nips nips2003 nips2003-7 knowledge-graph by maker-knowledge-mining

7 nips-2003-A Functional Architecture for Motion Pattern Processing in MSTd


Source: pdf

Author: Scott A. Beardsley, Lucia M. Vaina

Abstract: Psychophysical studies suggest the existence of specialized detectors for component motion patterns (radial, circular, and spiral), that are consistent with the visual motion properties of cells in the dorsal medial superior temporal area (MSTd) of non-human primates. Here we use a biologically constrained model of visual motion processing in MSTd, in conjunction with psychophysical performance on two motion pattern tasks, to elucidate the computational mechanisms associated with the processing of widefield motion patterns encountered during self-motion. In both tasks discrimination thresholds varied significantly with the type of motion pattern presented, suggesting perceptual correlates to the preferred motion bias reported in MSTd. Through the model we demonstrate that while independently responding motion pattern units are capable of encoding information relevant to the visual motion tasks, equivalent psychophysical performance can only be achieved using interconnected neural populations that systematically inhibit non-responsive units. These results suggest the cyclic trends in psychophysical performance may be mediated, in part, by recurrent connections within motion pattern responsive areas whose structure is a function of the similarity in preferred motion patterns and receptive field locations between units. 1 In trod u ction A major challenge in computational neuroscience is to elucidate the architecture of the cortical circuits for sensory processing and their effective role in mediating behavior. In the visual motion system, biologically constrained models are playing an increasingly important role in this endeavor by providing an explanatory substrate linking perceptual performance and the visual properties of single cells. Single cell studies indicate the presence of complex interconnected structures in middle temporal and primary visual cortex whose most basic horizontal connections can impart considerable computational power to the underlying neural population [1, 2]. Combined psychophysical and computational studies support these findings Figure 1: a) Schematic of the graded motion pattern (GMP) task. Discrimination pairs of stimuli were created by perturbing the flow angle (φ) of each 'test' motion (with average dot speed, vav), by ±φp in the stimulus space spanned by radial and circular motions. b) Schematic of the shifted center-of-motion (COM) task. Discrimination pairs of stimuli were created by shifting the COM of the ‘test’ motion to the left and right of a central fixation point. For each motion pattern the COM was shifted within the illusory inner aperture and was never explicitly visible. and suggest that recurrent connections may play a significant role in encoding the visual motion properties associated with various psychophysical tasks [3, 4]. Using this methodology our goal is to elucidate the computational mechanisms associated with the processing of wide-field motion patterns encountered during self-motion. In the human visual motion system, psychophysical studies suggest the existence of specialized detectors for the motion pattern components (i.e., radial, circular and spiral motions) associated with self-motion [5, 6]. Neurophysiological studies reporting neurons sensitive to motion patterns in the dorsal medial superior temporal area (MSTd) support the existence of such mechanisms [7-10], and in conjunction with psychophysical studies suggest a strong link between the patterns of neural activity and motion-based perceptual performance [11, 12]. Through the combination of human psychophysical performance and biologically constrained modeling we investigate the computational role of simple recurrent connections within a population of MSTd-like units. Based on the known visual motion properties within MSTd we ask what neural structures are computationally sufficient to encode psychophysical performance on a series of motion pattern tasks. 2 M o t i o n pa t t e r n d i sc r i m i n a t i o n Using motion pattern stimuli consistent with previous studies [5, 6], we have developed a set of novel psychophysical tasks designed to facilitate a more direct comparison between human perceptual performance and the visual motion properties of cells in MSTd that have been found to underlie the discrimination of motion patterns [11, 12]. The psychophysical tasks, referred to as the graded motion pattern (GMP) and shifted center-of-motion (COM) tasks, are outlined in Fig. 1. Using a temporal two-alternative-forced-choice task we measured discrimination thresholds to global changes in the patterns of complex motion (GMP task), [13], and shifts in the center-of-motion (COM task). Stimuli were presented with central fixation using a constant stimulus paradigm and consisted of dynamic random dot displays presented in a 24o annular region (central 4o removed). In each task, the stimulus duration was randomly perturbed across presentations (440±40 msec) to control for timing-based cues, and dots moved coherently through a radial speed Figure 2: a) GMP thresholds across 8 'test' motions at two mean dot speeds for two observers. Performance varied continuously with thresholds for radial motions (φ=0, 180o) significantly lower than those for circular motions (φ=90,270o), (p<0.001; t(37)=3.39). b) COM thresholds at three mean dot speeds for two observers. As with the GMP task, performance varied continuously with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47). gradient in directions consistent with the global motion pattern presented. Discrimination thresholds were obtained across eight ‘test’ motions corresponding to expansion, contraction, CW and CCW rotation, and the four intermediate spiral motions. To minimize adaptation to specific motion patterns, opposing motions (e.g., expansion/ contraction) were interleaved across paired presentations. 2.1 Results Discrimination thresholds are reported here from a subset of the observer population consisting of three experienced psychophysical observers, one of which was naïve to the purpose of the psychophysical tasks. For each condition, performance is reported as the mean and standard error averaged across 8-12 thresholds. Across observers and dot speeds GMP thresholds followed a distinct trend in the stimulus space [13], with radial motions (expansion/contraction) significantly lower than circular motions (CW/CCW rotation), (p<0.001; t(37)=3.39), (Fig. 2a). While thresholds for the intermediate spiral motions were not significantly different from circular motions (p=0.223, t(60)=0.74), the trends across 'test' motions were well fit within the stimulus space (SB: r>0.82, SC: r>0.77) by sinusoids whose period and phase were 196 ± 10o and -72 ± 20o respectively (Fig. 1a). When the radial speed gradient was removed by randomizing the spatial distribution of dot speeds, threshold performance increased significantly across observers (p<0.05; t(17)=1.91), particularly for circular motions (p<0.005; t(25)=3.31), (data not shown). Such performance suggests a perceptual contribution associated with the presence of the speed gradient and is particularly interesting given the fact that the speed gradient did not contribute computationally relevant information to the task. However, the speed gradient did convey information regarding the integrative structure of the global motion field and as such suggests a preference of the underlying motion mechanisms for spatially structured speed information. Similar trends in performance were observed in the COM task across observers and dot speeds. Discrimination thresholds varied continuously as a function of the 'test' motion with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47) and could be well fit by a sinusoidal trend line (e.g. SB at 3 deg/s: r>0.91, period = 178 ± 10 o and phase = -70 ± 25o), (Fig. 2b). 2.2 A local or global task? The consistency of the cyclic threshold profile in stimuli that restricted the temporal integration of individual dot motions [13], and simultaneously contained all directions of motion, generally argues against a primary role for local motion mechanisms in the psychophysical tasks. While the psychophysical literature has reported a wide variety of “local” motion direction anisotropies whose properties are reminiscent of the results observed here, e.g. [14], all would predict equivalent thresholds for radial and circular motions for a set of uniformly distributed and/or spatially restricted motion direction mechanisms. Together with the computational impact of the speed gradient and psychophysical studies supporting the existence of wide-field motion pattern mechanisms [5, 6], these results suggest that the threshold differences across the GMP and COM tasks may be associated with variations in the computational properties across a series of specialized motion pattern mechanisms. 3 A computational model The similarities between the motion pattern stimuli used to quantify human perception and the visual motion properties of cells in MSTd suggests that MSTd may play a computational role in the psychophysical tasks. To examine this hypothesis, we constructed a population of MSTd-like units whose visual motion properties were consistent with the reported neurophysiology (see [13] for details). Across the population, the distribution of receptive field centers was uniform across polar angle and followed a gamma distribution Γ(5,6) across eccenticity [7]. For each unit, visual motion responses followed a gaussian tuning profile as a function of the stimulus flow angle G( φ), (σi=60±30o; [10]), and the distance of the stimulus COM from the unit’s receptive field center Gsat(xi, yi, σs=19o), Eq. 1, such that its preferred motion response was position invariant to small shifts in the COM [10] and degraded continuously for large shifts [9]. Within the model, simulations were categorized according to the distribution of preferred motions represented across the population (one reported in MSTd and a uniform control). The first distribution simulated an expansion bias in which the density of preferred motions decreased symmetrically from expansions to contraction [10]. The second distribution simulated a uniform preference for all motions and was used as a control to quantify the effects of an expansion bias on psychophysical performance. Throughout the paper we refer to simulations containing these distributions as ‘Expansion-biased’ and ‘Uniform’ respectively. 3.1 Extracting perceptual estimates from the neural code For each stimulus presentation, the ith unit’s response was calculated as the average firing rate, Ri, from the product of its motion pattern and spatial tuning profiles, ( ) Ri = Rmax G min[φ − φi ] ,σ ti G sati (x− xi , y − y i ,σ s ) + P (λ = 12 ) (1) where Rmax is the maximum preferred stimulus response (spikes/s), min[ ] refers to the minimum angular distance between the stimulus flow angle φ and the unit’s preferred motion φi, Gsat is the unit’s spatial tuning profile saturated within the central 5±3o, σti and σs are the standard deviations of the unit’s motion pattern and Figure 3: Model vs. psychophysical performance for independently responding units. Model thresholds are reported as the average (±1 S.E.) across five simulated populations. a) GMP thresholds were highest for contracting motions and lowest for expanding motions across all Expansion-biased populations. b) Comparable trends in performance were observed for COM thresholds. Comparison with the Uniform control simulations in both tasks (2000 units shown here) indicates that thresholds closely followed the distribution of preferred motions simulated within the model. spatial tuning profiles respectively, (xi,yi) is the spatial location of the unit’s receptive field center, (x,y) is the spatial location of the stimulus COM, and P(λ=12) is the background activity simulated as an uncorrelated Poisson process. The psychophysical tasks were simulated using a modified center-of-gravity ^ approach to decode estimates of the stimulus properties, i.e. flow angle (φ ) and ˆ ˆ COM location in the visual field (x, y ) , from the neural population   ∑ xi Ri ∑ y i Ri v  , i , ∑ φ i Ri  ∑ Ri i   i i   i (xˆ, yˆ , φˆ) =  i∑ R  (2) v where φi is the unit vector in the stimulus space (Fig. 1a) corresponding to the unit’s preferred motion. For each set of paired stimuli, psychophysical judgments were made by comparing the estimated stimulus properties according to the discrimination criteria, specified in the psychophysical tasks. As with the psychophysical experiments, discrimination thresholds were computed using a leastsquares fit to percent correct performance across constant stimulus levels. 3.2 Simulation 1: Independent neural responses In the first series of simulations, GMP and COM thresholds were quantified across three populations (500, 1000, and 2000 units) of independently responding units for each simulated distribution (Expansion-biased and Uniform). Across simulations, both the range in thresholds and their trends across ‘test’ motions were compared with human psychophysical performance to quantify the effects of population size and an expansion biased preferred motion distribution on model performance. Over the psychophysical range of interest (φp ± 7o), GMP thresholds for contracting motions were at chance across all Expansion-biased populations, (Fig. 3a). While thresholds for expanding motions were generally consistent with those for human observers, those for circular motions remained significantly higher for all but the largest populations. Similar trends in performance were observed for the COM task, (Fig. 3b). Here the range of COM thresholds was well matched with human performance for simulations containing 1000 units, however, the trends across motion patterns remained inconsistent even for the largest populations. Figure 4: Proposed recurrent connection profile between motion pattern units. a) Across the motion pattern space connection strength followed an inverse gaussian profile such that the ith unit (with preferred motion φi) systematically inhibited units with anti-preferred motions centered at 180+φi. b) Across the visual field connection strength followed a difference-of-gaussians profile as a function of the relative distance between receptive field centers such that spatially local units are mutually excitatory (σRe=10o) and more distant units were mutually inhibitory (σRi=80o). For simulations containing a uniform distribution of preferred motions, the threshold range was consistent with human performance on both tasks, however, the trend across motion patterns was generally flat. What variability did occur was due primarily to the discrete sampling of preferred motions across the population. Comparison of the discrimination thresholds for the Expansion-biased and Uniform populations indicates that the trend across thresholds was closely matched to the underlying distributions of preferred motions. This result in due in part to the nearequal weighting of independently responding units and can be explained to a first approximation by the proportional increase in the signal-to-noise ratio across the population as a function of the density of units responsive to a given 'test' motion. 3.3 Simulation 2: An interconnected neural structure In a second series of simulations, we examined the computational effect of adding recurrent connections between units. If the distribution of preferred motions in MSTd is in fact biased towards expansions, as the neurophysiology suggests, it seems unlikely that independent estimates of the visual motion information would be sufficient to yield the threshold profiles observed in the psychophysical tasks. We hypothesize that a simple fixed architecture of excitatory and/or inhibitory connections is sufficient to account for the cyclic trends in discrimination thresholds. Specifically, we propose that a recurrent connection profile whose strength varies as a function of (a) the similarity between preferred motion patterns and (b) the distance between receptive field centers, is computationally sufficient to recover the trends in GMP/COM performance (Fig. 4), wij = S R e − ( xi − x j )2 + ( yi − y j )2 2 2σ R e − SR e 2 − −(min[ φi − φ j ])2 ( xi − x j )2 + ( yi − y j )2 2 2 σ Ri − Sφ e 2σ I2 (3) Figure 5: Model vs. psychophysical performance for populations containing recurrent connections (σI=80o). As the number of units increased for Expansionbiased populations, discrimination thresholds decreased to psychophysical levels and the sinusoidal trend in thresholds emerged for both the (a) GMP and (b) COM tasks. Sinusoidal trends were established for as few as 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were (193.8 ± 11.7o, -70.0 ± 22.6o) and (168.2 ± 13.7o, -118.8 ± 31.8o) for the GMP and COM tasks respectively. where wij is the strength of the recurrent connection between ith and jth units, (xi,yi) and (xj,yj) denote the spatial locations of their receptive field centers, σRe (=10o) and σRi (=80o) together define the spatial extent of a difference-of-gaussians interaction between receptive field centers, and SR and Sφ scale the connection strength. To examine the effects of the spread of motion pattern-specific inhibition and connection strength in the model, σI, Sφ, and SR were considered free parameters. Within the parameter space used to define recurrent connections (i.e., σI, Sφ and SR), Monte Carlo simulations of Expansion-biased model performance (1000 units) yielded regions of high correlation on both tasks (with respect to the psychophysical thresholds, r>0.7) that were consistent across independently simulated populations. Typically these regions were well defined over a broad range such that there was significant overlap between tasks (e.g., for the GMP task (SR=0.03), σI=[45,120o], Sφ=[0.03,0.3] and for the COM task (σI=80o), Sφ = [0.03,0.08], SR = [0.005, 0.04]). Fig. 5 shows averaged threshold performance for simulations of interconnected units drawn from the highly correlated regions of the (σI, Sφ, SR) parameter space. For populations not explicitly examined in the Monte Carlo simulations connection strengths (Sφ, SR) were scaled inversely with population size to maintain an equivalent level of recurrent activity. With the incorporation of recurrent connections, the sinusoidal trend in GMP and COM thresholds emerged for Expansion-biased populations as the number of units increased. In both tasks the cyclic threshold profiles were established for 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were consistent with human performance. Unlike the Expansion-biased populations, Uniform populations were not significantly affected by the presence of recurrent connections (Fig. 5). Both the range in thresholds and the flat trend across motion patterns were well matched to those in Section 3.2. Together these results suggest that the sinusoidal trends in GMP and COM performance may be mediated by the combined contribution of the recurrent interconnections and the bias in preferred motions across the population. 4 D i s c u s s i on Using a biologically constrained computational model in conjunction with human psychophysical performance on two motion pattern tasks we have shown that the visual motion information encoded across an interconnected population of cells responsive to motion patterns, such as those in MSTd, is computationally sufficient to extract perceptual estimates consistent with human performance. Specifically, we have shown that the cyclic trend in psychophysical performance observed across tasks, (a) cannot be reproduced using populations of independently responding units and (b) is dependent, in part, on the presence of an expanding motion bias in the distribution of preferred motions across the neural population. The model’s performance suggests the presence of specific recurrent structures within motion pattern responsive areas, such as MSTd, whose strength varies as a function of the similarity between preferred motion patterns and the distance between receptive field centers. While such structures have not been explicitly examined in MSTd and other higher visual motion areas there is anecdotal support for the presence of inhibitory connections [8]. Together, these results suggest that robust processing of the motion patterns associated with self-motion and optic flow may be mediated, in part, by recurrent structures in extrastriate visual motion areas whose distributions of preferred motions are biased strongly in favor of expanding motions. Acknowledgments This work was supported by National Institutes of Health grant EY-2R01-07861-13 to L.M.V. References [1] Malach, R., Schirman, T., Harel, M., Tootell, R., & Malonek, D., (1997), Cerebral Cortex, 7(4): 386-393. [2] Gilbert, C. D., (1992), Neuron, 9: 1-13. [3] Koechlin, E., Anton, J., & Burnod, Y., (1999), Biological Cybernetics, 80: 2544. [4] Stemmler, M., Usher, M., & Niebur, E., (1995), Science, 269: 1877-1880. [5] Burr, D. C., Morrone, M. C., & Vaina, L. M., (1998), Vision Research, 38(12): 1731-1743. [6] Meese, T. S. & Harris, S. J., (2002), Vision Research, 42: 1073-1080. [7] Tanaka, K. & Saito, H. A., (1989), Journal of Neurophysiology, 62(3): 626-641. [8] Duffy, C. J. & Wurtz, R. H., (1991), Journal of Neurophysiology, 65(6): 13461359. [9] Duffy, C. J. & Wurtz, R. H., (1995), Journal of Neuroscience, 15(7): 5192-5208. [10] Graziano, M. S., Anderson, R. A., & Snowden, R., (1994), Journal of Neuroscience, 14(1): 54-67. [11] Celebrini, S. & Newsome, W., (1994), Journal of Neuroscience, 14(7): 41094124. [12] Celebrini, S. & Newsome, W. T., (1995), Journal of Neurophysiology, 73(2): 437-448. [13] Beardsley, S. A. & Vaina, L. M., (2001), Journal of Computational Neuroscience, 10: 255-280. [14] Matthews, N. & Qian, N., (1999), Vision Research, 39: 2205-2211.

Reference: text


Summary: the most important sentenses genereted by tfidf model

sentIndex sentText sentNum sentScore

1 edu Abstract Psychophysical studies suggest the existence of specialized detectors for component motion patterns (radial, circular, and spiral), that are consistent with the visual motion properties of cells in the dorsal medial superior temporal area (MSTd) of non-human primates. [sent-7, score-1.256]

2 Here we use a biologically constrained model of visual motion processing in MSTd, in conjunction with psychophysical performance on two motion pattern tasks, to elucidate the computational mechanisms associated with the processing of widefield motion patterns encountered during self-motion. [sent-8, score-2.018]

3 In both tasks discrimination thresholds varied significantly with the type of motion pattern presented, suggesting perceptual correlates to the preferred motion bias reported in MSTd. [sent-9, score-1.782]

4 These results suggest the cyclic trends in psychophysical performance may be mediated, in part, by recurrent connections within motion pattern responsive areas whose structure is a function of the similarity in preferred motion patterns and receptive field locations between units. [sent-11, score-2.109]

5 1 In trod u ction A major challenge in computational neuroscience is to elucidate the architecture of the cortical circuits for sensory processing and their effective role in mediating behavior. [sent-12, score-0.087]

6 In the visual motion system, biologically constrained models are playing an increasingly important role in this endeavor by providing an explanatory substrate linking perceptual performance and the visual properties of single cells. [sent-13, score-0.794]

7 Single cell studies indicate the presence of complex interconnected structures in middle temporal and primary visual cortex whose most basic horizontal connections can impart considerable computational power to the underlying neural population [1, 2]. [sent-14, score-0.343]

8 Combined psychophysical and computational studies support these findings Figure 1: a) Schematic of the graded motion pattern (GMP) task. [sent-15, score-0.878]

9 Discrimination pairs of stimuli were created by perturbing the flow angle (φ) of each 'test' motion (with average dot speed, vav), by ±φp in the stimulus space spanned by radial and circular motions. [sent-16, score-0.931]

10 Discrimination pairs of stimuli were created by shifting the COM of the ‘test’ motion to the left and right of a central fixation point. [sent-18, score-0.525]

11 For each motion pattern the COM was shifted within the illusory inner aperture and was never explicitly visible. [sent-19, score-0.508]

12 and suggest that recurrent connections may play a significant role in encoding the visual motion properties associated with various psychophysical tasks [3, 4]. [sent-20, score-1.2]

13 Using this methodology our goal is to elucidate the computational mechanisms associated with the processing of wide-field motion patterns encountered during self-motion. [sent-21, score-0.589]

14 In the human visual motion system, psychophysical studies suggest the existence of specialized detectors for the motion pattern components (i. [sent-22, score-1.515]

15 , radial, circular and spiral motions) associated with self-motion [5, 6]. [sent-24, score-0.167]

16 Through the combination of human psychophysical performance and biologically constrained modeling we investigate the computational role of simple recurrent connections within a population of MSTd-like units. [sent-26, score-0.707]

17 Based on the known visual motion properties within MSTd we ask what neural structures are computationally sufficient to encode psychophysical performance on a series of motion pattern tasks. [sent-27, score-1.476]

18 The psychophysical tasks, referred to as the graded motion pattern (GMP) and shifted center-of-motion (COM) tasks, are outlined in Fig. [sent-29, score-0.843]

19 Using a temporal two-alternative-forced-choice task we measured discrimination thresholds to global changes in the patterns of complex motion (GMP task), [13], and shifts in the center-of-motion (COM task). [sent-31, score-0.872]

20 Stimuli were presented with central fixation using a constant stimulus paradigm and consisted of dynamic random dot displays presented in a 24o annular region (central 4o removed). [sent-32, score-0.177]

21 In each task, the stimulus duration was randomly perturbed across presentations (440±40 msec) to control for timing-based cues, and dots moved coherently through a radial speed Figure 2: a) GMP thresholds across 8 'test' motions at two mean dot speeds for two observers. [sent-33, score-1.142]

22 Performance varied continuously with thresholds for radial motions (φ=0, 180o) significantly lower than those for circular motions (φ=90,270o), (p<0. [sent-34, score-1.274]

23 b) COM thresholds at three mean dot speeds for two observers. [sent-37, score-0.347]

24 As with the GMP task, performance varied continuously with thresholds for radial motions significantly lower than those for circular motions, (p<0. [sent-38, score-0.972]

25 gradient in directions consistent with the global motion pattern presented. [sent-41, score-0.538]

26 Discrimination thresholds were obtained across eight ‘test’ motions corresponding to expansion, contraction, CW and CCW rotation, and the four intermediate spiral motions. [sent-42, score-0.757]

27 To minimize adaptation to specific motion patterns, opposing motions (e. [sent-43, score-0.796]

28 1 Results Discrimination thresholds are reported here from a subset of the observer population consisting of three experienced psychophysical observers, one of which was naïve to the purpose of the psychophysical tasks. [sent-47, score-0.961]

29 For each condition, performance is reported as the mean and standard error averaged across 8-12 thresholds. [sent-48, score-0.195]

30 Across observers and dot speeds GMP thresholds followed a distinct trend in the stimulus space [13], with radial motions (expansion/contraction) significantly lower than circular motions (CW/CCW rotation), (p<0. [sent-49, score-1.566]

31 While thresholds for the intermediate spiral motions were not significantly different from circular motions (p=0. [sent-53, score-1.183]

32 74), the trends across 'test' motions were well fit within the stimulus space (SB: r>0. [sent-55, score-0.698]

33 77) by sinusoids whose period and phase were 196 ± 10o and -72 ± 20o respectively (Fig. [sent-57, score-0.037]

34 When the radial speed gradient was removed by randomizing the spatial distribution of dot speeds, threshold performance increased significantly across observers (p<0. [sent-59, score-0.554]

35 Such performance suggests a perceptual contribution associated with the presence of the speed gradient and is particularly interesting given the fact that the speed gradient did not contribute computationally relevant information to the task. [sent-64, score-0.198]

36 However, the speed gradient did convey information regarding the integrative structure of the global motion field and as such suggests a preference of the underlying motion mechanisms for spatially structured speed information. [sent-65, score-1.157]

37 Similar trends in performance were observed in the COM task across observers and dot speeds. [sent-66, score-0.373]

38 Discrimination thresholds varied continuously as a function of the 'test' motion with thresholds for radial motions significantly lower than those for circular motions, (p<0. [sent-67, score-1.634]

39 47) and could be well fit by a sinusoidal trend line (e. [sent-69, score-0.189]

40 The consistency of the cyclic threshold profile in stimuli that restricted the temporal integration of individual dot motions [13], and simultaneously contained all directions of motion, generally argues against a primary role for local motion mechanisms in the psychophysical tasks. [sent-76, score-1.415]

41 While the psychophysical literature has reported a wide variety of “local” motion direction anisotropies whose properties are reminiscent of the results observed here, e. [sent-77, score-0.827]

42 [14], all would predict equivalent thresholds for radial and circular motions for a set of uniformly distributed and/or spatially restricted motion direction mechanisms. [sent-79, score-1.26]

43 3 A computational model The similarities between the motion pattern stimuli used to quantify human perception and the visual motion properties of cells in MSTd suggests that MSTd may play a computational role in the psychophysical tasks. [sent-81, score-1.493]

44 To examine this hypothesis, we constructed a population of MSTd-like units whose visual motion properties were consistent with the reported neurophysiology (see [13] for details). [sent-82, score-0.849]

45 Across the population, the distribution of receptive field centers was uniform across polar angle and followed a gamma distribution Γ(5,6) across eccenticity [7]. [sent-83, score-0.547]

46 For each unit, visual motion responses followed a gaussian tuning profile as a function of the stimulus flow angle G( φ), (σi=60±30o; [10]), and the distance of the stimulus COM from the unit’s receptive field center Gsat(xi, yi, σs=19o), Eq. [sent-84, score-1.093]

47 1, such that its preferred motion response was position invariant to small shifts in the COM [10] and degraded continuously for large shifts [9]. [sent-85, score-0.653]

48 Within the model, simulations were categorized according to the distribution of preferred motions represented across the population (one reported in MSTd and a uniform control). [sent-86, score-0.818]

49 The first distribution simulated an expansion bias in which the density of preferred motions decreased symmetrically from expansions to contraction [10]. [sent-87, score-0.636]

50 The second distribution simulated a uniform preference for all motions and was used as a control to quantify the effects of an expansion bias on psychophysical performance. [sent-88, score-0.78]

51 Throughout the paper we refer to simulations containing these distributions as ‘Expansion-biased’ and ‘Uniform’ respectively. [sent-89, score-0.06]

52 Model thresholds are reported as the average (±1 S. [sent-93, score-0.277]

53 a) GMP thresholds were highest for contracting motions and lowest for expanding motions across all Expansion-biased populations. [sent-96, score-1.11]

54 b) Comparable trends in performance were observed for COM thresholds. [sent-97, score-0.13]

55 Comparison with the Uniform control simulations in both tasks (2000 units shown here) indicates that thresholds closely followed the distribution of preferred motions simulated within the model. [sent-98, score-1.066]

56 spatial tuning profiles respectively, (xi,yi) is the spatial location of the unit’s receptive field center, (x,y) is the spatial location of the stimulus COM, and P(λ=12) is the background activity simulated as an uncorrelated Poisson process. [sent-99, score-0.477]

57 The psychophysical tasks were simulated using a modified center-of-gravity ^ approach to decode estimates of the stimulus properties, i. [sent-100, score-0.517]

58 flow angle (φ ) and ˆ ˆ COM location in the visual field (x, y ) , from the neural population   ∑ xi Ri ∑ y i Ri v  , i , ∑ φ i Ri  ∑ Ri i   i i   i (xˆ, yˆ , φˆ) =  i∑ R  (2) v where φi is the unit vector in the stimulus space (Fig. [sent-102, score-0.454]

59 For each set of paired stimuli, psychophysical judgments were made by comparing the estimated stimulus properties according to the discrimination criteria, specified in the psychophysical tasks. [sent-104, score-0.846]

60 As with the psychophysical experiments, discrimination thresholds were computed using a leastsquares fit to percent correct performance across constant stimulus levels. [sent-105, score-0.968]

61 2 Simulation 1: Independent neural responses In the first series of simulations, GMP and COM thresholds were quantified across three populations (500, 1000, and 2000 units) of independently responding units for each simulated distribution (Expansion-biased and Uniform). [sent-107, score-0.71]

62 Across simulations, both the range in thresholds and their trends across ‘test’ motions were compared with human psychophysical performance to quantify the effects of population size and an expansion biased preferred motion distribution on model performance. [sent-108, score-1.895]

63 Over the psychophysical range of interest (φp ± 7o), GMP thresholds for contracting motions were at chance across all Expansion-biased populations, (Fig. [sent-109, score-1.05]

64 While thresholds for expanding motions were generally consistent with those for human observers, those for circular motions remained significantly higher for all but the largest populations. [sent-111, score-1.235]

65 Similar trends in performance were observed for the COM task, (Fig. [sent-112, score-0.13]

66 Here the range of COM thresholds was well matched with human performance for simulations containing 1000 units, however, the trends across motion patterns remained inconsistent even for the largest populations. [sent-114, score-1.142]

67 Figure 4: Proposed recurrent connection profile between motion pattern units. [sent-115, score-0.752]

68 a) Across the motion pattern space connection strength followed an inverse gaussian profile such that the ith unit (with preferred motion φi) systematically inhibited units with anti-preferred motions centered at 180+φi. [sent-116, score-1.794]

69 b) Across the visual field connection strength followed a difference-of-gaussians profile as a function of the relative distance between receptive field centers such that spatially local units are mutually excitatory (σRe=10o) and more distant units were mutually inhibitory (σRi=80o). [sent-117, score-0.811]

70 For simulations containing a uniform distribution of preferred motions, the threshold range was consistent with human performance on both tasks, however, the trend across motion patterns was generally flat. [sent-118, score-1.112]

71 What variability did occur was due primarily to the discrete sampling of preferred motions across the population. [sent-119, score-0.627]

72 Comparison of the discrimination thresholds for the Expansion-biased and Uniform populations indicates that the trend across thresholds was closely matched to the underlying distributions of preferred motions. [sent-120, score-1.099]

73 This result in due in part to the nearequal weighting of independently responding units and can be explained to a first approximation by the proportional increase in the signal-to-noise ratio across the population as a function of the density of units responsive to a given 'test' motion. [sent-121, score-0.539]

74 3 Simulation 2: An interconnected neural structure In a second series of simulations, we examined the computational effect of adding recurrent connections between units. [sent-123, score-0.262]

75 If the distribution of preferred motions in MSTd is in fact biased towards expansions, as the neurophysiology suggests, it seems unlikely that independent estimates of the visual motion information would be sufficient to yield the threshold profiles observed in the psychophysical tasks. [sent-124, score-1.522]

76 We hypothesize that a simple fixed architecture of excitatory and/or inhibitory connections is sufficient to account for the cyclic trends in discrimination thresholds. [sent-125, score-0.364]

77 Specifically, we propose that a recurrent connection profile whose strength varies as a function of (a) the similarity between preferred motion patterns and (b) the distance between receptive field centers, is computationally sufficient to recover the trends in GMP/COM performance (Fig. [sent-126, score-1.304]

78 As the number of units increased for Expansionbiased populations, discrimination thresholds decreased to psychophysical levels and the sinusoidal trend in thresholds emerged for both the (a) GMP and (b) COM tasks. [sent-129, score-1.179]

79 Sinusoidal trends were established for as few as 1000 units and were well fit (r>0. [sent-130, score-0.258]

80 To examine the effects of the spread of motion pattern-specific inhibition and connection strength in the model, σI, Sφ, and SR were considered free parameters. [sent-141, score-0.531]

81 Within the parameter space used to define recurrent connections (i. [sent-142, score-0.219]

82 , σI, Sφ and SR), Monte Carlo simulations of Expansion-biased model performance (1000 units) yielded regions of high correlation on both tasks (with respect to the psychophysical thresholds, r>0. [sent-144, score-0.484]

83 Typically these regions were well defined over a broad range such that there was significant overlap between tasks (e. [sent-146, score-0.075]

84 5 shows averaged threshold performance for simulations of interconnected units drawn from the highly correlated regions of the (σI, Sφ, SR) parameter space. [sent-157, score-0.301]

85 For populations not explicitly examined in the Monte Carlo simulations connection strengths (Sφ, SR) were scaled inversely with population size to maintain an equivalent level of recurrent activity. [sent-158, score-0.403]

86 With the incorporation of recurrent connections, the sinusoidal trend in GMP and COM thresholds emerged for Expansion-biased populations as the number of units increased. [sent-159, score-0.753]

87 In both tasks the cyclic threshold profiles were established for 1000 units and were well fit (r>0. [sent-160, score-0.35]

88 9) by sinusoids whose periods and phases were consistent with human performance. [sent-161, score-0.107]

89 Unlike the Expansion-biased populations, Uniform populations were not significantly affected by the presence of recurrent connections (Fig. [sent-162, score-0.427]

90 Both the range in thresholds and the flat trend across motion patterns were well matched to those in Section 3. [sent-164, score-0.99]

91 Together these results suggest that the sinusoidal trends in GMP and COM performance may be mediated by the combined contribution of the recurrent interconnections and the bias in preferred motions across the population. [sent-166, score-1.021]

92 The model’s performance suggests the presence of specific recurrent structures within motion pattern responsive areas, such as MSTd, whose strength varies as a function of the similarity between preferred motion patterns and the distance between receptive field centers. [sent-169, score-1.63]

93 While such structures have not been explicitly examined in MSTd and other higher visual motion areas there is anecdotal support for the presence of inhibitory connections [8]. [sent-170, score-0.635]

94 Together, these results suggest that robust processing of the motion patterns associated with self-motion and optic flow may be mediated, in part, by recurrent structures in extrastriate visual motion areas whose distributions of preferred motions are biased strongly in favor of expanding motions. [sent-171, score-1.801]


similar papers computed by tfidf model

tfidf for this paper:

wordName wordTfidf (topN-words)

[('motion', 0.456), ('motions', 0.34), ('psychophysical', 0.311), ('com', 0.281), ('thresholds', 0.244), ('gmp', 0.231), ('mstd', 0.2), ('preferred', 0.163), ('recurrent', 0.126), ('across', 0.124), ('circular', 0.118), ('units', 0.112), ('populations', 0.112), ('discrimination', 0.109), ('field', 0.098), ('trends', 0.092), ('significantly', 0.092), ('stimulus', 0.088), ('sr', 0.086), ('visual', 0.082), ('trend', 0.078), ('radial', 0.077), ('tasks', 0.075), ('profile', 0.075), ('connections', 0.069), ('interconnected', 0.067), ('receptive', 0.064), ('patterns', 0.063), ('population', 0.062), ('dot', 0.062), ('vaina', 0.062), ('simulations', 0.06), ('ri', 0.059), ('flow', 0.057), ('observers', 0.057), ('sinusoidal', 0.057), ('fit', 0.054), ('sufficient', 0.054), ('responsive', 0.054), ('pattern', 0.052), ('responding', 0.051), ('spiral', 0.049), ('perceptual', 0.048), ('neurophysiology', 0.047), ('profiles', 0.045), ('connection', 0.043), ('simulated', 0.043), ('stimuli', 0.042), ('speed', 0.042), ('centers', 0.041), ('speeds', 0.041), ('contraction', 0.04), ('human', 0.04), ('cyclic', 0.04), ('spatial', 0.038), ('mechanisms', 0.038), ('performance', 0.038), ('sinusoids', 0.037), ('uniform', 0.036), ('unit', 0.036), ('studies', 0.035), ('boston', 0.035), ('biologically', 0.034), ('continuously', 0.034), ('reported', 0.033), ('elucidate', 0.032), ('specialized', 0.032), ('strength', 0.032), ('angle', 0.031), ('expanding', 0.031), ('beardsley', 0.031), ('celebrini', 0.031), ('contracting', 0.031), ('duffy', 0.031), ('gsat', 0.031), ('newsome', 0.031), ('wurtz', 0.031), ('consistent', 0.03), ('varied', 0.029), ('followed', 0.029), ('mediated', 0.029), ('presence', 0.028), ('neuroscience', 0.028), ('suggest', 0.027), ('role', 0.027), ('properties', 0.027), ('fixation', 0.027), ('rmax', 0.027), ('matched', 0.025), ('spatially', 0.025), ('bias', 0.025), ('tuning', 0.025), ('expansion', 0.025), ('define', 0.024), ('dorsal', 0.024), ('emerged', 0.024), ('graded', 0.024), ('threshold', 0.024), ('existence', 0.024), ('independently', 0.024)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 1.0000004 7 nips-2003-A Functional Architecture for Motion Pattern Processing in MSTd

Author: Scott A. Beardsley, Lucia M. Vaina

Abstract: Psychophysical studies suggest the existence of specialized detectors for component motion patterns (radial, circular, and spiral), that are consistent with the visual motion properties of cells in the dorsal medial superior temporal area (MSTd) of non-human primates. Here we use a biologically constrained model of visual motion processing in MSTd, in conjunction with psychophysical performance on two motion pattern tasks, to elucidate the computational mechanisms associated with the processing of widefield motion patterns encountered during self-motion. In both tasks discrimination thresholds varied significantly with the type of motion pattern presented, suggesting perceptual correlates to the preferred motion bias reported in MSTd. Through the model we demonstrate that while independently responding motion pattern units are capable of encoding information relevant to the visual motion tasks, equivalent psychophysical performance can only be achieved using interconnected neural populations that systematically inhibit non-responsive units. These results suggest the cyclic trends in psychophysical performance may be mediated, in part, by recurrent connections within motion pattern responsive areas whose structure is a function of the similarity in preferred motion patterns and receptive field locations between units. 1 In trod u ction A major challenge in computational neuroscience is to elucidate the architecture of the cortical circuits for sensory processing and their effective role in mediating behavior. In the visual motion system, biologically constrained models are playing an increasingly important role in this endeavor by providing an explanatory substrate linking perceptual performance and the visual properties of single cells. Single cell studies indicate the presence of complex interconnected structures in middle temporal and primary visual cortex whose most basic horizontal connections can impart considerable computational power to the underlying neural population [1, 2]. Combined psychophysical and computational studies support these findings Figure 1: a) Schematic of the graded motion pattern (GMP) task. Discrimination pairs of stimuli were created by perturbing the flow angle (φ) of each 'test' motion (with average dot speed, vav), by ±φp in the stimulus space spanned by radial and circular motions. b) Schematic of the shifted center-of-motion (COM) task. Discrimination pairs of stimuli were created by shifting the COM of the ‘test’ motion to the left and right of a central fixation point. For each motion pattern the COM was shifted within the illusory inner aperture and was never explicitly visible. and suggest that recurrent connections may play a significant role in encoding the visual motion properties associated with various psychophysical tasks [3, 4]. Using this methodology our goal is to elucidate the computational mechanisms associated with the processing of wide-field motion patterns encountered during self-motion. In the human visual motion system, psychophysical studies suggest the existence of specialized detectors for the motion pattern components (i.e., radial, circular and spiral motions) associated with self-motion [5, 6]. Neurophysiological studies reporting neurons sensitive to motion patterns in the dorsal medial superior temporal area (MSTd) support the existence of such mechanisms [7-10], and in conjunction with psychophysical studies suggest a strong link between the patterns of neural activity and motion-based perceptual performance [11, 12]. Through the combination of human psychophysical performance and biologically constrained modeling we investigate the computational role of simple recurrent connections within a population of MSTd-like units. Based on the known visual motion properties within MSTd we ask what neural structures are computationally sufficient to encode psychophysical performance on a series of motion pattern tasks. 2 M o t i o n pa t t e r n d i sc r i m i n a t i o n Using motion pattern stimuli consistent with previous studies [5, 6], we have developed a set of novel psychophysical tasks designed to facilitate a more direct comparison between human perceptual performance and the visual motion properties of cells in MSTd that have been found to underlie the discrimination of motion patterns [11, 12]. The psychophysical tasks, referred to as the graded motion pattern (GMP) and shifted center-of-motion (COM) tasks, are outlined in Fig. 1. Using a temporal two-alternative-forced-choice task we measured discrimination thresholds to global changes in the patterns of complex motion (GMP task), [13], and shifts in the center-of-motion (COM task). Stimuli were presented with central fixation using a constant stimulus paradigm and consisted of dynamic random dot displays presented in a 24o annular region (central 4o removed). In each task, the stimulus duration was randomly perturbed across presentations (440±40 msec) to control for timing-based cues, and dots moved coherently through a radial speed Figure 2: a) GMP thresholds across 8 'test' motions at two mean dot speeds for two observers. Performance varied continuously with thresholds for radial motions (φ=0, 180o) significantly lower than those for circular motions (φ=90,270o), (p<0.001; t(37)=3.39). b) COM thresholds at three mean dot speeds for two observers. As with the GMP task, performance varied continuously with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47). gradient in directions consistent with the global motion pattern presented. Discrimination thresholds were obtained across eight ‘test’ motions corresponding to expansion, contraction, CW and CCW rotation, and the four intermediate spiral motions. To minimize adaptation to specific motion patterns, opposing motions (e.g., expansion/ contraction) were interleaved across paired presentations. 2.1 Results Discrimination thresholds are reported here from a subset of the observer population consisting of three experienced psychophysical observers, one of which was naïve to the purpose of the psychophysical tasks. For each condition, performance is reported as the mean and standard error averaged across 8-12 thresholds. Across observers and dot speeds GMP thresholds followed a distinct trend in the stimulus space [13], with radial motions (expansion/contraction) significantly lower than circular motions (CW/CCW rotation), (p<0.001; t(37)=3.39), (Fig. 2a). While thresholds for the intermediate spiral motions were not significantly different from circular motions (p=0.223, t(60)=0.74), the trends across 'test' motions were well fit within the stimulus space (SB: r>0.82, SC: r>0.77) by sinusoids whose period and phase were 196 ± 10o and -72 ± 20o respectively (Fig. 1a). When the radial speed gradient was removed by randomizing the spatial distribution of dot speeds, threshold performance increased significantly across observers (p<0.05; t(17)=1.91), particularly for circular motions (p<0.005; t(25)=3.31), (data not shown). Such performance suggests a perceptual contribution associated with the presence of the speed gradient and is particularly interesting given the fact that the speed gradient did not contribute computationally relevant information to the task. However, the speed gradient did convey information regarding the integrative structure of the global motion field and as such suggests a preference of the underlying motion mechanisms for spatially structured speed information. Similar trends in performance were observed in the COM task across observers and dot speeds. Discrimination thresholds varied continuously as a function of the 'test' motion with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47) and could be well fit by a sinusoidal trend line (e.g. SB at 3 deg/s: r>0.91, period = 178 ± 10 o and phase = -70 ± 25o), (Fig. 2b). 2.2 A local or global task? The consistency of the cyclic threshold profile in stimuli that restricted the temporal integration of individual dot motions [13], and simultaneously contained all directions of motion, generally argues against a primary role for local motion mechanisms in the psychophysical tasks. While the psychophysical literature has reported a wide variety of “local” motion direction anisotropies whose properties are reminiscent of the results observed here, e.g. [14], all would predict equivalent thresholds for radial and circular motions for a set of uniformly distributed and/or spatially restricted motion direction mechanisms. Together with the computational impact of the speed gradient and psychophysical studies supporting the existence of wide-field motion pattern mechanisms [5, 6], these results suggest that the threshold differences across the GMP and COM tasks may be associated with variations in the computational properties across a series of specialized motion pattern mechanisms. 3 A computational model The similarities between the motion pattern stimuli used to quantify human perception and the visual motion properties of cells in MSTd suggests that MSTd may play a computational role in the psychophysical tasks. To examine this hypothesis, we constructed a population of MSTd-like units whose visual motion properties were consistent with the reported neurophysiology (see [13] for details). Across the population, the distribution of receptive field centers was uniform across polar angle and followed a gamma distribution Γ(5,6) across eccenticity [7]. For each unit, visual motion responses followed a gaussian tuning profile as a function of the stimulus flow angle G( φ), (σi=60±30o; [10]), and the distance of the stimulus COM from the unit’s receptive field center Gsat(xi, yi, σs=19o), Eq. 1, such that its preferred motion response was position invariant to small shifts in the COM [10] and degraded continuously for large shifts [9]. Within the model, simulations were categorized according to the distribution of preferred motions represented across the population (one reported in MSTd and a uniform control). The first distribution simulated an expansion bias in which the density of preferred motions decreased symmetrically from expansions to contraction [10]. The second distribution simulated a uniform preference for all motions and was used as a control to quantify the effects of an expansion bias on psychophysical performance. Throughout the paper we refer to simulations containing these distributions as ‘Expansion-biased’ and ‘Uniform’ respectively. 3.1 Extracting perceptual estimates from the neural code For each stimulus presentation, the ith unit’s response was calculated as the average firing rate, Ri, from the product of its motion pattern and spatial tuning profiles, ( ) Ri = Rmax G min[φ − φi ] ,σ ti G sati (x− xi , y − y i ,σ s ) + P (λ = 12 ) (1) where Rmax is the maximum preferred stimulus response (spikes/s), min[ ] refers to the minimum angular distance between the stimulus flow angle φ and the unit’s preferred motion φi, Gsat is the unit’s spatial tuning profile saturated within the central 5±3o, σti and σs are the standard deviations of the unit’s motion pattern and Figure 3: Model vs. psychophysical performance for independently responding units. Model thresholds are reported as the average (±1 S.E.) across five simulated populations. a) GMP thresholds were highest for contracting motions and lowest for expanding motions across all Expansion-biased populations. b) Comparable trends in performance were observed for COM thresholds. Comparison with the Uniform control simulations in both tasks (2000 units shown here) indicates that thresholds closely followed the distribution of preferred motions simulated within the model. spatial tuning profiles respectively, (xi,yi) is the spatial location of the unit’s receptive field center, (x,y) is the spatial location of the stimulus COM, and P(λ=12) is the background activity simulated as an uncorrelated Poisson process. The psychophysical tasks were simulated using a modified center-of-gravity ^ approach to decode estimates of the stimulus properties, i.e. flow angle (φ ) and ˆ ˆ COM location in the visual field (x, y ) , from the neural population   ∑ xi Ri ∑ y i Ri v  , i , ∑ φ i Ri  ∑ Ri i   i i   i (xˆ, yˆ , φˆ) =  i∑ R  (2) v where φi is the unit vector in the stimulus space (Fig. 1a) corresponding to the unit’s preferred motion. For each set of paired stimuli, psychophysical judgments were made by comparing the estimated stimulus properties according to the discrimination criteria, specified in the psychophysical tasks. As with the psychophysical experiments, discrimination thresholds were computed using a leastsquares fit to percent correct performance across constant stimulus levels. 3.2 Simulation 1: Independent neural responses In the first series of simulations, GMP and COM thresholds were quantified across three populations (500, 1000, and 2000 units) of independently responding units for each simulated distribution (Expansion-biased and Uniform). Across simulations, both the range in thresholds and their trends across ‘test’ motions were compared with human psychophysical performance to quantify the effects of population size and an expansion biased preferred motion distribution on model performance. Over the psychophysical range of interest (φp ± 7o), GMP thresholds for contracting motions were at chance across all Expansion-biased populations, (Fig. 3a). While thresholds for expanding motions were generally consistent with those for human observers, those for circular motions remained significantly higher for all but the largest populations. Similar trends in performance were observed for the COM task, (Fig. 3b). Here the range of COM thresholds was well matched with human performance for simulations containing 1000 units, however, the trends across motion patterns remained inconsistent even for the largest populations. Figure 4: Proposed recurrent connection profile between motion pattern units. a) Across the motion pattern space connection strength followed an inverse gaussian profile such that the ith unit (with preferred motion φi) systematically inhibited units with anti-preferred motions centered at 180+φi. b) Across the visual field connection strength followed a difference-of-gaussians profile as a function of the relative distance between receptive field centers such that spatially local units are mutually excitatory (σRe=10o) and more distant units were mutually inhibitory (σRi=80o). For simulations containing a uniform distribution of preferred motions, the threshold range was consistent with human performance on both tasks, however, the trend across motion patterns was generally flat. What variability did occur was due primarily to the discrete sampling of preferred motions across the population. Comparison of the discrimination thresholds for the Expansion-biased and Uniform populations indicates that the trend across thresholds was closely matched to the underlying distributions of preferred motions. This result in due in part to the nearequal weighting of independently responding units and can be explained to a first approximation by the proportional increase in the signal-to-noise ratio across the population as a function of the density of units responsive to a given 'test' motion. 3.3 Simulation 2: An interconnected neural structure In a second series of simulations, we examined the computational effect of adding recurrent connections between units. If the distribution of preferred motions in MSTd is in fact biased towards expansions, as the neurophysiology suggests, it seems unlikely that independent estimates of the visual motion information would be sufficient to yield the threshold profiles observed in the psychophysical tasks. We hypothesize that a simple fixed architecture of excitatory and/or inhibitory connections is sufficient to account for the cyclic trends in discrimination thresholds. Specifically, we propose that a recurrent connection profile whose strength varies as a function of (a) the similarity between preferred motion patterns and (b) the distance between receptive field centers, is computationally sufficient to recover the trends in GMP/COM performance (Fig. 4), wij = S R e − ( xi − x j )2 + ( yi − y j )2 2 2σ R e − SR e 2 − −(min[ φi − φ j ])2 ( xi − x j )2 + ( yi − y j )2 2 2 σ Ri − Sφ e 2σ I2 (3) Figure 5: Model vs. psychophysical performance for populations containing recurrent connections (σI=80o). As the number of units increased for Expansionbiased populations, discrimination thresholds decreased to psychophysical levels and the sinusoidal trend in thresholds emerged for both the (a) GMP and (b) COM tasks. Sinusoidal trends were established for as few as 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were (193.8 ± 11.7o, -70.0 ± 22.6o) and (168.2 ± 13.7o, -118.8 ± 31.8o) for the GMP and COM tasks respectively. where wij is the strength of the recurrent connection between ith and jth units, (xi,yi) and (xj,yj) denote the spatial locations of their receptive field centers, σRe (=10o) and σRi (=80o) together define the spatial extent of a difference-of-gaussians interaction between receptive field centers, and SR and Sφ scale the connection strength. To examine the effects of the spread of motion pattern-specific inhibition and connection strength in the model, σI, Sφ, and SR were considered free parameters. Within the parameter space used to define recurrent connections (i.e., σI, Sφ and SR), Monte Carlo simulations of Expansion-biased model performance (1000 units) yielded regions of high correlation on both tasks (with respect to the psychophysical thresholds, r>0.7) that were consistent across independently simulated populations. Typically these regions were well defined over a broad range such that there was significant overlap between tasks (e.g., for the GMP task (SR=0.03), σI=[45,120o], Sφ=[0.03,0.3] and for the COM task (σI=80o), Sφ = [0.03,0.08], SR = [0.005, 0.04]). Fig. 5 shows averaged threshold performance for simulations of interconnected units drawn from the highly correlated regions of the (σI, Sφ, SR) parameter space. For populations not explicitly examined in the Monte Carlo simulations connection strengths (Sφ, SR) were scaled inversely with population size to maintain an equivalent level of recurrent activity. With the incorporation of recurrent connections, the sinusoidal trend in GMP and COM thresholds emerged for Expansion-biased populations as the number of units increased. In both tasks the cyclic threshold profiles were established for 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were consistent with human performance. Unlike the Expansion-biased populations, Uniform populations were not significantly affected by the presence of recurrent connections (Fig. 5). Both the range in thresholds and the flat trend across motion patterns were well matched to those in Section 3.2. Together these results suggest that the sinusoidal trends in GMP and COM performance may be mediated by the combined contribution of the recurrent interconnections and the bias in preferred motions across the population. 4 D i s c u s s i on Using a biologically constrained computational model in conjunction with human psychophysical performance on two motion pattern tasks we have shown that the visual motion information encoded across an interconnected population of cells responsive to motion patterns, such as those in MSTd, is computationally sufficient to extract perceptual estimates consistent with human performance. Specifically, we have shown that the cyclic trend in psychophysical performance observed across tasks, (a) cannot be reproduced using populations of independently responding units and (b) is dependent, in part, on the presence of an expanding motion bias in the distribution of preferred motions across the neural population. The model’s performance suggests the presence of specific recurrent structures within motion pattern responsive areas, such as MSTd, whose strength varies as a function of the similarity between preferred motion patterns and the distance between receptive field centers. While such structures have not been explicitly examined in MSTd and other higher visual motion areas there is anecdotal support for the presence of inhibitory connections [8]. Together, these results suggest that robust processing of the motion patterns associated with self-motion and optic flow may be mediated, in part, by recurrent structures in extrastriate visual motion areas whose distributions of preferred motions are biased strongly in favor of expanding motions. Acknowledgments This work was supported by National Institutes of Health grant EY-2R01-07861-13 to L.M.V. References [1] Malach, R., Schirman, T., Harel, M., Tootell, R., & Malonek, D., (1997), Cerebral Cortex, 7(4): 386-393. [2] Gilbert, C. D., (1992), Neuron, 9: 1-13. [3] Koechlin, E., Anton, J., & Burnod, Y., (1999), Biological Cybernetics, 80: 2544. [4] Stemmler, M., Usher, M., & Niebur, E., (1995), Science, 269: 1877-1880. [5] Burr, D. C., Morrone, M. C., & Vaina, L. M., (1998), Vision Research, 38(12): 1731-1743. [6] Meese, T. S. & Harris, S. J., (2002), Vision Research, 42: 1073-1080. [7] Tanaka, K. & Saito, H. A., (1989), Journal of Neurophysiology, 62(3): 626-641. [8] Duffy, C. J. & Wurtz, R. H., (1991), Journal of Neurophysiology, 65(6): 13461359. [9] Duffy, C. J. & Wurtz, R. H., (1995), Journal of Neuroscience, 15(7): 5192-5208. [10] Graziano, M. S., Anderson, R. A., & Snowden, R., (1994), Journal of Neuroscience, 14(1): 54-67. [11] Celebrini, S. & Newsome, W., (1994), Journal of Neuroscience, 14(7): 41094124. [12] Celebrini, S. & Newsome, W. T., (1995), Journal of Neurophysiology, 73(2): 437-448. [13] Beardsley, S. A. & Vaina, L. M., (2001), Journal of Computational Neuroscience, 10: 255-280. [14] Matthews, N. & Qian, N., (1999), Vision Research, 39: 2205-2211.

2 0.17593934 37 nips-2003-Automatic Annotation of Everyday Movements

Author: Deva Ramanan, David A. Forsyth

Abstract: This paper describes a system that can annotate a video sequence with: a description of the appearance of each actor; when the actor is in view; and a representation of the actor’s activity while in view. The system does not require a fixed background, and is automatic. The system works by (1) tracking people in 2D and then, using an annotated motion capture dataset, (2) synthesizing an annotated 3D motion sequence matching the 2D tracks. The 3D motion capture data is manually annotated off-line using a class structure that describes everyday motions and allows motion annotations to be composed — one may jump while running, for example. Descriptions computed from video of real motions show that the method is accurate.

3 0.12453744 106 nips-2003-Learning Non-Rigid 3D Shape from 2D Motion

Author: Lorenzo Torresani, Aaron Hertzmann, Christoph Bregler

Abstract: This paper presents an algorithm for learning the time-varying shape of a non-rigid 3D object from uncalibrated 2D tracking data. We model shape motion as a rigid component (rotation and translation) combined with a non-rigid deformation. Reconstruction is ill-posed if arbitrary deformations are allowed. We constrain the problem by assuming that the object shape at each time instant is drawn from a Gaussian distribution. Based on this assumption, the algorithm simultaneously estimates 3D shape and motion for each time frame, learns the parameters of the Gaussian, and robustly fills-in missing data points. We then extend the algorithm to model temporal smoothness in object shape, thus allowing it to handle severe cases of missing data. 1

4 0.1129192 10 nips-2003-A Low-Power Analog VLSI Visual Collision Detector

Author: Reid R. Harrison

Abstract: We have designed and tested a single-chip analog VLSI sensor that detects imminent collisions by measuring radially expansive optic flow. The design of the chip is based on a model proposed to explain leg-extension behavior in flies during landing approaches. A new elementary motion detector (EMD) circuit was developed to measure optic flow. This EMD circuit models the bandpass nature of large monopolar cells (LMCs) immediately postsynaptic to photoreceptors in the fly visual system. A 16 × 16 array of 2-D motion detectors was fabricated on a 2.24 mm × 2.24 mm die in a standard 0.5-µm CMOS process. The chip consumes 140 µW of power from a 5 V supply. With the addition of wide-angle optics, the sensor is able to detect collisions around 500 ms before impact in complex, real-world scenes. 1

5 0.1032171 69 nips-2003-Factorization with Uncertainty and Missing Data: Exploiting Temporal Coherence

Author: Amit Gruber, Yair Weiss

Abstract: The problem of “Structure From Motion” is a central problem in vision: given the 2D locations of certain points we wish to recover the camera motion and the 3D coordinates of the points. Under simplified camera models, the problem reduces to factorizing a measurement matrix into the product of two low rank matrices. Each element of the measurement matrix contains the position of a point in a particular image. When all elements are observed, the problem can be solved trivially using SVD, but in any realistic situation many elements of the matrix are missing and the ones that are observed have a different directional uncertainty. Under these conditions, most existing factorization algorithms fail while human perception is relatively unchanged. In this paper we use the well known EM algorithm for factor analysis to perform factorization. This allows us to easily handle missing data and measurement uncertainty and more importantly allows us to place a prior on the temporal trajectory of the latent variables (the camera position). We show that incorporating this prior gives a significant improvement in performance in challenging image sequences. 1

6 0.093086623 67 nips-2003-Eye Micro-movements Improve Stimulus Detection Beyond the Nyquist Limit in the Peripheral Retina

7 0.086708352 140 nips-2003-Nonlinear Processing in LGN Neurons

8 0.073892497 16 nips-2003-A Recurrent Model of Orientation Maps with Simple and Complex Cells

9 0.07318522 85 nips-2003-Human and Ideal Observers for Detecting Image Curves

10 0.070742309 35 nips-2003-Attractive People: Assembling Loose-Limbed Models using Non-parametric Belief Propagation

11 0.067289859 45 nips-2003-Circuit Optimization Predicts Dynamic Networks for Chemosensory Orientation in Nematode C. elegans

12 0.062411867 51 nips-2003-Design of Experiments via Information Theory

13 0.060233261 160 nips-2003-Prediction on Spike Data Using Kernel Algorithms

14 0.059654925 43 nips-2003-Bounded Invariance and the Formation of Place Fields

15 0.05469862 119 nips-2003-Local Phase Coherence and the Perception of Blur

16 0.05179872 49 nips-2003-Decoding V1 Neuronal Activity using Particle Filtering with Volterra Kernels

17 0.051298827 95 nips-2003-Insights from Machine Learning Applied to Human Visual Classification

18 0.050560091 56 nips-2003-Dopamine Modulation in a Basal Ganglio-Cortical Network of Working Memory

19 0.049895287 175 nips-2003-Sensory Modality Segregation

20 0.049202729 22 nips-2003-An Improved Scheme for Detection and Labelling in Johansson Displays


similar papers computed by lsi model

lsi for this paper:

topicId topicWeight

[(0, -0.125), (1, 0.034), (2, 0.161), (3, -0.003), (4, -0.093), (5, -0.062), (6, 0.099), (7, 0.055), (8, 0.033), (9, -0.054), (10, -0.014), (11, 0.149), (12, 0.046), (13, 0.138), (14, 0.091), (15, 0.094), (16, 0.102), (17, -0.108), (18, -0.025), (19, -0.191), (20, 0.038), (21, -0.254), (22, -0.007), (23, 0.028), (24, -0.014), (25, -0.015), (26, 0.035), (27, -0.021), (28, 0.117), (29, -0.147), (30, 0.015), (31, 0.103), (32, -0.025), (33, 0.122), (34, 0.047), (35, 0.02), (36, -0.017), (37, 0.017), (38, -0.01), (39, 0.117), (40, 0.061), (41, -0.038), (42, -0.01), (43, -0.001), (44, -0.037), (45, 0.092), (46, -0.064), (47, 0.053), (48, 0.032), (49, -0.004)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.98901916 7 nips-2003-A Functional Architecture for Motion Pattern Processing in MSTd

Author: Scott A. Beardsley, Lucia M. Vaina

Abstract: Psychophysical studies suggest the existence of specialized detectors for component motion patterns (radial, circular, and spiral), that are consistent with the visual motion properties of cells in the dorsal medial superior temporal area (MSTd) of non-human primates. Here we use a biologically constrained model of visual motion processing in MSTd, in conjunction with psychophysical performance on two motion pattern tasks, to elucidate the computational mechanisms associated with the processing of widefield motion patterns encountered during self-motion. In both tasks discrimination thresholds varied significantly with the type of motion pattern presented, suggesting perceptual correlates to the preferred motion bias reported in MSTd. Through the model we demonstrate that while independently responding motion pattern units are capable of encoding information relevant to the visual motion tasks, equivalent psychophysical performance can only be achieved using interconnected neural populations that systematically inhibit non-responsive units. These results suggest the cyclic trends in psychophysical performance may be mediated, in part, by recurrent connections within motion pattern responsive areas whose structure is a function of the similarity in preferred motion patterns and receptive field locations between units. 1 In trod u ction A major challenge in computational neuroscience is to elucidate the architecture of the cortical circuits for sensory processing and their effective role in mediating behavior. In the visual motion system, biologically constrained models are playing an increasingly important role in this endeavor by providing an explanatory substrate linking perceptual performance and the visual properties of single cells. Single cell studies indicate the presence of complex interconnected structures in middle temporal and primary visual cortex whose most basic horizontal connections can impart considerable computational power to the underlying neural population [1, 2]. Combined psychophysical and computational studies support these findings Figure 1: a) Schematic of the graded motion pattern (GMP) task. Discrimination pairs of stimuli were created by perturbing the flow angle (φ) of each 'test' motion (with average dot speed, vav), by ±φp in the stimulus space spanned by radial and circular motions. b) Schematic of the shifted center-of-motion (COM) task. Discrimination pairs of stimuli were created by shifting the COM of the ‘test’ motion to the left and right of a central fixation point. For each motion pattern the COM was shifted within the illusory inner aperture and was never explicitly visible. and suggest that recurrent connections may play a significant role in encoding the visual motion properties associated with various psychophysical tasks [3, 4]. Using this methodology our goal is to elucidate the computational mechanisms associated with the processing of wide-field motion patterns encountered during self-motion. In the human visual motion system, psychophysical studies suggest the existence of specialized detectors for the motion pattern components (i.e., radial, circular and spiral motions) associated with self-motion [5, 6]. Neurophysiological studies reporting neurons sensitive to motion patterns in the dorsal medial superior temporal area (MSTd) support the existence of such mechanisms [7-10], and in conjunction with psychophysical studies suggest a strong link between the patterns of neural activity and motion-based perceptual performance [11, 12]. Through the combination of human psychophysical performance and biologically constrained modeling we investigate the computational role of simple recurrent connections within a population of MSTd-like units. Based on the known visual motion properties within MSTd we ask what neural structures are computationally sufficient to encode psychophysical performance on a series of motion pattern tasks. 2 M o t i o n pa t t e r n d i sc r i m i n a t i o n Using motion pattern stimuli consistent with previous studies [5, 6], we have developed a set of novel psychophysical tasks designed to facilitate a more direct comparison between human perceptual performance and the visual motion properties of cells in MSTd that have been found to underlie the discrimination of motion patterns [11, 12]. The psychophysical tasks, referred to as the graded motion pattern (GMP) and shifted center-of-motion (COM) tasks, are outlined in Fig. 1. Using a temporal two-alternative-forced-choice task we measured discrimination thresholds to global changes in the patterns of complex motion (GMP task), [13], and shifts in the center-of-motion (COM task). Stimuli were presented with central fixation using a constant stimulus paradigm and consisted of dynamic random dot displays presented in a 24o annular region (central 4o removed). In each task, the stimulus duration was randomly perturbed across presentations (440±40 msec) to control for timing-based cues, and dots moved coherently through a radial speed Figure 2: a) GMP thresholds across 8 'test' motions at two mean dot speeds for two observers. Performance varied continuously with thresholds for radial motions (φ=0, 180o) significantly lower than those for circular motions (φ=90,270o), (p<0.001; t(37)=3.39). b) COM thresholds at three mean dot speeds for two observers. As with the GMP task, performance varied continuously with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47). gradient in directions consistent with the global motion pattern presented. Discrimination thresholds were obtained across eight ‘test’ motions corresponding to expansion, contraction, CW and CCW rotation, and the four intermediate spiral motions. To minimize adaptation to specific motion patterns, opposing motions (e.g., expansion/ contraction) were interleaved across paired presentations. 2.1 Results Discrimination thresholds are reported here from a subset of the observer population consisting of three experienced psychophysical observers, one of which was naïve to the purpose of the psychophysical tasks. For each condition, performance is reported as the mean and standard error averaged across 8-12 thresholds. Across observers and dot speeds GMP thresholds followed a distinct trend in the stimulus space [13], with radial motions (expansion/contraction) significantly lower than circular motions (CW/CCW rotation), (p<0.001; t(37)=3.39), (Fig. 2a). While thresholds for the intermediate spiral motions were not significantly different from circular motions (p=0.223, t(60)=0.74), the trends across 'test' motions were well fit within the stimulus space (SB: r>0.82, SC: r>0.77) by sinusoids whose period and phase were 196 ± 10o and -72 ± 20o respectively (Fig. 1a). When the radial speed gradient was removed by randomizing the spatial distribution of dot speeds, threshold performance increased significantly across observers (p<0.05; t(17)=1.91), particularly for circular motions (p<0.005; t(25)=3.31), (data not shown). Such performance suggests a perceptual contribution associated with the presence of the speed gradient and is particularly interesting given the fact that the speed gradient did not contribute computationally relevant information to the task. However, the speed gradient did convey information regarding the integrative structure of the global motion field and as such suggests a preference of the underlying motion mechanisms for spatially structured speed information. Similar trends in performance were observed in the COM task across observers and dot speeds. Discrimination thresholds varied continuously as a function of the 'test' motion with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47) and could be well fit by a sinusoidal trend line (e.g. SB at 3 deg/s: r>0.91, period = 178 ± 10 o and phase = -70 ± 25o), (Fig. 2b). 2.2 A local or global task? The consistency of the cyclic threshold profile in stimuli that restricted the temporal integration of individual dot motions [13], and simultaneously contained all directions of motion, generally argues against a primary role for local motion mechanisms in the psychophysical tasks. While the psychophysical literature has reported a wide variety of “local” motion direction anisotropies whose properties are reminiscent of the results observed here, e.g. [14], all would predict equivalent thresholds for radial and circular motions for a set of uniformly distributed and/or spatially restricted motion direction mechanisms. Together with the computational impact of the speed gradient and psychophysical studies supporting the existence of wide-field motion pattern mechanisms [5, 6], these results suggest that the threshold differences across the GMP and COM tasks may be associated with variations in the computational properties across a series of specialized motion pattern mechanisms. 3 A computational model The similarities between the motion pattern stimuli used to quantify human perception and the visual motion properties of cells in MSTd suggests that MSTd may play a computational role in the psychophysical tasks. To examine this hypothesis, we constructed a population of MSTd-like units whose visual motion properties were consistent with the reported neurophysiology (see [13] for details). Across the population, the distribution of receptive field centers was uniform across polar angle and followed a gamma distribution Γ(5,6) across eccenticity [7]. For each unit, visual motion responses followed a gaussian tuning profile as a function of the stimulus flow angle G( φ), (σi=60±30o; [10]), and the distance of the stimulus COM from the unit’s receptive field center Gsat(xi, yi, σs=19o), Eq. 1, such that its preferred motion response was position invariant to small shifts in the COM [10] and degraded continuously for large shifts [9]. Within the model, simulations were categorized according to the distribution of preferred motions represented across the population (one reported in MSTd and a uniform control). The first distribution simulated an expansion bias in which the density of preferred motions decreased symmetrically from expansions to contraction [10]. The second distribution simulated a uniform preference for all motions and was used as a control to quantify the effects of an expansion bias on psychophysical performance. Throughout the paper we refer to simulations containing these distributions as ‘Expansion-biased’ and ‘Uniform’ respectively. 3.1 Extracting perceptual estimates from the neural code For each stimulus presentation, the ith unit’s response was calculated as the average firing rate, Ri, from the product of its motion pattern and spatial tuning profiles, ( ) Ri = Rmax G min[φ − φi ] ,σ ti G sati (x− xi , y − y i ,σ s ) + P (λ = 12 ) (1) where Rmax is the maximum preferred stimulus response (spikes/s), min[ ] refers to the minimum angular distance between the stimulus flow angle φ and the unit’s preferred motion φi, Gsat is the unit’s spatial tuning profile saturated within the central 5±3o, σti and σs are the standard deviations of the unit’s motion pattern and Figure 3: Model vs. psychophysical performance for independently responding units. Model thresholds are reported as the average (±1 S.E.) across five simulated populations. a) GMP thresholds were highest for contracting motions and lowest for expanding motions across all Expansion-biased populations. b) Comparable trends in performance were observed for COM thresholds. Comparison with the Uniform control simulations in both tasks (2000 units shown here) indicates that thresholds closely followed the distribution of preferred motions simulated within the model. spatial tuning profiles respectively, (xi,yi) is the spatial location of the unit’s receptive field center, (x,y) is the spatial location of the stimulus COM, and P(λ=12) is the background activity simulated as an uncorrelated Poisson process. The psychophysical tasks were simulated using a modified center-of-gravity ^ approach to decode estimates of the stimulus properties, i.e. flow angle (φ ) and ˆ ˆ COM location in the visual field (x, y ) , from the neural population   ∑ xi Ri ∑ y i Ri v  , i , ∑ φ i Ri  ∑ Ri i   i i   i (xˆ, yˆ , φˆ) =  i∑ R  (2) v where φi is the unit vector in the stimulus space (Fig. 1a) corresponding to the unit’s preferred motion. For each set of paired stimuli, psychophysical judgments were made by comparing the estimated stimulus properties according to the discrimination criteria, specified in the psychophysical tasks. As with the psychophysical experiments, discrimination thresholds were computed using a leastsquares fit to percent correct performance across constant stimulus levels. 3.2 Simulation 1: Independent neural responses In the first series of simulations, GMP and COM thresholds were quantified across three populations (500, 1000, and 2000 units) of independently responding units for each simulated distribution (Expansion-biased and Uniform). Across simulations, both the range in thresholds and their trends across ‘test’ motions were compared with human psychophysical performance to quantify the effects of population size and an expansion biased preferred motion distribution on model performance. Over the psychophysical range of interest (φp ± 7o), GMP thresholds for contracting motions were at chance across all Expansion-biased populations, (Fig. 3a). While thresholds for expanding motions were generally consistent with those for human observers, those for circular motions remained significantly higher for all but the largest populations. Similar trends in performance were observed for the COM task, (Fig. 3b). Here the range of COM thresholds was well matched with human performance for simulations containing 1000 units, however, the trends across motion patterns remained inconsistent even for the largest populations. Figure 4: Proposed recurrent connection profile between motion pattern units. a) Across the motion pattern space connection strength followed an inverse gaussian profile such that the ith unit (with preferred motion φi) systematically inhibited units with anti-preferred motions centered at 180+φi. b) Across the visual field connection strength followed a difference-of-gaussians profile as a function of the relative distance between receptive field centers such that spatially local units are mutually excitatory (σRe=10o) and more distant units were mutually inhibitory (σRi=80o). For simulations containing a uniform distribution of preferred motions, the threshold range was consistent with human performance on both tasks, however, the trend across motion patterns was generally flat. What variability did occur was due primarily to the discrete sampling of preferred motions across the population. Comparison of the discrimination thresholds for the Expansion-biased and Uniform populations indicates that the trend across thresholds was closely matched to the underlying distributions of preferred motions. This result in due in part to the nearequal weighting of independently responding units and can be explained to a first approximation by the proportional increase in the signal-to-noise ratio across the population as a function of the density of units responsive to a given 'test' motion. 3.3 Simulation 2: An interconnected neural structure In a second series of simulations, we examined the computational effect of adding recurrent connections between units. If the distribution of preferred motions in MSTd is in fact biased towards expansions, as the neurophysiology suggests, it seems unlikely that independent estimates of the visual motion information would be sufficient to yield the threshold profiles observed in the psychophysical tasks. We hypothesize that a simple fixed architecture of excitatory and/or inhibitory connections is sufficient to account for the cyclic trends in discrimination thresholds. Specifically, we propose that a recurrent connection profile whose strength varies as a function of (a) the similarity between preferred motion patterns and (b) the distance between receptive field centers, is computationally sufficient to recover the trends in GMP/COM performance (Fig. 4), wij = S R e − ( xi − x j )2 + ( yi − y j )2 2 2σ R e − SR e 2 − −(min[ φi − φ j ])2 ( xi − x j )2 + ( yi − y j )2 2 2 σ Ri − Sφ e 2σ I2 (3) Figure 5: Model vs. psychophysical performance for populations containing recurrent connections (σI=80o). As the number of units increased for Expansionbiased populations, discrimination thresholds decreased to psychophysical levels and the sinusoidal trend in thresholds emerged for both the (a) GMP and (b) COM tasks. Sinusoidal trends were established for as few as 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were (193.8 ± 11.7o, -70.0 ± 22.6o) and (168.2 ± 13.7o, -118.8 ± 31.8o) for the GMP and COM tasks respectively. where wij is the strength of the recurrent connection between ith and jth units, (xi,yi) and (xj,yj) denote the spatial locations of their receptive field centers, σRe (=10o) and σRi (=80o) together define the spatial extent of a difference-of-gaussians interaction between receptive field centers, and SR and Sφ scale the connection strength. To examine the effects of the spread of motion pattern-specific inhibition and connection strength in the model, σI, Sφ, and SR were considered free parameters. Within the parameter space used to define recurrent connections (i.e., σI, Sφ and SR), Monte Carlo simulations of Expansion-biased model performance (1000 units) yielded regions of high correlation on both tasks (with respect to the psychophysical thresholds, r>0.7) that were consistent across independently simulated populations. Typically these regions were well defined over a broad range such that there was significant overlap between tasks (e.g., for the GMP task (SR=0.03), σI=[45,120o], Sφ=[0.03,0.3] and for the COM task (σI=80o), Sφ = [0.03,0.08], SR = [0.005, 0.04]). Fig. 5 shows averaged threshold performance for simulations of interconnected units drawn from the highly correlated regions of the (σI, Sφ, SR) parameter space. For populations not explicitly examined in the Monte Carlo simulations connection strengths (Sφ, SR) were scaled inversely with population size to maintain an equivalent level of recurrent activity. With the incorporation of recurrent connections, the sinusoidal trend in GMP and COM thresholds emerged for Expansion-biased populations as the number of units increased. In both tasks the cyclic threshold profiles were established for 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were consistent with human performance. Unlike the Expansion-biased populations, Uniform populations were not significantly affected by the presence of recurrent connections (Fig. 5). Both the range in thresholds and the flat trend across motion patterns were well matched to those in Section 3.2. Together these results suggest that the sinusoidal trends in GMP and COM performance may be mediated by the combined contribution of the recurrent interconnections and the bias in preferred motions across the population. 4 D i s c u s s i on Using a biologically constrained computational model in conjunction with human psychophysical performance on two motion pattern tasks we have shown that the visual motion information encoded across an interconnected population of cells responsive to motion patterns, such as those in MSTd, is computationally sufficient to extract perceptual estimates consistent with human performance. Specifically, we have shown that the cyclic trend in psychophysical performance observed across tasks, (a) cannot be reproduced using populations of independently responding units and (b) is dependent, in part, on the presence of an expanding motion bias in the distribution of preferred motions across the neural population. The model’s performance suggests the presence of specific recurrent structures within motion pattern responsive areas, such as MSTd, whose strength varies as a function of the similarity between preferred motion patterns and the distance between receptive field centers. While such structures have not been explicitly examined in MSTd and other higher visual motion areas there is anecdotal support for the presence of inhibitory connections [8]. Together, these results suggest that robust processing of the motion patterns associated with self-motion and optic flow may be mediated, in part, by recurrent structures in extrastriate visual motion areas whose distributions of preferred motions are biased strongly in favor of expanding motions. Acknowledgments This work was supported by National Institutes of Health grant EY-2R01-07861-13 to L.M.V. References [1] Malach, R., Schirman, T., Harel, M., Tootell, R., & Malonek, D., (1997), Cerebral Cortex, 7(4): 386-393. [2] Gilbert, C. D., (1992), Neuron, 9: 1-13. [3] Koechlin, E., Anton, J., & Burnod, Y., (1999), Biological Cybernetics, 80: 2544. [4] Stemmler, M., Usher, M., & Niebur, E., (1995), Science, 269: 1877-1880. [5] Burr, D. C., Morrone, M. C., & Vaina, L. M., (1998), Vision Research, 38(12): 1731-1743. [6] Meese, T. S. & Harris, S. J., (2002), Vision Research, 42: 1073-1080. [7] Tanaka, K. & Saito, H. A., (1989), Journal of Neurophysiology, 62(3): 626-641. [8] Duffy, C. J. & Wurtz, R. H., (1991), Journal of Neurophysiology, 65(6): 13461359. [9] Duffy, C. J. & Wurtz, R. H., (1995), Journal of Neuroscience, 15(7): 5192-5208. [10] Graziano, M. S., Anderson, R. A., & Snowden, R., (1994), Journal of Neuroscience, 14(1): 54-67. [11] Celebrini, S. & Newsome, W., (1994), Journal of Neuroscience, 14(7): 41094124. [12] Celebrini, S. & Newsome, W. T., (1995), Journal of Neurophysiology, 73(2): 437-448. [13] Beardsley, S. A. & Vaina, L. M., (2001), Journal of Computational Neuroscience, 10: 255-280. [14] Matthews, N. & Qian, N., (1999), Vision Research, 39: 2205-2211.

2 0.69786006 37 nips-2003-Automatic Annotation of Everyday Movements

Author: Deva Ramanan, David A. Forsyth

Abstract: This paper describes a system that can annotate a video sequence with: a description of the appearance of each actor; when the actor is in view; and a representation of the actor’s activity while in view. The system does not require a fixed background, and is automatic. The system works by (1) tracking people in 2D and then, using an annotated motion capture dataset, (2) synthesizing an annotated 3D motion sequence matching the 2D tracks. The 3D motion capture data is manually annotated off-line using a class structure that describes everyday motions and allows motion annotations to be composed — one may jump while running, for example. Descriptions computed from video of real motions show that the method is accurate.

3 0.52789491 69 nips-2003-Factorization with Uncertainty and Missing Data: Exploiting Temporal Coherence

Author: Amit Gruber, Yair Weiss

Abstract: The problem of “Structure From Motion” is a central problem in vision: given the 2D locations of certain points we wish to recover the camera motion and the 3D coordinates of the points. Under simplified camera models, the problem reduces to factorizing a measurement matrix into the product of two low rank matrices. Each element of the measurement matrix contains the position of a point in a particular image. When all elements are observed, the problem can be solved trivially using SVD, but in any realistic situation many elements of the matrix are missing and the ones that are observed have a different directional uncertainty. Under these conditions, most existing factorization algorithms fail while human perception is relatively unchanged. In this paper we use the well known EM algorithm for factor analysis to perform factorization. This allows us to easily handle missing data and measurement uncertainty and more importantly allows us to place a prior on the temporal trajectory of the latent variables (the camera position). We show that incorporating this prior gives a significant improvement in performance in challenging image sequences. 1

4 0.51504058 140 nips-2003-Nonlinear Processing in LGN Neurons

Author: Vincent Bonin, Valerio Mante, Matteo Carandini

Abstract: According to a widely held view, neurons in lateral geniculate nucleus (LGN) operate on visual stimuli in a linear fashion. There is ample evidence, however, that LGN responses are not entirely linear. To account for nonlinearities we propose a model that synthesizes more than 30 years of research in the field. Model neurons have a linear receptive field, and a nonlinear, divisive suppressive field. The suppressive field computes local root-meansquare contrast. To test this model we recorded responses from LGN of anesthetized paralyzed cats. We estimate model parameters from a basic set of measurements and show that the model can accurately predict responses to novel stimuli. The model might serve as the new standard model of LGN responses. It specifies how visual processing in LGN involves both linear filtering and divisive gain control. 1 In t rod u ct i on According to a widely held view, neurons in lateral geniculate nucleus (LGN) operate linearly (Cai et al., 1997; Dan et al., 1996). Their response L(t) is the convolution of the map of stimulus contrast S(x,t) with a receptive field F(x,t): L (t ) = [ S∗ F ] ( 0, t ) The receptive field F(x,t) is typically taken to be a difference of Gaussians in space (Rodieck, 1965) and a difference of Gamma functions in time (Cai et al., 1997). This linear model accurately predicts the selectivity of responses for spatiotemporal frequency as measured with gratings (Cai et al., 1997; Enroth-Cugell and Robson, 1966). It also predicts the main features of responses to complex dynamic video sequences (Dan et al., 1996). 150 spikes/s Data Model Figure 1. Response of an LGN neuron to a dynamic video sequence along with the prediction made by the linear model. Stimuli were sequences from Walt Disney’s “Tarzan”. From Mante et al. (2002). The linear model, however, suffers from limitations. For example, consider the response of an LGN neuron to a complex dynamic video sequences (Figure 1). The response is characterized by long periods of relative silence interspersed with brief events of high firing rate (Figure 1, thick traces). The linear model (Figure 1, thin traces) successfully predicts the timing of these firing events but fails to account for their magnitude (Mante et al., 2002). The limitations of the linear model are not surprising since there is ample evidence that LGN responses are nonlinear. For instance, responses to drifting gratings saturate as contrast is increased (Sclar et al., 1990) and are reduced, or masked, by superposition of a second grating (Bonin et al., 2002). Moreover, responses are selective for stimulus size (Cleland et al., 1983; Hubel and Wiesel, 1961; Jones and Sillito, 1991) in a nonlinear manner (Solomon et al., 2002). We propose that these and other nonlinearities can be explained by a nonlinear model incorporating a nonlinear suppressive field. The qualitative notion of a suppressive field was proposed three decades ago by Levick and collaborators (1972). We propose that the suppressive field computes local root-mean-square contrast, and operates divisively on the receptive field output. Basic elements of this model appeared in studies of contrast gain control in retina (Shapley and Victor, 1978) and in primary visual cortex (Cavanaugh et al., 2002; Heeger, 1992; Schwartz and Simoncelli, 2001). Some of these notions have been applied to LGN (Solomon et al., 2002), to fit responses to a limited set of stimuli with tailored parameter sets. Here we show that a single model with fixed parameters predicts responses to a broad range of stimuli. 2 Mod el In the model (Figure 2), the linear response of the receptive field L(t) is divided by the output of the suppressive field. The latter is a measure of local root-mean-square contrast c local. The result of the division is a generator potential V (t ) = Vmax L (t ) c50 + clocal , where c 50 is a constant. F(x,t) Stimulus S(x,t) V0 L(t) Receptive Field R(t) Firing rate Rectification H(x,t) S*(x,t) c50 clocal Suppressive Field Filter Figure 2. Nonlinear model of LGN responses. The suppressive field operates on a filtered version of the stimulus, S*=S*H, where H is a linear filter and * denotes convolution. The squared output of the suppressive field is the local mean square (the local variance) of the filtered stimulus: clocal = S ( x, t ) G ( x ) dx dt , 2 * 2 ∫∫ where G(x) is a 2-dimensional Gaussian. Firing rate is a rectified version of generator potential, with threshold V thresh:   R(t ) = V (t ) − Vthresh + .   To test the nonlinear model, we recorded responses from neurons in the LGN of anesthetized paralyzed cats. Methods for these recordings were described elsewhere (Freeman et al., 2002). 3 Resu l t s We proceed in two steps: first we estimate model parameters by fitting the model to a large set of canonical data; second we fix model parameters and evaluate the model by predicting responses to a novel set of stimuli. B 60 40 20 0 0.01 0.2 4 Spatial Frequency (cpd) Response (spikes/s) Response (spikes/s) A 80 60 40 20 0 0.5 5 50 Temporal Frequency (Hz) Figure 3. Estimating the receptive field in an example LGN cell. Stimuli are gratings varying in spatial (A) and temporal (B) frequency. Responses are the harmonic component of spike trains at the grating temporal frequency. Error bars represent standard deviation of responses. Curves indicate model fit. Response (spikes/s) 0.25 0.50 0.75 Test contrast 100 80 60 40 20 0 1.00 100 D 80 60 40 20 0 0.5 Response (spikes/s) 50 0 C B 100 Response (spikes/s) Response (spikes/s) A 0.25 0.50 0.75 1.00 Mask contrast 100 80 60 40 20 4.0 32.0 Mask diameter (deg) 0 0.01 0.20 4.00 Mask spatial frequency (cpd) Figure 4. Estimating the suppressive field in the example LGN cell. Stimuli are sums of a test grating and a mask grating. Responses are the harmonic component of spike trains at the temporal frequency of test. A: Responses to test alone. B-D: Responses to test+mask as function of three mask attributes: contrast (B), diameter (C) and spatial frequency (D). Gray areas indicate baseline response (test alone, 50% contrast). Dashed curves are predictions of linear model. Solid curves indicate fit of nonlinear model. 3.1 C h a r a c te r i z i n g t he r e c e p ti v e fi e l d We obtain the parameters of the receptive field F(x,t) from responses to large drifting gratings (Figure 3). These stimuli elicit approximately constant output in the suppressive field, so they allow us to characterize the receptive field. Responses to different spatial frequencies constrain F(x,t) in space (Figure 3A). Responses to different temporal frequencies constrain F(x,t) in time (Figure 3B). 3.2 C h a r a c te r i z i n g t he s u p p r e s s i v e f i e l d To characterize the divisive stage, we start by measuring how responses saturate at high contrast (Figure 4A). A linear model cannot account for this contrast saturation (Figure 4A, dashed curve). The nonlinear model (Figure 4A, solid curve) captures saturation because increases in receptive field output are attenuated by increases in suppressive field output. At low contrast, no saturation is observed because the output of the suppressive field is dominated by the constant c 50. From these data we estimate the value of c50. To obtain the parameters of the suppressive field, we recorded responses to sums of two drifting gratings (Figure 4B-D): an optimal test grating at 50% contrast, which elicits a large baseline response, and a mask grating that modulates this response. Test and mask temporal frequencies are incommensurate so that they temporally label a test response (at the frequency of the test) and a mask response (at the frequency of the mask) (Bonds, 1989). We vary mask attributes and study how they affect the test responses. Increasing mask contrast progressively suppresses responses (Figure 4B). The linear model fails to account for this suppression (Figure 4B, dashed curve). The nonlinear model (Figure 4B, solid curve) captures it because increasing mask contrast increases the suppressive field output while the receptive field output (at the temporal frequency of the test) remains constant. With masks of low contrast there is little suppression because the output of the suppressive field is dominated by the constant c 50 . Similar effects are seen if we increase mask diameter. Responses decrease until they reach a plateau (Figure 4C). A linear model predicts no decrease (Figure 4C, dashed curve). The nonlinear model (Figure 4C, solid curve) captures it because increasing mask diameter increases the suppressive field output while it does not affect the receptive field output. A plateau is reached once masks extend beyond the suppressive field. From these data we estimate the size of the Gaussian envelope G(x) of the suppressive field. Finally, the strength of suppression depends on mask spatial frequency (Figure 4D). At high frequencies, no suppression is elicited. Reducing spatial frequency increases suppression. This dependence of suppression on spatial frequency is captured in the nonlinear model by the filter H(x,t). From these data we estimate the spatial characteristics of the filter. From similar experiments involving different temporal frequencies (not shown), we estimate the filter’s selectivity for temporal frequency. 3.3 P r e d i c ti n g r e s p o n s e s t o n o v e l s ti m u l i We have seen that with a fixed set of parameters the model provides a good fit to a large set of measurements (Figure 3 and Figure 4). We now test whether the model predicts responses to a set of novel stimuli: drifting gratings varying in contrast and diameter. Responses to high contrast stimuli exhibit size tuning (Figure 5A, squares): they grow with size for small diameters, reach a maximum value at intermediate diameter and are reduced for large diameters (Jones and Sillito, 1991). Size tuning , however, strongly depends on stimulus contrast (Solomon et al., 2002): no size tuning is observed at low contrast (Figure 5A, circles). The model predicts these effects (Figure 5A, curves). For large, high contrast stimuli the output of the suppressive field is dominated by c local, resulting in suppression of responses. At low contrast, c local is much smaller than c50, and the suppressive field does not affect responses. Similar considerations can be made by plotting these data as a function of contrast (Figure 5B). As predicted by the nonlinear model (Figure 5B, curves), the effect of increasing contrast depends on stimulus size: responses to large stimuli show strong saturation (Figure 5B, squares), whereas responses to small stimuli grow linearly (Figure 5B, circles). The model predicts these effects because only large, high contrast stimuli elicit large enough responses from the suppressive field to cause suppression. For small, low contrast stimuli, instead, the linear model is a good approximation. B 100 Response (spikes/s) A 80 60 40 20 0 0.50 4.00 32.00 Diameter (deg) 0.00 0.25 0.50 0.75 1.00 Contrast Figure 5. Predicting responses to novel stimuli in the example LGN cell. Stimuli are gratings varying in diameter and contrast, and responses are harmonic component of spike trains at grating temporal frequency. Curves show model predictions based on parameters as estimated in previous figures, not fitted to these data. A: Responses as function of diameter for different contrasts. B: Responses as function of contrast for different diameters. 3.4 M o d e l pe r f or m a nc e To assess model performance across neurons we calculate the percentage of variance in the data that is explained by the model (see Freeman et al., 2002 for methods). The model provides good fits to the data used to characterize the suppressive field (Figure 4), explaining more than 90% of the variance in the data for 9/13 cells (Figure 6A). Model parameters are then held fixed, and the model is used to predict responses to gratings of different contrast and diameter (Figure 5). The model performs well, explaining in 10/13 neurons above 90% of the variance in these novel data (Figure 6B, shaded histogram). The agreement between the quality of the fits and the quality of the predictions suggests that model parameters are well constrained and rules out a role of overfitting in determining the quality of the fits. To further confirm the performance of the model, in an additional 54 cells we ran a subset of the whole protocol, involving only the experiment for characterizing the receptive field (Figure 3), and the experiment involving gratings of different contrast and diameter (Figure 5). For these cells we estimate the suppressive field by fitting the model directly to the latter measurements. The model explains above 90% of the variance in these data in 20/54 neurons and more than 70% in 39/54 neurons (Figure 6B, white histogram). Considering the large size of the data set (more than 100 stimuli, requiring several hours of recordings per neuron) and the small number of free parameters (only 6 for the purpose of this work), the overall, quality of the model predictions is remarkable. Estimating the suppressive field A # cells 6 n=13 4 2 0 Size tuning at different contrasts 15 n=54 10 # cells B 5 0 0 50 100 Explained variance (%) Figure 6. Percentage of variance in data explained by model. A: Experiments to estimate the suppressive field. B: Experiments to test the model. Gray histogram shows quality of predictions. White histogram shows quality of fits. 4 Co n cl u si o n s The nonlinear model provides a unified description of visual processing in LGN neurons. Based on a fixed set of parameters, it can predict both linear properties (Figure 3), as well as nonlinear properties such as contrast saturation (Figure 4A) and masking (Figure 4B-D). Moreover, once the parameters are fixed, it predicts responses to novel stimuli (Figure 5). The model explains why responses are tuned for stimulus size at high contrast but not at low contrast, and it correctly predicts that only responses to large stimuli saturate with contrast, while responses to small stimuli grow linearly. The model implements a form of contrast gain control. A possible purpose for this gain control is to increase the range of contrast that can be transmitted given the limited dynamic range of single neurons. Divisive gain control may also play a role in population coding: a similar model applied to responses of primary visual cortex was shown to maximize independence of the responses across neurons (Schwartz and Simoncelli, 2001). We are working towards improving the model in two ways. First, we are characterizing the dynamics of the suppressive field, e.g. to predict how it responds to transient stimuli. Second, we are testing the assumption that the suppressive field computes root-mean-square contrast, a measure that solely depends on the secondorder moments of the light distribution. Our ultimate goal is to predict responses to complex stimuli such as those shown in Figure 1 and quantify to what degree the nonlinear model improves on the predictions of the linear model. Determining the role of visual nonlinearities under more natural stimulation conditions is also critical to understanding their function. The nonlinear model synthesizes more than 30 years of research. It is robust, tractable and generalizes to arbitrary stimuli. As a result it might serve as the new standard model of LGN responses. Because the nonlinearities we discussed are already present in the retina (Shapley and Victor, 1978), and tend to get stronger as one ascends the visual hierarchy (Sclar et al., 1990), it may also be used to study how responses take shape from one stage to another in the visual system. A c k n o w l e d g me n t s This work was supported by the Swiss National Science Foundation and by the James S McDonnell Foundation 21st Century Research Award in Bridging Brain, Mind & Behavior. References Bonds, A. B. (1989). Role of inhibition in the specification of orientation selectivity of cells in the cat striate cortex. Vis Neurosci 2, 41-55. Bonin, V., Mante, V., and Carandini, M. (2002). The contrast integration field of cat LGN neurons. Program No. 352.16. In Abstract Viewer/Itinerary Planner (Washington, DC, Society for Neuroscience). Cai, D., DeAngelis, G. C., and Freeman, R. D. (1997). Spatiotemporal receptive field organization in the lateral geniculate nucleus of cats and kittens. J Neurophysiol 78, 10451061. Cavanaugh, J. R., Bair, W., and Movshon, J. A. (2002). Selectivity and spatial distribution of signals from the receptive field surround in macaque v1 neurons. J Neurophysiol 88, 25472556. Cleland, B. G., Lee, B. B., and Vidyasagar, T. R. (1983). Response of neurons in the cat's lateral geniculate nucleus to moving bars of different length. J Neurosci 3, 108-116. Dan, Y., Atick, J. J., and Reid, R. C. (1996). Efficient coding of natural scenes in the lateral geniculate nucleus: experimental test of a computational theory. J Neurosci 16, 3351-3362. Enroth-Cugell, C., and Robson, J. G. (1966). The contrast sensitivity of retinal ganglion cells of the cat. J Physiol (Lond) 187, 517-552. Freeman, T., Durand, S., Kiper, D., and Carandini, M. (2002). Suppression without Inhibition in Visual Cortex. Neuron 35, 759. Heeger, D. J. (1992). Normalization of cell responses in cat striate cortex. Vis Neurosci 9, 181-197. Hubel, D., and Wiesel, T. N. (1961). Integrative action in the cat's lateral geniculate body. J Physiol (Lond) 155, 385-398. Jones, H. E., and Sillito, A. M. (1991). The length-response properties of cells in the feline dorsal lateral geniculate nucleus. J Physiol (Lond) 444, 329-348. Levick, W. R., Cleland, B. G., and Dubin, M. W. (1972). Lateral geniculate neurons of cat: retinal inputs and physiology. Invest Ophthalmol 11, 302-311. Mante, V., Bonin, V., and Carandini, M. (2002). Responses of cat LGN neurons to plaids and movies. Program No. 352.15. In Abstract Viewer/Itinerary Planner (Washington, DC, Society for Neuroscience). Rodieck, R. W. (1965). Quantitative analysis of cat retina ganglion cell response to visual stimuli. Vision Res 5, 583-601. Schwartz, O., and Simoncelli, E. P. (2001). Natural signal statistics and sensory gain control. Nat Neurosci 4, 819-825. Sclar, G., Maunsell, J. H. R., and Lennie, P. (1990). Coding of image contrast in central visual pathways of the macaque monkey. Vision Res 30, 1-10. Shapley, R. M., and Victor, J. D. (1978). The effect of contrast on the transfer properties of cat retinal ganglion cells. J Physiol 285, 275-298. Solomon, S. G., White, A. J., and Martin, P. R. (2002). Extraclassical receptive field properties of parvocellular, magnocellular, and koniocellular cells in the primate lateral geniculate nucleus. J Neurosci 22, 338-349.

5 0.46817517 10 nips-2003-A Low-Power Analog VLSI Visual Collision Detector

Author: Reid R. Harrison

Abstract: We have designed and tested a single-chip analog VLSI sensor that detects imminent collisions by measuring radially expansive optic flow. The design of the chip is based on a model proposed to explain leg-extension behavior in flies during landing approaches. A new elementary motion detector (EMD) circuit was developed to measure optic flow. This EMD circuit models the bandpass nature of large monopolar cells (LMCs) immediately postsynaptic to photoreceptors in the fly visual system. A 16 × 16 array of 2-D motion detectors was fabricated on a 2.24 mm × 2.24 mm die in a standard 0.5-µm CMOS process. The chip consumes 140 µW of power from a 5 V supply. With the addition of wide-angle optics, the sensor is able to detect collisions around 500 ms before impact in complex, real-world scenes. 1

6 0.46332237 106 nips-2003-Learning Non-Rigid 3D Shape from 2D Motion

7 0.44116405 67 nips-2003-Eye Micro-movements Improve Stimulus Detection Beyond the Nyquist Limit in the Peripheral Retina

8 0.4237504 56 nips-2003-Dopamine Modulation in a Basal Ganglio-Cortical Network of Working Memory

9 0.36481187 175 nips-2003-Sensory Modality Segregation

10 0.36282909 43 nips-2003-Bounded Invariance and the Formation of Place Fields

11 0.36115244 16 nips-2003-A Recurrent Model of Orientation Maps with Simple and Complex Cells

12 0.30373532 127 nips-2003-Mechanism of Neural Interference by Transcranial Magnetic Stimulation: Network or Single Neuron?

13 0.29773134 45 nips-2003-Circuit Optimization Predicts Dynamic Networks for Chemosensory Orientation in Nematode C. elegans

14 0.27327073 22 nips-2003-An Improved Scheme for Detection and Labelling in Johansson Displays

15 0.27185994 35 nips-2003-Attractive People: Assembling Loose-Limbed Models using Non-parametric Belief Propagation

16 0.25522003 130 nips-2003-Model Uncertainty in Classical Conditioning

17 0.25459516 185 nips-2003-The Doubly Balanced Network of Spiking Neurons: A Memory Model with High Capacity

18 0.24735837 25 nips-2003-An MCMC-Based Method of Comparing Connectionist Models in Cognitive Science

19 0.22682333 85 nips-2003-Human and Ideal Observers for Detecting Image Curves

20 0.22526693 51 nips-2003-Design of Experiments via Information Theory


similar papers computed by lda model

lda for this paper:

topicId topicWeight

[(0, 0.016), (11, 0.013), (29, 0.024), (30, 0.018), (35, 0.021), (40, 0.389), (53, 0.113), (64, 0.014), (69, 0.012), (71, 0.032), (76, 0.034), (85, 0.081), (91, 0.13)]

similar papers list:

simIndex simValue paperId paperTitle

same-paper 1 0.86817384 7 nips-2003-A Functional Architecture for Motion Pattern Processing in MSTd

Author: Scott A. Beardsley, Lucia M. Vaina

Abstract: Psychophysical studies suggest the existence of specialized detectors for component motion patterns (radial, circular, and spiral), that are consistent with the visual motion properties of cells in the dorsal medial superior temporal area (MSTd) of non-human primates. Here we use a biologically constrained model of visual motion processing in MSTd, in conjunction with psychophysical performance on two motion pattern tasks, to elucidate the computational mechanisms associated with the processing of widefield motion patterns encountered during self-motion. In both tasks discrimination thresholds varied significantly with the type of motion pattern presented, suggesting perceptual correlates to the preferred motion bias reported in MSTd. Through the model we demonstrate that while independently responding motion pattern units are capable of encoding information relevant to the visual motion tasks, equivalent psychophysical performance can only be achieved using interconnected neural populations that systematically inhibit non-responsive units. These results suggest the cyclic trends in psychophysical performance may be mediated, in part, by recurrent connections within motion pattern responsive areas whose structure is a function of the similarity in preferred motion patterns and receptive field locations between units. 1 In trod u ction A major challenge in computational neuroscience is to elucidate the architecture of the cortical circuits for sensory processing and their effective role in mediating behavior. In the visual motion system, biologically constrained models are playing an increasingly important role in this endeavor by providing an explanatory substrate linking perceptual performance and the visual properties of single cells. Single cell studies indicate the presence of complex interconnected structures in middle temporal and primary visual cortex whose most basic horizontal connections can impart considerable computational power to the underlying neural population [1, 2]. Combined psychophysical and computational studies support these findings Figure 1: a) Schematic of the graded motion pattern (GMP) task. Discrimination pairs of stimuli were created by perturbing the flow angle (φ) of each 'test' motion (with average dot speed, vav), by ±φp in the stimulus space spanned by radial and circular motions. b) Schematic of the shifted center-of-motion (COM) task. Discrimination pairs of stimuli were created by shifting the COM of the ‘test’ motion to the left and right of a central fixation point. For each motion pattern the COM was shifted within the illusory inner aperture and was never explicitly visible. and suggest that recurrent connections may play a significant role in encoding the visual motion properties associated with various psychophysical tasks [3, 4]. Using this methodology our goal is to elucidate the computational mechanisms associated with the processing of wide-field motion patterns encountered during self-motion. In the human visual motion system, psychophysical studies suggest the existence of specialized detectors for the motion pattern components (i.e., radial, circular and spiral motions) associated with self-motion [5, 6]. Neurophysiological studies reporting neurons sensitive to motion patterns in the dorsal medial superior temporal area (MSTd) support the existence of such mechanisms [7-10], and in conjunction with psychophysical studies suggest a strong link between the patterns of neural activity and motion-based perceptual performance [11, 12]. Through the combination of human psychophysical performance and biologically constrained modeling we investigate the computational role of simple recurrent connections within a population of MSTd-like units. Based on the known visual motion properties within MSTd we ask what neural structures are computationally sufficient to encode psychophysical performance on a series of motion pattern tasks. 2 M o t i o n pa t t e r n d i sc r i m i n a t i o n Using motion pattern stimuli consistent with previous studies [5, 6], we have developed a set of novel psychophysical tasks designed to facilitate a more direct comparison between human perceptual performance and the visual motion properties of cells in MSTd that have been found to underlie the discrimination of motion patterns [11, 12]. The psychophysical tasks, referred to as the graded motion pattern (GMP) and shifted center-of-motion (COM) tasks, are outlined in Fig. 1. Using a temporal two-alternative-forced-choice task we measured discrimination thresholds to global changes in the patterns of complex motion (GMP task), [13], and shifts in the center-of-motion (COM task). Stimuli were presented with central fixation using a constant stimulus paradigm and consisted of dynamic random dot displays presented in a 24o annular region (central 4o removed). In each task, the stimulus duration was randomly perturbed across presentations (440±40 msec) to control for timing-based cues, and dots moved coherently through a radial speed Figure 2: a) GMP thresholds across 8 'test' motions at two mean dot speeds for two observers. Performance varied continuously with thresholds for radial motions (φ=0, 180o) significantly lower than those for circular motions (φ=90,270o), (p<0.001; t(37)=3.39). b) COM thresholds at three mean dot speeds for two observers. As with the GMP task, performance varied continuously with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47). gradient in directions consistent with the global motion pattern presented. Discrimination thresholds were obtained across eight ‘test’ motions corresponding to expansion, contraction, CW and CCW rotation, and the four intermediate spiral motions. To minimize adaptation to specific motion patterns, opposing motions (e.g., expansion/ contraction) were interleaved across paired presentations. 2.1 Results Discrimination thresholds are reported here from a subset of the observer population consisting of three experienced psychophysical observers, one of which was naïve to the purpose of the psychophysical tasks. For each condition, performance is reported as the mean and standard error averaged across 8-12 thresholds. Across observers and dot speeds GMP thresholds followed a distinct trend in the stimulus space [13], with radial motions (expansion/contraction) significantly lower than circular motions (CW/CCW rotation), (p<0.001; t(37)=3.39), (Fig. 2a). While thresholds for the intermediate spiral motions were not significantly different from circular motions (p=0.223, t(60)=0.74), the trends across 'test' motions were well fit within the stimulus space (SB: r>0.82, SC: r>0.77) by sinusoids whose period and phase were 196 ± 10o and -72 ± 20o respectively (Fig. 1a). When the radial speed gradient was removed by randomizing the spatial distribution of dot speeds, threshold performance increased significantly across observers (p<0.05; t(17)=1.91), particularly for circular motions (p<0.005; t(25)=3.31), (data not shown). Such performance suggests a perceptual contribution associated with the presence of the speed gradient and is particularly interesting given the fact that the speed gradient did not contribute computationally relevant information to the task. However, the speed gradient did convey information regarding the integrative structure of the global motion field and as such suggests a preference of the underlying motion mechanisms for spatially structured speed information. Similar trends in performance were observed in the COM task across observers and dot speeds. Discrimination thresholds varied continuously as a function of the 'test' motion with thresholds for radial motions significantly lower than those for circular motions, (p<0.001; t(37)=4.47) and could be well fit by a sinusoidal trend line (e.g. SB at 3 deg/s: r>0.91, period = 178 ± 10 o and phase = -70 ± 25o), (Fig. 2b). 2.2 A local or global task? The consistency of the cyclic threshold profile in stimuli that restricted the temporal integration of individual dot motions [13], and simultaneously contained all directions of motion, generally argues against a primary role for local motion mechanisms in the psychophysical tasks. While the psychophysical literature has reported a wide variety of “local” motion direction anisotropies whose properties are reminiscent of the results observed here, e.g. [14], all would predict equivalent thresholds for radial and circular motions for a set of uniformly distributed and/or spatially restricted motion direction mechanisms. Together with the computational impact of the speed gradient and psychophysical studies supporting the existence of wide-field motion pattern mechanisms [5, 6], these results suggest that the threshold differences across the GMP and COM tasks may be associated with variations in the computational properties across a series of specialized motion pattern mechanisms. 3 A computational model The similarities between the motion pattern stimuli used to quantify human perception and the visual motion properties of cells in MSTd suggests that MSTd may play a computational role in the psychophysical tasks. To examine this hypothesis, we constructed a population of MSTd-like units whose visual motion properties were consistent with the reported neurophysiology (see [13] for details). Across the population, the distribution of receptive field centers was uniform across polar angle and followed a gamma distribution Γ(5,6) across eccenticity [7]. For each unit, visual motion responses followed a gaussian tuning profile as a function of the stimulus flow angle G( φ), (σi=60±30o; [10]), and the distance of the stimulus COM from the unit’s receptive field center Gsat(xi, yi, σs=19o), Eq. 1, such that its preferred motion response was position invariant to small shifts in the COM [10] and degraded continuously for large shifts [9]. Within the model, simulations were categorized according to the distribution of preferred motions represented across the population (one reported in MSTd and a uniform control). The first distribution simulated an expansion bias in which the density of preferred motions decreased symmetrically from expansions to contraction [10]. The second distribution simulated a uniform preference for all motions and was used as a control to quantify the effects of an expansion bias on psychophysical performance. Throughout the paper we refer to simulations containing these distributions as ‘Expansion-biased’ and ‘Uniform’ respectively. 3.1 Extracting perceptual estimates from the neural code For each stimulus presentation, the ith unit’s response was calculated as the average firing rate, Ri, from the product of its motion pattern and spatial tuning profiles, ( ) Ri = Rmax G min[φ − φi ] ,σ ti G sati (x− xi , y − y i ,σ s ) + P (λ = 12 ) (1) where Rmax is the maximum preferred stimulus response (spikes/s), min[ ] refers to the minimum angular distance between the stimulus flow angle φ and the unit’s preferred motion φi, Gsat is the unit’s spatial tuning profile saturated within the central 5±3o, σti and σs are the standard deviations of the unit’s motion pattern and Figure 3: Model vs. psychophysical performance for independently responding units. Model thresholds are reported as the average (±1 S.E.) across five simulated populations. a) GMP thresholds were highest for contracting motions and lowest for expanding motions across all Expansion-biased populations. b) Comparable trends in performance were observed for COM thresholds. Comparison with the Uniform control simulations in both tasks (2000 units shown here) indicates that thresholds closely followed the distribution of preferred motions simulated within the model. spatial tuning profiles respectively, (xi,yi) is the spatial location of the unit’s receptive field center, (x,y) is the spatial location of the stimulus COM, and P(λ=12) is the background activity simulated as an uncorrelated Poisson process. The psychophysical tasks were simulated using a modified center-of-gravity ^ approach to decode estimates of the stimulus properties, i.e. flow angle (φ ) and ˆ ˆ COM location in the visual field (x, y ) , from the neural population   ∑ xi Ri ∑ y i Ri v  , i , ∑ φ i Ri  ∑ Ri i   i i   i (xˆ, yˆ , φˆ) =  i∑ R  (2) v where φi is the unit vector in the stimulus space (Fig. 1a) corresponding to the unit’s preferred motion. For each set of paired stimuli, psychophysical judgments were made by comparing the estimated stimulus properties according to the discrimination criteria, specified in the psychophysical tasks. As with the psychophysical experiments, discrimination thresholds were computed using a leastsquares fit to percent correct performance across constant stimulus levels. 3.2 Simulation 1: Independent neural responses In the first series of simulations, GMP and COM thresholds were quantified across three populations (500, 1000, and 2000 units) of independently responding units for each simulated distribution (Expansion-biased and Uniform). Across simulations, both the range in thresholds and their trends across ‘test’ motions were compared with human psychophysical performance to quantify the effects of population size and an expansion biased preferred motion distribution on model performance. Over the psychophysical range of interest (φp ± 7o), GMP thresholds for contracting motions were at chance across all Expansion-biased populations, (Fig. 3a). While thresholds for expanding motions were generally consistent with those for human observers, those for circular motions remained significantly higher for all but the largest populations. Similar trends in performance were observed for the COM task, (Fig. 3b). Here the range of COM thresholds was well matched with human performance for simulations containing 1000 units, however, the trends across motion patterns remained inconsistent even for the largest populations. Figure 4: Proposed recurrent connection profile between motion pattern units. a) Across the motion pattern space connection strength followed an inverse gaussian profile such that the ith unit (with preferred motion φi) systematically inhibited units with anti-preferred motions centered at 180+φi. b) Across the visual field connection strength followed a difference-of-gaussians profile as a function of the relative distance between receptive field centers such that spatially local units are mutually excitatory (σRe=10o) and more distant units were mutually inhibitory (σRi=80o). For simulations containing a uniform distribution of preferred motions, the threshold range was consistent with human performance on both tasks, however, the trend across motion patterns was generally flat. What variability did occur was due primarily to the discrete sampling of preferred motions across the population. Comparison of the discrimination thresholds for the Expansion-biased and Uniform populations indicates that the trend across thresholds was closely matched to the underlying distributions of preferred motions. This result in due in part to the nearequal weighting of independently responding units and can be explained to a first approximation by the proportional increase in the signal-to-noise ratio across the population as a function of the density of units responsive to a given 'test' motion. 3.3 Simulation 2: An interconnected neural structure In a second series of simulations, we examined the computational effect of adding recurrent connections between units. If the distribution of preferred motions in MSTd is in fact biased towards expansions, as the neurophysiology suggests, it seems unlikely that independent estimates of the visual motion information would be sufficient to yield the threshold profiles observed in the psychophysical tasks. We hypothesize that a simple fixed architecture of excitatory and/or inhibitory connections is sufficient to account for the cyclic trends in discrimination thresholds. Specifically, we propose that a recurrent connection profile whose strength varies as a function of (a) the similarity between preferred motion patterns and (b) the distance between receptive field centers, is computationally sufficient to recover the trends in GMP/COM performance (Fig. 4), wij = S R e − ( xi − x j )2 + ( yi − y j )2 2 2σ R e − SR e 2 − −(min[ φi − φ j ])2 ( xi − x j )2 + ( yi − y j )2 2 2 σ Ri − Sφ e 2σ I2 (3) Figure 5: Model vs. psychophysical performance for populations containing recurrent connections (σI=80o). As the number of units increased for Expansionbiased populations, discrimination thresholds decreased to psychophysical levels and the sinusoidal trend in thresholds emerged for both the (a) GMP and (b) COM tasks. Sinusoidal trends were established for as few as 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were (193.8 ± 11.7o, -70.0 ± 22.6o) and (168.2 ± 13.7o, -118.8 ± 31.8o) for the GMP and COM tasks respectively. where wij is the strength of the recurrent connection between ith and jth units, (xi,yi) and (xj,yj) denote the spatial locations of their receptive field centers, σRe (=10o) and σRi (=80o) together define the spatial extent of a difference-of-gaussians interaction between receptive field centers, and SR and Sφ scale the connection strength. To examine the effects of the spread of motion pattern-specific inhibition and connection strength in the model, σI, Sφ, and SR were considered free parameters. Within the parameter space used to define recurrent connections (i.e., σI, Sφ and SR), Monte Carlo simulations of Expansion-biased model performance (1000 units) yielded regions of high correlation on both tasks (with respect to the psychophysical thresholds, r>0.7) that were consistent across independently simulated populations. Typically these regions were well defined over a broad range such that there was significant overlap between tasks (e.g., for the GMP task (SR=0.03), σI=[45,120o], Sφ=[0.03,0.3] and for the COM task (σI=80o), Sφ = [0.03,0.08], SR = [0.005, 0.04]). Fig. 5 shows averaged threshold performance for simulations of interconnected units drawn from the highly correlated regions of the (σI, Sφ, SR) parameter space. For populations not explicitly examined in the Monte Carlo simulations connection strengths (Sφ, SR) were scaled inversely with population size to maintain an equivalent level of recurrent activity. With the incorporation of recurrent connections, the sinusoidal trend in GMP and COM thresholds emerged for Expansion-biased populations as the number of units increased. In both tasks the cyclic threshold profiles were established for 1000 units and were well fit (r>0.9) by sinusoids whose periods and phases were consistent with human performance. Unlike the Expansion-biased populations, Uniform populations were not significantly affected by the presence of recurrent connections (Fig. 5). Both the range in thresholds and the flat trend across motion patterns were well matched to those in Section 3.2. Together these results suggest that the sinusoidal trends in GMP and COM performance may be mediated by the combined contribution of the recurrent interconnections and the bias in preferred motions across the population. 4 D i s c u s s i on Using a biologically constrained computational model in conjunction with human psychophysical performance on two motion pattern tasks we have shown that the visual motion information encoded across an interconnected population of cells responsive to motion patterns, such as those in MSTd, is computationally sufficient to extract perceptual estimates consistent with human performance. Specifically, we have shown that the cyclic trend in psychophysical performance observed across tasks, (a) cannot be reproduced using populations of independently responding units and (b) is dependent, in part, on the presence of an expanding motion bias in the distribution of preferred motions across the neural population. The model’s performance suggests the presence of specific recurrent structures within motion pattern responsive areas, such as MSTd, whose strength varies as a function of the similarity between preferred motion patterns and the distance between receptive field centers. While such structures have not been explicitly examined in MSTd and other higher visual motion areas there is anecdotal support for the presence of inhibitory connections [8]. Together, these results suggest that robust processing of the motion patterns associated with self-motion and optic flow may be mediated, in part, by recurrent structures in extrastriate visual motion areas whose distributions of preferred motions are biased strongly in favor of expanding motions. Acknowledgments This work was supported by National Institutes of Health grant EY-2R01-07861-13 to L.M.V. References [1] Malach, R., Schirman, T., Harel, M., Tootell, R., & Malonek, D., (1997), Cerebral Cortex, 7(4): 386-393. [2] Gilbert, C. D., (1992), Neuron, 9: 1-13. [3] Koechlin, E., Anton, J., & Burnod, Y., (1999), Biological Cybernetics, 80: 2544. [4] Stemmler, M., Usher, M., & Niebur, E., (1995), Science, 269: 1877-1880. [5] Burr, D. C., Morrone, M. C., & Vaina, L. M., (1998), Vision Research, 38(12): 1731-1743. [6] Meese, T. S. & Harris, S. J., (2002), Vision Research, 42: 1073-1080. [7] Tanaka, K. & Saito, H. A., (1989), Journal of Neurophysiology, 62(3): 626-641. [8] Duffy, C. J. & Wurtz, R. H., (1991), Journal of Neurophysiology, 65(6): 13461359. [9] Duffy, C. J. & Wurtz, R. H., (1995), Journal of Neuroscience, 15(7): 5192-5208. [10] Graziano, M. S., Anderson, R. A., & Snowden, R., (1994), Journal of Neuroscience, 14(1): 54-67. [11] Celebrini, S. & Newsome, W., (1994), Journal of Neuroscience, 14(7): 41094124. [12] Celebrini, S. & Newsome, W. T., (1995), Journal of Neurophysiology, 73(2): 437-448. [13] Beardsley, S. A. & Vaina, L. M., (2001), Journal of Computational Neuroscience, 10: 255-280. [14] Matthews, N. & Qian, N., (1999), Vision Research, 39: 2205-2211.

2 0.78078222 16 nips-2003-A Recurrent Model of Orientation Maps with Simple and Complex Cells

Author: Paul Merolla, Kwabena A. Boahen

Abstract: We describe a neuromorphic chip that utilizes transistor heterogeneity, introduced by the fabrication process, to generate orientation maps similar to those imaged in vivo. Our model consists of a recurrent network of excitatory and inhibitory cells in parallel with a push-pull stage. Similar to a previous model the recurrent network displays hotspots of activity that give rise to visual feature maps. Unlike previous work, however, the map for orientation does not depend on the sign of contrast. Instead, signindependent cells driven by both ON and OFF channels anchor the map, while push-pull interactions give rise to sign-preserving cells. These two groups of orientation-selective cells are similar to complex and simple cells observed in V1. 1 Orientation Maps Neurons in visual areas 1 and 2 (V1 and V2) are selectively tuned for a number of visual features, the most pronounced feature being orientation. Orientation preference of individual cells varies across the two-dimensional surface of the cortex in a stereotyped manner, as revealed by electrophysiology [1] and optical imaging studies [2]. The origin of these preferred orientation (PO) maps is debated, but experiments demonstrate that they exist in the absence of visual experience [3]. To the dismay of advocates of Hebbian learning, these results suggest that the initial appearance of PO maps rely on neural mechanisms oblivious to input correlations. Here, we propose a model that accounts for observed PO maps based on innate noise in neuron thresholds and synaptic currents. The network is implemented in silicon where heterogeneity is as ubiquitous as it is in biology. 2 Patterned Activity Model Ernst et al. have previously described a 2D rate model that can account for the origin of visual maps [4]. Individual units in their network receive isotropic feedforward input from the geniculate and recurrent connections from neighboring units in a Mexican hat profile, described by short-range excitation and long-range inhibition. If the recurrent connections are sufficiently strong, hotspots of activity (or ‘bumps’) form periodically across space. In a homogeneous network, these bumps of activity are equally stable at any position in the network and are free to wander. Introducing random jitter to the Mexican hat connectivity profiles breaks the symmetry and reduces the number of stable states for the bumps. Subsequently, the bumps are pinned down at the locations that maximize their net local recurrent feedback. In this regime, moving gratings are able to shift the bumps away from their stability points such that the responses of the network resemble PO maps. Therefore, the recurrent network, given an ample amount of noise, can innately generate its own orientation specificity without the need for specific hardwired connections or visually driven learning rules. 2.1 Criticisms of the Bump model We might posit that the brain uses a similar opportunistic model to derive and organize its feature maps – but the parallels between the primary visual cortex and the Ernst et al. bump model are unconvincing. For instance, the units in their model represent the collective activity of a column, reducing the network dynamics to a firing-rate approximation. But this simplification ignores the rich temporal dynamics of spiking networks, which are known to affect bump stability. More fundamentally, there is no role for functionally distinct neuron types. The primary criticism of the Ernst et al.’s bump model is that its input only consists of a luminance channel, and it is not obvious how to replace this channel with ON and OFF rectified channels to account for simple and complex cells. One possibility would be to segregate ON-driven and OFF-driven cells (referred to as simple cells in this paper) into two distinct recurrent networks. Because each network would have its own innate noise profile, bumps would form independently. Consequently, there is no guarantee that ON-driven maps would line up with OFF-driven maps, which would result in conflicting orientation signals when these simple cells converge onto sign-independent (complex) cells. 2.2 Simple Cells Solve a Complex Problem To ensure that both ON-driven and OFF-driven simple cells have the same orientation maps, both ON and OFF bumps must be computed in the same recurrent network so that they are subjected to the same noise profile. We achieve this by building our recurrent network out of cells that are sign-independent; that is both ON and OFF channels drive the network. These cells exhibit complex cell-like behavior (and are referred to as complex cells in this paper) because they are modulated at double the spatial frequency of a sinusoidal grating input. The simple cells subsequently derive their responses from two separate signals: an orientation selective feedback signal from the complex cells indicating the presence of either an ON or an OFF bump, and an ON–OFF selection signal that chooses the appropriate response flavor. Figure 1 left illustrates the formation of bumps (highlighted cells) by a recurrent network with a Mexican hat connectivity profile. Extending the Ernst et al. model, these complex bumps seed simple bumps when driven by a grating. Simple bumps that match the sign of the input survive, whereas out-of-phase bumps are extinguished (faded cells) by push-pull inhibition. Figure 1 right shows the local connections within a microcircuit. An EXC (excitatory) cell receives excitatory input from both ON and OFF channels, and projects to other EXC (not shown) and INH (inhibitory) cells. The INH cell projects back in a reciprocal configuration to EXC cells. The divergence is indicated in left. ON-driven and OFF-driven simple cells receive input in a push-pull configuration (i.e., ON cells are excited by ON inputs and inhibited by OFF inputs, and vise-versa), while additionally receiving input from the EXC–INH recurrent network. In this model, we implement our push-pull circuit using monosynaptic inhibitory connections, despite the fact that geniculate input is strictly excitatory. This simplification, while anatomically incorrect, yields a more efficient implementation that is functionally equivalent. ON Input Luminance OFF Input left right EXC EXC Divergence INH INH Simple Cells Complex Cells ON & OFF Input ON OFF OFF Space Figure 1: left, Complex and simple cell responses to a sinusoidal grating input. Luminance is transformed into ON (green) and OFF (red) pathways by retinal processing. Complex cells form a recurrent network through excitatory and inhibitory projections (yellow and blue lines, respectively), and clusters of activity occur at twice the spatial frequency of the grating. ON input activates ON-driven simple cells (bright green) and suppresses OFF-driven simple cells (faded red), and vise-versa. right, The bump model’s local microcircuit: circles represent neurons, curved lines represent axon arbors that end in excitatory synapses (v shape) or inhibitory synapses (open circles). For simplicity, inhibitory interneurons were omitted in our push-pull circuit. 2.3 Mathematical Description • The neurons in our network follow the equation CV = −∑ ∂(t − tn) + I syn − I KCa − I leak , • n where C is membrane capacitance, V is the temporal derivative of the membrane voltage, δ(·) is the Dirac delta function, which resets the membrane at the times tn when it crosses threshold, Isyn is synaptic current from the network, and Ileak is a constant leak current. Neurons receive synaptic current of the form: ON I syn = w+ I ON − w− I OFF + wEE I EXC − wEI I INH , EXC I syn = w+ ( I ON + I OFF ) + wEE I EXC − wEI I INH + I back , OFF INH I syn = w+ I OFF − w− I ON + wEE I EXC − wEI I INH , I syn = wIE I EXC where w+ is the excitatory synaptic strength for ON and OFF input synapses, w- is the strength of the push-pull inhibition, wEE is the synaptic strength for EXC cell projections to other EXC cells, wEI is the strength of INH cell projections to EXC cells, wIE is the strength of EXC cell projections to INH cells, Iback is a constant input current, and I{ON,OFF,EXC,INH} account for all impinging synapses from each of the four cell types. These terms are calculated for cell i using an arbor function that consists of a spatial weighting J(r) and a post-synaptic current waveform α(t): k ∑ J (i − k ) ⋅ α (t − t n ) , where k spans all cells of a given type and n indexes their spike k ,n times. The spatial weighting function is described by J (i − k ) = exp( − i − k σ ) , with σ as the space constant. The current waveform, which is non-zero for t>0, convolves a 1 t function with a decaying exponential: α (t ) = (t τ c + α 0 ) −1 ∗ exp(− t τ e ) , where τc is the decay-rate, and τe is the time constant of the exponential. Finally, we model spike-rate adaptation with a calcium-dependent potassium-channel (KCa), which integrates Ca triggered by spikes at times tn with a gain K and a time constant τk, as described by I KCa = ∑ K exp(tn − t τ k ) . n 3 Silicon Implementation We implemented our model in silicon using the TSMC (Taiwan Semiconductor Manufacturing Company) 0.25µm 5-metal layer CMOS process. The final chip consists of a 2-D core of 48x48 pixels, surrounded by asynchronous digital circuitry that transmits and receives spikes in real-time. Neurons that reach threshold within the array are encoded as address-events and sent off-chip, and concurrently, incoming address-events are sent to their appropriate synapse locations. This interface is compatible with other spike-based chips that use address-events [5]. The fabricated bump chip has close to 460,000 transistors packed in 10 mm2 of silicon area for a total of 9,216 neurons. 3.1 Circuit Design Our neural circuit was morphed into hardware using four building blocks. Figure 2 shows the transistor implementation for synapses, axonal arbors (diffuser), KCa analogs, and neurons. The circuits are designed to operate in the subthreshold region (except for the spiking mechanism of the neuron). Noise is not purposely designed into the circuits. Instead, random variations from the fabrication process introduce significant deviations in I-V curves of theoretically identical MOS transistors. The function of the synapse circuit is to convert a brief voltage pulse (neuron spike) into a postsynaptic current with biologically realistic temporal dynamics. Our synapse achieves this by cascading a current-mirror integrator with a log-domain low-pass filter. The current-mirror integrator has a current impulse response that decays as 1 t (with a decay rate set by the voltage τc and an amplitude set by A). This time-extended current pulse is fed into a log-domain low-pass filter (equivalent to a current-domain RC circuit) that imposes a rise-time on the post-synaptic current set by τe. ON and OFF input synapses receive presynaptic spikes from the off-chip link, whereas EXC and INH synapses receive presynaptic spikes from local on-chip neurons. Synapse Je Diffuser Ir A Ig Jc KCa Analog Neuron Jk Vmem Vspk K Figure 2: Transistor implementations are shown for a synapse, diffuser, KCa analog, and neuron (simplified), with circuit insignias in the top-left of each box. The circuits they interact with are indicated (e.g. the neuron receives synaptic current from the diffuser as well as adaptation current from the KCa analog; the neuron in turn drives the KCa analog). The far right shows layout for one pixel of the bump chip (vertical dimension is 83µm, horizontal is 30 µm). The diffuser circuit models axonal arbors that project to a local region of space with an exponential weighting. Analogous to resistive divider networks, diffusers [6] efficiently distribute synaptic currents to multiple targets. We use four diffusers to implement axonal projections for: the ON pathway, which excites ON and EXC cells and inhibits OFF cells; the OFF pathway, which excites OFF and EXC cells and inhibits ON cells; the EXC cells, which excite all cell types; and the INH cells, which inhibits EXC, ON, and OFF cells. Each diffuser node connects to its six neighbors through transistors that have a pseudo-conductance set by σr, and to its target site through a pseudo-conductance set by σg; the space-constant of the exponential synaptic decay is set by σr and σg’s relative levels. The neuron circuit integrates diffuser currents on its membrane capacitance. Diffusers either directly inject current (excitatory), or siphon off current (inhibitory) through a current-mirror. Spikes are generated by an inverter with positive feedback (modified from [7]), and the membrane is subsequently reset by the spike signal. We model a calcium concentration in the cell with a KCa analog. K controls the amount of calcium that enters the cell per spike; the concentration decays exponentially with a time constant set by τk. Elevated charge levels activate a KCa-like current that throttles the spike-rate of the neuron. 3.2 Experimental Setup Our setup uses either a silicon retina [8] or a National Instruments DIO (digital input–output) card as input to the bump chip. This allows us to test our V1 model with real-time visual stimuli, similar to the experimental paradigm of electrophysiologists. More specifically, the setup uses an address-event link [5] to establish virtual point-to-point connectivity between ON or OFF ganglion cells from the retina chip (or DIO card) with ON or OFF synapses on the bump chip. Both the input activity and the output activity of the bump chip is displayed in real-time using receiver chips, which integrate incoming spikes and displays their rates as pixel intensities on a monitor. A logic analyzer is used to capture spike output from the bump chip so it can be further analyzed. We investigated responses of the bump chip to gratings moving in sixteen different directions, both qualitatively and quantitatively. For the qualitative aspect, we created a PO map by taking each cell’s average activity for each stimulus direction and computing the vector sum. To obtain a quantitative measure, we looked at the normalized vector magnitude (NVM), which reveals the sharpness of a cell’s tuning. The NVM is calculated by dividing the vector sum by the magnitude sum for each cell. The NVM is 0 if a cell responds equally to all orientations, and 1 if a cell’s orientation selectivity is perfect such that it only responds at a single orientation. 4 Results We presented sixteen moving gratings to the network, with directions ranging from 0 to 360 degrees. The spatial frequency of the grating is tuned to match the size of the average bump, and the temporal frequency is 1 Hz. Figure 3a shows a resulting PO map for directions from 180 to 360 degrees, looking at the inhibitory cell population (the data looks similar for other cell types). Black contours represent stable bump regions, or equivalently, the regions that exceed a prescribed threshold (90 spikes) for all directions. The PO map from the bump chip reveals structure that resembles data from real cortex. Nearby cells tend to prefer similar orientations except at fractures. There are even regions that are similar to pinwheels (delimited by a white rectangle). A PO is a useful tool to describe a network’s selectivity, but it only paints part of the picture. So we have additionally computed a NVM map and a NVM histogram, shown in Figure 3b and 3c respectively. The NVM map shows that cells with sharp selectivity tend to cluster, particularly around the edge of the bumps. The histogram also reveals that the distribution of cell selectivity across the network varies considerably, skewed towards broadly tuned cells. We also looked at spike rasters from different cell-types to gain insight into their phase relationship with the stimulus. In particular, we present recordings for the site indicated by the arrow (see Figure 3a) for gratings moving in eight directions ranging from 0 to 360 degrees in 45-degree increments (this location was chosen because it is in the vicinity of a pinwheel, is reasonably selective, and shows considerable modulation in its firing rate). Figure 4 shows the luminance of the stimulus (bottom sinusoids), ON- (cyan) and OFF-input (magenta) spike trains, and the resulting spike trains from EXC (yellow), INH (blue), ON- (green), and OFFdriven (red) cell types for each of the eight directions. The center polar plot summarizes the orientation selectivity for each cell-type by showing the normalized number of spikes for each stimulus. Data is shown for one period. Even though all cells-types are selective for the same orientation (regardless of grating direction), complex cell responses tend to be phase-insensitive while the simple cell responses are modulated at the fundamental frequency. It is worth noting that the simple cells have sharper orientation selectivity compared to the complex cells. This trend is characteristic of our data. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 300 250 200 150 100 50 20 40 60 80 100 120 140 160 180 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Figure 3: (a) PO map for the inhibitory cell population stimulated with eight different directions from 180 to 360 degrees (black represents no activity, contours delineate regions that exceed 90 spikes for all stimuli). Normalized vector magnitude (NVM) data is presented as (b) a map and (c) a histogram. Figure 4: Spike rasters and polar plot for 8 directions ranging from 0 to 360 degrees. Each set of spike rasters represent from bottom to top, ON- (cyan) and OFF-input (magenta), INH (yellow), EXC (blue), and ON- (green) and OFF-driven (red). The stimulus period is 1 sec. 5 Discussion We have implemented a large-scale network of spiking neurons in a silicon chip that is based on layer 4 of the visual cortex. The initial testing of the network reveals a PO map, inherited from innate chip heterogeneities, resembling cortical maps. Our microcircuit proposes a novel function for complex-like cells; that is they create a sign-independent orientation selective signal, which through a push-pull circuit creates sharply tuned simple cells with the same orientation preference. Recently, Ringach et al. surveyed orientation selectivity in the macaque [9]. They observed that, in a population of V1 neurons (N=308) the distribution of orientation selectivity is quite broad, having a median NVM of 0.39. We have measured median NVM’s ranging from 0.25 to 0.32. Additionally, Ringach et al. found a negative correlation between spontaneous firing rate and NVM. This is consistent with our model because cells closer to the center of the bump have higher firing rates and broader tuning. While the results from the bump chip are promising, our maps are less consistent and noisier than the maps Ernst et al. have reported. We believe this is because our network is tuned to operate in a fluid state where bumps come on, travel a short distance and disappear (motivated by cortical imaging studies). But excessive fluidity can cause non-dominant bumps to briefly appear and adversely shift the PO maps. We are currently investigating the role of lateral connections between bumps as a means to suppress these spontaneous shifts. The neural mechanisms that underlie the orientation selectivity of V1 neurons are still highly debated. This may be because neuron responses are not only shaped by feedforward inputs, but are also influenced at the network level. If modeling is going to be a useful guide for electrophysiologists, we must model at the network level while retaining cell level detail. Our results demonstrate that a spike-based neuromorphic system is well suited to model layer 4 of the visual cortex. The same approach may be used to build large-scale models of other cortical regions. References 1. Hubel, D. and T. Wiesel, Receptive firelds, binocular interaction and functional architecture in the cat's visual cortex. J. Physiol, 1962. 160: p. 106-154. 2. Blasdel, G.G., Orientation selectivity, preference, and continuity in monkey striate cortex. J Neurosci, 1992. 12(8): p. 3139-61. 3. Crair, M.C., D.C. Gillespie, and M.P. Stryker, The role of visual experience in the development of columns in cat visual cortex. Science, 1998. 279(5350): p. 566-70. 4. Ernst, U.A., et al., Intracortical origin of visual maps. Nat Neurosci, 2001. 4(4): p. 431-6. 5. Boahen, K., Point-to-Point Connectivity. IEEE Transactions on Circuits & Systems II, 2000. vol 47 no 5: p. 416-434. 6. Boahen, K. and Andreou. A contrast sensitive silicon retina with reciprocal synapses. in NIPS91. 1992: IEEE. 7. Culurciello, E., R. Etienne-Cummings, and K. Boahen, A Biomorphic Digital Image Sensor. IEEE Journal of Solid State Circuits, 2003. vol 38 no 2: p. 281-294. 8. Zaghloul, K., A silicon implementation of a novel model for retinal processing, in Neuroscience. 2002, UPENN: Philadelphia. 9. Ringach, D.L., R.M. Shapley, and M.J. Hawken, Orientation selectivity in macaque V1: diversity and laminar dependence. J Neurosci, 2002. 22(13): p. 5639-51.

3 0.44231415 4 nips-2003-A Biologically Plausible Algorithm for Reinforcement-shaped Representational Learning

Author: Maneesh Sahani

Abstract: Significant plasticity in sensory cortical representations can be driven in mature animals either by behavioural tasks that pair sensory stimuli with reinforcement, or by electrophysiological experiments that pair sensory input with direct stimulation of neuromodulatory nuclei, but usually not by sensory stimuli presented alone. Biologically motivated theories of representational learning, however, have tended to focus on unsupervised mechanisms, which may play a significant role on evolutionary or developmental timescales, but which neglect this essential role of reinforcement in adult plasticity. By contrast, theoretical reinforcement learning has generally dealt with the acquisition of optimal policies for action in an uncertain world, rather than with the concurrent shaping of sensory representations. This paper develops a framework for representational learning which builds on the relative success of unsupervised generativemodelling accounts of cortical encodings to incorporate the effects of reinforcement in a biologically plausible way. 1

4 0.4388389 65 nips-2003-Extending Q-Learning to General Adaptive Multi-Agent Systems

Author: Gerald Tesauro

Abstract: Recent multi-agent extensions of Q-Learning require knowledge of other agents’ payoffs and Q-functions, and assume game-theoretic play at all times by all other agents. This paper proposes a fundamentally different approach, dubbed “Hyper-Q” Learning, in which values of mixed strategies rather than base actions are learned, and in which other agents’ strategies are estimated from observed actions via Bayesian inference. Hyper-Q may be effective against many different types of adaptive agents, even if they are persistently dynamic. Against certain broad categories of adaptation, it is argued that Hyper-Q may converge to exact optimal time-varying policies. In tests using Rock-Paper-Scissors, Hyper-Q learns to significantly exploit an Infinitesimal Gradient Ascent (IGA) player, as well as a Policy Hill Climber (PHC) player. Preliminary analysis of Hyper-Q against itself is also presented. 1

5 0.43531907 20 nips-2003-All learning is Local: Multi-agent Learning in Global Reward Games

Author: Yu-han Chang, Tracey Ho, Leslie P. Kaelbling

Abstract: In large multiagent games, partial observability, coordination, and credit assignment persistently plague attempts to design good learning algorithms. We provide a simple and efficient algorithm that in part uses a linear system to model the world from a single agent’s limited perspective, and takes advantage of Kalman filtering to allow an agent to construct a good training signal and learn an effective policy. 1

6 0.4321427 161 nips-2003-Probabilistic Inference in Human Sensorimotor Processing

7 0.43081874 107 nips-2003-Learning Spectral Clustering

8 0.42881781 57 nips-2003-Dynamical Modeling with Kernels for Nonlinear Time Series Prediction

9 0.42874601 68 nips-2003-Eye Movements for Reward Maximization

10 0.4253628 73 nips-2003-Feature Selection in Clustering Problems

11 0.42449337 79 nips-2003-Gene Expression Clustering with Functional Mixture Models

12 0.4244498 49 nips-2003-Decoding V1 Neuronal Activity using Particle Filtering with Volterra Kernels

13 0.42442897 24 nips-2003-An Iterative Improvement Procedure for Hierarchical Clustering

14 0.42395657 93 nips-2003-Information Dynamics and Emergent Computation in Recurrent Circuits of Spiking Neurons

15 0.42290592 30 nips-2003-Approximability of Probability Distributions

16 0.42224532 125 nips-2003-Maximum Likelihood Estimation of a Stochastic Integrate-and-Fire Neural Model

17 0.42181322 78 nips-2003-Gaussian Processes in Reinforcement Learning

18 0.41946754 104 nips-2003-Learning Curves for Stochastic Gradient Descent in Linear Feedforward Networks

19 0.41849375 54 nips-2003-Discriminative Fields for Modeling Spatial Dependencies in Natural Images

20 0.4179171 126 nips-2003-Measure Based Regularization