second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 ·...

20
Second-order receptive fields reveal multidigit interactions in area 3b of the macaque monkey Pramodsingh H. Thakur, Paul J. Fitzgerald, and Steven S. Hsiao Departments of Neuroscience and Biomedical Engineering, Zanvyl Krieger Mind/Brain Institute and Kennedy-Krieger Institute, Johns Hopkins University, Baltimore, Maryland Submitted 29 November 2010; accepted in final form 28 March 2012 Thakur PH, Fitzgerald PJ, Hsiao SS. Second-order receptive fields reveal multidigit interactions in area 3b of the macaque monkey. J Neurophysiol 108: 243–262, 2012. First published March 28, 2012; doi:10.1152/jn.01022.2010.—Linear receptive field (RF) models of area 3b neurons reveal a three-component structure: a central excit- atory region flanked by two inhibitory regions that are spatially and temporally nonoverlapping with the excitation. Previous studies also report that there is an “infield” inhibitory region throughout the neuronal RF, which is a nonlinear interactive (second order) effect whereby stimuli lagging an input to the excitatory region are sup- pressed. Thus linear models may be inaccurate approximations of the neurons’ true RFs. In this study, we characterize the RFs of area 3b neurons, using a second-order quadratic model. Data were collected from 80 neurons of two awake, behaving macaque monkeys while a random dot pattern was scanned simultaneously across the distal pads of digits D2, 3, and 4. We used an iterative method derived from matching pursuit to identify a set of linear and nonlinear terms with significant effects on the neuronal response. For most neurons (65/80), the linear component of the quadratic RF was characterized by a single excitatory region on the dominant digit. Interactions within the dominant digit were characterized by two quadratic filters that capture the spatial aspects of the interactive infield inhibition. Interactions between the dominant (most responsive) digit and its adjacent digit(s) formed the largest class of cross-digit interactions. The results dem- onstrate that a significant part of area 3b responses is due to nonlinear mechanisms, and furthermore, the data support the notion that area 3b neurons have “nonclassical RF”-like input from adjacent fingers, indicating that area 3b plays a role in integrating shape inputs across digits. somatosensory; tactile information; infield inhibition; generalized lin- ear model; point process model THERE IS STRONG EVIDENCE that nonlinear mechanisms play a significant role in the initial processing of information in the sensory systems. In this report we investigate the nonlinear receptive field (RF) properties of neurons in area 3b of primary somatosensory cortex (SI) and its relevance to two aspects of somatosensory function. The first is the role of nonlinear mechanisms in two-dimensional (2D) form processing from a single finger. Previous studies have shown that somatosensory inputs related to the perception of 2D form are transformed from an initial isomorphic representation in the periphery into partially transformed responses in SI that are selective to spatial features of the stimuli (Hsiao et al. 2006; Phillips et al. 1988). One way of understanding these transforms is to under- stand the RF structures of the neurons. One of the earliest attempts at characterizing the RF struc- tures of neurons in area 3b of SI involved indenting the RF with temporally and spatially separated paired probes (Gardner 1984; Gardner and Costanzo 1980a, 1980b). These studies revealed that the RF has two components, a central excitatory core with a latency of 10 ms and a spatially overlapping inhibitory field called infield inhibition that is delayed with respect to the excitatory core. Pharmacological blockade of GABAergic transmission increases the size of the RF excit- atory area, as well as the maximum response evoked at the RF center (Alloway and Burton 1991; Dykes et al. 1984). This suggests that infield inhibition is mediated by cortical interneu- rons, whereas the central excitatory core represents the tha- lamic feed-forward input. Thus understanding the respective contributions of excitatory core and infield inhibition to the neural responses may elucidate the roles that feed-forward inputs and local intracortical circuits play in imparting feature selectivity. Infield inhibition has been also shown in intracellular re- cordings as delayed inhibitory postsynaptic potentials (IPSPs) (Andersson 1965; Innocenti and Manzoni 1972; Whitehorn and Towe 1968) and accounts for the decreased response to a test stimulus preceded by a conditioning stimulus within the RF (Gardner and Costanzo 1980a, 1980b; Laskin and Spencer 1979). Since it is only evoked by paired stimuli, and not revealed when single punctate probes are used as stimuli, which show area 3b RFs as simple homogeneous excitatory regions (Iwamura et al. 1983; Mountcastle and Powell 1959; Sur 1980, Sur et al. 1984), infield inhibition is an interactive second-order effect. Linear RF models estimated from responses to multiple simultaneous indentations, via either scanned em- bossed random dot patterns (DiCarlo et al. 1998) or random probe indentations (Sripati et al. 2006), capture inhibition via negative linear weights, but only in spatiotemporal regions beyond the excitatory core. These studies show two components of inhi- bition. The portion of infield inhibition that is temporally coincident with the excitatory core but spatially extends be- yond the excitation is revealed as the temporally coincident component and gives the neurons a center-surround-like struc- ture. The portion of infield inhibition that persists after the excitation has subsided is expressed as the lagging inhibitory component. However, since they are unable to discriminate between the linear excitatory effects and nonlinear inhibitory second-order effects in spatiotemporal regions where these two effects overlap, these linear models approximate the excitatory and inhibitory effects via an average of the two. Since excita- tion is stronger than inhibition, the resulting linear weights are positive in this region of overlap, and the representation of Address for reprint requests and other correspondence: S. Hsiao, Zanvyl Krieger Mind/Brain Inst., Johns Hopkins Univ., 338 Krieger Hall, 3400 N Charles St., Baltimore, MD 21218 (e-mail: [email protected]). J Neurophysiol 108: 243–262, 2012. First published March 28, 2012; doi:10.1152/jn.01022.2010. 243 0022-3077/12 Copyright © 2012 the American Physiological Society www.jn.org by 10.220.33.5 on November 2, 2016 http://jn.physiology.org/ Downloaded from

Upload: others

Post on 22-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

Second-order receptive fields reveal multidigit interactions in area 3bof the macaque monkey

Pramodsingh H. Thakur, Paul J. Fitzgerald, and Steven S. HsiaoDepartments of Neuroscience and Biomedical Engineering, Zanvyl Krieger Mind/Brain Institute and Kennedy-KriegerInstitute, Johns Hopkins University, Baltimore, Maryland

Submitted 29 November 2010; accepted in final form 28 March 2012

Thakur PH, Fitzgerald PJ, Hsiao SS. Second-order receptivefields reveal multidigit interactions in area 3b of the macaque monkey.J Neurophysiol 108: 243–262, 2012. First published March 28, 2012;doi:10.1152/jn.01022.2010.—Linear receptive field (RF) models ofarea 3b neurons reveal a three-component structure: a central excit-atory region flanked by two inhibitory regions that are spatially andtemporally nonoverlapping with the excitation. Previous studies alsoreport that there is an “infield” inhibitory region throughout theneuronal RF, which is a nonlinear interactive (second order) effectwhereby stimuli lagging an input to the excitatory region are sup-pressed. Thus linear models may be inaccurate approximations of theneurons’ true RFs. In this study, we characterize the RFs of area 3bneurons, using a second-order quadratic model. Data were collectedfrom 80 neurons of two awake, behaving macaque monkeys while arandom dot pattern was scanned simultaneously across the distal padsof digits D2, 3, and 4. We used an iterative method derived frommatching pursuit to identify a set of linear and nonlinear terms withsignificant effects on the neuronal response. For most neurons (65/80),the linear component of the quadratic RF was characterized by asingle excitatory region on the dominant digit. Interactions within thedominant digit were characterized by two quadratic filters that capturethe spatial aspects of the interactive infield inhibition. Interactionsbetween the dominant (most responsive) digit and its adjacent digit(s)formed the largest class of cross-digit interactions. The results dem-onstrate that a significant part of area 3b responses is due to nonlinearmechanisms, and furthermore, the data support the notion that area 3bneurons have “nonclassical RF”-like input from adjacent fingers,indicating that area 3b plays a role in integrating shape inputs acrossdigits.

somatosensory; tactile information; infield inhibition; generalized lin-ear model; point process model

THERE IS STRONG EVIDENCE that nonlinear mechanisms play asignificant role in the initial processing of information in thesensory systems. In this report we investigate the nonlinearreceptive field (RF) properties of neurons in area 3b of primarysomatosensory cortex (SI) and its relevance to two aspects ofsomatosensory function. The first is the role of nonlinearmechanisms in two-dimensional (2D) form processing from asingle finger. Previous studies have shown that somatosensoryinputs related to the perception of 2D form are transformedfrom an initial isomorphic representation in the periphery intopartially transformed responses in SI that are selective tospatial features of the stimuli (Hsiao et al. 2006; Phillips et al.1988). One way of understanding these transforms is to under-stand the RF structures of the neurons.

One of the earliest attempts at characterizing the RF struc-tures of neurons in area 3b of SI involved indenting the RFwith temporally and spatially separated paired probes (Gardner1984; Gardner and Costanzo 1980a, 1980b). These studiesrevealed that the RF has two components, a central excitatorycore with a latency of �10 ms and a spatially overlappinginhibitory field called infield inhibition that is delayed withrespect to the excitatory core. Pharmacological blockade ofGABAergic transmission increases the size of the RF excit-atory area, as well as the maximum response evoked at the RFcenter (Alloway and Burton 1991; Dykes et al. 1984). Thissuggests that infield inhibition is mediated by cortical interneu-rons, whereas the central excitatory core represents the tha-lamic feed-forward input. Thus understanding the respectivecontributions of excitatory core and infield inhibition to theneural responses may elucidate the roles that feed-forwardinputs and local intracortical circuits play in imparting featureselectivity.

Infield inhibition has been also shown in intracellular re-cordings as delayed inhibitory postsynaptic potentials (IPSPs)(Andersson 1965; Innocenti and Manzoni 1972; Whitehorn andTowe 1968) and accounts for the decreased response to a teststimulus preceded by a conditioning stimulus within the RF(Gardner and Costanzo 1980a, 1980b; Laskin and Spencer1979). Since it is only evoked by paired stimuli, and notrevealed when single punctate probes are used as stimuli,which show area 3b RFs as simple homogeneous excitatoryregions (Iwamura et al. 1983; Mountcastle and Powell 1959;Sur 1980, Sur et al. 1984), infield inhibition is an interactivesecond-order effect. Linear RF models estimated from responsesto multiple simultaneous indentations, via either scanned em-bossed random dot patterns (DiCarlo et al. 1998) or random probeindentations (Sripati et al. 2006), capture inhibition via negativelinear weights, but only in spatiotemporal regions beyond theexcitatory core. These studies show two components of inhi-bition. The portion of infield inhibition that is temporallycoincident with the excitatory core but spatially extends be-yond the excitation is revealed as the temporally coincidentcomponent and gives the neurons a center-surround-like struc-ture. The portion of infield inhibition that persists after theexcitation has subsided is expressed as the lagging inhibitorycomponent. However, since they are unable to discriminatebetween the linear excitatory effects and nonlinear inhibitorysecond-order effects in spatiotemporal regions where these twoeffects overlap, these linear models approximate the excitatoryand inhibitory effects via an average of the two. Since excita-tion is stronger than inhibition, the resulting linear weights arepositive in this region of overlap, and the representation of

Address for reprint requests and other correspondence: S. Hsiao, ZanvylKrieger Mind/Brain Inst., Johns Hopkins Univ., 338 Krieger Hall, 3400 NCharles St., Baltimore, MD 21218 (e-mail: [email protected]).

J Neurophysiol 108: 243–262, 2012.First published March 28, 2012; doi:10.1152/jn.01022.2010.

2430022-3077/12 Copyright © 2012 the American Physiological Societywww.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 2: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

infield inhibition is lost. To fully disentangle the effects ofexcitation and inhibition, especially in regions of spatiotempo-ral overlap, it is essential to extend the linear models to includesecond-order terms to accurately capture interactive infieldinhibitory effects.

The second aspect of somatosensory processing investigatedhere is multidigit interaction. Although neurons in the digitregion of area 3b have traditionally been thought to be drivenexclusively by inputs from single finger pads, optical imagingstudies (Chen et al. 2001, 2005) indicate that subthresholdinputs to area 3b neurons may extend over an area substantiallylarger than suprathreshold inputs. Also, cortical plasticity stud-ies (Wang et al. 1995) have shown that coincident presentationof stimuli on adjacent digits expands neuronal RFs in area 3bto span multiple digits, indicating that an underlying substratefor multidigit interactions may exist in area 3b. Recently, Reedet al. (2006) showed, using multiple probe stimuli, that re-sponses of neurons with palmar RFs can be modulated bystimuli outside the classical RF, like nonclassical RF effectsobserved in the visual system (see Allman et al. 1985). The RFmechanisms that underlie these nonclassical effects in area 3bneurons are not known.

In this study, we estimate multidigit second-order RFs ofarea 3b neurons. Neural responses were obtained with methodsused previously (DiCarlo et al. 1998) except that random dotpatterns were scanned simultaneously across three distal pads(i.e., digits D2–4) of the awake macaque. RFs were thenestimated by using a point process model that includes bothlinear and second-order (quadratic) terms. To circumvent thedimensional explosion due to inclusion of quadratic terms inthe model, we used an iterative method based on matchingpursuit (Mallat and Zhang 1993) to determine a subset of linearand quadratic terms that have the most salient effects on theneuronal responses. We then investigated how these nonlinearproperties affect the neural responses.

METHODS

Animals and Surgery

Single-unit recordings were made with standard methods fromthree cortical hemispheres in two male rhesus monkeys (M. mulatta)weighing 5–7 kg. Briefly, before the recordings, animals were trainedto sit quietly in a primate chair (see below). Surgery was thenperformed to secure a head-holding device and recording chambers(19-mm diameter) to the skull. Surgical anesthesia was induced withketamine HCl (20 mg/kg im) and maintained with pentobarbital(10–25 mg·kg�1·h�1 iv). All surgical procedures were done understerile conditions and were reviewed and approved by the JohnsHopkins Animal Care and Use Committee and in accordance with therules and regulations of the Society for Neuroscience.

General Recording Techniques

All recordings were performed in awake, behaving monkeys. Themonkeys were trained to sit in a primate chair with their handsrestrained and were rewarded with sips of water at random intervals(about once every 7 s) while tactile stimuli (see below) were presentedto the distal pads of digits D2, D3, and D4. On each recording day, amultielectrode microdrive (Mountcastle et al. 1991) was loaded withseven quartz-coated platinum-tungsten (90:10) microelectrodes (shaftdiameter 80 �m, tip diameter 4 �m, impedance 1–5 M� at 1,000 Hz).On approximately half of the recording days, microelectrodes werecoated with fluorescent dye (DiI; Molecular Probes, Eugene, OR) to

facilitate later histological localization of the recording sites (DiCarloet al. 1996). For all recordings, the tip of the microdrive was insertedinto the recording chamber and oriented with the microelectrodesnormal to the dura mater.

The seven microelectrodes were oriented at an angle of 45° to themediolateral axis (parallel to the central sulcus) with center-to-centerspacings of 400 �m. On each recording day, the microelectrodes werepositioned within the recording chamber with micrometer-controlledtranslation stages (model 430; Newport, Irvine, CA) (accuracy within5 �m).

To record in SI, we used a stereotaxic instrument to first center therecording chamber over the central sulcus [Horsley-Clarke (HC)coordinates: anterior 6. lateral 21]. The central sulcus was then locatedwithin the recording chamber based on the responses encountered onthe microelectrodes. Responses anterior to the sulcus are motor,whereas responses posterior to the sulcus are somatosensory. Once thelocation of the sulcus was determined, the microelectrode array wasadjusted to span the digit representation, just posterior to the centralsulcus.

Because SI lies in the postcentral gyrus, we recorded first superfi-cially before driving the electrodes deeper into cortex. Thus, for eachposition of the array within the chamber, recordings were first madein area 1, then on the next day in area 3b, and then on the third dayin area 3a. After sampling the responses from the three areas, themicroelectrode array was moved systematically along the postcentralgyrus until neurons ceased to be driven by the digits. Determination ofwhich area we were recording from was based on the progression ofRFs from the distal pads to the proximal pads and then to the distalpads as the electrodes passed from area 1 to area 3b and confirmedpostmortem. Results in this report are based on the data collected fromarea 3b.

Data Collection

The signals from the seven channels were band-pass filtered andfed into a spike sorter (Alpha Omega), which detected and classifiedaction potentials based on their similarity to spike templates. Neuronsthat met the following criteria were selected for analysis: 1) the actionpotentials were well isolated from noise and from other spike tem-plates; 2) the neuronal firing was stable throughout the stimulusprotocol; and 3) the neuronal RF, as determined by handheld probes,was located on the distal pads.

Stimulus

The stimuli consisted of raised dot patterns fabricated from sheetsof photosensitive plastic that are water-soluble until exposed to UVlight (Toyoba Printing Plastics, EF series). The patterns are producedby laying a photographic negative of the pattern over the plastic sheetand exposing it to UV light, which polymerizes and hardens theexposed plastic. The portion of the surface layer not exposed to UVlight is scrubbed off lightly in water, leaving a raised pattern whosethickness equals that of the original water-soluble layer (380 �m). Thestimulus pattern was specified in Postscript and printed at 3,386dots/in. After construction, the sheet containing the stimulus patternswas wrapped around and glued to a cylindrical drum, 640 mm incircumference, which was mounted on a rotating drum stimulator(Johnson and Phillips 1988).

The random dot patterns were designed to be similar to the patternsused in a previous study (DiCarlo et al. 1998). Each dot was 380 �mhigh (in relief) and 400 �m in diameter at its top, with sides thatsloped away at 60° relative to the surface of the drum. The location ofthe center of each dot was determined by selecting two randomnumbers from a uniform random number generator and scaling themto the height (40 mm) and width (560 mm) of the pattern. The dotcenters were specified to a precision of 1 �m, and dots were allowedto overlap when the random number generator placed them closer than

244 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 3: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

the dot diameter. Dots were added until the average dot densityequaled 10 dots/cm2.

After one or more neurons with RFs (as determined by qualitativemapping) on the distal pads of digit D2, D3, or D4 were isolated, thedrum with the random dot pattern was lowered and positioned so thatit simultaneously scanned across the distal pads of digits D2, D3, andD4. The pattern was scanned in the proximal to distal direction at aspeed of 25 mm/s, while a torque motor maintained a contact force of90g. After each complete rotation, the drum was shifted laterally by200 �m, thus presenting a slightly different pattern during successivescans. Data were collected continuously for a minimum of 30 scans,or �13 min. The contact patch of the finger on the drum wasmonitored with a metallic “hot” pin (1-mm diameter) placed on thesurface of the drum. Contact of the finger with this pin triggered aseries of square wave pulses that were stored as events in the datastream. This pulse signal was then used to align the section of stimuluspattern that was in contact with the pads at any given time step withthe neural response.

Linear and Quadratic Receptive Field Estimation

The recorded spike train was binned (32-ms bin width) to yield aspike count vector (Rob � |r1, r2, . . .rT|) (see Fig. 1). Each stimulatedpad was divided into subregions of size 800 �m � 800 �m, yielding432 subregions (3 pads � 144 subregions each) that could potentiallyaffect the neural response. For a linear model, each subregion wasassociated with a weight wi that defines its linear excitatory orinhibitory influence on the neural response. The neural firing rate atthe nth step is related to the linear combination (�n) of the correspond-ing weights at each subregion multiplied by the number of dots xi(n)in that subregion at the nth time step (see Fig. 1A). Specifically,

�n � w0 � �i�1

432

xi�n�wi (1)

In this model, excitation and inhibition are captured by positive andnegative weights, respectively.

The model was estimated in two steps. First, a subset of terms orpixels (usually �100) that had a salient contribution to the neuralresponse was identified with a dimensionality reduction method basedon matching pursuit (see APPENDIX). Weights were then estimated byfitting a generalized linear model (GLM) (function glmfit, MATLAB;MathWorks) with a log-link function between the selected subset ofterms and observed spike count vector. This assumes a Poissondistribution of spike counts [i.e., rn � Poisson(e�n)], yielding alog-linear model in input covariates or subregions (x). Poisson as-sumption has been shown to be reasonable with a sufficient number ofrepeated trials or at intervals beyond refractory periods (Kass et al.2005; Ventura et al. 2002).

For the quadratic model, in addition to the linear weights asdescribed above, quadratic weights (wij) were defined to capture everypair of possible interactions between subregions (i,j) across the threepads (see Fig. 1B). Such weights denote the interaction between thespecific pairs of subregions, and their values indicate the effect ofstimulating the two subregions simultaneously on the neural response.Specifically, the model was defined as

�n � w0 � �i�1

432

wixi�n� � �i�1

432

�j�1

432

wijxi�n�xj�n� (2)

where xi and xj are the number of dots stimulating the ith and jthsubregions during the nth time step. The second-order weight wij

captures the effect of simultaneously indenting subregions i and j onthe neural response. A negative (positive) weight indicates suppres-sive (facilitative) interaction due to a simultaneous stimulation ofsubregions i and j. On the other hand, the contribution of this term iszero if either of the two subregions i and j is not stimulated. Thus,under this model, any linear effect will be captured by the linear

weights wi, which we refer to as the linear component of the quadraticmodel, whereas the interactive effects between pairs of subregions arecaptured by the second-order weights wij. If the infield inhibition is alinear effect it will be captured by negative linear weights, but if it isan interactive effect it will be captured by the second-order weights wij

and will not be found in the linear component of the quadratic model.Like the linear model, the quadratic model was estimated in two

steps. First, a set of salient terms was identified with matching pursuit,followed by an estimate of the weights corresponding to those terms

Fig. 1. Design of linear kernels. A demonstrates how the linear kernels and ratevector (Rob) were constructed for the random dot data. Discrete time steps wereobtained by binning the entire duration of the experiment with suitable timebins. The nth entry in the rate vector array equals the number of spikes duringthe nth time step. Xab and Xac are two 1 pixel � 1 pixel kernels that span pixelsat coordinates (a,b) and (a,c), respectively. The nth entry of each vector equalsthe number of dots within the areas spanned by the vectors during the nth timestep. During the 1st illustrated frame, there is a dot at (a,b) but no dot at (a,c).Hence the corresponding entries in vectors Xab and Xac during this time step are1 and 0, respectively. During the 2nd illustrated frame, there are no dots at(a,b) but a dot at (a,c). Hence the entries in the 2 vectors at this time step are0 and 1, respectively. B shows construction of the quadratic kernels. The nthentry of each vector equaled the product of the number of dots in the 2subregions (i,j) and (i,k) spanned by that vector during the nth time step.

245MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 4: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

via fitting a GLM, yielding a log-quadratic model in input covariatesor subregions (x) [i.e., rn � Poisson(e�n)].

Linear and Quadratic Kernels Spanning Multiple Subregions

With an average dot density of 10 dots/cm2 (see above), theprobability of stimulating a subregion at a given instant of time is P �10/(12.5 � 12.5) � 0.064. With 598 bins each from 30 scans (N �598 � 30 � 17,984 unique equations for fitting), the expected numberof nonzero occurrences of stimulus dots at any subregion over theperiod analyzed is N � P � 1,148. However, the number of simul-taneous occurrences of nonzero dot counts over two subregions(occurrences of simultaneous stimulation to uncover quadratic re-sponse) is N � P � P � 74. Because of such low prevalence of pairedstimulation, the errors in estimating quadratic weights are likely to behigher. To improve the reliability of quadratic weights, we expandedour dictionary to include “kernels” that span multiple subregions (upto 5 subregions � 5 subregions or 4 mm � 4 mm) and used matchingpursuit to iteratively pull out the best matching linear or quadratickernel at any stage. For example, a linear kernel spanning 2 � 2subregions was defined as the total number of dots within all 4subregions spanned. A quadratic kernel spanning 2 � 2 subregionswas defined as the product of the number of dots within the two 2 �2 regions represented by that kernel. Mathematically, using suchmultiple subregion spanning kernels translates into assuming thatweights are similar for all subregions spanned (and hence combiningthem and estimating a single weight instead). For example, a linearkernel BL spanning subregions x1, x2, x3, and x4 involves the followingapproximation:

w1x1 � w2x2 � w3x3 � w4x4 � w�x1 � x2 � x3 � x4� � wBL

Similarly, a 2 � 2 quadratic kernel Bq representing a “combined”interaction between subregions x1, x2, x3, and x4 and subregions x5, x6,x7, and x8 involves the following approximation:

w15x1x5 � w16x1x6 � w17x1x7 � w18x1x8 � . . . �w45x4x5 � w46x4x6

� w47x4x7 � w48x4x8 � w�x1 � x2 � x3 � x4��x5 � x6 � x7 � x8�� wBq

Although the prevalence of paired stimulation of subregions is low,the prevalence increases dramatically as we consider kernels spanningmultiple subregions. For example, the probability of finding at leastone dot within 4 subregions is P � 10/(6.25 � 6.25) � 0.256. As aresult, the expected number of occurrences of nonzero terms for a 2 �2 quadratic kernel is N � P � P � 1,174, which is similar to a 1 �1 linear kernel spanning just one subregion. For our analysis linearkernels spanning 1 � 1, 2 � 2, 3 � 3, 4 � 4, and 5 � 5 subregionsand quadratic kernels spanning 2 � 2, 3 � 3, 4 � 4, and 5 � 5subregions were included in the dictionary for matching pursuit; 1 �1 quadratic kernels were excluded because of their low prevalence.

Multiple Instances of Linear and Quadratic Models via RandomizedMatching Pursuit

Matching pursuit is an iterative algorithm that finds the best-fittingterms at any iteration from a dictionary of kernels. At the end of eachiteration, it projects the remaining terms in the dictionary into a spaceorthogonal to the response vector, so that at subsequent iterations thealgorithm picks up terms that best capture the portion of the responsevector not accounted for up to the current iteration (see APPENDIX).Given the iterative nature of the algorithm, a noisy term picked up atany step has a bearing on all subsequent terms picked up by thealgorithm. Further, since the algorithm is fitting to a spike countvector, which is a single observation of a random process, it is likelyto pick up noise at any step. To minimize the effects of noisy termsbeing picked up at any given step, we obtained multiple instances ofeach type of model by using randomization. Rather than picking the

kernel with the maximum projection at every iteration, we randomlyselected a kernel with projection �75% of the maximum projection.Because of the randomization, the algorithm selects a different se-quence of vectors at each instance. We hypothesize that the signalterms or kernels are selected more often across the different instancessince they should depend on the systematic component of the responsecommon across all data sets, whereas a given noisy kernel will beselected only on a few runs since the noise component of the responsewill be dependent on the current rate vector at that iteration. The finalweights obtained for each type of model were averaged across 25 suchinstances.

Expressing Second-Order Terms as Quadratic Filters

The excitatory and inhibitory effects of a linear RF can be easilyvisualized by plotting the weights at the corresponding spatial loca-tions. However, second-order interactions between different spatiallocations cannot be visualized directly from the quadratic terms. Toobtain a visual display of the pattern of interactions between variousspatial locations, we obtained an equivalent representation of Eq. 2 interms of 2D quadratic filters. Specifically, Eq. 2 can be rewritten inmatrix form:

�n � w0 � AnTWL � An

TWQAn (3)

where

An ��x1�n�x2�n�

x432�n��, WL ��

w1

w2

w432

� and WQ

��w11 w12 ⁄ 2 · · · w1,432 ⁄ 2

w12 ⁄ 2 w22 · · · w2,432 ⁄ 2

� � · · · �

w1,432 ⁄ 2 w2,432 ⁄ 2 · · · w432,432

�WQ is known as the Hessian matrix and represents the interactiveeffect between pairs of subregions in a compact matrix form.

To obtain a visual display of the pattern of interactions betweenvarious spatial locations, an equivalent representation of the Hessianmatrix can be obtained in terms of its eigenvalues (�i) and eigenvec-tors (qi):

WQ � �i�1

432

�iqiqiT

Substituting in Eq. 3,

�n � w0 � AnTWL � �

i�1

432

�i�AnTqi�2 � w0 � An

TWL

� �i�1

432 �i

��i��An

T����i�qi�2(4)

The term�i

|�i|is 1 if �i is positive and �1 if negative and determines

whether the contribution of a particular eigenvector is suppressive(negative) or facilitative (positive).

Thus the eigenvectors, scaled by the root of the absolute value ofthe corresponding eigenvalue (i.e., ��iqi), can be considered asquadratic filters that contribute to the response by the square of theconvolution of these filters with the input stimulus pattern. Because ofthe eventual squaring operation, the overall sign of any quadratic filteris immaterial. (If a quadratic filter is multiplied by �1, so that thesigns of the weights of all subregions are reversed, the contribution ofthe filter to the response remains unchanged.) Thus the effect of agiven quadratic filter depends on the sign and magnitude of thecorresponding eigenvalue, as well as the relative signs of the weights

246 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 5: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

of different subregions in the quadratic filter. If a pair of subregions inthe quadratic filter has the same sign, then a negative eigenvalueimplies suppression, whereas a positive eigenvalue implies facilita-tion. On the other hand, if subregions have opposite signs, then anegative eigenvalue implies facilitative interaction, whereas a positiveeigenvalue implies a suppressive interaction (see APPENDIX for detailson interpretation). These quadratic filters are analogous to second-order RF structures obtained via second-order methods such as spike-triggered covariance.

Separating Within-Digit and Cross-Digit Interactions

Instead of obtaining quadratic filters from the Hessian matrixdirectly, we transformed the matrix in such a way that it enables adirect and straightforward quantification, comparison, and interpreta-tion of the different types of within-digit and cross-digit interactions.For every neuron we classified each of the three pads as 1) maximallyresponsive digit (Dmax: defined as the pad containing the subregiongiving the highest linear response), 2) digit(s) adjacent to the maxi-mally responsive digit (Dadj), and 3) nonadjacent digit (Dnonadj). Thequadratic weights representing the interaction between two subregionswere sorted into six classes of within-digit or cross-digit interactionsdepending upon the locations of the two subregions on the digits,namely, 1) Dmax � Dmax: both the subregions on Dmax; 2) Dadj �Dadj: both the subregions on Dadj; 3) Dnonadj � Dnonadj: both thesubregions on Dnonadj; 4) Dmax � Dadj: one subregion each on Dmax

and Dadj; 5) Dmax � Dnonadj: one subregion each on Dmax and Dnonadj;and 6) Dadj � Dnonadj: one subregion each on Dadj and Dnonadj.

Next, the Hessian matrix was expressed as a combination of sixcomponent matrices that represent specific types of within-digit orcross-digit interaction. Specifically,

WQ � WDmax�Dmax� WDadj�Dadj

� WDnonadj�Dnonadj� WDmax�Dadj

� WDadj�Dnonadj� WDmax�Dnonadj

The (i,j)th element of the component matrix WDyxDz (whichrepresents the interaction between digits y and z) is equal to thecorresponding element of WQ if subregion i belongs to digit y andsubregion j belongs to digit z or subregion i belongs to digit z andsubregion j belongs to digit y and is zero otherwise. Thus this is asimple redistribution or sorting of the various terms of WQ into sixgroups (6 matrices) based on the location of the subregions that eachterm represents. Each component matrix, when subjected to an eigen-vector/eigenvalue decomposition, yields quadratic filters that capturea specific type of interaction represented by the parent componentmatrix WDyxDz.

The quadratic filters that correspond to within-pad interactions on themaximally responsive digit (Dmax � Dmax) were obtained by setting allthe quadratic weights, except the weights representing the Dmax � Dmax

type of interactions, to zero before decomposing the Hessian matrix intoeigenvalues and eigenvectors. Similarly, the quadratic filters representingthe Dmax � Dadj type of interactions were obtained by setting all thequadratic weights, except the weights representing the Dmax � Dadj typeof interactions, to zero before decomposing the Hessian matrix.

Comparing Strengths of Different Types of Interactions

The relative strength of the six classes of interactions was quanti-fied by the ratio of the sum of squares of the quadratic weightsbelonging to a particular type of interaction to the sum of squares ofall the quadratic weights (called fraction of quadratic energy capturedby a particular type of interaction). The significance of each type ofinteraction was assessed by comparing the distribution with a nulldistribution obtained by repeating the entire model estimation proce-dure with the stimulus pattern on the adjacent digits shuffled along thetime axis in the regression equations.

Strength of a Quadratic Filter

We again used measures based upon the sum of squares ofquadratic weights to characterize the fraction of total quadraticstrength accounted for by a given quadratic filter. If WDyxDz repre-sents the Hessian matrix constructed from the weights correspond-ing to a particular type of interaction (DyxDz), then the sum ofsquares of all the weights corresponding to this particular interac-tion can be written as

�WDyxDz�2 � �WDyxDz

T WDyxDz� � �� �i�1

n�iqiqi

T�T� �i�1

n�iqiqi

T�� � �i�1

n�i

2

Also, the norm of the projection of a quadratic filter (defined asQFiQFi

T) is �i2. Hence, the fraction of the quadratic strength of a

component matrix accounted for by a given quadratic filter can be

measured by the ratio�i

2

�i�1n �i

2.

Homogeneity of a Quadratic Filter

Since the relative signs of different regions of a quadratic filterdetermine the nature of interactive contribution of the filter to theneural response (see APPENDIX), we estimated a homogeneity metricfor each filter to characterize whether a given filter had similarlysigned pixels throughout the RF or a balance of positive and negativeregions. The RF homogeneity index measured (on a scale of 0–1) therelative strengths of regions with positive and negative signs withinthe quadratic filter. Specifically, if n1 and n2 are the sums of squaresof positive and negative weights, respectively, then we define thehomogeneity index as

HI ��n1 � n2�

n1 � n2

HI � 0 indicates that the positive and negative regions within thequadratic filter have equal strengths, whereas HI � 1 indicates that allof the nonzero pixels of that quadratic filter have similar signs.

The strength of the contribution of a quadratic filter was measuredas the ratio of the square of the corresponding eigenvalue to the sumof squares of all the quadratic weights representing that particular typeof interaction.

Cross Validation

The data set was divided into two separate sets for the purposes ofestimating free parameters (80%) and cross-validation (20%). Weused the fraction of the variance in the 20% of data set aside that canbe accounted for by a given model as a measure of predictive r2.Specifically, if yi is the observed response at the ith time step and yi

is the response predicted by a given model, we obtained the total sumof squares (TSS) and the residual sum of squares (RSS) as

TSS � �i�1

N yi �1

N�j�1

N

yj�2

RSS � �i�1

N

�yi � yi�2

In addition, we obtained an estimate of the variance attributed to noise(NSS) by assuming successive scans as repeats (see DiCarlo et al.1998 for details).

NSS � �i�1

N

Var�Yi�

where Yi � [yi, yi�T, yiT] Here T is the number of steps in each scan.Predictive r2 was defined as

247MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 6: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

r2 �TSS � RSS

TSS � NSS

The procedure was repeated 150 times, using a different set of 20%of the data each time. For each combination of 80–20 (data estima-tion-cross validation), model response yi was averaged across 25instances of the model obtained by randomization (see above). Thefinal predictive r2, linear and quadratic weights, and the predictedresponse reported in this article were averaged across all 150 repeats.

Linear Model with Spiking History

To test whether the improvements observed with a quadratic modelover a linear model could be attributable to known neural nonlineari-ties such as spiking history (Paninski et al. 2004; Pillow and Simon-celli 2003; Truccolo et al. 2005, 2010), we also evaluated the good-ness of fit of the linear model (Eq. 2) with terms added to capturespiking history. Specifically, the linear models derived by matchingpursuit as described above were expanded by including a linearcombination of terms denoting the effects of spiking history at a delay as follows:

�n � w0 � � xi�n�wi � ��1

max

wh · rn�

Here wh captured the weight/contribution of the neural firing at ahistory of ms.

The model was separately evaluated for multiple levels of spikinghistory such as max � 32 ms, max � 160 ms, and max � 320 ms byfitting a GLM, again yielding a log-linear model in linear terms andspiking history [i.e., rn � Poisson(e�n)]. Note that unlike previousmodels, which let the algorithm iteratively pick salient terms from adictionary of potential candidates, we simply appended terms forspiking history to the linear model obtained earlier (i.e., Eq. 1). As aresult, if spiking history does not contribute to neuronal firing, it maylead to overfitting and hence poorer performance on cross validation.

RESULTS

Comparing Responses Predicted by Linear and QuadraticReceptive Fields

Figure 2 shows the responses of the linear model and thequadratic model (consisting of linear and quadratic terms asdescribed in Eq. 2) of a typical 3b neuron. The linear RF (Fig.2A) indicates that the major drive for this neuron is from D3.Consistent with previous studies (DiCarlo et al. 1998), thelinear RF on this digit comprises an excitatory region (red)surrounded by an inhibitory region (blue). The inhibitoryregion is shifted distally on the finger pad with respect to theexcitatory region, which could be due to spatial effects or dueto inhibition that is delayed with respect to the excitation(DiCarlo et al. 1998; Gardner and Costanzo 1980b). Since themodel fitting procedure correlates the responses with the in-stantaneous pattern of dots on the finger pad, this temporallytrailing inhibition can only manifest in the linear RF by beingshifted distally along the scan direction. However, trailinginhibition is an interactive effect between two dots separatedalong the temporal axis (Gardner and Costanzo 1980b). Sincea linear RF, by definition, is restricted to just the linear terms,the fitting procedure accounts for this inhibitory effect bymeans of negative weights (DiCarlo et al. 1998) in regions onthe pad surrounding the excitatory spot. When quadratic termsare included in the model, the linear component is composed

solely of an excitatory area (Fig. 2B), with the trailing andinfield inhibition being captured by the quadratic terms.

The first raster of Fig. 2C displays the observed response asa raw spatial event plot (SEP, consisting of 30 successive scansof the drum plotted as 30 rows). Note that the raw SEP reflectsneuronal all-or-none firings, which are probabilistic events. Tovisualize the underlying firing rate, the second raster plotdisplays the same SEP convolved with a 2D Gaussian ( � 1pixel). The next two raster plots correspond to the responses ofthe linear (4th raster) and quadratic (3rd raster) RF over thedata set. The bottom band (5th raster) displays the differencebetween the responses of the quadratic and the linear RFs.Regions where the quadratic model outputs a higher (lower)response than the linear model are indicated in red (blue).Overall, both the linear and quadratic responses qualitativelymatch well with the observed responses, suggesting that thelinear models capture much of the primary responses of theneuron. However, at a finer scale, there are significant differ-ences between the predictions of the two models. Examples ofsuch differences are illustrated via enlarged images of threeregions (namely, A, B, and C) in Fig. 2D.

The number of dots stimulating the excitatory region of theRF is higher at region B than region A. Hence the linear RFoutputs a proportionately higher response at bins around B(compare the overall intensity of the top 2 panels in the 4thcolumn of Fig. 2D). However, as indicated by the convolvedSEP of Fig. 2D, the observed responses are unaffected by thedot density. This is because, unlike peripheral neural responsesthat increase their firing rate logarithmically as the average dotdensity within the RF increases, area 3b neurons are largelyunaffected by the average dot density (Sripati et al. 2006). Thequadratic terms correct for this difference by increasing themodel response around A (red region in the top panel in the lastcolumn of Fig. 2D) and lowering the model response around B(blue region in the center of last column of Fig. 2D) such thatthe responses are similar around A and B, which matches thesimilarity of observed responses around A and B. As a result,the quadratic RF model provides a closer match to the observedresponses and better mimics the invariance in the neuralresponses to changes in average dot density.

Region C in Fig. 2D corresponds to a time step when thenumber of dots in the inhibitory region of the linear RF exceedsthe number of dots in the excitatory region. As indicated by thebottom panel of the fourth column in Fig. 2D, the linear RFpredicts no response around this region in the raster, althoughthe observed response shows neuronal firing (1st and 2ndcolumns, last row). This suggests that the inhibition in thelinear RF is a result of a linear approximation to trailinginhibitory effects, which is in fact an interactive or quadraticeffect. A quadratic filter is better at capturing the inhibition bysecond-order weights, and hence the linear component of thequadratic RF (Fig. 2B) differs from the purely linear model andlacks the distal inhibitory region. The result is a better match tothe observed data (bottom panel of the 3rd column in Fig. 2D),as the quadratic model predicts a nonzero response around Csimilar to the observed response.

Characterization of Linear Component of Quadratic Field

The linear components were classified based on the strengthsof the positive and negative weights within the three stimulated

248 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 7: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

pads, with respect to the maximum weight. For most neurons(65/80), the negative weights within the dominant digit or theweights on the nondominant digits did not exceed 10% of themaximum weight, indicating that the linear response is domi-nated by a single excitatory region exclusively on a single pad(Fig. 3A, left). A few neurons had an inhibitory region sur-

rounding the excitatory region (5/80) (Fig. 3A, 2nd from left),another excitatory region on an adjacent pad (multidigit) (2/80)(Fig. 3A, center), distinct multiple regions of excitation on asingle pad (1/80) (Fig. 3A, 2nd from right), an inhibitory regionon some other pad (2/80) (Fig. 3A, right), or no linear excitatorycomponent at all (5/80) (not shown). Figure 3B displays the

Fig. 2. Comparison of responses of linear and quadratic receptive fields (RFs). A: structure of a linear RF of a sample neuron. Red (blue) indicates excitation(inhibition). The 3 squares indicate 1 cm � 1 cm regions around the centers of distal pads of digits D2, D3, and D4. B: the linear component of the quadraticRF for the same neuron. C: a section of the random dot stimuli (top) and the raw response as a spatial event plot (SEP) (1st raster). An estimate of the underlyingfiring rate is obtained by convolving the SEP with a 2-dimensional (2D) Gaussian filter of width ( � 1 pixel) (2nd raster). The next 2 rasters show the responsesof the linear RF (3rd raster) and the quadratic RF (4th raster). Each raster comprises 30 rows corresponding to 30 scans of 9.6 s each. Each successive scan wasobtained by shifting the stimulus laterally by 200 �m. The last row (5th raster) shows the difference in the response predicted by the quadratic RF over that ofthe linear RF. Regions where the quadratic model outputs a higher (lower) response than the linear model are indicated in red (blue). At time steps A and B, theexcitatory region of the linear RF was stimulated by 1 and 5 dots, respectively. At time step C, the number of dots in the inhibitory region of the linear RFexceeded the number of dots in the excitatory region of the linear RF. D: 1st column shows the stimulus pattern overlaid with the linear RF of the neuron attime steps A, B, and C. Columns 2–5 show 11 � 11 arrays of bins around A, B, and C for observed responses (2nd column), responses of the quadraticRF (3rd column) and linear RF (4th column), and the difference between the 2 model responses (5th column). Black dots in the panels correspond to timebins A, B, and C.

249MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 8: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

distribution of the excitatory area measured by counting thesubregions with a positive weight. The mean area was 10.87 mm2

(min 3 mm2, max 21 mm2), which is similar to previous reports(DiCarlo et al. 1998: mean 14 mm2, min 3 mm2, max 43 mm2).

The linear excitatory components were further characterizedby fitting ellipses to the excitatory region (Fig. 3A). Wemeasured the eccentricity of the fitted ellipse as the ratio of thedifference between the lengths of the major and minor axes tothe sum of the lengths of the two axes (0 � circular, 1 �sharply elliptical). The distribution of eccentricities across thepopulation (Fig. 3C) has a mean of 0.09 (min 0.004; max 0.53),which is significantly less than 0.5 (P � 0.01; 1-sample t-test),indicating that the linear components of the excitatory part ofthe quadratic RF of many area 3b neurons are not sharplyelliptical. The orientation of the major axis denotes the direc-tion that the ellipse is oriented along. The distribution oforientations across the population has a peak around 90° (Fig.3D), which corresponds to the distal-proximal axis, althoughthe distribution is not significantly different from a uniformcircular distribution (Raleigh test, P � 0.5).

Thus the salient characteristic of the linear component of thequadratic RF across the population was a mildly ellipticalsingle-digit excitatory hot spot.

Quadratic Interactions Within and Between Digits

The quadratic terms were separated into six classes that weredefined relative to the maximally responsive digit (see METH-ODS). The relative strength of each type of interaction is shownin Fig. 4A. As expected, the fraction of quadratic strengthwithin a single digit is largest for the maximally responsivedigit (34%). The adjacent and nonadjacent digits by themselvesaccount for a small fraction of the total quadratic strength (9%and 5%, respectively). This is reasonable since neurons in area3b are classically thought to have RFs confined to a single

digit. Surprisingly, a large fraction of the total quadraticstrength is accounted for by interactions between digits (45%).The largest interdigit interactions are between the maximallyresponsive digit and its adjacent digit (24%), followed by theinteractions between the maximally responsive digit and thenonadjacent digit (12%) and interactions between the adjacentdigit and the nonadjacent digit (9%). These distributions implythat there is a systematic reduction in the interaction within andbetween digits as we progressively move away from the centerof the neuron’s RF. The interactions within the maximallyresponsive digit and its adjacent digits account for nearly 60%of the total quadratic input to the response. These two compo-nents are characterized in greater detail in the next section.

Testing Validity of Model Fits

In any system identification framework, the response ob-tained to any set of inputs can be characterized as having asystematic or driven component and a noise component. Theaim of a system identification procedure is to capture as muchof the systematic component without fitting the noise. Modelscan be tested for noise fitting with cross validation. However,given that we use matching pursuit with a very large dictionaryof potential terms there is a possibility that the algorithm willfind some combination of terms that fits the systematic com-ponent that does not have any functional meaning.

To test whether the quadratic terms that we observed arefunctionally relevant or are an artifact of blind curve-fitting, weobtained a null distribution of fractions of quadratic energiesby repeating the model estimation procedure with the stimuluspattern on the Dadj and Dnonadj shuffled along the time axis inthe regression framework. Shuffling the stimulus pattern ondigits other than the maximally responsive digit removes sys-tematic relationships between the neural responses and theactual stimulus pattern present on the digits. As such, quadratic

Fig. 3. Characterization of the linear component of the quadratic RF. A: the different types of linear components along with their frequency encountered in thepopulation. A majority of the neurons were characterized by a single excitatory component on the dominant digit (1st panel), and just a few had excitatory spotson multiple digits (3rd panel), multiple spots on a single digit (4th panel), or a single inhibitory spot (5th panel). The ellipses that were fitted to characterize theexcitatory spot are also shown. B: distribution of the areas of the excitatory components measured by counting the pixels with positive linear weights.C: distribution of the eccentricities (measured as ratio of difference of lengths of the major and minor axes to their sum) of the fitted ellipse. D: distribution ofthe orientation of the fitted ellipses obtained as the orientation of the major axis.

250 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 9: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

terms related to Dadj and Dnonadj that are picked up after theshuffling are most likely a result of blind curve-fitting.

Figure 4B compares the null distribution (red curve) and theactual distributions (blue curves) of quadratic energies for thefive types of interactions related to Dadj and Dnonadj. The nulland actual distributions for Dmax � Dadj, Dmax � Dnonadj,Dnonadj � Dnonadj, and Dadj � Dadj types of interactions aresignificantly different (P � 0.0001; paired t-test), suggestingthat these classes of quadratic interactions we observed arenot fully attributable to noise. The null and actual distribu-tions for the Dnonadj � Dnonadj type of interaction were not

significantly different (P � 0.4; paired t-test) (Fig. 4B, topright), indicating that the within-digit interaction on thenonadjacent digits is small and is no better than the fit to thenoise itself.

Within-Digit Quadratic Component on the MaximallyResponsive Digit

We characterized the quadratic component on Dmax byestimating the quadratic filters corresponding to the mostnegative (termed filter Q1), and most positive (termed filter

Fig. 4. Comparison of the different types of quadratic interactions. A: distributions of the fraction of quadratic energies captured by the 3 classes of within-digitinteractions, namely, Dmax � Dmax (maximally responsive digit), Dadj � Dadj (pads adjacent to maximally responsive digit), and Dnonadj � Dnonadj (pad notadjacent to maximally responsive digit), and the 3 types of cross-digit interactions, namely, Dmax � Dadj, Dmax � Dnonadj, and Dadj � Dnonadj, with 2 differentmethods. Fraction of quadratic strength was measured as the ratio of the sum of the squares of quadratic weights corresponding to each type of interaction tothe sum of the squares of all the quadratic weights (see APPENDIX). B: cumulative plots comparing the distributions of fractions of quadratic strengths (blue) withthose expected by chance (red). Chance-level distributions were obtained by shuffling the stimulus patterns on Dadj and Dnonadj across time in the model estimationprocedure (see METHODS).

251MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 10: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

Q2) eigenvalues of the Dmax � Dmax quadratic componentmatrix.

Filter Q1. Figure 5A shows a distribution of the RF homo-geneity index (HI), which measures the relative strengths ofregions with positive and negative signs within the quadraticfilter (see METHODS) for filter Q1 across all the neurons. Thedistribution is significantly skewed toward 1 (H0: median �0.5, P � 0.0001, Wilcoxon signed-rank test), indicating thatfor a majority of the neurons the nonzero subregions of filterQ1 tend to have similar signs. This is further illustrated byplotting filter Q1 for three sample neurons from three differentlocations on the distribution. Neuron C, which is drawn from aregion with high HI (HI � 1), is characterized by a single largeregion with a negative sign. Neuron B, drawn from the center(HI � 0.55), again has a dominant large region with a positivesign, flanked by a small region of negative sign on the proximalside. Neuron A, which represents a small population of neurons

with low HI (HI � 0.17), has two regions with opposite signsaligned along the proximal to distal or scan direction.

Filter Q1 was further characterized by segmenting the larg-est contiguous lobe. Figure 5B, which plots the offset betweenthe center of this lobe and the center of the correspondinglinear component, indicates that for most neurons the center ofthis lobe spatially overlaps with the linear component (mean xoffset 0.33 0.94 mm: mean y offset 0.14 1.13 mm). Themean area across the population (Fig. 5C) is 51.4 mm2, whichis significantly greater than the area spanned by the linearcomponent (mean area 10.87 mm2) (P � 0.0001; paired t-test).This is further illustrated in Fig. 5D, which plots the ratio of thetwo areas. The linear component spans only �30% of the areaspanned by the largest similarly signed contiguous region ofQ1 across the population.

We characterized the shape of this region by fitting an ellipseand estimating the eccentricity (Fig. 5E) and orientation (Fig.

Fig. 5. Characterization of filter (Q1) corresponding to most negative eigenvalue of Dmax � Dmax interaction matrix. A: distribution of homogeneity index thatmeasured the relative strengths of positive and negative regions within the filter on a scale of 0 (balance between positive and negative regions) to 1 (dominatedby a single sign). Sample filters from 3 different ranges of the distribution are shown at bottom. B: difference between the centers of the single largest lobe ofthis quadratic filter and the center of the excitatory component of the linear component. C: distribution of area of the single largest lobe. D: distribution of theratios of the area of linear component to the area of the single largest lobe of filter Q1. E: distribution of the eccentricities of ellipses fitted to the single largestlobe of filter Q1 across the population. F: distribution of orientation of the fitted ellipse (measured as orientation of major axis of fitted ellipse). G: distributionof the fraction of Dmax � Dmax quadratic energy captured by filter Q1, measured as the ratio of the square of the eigenvalue associated with this filter (equalsthe Frobenius norm of the projection matrix associated with this filter) to the sum of squares of all Dmax � Dmax quadratic weights (see APPENDIX).

252 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 11: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

5F) of the fitted ellipses. The mean eccentricity across thepopulation is 0.17, which is significantly higher than the meaneccentricity of the linear component (P � 0.0001; pairedt-test). The distribution of orientation of the major axis had apeak at 90°, which corresponds to the distal-proximal axis, andthe distribution is significantly different from a uniform circu-lar distribution (Raleigh test, P � 0.01). Figure 5G shows adistribution of the fraction of the Dmax � Dmax type ofquadratic strength accounted for by this filter (see METHODS).On average, this filter accounts for �59% of the total within-pad quadratic strength on the maximally responsive digit.

Thus the salient characteristic of filter Q1 across the popu-lation was a single contiguous region, either positive or nega-tive, that was larger in size, more elliptical, and spatiallyoverlapped with the linear component of the quadratic RF.

Filter Q2. In contrast to Q1, the distribution of the homo-geneity index for filter Q2 (Fig. 6A) is significantly skewedtoward 0 (H0: median � 0.5, P � 0.0001, Wilcoxon signed-rank test), indicating that for a majority of the neurons thestrengths of regions with positive and negative signs arebalanced. This is illustrated by plotting filter Q2 for threesample neurons from three different locations on the distribu-

tion. Neurons A and B, which are drawn from regions with lowto moderate HIs (0.14 and 0.53, respectively) and are repre-sentative of the majority of the neurons, are characterized by apositive and a negative lobe each aligned nearly along thedistal-proximal axis. Neuron B, drawn from the center (HI �0.53), has, in addition to the two lobes, a few small regions inthe surrounding space. Neuron C, which represents the smallpopulation of neurons with high HI (HI � 0.78), has a largepositive region flanked by a few weak negative regions.

Filter Q2 was characterized by segmenting the largest con-tiguous positive and negative lobes. We measured the fractionof the total Q2 filter structure captured by these two lobes bytaking the ratio of the sum of squares of weights of the pixelswithin the two lobes to the sum of squares of all nonzeroweights within the filter (Fig. 6B). Across the population, thetwo lobes capture �86% of the total structure of filter Q2.Figure 6C, which plots the offset between the mean of thecenters of the two lobes and the center of the correspondinglinear component, indicates that, like filter Q1, for most neu-rons the two lobes spatially overlap with the linear component(mean x offset 0.58 1.4 mm; mean y offset 0.5 1.7 mm).Figure 6D shows the distribution of orientation of the line

Fig. 6. Characterization of filter (Q2) corresponding to most positive eigenvalue of Dmax � Dmax interaction matrix. A: distribution of homogeneity index thatmeasured the relative strengths of positive and negative regions within the filter on a scale of 0 (balance between positive and negative regions) to 1 (dominatedby a single sign). Sample filters from 3 different ranges of the distribution are shown at bottom. B: fraction of the total structure on this filter captured by thesingle largest positive and single largest negative lobe. C: difference between the mean of the centers of largest positive and largest negative lobes of filter Q2and the center of the excitatory component of the linear component. D: orientations of the line joining the centers of the 2 lobes. E: distribution of the fractionof Dmax � Dmax quadratic energy captured by filter Q2. F: distribution of the fraction of Dmax � Dmax quadratic energy captured by both filters Q1 and Q2.

253MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 12: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

joining the centers of the two lobes. The distribution has a peakat 90°, which corresponds to the distal-proximal axis, and thedistribution is significantly different from a uniform circulardistribution (Raleigh test, P � 0.05). On average, this filteraccounts for �12% of the total within-pad quadratic strengthon the maximally responsive digit (Fig. 6E). Filters Q1 and Q2together capture �71% of the total quadratic energy on themaximally responsive digit (Fig. 6F).

Thus filter Q2 was characterized across the population by apair of lobes, opposite in sign and aligned along the proximal-distal axis or the scan direction.

Functional Roles of Quadratic Filters

To illustrate the roles played by the filters characterized above,we compared output responses of the linear component and thetwo quadratic filters Q1 and Q2 with the observed responses fora sample neuron (Fig. 7). The outputs for the quadratic filtersare obtained by squaring the convolution of the filters withthe instantaneous stimulus patterns scaled by the corre-sponding eigenvalue (see METHODS). Since the squaring op-eration always results in a positive number, the sign of theoutput response (i.e., whether the effect is facilitative or

Fig. 7. Contributions of the linear component and quadratic filters Q1 and Q2 to the overall quadratic RF response. Top left: estimate of the underlying firingrate obtained by convolving the raw SEPs with a 2D Gaussian ( � 1 pixel) (1st raster), the complete output of the quadratic model (2nd raster), outputs of justthe linear component (3rd raster), quadratic filter corresponding to the most negative eigenvalue, i.e., Q1 (4th raster), and quadratic filter corresponding to themost positive eigenvalue, i.e., Q2 (5th raster). At time step A, the linear component of the quadratic RF for this neuron was stimulated by a single dot. At timesteps B and C, the multiple dots stimulating the neuron’s RF were either in a single region of Q2 (C) or spread evenly between the 2 regions of Q2 (B). Blackcircles highlight time steps A, B, and C in each raster.

254 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 13: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

suppressive) is determined by the sign of the correspondingeigenvalue. Q1 provides a suppressive interaction, which isevident by the negative contributions (blue) shown in Fig. 7(4th raster). In contrast, the contribution of Q2 lacks anynegative predictions (blue color) because of the positivesign of the eigenvalue (Fig. 7, 5th raster).

The roles of the three filters in shaping the neural responseare further illustrated via three time bins A, B, and C. Thepanels on the top right and bottom of Fig. 7 show the stimuluspatterns during these time bins overlaid with the linear com-ponent and the two quadratic filters for the sample neuron. Thecorresponding locations of these time points are highlighted byblack circles on each raster. The neuron is stimulated by asingle dot at A but multiple dots at B. Since a linear filterlinearly scales its output with the number of dots, the responseof the linear component is significantly higher at B than A(compare the 2 left circles on the 3rd raster in Fig. 7).However, the observed responses around these bins, as high-lighted by the corresponding circles on the first raster, indicatethat the neuron is slightly inhibited around B compared with A.Filter Q1 corrects for this discrepancy by contributing a strongsuppressive response at B (compare the corresponding circleson the 3rd raster).

The example denoted by region C also corresponds to a timebin when the stimulus consisted of multiple dots in the respon-sive region. However, unlike B, which corresponds to aninstance when the multiple dots are spread evenly between thetwo lobes of filter Q2, C corresponds to an instance when themultiple dots were located within one of the two lobes of filterQ2. Q2 provides no response at B but provides a strongfacilitative contribution at C (last 2 panels in last column, Fig.7). Thus Q2 masks or reduces the suppression provided by Q1if a pair of dots is located within one of the two lobes of Q2 butleaves the suppression unaffected if the two dots happen to bein the two lobes. For the majority of the neurons, the two lobesof Q2 are nearly aligned along the proximal-distal direction(Fig. 7), which is also the scan direction. This shows that filterQ2 does not mask the suppression provided by Q1 if the pairof dots is aligned along the scan direction.

Thus filter Q1 provides a nonspecific suppression in a regionoverlapping with the neuronal RF that scales quadratically withdot density within the RF and offsets the linear scaling with dotdensity provided by the linear component of the RF. Filter Q2modulates the suppression provided by Q1 depending uponwhere the dots are located with respect to the scan axis. Itreduces the suppression provided by Q1 if the multiple dots arelocated orthogonal to the scan axis but leaves the suppressionunaffected if the dots are located along the scan axis. In otherwords, filter Q1 captures the spatial extent of infield inhibition,whereas filter Q2 captures the temporal or trailing effects ofinhibition along the scan direction.

Characterizing Quadratic Interaction Between Digits

As mentioned above, the interaction between the maximallyresponsive digit and its adjacent digits was the second largestquadratic component across the population. In this section, wecharacterize this component by looking at the quadratic filterscorresponding to the most positive and most negative eigen-values of the Hessian matrix that represents this type ofinteraction. These filters are complementary with respect to

each other in the sense that they have identical spatial structure,except that the overall sign on one of the digits is reversed (seesample examples in Fig. 8A). Thus we will characterize thespatial properties of just one of the two filters.

Figure 8A shows the distribution of the homogeneity indexfor the components of this filter on the maximally responsivedigit (top) and the adjacent digit (bottom). The two distribu-tions are significantly skewed toward 1 (P � 0.0001, Wilcoxonrank sum test), indicating that for the majority of the neuronsthe components of these filters on each pad tend to have similarsigns. If the components on the two digits are opposite for thefilter corresponding to the most negative eigenvalue (or similarfor the filter corresponding to the most positive eigenvalue,since the two filters have complementary structures), it indi-cates facilitative interaction (see APPENDIX for interpretation).The majority of neurons with a linear excitatory component onDmax (52/75) had facilitative interaction between Dmax andDadj, as indicated by the relative signs of the largest lobes of thefilters on the two digits. Figure 8B, which plots the ratio of areaspanned by the linear component to the area spanned by thelargest lobe of this filter on the maximally responsive digit,indicates that the linear component spans only �33% of thearea spanned by the largest lobe. Figure 8C plots the differencebetween the center of the largest lobe on the maximallyresponsive digit and the center of the corresponding linearcomponent. Compared with filters Q1 and Q2 describedabove, the largest lobe of this filter has a greater offset withrespect to the linear component (mean x offset 0.66 1.79mm, mean y offset 0.82 2.2 mm). Across the population,these filters capture �30% of the total Dmax � Dadj type ofinteraction (Fig. 8D).

Figure 8E illustrates the contribution of the multidigit qua-dratic terms in the response SEP of a sample neuron. The toptwo rasters plot the raw SEP and an estimate of the underlyingfiring rate [obtained by convolving with a 2D Gaussian ( � 1pixel)], while the third and fourth rasters plot the outputs ofsingle-digit quadratic and multidigit quadratic models, respec-tively. The last raster we plot the difference between theoutputs of the two models. Red spots (highlighted by bluecircles) on the last raster indicate regions on the SEP where themultidigit RF response was significantly greater than the sin-gle-digit RF. The corresponding locations on the neuronalresponse (highlighted by red circles) with the difference plotshows that a majority of the differences correlate with regionswith bursts of high firing. This indicates that the presence ofcoincident stimuli on the maximally responsive digit and theadjacent digit (activating the Dmax � Dadj interaction terms)tends to increase the local firing rate to specific features.

Predictive Power of Models

We used cross validation to compare the predictive power ofthe quadratic model with that of the linear model by predictingresponses to 20% of the data based on weights estimated from80% of the data. Because of the approximation built into theestimate of the portion of the total sum of squares attributable tonoise (i.e., NSS; see METHODS), individual estimates of predictiver2 may not be bounded between �1 and 1 like traditional regres-sion. For example, individual samples in Fig. 9A, which showsthe distribution of 150 samples of predictive r2 for the linear(blue) and quadratic (red) models for an example neuron, may

255MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 14: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

256 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 15: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

exceed 1 depending upon the errors in the estimate of NSS forthat data set. However, on a population level across multipleruns, we expect them to converge to the underlying truepredictive r2. Thus absolute values of predictive r2 for anymodel may be prone to errors (see DiCarlo et al. 1998 for adescription of the error), but a comparison across models onthe same data set is independent of such errors, since theywould be common across all models. Thus, even though wesee considerable variability in estimates of predictive r2 fora particular model, the distribution for the quadratic modelfor the neuron in Fig. 9A is significantly greater than thedistribution for the linear model, indicating that the qua-dratic model has significantly more predictive content forthis neuron. Figure 9B compares the median predictive r2

across 150 samples for the linear model with the quadraticmodel for all neurons. For essentially all of the neurons(79/80), the quadratic model performs substantially betterthan the linear model, indicating that a quadratic model is

systematically better at capturing the interactive infieldinhibition and nonclassical RF effects.

The iterative algorithm used to select a subset of terms forthe linear and quadratic models was terminated based upon astop condition derived from an expected �2-distribution ofchance level projections (see APPENDIX). Using an identical stopcondition (� � 0.01; see APPENDIX) for the linear and quadraticmodels (Fig. 9B) maintains parity between the two models butresults in a larger number of terms in the quadratic model (Fig.9C). Smaller order or number of free parameters of the linearmodel could potentially yield poorer performance due to un-derfit. We obtained a variant of the linear model by letting theiterative algorithm run until the number of terms equaled thoseof the corresponding quadratic model (order-matched linearmodel). Although matching the order to the quadratic modelprovided a modest improvement in the predictive r2 (Fig. 9D),it is still substantially lower than that of the quadratic model,indicating that the improved fit provided by the quadratic

Fig. 8. Characterization of the interaction between the maximally responsive digit and its adjacent digit. A: distribution of homogeneity index linear componentsof the quadratic RF for 2 sample neurons. For the 2 neurons, quadratic filters corresponding to the most negative eigenvalue (1st column) and the most positiveeigenvalue (2nd column) of the Dmax � Dadj component matrix are shown together with the corresponding linear component (3rd column). B: ratio of the areaof the quadratic filters on the maximally responsive digit to the excitatory area of the linear component on the maximally responsive digit. C: difference betweenthe centers of the part of the quadratic filter on the maximally responsive digit and the centers of the excitatory regions of the linear component. D: distributionof the quadratic strength of these filters, measured as the ratio of the sum of the squares of the eigenvalues associated with these filters for a neuron (equals theFrobenius norm of the projection matrix associated with this filter) to the square of the norm of the Dmax � Dadj component matrix. E: raw SEP for a sampleresponse showing observed response (1st raster), an estimate of the underlying firing rate obtained by convolving the SEP with a 2D Gaussian ( � 1 pixel)(2nd raster), single-digit RF output (3rd raster), and multidigit RF output (4th raster). The last raster plots the difference between the outputs of the multidigitand single-digit RFs.

Fig. 9. Cross validation performance of the linear and quadratic RFs. The weights were estimated using 80% of the data. Cross validation was measured as thefraction of the variance of the remaining 20% of data that is accounted for by a given model. A shows the distribution of predictive r2 for the linear (blue) andquadratic (red) models for a sample neuron across the 150 repeats using a unique set of 20% data for prediction over each repeat. B compares the average (median)predictive r2 for the linear (x-axis) and quadratic (y-axis) models across all neurons. C compares the average number of terms (free parameters) in the linear(x-axis) and quadratic (y-axis) models. D compares the predictive r2 between the original linear model with an extended version of the linear model with numberof free parameters matched to the corresponding quadratic model for that neuron. E compares the predictive r2 for the original linear model with versions of thelinear model that incorporated spiking history over 32 ms (red), 160 ms (green), and 320 ms (blue).

257MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 16: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

model cannot be attributed to differences in order of themodels.

To test whether the quadratic effect is attributable to knownneural nonlinearities such as spiking history, we compared theoriginal linear model with variants of the linear model withterms explicitly accounting for spiking history (Fig. 9E) up to32 ms (red), 160 ms (green), and 320 ms (blue). Adding termsto account for spiking history degrades the model, resulting inpoorer fits and indicating that the interactive effects in area 3bneurons captured by the quadratic terms are not attributable tospiking history.

DISCUSSION

In this study, we investigated the nonlinear interactions inthe responses of area 3b neurons by estimating quadraticRFs with linear and quadratic terms spanning the distal padsof digits D2– 4. Unlike previous studies (DiCarlo et al.1998; Sripati et al. 2006) that used a multiple regressionframework between linear terms and a neuronal spike countto estimate linear RF, we used a GLM with a log-linkfunction that yields a log-linear relationship between theinput subregions and firing rate. A GLM fit may be bettergiven the Poisson nature of spike counts; however, weobtained similar results when we used a multiple regressionframework like the previous studies (results not shown).Since the inclusion of second-order terms in the quadraticmodel potentially causes a dimensional explosion in thenumber of terms in the model, we first estimated, using amethod based on matching pursuit, a subset of the quadraticterms that contribute significantly to the observed response.We find that, for a majority of the neurons, the linearcomponent of the RF consisted of a single excitatory regionon the maximally responsive digit. Furthermore, we showthat there are quadratic interactions on the maximally re-sponsive digit, as well as quadratic interactions between themaximally responsive digit and its adjacent digit(s).

About 71% of the quadratic interactions on the maximallyresponsive digit were captured by two filters that correspondphysiologically to previously reported nonlinear inhibitoryfields, namely, infield suppression and trailing inhibition (Di-Carlo and Johnson 2000; Gardner 1984). Furthermore, theinteraction between the maximally responsive digit and itsadjacent digit(s) was characterized by a pair of filters thattogether represented facilitative or suppressive modulations ofthe response on the dominant digit.

Since the method that we have derived from matchingpursuit is statistical, it cannot completely eliminate pickingup terms that fit noise. Thus at least a part of the observedmultidigit interactions may still be attributable to noise.However, we believe that this contribution is small because1) the strengths of different types of quadratic energies (Fig.4A) show a systematic reduction in interactions as we move awayfrom the center of the RF and terms fitting to noise would tend todistribute randomly across the RF and 2) the distributions ofobserved fractions of quadratic energies are significantly greaterthan a chance distribution obtained by fitting the response toarbitrary dot patterns obtained by shuffling the dot patterns on theadjacent and nonadjacent fingers (Fig. 4B).

Previous Studies

Previous studies in area 3b have revealed overlapping ex-citatory and inhibitory inputs throughout the RF of the neuron(Alloway and Burton 1991; Dykes et al. 1984). However,infield inhibition is rarely revealed by single punctate probes(Iwamura et al. 1983; Mountcastle and Powell 1959; Sur 1980;Sur et al. 1984), which show area 3b RFs as homogeneousexcitatory regions (Sur 1980). This is in agreement with thepredictions of our quadratic RFs. Since a single stimuluscannot evoke interactive components, the observed RF will bedominated by the linear component of the quadratic RF, whichis a single excitatory region for most neurons.

Linear RFs estimated with a scanned random dot pattern(Dicarlo et al. 1998) or multiple probes randomized in spaceand time (Sripati et al. 2006) do reveal a significant amount ofinhibition, but a linear RF fails to capture that part of inhibitionthat spatially and temporally overlaps with the excitatory core.Since the excitatory core is stronger than the overlappinginhibition, the weights are positive in the overlapping region.By including quadratic terms, we have disentangled the feed-forward excitatory drive (excitatory core), captured as thelinear excitatory component, and the infield inhibition, whichmanifests as spatially overlapping suppressive quadratic filters.These filters span an area larger than the linear componentacross the population, which is consistent with previous reportssuggesting that the inhibition extends over an area larger thanthe revealed excitatory region (Alloway et al. 1989; Allowayand Burton 1991; DiCarlo et al. 1998; Gardner and Costanzo1980b). A linear RF reveals only parts of this infield inhibitionas a temporally coincident but spatially offset inhibitory com-ponent or a spatially coincident but temporally laggingcomponent (DiCarlo and Johnson 2000; Sripati et al. 2006).Since the interactive inhibition in the overlapping region isa systematic effect that the linear models fail to capture butis accounted for by the second-order model, quadratic mod-els are better descriptions of the response than simple linearmodels (Fig. 9B).

Another factor that may contribute to the infield inhibition,especially in regions of small RF sizes such as the distal pads,is skin mechanics. The role of skin mechanics is evident in theperiphery as surround inhibition (Sripati et al. 2006; Vega-Bermudez and Johnson 1999) in the RFs of SA1 afferents,although there are no known synaptic mechanisms in theperiphery that could lead to lateral interactions. Since thesevery same afferents constitute the major input of area 3bneurons, they are likely to pass on the skin mechanical effectsin their response.

In addition to spatial interactions, studies have revealed atemporal effect of infield inhibition in the form of a reducedresponse to the second of a pair of temporally separated stimuli(Gardner and Costanzo 1980b). Since we stimulated the skinwith a constant scan velocity and direction (proximal-distal),effects along the temporal axis manifest as spatial effects alongthe distal-proximal axis. Thus the filter corresponding to themost positive eigenvalue, which overlaps with the linear com-ponent and consists of two lobes of opposite sign separatedalong the distal-proximal axis, captures the temporal aspect ofinfield inhibition in the form of suppressive interaction be-tween the two regions of opposite sign separated along thedistal-proximal direction. We hypothesize that if the scan

258 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 17: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

direction is varied the two lobes of filter Q2 will move alongthe scan direction, whereas if the scan velocity is increased(decreased) the distance between the two lobes will increase(decrease). Furthermore, if a randomized spatiotemporal stim-ulus is used instead of a scanned stimulus to obtain a spatio-temporal quadratic RF, the two lobes will lie along the timeaxis. This is analogous to the fact that the trailing componentof infield inhibition manifests (though partly) as a “replacinginhibition” along the temporal axis if a spatiotemporal stimulusis used (Sripati et al. 2006) but manifests as a spatially offsetcomponent of inhibition along the scan direction if a scannedstimulus is used (DiCarlo and Johnson 2000).

Pharmacological blockade of GABAergic transmission re-veals an underlying excitatory drive from adjacent digits. Thisdrive is attributable to intracortical lateral connections due togreater response latencies (Alloway and Burton 1991). How-ever, given that �10% of the neurons in area 3b show multi-digit responses to punctate probes (Sur 1980), the excitatorydrive from adjacent digits is balanced by the inhibitory drivepotentially through GABAergic interneurons. Our results sug-gest that this balance between lateral excitation and inhibitionmay be slightly offset. Such an offset may be the result of ashort delay due to the extra synapse involved with GABAergicinterneurons. This offset may not be sufficient to drive theneuron to a stimulus on the adjacent digit but may modulate themembrane potential so that the neuron fires more vigorously toa coincident volley of excitatory postsynaptic potentials (EP-SPs) arising from the stimulation of the maximally responsivedigit. Such a modulatory effect may be manifested as thequadratic interaction between the maximally responsive digitand its adjacent digits.

Functional Significance

Our results suggest that the transformation performed byarea 3b neurons on the input spatial pattern has a significantnonlinear component, which may play an important role inshaping the feature selectivity. The linear excitatory core inarea 3b is weakly elliptical and may have a role in shapingfeature selectivity such as orientation selectivity. Thus orien-tation selectivity observed in area 3b (Bensmaia et al. 2008;Hyvarinen and Poranen 1978; Pubols and Leroy 1977; Warrenet al. 1986) may arise because of the eccentricities observed inthe spatially overlapping quadratic filters. A bar oriented alongthe major axis of this filter will result in a greater interactivesuppression (and weaker response) than a bar oriented alongthe minor axis.

The presence of a significant interaction between the maxi-mally responsive digit and its adjacent digits implies thatintegration of tactile information across pads begins as early asarea 3b. For a stimulus spanning multiple digits, the “exten-sion” of the stimulus on the adjacent digits provides a contex-tual modulation of the response to the part of the stimuluswithin the dominant digit. As studies in the visual system haveshown, such a contextual modulation may be critical for therepresentation of extended contours (Fitzpatrick 2000; Gilbertand Wiesel 1990), corners (Silito et al. 1995), or local curva-ture (Krieger and Zetzsche 1996; Wilson and Richards 1992)and thus provides a first step of transforming the isomorphicinput from the peripheral afferents to a nonisomorphic formthat underlies perception.

This cross-digit interaction may also be the basis for the RFchanges observed in cortical plasticity studies (Wang et al.1995). Conditioning through a coincident presentation of stim-uli on adjacent digits could strengthen the lateral excitatoryinputs from adjacent digits. This long-term potentiation willlead to an increase in the drive from lateral inputs and cause theneuron to fire to any stimulus on the adjacent digit, causing thelinear component of the RF to spread to adjacent digits.

We conclude that nonlinear mechanisms play a major role inshaping somatosensory neural responses as early as area 3b.The nonlinear effects function to shape the RFs to be selectiveto local spatial variations in form and give neurons in area 3bselectivity to local simple features such as orientation andmotion. This underlies the perception of local spatial form andRFs that form the basis for roughness perception (Hsiao et al.2003).

APPENDIX

Matching Pursuit to Select Subset of Terms with SalientContribution to Neural Response

The analytical tool used in this paper to circumvent the problem of“dimensional explosion” due to the inclusion of quadratic terms is aniterative algorithm derived from matching pursuit (Mallat and Zhang1993). We assume that underlying firing rate (RTrue) of the neurongoverning the spiking response is related to the input (X) through alinear combination of some kernels B(X) � [B1(X), B2(X), . . . , Bq(X)].Specifically,

RTrue � �i�1

q

Bi�X�Wi (A1)

Given n observations in an experiment, RTrue as well as the kernelsBi(X) can be thought of as n-dimensional vectors [with jth entry of thekernel vector Bi(X) computed from the input stimulus pattern duringthe jth time step]. These kernels can be thought of as nonlinear featuremaps that transform the input into a feature space, in which neuronalresponse is linear (Wu and Gallant 2004). The nature of kernelsdepends upon the choice of basis set over which the user wishes toexpress the firing rate (RTrue). For example, if we want to estimate thelinear RF, then B(X) is simply the linear terms X spanning the distalpads. If we wish to express the function in terms of a Volterra series,then B(X) consists of various powers and products of powers of X. Themodel is known exactly if we identify the set of kernels B(X) andestimate the corresponding weights w � [w1, w2, . . . , wq]. In general,this is a hard problem since we usually have an infinite set of possiblekernels and a finite amount of data. The proposed algorithm circum-vents this problem by building the model in steps from a constant term.At every iteration, the algorithm picks up and adds to B(X) a kernel vectorfrom an overcomplete dictionary D that maximizes the projection of theobserved spike train. At the end of each iteration, the remainingkernel vectors in D are orthogonalized with respect to the selectedkernel. This ensures that the dictionary always spans a spaceorthogonal to the space spanned by the selected kernels in B, andhence always picks up that part of the spike response not alreadyaccounted for by the current running model. The iterations areterminated when the maximum projection at a given step is notstatistically significant.

Note 1. The idea of orthogonalizing the entire dictionary D withrespect to the selected kernel is a daunting proposition. However, itcan be achieved by projecting the observation vector R into theorthogonal space instead (see algorithm below for details).

Algorithm. The spike trains collected in the experiments are con-verted into an array of spike count vectors (observation vector) bybinning at a suitable bin interval. Let R (n � 1) be the “current”

259MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 18: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

observation vector (given n observations) and D � {D1, D2, . . . , Dm}be the dictionary of all vectors that we wish to test the observedresponse against. Let N � ��D1�, �D2�, . . .,�Dm� be a vector holdingnorms of the dictionary elements. Let B be the set of selected vectorsand ws be the corresponding projections. (Note that this is not thesame as weights w in Eq. 1. They converge only when dictionaryelements are samples of Gaussian white noise, in which case themethod resembles spike-triggered averaging). The steps involved inthe algorithm are as follows:

1) Project R onto the dictionary elements and divide by the normvector to obtain the current projections.

pi � RTDi ⁄ Ni

2) Choose the vector (Ds) with maximum projection energy (ps2) as

the current best estimate. Obtain the component of Ds orthogonal tothe vectors in B (call it Ds

�) and include it in B (i.e., B � [B, Ds�]). This

ensures that all vectors in B are mutually orthogonal.3) Update the projection vector, observation vector, and norm

vector as follows:

wcurr � RTDs� ⁄ �Ds

��

ws � �ws, wcurrR � R � wcurrDs

� ⁄ �Ds��

Ni � �Ni2 � �Di

TDs��2 ⁄ �Ds

��2

During the update, the observation vector (R) is replaced by the compo-nent of R that is orthogonal to the selected vector Ds

�. The update in normvector ensures that at the next step we compute the projection of R on thecomponents of the dictionary elements orthogonal to Ds

�, without explic-itly computing the orthogonal component of each of the dictionaryelements.

4) If the current projection energy (ps2) is less than a predetermined

threshold (see below) then stop; otherwise, repeat the procedure bygoing back to step 1.

At the first iteration, we choose D0, a vector of ones, as the selectedvector. This ensures that the observation vector R and all the elementsof the dictionary have a mean of zero from the second iterationonwards and that the vectors that are selected by the algorithm capturethe variability in the response not already accounted for by the mean.This also enables us to test directly our new model against the nullmodel that the response can be explained simply by the mean firingrate.

Stop condition. At every step, the algorithm tries to find a basisfunction that best captures the response vector. If unchecked, thealgorithm will continue until it finds n linearly independent vectorsfrom the dictionary that span a space containing the responsevector R, thus providing an exact solution to the system ofequations. In theory, if R is a vector of true firing rates it will havevery small projections (zero in the absence of static nonlinearity)on those kernels in the dictionary that do not modulate theresponse. Hence, the algorithm will never pick up such vectors.However, the observation vector R consists of samples of a randomvariable and is noisy. It will have some nonzero projection onvectors that do not contribute to the response. The following is atheoretical framework to optimally terminate the algorithm anddetermines when the projection on a particular vector is notsignificantly better than chance.

The projection of the observation vector R on Di is p � DiTR⁄ �Di�. If

R is sufficiently big (large number of observations), successive stim-ulus presentations are random, and we assume a Poisson spike count,then by the central limit theorem p tends to Gaussian with mean� � Di

TRtrue⁄ �Di� and variance 2 � �j�1n dij

2rjtrue⁄ �Di�

2, where Rtrue �[r1, r2, . . . , rn]T is the underlying true rate vector. If Di is orthogonalto RTrue (Di has no role in the neuronal response), then � � 0 andp2/2 follows a �2-distribution with 1 degree of freedom.

Thus at the end of every iteration we can formulate a null hypoth-esis that the current estimate of rate vector R (obtained as thebest-fitting linear combination of the kernels selected up to the current

iteration passed through a static nonlinearity, i.e., R� � �B·wsT�) is

the best estimate of the firing rate underlying the observed neuralresponse. Under this null hypothesis, projection energy on any otherkernel (Di) in the dictionary is simply a sample of a zero-mean�2-distribution with variance 2 � �j�1

n dij2 rj

�⁄ �Di�2. Hence, to termi-

nate the algorithm at a particular level �, the maximum projectionenergy at any iteration can be compared to the cumulative value ofthis �2-distribution at a level �. A cutoff of � � 0.01 was used in thisstudy.

Estimating weights. Once a suitable set of basis vectors consistingof linear and nonlinear kernels B(X) is identified, weights w areobtained by fitting a GLM with a log-link function. Usually, thenumber of kernels picked up by the algorithm is relatively small.Hence, fitting a GLM is tractable. Note that a GLM fit yields a modelthat is log-linear in the selected terms, which differs from the originallinear framework between firing rate and kernels used to iterativelyselect the subset of salient terms (Eq. A1). Alternatively, a linearleast-squares method could be used to fit the observed spike countvector and selected kernels with explicit static nonlinearity, whichmay maintain the linear relationship between the selected terms andneuronal firing rate but may be suboptimal given the Poisson nature ofspike counts.

Interpreting Quadratic Filters

The interpretation of second-order interaction implied by a givenquadratic filter can be understood by considering the following simpleexample. Assume that we have a 3-pixel or 3-term system given byX � [X1, X2, . . . , X3]T. Let us assume that the interaction betweenthese 3 pixels is given by the following Hessian matrix:

H � �0 0 1

0 0 0

1 0 0�

This Hessian implies that there is a positive or facilitative interactionbetween pixels X1 and X3. An eigenvector/eigenvalue decompositionof this matrix gives two eigenvectors with nonzero eigenvalues,namely,

1. E1 � � 1

�2, 0, �

1

�2 T

�with eigenvalue � 1�

2. E2 � � 1

�2, 0,

1

�2 T

�with eigenvalue 1�

Thus a facilitative quadratic interaction between two regions canmanifest in two ways in the quadratic filter. If the eigenvalue isnegative, then the two regions in the filter have opposite signs. If theeigenvalue is positive, then the two regions have identical signs.

Similarly, the Hessian

H � �0 0 �1

0 0 0

�1 0 0�

which implies a suppressive interaction between pixels X1 and X3, hasthe following two eigenvalues:

1. E1 � � 1

�2, 0,

1

�2 T

�with eigenvalue � 1�

260 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 19: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

2. E2 � � 1

�2, 0, �

1

�2 T

�with eigenvalue 1�

Thus if two regions have suppressive quadratic interaction, they haveidentical signs if the corresponding eigenvalue is negative and oppo-site signs if the eigenvalue is positive.

ACKNOWLEDGMENTS

We thank Bill Nash and Bill Quinlan for designing the drum stimulus. Wethank Justin Killebrew, Eric T. Carlson, Frank Dammann, Zhicheng Lai, andSung Soo Kim for their help in setting up the experiments and analysis.

GRANTS

This work was supported by National Institute of Neurological Disordersand Stroke Grant NS-34086.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the author(s).

AUTHOR CONTRIBUTIONS

Author contributions: P.H.T., P.J.F., and S.S.H. conception and design ofresearch; P.H.T., P.J.F., and S.S.H. performed experiments; P.H.T. analyzeddata; P.H.T., P.J.F., and S.S.H. interpreted results of experiments; P.H.T. andP.J.F. prepared figures; P.H.T. drafted manuscript; P.H.T., P.J.F., and S.S.H.edited and revised manuscript; P.H.T. and S.S.H. approved final version ofmanuscript.

REFERENCES

Allman J, Miezin F, McGuiness E. Stimulus-specific responses from beyondthe classical receptive field: neurophysiological mechanisms for local-globalcomparisons on visual neurons. Annu Rev Neurosci 8: 407–430, 1985.

Alloway KD, Rosenthal P, Burton H. Quantitative measurements of recep-tive field changes during antagonism of GABAergic transmission in primarysomatosensory cortex in cat. Exp Brain Res 78: 514–532, 1989.

Alloway KD, Burton H. Differential effects of GABA and bicuculline onrapidly- and slowly-adapting neurons in primary somatosensory cortex ofprimates. Exp Brain Res 85: 598–610, 1991.

Andersson SA. Intracellular postsynaptic potentials in the somatosensorycortex of the cat. Nature 205: 297–298, 1965.

Bankman IN, Johnson KO, Hsiao SS. Neural image transformation in thesomatosensory system of the monkey: comparison of neurophysiologicalobservations with responses in a neural network model. Cold Spring HarbSymp Quant Biol 55: 611–620, 1990.

Bensmaia SJ, Denchev PV, Dammann JF 3rd, Craig JC, Hsiao SS. Therepresentation of stimulus orientation in the early stages of somatosensoryprocessing. J Neurosci 28: 776–786, 2008.

Chen LM, Friedman RM, Ramsden BM, LaMotte RH, Roe AW. Fine-scale organization of SI (area 3b) in the squirrel monkey revealed withintrinsic optical imaging. J Neurophysiol 86: 3011–2029, 2001.

Chen LM, Friedman RM, Roe AW. Optical imaging of SI topography inanesthetized and awake squirrel monkeys. J Neurosci 25: 7648–7659, 2005.

DiCarlo JJ, Lane JW, Hsiao SS, Johnson KO. Marking microelectrodepenetrations with fluorescent dyes. J Neurosci Methods 64: 75–81, 1996.

DiCarlo JJ, Johnson KO, Hsiao SS. Structure of receptive fields in area 3bof primary somatosensory cortex in the alert monkey. J Neurosci 18:2626–2645, 1998.

DiCarlo JJ, Johnson KO. Velocity invariance of receptive field structure insomatosensory cortical area 3b of the alert monkey. J Neurosci 19: 401–419, 1999.

DiCarlo JJ, Johnson KO. Spatial and temporal structure of receptive fields inprimate somatosensory area 3b: effects of stimulus scanning direction andorientation. J Neurosci 20: 495–510, 2000.

DiCarlo JJ, Johnson KO. Receptive field structure in cortical area 3b of thealert monkey. Behav Brain Res 135: 167–178, 2002.

Dykes RW, Landry P, Metherate RS, Hicks TP. Functional role of GABAin cat primary somatosensory cortex: shaping receptive fields of corticalneurons. J Neurophysiol 52: 1066–1093, 1984.

Fitzpatrick D. Seeing beyond the receptive field in primary visual cortex.Curr Opin Neurobiol 10: 438–443, 2000.

Gardner EP. Cortical neuronal mechanisms underlying the perception ofmotion across the skin. In: Somatosensory Mechanisms, edited by von EulerC, Franzen O, Lindblom U, Ottoson D. London: Macmillan, 1984, p.93–112.

Gardner EP, Costanzo RM. Spatial integration of multiple-point stimuli inprimary somatosensory cortical receptive fields of alert monkeys. J Neuro-physiol 43: 420–443, 1980a.

Gardner EP, Costanzo RM. Temporal integration of multiple-point stimuli inprimary somatosensory cortical receptive fields of alert monkeys. J Neuro-physiol 43: 444–468, 1980b.

Gilbert CD, Wiesel TN. The influence of contextual stimuli on the orientationselectivity of cells in primary visual cortex of the cat. Vision Res 30:1689–1701, 1990.

Hsiao SS, Johnson KO, Yoshioka T. Processing of tactile information in theprimate brain. In: Comprehensive Handbook of Psychology, Volume 3:Biological Psychology, edited by Gallagher M, Nelson RJ. New York:Wiley, 2003, p. 211–236.

Hsiao SS, Fitzgerald PJ, Thakur PH, Denchev P, Yoshioka T. Receptivefields of somatosensory neurons. New Encyclopedia of Neuroscience. Am-sterdam: Elsevier, 2006.

Hyvarinen J, Poranen A. Movement-sensitive and direction and orientation-selective cutaneous receptive fields in the hand area of the post-central gyrusin monkeys. J Physiol 283: 523–537, 1978.

Innocenti GM, Manzoni T. Response patterns of somatosensory corticalneurones to peripheral stimuli. An intracellular study. Arch Ital Biol 110:322–347, 1972.

Iwamura Y, Tanaka M, Sakamoto M, Hikosaka O. Functional subdivisionsrepresenting different finger regions in area 3 of the first somatosensorycortex of the conscious monkey. Exp Brain Res 51: 315–326, 1983.

Johnson KO, Phillips JR. A rotating drum stimulator for scanning embossedpatterns and textures across the skin. J Neurosci Methods 22: 221–231,1988.

Johnson KO, Hsiao SS, Twombly IA. Neural mechanisms of tactile formrecognition. In: The Cognitive Sciences, edited by Gazzaniga M. Cambridge,MA: MIT Press, 1995, p. 253–268.

Jones EG. Connectivity of the primate sensory-motor cortex. In: CerebralCortex, edited by Jones EG, Peters A. New York: Plenum, 1986, vol. 5, p.113–183.

Jones EG, Coulter JD, Hendry SH. Intracortical connectivity of architectonicfields in the somatic sensory, motor and parietal cortex of monkeys. J CompNeurol 181: 291–347, 1978.

Jones EG, Friedman DP. Projection pattern of functional components ofthalamic ventrobasal complex on monkey somatosensory cortex. J Neuro-physiol 48: 521–544, 1982.

Kass RE, Ventura V, Brown EN. Statistical issues in the analysis of neuronaldata. J Neurophysiol 94: 8–25, 2005.

Krieger G, Zetzsche C. Nonlinear image operators for the evaluation of localintrinsic dimensionality. IEEE Trans Image Process 5: 1026–1041, 1996.

Laskin SE, Spencer WA. Cutaneous masking. II. Geometry of excitatory andinhibitory receptive fields of single units in somatosensory cortex of the cat.J Neurophysiol 42: 1061–1082, 1979.

Mallat S, Zhang Z. Matching pursuit with time frequency dictionaries. IEEETrans Signal Process 41: 3397–3415, 1993.

Mountcastle VB. Central nervous system mechanisms in mechanoreceptivesensibility. In: Handbook of Physiology. The Nervous System. Sensory Pro-cesses. Bethesda, MD: Am Physiol Soc, 1984, sect. 1, vol. III, pt. 2, p. 789–878.

Mountcastle VB. The Sensory Hand. Neural Mechanisms of Somatic Sensa-tion. Cambridge, MA: Harvard Univ. Press, 2005.

Mountcastle VB, Powell TP. Neural mechanisms subserving cutaneous sensibil-ity, with special reference to the role of afferent inhibition in sensory perceptionand discrimination. Bull Johns Hopkins Hosp 105: 201–232, 1959.

Mountcastle VB, Reitboeck HJ, Poggio GF, Steinmetz MA. Adaptation ofthe Reitboeck method of multiple microelectrode recording to the neocortexof the waking monkey. J Neurosci Methods 36: 77–84, 1991.

Paninski L, Pillow JW, Simoncelli EP. Maximum likelihood estimation of astochastic integrate-and-fire neural encoding model. Neural Comp 16: 2533–2561, 2004.

Phillips JR, Johnson KO, Hsiao SS. Spatial pattern representation andtransformation in monkey somatosensory cortex. Proc Natl Acad Sci USA85: 1317–1321, 1988.

Pillow JW, Simoncelli EP. Biases in white noise analysis due to non-Poissonspike generation. Neurocomputing 52: 109–115, 2003.

261MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from

Page 20: Second-order receptive fields reveal multidigit interactions in ... - … · 2016-11-03 · Second-order receptive fields reveal multidigit interactions in area 3b of the macaque

Poggio GF, Mountcastle VB. A study of the functional contributions of thelemniscal and spinothalamic systems to somatic sensibility. Bull JohnsHopkins Hosp 108: 266–316, 1960.

Pubols LM, Leroy RF. Orientation detectors in the primary somatosensoryneocortex of the raccoon. Brain Res 129: 61–74, 1977.

Reed JL, Qi H, Zhou Z, Burish MJ, Bernard MR, Kajikawa Y, Pouget P,Bonds AB, Kaas JH. Multielectrode recordings of neurons in primarysomatosensory cortex of owl monkeys during skin indentation with dualprobes inside and outside the classical receptive field (Abstract). SocNeurosci Abstr 2006: 53.20/O12, 2006.

Shanks MF, Pearson RC, Powell TP. The ipsilateral cortico-cortical con-nexions between the cytoarchitectonic subdivisions of the primary somaticsensory cortex in the monkey. Brain Res 356: 67–88, 1985.

Sillito AM, Grieve KL, Jones HE, Cudeiro J, Davis J. Visual corticalmechanisms detecting local focal orientation discontinuities. Nature 378:492–496, 1995.

Sripati AP, Yoshioka T, Denchev P, Hsiao SS, Johnson KO. Spatiotemporalreceptive fields of peripheral afferents and cortical area 3b and 1 neurons inthe primate somatosensory system. J Neurosci 26: 2101–2114, 2006.

Sur M. Receptive fields of neurons in areas 3b and 1 of somatosensory cortexin monkeys. Brain Res 198: 465–471, 1980.

Sur M, Wall JT, Kaas JH. Modular distribution of neurons with slowlyadapting and rapidly adapting responses in area 3b of somatosensory cortexin monkeys. J Neurophysiol 51: 724–744, 1984.

Truccolo W, Eden UT, Fellows MR, Donoghue JP, Brown EN. A pointprocess framework for relating neural spiking activity to spiking history,

neural ensemble, and extrinsic covariate effects. J Neurophysiol 93: 1074–1089, 2005.

Truccolo W, Hochberg LR, Donoghue JP. Collective dynamics in humanand monkey sensorimotor cortex: predicting single neuron spikes. NatNeurosci 13: 105–111, 2010.

Vega-Bermudez F, Johnson KO. Surround suppression in the responses ofprimate SA1 and RA mechanoreceptive afferents mapped with a probearray. J Neurophysiol 81: 2711–2719, 1999.

Ventura V, Carta R, Kass RE, Gettner SN, Olson CR. Statistical analysisof temporal evolution in single-neuron firing rates. Biostatistics 3: 1–20,2002.

Wang X, Merzenich MM, Sameshima K and Jenkins WM. Remodelling ofhand representation in adult cortex determined by timing of tactile stimu-lation. Nature 378: 71–75, 1995.

Warren S, Hamalainen HA, Gardner EP. Objective classification of motion-and direction-sensitive neurons in primary somatosensory cortex of awakemonkeys. J Neurophysiol 56: 598–622, 1986.

Whitehorn D, Towe AL. Postsynaptic potential patterns evoked upon cells insensorimotor cortex of cat by stimulation at the periphery. Exp Neurol 22:222–242, 1968.

Wilson HR, Richards WA. Curvature and separation discrimination at textureboundaries. J Opt Soc Am A 9: 1653–1662, 1992.

Wu MCK, Gallant JL. What’s beyond second order? Kernel regressiontechniques for nonlinear functional characterization of visual neurons (Ab-stract). Soc Neurosci Abstr, 2004.

262 MULTIDIGIT QUADRATIC RECEPTIVE FIELDS OF AREA 3b NEURONS

J Neurophysiol • doi:10.1152/jn.01022.2010 • www.jn.org

by 10.220.33.5 on Novem

ber 2, 2016http://jn.physiology.org/

Dow

nloaded from