5
Cygwin インストール⼿順(ネットワークインストールの場合)Ver.3.1.6 (2020.08) (1)ブラウザで Cygwin 公式サイトを表⽰し、”Install Cygwin の setup リンクをクリック ⇒操作選択で実⾏を押す。 (2)「はい」を押し、デバイス変更許可をする。 (3)Cygwin Setup 画⾯で次へ

Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

117

BibliographyAarts, E., Verhage, M., Veenvliet, J. V, Dolan, C. V, and van der Sluis, S. (2014). A solution to dependency: using

multilevel analysis to accommodate nested data. Nat. Neurosci. 17, 491–496.

Abbott, L.F., and Nelson, S.B. (2000). Synaptic plasticity: taming the beast. Nat. Neurosci. 3 Suppl, 1178–1183.

Abbott, L.F., and Regehr, W.G. (2004). Synaptic computation. Nature 431, 796–803.

Abbott, L.F., Varela, J. a, Sen, K., and Nelson, S.B. (1997). Synaptic depression and cortical gain control. Science 275, 220–224.

Abraham, W.C. (2003). How long will long-term potentiation last? Philos. Trans. R. Soc. Lond. B. Biol. Sci. 358, 735–744.

van Aerde, K.I., Mann, E.O., Canto, C.B., Heistek, T.S., Linkenkaer-Hansen, K., Mulder, A.B., van der Roest, M., Paulsen, O., Brussaard, A.B., and Mansvelder, H.D. (2009). Flexible spike timing of layer 5 neurons during dynamic beta oscillation shifts in rat prefrontal cortex. J. Physiol. 587, 5177–5196.

Albuquerque, E.X., Pereira, E.F., Mike, a, Eisenberg, H.M., Maelicke, a, and Alkondon, M. (2000). Neuronal nico-tinic receptors in synaptic functions in humans and rats: physiological and clinical relevance. Behav. Brain Res. 113, 131–141.

Alkondon, M., Pereira, E.F., Eisenberg, H.M., and Albuquerque, E.X. (2000). Nicotinic receptor activation in human cerebral cortical interneurons: a mechanism for inhibition and disinhibition of neuronal networks. J. Neurosci. 20, 66–75.

Allen, C., and Stevens, C.F. (1994). An evaluation of causes for unreliability of synaptic transmission. Proc. Natl. Acad. Sci. U. S. A. 91, 10380–10383.

Altemus, K.L., Lavenex, P., Ishizuka, N., and Amaral, D.G. (2005). Morphological characteristics and electrophysi-ological properties of CA1 pyramidal neurons in macaque monkeys. Neuroscience 136, 741–756.

Anderson, K., Bones, B., Robinson, B., Hass, C., Lee, H., Ford, K., Roberts, T.A., and Jacobs, B. (2009). The morphology of supragranular pyramidal neurons in the human insular cortex: A quantitative golgi study. Cereb. Cortex 19, 2131–2144.

Andjelic, S., Gallopin, T., Cauli, B., Hill, E.L., Roux, L., Badr, S., Hu, E., Tamás, G., and Lambolez, B. (2009). Glutama-tergic nonpyramidal neurons from neocortical layer VI and their comparison with pyramidal and spiny stellate neurons. J. Neurophysiol. 101, 641–654.

Page 2: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

118

Bibliography

Arimatsu, Y., Ishida, M., Kaneko, T., Ichinose, S., and Omori, A. (2003). Organization and development of cortico-cortical associative neurons expressing the orphan nuclear receptor Nurr1. J. Comp. Neurol. 466, 180–196.

Arroyo, S., Bennett, C., Aziz, D., Brown, S.P., and Hestrin, S. (2012). Prolonged disynaptic inhibition in the cortex mediated by slow, non-α7 nicotinic excitation of a specific subset of cortical interneurons. J. Neurosci. 32, 3859–3864.

Arroyo, S., Bennett, C., and Hestrin, S. (2014). Nicotinic modulation of cortical circuits. Front. Neural Circuits 8, 30.

Arsiero, M., Lüscher, H.-R., Lundstrom, B.N., and Giugliano, M. (2007). The impact of input fluctuations on the frequency-current relationships of layer 5 pyramidal neurons in the rat medial prefrontal cortex. J. Neurosci. 27, 3274–3284.

Ascoli, G. a (2006). Mobilizing the base of neuroscience data: the case of neuronal morphologies. Nat. Rev. Neurosci. 7, 318–324.

Azevedo, F. a C., Carvalho, L.R.B., Grinberg, L.T., Farfel, J.M., Ferretti, R.E.L., Leite, R.E.P., Filho, W.J., Lent, R., and Herculano-Houzel, S. (2009). Equal numbers of neuronal and nonneuronal cells make the human brain an isometrically scaled-up primate brain. J. Comp. Neurol. 513, 532–541.

Babadi, B., and Abbott, L.F. (2010). Intrinsic stability of temporally shifted spike-timing dependent plasticity. PLoS Comput. Biol. 6.

Bailey, C.D.C., Alves, N.C., Nashmi, R., De Biasi, M., and Lambe, E.K. (2012). Nicotinic α5 subunits drive develop-mental changes in the activation and morphology of prefrontal cortex layer VI neurons. Biol. Psychiatry 71, 120–128.

Balter, M. (2007). Brain Evolution Studies Go Micro. Science (80-. ). 315, 1208–1211.

Banerjee, A., Meredith, R.M., Rodríguez-Moreno, A., Mierau, S.B., Auberson, Y.P., and Paulsen, O. (2009). Double dissociation of spike timing-dependent potentiation and depression by subunit-preferring NMDA receptor antagonists in mouse barrel cortex. Cereb. Cortex 19, 2959–2969.

Barton, R.A., and Harvey, P.H. (2000). Mosaic evolution of brain structure in mammals. Nature 405, 1055–1058.

Batsikadze, G., Paulus, W., Grundey, J., Kuo, M.-F., and Nitsche, M. a (2014). Effect of the Nicotinic α4β2-receptor Partial Agonist Varenicline on Non-invasive Brain Stimulation-Induced Neuroplasticity in the Human Motor Cortex. Cereb. Cortex 3249–3259.

Bayés, À., Collins, M.O., Croning, M.D.R., van de Lagemaat, L.N., Choudhary, J.S., and Grant, S.G.N. (2012). Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic Proteins. PLoS One 7.

Bear, M.F., and Abraham, W.C. (1996). Long-Term Depression in Hippocampus. Annu. Rev. Neurosci. 19, 437–462.

De Beaumont, L., Tremblay, S., Poirier, J., Lassonde, M., and Théoret, H. (2012). Altered bidirectional plasticity and reduced implicit motor learning in concussed athletes. Cereb. Cortex 22, 112–121.

Bekkers, J.M., and Häusser, M. (2007). Targeted dendrotomy reveals active and passive contributions of the dendritic tree to synaptic integration and neuronal output. Proc. Natl. Acad. Sci. U. S. A. 104, 11447–11452.

Bell, C.C., Han, V.Z., Sugawara, Y., and Grant, K. (1997). Synaptic plasticity in a cerebellum-like structure depends on temporal order. Nature 387, 278–281.

Benavides-Piccione, R., Fernaud-Espinosa, I., Robles, V., Yuste, R., and DeFelipe, J. (2012). Age-Based Comparison of Human Dendritic Spine Structure Using Complete Three-Dimensional Reconstructions. Cereb. Cortex 23, 1798–1810.

Bender, V.A., Bender, K.J., Brasier, D.J., and Feldman, D.E. (2006). Two coincidence detectors for spike timing-dependent plasticity in somatosensory cortex. J. Neurosci. 26, 4166–4177.

Bennett, C., Arroyo, S., Berns, D., and Hestrin, S. (2012). Mechanisms generating dual-component nicotinic EPSCs in cortical interneurons. J. Neurosci. 32, 17287–17296.

Benwell, M. (1985). Central nicotine binding sites: A study of post-mortem stability. Neuropharmacology 24, 1135–1137.

Benwell, M.E., Balfour, D.J., and Anderson, J.M. (1988). Evidence that tobacco smoking increases the density of (-)-[3H]nicotine binding sites in human brain. J. Neurochem. 50, 1243–1247.

Bi, G.Q., and Poo, M.M. (1998). Synaptic modifications in cultured hippocampal neurons: dependence on spike timing, synaptic strength, and postsynaptic cell type. J. Neurosci. 18, 10464–10472.

Page 3: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

119

Bibliography

Bianchi, S., Stimpson, C.D., Bauernfeind, A.L., Schapiro, S.J., Baze, W.B., McArthur, M.J., Bronson, E., Hopkins, W.D., Semendeferi, K., Jacobs, B., et al. (2013). Dendritic morphology of pyramidal neurons in the chimpanzee neocortex: Regional specializations and comparison to humans. Cereb. Cortex 23, 2429–2436.

Bliss, T. V, and Collingridge, G.L. (1993). A synaptic model of memory: long-term potentiation in the hippocampus. Nature 361, 31–39.

Bloem, B., Schoppink, L., Rotaru, D.C., Faiz, A., Hendriks, P., Mansvelder, H.D., van de Berg, W.D.J., and Wouterlood, F.G. (2014). Topographic Mapping between Basal Forebrain Cholinergic Neurons and the Medial Prefrontal Cortex in Mice. J. Neurosci. 34, 16234–16246.

Borst, a, and Theunissen, F.E. (1999). Information theory and neural coding. Nat. Neurosci. 2, 947–957.

Bortone, D.S., Olsen, S.R., and Scanziani, M. (2014). Translaminar inhibitory cells recruited by layer 6 corticotha-lamic neurons suppress visual cortex. Neuron 82, 474–485.

Bozek, K., Wei, Y., Yan, Z., Liu, X., Xiong, J., Sugimoto, M., Tomita, M., Pääbo, S., Sherwood, C.C., Hof, P.R., et al. (2015). Organization and Evolution of Brain Lipidome Revealed by Large-Scale Analysis of Human, Chim-panzee, Macaque, and Mouse Tissues. Neuron 695–702.

Braak, H. (1980). Architectonics of the Human Telencephalic Cortex (Berlin: Heidelberg: Springer).

Breese, C.R., Marks, M.J., Logel, J., Adams, C.E., Sullivan, B., Collins, a C., and Leonard, S. (1997). Effect of smoking history on [3H]nicotine binding in human postmortem brain. J. Pharmacol. Exp. Ther. 282, 7–13.

Brunet, M., Guy, F., Pilbeam, D., Mackaye, H.T., Likius, A., Ahounta, D., Beauvilain, A., Blondel, C., Bocherens, H., Boisserie, J.-R., et al. (2002). A new hominid from the Upper Miocene of Chad, Central Africa. Nature 418, 145–151.

Bruno, R.M., Hahn, T.T.G., Wallace, D.J., de Kock, C.P.J., and Sakmann, B. (2009). Sensory experience alters specific branches of individual corticocortical axons during development. J. Neurosci. Off. J. Soc. Neurosci. 29, 3172–3181.

Buxhoeveden, D.P., Switala, A.E., Roy, E., Litaker, M., and Casanova, M.F. (2001). Morphological differences between minicolumns in human and nonhuman primate cortex. Am. J. Phys. Anthropol. 115, 361–371.

Buckmaster PS, Amaral DG (2001) Intracellular recording and labeling of mossy cells and proximal CA3 pyramidal cells in macaque monkeys. J Comp Neurol 430:264–281.

Caillard, O., Moreno, H., Schwaller, B., Llano, I., Celio, M.R., and Marty, a (2000). Role of the calcium-binding protein parvalbumin in short-term synaptic plasticity. Proc. Natl. Acad. Sci. U. S. A. 97, 13372–13377.

Callaway, E.M., and Yuste, R. (2002). Stimulating neurons with light. Curr. Opin. Neurobiol. 12, 587–592.

Caporale, N., and Dan, Y. (2008). Spike timing-dependent plasticity: a Hebbian learning rule. Annu. Rev. Neurosci. 31, 25–46.

Carnevale, N.T., and Hines, M.L. (2006). The NEURON Book. Neuron 30, 457.

Castillo, J. Del, and Katz, B. (1954). Quantal components of the end-plate potential. J. Physiol. 124, 560–573.

Catterall, W.A. (2011). Voltage-Gated Calcium Channels.

Chen, X., Leischner, U., Rochefort, N.L., Nelken, I., and Konnerth, A. (2011). Functional mapping of single spines in cortical neurons in vivo. Nature 475, 501–505.

Clowry, G., Molnár, Z., and Rakic, P. (2010). Renewed focus on the developing human neocortex. J. Anat. 217, 276–288.

Conde, V., Vollmann, H., Taubert, M., Sehm, B., Cohen, L.G., Villringer, A., and Ragert, P. (2013). Reversed timing-dependent associative plasticity in the human brain through interhemispheric interactions. J. Neurophysiol. 2260–2271.

Cooke, S.F., and Bliss, T.V.P. (2006). Plasticity in the human central nervous system. Brain 129, 1659–1673.

Couey, J.J., Meredith, R.M., Spijker, S., Poorthuis, R.B., Smit, A.B., Brussaard, A.B., and Mansvelder, H.D. (2007). Distributed network actions by nicotine increase the threshold for spike-timing-dependent plasticity in prefrontal cortex. Neuron 54, 73–87.

Cuntz, H., Forstner, F., Borst, A., and H??usser, M. (2010). One rule to grow them all: A general theory of neuronal branching and its practical application. PLoS Comput. Biol. 6.

Page 4: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

120

Bibliography

Dan, Y., and Poo, M.-M. (2004). Spike timing-dependent plasticity of neural circuits. Neuron 44, 23–30.

DeFelipe, J. (1999). Estimation of the number of synapses in the cerebral cortex: Methodological considerations. Cereb. Cortex 9, 722–732.

DeFelipe, J. (2011). The evolution of the brain, the human nature of cortical circuits, and intellectual creativity. Front. Neuroanat. 5, 29.

DeFelipe, J., Alonso-Nanclares, L., and Arellano, J.I. (2003). Microstructure of the neocortex: comparative aspects. J. Neurocytol. 31, 299–316.

Destexhe, A., Rudolph, M., Fellous, J.M., and Sejnowski, T.J. (2001). Fluctuating synaptic conductances recreate in vivo-like activity in neocortical neurons. Neuroscience 107, 13–24.

Destexhe, A., Rudolph, M., and Paré, D. (2003). The high-conductance state of neocortical neurons in vivo. Nat. Rev. Neurosci. 4, 739–751.

Dittman, J.S., and Regehr, W.G. (1998). Calcium dependence and recovery kinetics of presynaptic depression at the climbing fiber to Purkinje cell synapse. J. Neurosci. 18, 6147–6162.

Drachman, D.A. (2005). Do we have brain to spare? Neurology 64, 2004–2005.

Duan, H., Wearne, S.L., Morrison, J.H., and Hof, P.R. (2002). Quantitative analysis of the dendritic morphology of corticocortical projection neurons in the macaque monkey association cortex. Neuroscience 114, 349–359.

Duan, H., Wearne, S.L., Rocher, A.B., Macedo, A., Morrison, J.H., and Hof, P.R. (2003). Age-related dendritic and spine changes in corticocortically projecting neurons in macaque monkeys. Cereb. Cortex 13, 950–961.

Dudek, S.M., and Bear, M.F. (1993). Bidirectional long-term modification of synaptic effectiveness in the adult and immature hippocampus. J. Neurosci. 13, 2910–2918.

Egger, V., Feldmeyer, D., and Sakmann, B. (1999). Coincidence detection and changes of synaptic efficacy in spiny stellate neurons in rat barrel cortex. Nat. Neurosci. 2, 1098–1105.

Egger, V., Nevian, T., and Bruno, R.M. (2008). Subcolumnar dendritic and axonal organization of spiny stellate and star pyramid neurons within a barrel in rat somatosensory cortex. Cereb. Cortex 18, 876–889.

Elahi, B., Gunraj, C., and Chen, R. (2012). Short-interval intracortical inhibition blocks long-term potentiation induced by paired associative stimulation. J. Neurophysiol. 107, 1935–1941.

van Elburg, R. a J., and van Ooyen, A. (2010). Impact of dendritic size and dendritic topology on burst firing in pyramidal cells. PLoS Comput. Biol. 6, 1–19.

Elston, G.N. (2007). Specialization of the Neocortical Pyramidal Cell during Primate Evolution. In Evolution of Nervous Systems, J.H. Kaas, ed. (Oxford: Elsevier), pp. 191–242.

Elston, G.N., and Fujita, I. (2014). Pyramidal cell development: postnatal spinogenesis, dendritic growth, axon growth, and electrophysiology. Front. Neuroanat. 8, 1–20.

Elston, G.N., Benavides-Piccione, R., and DeFelipe, J. (2001). The pyramidal cell in cognition: a comparative study in human and monkey. J. Neurosci. 21, RC163.

Elston, G.N., Benavides-Piccione, R., Elston, A., Manger, P.R., and Defelipe, J. (2011). Pyramidal cells in prefrontal cortex of primates: marked differences in neuronal structure among species. Front. Neuroanat. 5, 2.

von Engelhardt, J., Eliava, M., Meyer, A.H., Rozov, A., and Monyer, H. (2007). Functional characterization of intrinsic cholinergic interneurons in the cortex. J. Neurosci. 27, 5633–5642.

von Engelhardt, J., Mack, V., Sprengel, R., Kavenstock, N., Li, K.W., Stern-Bach, Y., Smit, A.B., Seeburg, P.H., and Monyer, H. (2010). CKAMP44: A Brain-Specific Protein Attenuating Short-Term Synaptic Plasticity in the Dentate Gyrus. Science (80-. ). 327, 1518–1522.

Eyal, G., Verhoog, M., Testa-Silva, G., Deitcher, Y., Lodder, J., Benavides-Piccione, R., Morales, J., DeFelipe, J., De Kock, C., Mansvelder, H., et al. Human L2/3 pyramidal cells possess unique membrane and dendritic proper-ties with augmented computational capabilities. Submitt. Manuscr.

Eyal, G., Mansvelder, H.D., de Kock, C.P.J., and Segev, I. (2014). Dendrites impact the encoding capabilities of the axon. J. Neurosci. 34, 8063–8071.

Filevich, O., Salierno, M., and Etchenique, R. (2010). A caged nicotine with nanosecond range kinetics and visible light sensitivity. J. Inorg. Biochem. 104, 1248–1251.

Page 5: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

121

Bibliography

Finlay, B.L., and Darlington, R.B. (1995). Linked regularities in the development and evolution of mammalian brains. Science 268, 1578–1584.

Fino, E., and Venance, L. (2010). Spike-timing dependent plasticity in the striatum. Front. Synaptic Neurosci.

Fino, E., Paille, V., Cui, Y., Morera-Herreras, T., Deniau, J.-M., and Venance, L. (2010). Distinct coincidence detec-tors govern the corticostriatal spike timing-dependent plasticity. J. Physiol. 588, 3045–3062.

Fischl, B., and Dale, A.M. (2000). Measuring the thickness of the human cerebral cortex from magnetic resonance images. Proc. Natl. Acad. Sci. U. S. A. 97, 11050–11055.

Fonseca-Azevedo, K., and Herculano-Houzel, S. (2012). Metabolic constraint imposes tradeoff between body size and number of brain neurons in human evolution. Proc. Natl. Acad. Sci. 1–6.

Fourcaud-trocme, N., Hansel, D., Vreeswijk, C. Van, and Brunel, N. (2003). How Spike Generation Mechanisms Determine the Neuronal Response to Fluctuating Inputs. J. Neurosci. 23, 11628–11640.

Froemke, R.C., and Dan, Y. (2002). Spike-timing-dependent synaptic modification induced by natural spike trains. Nature 416, 433–438.

Froemke, R.C., Poo, M.-M., and Dan, Y. (2005). Spike-timing-dependent synaptic plasticity depends on dendritic location. Nature 434, 221–225.

Froemke, R.C., Letzkus, J.J., Kampa, B.M., Hang, G.B., and Stuart, G.J. (2010). Dendritic synapse location and neocortical spike-timingdependent plasticity. Front. Synaptic Neurosci.

Fucile, S. (2004). Ca2+ permeability of nicotinic acetylcholine receptors. Cell Calcium 35, 1–8.

Fuenzalida, M., Fernández de Sevilla, D., Couve, A., and Buño, W. (2010). Role of AMPA and NMDA receptors and back-propagating action potentials in spike timing-dependent plasticity. J. Neurophysiol. 103, 47–54.

Fuhrmann, G., Segev, I., Markram, H., and Tsodyks, M. (2002). Coding of temporal information by activity-depen-dent synapses. J. Neurophysiol. 87, 140–148.

Fuhrmann, G., Cowan, A., Segev, I., Tsodyks, M., and Stricker, C. (2004). Multiple mechanisms govern the dynamics of depression at neocortical synapses of young rats. J. Physiol. 557, 415–438.

Gabbott, P.L. a, Warner, T. a, Jays, P.R.L., Salway, P., and Busby, S.J. (2005). Prefrontal cortex in the rat: projections to subcortical autonomic, motor, and limbic centers. J. Comp. Neurol. 492, 145–177.

Ge, S., and Dani, J. a (2005). Nicotinic acetylcholine receptors at glutamate synapses facilitate long-term depres-sion or potentiation. J. Neurosci. 25, 6084–6091.

Gentet, L.J., Stuart, G.J., and Clements, J.D. (2000). Direct measurement of specific membrane capacitance in neurons. Biophys. J. 79, 314–320.

Genzen, J.R., and McGehee, D.S. (2003). Short- and long-term enhancement of excitatory transmission in the spinal cord dorsal horn by nicotinic acetylcholine receptors. Proc Natl Acad Sci U S A 100, 6807–6812.

Geschwind, D.H., and Rakic, P. (2013). Cortical evolution: Judge the brain by its cover. Neuron 80, 633–647.

Gonzalez-Burgos, G., Kroener, S., Seamans, J.K., Lewis, D. a, and Barrionuevo, G. (2005). Dopaminergic modula-tion of short-term synaptic plasticity in fast-spiking interneurons of primate dorsolateral prefrontal cortex. J. Neurophysiol. 94, 4168–4177.

Goriounova, N.A., and Mansvelder, H.D. (2012). Nicotine exposure during adolescence leads to short- and long-term changes in spike timing-dependent plasticity in rat prefrontal cortex. J. Neurosci. 32, 10484–10493.

Gotti, C., Zoli, M., and Clementi, F. (2006). Brain nicotinic acetylcholine receptors: native subtypes and their relevance. Trends Pharmacol. Sci. 27, 482–491.

Grady, S.R., Wageman, C.R., Patzlaff, N.E., and Marks, M.J. (2012). Low concentrations of nicotine differentially desensitize nicotinic acetylcholine receptors that include α5 or α6 subunits and that mediate synaptosomal neurotransmitter release. Neuropharmacology 62, 1935–1943.

Gray, R., Rajan, a S., Radcliffe, K. a, Yakehiro, M., and Dani, J. a (1996). Hippocampal synaptic transmission enhanced by low concentrations of nicotine. Nature 383, 713–716.

Griguoli, M., Cellot, G., and Cherubini, E. (2013). In hippocampal oriens interneurons anti-Hebbian long-term potentiation requires cholinergic signaling via α7 nicotinic acetylcholine receptors. J. Neurosci. 33, 1044–1049.

Page 6: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

122

Bibliography

Groh, A., Meyer, H.S., Schmidt, E.F., Heintz, N., Sakmann, B., and Krieger, P. (2010). Cell-type specific properties of pyramidal neurons in neocortex underlying a layout that is modifiable depending on the cortical area. Cereb. Cortex 20, 826–836.

Grundey, J., Thirugnanasambandam, N., Kaminsky, K., Drees, A., Skwirba, A.C., Lang, N., Paulus, W., and Nitsche, M. a (2012). Rapid effect of nicotine intake on neuroplasticity in non-smoking humans. Front. Pharmacol. 3, 186.

Gu, Z., and Yakel, J.L. (2011). Timing-dependent septal cholinergic induction of dynamic hippocampal synaptic plasticity. Neuron 71, 155–165.

Guillem, K., Bloem, B., Poorthuis, R.B., Loos, M., Smit, A.B., Maskos, U., Spijker, S., and Mansvelder, H.D. (2011). Nicotinic acetylcholine receptor β2 subunits in the medial prefrontal cortex control attention. Science 333, 888–891.

Gulledge, A.T., Park, S.B., Kawaguchi, Y., and Stuart, G.J. (2007). Heterogeneity of phasic cholinergic signaling in neocortical neurons. J. Neurophysiol. 97, 2215–2229.

Hamilton, T.J., Wheatley, B.M., Sinclair, D.B., Bachmann, M., Larkum, M.E., and Colmers, W.F. (2010). Dopamine modulates synaptic plasticity in dendrites of rat and human dentate granule cells. Proc. Natl. Acad. Sci. 107, 18185–18190.

Han, Jin-hee, Josselyn, S. (2009). Selective Erasure of fear memory. 323, 1492–1496.

Hangya, B., Ranade, S.P., Lorenc, M., and Kepecs, A. (2015). Central Cholinergic Neurons Are Rapidly Recruited by Reinforcement Feedback. Cell 162, 1155–1168.

Harris, K.D., and Shepherd, G.M.G. (2015). The neocortical circuit: themes and variations. Nat. Neurosci. 18, 170–181.

Harvey, P., and Krebs (1990). Comparing brains. Science (80-. ). 249, 140–146.

Hasan, M.T., Hernández-González, S., Dogbevia, G., Treviño, M., Bertocchi, I., Gruart, A., and Delgado-García, J.M. (2013). Role of motor cortex NMDA receptors in learning-dependent synaptic plasticity of behaving mice. Nat. Commun. 4, 2258.

Hasselmo, M.E. (2006). The role of acetylcholine in learning and memory. Curr. Opin. Neurobiol. 16, 710–715.

Hasselmo, M.E., and Howard Eichenbaum (2005). Hippocampal mechanisms for the context-dependent retrieval of episodes. Neural Networks 18, 1172–1190.

Häusser, M. (2001). Synaptic function: Dendritic democracy. Curr. Biol. 11.

Hay, Y. a., Lambolez, B., and Tricoire, L. (2015). Nicotinic Transmission onto Layer 6 Cortical Neurons Relies on Synaptic Activation of Non- 7 Receptors. Cereb. Cortex 2, 1–14.

Hebb, D.O. (1949). The Organization of Behaviour. Organization 62.

Hedrick, T., and Waters, J. (2015). Acetylcholine excites neocortical pyramidal neurons via nicotinic receptors. J. Neurophysiol. 2195–2209.

Heishman, S.J., Kleykamp, B. a., and Singleton, E.G. (2010). Meta-analysis of the acute effects of nicotine and smoking on human performance. Psychopharmacology (Berl). 210, 453–469.

Herculano-Houzel, S. (2009). The human brain in numbers: a linearly scaled-up primate brain. Front. Hum. Neurosci. 3, 31.

Herculano-Houzel, S. (2012). The remarkable, yet not extraordinary, human brain as a scaled-up primate brain and its associated cost. Proc. Natl. Acad. Sci. 109, 10661–10668.

Herculano-Houzel, S., Manger, P.R., and Kaas, J.H. (2014). Brain scaling in mammalian evolution as a consequence of concerted and mosaic changes in numbers of neurons and average neuronal cell size. Front. Neuroanat. 8, 1–28.

Hofman, M.A. (1985). Size and Shape of the Cerebral Cortex in Mammals. Part I. Brain. Behav. Evol. 27, 28–40.

Hofman, M.A. (1988). Size and Shape of the Cerebral Cortex in Mammals. Part II. Brain. Behav. Evol. 32, 17–26.

Hogg, R.C., Raggenbass, M., and Bertrand, D. (2003). Nicotinic acetylcholine receptors: from structure to brain function. Rev. Physiol. Biochem. Pharmacol. 147, 1–46.

Horikawa, K., and Armstrong, W.E.E. (1988). A versatile means of intracellular labeling: injection of biocytin and its detection with avidin conjugates. J. Neurosci. Methods 25, 1–11.

Page 7: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

123

Bibliography

Howe, W.M., Berry, A.S., Francois, J., Gilmour, G., Carp, J.M., Tricklebank, M., Lustig, C., and Sarter, M. (2013). Prefrontal cholinergic mechanisms instigating shifts from monitoring for cues to cue-guided performance: converging electrochemical and FMRI evidence from rats and humans. J. Neurosci. 33, 8742–8752.

Hox, J. (2002). Multilevel analysis: techniques and applications.

Huang, M., Volgushev, M., and Wolf, F. (2012). A small fraction of strongly cooperative sodium channels boosts neuronal encoding of high frequencies. PLoS One 7.

Huang, S., Rozas, C., Trevino, M., Contreras, J., Yang, S., Song, L., Yoshioka, T., Lee, H.-K., and Kirkwood, A. (2014). Associative Hebbian Synaptic Plasticity in Primate Visual Cortex. J. Neurosci. 34, 7575–7579.

Ilin, V., Malyshev, A., Wolf, F., and Volgushev, M. (2013). Fast Computations in Cortical Ensembles Require Rapid Initiation of Action Potentials. J. Neurosci. 33, 2281–2292.

Insel, T., Landis, S., and Collins, F. (2013). The NIH brain initiative. Science (80-. ). 340, 687–688.

Ison, M.J., Quian Quiroga, R., and Fried, I. (2015). Rapid Encoding of New Memories by Individual Neurons in the Human Brain. Neuron 87, 220–230.

Jack, J.J.B., Noble, D., and Tsien, R.W. (1975). Electric Current Flow in Excitable Cells. (Clarendon Press Oxford).

Jacobs, B., Driscoll, L., and Schall, M. (1997). Life-span dendritic and spine changes in areas 10 and 18 of human cortex: A quantitative golgi study. J. Comp. Neurol. 386, 661–680.

Jacobs, B., Schall, M., Prather, M., Kapler, E., Driscoll, L., Baca, S., Jacobs, J., Ford, K., Wainwright, M., and Treml, M. (2001). Regional dendritic and spine variation in human cerebral cortex: a quantitative golgi study. Cereb. Cortex 11, 558–571.

Janmaat, K.R.L., Byrne, R.W., and Zuberbühler, K. (2006). Primates take weather into account when searching for fruits. Curr. Biol. 16, 1232–1237.

Janmaat, K.R.L., Ban, S.D., and Boesch, C. (2013). Taï chimpanzees use botanical skills to discover fruit: what we can learn from their mistakes. Anim. Cogn.

Ji, D., Lape, R., and Dani, J.A. (2001). Timing and Location of Nicotinic Activity Enhances or Depresses Hippo-campal Synaptic Plasticity. Neuron 31, 131–141.

John, J.L.S., Rosene, D.L., and Luebke, J.I. (1997). Morphology and electrophysiology of dentate granule cells in the rhesus monkey: Comparison with the rat. J. Comp. Neurol. 387, 136–147.

Johnson, M.B., Kawasawa, Y.I., Mason, C.E., Krsnik, Z., Coppola, G., Bogdanović, D., Geschwind, D.H., Mane, S.M., State, M.W., and Sestan, N. (2009). Functional and evolutionary insights into human brain development through global transcriptome analysis. Neuron 62, 494–509.

Jones, I.W., and Wonnacott, S. (2004). Precise localization of alpha7 nicotinic acetylcholine receptors on glutama-tergic axon terminals in the rat ventral tegmental area. J. Neurosci. 24, 11244–11252.

Josselyn, S. a. (2010). Continuing the search for the engram: Examining the mechanism of fear memories. J. Psychiatry Neurosci. 35, 221–228.

Kampa, B.M., Letzkus, J.J., and Stuart, G.J. (2006). Requirement of dendritic calcium spikes for induction of spike-timing-dependent synaptic plasticity. J. Physiol. 574, 283–290.

Kampa, B.M., Letzkus, J.J., and Stuart, G.J. (2007). Dendritic mechanisms controlling spike-timing-dependent synaptic plasticity. Trends Neurosci. 30, 456–463.

Kang, J. Il, Huppé-Gourgues, F., and Vaucher, E. (2014). Boosting visual cortex function and plasticity with acetylcholine to enhance visual perception. Front. Syst. Neurosci. 8, 1–14.

Kassam, S.M., Herman, P.M., Goodfellow, N.M., Alves, N.C., and Lambe, E.K. (2008). Developmental excitation of corticothalamic neurons by nicotinic acetylcholine receptors. J. Neurosci. 28, 8756–8764.

Kasthuri, N., Hayworth, K.J., Berger, D.R., Schalek, R.L., Conchello, J.A., Knowles-Barley, S., Lee, D., Vázquez-Reina, A., Kaynig, V., Jones, T.R., et al. (2015). Saturated Reconstruction of a Volume of Neocortex. Cell 162, 648–661.

Kato, H., Ito, Z., Matsuoka, S., and Sakurai, Y. (1973). ELECTRICAL ACTIVITIES OF NEURONS IN THE SLICED HUMAN CORTEX IN VITRO. Immunology 457–462.

Kawaguchi, Y., and Kubota, Y. (1997). GABAergic cell subtypes and their synaptic connections in rat frontal cortex. Cereb. Cortex 7, 476–486.

Page 8: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

124

Bibliography

Kilgard, M.P., and Merzenich, M.M. (1998). Cortical map reorganization enabled by nucleus basalis activity. Science 279, 1714–1718.

Kim, H.G., and Connors, B.W. (1993). Apical dendrites of the neocortex: correlation between sodium- and calcium-dependent spiking and pyramidal cell morphology. J. Neurosci. 13, 5301–5311.

Klug, A., Borst, J.G., Carlson, B.A., Kopp-Scheinpflug, C., Klyachko, V.A., and Xu-Friedman, M.A. (2012). How do short-term changes at synapses fine-tune information processing? J Neurosci 32, 14058–14063.

Klyachko, V.A., and Stevens, C.F. (2006). Excitatory and feed-forward inhibitory hippocampal synapses work syner-gistically as an adaptive filter of natural spike trains. PLoS Biol. 4, 1187–1200.

Koch, G., Ponzo, V., Di Lorenzo, F., Caltagirone, C., and Veniero, D. (2013). Hebbian and Anti-Hebbian Spike-Timing-Dependent Plasticity of Human Cortico-Cortical Connections. J. Neurosci. 33, 9725–9733.

Köndgen, H., Geisler, C., Fusi, S., Wang, X.J., Lüscher, H.R., and Giugliano, M. (2008). The dynamical response properties of neocortical neurons to temporally modulated noisy inputs in vitro. Cereb. Cortex 18, 2086–2097.

Kumari, V., Gray, J. a., Ffytche, D.H., Mitterschiffthaler, M.T., Das, M., Zachariah, E., Vythelingum, G.N., Williams, S.C.R., Simmons, A., and Sharma, T. (2003). Cognitive effects of nicotine in humans: An fMRI study. Neuro-image 19, 1002–1013.

Kuryatov, A., Onksen, J., and Lindstrom, J. (2008). Roles of accessory subunits in α4β2* nicotinic receptors. Mol. Pharmacol. 74, 132–143.

Kwan, K.Y., Lam, M.M.S., Johnson, M.B., Dube, U., Shim, S., Ra??in, M.R., Sousa, A.M.M., Fertuzinhos, S., Chen, J.G., Arellano, J.I., et al. (2012). Species-dependent posttranscriptional regulation of NOS1 by FMRP in the developing cerebral cortex. Cell 149, 899–911.

Lambe, E.K., Picciotto, M.R., and Aghajanian, G.K. (2003). Nicotine induces glutamate release from thalamocor-tical terminals in prefrontal cortex. Neuropsychopharmacology 28, 216–225.

Larkum, M. (2013). A cellular mechanism for cortical associations: An organizing principle for the cerebral cortex. Trends Neurosci. 36, 141–151.

Larkum, M.E., Zhu, J.J., and Sakmann, B. (1999a). A new cellular mechanism for coupling inputs arriving at different cortical layers. Nature 398, 338–341.

Larkum, M.E., Kaiser, K.M., and Sakmann, B. (1999b). Calcium electrogenesis in distal apical dendrites of layer 5 pyramidal cells at a critical frequency of back-propagating action potentials. Proc. Natl. Acad. Sci. U. S. A. 96, 14600–14604.

Larkum, M.E., Zhu, J.J., and Sakmann, B. (2001). Dendritic mechanisms underlying the coupling of the dendritic with the axonal action potential initiation zone of adult rat layer 5 pyramidal neurons. J. Physiol. 533, 447–466.

Larkum, M.E., Nevian, T., Sandler, M., Polsky, A., and Schiller, J. (2009). Synaptic integration in tuft dendrites of layer 5 pyramidal neurons: a new unifying principle. Science 325, 756–760.

Lavzin, M., Rapoport, S., Polsky, A., Garion, L., and Schiller, J. (2012). Nonlinear dendritic processing determines angular tuning of barrel cortex neurons in vivo. Nature 490, 397–401.

Ledergerber, D., and Larkum, M.E. (2010). Properties of layer 6 pyramidal neuron apical dendrites. J. Neurosci. 30, 13031–13044.

Lee, M.G., Hassani, O.K., Alonso, A., and Jones, B.E. (2005). Cholinergic basal forebrain neurons burst with theta during waking and paradoxical sleep. J. Neurosci. 25, 4365–4369.

Letzkus, J.J., Kampa, B.M., and Stuart, G.J. (2006). Learning rules for spike timing-dependent plasticity depend on dendritic synapse location. J. Neurosci. 26, 10420–10429.

Letzkus, J.J., Kampa, B.M., and Stuart, G.J. (2007). Does spike timing-dependent synaptic plasticity underlie memory formation? Clin. Exp. Pharmacol. Physiol. 34, 1070–1076.

Letzkus, J.J., Wolff, S.B.E., Meyer, E.M.M., Tovote, P., Courtin, J., Herry, C., and Lüthi, A. (2011). A disinhibitory microcircuit for associative fear learning in the auditory cortex. Nature 480, 331–335.

Levy, W.B., and Steward, O. (1983). Temporal contiguity requirements for long-term associative potentiation/depression in the hippocampus. Neuroscience 8, 791–797.

Linaro, D., Couto, J., and Giugliano, M. (2014). Command-line cellular electrophysiology for conventional and real-time closed-loop experiments. J. Neurosci. Methods 230, 5–19.

Page 9: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

125

Bibliography

Liu, X., Ramirez, S., Pang, P.T., Puryear, C.B., Govindarajan, A., Deisseroth, K., and Tonegawa, S. (2012). Optoge-netic stimulation of a hippocampal engram activates fear memory recall. Nature 484, 381–385.

Loebel, A., Silberberg, G., Helbig, D., Markram, H., Tsodyks, M., and Richardson, M.J.E. (2009). Multiquantal release underlies the distribution of synaptic efficacies in the neocortex. Front. Comput. Neurosci. 3, 27.

Logan, C.A. (1999). The altered rationale for the choice of a standard animal in experimental psychology: Henry H. Donaldson, Adolf Meyer, and “the” albino rat. Hist. Psychol. 2, 3–24.

Lu, M.-K., Tsai, C.-H., and Ziemann, U. (2012). Cerebellum to motor cortex paired associative stimulation induces bidirectional STDP-like plasticity in human motor cortex. Front. Hum. Neurosci. 6, 1–9.

Luebke, J.I., Medalla, M., Amatrudo, J.M., Weaver, C.M., Crimins, J.L., Hunt, B., Hof, P.R., and Peters, A. (2013). Age-Related Changes to Layer 3 Pyramidal Cells in the Rhesus Monkey Visual Cortex. Cereb. Cortex 3, 1–15.

Madisen, L., Mao, T., Koch, H., Zhuo, J., Berenyi, A., Fujisawa, S., Hsu, Y.-W. a, Garcia, A.J., Gu, X., Zanella, S., et al. (2012). A toolbox of Cre-dependent optogenetic transgenic mice for light-induced activation and silencing. Nat. Neurosci. 15, 793–802.

Magee, J.C. (2000). Dendritic integration of excitatory synaptic input. Nat. Rev. Neurosci. 1, 181–190.

Magee, J.C., and Cook, E.P. (2000). Somatic EPSP amplitude is independent of synapse location in hippocampal pyramidal neurons. Nat. Neurosci. 3, 895–903.

Magee, J.C., and Johnston, D. (1997). A synaptically controlled, associative signal for Hebbian plasticity in hippo-campal neurons. Science 275, 209–213.

Mainen, Z.F., and Sejnowski, T.J. (1996). Influence of dendritic structure on firing pattern in model neocortical neurons. Nature 382, 363–366.

Major, G., Larkum, M.E., and Schiller, J. (2013). Active properties of neocortical pyramidal neuron dendrites. Annu. Rev. Neurosci. 36, 1–24.

Malenka, R.C., and Nicoll, R. a. (1993). NMDA-receptor-dependent synaptic plasticity: Multiple forms and mecha-nisms. Trends Neurosci. 16, 521–527.

Manger, P.R., Spocter, M. a., and Patzke, N. (2013). The evolutions of large brain size in mammals: The “Over-700-gram club quartet.” Brain. Behav. Evol. 82, 68–78.

Mann, E.O., Suckling, J.M., Hajos, N., Greenfield, S.A., and Paulsen, O. (2005). Perisomatic feedback inhibi-tion underlies cholinergically induced fast network oscillations in the rat hippocampus in vitro. Neuron 45, 105–117.

Mansvelder, H.D., and McGehee, D.S. (2000). Long-term potentiation of excitatory inputs to brain reward areas by nicotine. Neuron 27, 349–357.

Markram, H. (2012). The Human Brain Project. Sci. Am. 306, 50–55.

Markram, H., Lübke, J., Frotscher, M., Sakmann, B., Lubke, J., Frotscher, M., and Sakmann, B. (1997). Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science 275, 213–215.

Markram, H., Wang, Y., and Tsodyks, M. (1998a). Differential signaling via the same axon of neocortical pyramidal neurons. Proc. Natl. Acad. Sci. U. S. A. 95, 5323–5328.

Markram, H., Gupta, a, Uziel, a, Wang, Y., and Tsodyks, M. (1998b). Information processing with frequency-dependent synaptic connections. Neurobiol. Learn. Mem. 70, 101–112.

Marx, M., and Feldmeyer, D. (2013). Morphology and physiology of excitatory neurons in layer 6b of the somato-sensory rat barrel cortex. Cereb. Cortex 23, 2803–2817.

Mayer, M.L., Westbrook, G.L., and Guthrie, P.B. (1984). Voltage-dependent block by Mg2+ of NMDA responses in spinal cord neurones. Nature 309, 261–263.

McCormick, D. a, Shu, Y., and Yu, Y. (2007). Neurophysiology: Hodgkin and Huxley model--still standing? Nature 445, E1–E2; discussion E2–E3.

McGehee, D., Heath, M., Gelber, S., Devay, P., and Role, L. (1995). Nicotine enhancement of fast excitatory synaptic transmission in CNS by presynaptic receptors. Science (80-. ). 269, 1692–1696.

Meredith, R.M., Floyer-Lea, A.M., and Paulsen, O. (2003). Maturation of long-term potentiation induction rules in rodent hippocampus: role of GABAergic inhibition. J. Neurosci. 23, 11142–11146.

Page 10: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

126

Bibliography

Meredith, R.M., Holmgren, C.D., Weidum, M., Burnashev, N., and Mansvelder, H.D. (2007). Increased threshold for spike-timing-dependent plasticity is caused by unreliable calcium signaling in mice lacking fragile X gene FMR1. Neuron 54, 627–638.

Meyer, H.S., Wimmer, V.C., Hemberger, M., Bruno, R.M., De Kock, C.P.J., Frick, A., Sakmann, B., and Helmstaedter, M. (2010). Cell type-specific thalamic innervation in a column of rat vibrissal cortex. Cereb. Cortex 20, 2287–2303.

De Miguel, C., and Henneberg, M. (2001). Variation in hominid brain size: how much is due to method? Homo 52, 3–58.

Mink, J.W., Blumenschine, R.J., and Adams, D.B. (1981). Ratio of central nervous system to body metabolism in vertebrates: its constancy and functional basis. Am. J. Physiol. 241, R203–R212.

Mohan, H., Verhoog, M.B., Doreswamy, K.K., Eyal, G., Aardse, R., Lodder, B.N., Goriounova, N. a., Asamoah, B., B. Brakspear, a. B.C., Groot, C., et al. (2015). Dendritic and Axonal Architecture of Individual Pyramidal Neurons across Layers of Adult Human Neocortex. Cereb. Cortex 25, 4839–4853.

Molnár, G., Oláh, S., Komlósi, G., Füle, M., Szabadics, J., Varga, C., Barzó, P., and Tamás, G. (2008). Complex Events Initiated by Individual Spikes in the Human Cerebral Cortex. PLoS Biol. 6, e222.

Morales, J., Benavides-Piccione, R., Dar, M., Fernaud, I., Rodriguez, A., Anton-Sanchez, L., Bielza, C., Larranaga, P., DeFelipe, J., Yuste, R., et al. (2014). Random positions of dendritic spines in human cerebral cortex. J. Neurosci. 34, 10078–10084.

Morishita, H., Miwa, J.M., Heintz, N., and Hensch, T.K. (2010). Lynx1, a cholinergic brake, limits plasticity in adult visual cortex. Science 330, 1238–1240.

Müller-Dahlhaus, F., Ziemann, U., and Classen, J. (2010). Plasticity resembling spike-timing dependent synaptic plasticity: the evidence in human cortex. Front. Synaptic Neurosci. 2, 34.

Müller-Dahlhaus, J.F.M., Orekhov, Y., Liu, Y., and Ziemann, U. (2008). Interindividual variability and age-depen-dency of motor cortical plasticity induced by paired associative stimulation. Exp. Brain Res. 187, 467–475.

Multani, P., Myers, R.H., Blume, H.W., Schomer, D.L., and Sotrel, A. (1994). Neocortical Dendritic Pathology in Human Partial Epilepsy: A Quantitative Golgi Study. Epilepsia 35, 728–736.

Narayanan, R.T., Egger, R., Johnson, A.S., Mansvelder, H.D., Sakmann, B., de Kock, C.P.J., and Oberlaender, M. (2015). Beyond Columnar Organization: Cell Type- and Target Layer-Specific Principles of Horizontal Axon Projection Patterns in Rat Vibrissal Cortex. Cereb. Cortex bhv053 – .

Naundorf, B., Wolf, F., and Volgushev, M. (2006). Unique features of action potential initiation in cortical neurons. Nature 440, 1060–1063.

Navarrete, A., van Schaik, C.P., and Isler, K. (2011). Energetics and the evolution of human brain size. Nature 480, 91–93.

Nemenman, I., Lewen, G.D., Bialek, W., and Van Steveninck, R.R.D.R. (2008). Neural coding of natural stimuli: Information at sub-millisecond resolution. PLoS Comput. Biol. 4.

Neuromorpho.org http://neuromorpho.org/neuroMorpho/index. jsp.

Nevian, T., and Sakmann, B. (2006). Spine Ca2+ signaling in spike-timing-dependent plasticity. J. Neurosci. 26, 11001–11013.

Nevian, T., Larkum, M.E., Polsky, A., and Schiller, J. (2007). Properties of basal dendrites of layer 5 pyramidal neurons: a direct patch-clamp recording study. Nat. Neurosci. 10, 206–214.

Nimchinsky, E.A., Gilissen, E., Allman, J.M., Perl, D.P., Erwin, J.M., and Hof, P.R. (1999). A neuronal morphologic type unique to humans and great apes. Proc. Natl. Acad. Sci. U. S. A. 96, 5268–5273.

Nishimura, Y., Perlmutter, S.I., Eaton, R.W., and Fetz, E.E. (2013). Spike-timing-dependent plasticity in primate corticospinal connections induced during free behavior. Neuron 80, 1301–1309.

Nithianantharajah, J., Komiyama, N.H., McKechanie, A., Johnstone, M., Blackwood, D.H., St Clair, D., Emes, R.D., van de Lagemaat, L.N., Saksida, L.M., Bussey, T.J., et al. (2013). Synaptic scaffold evolution generated compo-nents of vertebrate cognitive complexity. Nat. Neurosci. 16, 16–24.

Nonaka, A., Toyoda, T., Miura, Y., Hitora-Imamura, N., Naka, M., Eguchi, M., Yamaguchi, S., Ikegaya, Y., Matsuki, N., and Nomura, H. (2014). Synaptic plasticity associated with a memory engram in the basolateral amygdala. J. Neurosci. 34, 9305–9309.

Page 11: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

127

Bibliography

Nowak, L., Bregestovski, P., Ascher, P., Herbet, A., and Prochiantz, A. (1984). Magnesium gates glutamate-activated channels in mouse central neurones. Nature 307, 462–465.

Oberheim, N.A., Takano, T., Han, X., He, W., Lin, J.H.C., Wang, F., Xu, Q., Wyatt, J.D., Pilcher, W., Ojemann, J.G., et al. (2009). Uniquely hominid features of adult human astrocytes. J. Neurosci. 29, 3276–3287.

Oberlaender, M., Boudewijns, Z.S.R.M., Kleele, T., Mansvelder, H.D., Sakmann, B., and de Kock, C.P.J. (2011). Three-dimensional axon morphologies of individual layer 5 neurons indicate cell type-specific intracortical pathways for whisker motion and touch. Proc. Natl. Acad. Sci. U. S. A. 108, 4188–4193.

Ochoa, E.L., and O’Shea, S.M. (1994). Concomitant protein phosphorylation and endogenous acetylcholine release induced by nicotine: dependency on neuronal nicotinic receptors and desensitization. Cell. Mol. Neurobiol. 14, 315–340.

Olsen, S.R., Bortone, D.S., Adesnik, H., and Scanziani, M. (2012). Gain control by layer six in cortical circuits of vision. Nature 483, 47–52.

Ong, W.Y., and Garey, L.J. (1990). Pyramidal neurons are immunopositive for peptides, but not GABA, in the temporal cortex of the macaque monkey (Macaca fascicularis). 533, 24–41.

Paille, V., Fino, E., Du, K., Morera-Herreras, T., Perez, S., Kotaleski, J.H., and Venance, L. (2013). GABAergic circuits control spike-timing-dependent plasticity. J. Neurosci. 33, 9353–9363.

Pakkenberg, B., Pelvig, D., Marner, L., Bundgaard, M.J., Gundersen, H.J.G., Nyengaard, J.R., and Regeur, L. (2003). Aging and the human neocortex. In Experimental Gerontology, pp. 95–99.

Palmer, L.M., Shai, A.S., Reeve, J.E., Anderson, H.L., Paulsen, O., and Larkum, M.E. (2014). NMDA spikes enhance action potential generation during sensory input. Nat. Neurosci. 17, 383–390.

Parikh, V., Kozak, R., Martinez, V., and Sarter, M. (2007). Prefrontal Acetylcholine Release Controls Cue Detection on Multiple Timescales. Neuron 56, 141–154.

Pawlak, V., and Kerr, J.N.D. (2008). Dopamine receptor activation is required for corticostriatal spike-timing-dependent plasticity. J. Neurosci. 28, 2435–2446.

Pawlak, V., Wickens, J.R., Kirkwood, A., and Kerr, J.N.D. (2010). Timing is not everything: Neuromodulation opens the STDP gate. Front. Synaptic Neurosci. 2, 1–14.

van Pelt, J., van Ooyen, A., and Uylings, H.B.M. (2014). Axonal and dendritic density field estimation from incom-plete single-slice neuronal reconstructions. Front. Neuroanat. 8, 54.

Perry, D.C., Dávila-García, M.I., Stockmeier, C. a, and Kellar, K.J. (1999). Increased nicotinic receptors in brains from smokers: membrane binding and autoradiography studies. J. Pharmacol. Exp. Ther. 289, 1545–1552.

Petanjek, Z., Judaš, M., Kostović, I., and Uylings, H.B.M. (2008). Lifespan alterations of basal dendritic trees of pyramidal neurons in the human prefrontal cortex: A layer-specific pattern. Cereb. Cortex 18, 915–929.

Petanjek, Z., Judas, M., Simic, G., Rasin, M.R., Uylings, H.B.M., Rakic, P., and Kostovic, I. (2011). Extraordinary neoteny of synaptic spines in the human prefrontal cortex. Proc. Natl. Acad. Sci. U. S. A. 108, 13281–13286.

Peters, a, and Feldman, M.L. (1976). The projection of the lateral geniculate nucleus to area 17 of the rat cerebral cortex. I. General description. J. Neurocytol. 6, 669–689.

Petreanu, L., Huber, D., Sobczyk, A., and Svoboda, K. (2007). Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nat. Neurosci. 10, 663–668.

Petreanu, L., Mao, T., Sternson, S.M., and Svoboda, K. (2009). The subcellular organization of neocortical excit-atory connections. Nature 457, 1142–1145.

Pilbeam, D., and Gould, S.J. (1974). Size and Scaling in Human Evolution. Science (80-. ). 186, 892–901.

Poirazi, P., and Mel, B.W. (2001). Impact of active dendrites and structural plasticity on the memory capacity of neural tissue. Neuron 29, 779–796.

Poorthuis, R.B., Bloem, B., Schak, B., Wester, J., De Kock, C.P.J., and Mansvelder, H.D. (2013a). Layer-specific modulation of the prefrontal cortex by nicotinic acetylcholine receptors. Cereb. Cortex 23, 148–161.

Poorthuis, R.B., Bloem, B., Verhoog, M.B., and Mansvelder, H.D. (2013b). Layer-specific interference with cholin-ergic signaling in the prefrontal cortex by smoking concentrations of nicotine. J. Neurosci. 33, 4843–4853.

Poorthuis, R.B., Enke, L., and Letzkus, J.J. (2014). Cholinergic circuit modulation through differential recruitment of neocortical interneuron types during behaviour. J. Physiol. 00, 1–10.

Page 12: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

128

Bibliography

Povysheva N V, Zaitsev a V, Rotaru DC, Gonzalez-Burgos G, Lewis D a, Krimer LS (2008) Parvalbumin-positive basket interneurons in monkey and rat prefrontal cortex. J Neurophysiol 100:2348–2360.

Povysheva N V, Zaitsev A V, Kröner S, Krimer OA, Rotaru DC, Gonzalez-Burgos G, Lewis DA, Krimer LS (2007) Electrophysiological differences between neurogliaform cells from monkey and rat prefrontal cortex. J Neuro-physiol 97:1030–1039 Available at: http://www.ncbi.nlm.nih.gov/pubmed/17122314.

Premack, D. (2007). Human and animal cognition: continuity and discontinuity. Proc. Natl. Acad. Sci. U. S. A. 104, 13861–13867.

Press, W.H., Teukolsky, S. a, Vetterling, W.T., and Flannery, B.P. (1992). Numerical Recipes in C.

Preuss, T.M. (2000). Taking the measure of diversity: comparative alternatives to the model-animal paradigm in cortical neuroscience. Brain. Behav. Evol. 55, 287–299.

Preuss, T.M. (2010). Reinventing Primate Neuroscience for the Twenty-First Century. In Primate Neuroethology, M. Platt, and A. Ghazanfar, eds. (Oxford University Press), pp. 422–453.

Proulx, E., Piva, M., Tian, M.K., Bailey, C.D.C., and Lambe, E.K. (2014). Nicotinic acetylcholine receptors in atten-tion circuitry: the role of layer VI neurons of prefrontal cortex. Cell. Mol. Life Sci. 71, 1225–1244.

Radonjic, N. V., Ayoub, A.E., Memi, F., Yu, X., Maroof, A., Jakovcevski, I., Anderson, S.A., Rakic, P., and Zecevic, N. (2014). Diversity of Cortical Interneurons in Primates: The Role of the Dorsal Proliferative Niche. Cell Rep. 9, 2139–2151.

Rakic, P. (2009). Evolution of the neocortex: a perspective from developmental biology. Nat. Rev. Neurosci. 10, 724–735.

Rall, W. (1959). Branching dendritic trees and motoneuron membrane resistivity. Exp. Neurol. 1, 491–527.

Rall, W. (1964). Theoretical significance of dendritic trees for neuronal input-output relations. Neural Theory Model. 73–97.

Rall, W. (1967). Distinguishing Theoretical Synaptic Poten- Tials Computed for Different Soma-Den- Dritic Distribu-tions of Synaptic. J Neurophysiol.

Ravid, R., and Swaab, D.F. (1993). The Netherlands brain bank--a clinico-pathological link in aging and dementia research. J. Neural Transm. Suppl. 39, 143–153.

Reader, S.M., Hager, Y., and Laland, K.N. (2011). The evolution of primate general and cultural intelligence. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 366, 1017–1027.

Remy, S., and Spruston, N. (2007). Dendritic spikes induce single-burst long-term potentiation. Proc. Natl. Acad. Sci. U. S. A. 104, 17192–17197.

Rieke, F., Warland, D., De Ruyter van Steveninck, R., and Bialek, W. (1999). Spikes (Cambridge (Massachusetts): The MIT Press).

Rodríguez-Moreno, A., and Paulsen, O. (2008). Spike timing-dependent long-term depression requires presyn-aptic NMDA receptors. Nat. Neurosci. 11, 744–745.

Rodríguez-Moreno, A., Banerjee, A., and Paulsen, O. (2010). Presynaptic NMDA receptors and spike timing-dependent depression at cortical synapses. Front. Synaptic Neurosci.

Rosoklija, G.B., Petrushevski, V.M., Stankov, A., Dika, A., Jakovski, Z., Pavlovski, G., Davcheva, N., Lipkin, R., Schnieder, T., Scobie, K., et al. (2014). Reliable and durable Golgi staining of brain tissue from human autopsies and experimental animals. J. Neurosci. Methods 230, 20–29.

Rotaru DC, Olezene C, Miyamae T, Povysheva N V, Zaitsev A V, Lewis DA, Gonzalez-Burgos G (2015) Functional properties of GABA synaptic inputs onto GABA neurons in monkey prefrontal cortex. J Neurophysiol 113:1850–1861 Available at: http://www.ncbi.nlm.nih.gov/pubmed/25540225.

Roth, G., and Dicke, U. (2005). Evolution of the brain and intelligence. Trends Cogn. Sci. 9, 250–257.

Rotman, Z., Deng, P.-Y.P.-Y., and Klyachko, V. a. (2011). Short-Term Plasticity Optimizes Synaptic Information Trans-mission. J. Neurosci. 31, 14800–14809.

Rubin, J.E., Gerkin, R.C., Bi, G., Chow, C.C., and Jonathan, E. (2005). Calcium Time Course as a Signal for Spike-Timing – Dependent Plasticity. J. Neurophysiol. 2600–2613.

de Ruiter, J.P., and Uylings, H.B.M. (1987). Morphometric and dendritic analysis of fascia dentata granule cells in human aging and senile dementia. Brain Res. 402, 217–229.

Page 13: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

129

Bibliography

Ryan, T.J., and Grant, S.G.N. (2009). The origin and evolution of synapses. Nat. Rev. Neurosci. 10, 701–712.

Ryan, T.J., Roy, D.S., Pignatelli, M., Arons, A., and Tonegawa, S. (2015). Engram cells retain memory under retro-grade amnesia. Science 348, 1007–1013.

Sabatini, B.L., and Svoboda, K. (2000). Analysis of calcium channels in single spines using optical fluctuation analysis. Nature 408, 589–593.

Salinas, E., and Sejnowski, T.J. (2001). Correlated neuronal activity and the flow of neural information. Nat. Rev. Neurosci. 2, 539–550.

Samochocki, M., Höffle, A., Fehrenbacher, A., Jostock, R., Ludwig, J., Christner, C., Radina, M., Zerlin, M., Ullmer, C., Pereira, E.F.R., et al. (2003). Galantamine is an allosterically potentiating ligand of neuronal nicotinic but not of muscarinic acetylcholine receptors. J. Pharmacol. Exp. Ther. 305, 1024–1036.

Sarid, L., Bruno, R., Sakmann, B., Segev, I., and Feldmeyer, D. (2007). Modeling a layer 4-to-layer 2/3 module of a single column in rat neocortex: interweaving in vitro and in vivo experimental observations. Proc. Natl. Acad. Sci. U. S. A. 104, 16353–16358.

Sarter, M., Parikh, V., and Howe, W.M. (2009). nAChR agonist-induced cognition enhancement: integration of cognitive and neuronal mechanisms. Biochem. Pharmacol. 78, 658–667.

Sawaguchi, T., and Kudo, H. (1990). Neocortical development and social structure in primates. Primates 31, 283–289.

Sawaki, L., Yaseen, Z., Kopylev, L., and Cohen, L.G. (2003). Age-dependent changes in the ability to encode a novel elementary motor memory. Ann. Neurol. 53, 521–524.

Schiller, J., Major, G., Koester, H.J., and Schiller, Y. (2000). NMDA spikes in basal dendrites of cortical pyramidal neurons. Nature 404, 285–289.

Schwartzkroin, P. a, and Prince, D. a (1976). Microphysiology of human cerebral cortex studied in vitro. Brain Res. 115, 497–500.

Segev, I., and Rall, W. (1998). Excitable dendrites and spines: Earlier theoretical insights elucidate recent direct observations. Trends Neurosci. 21, 453–460.

Semon, R. (1921). The mneme.

Seol, G.H., Ziburkus, J., Huang, S., Song, L., Kim, I.T., Takamiya, K., Huganir, R.L., Lee, H.K., and Kirkwood, A. (2007). Neuromodulators Control the Polarity of Spike-Timing-Dependent Synaptic Plasticity. Neuron 55, 919–929.

Shannon, C.E. (1948). A mathematical theory of communication. Bell Syst. Tech. J. 27, 379–423.

Shepherd, G.M.G. (2003). The synaptic organisation of the brain (New York: Oxford University Press).

Shepherd, G.M.G., Stepanyants, A., Bureau, I., Chklovskii, D., and Svoboda, K. (2005). Geometric and functional organization of cortical circuits. Nat. Neurosci. 8, 782–790.

Shettleworth, S.J. (2012). Modularity, comparative cognition and human uniqueness. Philos. Trans. R. Soc. B Biol. Sci. 367, 2794–2802.

Shouval, H.Z., Bear, M.F., and Cooper, L.N. (2002). A unified model of NMDA receptor-dependent bidirectional synaptic plasticity. Proc. Natl. Acad. Sci. U. S. A. 99, 10831–10836.

Sihver, W., Gillberg, P.G., and Nordberg, a. (1998). Laminar distribution of nicotinic receptor subtypes in human cerebral cortex as determined by [3H](-)nicotine, [3H]cytisine and [3H]epibatidine in vitro autoradiography. Neuroscience 85, 1121–1133.

Silva, A.J., Zhou, Y., Rogerson, T., Shobe, J., and Balaji, J. (2009). Molecular and cellular approaches to memory allocation in neural circuits. Science 326, 391–395.

Sjostrom, P.J., Rancz, E. a D.E.A., Roth, A., Hausser, M., Ha, M., and Sjo, P.J. (2008). Dendritic Excitability and Synaptic Plasticity. Physiol. Rev. 88, 769–840.

Sjöström, P.J., and Häusser, M. (2006). A cooperative switch determines the sign of synaptic plasticity in distal dendrites of neocortical pyramidal neurons. Neuron 51, 227–238.

Sjöström, P.J., Turrigiano, G.G., and Nelson, S.B. (2003). Neocortical LTD via coincident activation of presynaptic NMDA and cannabinoid receptors. Neuron 39, 641–654.

Smiley, J.F., Morrell, F., and Mesulam, M.M. (1997). Cholinergic synapses in human cerebral cortex: an ultrastruc-tural study in serial sections. Exp. Neurol. 144, 361–368.

Page 14: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

130

Bibliography

Softky, W.R., and Koch, C. (1993). The highly irregular firing of cortical cells is inconsistent with temporal integra-tion of random EPSPs. J. Neurosci. 13, 334–350.

Song, S., Miller, K.D., and Abbott, L.F. (2000). Competitive Hebbian learning through spike-timing-dependent synaptic plasticity. Nat. Neurosci. 3, 919–926.

Spruston, N. (2008). Pyramidal neurons: dendritic structure and synaptic integration. Nat. Rev. Neurosci. 9, 206–221.

Stefan, K., Kunesch, E., Cohen, L.G., Benecke, R., and Classen, J. (2000). Induction of plasticity in the human motor cortex by paired associative stimulation. Brain 123 Pt 3, 572–584.

Stefan, K., Kunesch, E., Benecke, R., Cohen, L.G., and Classen, J. (2002). Mechanisms of enhancement of human motor cortex excitability induced by interventional paired associative stimulation. J. Physiol. 543, 699–708.

Swaab, D., and Uylings, H. (1988). Potentialities and pitfalls in the use of human brain material in molecular neuroanatomy. In Molecular Neuroanatomy, (Elsevier Science Publishers B.V.), pp. 403–416.

Tapia, L., Kuryatov, A., and Lindstrom, J. (2007). Ca2+ permeability of the (alpha4)3(beta2)2 stoichiometry greatly exceeds that of (alpha4)2(beta2)3 human acetylcholine receptors. Mol. Pharmacol. 71, 769–776.

Testa-Silva, G., Verhoog, M.B., Goriounova, N.A., Loebel, A., Hjorth, J.J.J., Baayen, J.C., de Kock, C.P.J., and Mans-velder, H.D. (2010). Human synapses show a wide temporal window for spike-timing-dependent plasticity. Front. Synaptic Neurosci.

Testa-Silva, G., Loebel, A., Giugliano, M., De Kock, C.P.J., Mansvelder, H.D., and Meredith, R.M. (2012). Hyper-connectivity and slow synapses during early development of medial prefrontal cortex in a mouse model for mental retardation and Autism. Cereb. Cortex 22, 1333–1342.

Testa-Silva, G., Verhoog, M.B., Linaro, D., de Kock, C.P.J., Baayen, J.C., Meredith, R.M., De Zeeuw, C.I., Giugliano, M., and Mansvelder, H.D. (2014). High Bandwidth Synaptic Communication and Frequency Tracking in Human Neocortex. PLoS Biol. 12, e1002007.

Thabit, M.N., Ueki, Y., Koganemaru, S., Fawi, G., Fukuyama, H., and Mima, T. (2010). Movement-related cortical stimulation can induce human motor plasticity. J. Neurosci. 30, 11529–11536.

Thirugnanasambandam, N., Grundey, J., Adam, K., Drees, A., Skwirba, A.C., Lang, N., Paulus, W., and Nitsche, M. a (2011). Nicotinergic impact on focal and non-focal neuroplasticity induced by non-invasive brain stimulation in non-smoking humans. Neuropsychopharmacology 36, 879–886.

Thorndike, R.L. (1953). Who belongs in the family? Psychometrika 18, 267–276.

Tian, M.K., Bailey, C.D.C., and Lambe, E.K. (2014). Cholinergic excitation in mouse primary vs. associative cortex: Region-specific magnitude and receptor balance. Eur. J. Neurosci. 40, 2608–2618.

Tsodyks, M. V., and Markram, H. (1997). The neural code between neocortical pyramidal neurons depends on neurotransmitter release probability. Proc. Natl. Acad. Sci. U. S. A. 94, 719–723.

Tsodyks, M., Pawelzik, K., and Markram, H. (1998). Neural networks with dynamic synapses. Neural Comput. 10, 821–835.

Tu, B., Gu, Z., Shen, J.-X., Lamb, P.W., and Yakel, J.L. (2009). Characterization of a nicotine-sensitive neuronal population in rat entorhinal cortex. J. Neurosci. 29, 10436–10448.

Tzounopoulos, T., Kim, Y., Oertel, D., and Trussell, L.O. (2004). Cell-specific, spike timing-dependent plasticities in the dorsal cochlear nucleus. Nat. Neurosci. 7, 719–725.

Urban, N.N., Henze, D.A., Lewis, D.A., and Barrionuevo, G. (1996). Properties of LTP induction in the CA3 region of the primate hippocampus. Learn. Mem. 3, 86–95.

Varki, A., Geschwind, D.H., and Eichler, E.E. (2008). Explaining human uniqueness: genome interactions with envi-ronment, behaviour and culture. Nat. Rev. Genet. 9, 749–763.

Vélez-Fort, M., and Margrie, T.W. (2012). Cortical circuits: layer 6 is a gain changer. Curr. Biol. 22, R411–R413.

Vélez-Fort, M., Rousseau, C. V., Niedworok, C.J., Wickersham, I.R., Rancz, E.A., Brown, A.P.Y., Strom, M., and Margrie, T.W. (2014). The Stimulus Selectivity and Connectivity of Layer Six Principal Cells Reveals Cortical Microcircuits Underlying Visual Processing. Neuron 83, 1431–1443.

Verhoog, M.B., and Mansvelder, H.D. (2011). Presynaptic ionotropic receptors controlling and modulating the rules for spike timing-dependent plasticity. Neural Plast. 2011.

Page 15: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

131

Bibliography

Verhoog, M.B., Goriounova, N. a, Obermayer, J., Stroeder, J., Hjorth, J.J.J., Testa-Silva, G., Baayen, J.C., de Kock, C.P.J., Meredith, R.M., and Mansvelder, H.D. (2013). Mechanisms underlying the rules for associative plasticity at adult human neocortical synapses. J. Neurosci. 33, 17197–17208.

Vetter, P., Roth, a, and Häusser, M. (2001). Propagation of action potentials in dendrites depends on dendritic morphology. J. Neurophysiol. 85, 926–937.

de Waal, F.B.M. (2007). Chimpanzee politics: Power and sex among apes (JHU Press).

Wang, Y., Markram, H., Goodman, P.H., Berger, T.K., Ma, J., and Goldman-Rakic, P.S. (2006). Heterogeneity in the pyramidal network of the medial prefrontal cortex. Nat. Neurosci. 9, 534–542.

Wankerl, K., Weise, D., Gentner, R., Rumpf, J., and Classen, J. (2010). L-type voltage-gated Ca2+ channels: a single molecular switch for long-term potentiation/long-term depression-like plasticity and activity-dependent metaplasticity in humans. J. Neurosci. 30, 6197–6204.

Williams, R.W., and Herrup, K. (1988). The control of neuron number. Annu. Rev. Neurosci. 11, 423–453.

Williams, S.R., and Stuart, G.J. (2002). Dependence of EPSP efficacy on synapse location in neocortical pyramidal neurons. Science 295, 1907–1910.

Wittenberg, G.M., and Wang, S.S.-H. (2006). Malleability of spike-timing-dependent plasticity at the CA3-CA1 synapse. J. Neurosci. 26, 6610–6617.

Wolters, A., Sandbrink, F., Schlottmann, A., Kunesch, E., Stefan, K., Cohen, L.G., Benecke, R., and Classen, J. (2003). A temporally asymmetric Hebbian rule governing plasticity in the human motor cortex. J. Neurophysiol. 89, 2339–2345.

Wolters, A., Schmidt, A., Schramm, A., Zeller, D., Naumann, M., Kunesch, E., Benecke, R., Reiners, K., and Classen, J. (2005). Timing-dependent plasticity in human primary somatosensory cortex. J. Physiol. 565, 1039–1052.

Wrangham, R.W. (2009). Catching Fire: How Cooking Made Us Human (Profile Books).

Yamashita, K., Hirose, S., Kunimatsu, A., Aoki, S., Chikazoe, J., Jimura, K., Masutani, Y., Abe, O., Ohtomo, K., Miyas-hita, Y., et al. (2009). Formation of long-term memory representation in human temporal cortex related to pictorial paired associates. J. Neurosci. 29, 10335–10340.

Yu, Y., Shu, Y., and McCormick, D. a (2008). Cortical action potential backpropagation explains spike threshold variability and rapid-onset kinetics. J. Neurosci. 28, 7260–7272.

Yuste, R., and Tank, D.W. (1996). Dendritic integration in mammalian neurons, a century after Cajal. Neuron 16, 701–716.

Zaborszky, L., Csordas, A., Mosca, K., Kim, J., Gielow, M.R., Vadasz, C., and Nadasdy, Z. (2013). Neurons in the Basal Forebrain Project to the Cortex in a Complex Topographic Organization that Reflects Corticocortical Connectivity Patterns: An Experimental Study Based on Retrograde Tracing and 3D Reconstruction. Cereb. Cortex bht210.

Zaitsev a. V., Povysheva N V., Gonzalez-Burgos G, Lewis D a. (2012) Electrophysiological classes of layer 2/3 pyramidal cells in monkey prefrontal cortex. J Neurophysiol 108:595–609.

Zeba, M., Jovanov-Milosević, N., and Petanjek, Z. (2008). Quantitative analysis of basal dendritic tree of layer III pyramidal neurons in different areas of adult human frontal cortex. Coll. Antropol. 32 Suppl 1, 161–169.

Zeng, H., Shen, E.H., Hohmann, J.G., Oh, S.W., Bernard, A., Royall, J.J., Glattfelder, K.J., Sunkin, S.M., Morris, J.A., Guillozet-Bongaarts, A.L., et al. (2012). Large-scale cellular-resolution gene profiling in human neocortex reveals species-specific molecular signatures. Cell 149, 483–496.

Zenke, F., Hennequin, G., and Gerstner, W. (2013). Synaptic Plasticity in Neural Networks Needs Homeostasis with a Fast Rate Detector. PLoS Comput. Biol. 9.

Zhang, J.-C., Lau, P.-M., and Bi, G.-Q. (2009). Gain in sensitivity and loss in temporal contrast of STDP by dopami-nergic modulation at hippocampal synapses. Proc. Natl. Acad. Sci. U. S. A. 106, 13028–13033.

Zhao, S., Ting, J.T., Atallah, H.E., Qiu, L., Tan, J., Gloss, B., Augustine, G.J., Deisseroth, K., Luo, M., Graybiel, A.M., et al. (2011). Cell type–specific channelrhodopsin-2 transgenic mice for optogenetic dissection of neural circuitry function. Nat. Methods 8, 745–752.

Zhou, Y.-D., Acker, C.D., Netoff, T.I., Sen, K., and White, J. a (2005). Increasing Ca2+ transients by broadening postsynaptic action potentials enhances timing-dependent synaptic depression. Proc. Natl. Acad. Sci. U. S. A. 102, 19121–19125.

Zucker, R.S., and Regehr, W.G. (2002). Short-term synaptic plasticity. Annu. Rev. Physiol. 64, 355–405.

Page 16: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

132

SummaryOur advanced cognitive abilities are generally considered as defining our humanity, yet what adaptations our brain has undergone to gain these faculties remains one of the central ques-tions in neuroscience. Traditionally, answers to this question have been sought in the anatomy of our brain, pointing at its large size and large number of neurons. Yet upon further exami-nation, our brain appears to have little exceptional or extra-ordinary features to which our outstanding cognitive abilities can be straightforwardly attributed. The question arises then, to what extent our advanced cognitive capacities may also be down to adaptations on a smaller scale, at the level of neurons and circuits. In this thesis, we explore the electrophysiological properties of human neurons and synapses in an effort to identify which features our neurons share with rodents, and in which they differ. To achieve this, we use whole-cell recordings from living human pyramidal neurons in slices of brain tissue resected during neurosurgery.

The first question addressed is how the dendritic arborisations of human neocortical pyra-midal neurons compare to those of other species and how they affect dendritic signal propa-gation. Reconstructing close to a hundred biocytin-filled human neurons provided the first quantitative dataset on full dendritic trees of human pyramidal neurons. Comparison of their dendritic morphologies to those of mice and macaques showed that human layer 2/3 pyra-midal neurons have almost three times as much dendrite and a distinct branching architec-ture. The effect of the distinct human dendritic arborisation on passive signal propagation was explored using computer modelling, which revealed strong attenuation of signals in human dendrites. This indicates that human neurons may rely heavily on local dendritic computation, but mechanisms may also exist to compensate this signal attenuation to some extent, such as a substantially lower specific membrane capacitance.

The extensive dendritic arborisations of human neurons support a vast amount of synapses, but little is known about their physiology; the properties of short-term plasticity and infor-mation transfer at human cortical synapses have never been studied directly. We therefore proceeded to characterise short-term plasticity at synapses between human layer 2/3 cortical pyramidal neurons. Human cortical synapses were found to all show short-term depression, similar to rodent synapses. In contrast to rodent synapses however, they recover much more rapidly from depression and are therefore capable of transferring much more information. Thus, when it comes to receiving information, human cortical pyramidal neurons not only receive an enormous volume of input, made possible by their large dendritic arborisations, but the inputs that they receive are also higher in information content, owing to fast recov-ering synapses. In order for human neurons to make use of all this information, they will have to be sensitive to the associated rapid fluctuations in synaptic inputs and be able to react quickly with their firing of action potentials. We tested how reliably human neurons can do so using an experimental paradigm called frequency tracking. Human neurons indeed turn out to be capable of encoding much more rapid fluctuations in inputs into the timing of their action potential than juvenile or adult mouse neurons, an ability supported by the faster onset kinetics of their action potentials.

Page 17: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

133

Summary

Information is stored in the brain by activity-dependent modifications in the strength of connections between neurons. An important form of such neuron-level information storage, called spike timing-dependent plasticity, has not been investigated directly at a cellular level in the human brain. In the second part of this thesis, we therefore set out to characterise the rules, mechanisms and modulation of spike timing-dependent plasticity at human cortical synapses. We find that human synapses can undergo bidirectional changes in strength throughout adulthood with a wider and reversed temporal window compared to that gener-ally found in juvenile rodents. Employing pharmacological and calcium imaging techniques, we found synaptic potentiation and depression at human synapses is gated by postsynaptic NMDA receptors and that dendritic L-type voltage-gated calcium channels recruited by back-propagating action potentials are important for synaptic strengthening.

Spike timing-dependent plasticity rules are not fixed, but plastic themselves and can be altered by the actions of neuromodulators such as acetylcholine. This thesis ends with an investiga-tion focussed on nicotinic acetylcholine receptors, which show a distinct layer and cell-type specific expression in the neocortex. We aimed to identify to what extent this layer-specific expression translates to layer-specific modulation of spike timing-dependent plasticity rules. Starting in mouse medial prefrontal cortex, we found that endogenous acetylcholine release augments long-term potentiation of glutamatergic synapses on layer 6 pyramidal neurons by activating dendritic nicotinic receptors which amplify back-propagating action potentials. This is in contrast to layer 5 where long-term potentiation was shown before to be supressed by nicotine and so points to layer-specific control over spike timing-dependent plasticity rules by the cholinergic system. Returning to the human neocortex, we found comparable mechanisms were in place there, with functional nicotinic receptors having a similar laminar distribution and cholinergic modulation of synaptic plasticity being opposite in superficial versus deep cortical layers.

The findings presented in this thesis show that many basic electrophysiological features are shared between human and rodent neurons, but importantly, we have also come to identify a number of morphological and physiological differences which may strongly impact the way these cells process and storage information. When considering the origins of the cognitive capacities of the human brain, one therefore cannot merely point to its size and numbers of neurons, but must also take into account an array of neuron-level adaptations that may have large implications for the computational power of the system as a whole.

Page 18: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

134

Nederlandse samenvattingOnze intelligentie wordt doorgaans gezien als hetgeen wat ons mensen onderscheidt van andere dieren en wat ons mens-zijn definieert. Waar onze cognitieve vermogens vandaan komen blijft echter een van de grote vragen in de neurowetenschap. Vaak worden zij toege-schreven aan ons grote brein en de enorme hoeveelheid hersencellen, maar eigenlijk zijn er weinig anatomische aspecten van ons brein te noemen waarin wij werkelijk zo uniek zijn in vergelijking met andere dieren. De vraag is dus of onze intelligentie wellicht niet alleen uit een verschil in kwantiteit komt, maar ook een verschil in kwaliteit. Uiteindelijk worden al onze hersenfuncties ondersteund door hersencellen die met elektrische signaaltjes infor-matie uitwisselen in complex georganiseerde circuits. Zouden er op dit niveau wellicht eigen-schappen aan te wijzen zijn die onze intelligentie kunnen verklaren?

Op dit moment is er maar weinig bekend over de elektrische eigenschappen van de hersen-cellen in ons brein. De reden hiervoor is dat om deze te onderzoeken men levend hersenma-teriaal nodig heeft en dat is niet vaak voor handen. Gevolg is dat bijna al onze kennis over hoe hersencellen informatie verwerken en opslaan komt uit studies naar de hersenen van proefdieren, doorgaans ratten en muizen van een paar weken oud. Willen wij ons brein ooit helemaal begrijpen dan zullen we moeten toetsen in hoeverre de kennis opgedaan over het knaagdierenbrein van toepassing is op het onze. Ons laboratorium is een van de weinigen ter wereld met de kans dit te doen; zo nu en dan wordt er bij de neurochirurgie afdeling van het VU Medisch Centrum een stukje hersenweefsel weggesneden bij patiënten en dit weefsel kan, nadat het in dunne plakjes is gesneden, zo’n 12 uur in leven worden gehouden. Zo kunnen wij metingen doen aan levende menselijke hersencellen en hun elektrische eigenschappen eindelijk in detail bestuderen. Dit proefschrift presenteert het resultaat van 6 jaar onderzoek naar menselijke hersencellen dat met deze methode werd mogelijk gemaakt.

We beginnen met een heel basale vraag; hoe zien menselijke hersencellen er uit? Hersen-cellen hebben vanuit hun cellichaam grote uitlopers, de z.g. dendrieten, die functioneren als de ontvangstantennes van de cel. Hier komt informatie binnen van andere cellen via gespe-cialiseerde structuren die synapsen worden genoemd, die optreden waar de ene cel contact maakt met de ander. De dendrieten geven met hun vele vertakkingen de hersencel het kara-kteristieke uiterlijk van een ontwortelde boom, zoals ook op de voorkant van dit proefschrift is te zien. Verassend genoeg is er na meer dan 100 jaar neurowetenschap nog steeds geen duidelijk beeld van hoe deze er nu volledig uitzien. Het visualiseren van hersencellen in post-mortem materiaal gebeurt doorgaans in microscopisch dunne plakjes, veel dunner dan de spanwijdte van een hersencel, en geeft dus zelden een beeld van de complete dendritische ‘boom’. Ons preparaat laat cellen veel meer intact en zo hebben wij bijna honderd digitale 3D reconstructies weten te maken van haast complete menselijke hersencellen. Deze lieten zien dat menselijke hersencellen ongeveer drie keer zo groot zijn als die van een muis en zelfs andere apen soorten. Met hun veel grotere dendritische boom kunnen menselijke hersen-cellen dus veel meer informatie ontvangen dan de hersencellen van een muis.

Page 19: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

135

Nederlandse Samenvatting

Vervolgens hebben we bestudeerd hoe menselijke hersencellen met elkaar communiceren. Door twee cellen tegelijk te meten met een electrode, kon onderzocht worden hoe efficiënt informatie van de ene naar de andere cel via de synaps kon worden overgedragen. In veel aspecten bleken mensen- en muizensynapsen vergelijkbare fysiologische eigenschappen te hebben, maar er was ook een belangrijk verschil. Synapsen raken bij herhaalde activiteit steeds verder uitgeput, maar het bleek dat menselijke synapsen zeer snel bijkomen hiervan. Met dit vergrootte uithoudingsvermogen kan een menselijke synaps in een korte tijd veel meer infor-matie doorgeven dan die van een muis.

De ‘informatie’ die cellen uitwisselen neemt de vorm aan van een klein stroom signaaltje. Het cellichaam is de plek waar al deze inkomende signaaltjes bij elkaar worden opgeteld en waar een beslissing gemaakt wordt of er iets moet worden doorgegeven aan andere cellen. Dit gebeurt wanneer de stroompjes bij elkaar optellen tot een bepaalde drempelwaarde, waarna een actiepotentiaal optreedt. Dit is een soort kettingreactie waarbij er een grote stroomstoot wordt gegenereerd die langs het axon naar andere cellen stroomafwaarts schiet. Het hebben van zo veel synapsen, die allen veel informatie over kunnen dragen, stelt het cellichaam behoorlijk op de proef. We hebben daarom gekeken naar hoe snel een menselijke hersencel kan reageren op zo’n snelle stroom van informatie. De actiepotentialen van menselijke hersen-cellen bleken veel sneller uit de startblokken schieten, waardoor zij op veel fijnere fluctuaties in elektrische signalen kunnen reageren. Menselijke hersencellen zijn dus uitgerust met actie-potentialen die snel genoeg zijn om met de grote hoeveelheid informatie die zij binnen krijgen om te gaan.

In het tweede gedeelte van dit proefschrift zijn wij ingegaan op de cellulaire mechanismen die ten grondslag liggen aan ons geheugenvermogen. Herinneringen kunnen in ons brein worden opgeslagen door relatief kleine groepjes hersencellen. Een karakteristieke eigenschap van zo’n cel populatie die voor een herinnering zorgt is dat hun onderlinge verbindingen sterker zijn. Het vormen van een versterkte onderlinge verbinding vindt plaats door de synapsen tussen deze cellen sterker te maken. Dit fenomeen heet synaptische plasticiteit en treedt op wanneer twee verbonden cellen herhaaldelijk samen actief zijn; de synaps tussen de twee cellen merkt dit en ondergaat dan een verandering in sterkte.

Synaptische plasticiteit speelt waarschijnlijk een cruciale rol in ons leer- en geheugenver-mogen, maar desondanks is dit proces tot op heden nog nauwelijks direct op celniveau in het menselijk brein onderzocht. Wij hebben daarom de activiteitspatronen die synaptische plasticiteit opwekken nagebootst in menselijke hersenplakjes, om zo de precieze voorwaarden voor synaptische plasticiteit en de onderliggende moleculaire mechanismen in detail uit te zoeken. Wij vonden dat synaptische plasticiteit in het menselijk brein in een aantal opzichten vergelijkbaar is met dat in knaagdieren. Met name de onderliggende mechanismen, zoals de betrokkenheid van bepaalde receptoren en ion kanalen (de NMDA receptoren en L-type calcium kanalen) waren identiek. Toch was de relatie tussen welke verandering in synaptische sterkte optrad bij welk activiteitspatroon erg verschillend en in sommige opzichten zelfs compleet tegenovergesteld. Wat dit precies betekent is vooralsnog onduidelijk en zal in de toekomst met computer modellen moeten worden uitgezocht. Interessant is wel dat daar waar synaptische plasticiteit in veel dieren lijkt af te nemen bij volwassenwording, de synapsen in ons brein ons hele volwassen leven door sterker en zwakker gemaakt kunnen worden. Dit suggereert, wellicht geruststellend, dat je nooit te oud bent om iets te leren.

Page 20: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

136

Nederlandse Samenvatting

Hoe het wordt bepaald waar herinneringen worden opgeslagen, dus welke cellen sterker of zwakker met elkaar worden verbonden, is onderwerp van veel neurowetenschappelijk onder-zoek. Neuromodulatoren zoals bijvoorbeeld dopamine, serotonine en acetylcholine kunnen hier wellicht een rol in spelen. Wanneer deze stoffen in ons brein worden afgegeven kunnen zij het gedrag van hersencellen heel specifiek beïnvloeden en bepalen of synapsen worden versterkt of niet. Wij waren geïnteresseerd in hoe specifiek acetylcholine synaptische plastic-iteit in de menselijke hersenschors kan beïnvloeden. Acetylcholine activeert de nicotinerge acetylcholine receptor; deze heet zo omdat de nicotine in tabaksrook toevallig ook deze receptor activeert. Uit muizen onderzoek is bekend dat de nicotine receptor een heel specifiek verspreidingspatroon heeft over de verschillende cellagen van de hersenschors. In het laatste onderdeel van dit proefschrift laten wij zien dat dit verspreidingspatroon tot op zekere hoogte ook in mensen aanwezig is. Activatie van deze nicotine receptoren blijkt een cellaag-specifiek effect te hebben op synaptische plasticiteit; in de bovenste cellagen remmen zij de verster-king van synapsen, terwijl zij het in de diepste cellagen juist bevorderen. Acetylcholine heeft dus een cellaag-specifieke controle over welke synapsen wanneer versterken en kan zo het proces van geheugenopslag en het aanpassen van circuits in verschillende cel populaties van ons brein dirigeren.

Dit proefschrift laat zien dat er tussen mens en muis veel overeenkomsten, maar ook belan-grijke verschillen bestaan in de elektrische eigenschappen van hun hersencellen. Deze verschillen hebben een directe invloed op hoe informatie door de hersencellen van ons brein wordt verwerkt en opgeslagen. Om deze verschillen werkelijk op waarde te kunnen schatten is meer onderzoek nodig, met name naar de hersencellen van dieren die evolutionair, qua hersengrootte of qua intelligentie dichter bij de mens staan. Toch kunnen we nu al stellen dat er in de zoektocht naar de oorsprong van menselijke intelligentie niet alleen naar groot-schalige anatomische aspecten van ons brein gekeken moet worden; er bestaat namelijk een heel palet aan kleine adaptaties op celniveau die grote consequenties kunnen hebben voor de rekencapaciteit van ons brein in zijn geheel.

Page 21: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

137

Page 22: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

138

Dankwoord Dit proefschrift is het resultaat van inmiddels bijna zes jaar werken aan de VU. In deze tijd heb ik het voorrecht gehad om met veel bijzondere mensen samen te werken. Graag wil ik hier een aantal in het bijzonder bedanken voor hun rol in de totstandkoming van dit proefschrift.

Beste Huib, bedankt dat ik zo lang deel uit heb mogen maken van je lab. Ik heb er over de jaren echt een geweldige tijd gehad en veel geleerd. Ik weet nog goed hoe ik je voor het eerst ontmoette toen ik als bachelor student je kantoor instapte om een interview te doen met ‘een echte wetenschapper’. Ik was natuurlijk zenuwachtig, maar je ontspannen houding sloeg al snel op mij over. Die ‘no worries’ mentaliteit heb ik later ook altijd erg op prijs gesteld als dingen vast leken te zitten. Verder dank ik je ook voor alle kansen die ik heb gekregen om aan leuke projecten mee te doen. Met name het werken aan humaan hersenweefsel heeft mij vele spannende dagen en nachten vol adrenaline gegeven waar ik erg van heb genoten. In bredere zin ben ik misschien nog het meest dankbaar voor de vrijheid die ik kreeg in het onderzoek en het vertrouwen om mij mijn gang te laten gaan, ondanks dat het waarschijnlijk niet altijd even duidelijk geweest zal zijn waar ik nu precies mee bezig was.

Beste Hans, ik weet werkelijk niet wat ik zonder jouw hulp, ervaring en encyclopedische kennis over de praktijk van de electrofysiologie had moeten doen. De zes maanden dat ik in Oxford stage liep zonder jou waren erg moeilijk, met experimenten die maar niet wilden lukken. Het was treffend dat jij na drie zinnen al wist wat het probleem was en hoe het opgelost had kunnen worden (cesium-gluconaat i.p.v. cesium-chloride! Aargh…). Jouw aanwezigheid was dan ook een sterke reden om voor het INF te kiezen als promotieplek en ik ben blij dat ik dat heb gedaan; ik heb zo ontzettend veel van je geleerd de afgelopen jaren. Van je verhalen over de oude methoden heb ik altijd gesmuld, het was leuk om die te horen van iemand die er vanaf het eerste uur bij is geweest. Ik vind het daarom ook bijzonder passend dat mijn laatste experiment in het lab er een was met jou en een van die oude technieken, de nucleated patch. Hans, doe mij een plezier en zet je brein op papier; “A practical guide for the electrophysiolo-gist: techniques, tricks and troubleshooting” lijkt mij wel een aardige titel. Of begin anders ’s werelds eerste electrofysiologie museum met al je oude meetapparatuur, hand-gedreven slicers en prehistorische pipet-trekkers! Ik lig voor de deur te slapen.

Dear members of the INF, past and present; over the many years I’ve spent in the lab it has been a pleasure to meet and work with such an interesting and colourful array of people. All of you have helped shape my PhD years into what have certainly been some of the most fulfilling years of my life. Not once have I felt reluctant to go to work, or have I felt bored while on the job. I put this down to the unique INF feeling you all help create. Yes, other departments may sometimes jokingly refer to us as the autists when we don’t show up for drinks, but within our world I believe we are having an equally rich social experience as any other. The many endless behind-the-patch-rig discussions about experiments, the state of present-day science, global politics or nothing at all, the shameless nerd-talk over beers and raw cauliflower, and of course the ‘human nights’ with their one-of-a-kind atmosphere... These things will stay with me forever and I will miss them always. It is now, thinking back of all those years, that I come to realise how much precious memories have been made in that one corridor of the Earth and Life Sciences Faculty. It saddens me to think I’m leaving this behind, but I go with the feeling of having been part of something special. Thank you all.

Page 23: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

139

Ten slotte wil ik nog één speciaal persoon bedanken, mijn lieve vriendin Siri. Lieve Sier, ik ben zo dankbaar dat ik jou heb. Jouw liefde, passie en levenslust herinneren mij er altijd aan hoe mooi het leven is en hoeveel er is te ontdekken in de wereld buiten het lab. De onvoorwaardelijke steun die je me geeft houdt mij altijd op de been en zeker in de laatste jaren van mijn PhD, waarin ik besef dat ik een groot beroep heb gedaan op je begrip, is die onmisbaar geweest. Ik had het echt niet zonder jou kunnen doen. De tien jaar die wij nu samen zijn hebben mij als mens gevormd en zijn de meest waardevolle jaren van mijn leven, ik kan dan ook niet wachten op wat onze toekomst zal brengen! Ik hou van je.

Page 24: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

140

Page 25: Bibliography - research.vu.nl matter.pdf · Comparative Study of Human and Mouse Postsynaptic Proteomes Finds High Compositional Conservation and Abundance Differences for Key Synaptic

141

List of publicationsTesta-Silva, G., Verhoog, M.B., Goriounova, N.A., Loebel, A., Hjorth, J.J.J., Baayen, J.C., de Kock, C.P.J., and Mansvelder, H.D. (2010). Human synapses show a wide temporal window for spike-timing-dependent plasticity. Frontiers in Synaptic Neuroscience.

Verhoog, M.B., and Mansvelder, H.D. (2011). Presynaptic ionotropic receptors controlling and modulating the rules for spike timing-dependent plasticity. Neural Plasticity. 2011.

Poorthuis, R.B., Bloem, B., Verhoog, M.B., and Mansvelder, H.D. (2013). Layer-specific interfer-ence with cholinergic signalling in the prefrontal cortex by smoking concentrations of nicotine. The Journal of Neuroscience. 33, 4843–4853.

Verhoog, M.B.*, Goriounova, N.A.*, Obermayer, J., Stroeder, J., Hjorth, J.J.J., Testa-Silva, G., Baayen, J.C., de Kock, C.P.J., Meredith, R.M., and Mansvelder, H.D. (2013). Mechanisms under-lying the rules for associative plasticity at adult human neocortical synapses. The Journal of Neuroscience. 33, 17197–17208.

Testa-Silva, G., Verhoog, M.B., Linaro, D., de Kock, C.P.J., Baayen, J.C., Meredith, R.M., De Zeeuw, C.I., Giugliano, M., and Mansvelder, H.D. (2014). High bandwidth synaptic communica-tion and frequency tracking in human neocortex. PLoS Biology. 12, e1002007.

Mohan, H.*, Verhoog, M.B.*, Doreswamy, K.K., Eyal, G., Aardse, R., Lodder, B.N., Goriounova, N.A., Asamoah, B., Brakspear, A.B.C., Groot, C., van der Sluis, S., Testa-Silva, G., Obermayer, J., Boudewijns, Z.S.R.M., Narayanan, R.T., Baayen, J.C., Segev, I., Mansvelder, H.D. and de Kock, C.P.J. (2015). Dendritic and axonal architecture of individual pyramidal neurons across layers of adult human neocortex. Cerebral Cortex 25, 4839–4853.

Verhoog, M.B., Obermayer, J*, Kortleven, C.A.*, Wilbers, R., Wester, J., Baayen, J.C., De Kock, C.P.J., Meredith, R.M., and Mansvelder, H.D. Layer-specific cholinergic control of human and mouse cortical synaptic plasticity by pre- and postsynaptic nicotinic acetylcholine receptors. Under revision.

Eyal, G., Verhoog, M.B., Testa-Silva, G., Deitcher, Y., Lodder, J., Benavides-Piccione, R., Morales, J., DeFelipe, J., de Kock, C.P.J., Mansvelder, H.D., and Segev, I. Human L2/3 pyramidal cells possess unique membrane and dendritic properties with augmented computational capabili-ties. Manuscript in preparation.

* equal contribution