Oxford University Press.

Miller KS, Ross B (1993) An introduction to the fractional calculus and fractional differential

equations. New York: Wiley.

Oldham KB, Spanier J (2006) The fractional calculus : theory and applications of differentiation and

integration to arbitrary order. Mineola, N.Y.: Dover Publications.

Podlubny I (1999) Fractional differential equations : an introduction to fractional derivatives, fractional

differential equations, to methods of their solution and some of their applications. San Diego:

Academic Press.

Sokolov IM, Klafter J, Blumen A (2002) Fractional Kinetics. Physics Today:48-54.