Report

Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback Highlights d

High-field fMRI can resolve neural activity with laminar specificity in humans

d

Top-down and bottom-up signals lead to distinct laminar activation profiles in V1

Authors Peter Kok, Lauren J. Bains, Tim van Mourik, David G. Norris, Floris P. de Lange

Correspondence

d

Top-down signals selectively activate deep layers of V1

[email protected]

d

Bottom-up stimulus input activates all layers of V1

In Brief Using high-field fMRI, Kok et al. show that feedback signals evoked by a visual illusion selectively activate the deep layers of the primary visual cortex, demonstrating the potential for noninvasive in vivo recordings of neural activity with laminar specificity in humans.

Kok et al., 2016, Current Biology 26, 1–6 February 8, 2016 ª2016 Elsevier Ltd All rights reserved http://dx.doi.org/10.1016/j.cub.2015.12.038

Please cite this article in press as: Kok et al., Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback, Current Biology (2016), http://dx.doi.org/10.1016/j.cub.2015.12.038

Current Biology

Report Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback Peter Kok,1,* Lauren J. Bains,1,2 Tim van Mourik,1 David G. Norris,1,2,3 and Floris P. de Lange1 1Donders

Institute for Brain, Cognition and Behaviour, Radboud University, Kapittelweg 29, 6525 EN Nijmegen, the Netherlands L. Hahn Institute for Magnetic Resonance Imaging, UNESCO-Weltkulturerbe Zollverein, Leitstand Kokerei Zollverein, Kokereiallee 7, 45141 Essen, Germany 3MIRA Institute for Biomedical Technology and Technical Medicine, University of Twente, 7500 AE Enschede, the Netherlands *Correspondence: [email protected] http://dx.doi.org/10.1016/j.cub.2015.12.038 2Erwin

SUMMARY

In addition to bottom-up input, the visual cortex receives large amounts of feedback from other cortical areas [1–3]. One compelling example of feedback activation of early visual neurons in the absence of bottom-up input occurs during the famous Kanizsa illusion, where a triangular shape is perceived, even in regions of the image where there is no bottom-up visual evidence for it. This illusion increases the firing activity of neurons in the primary visual cortex with a receptive field on the illusory contour [4]. Feedback signals are largely segregated from feedforward signals within each cortical area, with feedforward signals arriving in the middle layer, while top-down feedback avoids the middle layers and predominantly targets deep and superficial layers [1, 2, 5, 6]. Therefore, the feedback-mediated activity increase in V1 during the perception of illusory shapes should lead to a specific laminar activity profile that is distinct from the activity elicited by bottom-up stimulation. Here, we used fMRI at high field (7 T) to empirically test this hypothesis, by probing the cortical response to illusory figures in human V1 at different cortical depths [7–14]. We found that, whereas bottom-up stimulation activated all cortical layers, feedback activity induced by illusory figures led to a selective activation of the deep layers of V1. These results demonstrate the potential for non-invasive recordings of neural activity with laminar specificity in humans and elucidate the role of top-down signals during perceptual processing. RESULTS We non-invasively examined the laminar activity profile of the human primary visual cortex (V1), using high-field (7 T) fMRI with high spatial resolution, in response to illusory figures (Figures 1A–1C). Specifically, we defined three equi-volume gray matter layers (superficial, middle, and deep) and determined

the proportion of each voxel’s volume in these layers (as well as in white matter and cerebrospinal fluid (CSF); Figures 1D and 1E). These layer ‘‘weights’’ were subsequently used in a spatial regression approach to determine layer-specific time courses of the blood-oxygen-level dependent (BOLD) signal in relevant parts of V1 [16]. We used a general linear model (GLM) of spatially distributed responses to unmix the signals from the different layers, even though these signals may be mixed in individual voxels (see Supplemental Experimental Procedures for technical details). The presence of an illusory triangle led to a selective increase of activity in regions of V1 whose receptive field was centered on the triangle (Figure 1C). This activation was only observed in the deep layers of V1 (t9 = 3.4, p = 0.0070), while the superficial and middle layers were not differentially activated (both t < 1, p > 0.50). On the other hand, bottom-up stimulation by a contrast-reversing checkerboard stimulus reliably activated all cortical layers, with most prominent activity observed in middle and superficial layers (all p < 0.005; Figure 2B). This qualitative difference in the laminar activity profile between bottom-up stimulation and top-down modulation led to a significant interaction between layer and condition (F2,18 = 4.1, p = 0.035), providing evidence for the differential activation of cortical layers by feedforward and feedback streams. In general, the BOLD time courses of the different gray matter layers were moderately correlated (r  0.4–0.5; Table S1). While this is likely in part due to neural activity propagating through the cortical column [17], leading to a co-activation of all the different layers, it may also partly reflect spurious correlations due to partial volume effects, as well as transfer of activation caused by blood flowing from deep to superficial layers. Therefore, in an additional analysis, we assessed each layer’s unique contribution by regressing out the time courses of the other two layers (see Supplemental Experimental Procedures). We found that the response to the illusory figure in the deep layers remained strongly present (t9 = 3.7, p = 0.0053; Figure S1A). In contrast, during bottom-up stimulation, the largest response was now observed in the middle layer (t9 = 5.0, p = 0.00072), while the deep and superficial layers did not show a response over and above the overall response (both t < 1, p > 0.30; Figure S1B). There was not only enhanced activity in regions of V1 whose receptive field was on the illusory triangle but also reduced activity in regions of V1 with receptive fields on the surrounding

Current Biology 26, 1–6, February 8, 2016 ª2016 Elsevier Ltd All rights reserved 1

Please cite this article in press as: Kok et al., Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback, Current Biology (2016), http://dx.doi.org/10.1016/j.cub.2015.12.038

Figure 1. Experimental Methods (A) Example Kanizsa stimulus, in which three out of four Pac-Man inducers were aligned such that an illusory triangle could be perceived. A fixation bull’s eye, containing a rapid letter stream, was presented 1.5 below and to the right of the center of the configuration. During the experiment, participants performed a target letter detection task. (B) In control configurations, the inducers were not aligned, and consequently, no illusory figure could be perceived. (C) Reconstruction of the difference in BOLD response evoked by illusory triangle configurations versus control configurations, averaged over participants. This reconstruction shows BOLD signals that have been transformed from cortical (i.e., voxel) space to visual (i.e., stimulus) space. This was achieved by multiplying each voxel’s population receptive field (a 2D Gaussian with center (x0, y0) and size s) by that voxel’s differential BOLD response (illusory figure versus control) and then summing over all voxels in a region (here, V1). Here, we collapsed over the four different illusory triangles by first rotating the reconstructions of the upward-, rightward-, and downward-pointing triangles 90 , 180 , and 270 counterclockwise, respectively, to move them into the reference frame of the leftward-pointing triangle, before averaging. See Supplemental Experimental Procedures (‘‘Retinotopic reconstruction of effects’’) for details. Dashed white triangles indicate the location of the illusory triangles. The increased BOLD response to the illusory figure and decreased response to the surrounding Pac-Man inducers replicated the results of [15]. (D) Sagittal view of the mean functional volume of a representative subject, overlaid with the boundaries between gray matter and CSF (red) and white matter (yellow). The black grid indicates the position and size (0.8 mm isotropic) of the functional voxels; green squares indicate visually active voxels (i.e., >20% variance explained during population receptive field (pRF) fitting; see Supplemental Experimental Procedures). (E) Schematic example of a voxel (red square) and the distribution of its volume over the three gray matter layers. This layer-volume distribution was determined for each voxel and used as the basis of a regression approach in order to obtain layer-specific BOLD time courses (see Supplemental Experimental Procedures).

inducer stimuli (t9 = 2.4, p = 0.042, collapsed over layers), replicating previous work [15] and suggesting that the effect of topdown feedback depends on whether it is met with bottom-up input. Therefore, the laminar profile of this effect is also of interest, as it reflects the effect of top-down feedback in the presence of bottom-up sensory signals, as opposed to the top-down signal in the absence of sensory input at the location of the illusory triangle. Notably, the laminar profile of this top-down effect (Figure 3A) did not significantly differ from the laminar profile evoked by sensory stimulation (Figure 3B; F2,18 = 1.2, p = 0.32; see also Figures S1C and S1D). Indeed, after applying the conservative laminar time course regression discussed above, the laminar profile of the inducer suppression effect differed significantly from the laminar profile of the excitatory illusory figure effect at the illusory triangle location (F2,18 = 3.8, p = 0.043; Figure S1E). In order to visualize the laminar-specific effects of illusory figure perception across the visual field, we reconstructed the BOLD signal into visual space at the three different cortical depths, for both the top-down illusory figure effect (i.e., Kanizsa trials versus control trials) and the bottom-up checkerboard stimulation (Figure 4). These reconstructions represent BOLD signals that have been transformed from cortical (i.e., voxel) space to visual (i.e., stimulus) space by multiplying each voxel’s population receptive field by that voxel’s BOLD response (for details, see Supplemental Experimental Procedures, ‘‘Retinotopic reconstruction of effects’’). It can be seen that the illusory figure enhancement is strongest in the deep layers and decreases as one moves up to the superficial layer (Figure 4A, cf.

Figure 2A), while the magnitude of bottom-up activity induced by the checkerboard is lowest in deep layers and increases in amplitude from deep to superficial layers (Figure 4B, cf. Figure 3A). The decreased response to the Pac-Man inducers shows a similar profile to bottom-up stimulation and was strongest in superficial layers (Figure 4A). DISCUSSION Layer-Specific Effects of Illusory Figures in V1 Using high-resolution fMRI at 7 T, we showed that a top-down activation of the visual cortex, induced by a visual illusion, selectively activates the deep cortical layers of V1. This stands in contrast to the neural response elicited by a bottom-up stimulus, which activates all layers and is strongest in the middle and superficial layers. The top-down suppression of the response to Pac-Man shapes that induce an illusory figure, compared to identical shapes that do not, shows a different laminar profile than the excitatory top-down illusory figure effect and is in fact more similar to the laminar profile of the bottom-up stimulus effect, being strongest in middle and superficial layers. These results suggest that the laminar profile of feedback effects depends on whether they interact with bottom-up inputs. In the current study, we divided the cortex in three equivolume layers, which do not necessarily correspond to the histological layers. However, in humans, these three laminar compartments are expected to correspond roughly to layers I–III, layer IV, and layers V–VI, respectively [18]. The selective activation of the deep layers of V1 due to the illusion appears

2 Current Biology 26, 1–6, February 8, 2016 ª2016 Elsevier Ltd All rights reserved

Please cite this article in press as: Kok et al., Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback, Current Biology (2016), http://dx.doi.org/10.1016/j.cub.2015.12.038

Figure 2. Laminar Profiles of the BOLD Response in the Illusory Figure Region

Figure 3. Laminar Profiles of the BOLD Response in the Inducer Region

(A) BOLD response to illusory figures (green) and control stimuli (red) in the region of V1 with receptive fields on the illusory figure, but not on the Pac-Man inducers. *p = 0.0070. (B) BOLD response to a contrast-reversing checkerboard in the same region of interest (ROI) as in (A). Error bars indicate within-subject SEM. See Figures S1 and S2 for additional analyses and Figure S3 for individual participants’ data.

(A) BOLD response to Pac-Man shapes when they induced an illusory figure (green) and when they did not (red) in the region of V1 with receptive fields on the inducers. Note that only physically identical Pac-Man shapes were compared (i.e., the bottom Pac-Man in the example). Results were averaged over the ROIs corresponding to the bottom and right Pac-Man locations only, since these ROIs contained the most voxels (compared to the left and top Pac-Man locations) due to cortical magnification. See Supplemental Experimental Procedures for details. *p = 0.017. (B) BOLD response to a contrast-reversing checkerboard in the same ROI as in (A). Error bars indicate within-subject SEM. See Figures S1 and S2 for additional analyses and Figure S3 for individual participants’ data.

in line with the known anatomical [2, 5] and functional [19] feedback connectivity from higher-order visual regions to the deep layers of V1. In addition to the deep layers, feedback connections are known to have dense terminations in layer I [2, 5]. Indeed, a recent laminar fMRI study of contextual effects in unstimulated human V1 reports effects in the most superficial layers [20]. It should be noted, however, that layer I is very thin, sparsely populated with neurons, and close to the pial veins on the cortical surface, making it challenging to detect layer-specific BOLD activation in this layer. In addition to layer I, contextual BOLD effects that are specific to the superficial layers could potentially also arise from layers II–III, known to be strong targets of horizontal connections [5, 21, 22]. Self et al. recently assessed layer-specific neural responses during figure-ground segregation using invasive laminar recordings in macaques [21]. In line with our findings, they report top-down figure filling-in in infragranular layer V. In addition, feedback effects are reported in superficial layer I and the upper part of layer II. An important difference with the current study is that Self and colleagues investigated the effect of top-down figure perception in the presence of bottom-up stimulation, while the illusory figure in the present study occurred in the absence of bottom-up input at the corresponding visual field locations. In fact, when we probed the laminar profile of the top-down suppression of activity evoked by the Pac-Man inducers, this effect was not restricted to the deep layers but appeared strongest in the superficial layers. This suggests that top-down feedback may evoke different laminar profiles depending on whether or not it interacts with bottom-up input. One potential mechanism of this interaction is through inhibitory connections from the deep layers to the granular layer IV [23–25], which in turn can cause a reduction throughout the entire cortical column as a result of the excitatory pathway from layer IV to layers II–III and

from layers II–III to layers V–VI [17]. In the absence of sensory input to layer IV, as is the case for the illusory triangle studied here, this modulation would not occur and top-down feedback signals would be restricted to the deep layers. An alternative explanation could be that the inducer suppression is the result of feedback connections terminating on inhibitory neurons in layer I, which in turn inhibit pyramidal neurons in layers II–III [26, 27]. This suggests that it is important to distinguish between driving and modulatory effects of backward connections when studying feedback signals in visual cortex. Hierarchical Perceptual Inference Our results are in line with theories that cast perception as hierarchical perceptual inference [28, 29], according to which higher-order sensory regions form hypotheses about the causes of current sensory inputs and send feedback to lower-order regions to ‘‘test’’ these hypotheses. In the context of Kanizsa figures, neurons in higher-order visual regions with large receptive fields (such as in the lateral occipital complex [30]) may ‘‘hypothesize’’ the presence of the (illusory) triangle based on the aligned wedges of the Pac-Man inducers and send feedback to neurons in lower-order visual regions (e.g., V1) that are expected to be active should such a triangle indeed be present. One proposed implementation of this process is (hierarchical) predictive coding [28, 31, 32], according to which each cortical region houses separate sub-populations of neurons coding for perceptual hypotheses (predictions) and mismatches between these hypotheses and bottom-up sensory

Current Biology 26, 1–6, February 8, 2016 ª2016 Elsevier Ltd All rights reserved 3

Please cite this article in press as: Kok et al., Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback, Current Biology (2016), http://dx.doi.org/10.1016/j.cub.2015.12.038

Figure 4. Reconstruction of the BOLD Signal at the Three Different Cortical Depths (A) Reconstruction of the difference in BOLD response evoked by illusory triangle configurations versus control configurations, separately for the three different cortical layers. The transformation of BOLD responses from cortical (i.e., voxel) space to visual (i.e., stimulus) space was achieved by multiplying each voxel’s population receptive field (a 2D Gaussian with center (x0, y0) and size s) by that voxel’s differential BOLD response (illusory figure versus control). In order to obtain reconstructions per cortical layer, each voxel’s receptive field was additionally weighted by the proportion of that voxel’s volume that was in the layer of interest. We collapsed over the four different illusory triangles by first rotating the reconstructions of the upward-, rightward-, and downward-pointing triangles 90 , 180 , and 270 counterclockwise, respectively, to move them into the reference frame of the leftward-pointing triangle, before averaging. See Supplemental Experimental Procedures (‘‘Retinotopic reconstruction of effects’’) for details. Dashed white triangles indicate the location of the illusory triangles. (B) Reconstruction of the BOLD response evoked by a checkerboard stimulus presented at the location at which illusory figures were sometimes induced (see Figure 2B). Note that these reconstructions only served visualization purposes; all statistical analyses were performed on the laminar profiles in Figures 2, 3, and S1.

input (prediction errors). These sub-populations are suggested to reside in different cortical layers, with prediction units being predominantly present in the deep layers and prediction error units in the middle and superficial layers [26]. In the context of the current study, the illusory figure can be seen as a perceptual hypothesis (prediction) and would thus be expected to be encoded by prediction units in the deep layers of V1. Conversely, at the location of the Pac-Man shapes that induce the illusion, the presence of an illusory figure elicits a perceptual hypothesis (‘‘a partially occluded black circle’’) that is met with consistent bottom-up input (a partial black circle). Such a match between top-down predictions and bottom-up signals would lead to a reduced prediction error response in middle and superficial layers, compared to when there are no such top-down predictions (as in the control configurations; Figure 1B). It should be noted that predictive coding is a process theory of perceptual processing that entails several phenomena—such as figure-ground segregation [33, 34] and biased competition [35, 36]—whose exact mechanisms have yet to be established. Therefore, more studies more specifically geared toward testing the effects of perceptual predictions [37, 38] with laminar resolution are needed. For instance, it would be of great interest to compare the laminar profiles of perceptual predictions and selective attention, as these are expected to affect separate sub-populations of neurons under these predictive coding accounts [36]. Additionally, future studies may pursue even greater specificity of neural responses by considering not only their amplitude but also their information content, using multivariate pattern analyses [39, 40]. For instance, such analyses could reveal whether the illusory figure effects presented here are specific to voxels whose orientation preference matches the orientation of the illusory contours. Indeed, a recent study combining multivariate pattern analysis with laminar fMRI demonstrates the potential power of this method [20].

Limitations of Laminar fMRI With laminar-specific fMRI, spatial resolution is a particular challenge. The average thickness of human V1 is 2.5 mm [41], which we subdivide into three equi-volume layers. It should be noted that the effective resolution is likely to be somewhat decreased by blurring due to factors like head motion, inaccuracies in the boundary registration, and cortical inhomogeneity. Importantly, we do not rely on each (0.8 mm) voxel’s volume being uniquely captured by one layer in order to obtain independent laminar signals. Instead, we performed a spatial regression, which explicitly attempts to unmix the signals from the different layers, even though these signals are mixed in individual voxels. This spatial unmixing is designed to extract signals with considerably less dependence on the actual voxel volume than the interpolation approach used in some previous laminar fMRI papers [8]. This is also expressed in our own data; the average correlation between the three gray matter layers is lower for time courses established using the regression approach (r  0.5) than for time courses at the same cortical depths established using the interpolation approach (r  0.8). Future studies may aim to increase the spatial resolution of laminar fMRI studies further, improving the ability to estimate independent laminar BOLD signals. Another matter of concern are interdependencies in the BOLD signal between the different layers, as a result of venous blood draining from the deeper layers toward the surface [8, 9, 12, 42, 43]. Given the direction of this flow, the more superficial layers are expected to contain a mixture of signals from different layers, while the signal in the deep layers is relatively free of such effects. These interdependencies, and the resulting difference in reliability for signals in the different layers, make it challenging to precisely ascertain the laminar origin of signals, particularly in the superficial layers. This argues for the importance of comparing laminar profiles of different effects (e.g., the checkerboard and illusory figure effects in the current study) rather than only considering the laminar profile of a particular effect per se.

4 Current Biology 26, 1–6, February 8, 2016 ª2016 Elsevier Ltd All rights reserved

Please cite this article in press as: Kok et al., Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback, Current Biology (2016), http://dx.doi.org/10.1016/j.cub.2015.12.038

We assessed the interdependencies between layers by performing an analysis in which the signals from the neighboring layers were regressed out to show each layer’s unique contribution to the signal (Figure S1). In the case of bottom-up stimulation, this additional regression particularly reduced the signal in the superficial layers, while the strongest signals remained in the middle layers (compare Figures 2B and 3B to S1B and S1D), suggesting drainage from middle layers to the surface. The illusory figure effect in the deep layers, however, was unaffected (compare Figures 2A and S1A). These results suggest that we can be confident about the origin of deep layer signals, compared to superficial signals. Finally, a challenge in interpreting laminar fMRI (and fMRI in general) results is the complex relationship between neuronal excitation and inhibition and the hemodynamic response. However, studies directly linking neuronal activity and hemodynamic responses in monkeys have reported a close correspondence between neural (synaptic) activity and the BOLD response, both for excitation [44] and inhibition [45]. Conclusions In sum, the current study reveals highly specific laminar effects related to a feedback signal in the primary visual cortex using fMRI, showing the potential of studying the activity dynamics of the cortical layers non-invasively in humans. This technique holds great promise for elucidating the interaction between bottom-up and top-down signals during perception, as well as in other areas of cognition such as memory [46]. EXPERIMENTAL PROCEDURES The experimental procedures are summarized briefly throughout the Results and are presented in detail in the Supplemental Experimental Procedures.

2. Harris, K.D., and Mrsic-Flogel, T.D. (2013). Cortical connectivity and sensory coding. Nature 503, 51–58. 3. Muckli, L., and Petro, L.S. (2013). Network interactions: non-geniculate input to V1. Curr. Opin. Neurobiol. 23, 195–201. 4. Lee, T.S., and Nguyen, M. (2001). Dynamics of subjective contour formation in the early visual cortex. Proc. Natl. Acad. Sci. USA 98, 1907–1911. 5. Rockland, K.S., and Pandya, D.N. (1979). Laminar origins and terminations of cortical connections of the occipital lobe in the rhesus monkey. Brain Res. 179, 3–20. 6. Wong-Riley, M. (1978). Reciprocal connections between striate and prestriate cortex in squirrel monkey as demonstrated by combined peroxidase histochemistry and autoradiography. Brain Res. 147, 159–164. 7. Koopmans, P.J., Barth, M., Orzada, S., and Norris, D.G. (2011). Multi-echo fMRI of the cortical laminae in humans at 7 T. Neuroimage 56, 1276–1285. 8. Koopmans, P.J., Barth, M., and Norris, D.G. (2010). Layer-specific BOLD activation in human V1. Hum. Brain Mapp. 31, 1297–1304. 9. Polimeni, J.R., Fischl, B., Greve, D.N., and Wald, L.L. (2010). Laminar analysis of 7T BOLD using an imposed spatial activation pattern in human V1. Neuroimage 52, 1334–1346. 10. Goense, J.B.M., and Logothetis, N.K. (2006). Laminar specificity in monkey V1 using high-resolution SE-fMRI. Magn. Reson. Imaging 24, 381–392. 11. Ress, D., Glover, G.H., Liu, J., and Wandell, B. (2007). Laminar profiles of functional activity in the human brain. Neuroimage 34, 74–84. 12. Chen, G., Wang, F., Gore, J.C., and Roe, A.W. (2013). Layer-specific BOLD activation in awake monkey V1 revealed by ultra-high spatial resolution functional magnetic resonance imaging. Neuroimage 64, 147–155. 13. Harel, N., Lin, J., Moeller, S., Ugurbil, K., and Yacoub, E. (2006). Combined imaging-histological study of cortical laminar specificity of fMRI signals. Neuroimage 29, 879–887. 14. Olman, C.A., Harel, N., Feinberg, D.A., He, S., Zhang, P., Ugurbil, K., and Yacoub, E. (2012). Layer-specific fMRI reflects different neuronal computations at different depths in human V1. PLoS ONE 7, e32536.

SUPPLEMENTAL INFORMATION

15. Kok, P., and de Lange, F.P. (2014). Shape perception simultaneously upand downregulates neural activity in the primary visual cortex. Curr. Biol. 24, 1531–1535.

Supplemental Information includes Supplemental Experimental Procedures, three figures, and one table and can be found with this article online at http://dx.doi.org/10.1016/j.cub.2015.12.038.

16. Van Mourik, T., Van Der Eerden, J.P.J.M., and Norris, D.G. (2015). Laminar time course extraction over extended cortical areas. Proceedings of the 23rd Annual Meeting of ISMRM 23, 2071.

AUTHOR CONTRIBUTIONS

17. Douglas, R.J., and Martin, K.A.C. (2004). Neuronal circuits of the neocortex. Annu. Rev. Neurosci. 27, 419–451.

P.K. and F.P.d.L. designed the experiment. P.K. and L.J.B. collected the data. T.v.M. and D.G.N. developed the laminar analysis technique. P.K. and T.v.M. analyzed the data. P.K., L.J.B., T.v.M., D.G.N., and F.P.d.L. wrote the paper.

18. de Sousa, A.A., Sherwood, C.C., Schleicher, A., Amunts, K., MacLeod, C.E., Hof, P.R., and Zilles, K. (2010). Comparative cytoarchitectural analyses of striate and extrastriate areas in hominoids. Cereb. Cortex 20, 966–981.

ACKNOWLEDGMENTS

19. Sandell, J.H., and Schiller, P.H. (1982). Effect of cooling area 18 on striate cortex cells in the squirrel monkey. J. Neurophysiol. 48, 38–48.

We thank Irati Markuerkiaga for helpful discussions. P.K. and F.P.d.L. were supported by grants from NWO (VIDI grant 452-13-016) and the James S. McDonnell Foundation (Understanding Human Cognition, 220020373). The institutional review board at Radboud Univeristy and the ethics committee of the University Duisburg-Essen approved the study.

20. Muckli, L., De Martino, F., Vizioli, L., Petro, L.S., Smith, F.W., Ugurbil, K., Goebel, R., and Yacoub, E. (2015). Contextual feedback to superficial layers of V1. Curr. Biol. 25, 2690–2695.

Received: August 28, 2015 Revised: November 7, 2015 Accepted: December 8, 2015 Published: January 28, 2016

22. Angelucci, A., and Bullier, J. (2003). Reaching beyond the classical receptive field of V1 neurons: horizontal or feedback axons? J. Physiol. Paris 97, 141–154.

21. Self, M.W., van Kerkoerle, T., Supe`r, H., and Roelfsema, P.R. (2013). Distinct roles of the cortical layers of area V1 in figure-ground segregation. Curr. Biol. 23, 2121–2129.

23. Thomson, A.M., and Bannister, A.P. (2003). Interlaminar connections in the neocortex. Cereb. Cortex 13, 5–14.

REFERENCES 1. Felleman, D.J., and Van Essen, D.C. (1991). Distributed hierarchical processing in the primate cerebral cortex. Cereb. Cortex 1, 1–47.

24. Ka¨tzel, D., Zemelman, B.V., Buetfering, C., Wo¨lfel, M., and Miesenbo¨ck, G. (2011). The columnar and laminar organization of inhibitory connections to neocortical excitatory cells. Nat. Neurosci. 14, 100–107.

Current Biology 26, 1–6, February 8, 2016 ª2016 Elsevier Ltd All rights reserved 5

Please cite this article in press as: Kok et al., Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback, Current Biology (2016), http://dx.doi.org/10.1016/j.cub.2015.12.038

25. Kim, J., Matney, C.J., Blankenship, A., Hestrin, S., and Brown, S.P. (2014). Layer 6 corticothalamic neurons activate a cortical output layer, layer 5a. J. Neurosci. 34, 9656–9664.

37. den Ouden, H.E.M., Friston, K.J., Daw, N.D., McIntosh, A.R., and Stephan, K.E. (2009). A dual role for prediction error in associative learning. Cereb. Cortex 19, 1175–1185.

26. Bastos, A.M., Usrey, W.M., Adams, R.A., Mangun, G.R., Fries, P., and Friston, K.J. (2012). Canonical microcircuits for predictive coding. Neuron 76, 695–711.

38. Kok, P., Jehee, J.F.M., and de Lange, F.P. (2012). Less is more: expectation sharpens representations in the primary visual cortex. Neuron 75, 265–270.

27. Chu, Z., Galarreta, M., and Hestrin, S. (2003). Synaptic interactions of latespiking neocortical neurons in layer 1. J. Neurosci. 23, 96–102.

39. Williams, M.A., Baker, C.I., Op de Beeck, H.P., Shim, W.M., Dang, S., Triantafyllou, C., and Kanwisher, N. (2008). Feedback of visual object information to foveal retinotopic cortex. Nat. Neurosci. 11, 1439–1445.

28. Mumford, D. (1992). On the computational architecture of the neocortex. II. The role of cortico-cortical loops. Biol. Cybern. 66, 241–251. 29. Lee, T.S., and Mumford, D. (2003). Hierarchical Bayesian inference in the visual cortex. J. Opt. Soc. Am. A Opt. Image Sci. Vis. 20, 1434–1448. 30. Kourtzi, Z., and Kanwisher, N. (2001). Representation of perceived object shape by the human lateral occipital complex. Science 293, 1506–1509. 31. Rao, R.P.N., and Ballard, D.H. (1999). Predictive coding in the visual cortex: a functional interpretation of some extra-classical receptive-field effects. Nat. Neurosci. 2, 79–87. 32. Brown, H., and Friston, K.J. (2012). Free-energy and illusions: the cornsweet effect. Front. Psychol. 3, 43. 33. Roelfsema, P.R., Lamme, V.A.F., Spekreijse, H., and Bosch, H. (2002). Figure-ground segregation in a recurrent network architecture. J. Cogn. Neurosci. 14, 525–537. 34. Kanai, R., Komura, Y., Shipp, S., and Friston, K. (2015). Cerebral hierarchies: predictive processing, precision and the pulvinar. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 20140169. 35. Desimone, R., and Duncan, J. (1995). Neural mechanisms of selective visual attention. Annu. Rev. Neurosci. 18, 193–222. 36. Feldman, H., and Friston, K.J. (2010). Attention, uncertainty, and free-energy. Front. Hum. Neurosci. 4, 215.

40. Kamitani, Y., and Tong, F. (2005). Decoding the visual and subjective contents of the human brain. Nat. Neurosci. 8, 679–685. 41. Fischl, B., and Dale, A.M. (2000). Measuring the thickness of the human cerebral cortex from magnetic resonance images. Proc. Natl. Acad. Sci. USA 97, 11050–11055. 42. Duvernoy, H.M., Delon, S., and Vannson, J.L. (1981). Cortical blood vessels of the human brain. Brain Res. Bull. 7, 519–579. 43. Heinzle, J., Koopmans, P.J., Den Ouden, H.E.M., Raman, S., and Stephan, K.E. (2016). A hemodynamic model for layered BOLD signals. Neuroimage 125, 556–570. 44. Goense, J.B.M., and Logothetis, N.K. (2008). Neurophysiology of the BOLD fMRI signal in awake monkeys. Curr. Biol. 18, 631–640. 45. Shmuel, A., Augath, M., Oeltermann, A., and Logothetis, N.K. (2006). Negative functional MRI response correlates with decreases in neuronal activity in monkey visual area V1. Nat. Neurosci. 9, 569–577. 46. Maass, A., Schu¨tze, H., Speck, O., Yonelinas, A., Tempelmann, C., Heinze, H.-J., Berron, D., Cardenas-Blanco, A., Brodersen, K.H., Stephan, K.E., and Du¨zel, E. (2014). Laminar activity in the hippocampus and entorhinal cortex related to novelty and episodic encoding. Nat. Commun. 5, 5547.

6 Current Biology 26, 1–6, February 8, 2016 ª2016 Elsevier Ltd All rights reserved

Current Biology Supplemental Information

Selective Activation of the Deep Layers of the Human Primary Visual Cortex by Top-Down Feedback Peter Kok, Lauren J. Bains, Tim van Mourik, David G. Norris, and Floris P. de Lange

Figure S1, related to Figures 2-3. Differential layer-specific BOLD responses, over and above the overall BOLD response. These response profiles were obtained by regressing out, for each layer, the time courses of the other two layers. (A) Differential BOLD response to illusory figures (green) and control stimuli (red) in the region of V1 with receptive fields on the illusory figure. *p = 0.0053. (B) Differential BOLD response to a contrast-reversing checkerboard in the same ROI as in panel A. (C) Differential BOLD response to the inducing Pac-Man shapes when they either induced an illusory figure (green) or when they did not (red), in the region of V1 with receptive fields on the inducers. *p = 0.037. (D) Differential BOLD response to a contrast-reversing checkerboard in the same ROI as in panel C. (E) Differential Laminar profiles of the illusory figure effect in the 'figure region' (blue), and the inducer suppression effect in the 'inducer region'. *p = 0.019, †p = 0.062. Error bars indicate within-subject SEM.

Figure S2, related to Figures 2-3. Layer-specific BOLD responses, obtained by interpolation analysis. These response profiles were obtained sampling the cortical depth at 50 points and estimating the BOLD signal at each depth through trilinear interpolation (see Supplemental Experimental Procedures). (A) Interpolated BOLD response to illusory figures (green) and control stimuli (red) in the region of V1 with receptive fields on the illusory figure. In order to perform similar statistical analyses as in the main analysis, we sampled this signal at three cortical depths corresponding to the three layers as defined in the spatial GLM. As in the main analysis, the laminar profile induced by the illusory figure differed significantly from that induced by a contrast-reversing checkerboard (F2,18 = 8.3, p = 0.0028), and the illusory figure effect was significantly present in the deep (t9 = 3.2, p = 0.010) but not the middle (t9 = 1.4, p = 0.20) and superficial (t9 = 1.1, p = 0.31) layers. (B) Interpolated BOLD response to a contrast-reversing checkerboard in the same ROI as in panel A. (C) Interpolated BOLD response to the inducing Pac-Man shapes when they either induced an illusory figure (green) or when they did not (red), in the region of V1 with receptive fields on the inducers. As in the main analysis, the suppression of the Pac-Men in the presence of an illusory figure (t9 = 2.4, p = 0.038, collapsed across layers) was most clearly visible in the superficial layers, as was the checkerboard response. There was a significant interaction between effect type (Illusory figure vs. Checkerboard and Layer; F2,18 = 5.6, p = 0.013), due to the slope being steeper for the checkerboard effect.(D) Interpolated BOLD response to a contrast-reversing checkerboard in the same ROI as in panel C. Error bars indicate within-subject SEM.

Figure S3, related to Figures 2-3. Layer-specific BOLD responses, individual participants’ data. Same data as in Figures 2 and 3, but displaying individual participant data (blue lines and dots), with group means superimposed (green x’s). (A) BOLD response to illusory figure in the region of V1 with receptive fields on the illusory triangle. (B) BOLD response to a contrast-reversing checkerboard in the same ROI as in panel A. (C) BOLD response to the inducing Pac-Man shapes when they either induced an illusory figure versus when they did not, in the region of V1 with receptive fields on the inducers. (D) BOLD response to a contrast-reversing checkerboard in the same ROI as in panel C.

Table S1, related to Figures 2-3. Correlations between cortical layers. Correlations between fMRI timecourses in the different cortical layers as estimated using the spatial GLM, averaged over scanner runs and over the three V1 ROIs used for the main analysis. Presented as group means (± standard deviation). Layer

WM

Deep

Middle

Superficial

CSF

WM Deep Middle Superficial CSF

– 0.34 ± 0.24 0.34 ± 0.20 0.42 ± 0.18 0.37 ± 0.19

– 0.46 ± 0.13 0.59 ± 0.20 0.40 ± 0.24

– 0.48 ± 0.21 0.36 ± 0.28

– 0.58 ± 0.10



Supplemental Experimental Procedures Participants Eleven healthy right-handed individuals (6 female, age 26 ± 4, mean ± SD) with normal or corrected-to-normal vision gave written informed consent to participate in this study, in accordance with the institutional guidelines of the local ethics committee (ethics committee of the University Duisburg-Essen). Data from one participant were not analysed due to a technical issue with the visual stimulation during data acquisition. Stimuli Stimuli were generated using MATLAB (MathWorks, Natick, MA, USA) and the Psychophysics Toolbox [S1], and were projected on a rear-projection screen using an EIKI (EIKI, Rancho Santa Margarita, CA) LC-X71 projector (1024 × 768 resolution, 60 Hz). During the illusory figure runs, stimuli consisted of four black circles with a wedge cut out (‘Pac-Men’; 3.6° diameter, 1.8 cd/m2) on a light gray background (1416 cd/m2) (Figure 1A). The stimuli were spaced such that the central 3.6×3.6° area did not contain any stimulus. A fixation bull’s-eye (0.7° diameter) was centred at 1.5° below and 1.5° to the right of the middle of the stimulus configuration. In all configurations, two inducers consisted of a black circle with a 45° wedge cut out, and two inducers had a 90° wedge cut out. In ‘Kanizsa’ configurations (Figure 1A), three inward facing inducers were aligned such that they induced the percept of an illusory triangle overlaying three black circles, while the fourth faced outward. ‘Control’ configurations consisted of inducers arranged such that no illusory figure was perceived (Figure 1B). In these configurations, either three out of four inducers faces inward but none of the inducers were aligned, or two out of four inducers faced inward and were partially aligned. In the illusory figure region (see ‘Specification of regions of interest’ for definition) there was no significant difference in BOLD response between the two different types of control trials, either collapsed over layers (t9 = -0.14, p = 0.89) or in any of the three individual layers (all p > 0.5), and we therefore collapsed over these trial types in the analysis. In total, we created 20 inducer configurations: 12 Kanizsa configurations and 8 controls. Each Kanizsa configuration contained one of four possible illusory triangles: leftward, rightward, upward or downward pointing. Trial length was either 17.04 s (5 TRs, first three participants) or 13.63 s (4 TRs, participants four to ten), during which the stimulus display alternated between four black whole circles (500 ms) and one of the 20 inducer configurations (500 ms). On a given trial, one inducer configuration was presented repeatedly for the duration of the trial. Additionally, a letter stream was presented at fixation throughout each run. Green letters (0.5°) were presented within the fixation bull’s eye, for 250 ms each, separated by 250 ms intervals during which only the fixation point was presented. Participants’ task was to detect the letters ‘X’ and ‘Z’ (target probability = 20%) in a stream of nontarget letters at fixation, thereby drawing attention away from the illusory shapes. In half of the experimental blocks, the letters were distorted by superimposing noise of the same colour (30% of pixels), increasing the difficulty of the letter detection task. Due to technical limitations, we could not record all button presses. During the blocks without noise, 97.5% ± 0.5% (mean ± se) of the recorded button presses occurred after a target was presented (i.e. hits; RT: 504 ms ± 10 ms), while 2.5% ± 0.5% of button presses occurred after no target was presented in the preceding 1000 ms (i.e. false alarms), suggesting that participants were successfully engaged by the task. During the blocks with added perceptual noise, 96.3% ± 0.9% of the recorded button presses were hits (RT: 518 ms ± 7 ms), while 3.8% ± 0.9% of button presses were false alarms. Since we found no significant differences between the two task difficulty levels in terms of BOLD signal in V1, we collapsed over these two conditions in all further analyses. As a baseline, we also included ‘fixation’ trials (of equal length as the other trial types), during which no black circles or inducers were presented, but during which the fixation point and letter stream persisted. In a separate scanner run during the same session, we presented participants with contrast-reversing checkerboard stimuli (10 Hz). The checkerboards contained checks of 0.4° width and height, each of which was either black or white randomly from one presentation (100 ms) to the next. Checkerboard contrast was either 50% or 80%, and since we found no significant differences between these conditions we collapsed over them for all further analyses. In separate trials, participants were presented with either four circular checkerboards (diameter = 3.6°, contrast fading linearly over the outer 0.5°) at the same locations at which the Pac-Men inducers were presented during the illusory figure runs (‘Outer checkerboard’, see Figure 3B), or with a single checkerboard (diameter = 3.6°, contrast fading linearly over the outer 0.5°) in the centre of the configuration (‘Inner checkerboard’, see Figure 2B), corresponding to the location where the illusory figure was perceived during the Kanizsa trials. Trial length was the same as in the illusory figure runs, as was the presence of the fixation point, including the letter stream and participants’ letter detection task. We also included fixation trials, during which no checkerboards were presented, but the fixation point and letter stream persisted.

Procedure In each block of the illusory figure runs, all 20 stimulus configurations were presented once, in random order, with fixation trials inserted after each fourth trial, yielding a block length of 426 s (first three participants) or 341 s (participants four to ten). The fixation point and letter stream were presented throughout the block. Participants performed three runs of two blocks each (30 s break between blocks). For one participant, data from one run had to be discarded due to a technical problem with visual stimulus presentation. In the checkerboard run, ‘Inner checkerboard’ and ‘Outer checkerboard’ trials were presented in interleaved order, with a fixation trial inserted after every fourth trial. Each block consisted of 10 trials of the two checkerboard types each, and 5 fixation trials, yielding a block length of 426 s (first three participants) or 341 s (participants four to ten). The fixation point and letter stream were presented throughout the block. Participants performed one runs of two blocks (30 s break between blocks). Before entering the scanner, participants received written instructions and practiced the letter detection task on a laptop. Population receptive field measurements After the main experiment, participants were presented with moving bar stimuli, in order to map the population receptive fields (pRFs) [S2] of voxels in early visual cortex, as well as allow polar angle mapping in order to delineate the borders between retinotopic areas in visual cortex [S3, 4]. During these runs, bars containing full contrast contrast-reversing checkerboards (2 Hz) moved across the screen in a circular aperture with a diameter of 12°. The bars moved in eight different directions (four cardinal and four diagonal directions) in 12 steps of 1°, one step per TR (3.408 s). Four blank fixation screens (13.6 s) were inserted after each of the cardinally moving bars. Throughout each run (~6.5 min), a coloured fixation dot was presented in the centre of the screen, changing colour (red to green and green to red) at random time points. Participants’ task was to press a button whenever this colour change occurred. All participants performed four runs of this task. fMRI acquisition parameters Functional images were acquired on a Magnetom 7T whole body MRI system (Siemens Healthcare GmbH, Erlangen, Germany) with a 32-channel head coil (Nova Medical, Wilmington, USA) combined with dielectric pads (Wyger M. Brink, Radiology, Leiden University Medical Center, Leiden, Netherlands [S5]), using a T2*-weighted 3D gradient-echo EPI sequence [S6] (TR = 3408 ms, TE = 28 ms, 48 transversal slices, voxel size 0.8×0.8×0.8 mm, 16° flip angle, field of view 192 × 192 mm, GRAPPA acceleration factor 4). Gradient maximum amplitude was 40 mT/m, the minimum gradient rise time was 200 μs, and the maximum slew rate was 200 T/m/s. Shimming was performed using the standard Siemens shimming procedure for 7T. Anatomical images were acquired using an MP2RAGE sequence [S7] (TR = 5000 ms, TE = 2.04 ms, voxel size 0.8×0.8×0.8 mm, field of view 240 × 240 mm, GRAPPA acceleration factor 2), yielding two inversion contrasts (TI1 = 900 ms, 4° flip angle , TI2 = 3200 ms, 6° flip angle). These two inversion contrasts were combined to yield a single T1-weighted image. Preprocessing of functional volumes SPM8 (http://www.fil.ion.ucl.ac.uk/spm, Wellcome Trust Centre for Neuroimaging, London, UK) was used for image preprocessing. The first four volumes of the each run were discarded to allow T1 equilibration. The functional volumes were cropped to remove the frontal part of the brain (in which more severe distortions occurred, due to our shimming volume being placed in the occipital lobe). The functional volumes were spatially realigned within scanner runs, and subsequently between runs, to correct for head movement. Additionally, in order to correct for changes in distortions in the phase-encoding direction in the EPI volumes occurring over the course of the experiment due to changes in head position, we calculated a non-linear (in the phase-encoding direction) registration [S8] from the mean functional volumes of the checkerboard and moving bar runs, respectively, to the mean functional volume of the illusory figure runs. These two non-linear registrations were applied to all functional volumes of the checkerboard and moving bars runs, respectively. This non-linear registration was restricted to correcting distortions in the phase encoding direction, and was described by the map:

T : (x, y, z) | (x, y  d(x, y, z), z) where d(x, y, z) is the displacement field and phase encoding is along the y-direction. The displacement field is parameterised as:

d ( x, y, z )   aijk X i ( x)Y j ( y ) Z k ( z ) i , j ,k

where Xi, Yj and Zk are the basis functions of the discrete cosine transform up to a certain order (here i = 6, j = 9, k = 9). This reduces the number of parameters to be estimated and ensures smoothness of the displacement field. The coefficients aijk were found by minimisation of the normalised mutual information of the to-be transformed EPI image and the target image, which has been coregistered with a rigid body transformation. Extraction and coregistration of cortical surfaces Freesurfer (http://surfer.nmr.mgh.harvard.edu/) was used to detect the boundaries between gray matter and white matter and cerebrospinal fluid (CSF), respectively, and generate inflated cortical surfaces, based on the T1-weighted anatomical image. The gray matter boundaries were subsequently registered to the mean functional volume in two steps: 1) a conventional rigid body boundary-based registration (BBR) [S9], and 2) recursive boundary-based registration [S10], BBR was applied recursively to increasingly smaller partitions of the cortical mesh. Here, we applied affine BBR with 7 degrees of freedom: rotation and translation along all three dimensions, and scaling along the phase-encoding direction. In each iteration, the cortical mesh is split into two, the optimal BBR transformations are found and applied to the respective parts. Subsequently, each part is split into two again and registered. The specificity increases at each stage, and corrects for local mismatches between the structural and the functional volumes that are due to magnetic field inhomogeneity related distortions. Here, we ran four such iterations. The splits are made along the cardinal axes of the volume, such that the number of vertices is equal for both parts. The plane for the second cut is orthogonal to the first, the third orthogonal to the first two. Having iterated through all three spatial dimensions, the final split made along the same axis as the first split. The median displacement was taken after running the recursive algorithm six times, in which different splitting orders where used, comprised of all six permutations of x, y and z. pRF estimation and retinotopic mapping The data from the moving bar runs were used to estimate the population receptive field (pRF) of each voxel in the functional volumes we obtained, using MrVista (http://white.stanford.edu/software/). Before estimation, the BOLD timecourses from the four runs were averaged together, yielding one timecourse per voxel. During estimation, a predicted BOLD signal is calculated from the known stimulus parameters and a model of the underlying neuronal population. The model of the neuronal population consisted of a two-dimensional Gaussian pRF, with parameters x0, y0, and σ, where x0 and y0 are the coordinates of the centre of the receptive field, and σ indicates its spread (standard deviation), or size. All parameters were stimulus-referred, and their units were degrees of visual angle. These parameters were adjusted to obtain the best possible fit of the predicted to the actual BOLD signal. The goodness of this fit was expressed as proportion of each voxel’s variance (in the averaged timecourse) explained by its pRF model. For all subsequent analyses, voxels were considered ‘visually active’ when at least 10% of their variance was explained by the pRF model. For all other voxels, pRF parameters were not considered reliable and thus not taken into consideration. For details of this procedure, see Dumoulin and Wandell (2008). This method has been shown to reconstruct the cortical visual field map more accurately than conventional retinotopic mapping methods, as well as produce pRF size estimates that agree well with electrophysiological receptive field measurements in monkey and human visual cortex. Once estimated, x0 and y0 can easily be converted to eccentricity and polar-angle measures, to allow retinotopic mapping. Polar-angle maps were overlaid on inflated cortical surfaces using Freesurfer to identify the boundaries of retinotopic areas in early visual cortex [S3, 4]. In this way, we delineated area V1 in both hemispheres. Specification of regions of interest We specified five regions of interest (ROIs), based on the pRF locations of the voxels. First, a ‘figure region’, consisting of visually active voxels (as indicated by a pRF model fit of > 10% variance explained, see ‘pRF estimation and retinotopic mapping’) whose receptive field centre (x0, y0) was on the area of the visual field containing the illusory figure (i.e. a circle with a diameter of 3.6° in the centre of the Pac-Man configurations), and at least 2σ away from any of the Pac-Man inducers (see Figure 2A). Second, four ‘inducer regions’, consisting of voxels with a receptive field centre on one of the four Pac-Man inducers (excluding voxels whose pRF was within 7.2° of the centre of the Pac-Man configuration, since these pRFs fell on the illusory figure itself on a subset of

trials) (see Figure 3A). These ROIs were subsequently projected into cortical surface space (using Freesurfer) and intersected with the V1 labels obtained through retinotopic mapping (see above). This resulted in 812 ± 246 (mean ± sd, range = 491 – 1208) ‘figure region’ voxels in V1. Due to cortical magnification, the ROIs corresponding to the inducers closest to fixation (i.e. the bottom and right inducers) contained more voxels (bottom: N = 515 ± 207, range = 237 – 900, right: N = 504 ± 231, range = 325 – 1094) than the ROIs for the inducers further away from fixation (i.e left and top inducer; left: N = 222 ± 66, range = 137 – 331, top: N = 162 ± 117, range = 53 – 436). Therefore, we decided to only consider the former two ROIs for further analysis, to ensure reliable extraction of laminar BOLD profiles. In order to sample the full cortical depth, we projected these ROIs from cortical surface space into volume space by labelling all voxels that intersected with the surface label over the full cortical depth (i.e. from the boundary between gray and white matter to the boundary between gray matter and CSF), plus a 50% extension into white matter and CSF, respectively. On the resulting volume-based ROIs, we imposed the following voxel exclusion criteria: for the ‘figure region’ ROI, we excluded any voxels that were both visually responsive (as indicated by a pRF model fit of > 10% variance explained, see ‘pRF estimation and retinotopic mapping’) and whose pRF was within 2σ of any of the Pac-Man inducers, as well as any voxels that responded to the ‘Outer checkerboard’ (at the liberal threshold of T > 1). These conservative criteria were applied to ensure that this ROI corresponded to the region of the visual field were an illusory figure was induced, but no bottom-up stimuli (i.e. Pac-Men) were presented. For the inducer ROIs, we simply excluded visually responsive voxels whose pRF was not on the corresponding inducer. This full sampling of the cortical depth resulted in 2229 ± 880 (range = 1107 – 3839) voxels in the ‘figure region’ ROI, 1181 ± 478 (range = 465 – 1954) voxels for the bottom Pac-Man inducer ROI and 1102 ± 258 (range = 870 – 1745) voxels for the right Pac-Man inducer ROI. Since the current study was geared towards measuring V1 with high resolution, unfortunately V2 was not well covered by the restricted field of view (particularly along the dorsal – ventral axis), preventing us from probing layer specific BOLD responses in V2. Definition of cortical layers Within a stretch of cortex, the ratios of the volumes of the different layers remains constant [S11, 12]. The effect of this is that the width of each layer depends on the curvature of the cortex at that location. Therefore, we divided the gray matter into three equi-volume layers, taking the curvature of the cortex into account [S11, 13, 14]. First, we computed the level set functions of the two cortical boundaries. A level set is a signed distance function (SDF), which is zero at the boundary, and has values that increase with distance from the boundary, being positive on one side and negative on the other. Once these SDFs are determined for the two cortical boundaries, we can determine intermediate surfaces by moving the zero level set, i.e. the boundary surface:

 d  (1   )   in     out ,   [ 0,1] where φd is the new boundary surface, φin is the white matter boundary and φout the pial boundary. Varying ρ, we can construct a level set of surfaces stratifying the cortex. A ‘layer’ is the volume between two neighbouring surfaces. If ρ is chosen to be constant, the resulting surface keeps a constant distance fraction from the segmented boundaries; an equidistant model. However, it is known [S12] that the actual cortical layers do not maintain a constant distance from the inner and pial surfaces, but instead vary with the curvature of the cortex, keeping a constant volume. Hence, an equi-volume model should construct laminae that are thick at high curvatures and thin at low curvatures. In other words, the equi-volume model first transforms a desired volume fraction α of the laminae into a distance fraction ρ, taking the local curvature of the pial and white matter surfaces into account. This procedure has been shown to be more accurate than an equidistant model, and is described in detail in [S11]. Here, we calculated two intermediate surface between the white matter and pial boundaries, yielding three gray matter layers. Based on this, we calculated four SDFs, containing for each functional voxel its distance to the boundaries between the five cortical compartments (white matter, CSF, and the three gray matter layers). This set of SDFs (or ‘level set’) allowed the calculation of the distribution of each voxel’s volume over the five compartments [S13]. This layervolume distribution provided the basis for the laminar GLM discussed below. Extraction of laminar timecourses

For each ROI, there was a matrix that described the distribution of the 5 layers (white matter, CSF, and three gray matter layers) over all n voxels within that ROI: its layer-volume distribution (Figure 1E). This matrix can be used to unmix the different layer contributions within the context of the General Linear Model (GLM) [S13], as follows:

Y = X ·B + U where Y is a [n × 1] vector of voxel values from an ROI, X is the [n × 5] layer-volume distribution matrix and B is a [5 × 1] vector of layer signals. The error term U is a [n × 1] vector of deviations from the least squares estimate. For each ROI and each functional volume, the layer signal B̂ was estimated by regressing Y against X, yielding five laminar timecourses per ROI and per functional run. In a control analysis, we extracted laminar timecourses using an interpolation method [S15] rather than a spatial GLM. Specifically, we sampled the gray matter at 50 equidistant points for each vertex pair, and estimated the signal at these points through trilinear interpolation. The results of this analysis qualitatively replicated those of the main analysis, and are presented in Figure S2. It should be noted that the correlation between the timecourses of the three gray matter layers as determined by the interpolation method (average r = 0.826) was substantially higher as compared to the spatial GLM method (average r = 0.511), suggesting increased unmixing of the layer signals by the GLM approach. This necessarily leads to the differences between the layers being less pronounced in the interpolation results. Estimating effects of interest per layer We tested for differences in laminar activity between conditions also within the context of the temporal GLM. For the illusory figure runs, regressors representing the different Pac-Man configurations and the fixation trials were constructed by convolving boxcars representing the onset and duration of the trials with SPM8’s canonical haemodynamic response function. Head motion parameters, as well as their derivatives and the square of the derivatives, were included as regressors of no interest [S16]. Finally, the data were high-pass filtered (cutoff = 256 s) to remove low-frequency signal drifts. A similar model was constructed for the checkerboard run, with separate regressors for ‘Inner checkerboard’, ‘Outer checkerboard’, and fixation trials. After estimating the GLM, the parameter estimates of the regressors of interest were converted to percent signal change by subtracting the parameter estimates of the fixation trials and dividing by the mean signal of each run, for both the illusory figure and the checkerboard runs. Parameter estimates for the illusory figure runs were averaged over runs. For the ‘figure region’ ROI, we separately averaged the parameter estimates of all trials in which an illusory figure was induced (‘Kanizsa trials’) and the parameter estimates of all ‘Control trials’ (Figure 2A). We also probed the effect of the ‘Inner checkerboard’ in this ROI (Figure 2B). For each ‘inducer region’ ROI, we separately averaged a subset of the Kanizsa trials and a subset of the Control trials. Specifically, we averaged the response to Pac-Men shapes that induced an illusory figure (e.g. the top PacMan in Figure 1A) and the response to the exact same Pac-Man shape when it did not induce an illusory figure (e.g. the top Pac-Man in Figure 1B). This was done to ensure that the bottom-up stimuli were exactly physically matched when studying the effect of the illusory figure (Figure 3A). We also probed the effect of the ‘Outer checkerboard’ in these ROIs (Figure 3B). For the ‘inducer region’ ROIs, results were averaged over the individual ROIs. Since we were interested in directly comparing the shapes of the laminar profiles of top-down and bottom-up effects in the primary visual cortex, the effects of ‘effect type’ (illusory figure vs. checkerboard) and ‘layer’ (deep, middle, and superficial) were assessed using two-way repeated measures ANOVAs. Since the effects of the illusory figure and the checkerboard had different magnitudes (cf. Figures 2A-B) and we were interested in comparing the shapes of their laminar profiles and not their magnitudes, all laminar profiles were z-scored (per participant) to ensure equal overall mean and variance, before being submitted to the ANOVA. This z-scoring entailed subtracting the mean percent signal change over all layers from each layer’s percent signal change, and dividing by the standard deviation of the percent signal change over layers. When we compared the laminar profiles of the illusory effects in the ‘figure region’ and the ‘inducer region’, no z-scoring was required since their magnitudes were similar. To test the significance of the individual effects per layer, paired sample t-tests were conducted. In general, the BOLD timecourses of the different gray matter layers showed substantial correlations (r~0.4-0.5, Table S1), likely in part due to activity travelling through the cortical column [S17], leading to a co-activation of the different layers, but also spuriously due to partial volume effects. Therefore, in an additional analysis, we assessed each layer’s unique contribution by regressing out the time courses of the other two gray matter layers prior to analysing the timecourses with a GLM, as discussed above. All other analysis steps were identical.

When testing within-subject (i.e. repeated measures) effects, such as in the present study, graphs showing regular (between-subject) standard errors of the mean (SEM) do not accurately reflect the reliability of the effects, since these SEMs include the between-subject variability [S18]. Therefore, within-subject SEMs were obtained by calculating, for each participant, the difference between that participant’s mean (over conditions) and the group mean (i.e. the subject effect), and subtracting this difference from the participant’s observations before calculating SEMs [S18]. Finally, to correct for positive covariance between conditions introduced by removing participant means, these SEMs are multiplied by M/(M-1), where M is the number of conditions [S19]. Error bars in all figures indicate within-subject standard error of the mean, per layer (SEM) [S18, 19]. Retinotopic reconstruction of effects In addition to ROI specification, the estimated pRF parameters allowed a straightforward and intuitive reconstruction of the BOLD effects from cortical space to visual space [S20] (Figure 1C). Each voxel’s receptive field can be represented by a two-dimensional Gaussian, with peak coordinates (x0, y0) and standard deviation σ. The reconstruction in visual space consisted of the sum of the 2D-Gaussians of all voxels in a given visual area, weighted by the voxels’ BOLD response: n

 a i 1

i

 bi   g x 0 i , y 0 i ,  i 

Where n is the number of voxels in a given area, ai and bi are responses to certain conditions, both in percent signal change from ‘fixation’, and g(x0i,y0i,σi) is the 2D-Gaussian defining the voxel’s receptive field. The rationale is the following: if a voxel in V1 is highly active, then this reflects activity in V1 neurons corresponding to the region of visual space modelled by the 2D-Gaussian. Therefore, in order to represent this activity in visual space, we multiplied the voxel’s 2D-Gaussian receptive field with its activity (i.e., BOLD signal) and projected the result on a 2D map of visual space. By doing this for all V1 voxels, we obtained a reconstruction of the BOLD signal in V1 in visual space. For example, to reconstruct the response to a leftward pointing Kanizsa configuration versus a control configuration, the contrast between the BOLD responses to these two conditions was calculated for each voxel, and these values were used as voxel weights in the reconstruction. For voxels with a receptive field centre within σ of one of the inducers, we contrasted only conditions that were matched exactly in terms of the type and orientation of the inducer presented in the receptive field of the voxel, to fully equate bottom-up stimulation. We also calculated a ‘baseline’ reconstruction of the visual field, with each voxel’s weight set to 1: n

 1  g x 0 , y 0 ,    1 i

i 1

i

i

We divided all reconstructions by this baseline reconstruction, pixel wise, in order to compensate for the relatively higher number of voxels near the fovea than in the periphery (i.e., cortical magnification; see e.g. Figure S1B in [S20]). Finally, all reconstructions were spatially smoothed with a 2-D Gaussian smoothing kernel with σ = 1.0°. In order to visualise the effects per cortical layer, we also calculated layer specific reconstructions. This was done simply by calculating a reconstruction per layer, in which each voxel’s receptive field was weighted by the proportion of its volume that was in the current layer: n

 a i 1

i

 bi   vi , j  g  x 0 i , y 0 i ,  i 

Where vi,j is the proportion of voxel i’s volume in layer j. As for the reconstructions above, the layer-specific reconstructions were divided by layer-specific baseline reconstructions. The reconstructions were plotted as heat maps, with red and blue shades indicating relative activations and deactivations, respectively. In order to collapse over the four different illusory triangles, we rotated the reconstructions of the upward, rightward, and downward pointing triangles 90°, 180°, and 270° anti-clockwise, respectively, to move them into the reference frame of the leftward pointing triangle (Figure 1C). We obtained such reconstructions for every single participant, and averaged these to obtain group averaged reconstructions.

Supplemental References S1.

Brainard, D.H. (1997). The Psychophysics Toolbox. Spatial Vision 10, 433-436.

S2.

Dumoulin, S.O., and Wandell, B.A. (2008). Population receptive field estimates in human visual cortex. NeuroImage 39, 647-660.

S3.

Sereno, M.I., Dale, A.M., Reppas, J.B., Kwong, K.K., Belliveau, J.W., Brady, T.J., Rosen, B.R., and Tootell, R.B. (1995). Borders of multiple visual areas in humans revealed by functional magnetic resonance imaging. Science 268, 889-893.

S4.

Engel, S.A., Glover, G.H., and Wandell, B.A. (1997). Retinotopic organization in human visual cortex and the spatial precision of functional MRI. Cereb Cortex 7, 181-192.

S5.

Teeuwisse, W.M., Brink, W.M., and Webb, A.G. (2012). Quantitative assessment of the effects of highpermittivity pads in 7 Tesla MRI of the brain. Magn Reson Med 67, 1285-1293.

S6.

Poser, B.A., Koopmans, P.J., Witzel, T., Wald, L.L., and Barth, M. (2010). Three dimensional echo-planar imaging at 7 Tesla. NeuroImage 51, 261-266.

S7.

Marques, J.P., Kober, T., Krueger, G., Van Der Zwaag, W., Van De Moortele, P.-F., and Gruetter, R. (2010). MP2RAGE, a self bias-field corrected sequence for improved segmentation and T1-mapping at high field. NeuroImage 49, 1271–1281.

S8.

Visser, E., Qin, S., and Zwiers, M. (2010). EPI distortion correction by constrained nonlinear coregistration improves group fMRI. Proc. 18th Ann. Mtg. of ISMRM 18, 3459.

S9.

Greve, D.N., and Fischl, B. (2009). Accurate and robust brain image alignment using boundary-based registration. NeuroImage 48, 63-72.

S10.

Van Mourik, T., Koopmans, P.J., and Norris, D.G. (2014). Improved cortical boundary registration for locally distorted fMRI. Proc. 22nd Ann. Mtg. of ISMRM 22, 0865.

S11.

Waehnert, M.D., Dinse, J., Weiss, M., Streicher, M.N., Waehnert, P., Geyer, S., Turner, R., and Bazin, P.L. (2014). Anatomically motivated modeling of cortical laminae. NeuroImage 93, 210-220.

S12.

Bok, S. (1929). Der Einfluß der in den Furchen und Windungen auftretenden Krümmungen der Großhirnrinde auf die Rindenarchitektur. Z Gesamte Neurol Psychiatr 12, 682-750.

S13.

Van Mourik, T., Van Der Eerden, J.P.J.M., and Norris, D.G. (2015). Laminar time course extraction over extended cortical areas. Proc. 23rd Ann. Mtg. of ISMRM 23, 2071.

S14.

Kleinnijenhuis, M., Van Mourik, T., Norris, D.G., Ruiter, D.J., Van Cappellen van Walsum, A.-M., and Barth, M. (2015). Diffusion tensor characteristics of gyrencephaly using high resolution diffusion MRI in vivo at 7T. NeuroImage 109, 378-387.

S15.

Koopmans, P.J., Barth, M., and Norris, D.G. (2010). Layer-Specific BOLD Activation in Human V1. Hum Brain Mapp 31, 1297-1304.

S16.

Lund, T.E., Norgaard, M.D., Rostrup, E., Rowe, J.B., and Paulson, O.B. (2005). Motion or activity: their role in intra- and inter-subject variation in fMRI. Neuroimage 26, 960-964.

S17.

Douglas, R.J., and Martin, K.A.C. (2004). Neuronal circuits of the neocortex. Annu Rev Neurosci 27, 419451.

S18.

Cousineau, D. (2005). Confidence intervals in within-subject designs: A simpler solution to Loftus and Masson's method. Tutorial in Quantitative Methods for Psychology 1, 42-45.

S19.

Morey, R.D. (2008). Confidence intervals from normalized data: A correction to Cousineau (2005). Tutorial in Quantitative Methods for Psychology 4, 61-64.

S20.

Kok, P., and De Lange, F.P. (2014). Shape perception simultaneously up- and downregulates neural activity in the primary visual cortex. Curr Biol 24, 1531-1535.

Selective Activation of the Deep Layers of the ... - Floris de Lange

Jan 28, 2016 - other areas of cognition such as memory [46]. ..... laptop. Population receptive field measurements. After the main experiment, participants were ...

2MB Sizes 0 Downloads 195 Views

Recommend Documents

Selective Activation of the Deep Layers of the Human ...
Jan 28, 2016 - ... K. (2010). Comparative cytoarchitectural ana- ..... functional volumes we obtained, using MrVista (http://white.stanford.edu/software/). Before ...

Uncorrected Proof - Floris de Lange
Abstract. □ Perception does not function as an isolated module but is tightly linked with other cognitive functions. Several studies have demonstrated an influence of language on motion percep- tion, but it remains debated at which level of process

Uncorrected Proof - Floris de Lange
Abstract. □ Perception does not function as an isolated module but is tightly linked with other cognitive functions. Several studies have demonstrated an influence of language on motion percep- tion, but it remains debated at which level of process

Low attention impairs optimal incorporation of prior ... - Floris de Lange
ferent visual discrimination experiments, when attention was directed .... Materials and methods ...... the data analysis, ensuring that we only considered subjects.

Language beyond action Journal of Physiology - Paris - Floris de Lange
We argue that the available empirical evidence does not appear to support these claims, ..... well suited for the processing of their most salient meaning aspects.

Language beyond action Journal of Physiology - Paris - Floris de Lange
namely the burst of energy generated during the release of the. 0928-4257/$ - see front ... A radical alternative to auditory accounts of speech perception was formulated by .... This implication is not supported by the facts. There is no decisive ..

Dynamic decoding of ongoing perception - Floris de Lange
imaging (fMRI) data using multivariate analysis methods (Norman. 36 ... Centre for Cognition, Cognitive Artificial Intelligence Department, Radboud University.

Neural correlates of observing joint actions with ... - Floris de Lange
Jun 5, 2015 - (>5 mm) or other technical difficulties at the time of data collection. .... of paired t-tests of contrast images to analyze the statistical difference ...

Expectations accelerate entry of visual stimuli into ... - Floris de Lange
Jun 26, 2015 - weaker evidence for confirmation (in a Bayesian scheme). Alternatively ... Here we report psychophysical data intended to .... Data analysis.

Expectations accelerate entry of visual stimuli into ... - Floris de Lange
Jun 26, 2015 - ployed in Experiments 3 and 4. Altogether, the experiment took approximately 60 min. Data analysis. In this experiment and in Experiment 2, ...

visual areas stimulation reduces resting state ... - Floris de Lange
Jul 24, 2013 - 50 articles, 20 of which can be accessed free at: This article cites ... possibility is in line with a number of recent studies that have demonstrated ...

Cerebral compensation during motor imagery in ... - Floris de Lange
conditions (left hands, right hands), each with a linear parametric modulation of .... The intensity threshold (i.e. at the voxel level) was set at p < 0.01 family-wise.

Early Visual Cortex as a Multiscale Cognitive ... - Floris de Lange
18 Jul 2016 - Abstract. Neurons in early visual cortical areas not only represent incoming visual in- formation but are also engaged by higher level cognitive processes, including attention, working memory, imagery, and decision-making. Are these cog

Attention induces conservative subjective biases in ... - Floris de Lange
Oct 23, 2011 - NATURE NEUROSCIENCE ADVANCE ONLINE PUBLICATION. 1 ... 3Department of Psychological and Brain Sciences, Johns Hopkins University, Baltimore, Maryland, USA. .... Lu, Z.L. & Dosher, B.A. Vision Res. ..... seated in a dimmed room about 60

Preference for Audiovisual Speech Congruency in ... - Floris de Lange
Auditory speech perception can be altered by concurrent visual information. .... METHODS .... analysis in an independent data set in which participants listened ...

visual areas stimulation reduces resting state ... - Floris de Lange
Jul 24, 2013 - Thus the final analysis included data from .... For the main analysis, we combined the differ-. 1812 ..... that transforms HR and FAR into z-scores.

Preference for Audiovisual Speech Congruency in ... - Floris de Lange
multisensory fusion (see Figure 1). ... fected by perceptual fusion of incongruent auditory and ..... Brugge, J. F., Volkov, I. O., Garell, P. C., Reale, R. A., &. Howard ...

Laminar fMRI: Applications for cognitive neuroscience - Floris de Lange
lates of a learned audiovisual association (tones and light) in infra- granular V1 layers (layers 5 and 6; Headley and Weinberger, 2015). With laminar fMRI, these ...

Indesign-The-Best-Of-Layers-Magazine-Tip-Of-The-Day.pdf
Page 1 of 3. Download ~-~-~-oo~~ eBook Indesign The Best Of Layers Magazine Tip Of The Day. [eBooks] Indesign The Best Of Layers Magazine Tip Of The. Day. INDESIGN THE BEST OF LAYERS MAGAZINE TIP OF THE DAY EBOOK AUTHOR BY JEFF. WITCHEL. Indesign The

Layers of Rock
Last spring, a team of geologists made a study of the rock layers exposed on the side of a cliff. Here are the notes one of the geologists made. “On this cliff, we can see six layers of sedimentary rock. There is an extrusion between the third and

Unconscious Activation of the Prefrontal No-Go Network - Journal of ...
Mar 17, 2010 - of 50 ms ranging from 100 to 700 ms. Inhibition rates were .... main experiment to rule out any effect of perceptual learning during the experiment. In this .... For illustration purposes, we show the BOLD time courses for all four ...

Modelling the influence of activation-induced apoptosis of CD4 ... - ORBi
sented by a set of nonlinear Ordinary Differential Equa- tions (ODEs) that aims at incorporating the influence of activation-induced apoptosis of CD4. + and CD8.

Modelling the influence of activation-induced apoptosis of CD4 ... - ORBi
Modelling the influence of activation-induced apoptosis of CD4. + and CD8. +. T-cells on the immune system response of a HIV infected patient. Guy-Bart Stan. ∗. , Florence Belmudes. #. , Raphaël Fonteneau. #. , Frederic Zeggwagh. +. ,. Marie-Anne

The influence of layers feed supplementation with ...
the eggs quality and selenium content in the egg ... of Organic Se in layers diet increased its content in the eggs. Key words: ..... selenium distribution in egg and.