Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-23T19:27:15.238Z Has data issue: false hasContentIssue false

An evolving view of retinogeniculate transmission

Published online by Cambridge University Press:  29 August 2017

ELIZABETH Y. LITVINA
Affiliation:
Department of Neurology, F.M. Kirby Neurobiology Center, Children’s Hospital, Boston, Boston, Massachusetts 02115 Program in Neuroscience, Harvard Medical School, Boston, Massachusetts 02115
CHINFEI CHEN*
Affiliation:
Department of Neurology, F.M. Kirby Neurobiology Center, Children’s Hospital, Boston, Boston, Massachusetts 02115 Program in Neuroscience, Harvard Medical School, Boston, Massachusetts 02115
*
*Address correspondence to: Chinfei Chen, F.M. Kirby Neurobiology Center Boston Children’s Hospital, 300 Longwood Avenue, CLS 12250 Boston, MA 02115. E-mail: chinfei.chen@childrens.harvard.edu
Rights & Permissions [Opens in a new window]

Abstract

The thalamocortical (TC) relay neuron of the dorsoLateral Geniculate Nucleus (dLGN) has borne its imprecise label for many decades in spite of strong evidence that its role in visual processing transcends the implied simplicity of the term “relay”. The retinogeniculate synapse is the site of communication between a retinal ganglion cell and a TC neuron of the dLGN. Activation of retinal fibers in the optic tract causes reliable, rapid, and robust postsynaptic potentials that drive postsynaptics spikes in a TC neuron. Cortical and subcortical modulatory systems have been known for decades to regulate retinogeniculate transmission. The dynamic properties that the retinogeniculate synapse itself exhibits during and after developmental refinement further enrich the role of the dLGN in the transmission of the retinal signal. Here we consider the structural and functional substrates for retinogeniculate synaptic transmission and plasticity, and reflect on how the complexity of the retinogeniculate synapse imparts a novel dynamic and influential capacity to subcortical processing of visual information.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2017 

More than a relay

The retinogeniculate synapse has been an invaluable workhorse for neuroscience research. In recent years, it has served as a powerful model for understanding the molecular and circuit-level mechanisms that influence normal development and disease (Bishop et al., Reference Bishop, Burke and Davis1959; Sengpiel & Kind, Reference Sengpiel and Kind2002; Kastner et al., Reference Kastner, Schneider and Wunderlich2006; Guido, Reference Guido2008; Kano & Hashimoto, Reference Kano and Hashimoto2009; Hong & Chen, Reference Hong and Chen2011; Stephan et al., Reference Stephan, Barres and Stevens2012; Kaplan, Reference Kaplan2014). The ability to independently label Retinal Ganglion Cells (RGCs) from opposite eyes has enabled the identification of cellular and molecular mechanisms of axon mapping, arbor pruning, and synapse elimination that drive the refinement of retinotopic maps and eye-specific lamination (Wong, Reference Wong1999; Luo & O’Leary, Reference Luo and O’Leary2005; Huberman et al., Reference Huberman, Feller and Chapman2008a ; Feller, Reference Feller2009; Kano & Hashimoto, Reference Kano and Hashimoto2009; Hong & Chen, Reference Hong and Chen2011). In vitro, the easily accessible bundle of RGC axons in the optic nerve provides a convenient means of selectively activating presynaptic inputs in studies of synaptic physiology and plasticity, and is also an excellent system for uncovering mechanisms of axon regeneration (Guido, Reference Guido2008; Hong & Chen, Reference Hong and Chen2011; Benowitz et al., Reference Benowitz, He and Goldberg2017). In vivo, the retinogeniculate synapse has been a prominent model system for the study of subcortical visual processing. Numerous studies have capitalized on the ease of manipulating visual stimulation, combined with the ability to simultaneously monitor the activity patterns of inputs and outputs of the thalamus to reliably demonstrate that the transfer of information from the retina to the visual cortex is the major function of the retinogeniculate synapse (Sherman, Reference Sherman2005; Usrey & Alitto, Reference Usrey and Alitto2015; Weyand, Reference Weyand2016).

Despite the tenacity of the term “relay” to describe the function of the dLGN, the retinogeniculate synapse does not simply transfer a copy of RGC activity patterns to the cortex. Over the past few decades, a number of retinogeniculate attributes have been shown to play a role in modifying visual information before conveying it to the cortex (Blitz et al., Reference Blitz, Foster and Regehr2004; Sherman, Reference Sherman2007; Usrey & Alitto, Reference Usrey and Alitto2015; Weyand, Reference Weyand2016). Simultaneous recording of the firing patterns of RGC inputs and TC target neurons in vivo has shown that the reliability of action potential generation in TC neurons depends on the local activity context, such as neuromodulatory signaling or the membrane potential. These factors can dictate or modulate the firing pattern (“tonic” or “burst”) of a TC neuron (McCormick & Bal, Reference McCormick and Bal1994; Usrey et al., Reference Usrey, Reppas and Reid1998; Sherman & Guillery, Reference Sherman and Guillery2002; Wang et al., Reference Wang, Wei, Vaingankar, Wang, Koepsell, Sommer and Hirsch2007). In some cases, a TC neuron is much more likely to generate a response to the second of a pair of visually-driven potentials arriving in quick succession (separated by less than 30 ms, Mastronarde, Reference Mastronarde1987; von Krosigk et al., Reference von Krosigk, Bal and McCormick1993; Usrey et al., Reference Usrey, Reppas and Reid1998; Levine & Cleland, Reference Levine and Cleland2001; Rowe & Fischer, Reference Rowe and Fischer2001; Carandini et al., Reference Carandini, Horton and Sincich2007; Weyand, Reference Weyand2007; Sincich et al., Reference Sincich, Adams, Economides and Horton2007, Reference Sincich, Horton and Sharpee2009; Alitto et al., Reference Alitto, Moore, Rathbun and Martin Usrey2011). This integrative function increases the amount of information encoded in thalamic spiking (Wang et al., Reference Wang, Hirsch and Sommer2010). In other cases, a single RGC impulse can result in a postsynaptic burst of multiple action potentials (Usrey et al., Reference Usrey, Reppas and Reid1998; Blitz & Regehr, Reference Blitz and Regehr2003).

Recent work is revealing new receptive field complexity and plasticity in the dLGN that further demonstrates significant thalamic processing of visual information en route to the cortex, at least in some species. Stimulus orientation selectivity is one salient example of a complex feature encoded in subpopulations of dLGN neurons in a variety of species: mouse (Marshel et al., Reference Marshel, Kaye, Nauhaus and Callaway2012; Piscopo et al., Reference Piscopo, El-Danaf, Huberman and Niell2013; Scholl et al., Reference Scholl, Tan, Corey and Priebe2013; Zhao et al., Reference Zhao, Chen, Liu and Cang2013) and rabbit (Levick et al., Reference Levick, Oyster and Takahashi2010; Hei et al., Reference Hei, Stoelzel, Zhuang, Bereshpolova, Huff, Alonso and Swadlow2014), with weaker orientation or direction bias occurring in the cat (Hubel & Wiesel, Reference Hubel and Wiesel1961; Daniels et al., Reference Daniels, Norman and Pettigrew1977; Levick & Thibos, Reference Levick and Thibos1980; Vidyasagar & Urbas, Reference Vidyasagar and Urbas1982; Soodak et al., Reference Soodak, Shapley and Kaplan1987; Shou & Leventhal, Reference Shou and Leventhal1989; Thompson et al., Reference Thompson, Zhou and Leventhal1994), squirrel (Zaltsman et al., Reference Zaltsman, Heimel and Van Hooser2015), and primate (Lee et al., Reference Lee, Creutzfeldt and Elepfandt1979; Smith et al., Reference Smith, Chino, Ridder, Kitagawa and Langston1990; Cheong et al., Reference Cheong, Tailby, Solomon and Martin2013). Complex feature selectivity persists in TC neurons after inactivation of the primary visual cortex, suggesting that the dLGN may compute orientation or direction selectivity rather than inherit it from cortical feedback (cat Vidyasagar & Urbas, Reference Vidyasagar and Urbas1982; mouse Zhao et al., Reference Zhao, Chen, Liu and Cang2013; Scholl et al., Reference Scholl, Tan, Corey and Priebe2013). Furthermore, cat TC neurons have a higher stimulus contrast sensitivity than their individual inputs, suggesting TC neurons can functionally integrate information from multiple RGC inputs (Rathbun et al., Reference Rathbun, Alitto, Warland and Usrey2016). The presence of binocularly innervated dLGN neurons in mice, cats, and primates further supports the possibility of convergence of multiple RGCs onto single geniculate neurons in mice, cat, and primate dLGN (Sanderson et al., Reference Sanderson, Bishop and Darian-Smith1971; Howarth et al., Reference Howarth, Walmsley and Brown2014; Zeater et al., Reference Zeater, Cheong, Solomon, Dreher and Martin2015; Rompani et al., Reference Rompani, Mullner, Wanner, Roth, Yonehara, Zhang and Roska2017). Here, we explore synaptic mechanisms and structural attributes that could support the diversity of features and functions now emerging in studies of the dLGN. In this review, we describe the structural properties of RGC axons and TC neurons, together with their biophysical features and plasticity mechanisms. These properties may combine to impart novel and adaptive functionality to the dLGN, and to dynamically regulate information flow between retina and cortex.

Retinogeniculate synaptic structure

To appreciate the functional complexity in retinogeniculate processing, we first describe the underlying synaptic structure. The architecture of the retinogeniculate synapse is conserved across species. Glutamate is the excitatory neurotransmitter packed into numerous round vesicles contained in large synaptic terminals along the axon (Montero & Wenthold, Reference Montero and Wenthold1989). A single retinal axon terminal can span ∼1–4 microns in diameter and contain multiple spatially distinct neurotransmitter release sites (cat and mouse studies: Famiglietti & Peters, Reference Famiglietti and Peters1972; Rafols & Valverde, Reference Rafols and Valverde1973; Sur & Sherman, Reference Sur and Sherman1982; Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987; Robson, Reference Robson1993; Bickford et al., Reference Bickford, Slusarczyk, Dilger, Krahe, Kucuk and Guido2010; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012; Morgan et al., Reference Morgan, Berger, Wetzel and Lichtman2016). RGC boutons contact a TC neuron near its cell body, synapsing directly onto the dendritic shaft or dendritic appendages that protrude from the proximal shaft or primary dendritic branch points (Rafols & Valverde, Reference Rafols and Valverde1973; Robson & Mason, Reference Robson and Mason1979; Wilson et al., Reference Wilson, Friedlander and Sherman1984; Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987; Bickford et al., Reference Bickford, Slusarczyk, Dilger, Krahe, Kucuk and Guido2010; Morgan et al., Reference Morgan, Berger, Wetzel and Lichtman2016). Retinal inputs account for only 5–10% of a TC neuron’s synaptic input, whereas cortical feedback projections from Layer 6 occupy the distal dendrites, providing as much as 50% of synaptic input (Wilson et al., Reference Wilson, Friedlander and Sherman1984; Montero, Reference Montero1991; Van Horn et al., Reference Van Horn, Erişir and Sherman2000). Nonetheless, the proximal position of retinogeniculate synapses along the dendrite, their synaptic structure with multiple release sites, and their large number of synaptic contacts drives powerful and reliable transmission that has earned the retinal input the moniker of “driver” to all other inputs’ “modulator” (Guillery & Sherman, Reference Guillery and Sherman2002).

The fine details of retinogeniculate connectivity reveal potential heterogeneity of circuit organization and, thereby, of function. A RGC axon contacts a TC neuron with as many as 59 terminals in cat or mouse (Fig. 1A–1C; Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987; Robson, Reference Robson1993; Morgan et al., Reference Morgan, Berger, Wetzel and Lichtman2016). The morphology of these numerous synaptic contacts between a single RGC axon and its target TC neuron can range from the simple to the complex (Jones & Powell, Reference Jones and Powell1969; Famiglietti & Peters, Reference Famiglietti and Peters1972; Hammer et al., Reference Hammer, Monavarfeshani, Lemon, Su and Fox2015; for finer categorization, see Lund & Cunningham, Reference Lund and Cunningham1972; Robson & Mason, Reference Robson and Mason1979; Morgan et al., Reference Morgan, Berger, Wetzel and Lichtman2016). Simple retinogeniculate contacts consist of one small or large crenulated bouton contacting a dendrite or dendritic appendage. Complex contacts comprise a glomerular structure containing multiple boutons from retinal, inhibitory, and neuromodulatory inputs (Robson & Mason, Reference Robson and Mason1979; Koch, Reference Koch1985; Sherman & Guillery, Reference Sherman and Guillery1996; Sherman, Reference Sherman2004). Although it is not known whether differences in bouton morphology correlate with specializations for retinogeniculate information transfer, work on cerebellar parallel fibers shows that bouton size can predict sensitivity to neuromodulation, and induction of LTP or LTD can induce plasticity in the size of hippocampal presynaptic boutons (Toni et al., Reference Toni, Buchs, Nikonenko, Bron and Muller1999; Becker et al., Reference Becker, Wierenga, Fonseca, Bonhoeffer and Nägerl2008; Zhang & Linden, Reference Zhang and Linden2009). Furthermore, the morphology of synaptic contacts is optimized to the requirements of sensory transmission in the retina, inner ear, and central auditory synapses (Taschenberger et al., Reference Taschenberger, Leão, Rowland, Spirou and von Gersdorff2002; Matthews & Fuchs, Reference Matthews and Fuchs2010; Freche et al., Reference Freche, Pannasch, Rouach and Holcman2011; Graydon et al., Reference Graydon, Cho, Diamond, Kachar, von Gersdorff and Grimes2014). It is therefore likely that the diversity of retinogeniculate contact morphologies reflects differences in their contribution to transmission. Indeed, different TC neuron types in cats exhibit biases in presynaptic morphology. Cat RGCs and TC neurons are distinguished into three categories (X, Y, W) based on the properties of their responses to visual stimuli, including the size of the receptive field and the degree of linearity of spatial summation, as well as morphological markers. The X and Y classifications refer to relatively homogeneous populations (and are often compared to primate M and P pathways), whereas the W (compared to K in primates) encompasses a more diverse set of cells with rarely-encountered physiological responses (Wilson et al., Reference Wilson, Rowe and Stone1976; Fukuda et al., Reference Fukuda, Hsiao, Watanabe and Ito1984; Felch & Van Hooser, Reference Felch and Van Hooser2012). Simple contacts dominate retinal input onto cat Y cells, whereas X cell dendritic appendages preferentially participate in complex synaptic structures (Robson & Mason, Reference Robson and Mason1979; Hamos et al., Reference Hamos, Van Horn, Raczkowski, Uhlrich and Sherman1985; Koch, Reference Koch1985; Sherman & Guillery, Reference Sherman and Guillery1996; Sherman, Reference Sherman2004). Similar distinctions among TC neurons have been observed in the mice by morphological analysis, but have not been associated with distinct patterns of synaptic structures nor delineated by physiology (Krahe et al., Reference Krahe, El-Danaf, Dilger, Henderson and Guido2011; El-Danaf et al., Reference El-Danaf, Krahe, Dilger, Bickford, Fox and Guido2015; Sriram et al., Reference Sriram, Meier and Reinagel2016).

Fig. 1. Synaptic structure shapes retinogeniculate transmission. (A) Tracing of an HRP-filled X-RGC arbor in the cat dLGN shows the location and morphology of a single branch (red box) of the X-RGC arbor used for EM reconstruction. This branch of the axon contacts 4 TC neurons out of 40 available neurons in the territory of the arbor. The remainder of the axon was not reconstructed, and likely contacts several other TC neurons. Bottom inset shows the location of the axonal arbor in the context of the cat LGN. Figure modified from Hamos et al. (Reference Hamos, Van Horn, Raczkowski and Sherman1987). Unmarked scale bar = 100 μm. (B, C) Reconstructed arbors of single RGC axons showing distribution of presynaptic boutons into dense clusters in the LGN of (B) an adult cat and (C) a p20 mouse. Note the clustering of boutons along the arbor. Image in B is modified from Robson et al. (Reference Robson1993), showing a segment of a RGC axon; Image in C is from Hong et al. (Reference Hong, Park, Litvina, Morales, Sanes and Chen2014), showing a BD-RGC axon. Scales bars are 100 μm. (D) A 3D reconstruction of a TC neuron dendrite and sites of contact between two neighboring RGC boutons from Budisantoso et al. (Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). In the top image, the dendrite and its appendages are depicted in blue, whereas pink and red sites label the postsynaptic densities of the two axons. In the bottom image, the structure of the terminals of two axons has been added. Spillover can occur between these two nearby terminals. (E) Evidence of spillover-mediated responses to the stimulation of a single RGC axon before eye opening. Two different synaptic responses were observed in response to single retinal fiber stimulation. Shown are recordings from TC neurons in whole cell voltage clamp at −70 mV in a dLGN slice in the presence of the NMDAR blocker, 20 µM CPP. On the left is an example of a retinogeniculate AMPAR EPSC with characteristic rapid rise time and decay kinetics (black trace). On the right is an atypical AMPAR EPSC response notable for significantly slower rise time and decay kinetics (black trace). The two types of EPSCs differ in their sensitivity to the low-affinity AMPAR antagonist, γ-DGG. Low affinity antagonists can be used to assess the relative concentration of glutamate in the synaptic cleft (Clements et al., Reference Clements, Lester, Tong, Jahr and Westbrook1992; Diamond & Jahr, Reference Diamond and Jahr1997). As γ-DGG competes with glutamate for binding to AMPAR, its efficacy of inhibition decreases with increasing glutamate concentration. γ-DGG has only a small effect on the amplitude of the fast EPSC, but dramatically reduces the amplitude of the slow EPSC (overlaid gray traces), consistent with lower peak glutamate concentration in the synaptic cleft of the slow EPSC. Because the EPSCs are evoked by minimal stimulation, the rapid EPSC represents a direct input from a single RGC axon that forms a direct synapse onto the voltage-clamped relay neuron, whereas the slow EPSC corresponds to the activation of a RGC axon that does not directly synapse onto the voltage-clamped neuron. Modified from Hauser et al. (Reference Hauser, Liu, Litvina and Chen2014). All figures reprinted with permission.

In addition to potential functional differences reflected in the morphology of presynaptic geniculate boutons, the close proximity of clustered multisynaptic boutons in the mature synapse makes possible novel interactions between glutamate transients originating in separate RGC inputs. Large simple boutons in the rat contain an average of 27 independent release sites, whereas each of the boutons in a glomerulus has approximately 6 (Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012; Hammer et al., Reference Hammer, Monavarfeshani, Lemon, Su and Fox2015). Notably, multiple RGCs may contribute to the same cluster or glomerulus of boutons (Hammer et al., Reference Hammer, Monavarfeshani, Lemon, Su and Fox2015; Morgan et al., Reference Morgan, Berger, Wetzel and Lichtman2016). At other CNS synapses, glia ensheath individual boutons and interdigitate into the synaptic cleft, preventing glutamate from diffusing, or “spilling over,” to neighboring boutons, in part through the action of glutamate transporters, which clear glutamate from the extracellular space (Diamond & Jahr, Reference Diamond and Jahr1997; Danbolt, Reference Danbolt2001; Tzingounis & Wadiche, Reference Tzingounis and Wadiche2007; Hauser et al., Reference Hauser, Edson, Hooks and Chen2013; Rimmele & Rosenberg, Reference Rimmele and Rosenberg2016). In contrast, glia do not interdigitate into the cleft of single RGC boutons nor within geniculate glomeruli, making the retinogeniculate connection conducive to glutamate spillover within and between individual boutons (Famiglietti & Peters, Reference Famiglietti and Peters1972; Robson & Mason, Reference Robson and Mason1979; Winfield et al., Reference Winfield, Hiorns and Powell1980; Mason, Reference Mason1982; Bickford et al., Reference Bickford, Slusarczyk, Dilger, Krahe, Kucuk and Guido2010). Although interbouton spillover has not been experimentally assessed in the mature dLGN, a simulation that demonstrated the likelihood of spillover between distant synapses within the same bouton also implies the possibility of spillover of glutamate between closely spaced boutons. Fig. 1D shows an example of two such close retinogeniculate boutons along one TC neuron dendrite, highlighting that glutamate released from one bouton can diffuse to postsynaptic release sites of the neighboring bouton (Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). At a developmental phase when boutons are less clustered (Sur et al., Reference Sur, Weller and Sherman1984; Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014) and glomeruli have not yet formed, retinogeniculate transmission exhibits extensive glutamate spillover between neighboring boutons. In fact, glutamate from the bouton of one RGC axon can spill over to the synaptic cleft of a neighboring RGC axon before eye opening (Hauser et al., Reference Hauser, Liu, Litvina and Chen2014). Glutamate spillover can be distinguished from direct retinogeniculate synaptic activation by the slower kinetics of the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR)-mediated Excitatory Postsynaptic Current (EPSC) and increased sensitivity to γ-DGG (gray traces), a low affinity AMPAR antagonist (Fig. 1E). These characteristics indicate that the receptors mediating the spillover current are exposed to a lower glutamate concentration than those mediating the direct EPSC. Glutamate spillover can therefore diminish synaptic specificity, but also result in complex and graded integration of information transmitted from different RGC inputs. This mechanism of transmission may be particularly relevant for high-frequency presynaptic activity (RGCs can reach up to 500 Hz, Nirenberg & Meister, Reference Nirenberg and Meister1997), which would promote the pooling and spillover of glutamate. Therefore, the intricate morphology of retinogeniculate contacts presents the possibility for a diversity of modes of retinogeniculate information transfer, which could vary as a function of preceding sensory or modulatory activity.

Short-term plasticity

The mature retinogeniculate synapse boasts multiple bouton contacts, each with many release sites, leading to a high probability of release (Yeow & Peterson, Reference Yeow and Peterson1991; Chen & Regehr, Reference Chen and Regehr2000; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). Combined with the proximity of retinal contacts to the cell body that minimizes potential dendritic filtering of the synaptic signal, these structural features give rise to an EPSC characterized by rapid kinetics (time constant of decay ∼2 ms) and large amplitudes in vitro (Chen & Regehr, Reference Chen and Regehr2000; example recording in Fig. 2A, p26–32). Studies in slices have identified several pre and postsynaptic mechanisms of short-term plasticity that modulate these EPSCs to further shape transmission based on the activity of the RGC input itself as well as other retinal and nonretinal inputs to that neuron. In shaping retinogeniculate transmission, these short-term plasticity mechanisms provide the means to dynamically modify the information transmitted from RGC to TC neuron output.

Fig. 2. Contributions of retinogeniculate short-term plasticity. (A) Representative traces of AMPAR and NMDAR mediated currents recorded before eye opening (left) and in a mature mouse (right) in response to the stimulation of the optic tract. Whole-cell voltage clamp recordings were performed with bicuculline to block GABAA-receptor mediated currents. At −70 mV holding potential, AMPARs mediate the fast activating and decaying current. AMPAR and NMDAR currents both contribute to the EPSCs recorded at +40 mV with AMPARs contributing to the rapid rise and the NMDAR currents contributing to the slow decay of the EPSC. The average amplitude of AMPAR currents increases over development. (B) 5-CT-mediated activation of serotonin receptors alters retinogeniculate short-term plasticity. Experiments were performed in retinogeniculate slices from mature mice. Top and bottom traces overlay pairs of retinogeniculate EPSCs evoked with varying ISI before (top) and after (bottom) the application of 5-CT to active 5HT-1 receptors expressed in presynaptic retinogeniculate boutons. Application of 5-CT reduces the amplitude of the first EPSC and relieves short-term depression, increasing the amplitude of the second EPSC preferentially at short interstimulus interval. (C) Physiologically relevant stimulation frequencies preferentially diminish the contribution of AMPARs to relay neuron firing. Current clamp recordings of action potential firing in response to trains of optic tract stimulation in the presence of AMPAR (NBQX) or NMDAR (CPP) antagonists. Holding potential −50 mV. Blockade of AMPARs alters the latency to first spike but only minimally reduces the overall number of spikes. In contrast, blockade of NMDARs abolished EPSC summation toward action potential firing; only the first stimulus evokes an action potential, reflecting the contribution of AMPARs that rapidly desensitize after the first pulse. Therefore, NMDAR currents can sustain action potential generation without AMPAR contribution. Adapted from (A) Chen and Regehr (Reference Chen and Regehr2000), B) Liu and Chen (Reference Liu and Chen2008) and (C) Augustinaite and Heggelund (Reference Augustinaite and Heggelund2007). All figures reprinted with permission.

A prominent feature of retinogeniculate transmission studied in vitro is short-term depression: the second of two impulses separated by a short interval generates a weaker response than the first. Both presynaptic mechanisms involving vesicle depletion and postsynaptic mechanisms including AMPAR desensitization and N-methyl-D-aspartate receptor (NMDAR) saturation contribute to this plasticity (Chen et al., Reference Chen, Blitz and Regehr2002; Blitz et al., Reference Blitz, Foster and Regehr2004; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). In contrast to the short-term depression observed in vitro in mice, extracellular recordings of TC neuron activity in vivo in cats and primates consistently show paired-stimulus enhancement, such that the second of two retinal impulses separated by a brief (<30 ms) interval is more effective at driving a postsynaptic action potential (Mastronarde, Reference Mastronarde1987; Usrey et al., Reference Usrey, Reppas and Reid1998; Levine & Cleland, Reference Levine and Cleland2001; Rowe & Fischer, Reference Rowe and Fischer2001; Carandini et al., Reference Carandini, Horton and Sincich2007; Rathbun et al., Reference Rathbun, Alitto and Weyand2007; Sincich et al., Reference Sincich, Adams, Economides and Horton2007, Reference Sincich, Horton and Sharpee2009). In part, this contradiction is due to the dependence of short-term depression on the recent history of activity at the synapse. Activation of the retinogeniculate synapse in slice usually follows a period of quiescence, whereas baseline spontaneous activity maintains synaptic transmission in a chronically depressed state in vivo (Levick & Williams, Reference Levick and Williams1964; Stoelzel et al., Reference Stoelzel, Huff, Bereshpolova, Alonso, Swadlow, Zhuang, Hei, Alonso and Swadlow2015). A train of spikes preceding optic tract stimulation in vitro attenuates the degree of synaptic depression (Seeburg et al., 2004; Augustinaite & Heggelund, Reference Augustinaite and Heggelund2007; Liu & Chen, Reference Liu and Chen2008). Further, neurotransmitter inputs from the brainstem and inhibitory neurons can modulate retinogeniculate transmission in a context-dependent manner in vivo. To understand the relationship between in vitro and in vivo manifestations of short-term plasticity, we review the prominent sources of modulation of retinogeniculate transmission.

Presynaptic modulation

Retinogeniculate transmission is known to be modulated presynaptically by a number of neurotransmitter receptors, including GABAB, serotonin 5HT1B (Chen & Regehr, Reference Chen and Regehr2003; Seeburg et al., Reference Seeburg, Liu and Chen2004), as well as adenosine A1 (Yang et al., Reference Yang, Hu, Huang and Chou2014) and metabotropic glutamate receptors (Hauser et al., Reference Hauser, Edson, Hooks and Chen2013; Lam & Sherman, Reference Lam and Sherman2013). Activation of the GABAB or 5-HT1B receptors strongly depresses neurotransmitter release and relieves short-term depression by decreasing the entry of calcium into the presynaptic terminal (see Fig. 2B; Chen & Regehr, Reference Chen and Regehr2003; Seeburg et al., Reference Seeburg, Liu and Chen2004). While these modulators decrease the strength of the retinogeniculate EPSC, they also alter the pattern of action potentials transmitted from the pre-to post-synaptic neuron. For example, activation of presynaptic 5-HT1B receptors leads to preferential transmission of high-frequency over lowfrequency activity, essentially acting as a high-pass filter (Seeburg et al., Reference Seeburg, Liu and Chen2004). In some cases, presynaptic modulation can be additive. The combined activation of 5HT1B and adenosine A1 receptors, can convert presynaptic depression into facilitation (Yang et al., Reference Yang, Hu, Huang and Chou2014). Therefore, the activity of neuromodulatory inputs in vivo can dynamically shape retinogeniculate information transfer by modulating the degree of short-term plasticity. It is not known whether the expression of presynaptic receptors differs between RGC types or RGC bouton morphologies. However, any such differences would add an additional layer of modularity to retinogeniculate transmission.

Postsynaptic modulation

In addition to presynaptic depression, postsynaptic glutamate receptor properties also contribute to short-term depression and shape the efficacy of retinogeniculate transmission. Postsynaptic AMPA and NMDA receptors both exhibit short-term depression, and perform complementary functions in retinogeniculate transmission.

AMPAR channel gating properties and the high density of their expression contribute to the large, rapid activation, and decay kinetics of the retinogeniculate EPSC (Tarusawa et al., Reference Tarusawa, Matsui, Budisantoso, Molnár, Watanabe, Matsui, Fukazawa and Shigemoto2009). Because AMPARs readily conduct at negative potentials, they are effective at initiating postsynaptic spiking, even from a relatively hyperpolarized membrane potential (Blitz & Regehr, Reference Blitz and Regehr2003; Augustinaite & Heggelund, Reference Augustinaite and Heggelund2007; Liu & Chen, Reference Liu and Chen2008). However, AMPARs desensitize upon exposure to glutamate and recover with a time constant of ∼100 ms (Chen et al., Reference Chen, Blitz and Regehr2002; Kielland & Heggelund, Reference Kielland and Heggelund2002). These properties lead to short-term depression of the AMPAR EPSC. Therefore, AMPARs contribute to the onset of an action potential train transmitted from a RGC, initiating robust short-latency spikes during low frequency activity, but cannot sustain the robust transmission of high-frequency retinogeniculate activity (Blitz & Regehr, Reference Blitz and Regehr2003). Fig. 2C demonstrates that pharmacological blockade of AMPAR with NBQX reduces the initial spikes in response to a stimulus train (Turner et al., Reference Turner, Leresche, Guyon, Soltesz and Crunelli1994; Augustinaite & Heggelund, Reference Augustinaite and Heggelund2007).

NMDA receptor-mediated currents have distinct kinetics and voltage dependence from AMPAR (Fig. 3A and 3B). At the retinogeniculate synapse, NMDARs exhibit short-term depression due to their high affinity to glutamate and receptor saturation (Traynelis et al., Reference Traynelis, Wollmuth, McBain, Menniti, Vance, Ogden, Hansen, Yuan, Myers and Dingledine2010). Recovery from NMDAR saturation occurs more quickly than from AMPAR desensitization (Chen et al., Reference Chen, Blitz and Regehr2002; Kielland & Heggelund, Reference Kielland and Heggelund2002). TC neuron NMDARs experience incomplete Mg2+ block at hyperpolarized potentials, and therefore conduct significant current at negative potentials (Liu & Chen, Reference Liu and Chen2008). Fig. 3B and 3C illustrates the contribution of NMDAR to transmission over development. NMDARs conduct current for many tens of milliseconds during a prolonged decay, which permits the summation of closely timed EPSCs, especially within the range of interstimulus intervals (ISIs) that exhibit paired-pulse enhancement in vivo, and supports multiple TC neuron spikes even in “tonic” mode (Chen et al., Reference Chen, Blitz and Regehr2002; Blitz & Regehr, Reference Blitz and Regehr2003; Augustinaite & Heggelund, Reference Augustinaite and Heggelund2007; Liu & Chen, Reference Liu and Chen2008; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). Blockade of NMDAR dramatically reduces retinogeniculate transmission in vivo (Sillito et al., Reference Sillito, Murphy, Salt and Moody1990; Kwon et al., Reference Kwon, Esguerra and Sur1991), and NMDAR current summation in vitro can even drive action potential firing in the presence of AMPAR blockers, though with less temporal precision (Fig. 2C; Chen et al., Reference Chen, Blitz and Regehr2002; Blitz & Regehr, Reference Blitz and Regehr2003; Augustinaite & Heggelund, Reference Augustinaite and Heggelund2007; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). In fact, the NMDAR component of the first EPSC may sufficiently depolarize a TC neuron to spike threshold, such that a small or depressed AMPAR current can shorten the latency to first spike (Kielland & Heggelund, Reference Kielland and Heggelund2002; Augustinaite & Heggelund, Reference Augustinaite and Heggelund2007; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). Therefore, summation of NMDAR currents enhances the probability of TC neuron spiking in response to the second and later RGC action potentials during a train. Together with modulation and postsynaptic integration (Carandini et al., Reference Carandini, Horton and Sincich2007), the properties of NMDA and AMPA receptor currents can reconcile the robust short-term depression seen in vitro with paired-pulse enhancement observed in vivo.

Fig. 3. Contribution of NMDAR-currents to retinogeniculate transmission over development. (A) NMDAR EPSCs recorded in the presence of the AMPAR blocker, NBQX, at +40 and −55 mV holding potentials in a p10 (left) and a p29 (right) retinogeniculate slice. Normalized traces are shown. Note the acceleration in NMDAR current decay time over development. (B) Example EPSCs recorded in young (top) and mature (bottom) TC neuron in slice before (left) and during (right) the application of NBQX. Holding potential, −55 mV. (C) NMDAR currents contribute more to the total retinogeniculate charge transfer at p9–11 than p26–32; however, even at the mature synapse, NMDARs contribute nearly half of the total charge transfer. Figure adapted from Liu and Chen (Reference Liu and Chen2008). All figures reprinted with permission.

An additional factor that could contribute to differences in in vitro and in vivo short-term plasticity is the expression of calcium-permeable AMPARs. The retinogeniculate synapse differs from other synapses in that the expression of calcium-permeable AMPARs increases over development (see Fig. 4A and 4B; Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012; Hauser et al., Reference Hauser, Liu, Litvina and Chen2014; Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014; compare to Kumar et al., Reference Kumar, Bacci, Kharazia and Huguenard2002; Soto et al., Reference Soto, Coombs, Kelly, Farrant and Cull-Candy2007). Calcium permeable AMPARs (those lacking the GluA2 subunit) exhibit stronger desensitization-mediated paired pulse depression (Budisantoso et al., Reference Budisantoso, Matsui, Kamasawa, Fukazawa and Shigemoto2012). However, depolarization-mediated reduction of polyamine block may partly rescue this effect (Rozov et al., Reference Rozov, Zilberter, Wollmuth and Burnashev1998; Soto et al., Reference Soto, Coombs, Kelly, Farrant and Cull-Candy2007), increasing the contribution from AMPARs to transmission later in a train of high frequency activity.

Fig. 4. Substrates for retinogeniculate plasticity. (A) Overlaid AMPAR current traces recorded from different holding potentials to assess the current voltage (IV) relationship. Currents in the presence of CPP to block NMDAR currents and with spermine in the internal solution to examine the degree of IV rectification. Calcium-permeable AMPARs exhibit a rectifying IV relationship. Traces were recorded at 20 mV increments from −60 to +60 mV holding potentials. Left-example obtained before eye opening; right, example from a mature slice. From Hauser et al. (Reference Hauser, Liu, Litvina and Chen2014). (B) Change in the average AMPAR EPSC IV relationship over development. Rectification of IV currents increases significantly from p9–11 to maturity, indicating a gradual increase in the contribution of CP-AMPARs to AMPAR-mediated currents. Modified from Hauser et al. (Reference Hauser, Liu, Litvina and Chen2014). Red: p9–11; blue: p15–16; black: p27–32. (C) Changes in AMPAR subunit composition in response to visual experience. The effect of visual deprivation from p20 (late-dark rear, LDR) or dark rearing from birth (chronic dark reared, CDR) on the AMPAR EPSC IV relationship. Rectification of AMPAR currents is reduced in LDR but not in chronically dark reared (CDR) mice when compared to normally reared mice (light rear, LR) mice. P = 0.03. Recordings performed at p27–32. Modified from Louros et al. (Reference Louros, Hooks, Litvina, Carvalho and Chen2014). (D) Comparison of the distribution of amplitudes of single fiber RGC inputs in juvenile (p27–34) and adult (p60+) mice show the persistence of weak (small-amplitude) inputs with age. Modified from Hong et al. (Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). All figures reprinted with permission.

The in vitro retinogeniculate preparation has permitted the identification of mechanisms regulating retinogeniculate synaptic transmission at an unmatched resolution. The synaptic mechanisms discussed above are part of a larger array of factors that affect retinogeniculate information transfer, including circuit elements that influence postsynaptic integration in the TC neuron: local and extrageniculate GABAergic circuits, reciprocal connectivity with the cortex, and brainstem modulatory inputs (reviewed in Sherman & Guillery, Reference Sherman and Guillery2002). Together, these synaptic and circuit mechanisms impart the dynamic features that regulate transmission of information at the retinogeniculate synapse. Interestingly, the paired-stimulus enhancement of retinogeniculate transmission in vivo acts not only to increase the signal-to-noise of retino-geniculo-cortical information transfer, but also encode emergent features in the spike code (Sincich et al., Reference Sincich, Horton and Sharpee2009; Rathbun et al., Reference Rathbun, Warland and Usrey2010; Wang et al., Reference Wang, Hirsch and Sommer2010). Thus these synaptic mechanisms may also impart novel functionality at the level of dLGN. Further, as these mechanisms can rapidly alter the contribution of a particular retinal input to postsynaptic spiking, they may regulate the contribution of strong versus weak RGC inputs to visual processing (discussed below).

Neurotransmission before eye-opening

The distinct features of the immature retinogeniculate synapse suggest that the immature dLGN carries out a different set of computations on incoming retinal information than the mature dLGN. Many of the modulatory circuits that shape transmission in the adult dLGN begin to innervate TC neurons shortly before eye-opening: corticogeniculate innervation is not complete in mice until p14 (Jacobs et al., Reference Jacobs, Campagnoni, Kampf, Reyes, Kalra, Handley, Xie, Hong-Hu, Spreur, Fisher and Campagnoni2007; Seabrook et al., Reference Seabrook, Eleftheriou, Krahe, Fox and Guido2013; Grant et al., Reference Grant, Hoerder-Suabedissen and Molnar2016), and cholinergic innervation develops over several postnatal weeks in cats (Carden et al., Reference Carden, Datskovskaia, Guido, Godwin and Bickford2000). GABAergic interneurons continue to be recruited into the dLGN at the end of the first postnatal week in mice, and GABAergic innervation in rodents and carnivores occurs gradually (Shatz and Kirkwood, Reference Shatz and Kirkwood1984; Ramoa & McCormick, Reference Ramoa and McCormick1994a ; Pirchio et al., Reference Pirchio, Turner, Williams, Asprodini and Crunelli1997; Ziburkus et al., Reference Ziburkus, Lo and Guido2003; Golding et al., Reference Golding, Pouchelon, Bellone, Murthy, Di Nardo, Govindan, Ogawa, Shimogori, Lüscher, Dayer and Jabaudon2014). In addition, presynaptic ultrastructural morphology and TC dendritic arbor complexity are immature at eye-opening (Bickford et al., Reference Bickford, Slusarczyk, Dilger, Krahe, Kucuk and Guido2010). Therefore, retinogeniculate transmission before eye-opening occurs in a very different environment than after circuits have matured.

Whereas mature TC neurons receive and integrate information from one or several strong retinal inputs that can reach several nA in amplitude, numerous weak inputs, measuring on average ∼40 pA in amplitude (peak AMPAR EPSC) innervate a TC neuron before eye opening (in vitro in mice; Hooks and Chen, Reference Hooks and Chen2006). Remarkably, retinogeniculate transmission to cortex does occur before input refinement: both the spontaneous patterns of activity (retinal waves) that prominently feature in the developing retina, and visually-evoked stimuli detectable through the closed eyelid influence the activity of the visual cortex (Katz & Shatz, Reference Katz and Shatz1996; Mooney et al., Reference Mooney, Penn, Gallego and Shatz1996; Feller, Reference Feller1999; Akerman et al., Reference Akerman, Smyth and Thompson2002; Hanganu et al., Reference Hanganu, Ben-Ari and Khazipov2006; Ackman et al., Reference Ackman, Burbridge and Crair2012). In slice, the synaptic charge transfer needed to drive TC neuron spiking before eye-opening is relatively small: an AMPAR EPSC with a peak amplitude of 120 pA is adequate (Liu & Chen, Reference Liu and Chen2008). The coincident activation of a subset of the dozen or more converging RGC inputs, perhaps relying on synchronous activity that dominates retinal activity during this period of development, can achieve this amplitude (Wong et al., Reference Wong, Meister and Shatz1993; Wong, Reference Wong1999; Butts & Rokhsar, Reference Butts and Rokhsar2001; Feller, Reference Feller2009).

Multiple mechanisms contribute to the efficacy of neurotransmission at the immature weak retinogeniculate synapse. A study using the slow calcium chelator, EGTA-AM, suggested that the distance between the presynaptic release machinery and calcium channels at retinal terminals is greater at immature presynaptic specializations, resulting in delayed or asynchronous release (Borst & Sakmann, Reference Borst and Sakmann1996; Hauser et al., Reference Hauser, Liu, Litvina and Chen2014). Additionally, the immature nervous system produces a slower action potential waveform (Taschenberger & von Gersdorff, Reference Taschenberger and von Gersdorff2000; Murphy & du Lac, Reference Murphy and du Lac2001) and a lower density of glutamate transporters in surrounding astrocytes (Thomas et al., Reference Thomas, Tian and Diamond2011). These properties of the immature synapse lead to the prolonged exposure of postsynaptic receptors to glutamate, promoting the integration of retinal EPSCs over a longer time scale than after maturation.

Several postsynaptic mechanisms also improve the integration of weak RGC inputs at young ages. The temporal window for postsynaptic summation is much greater before eye opening (Liu & Chen, Reference Liu and Chen2008). This is due in part to the higher input resistance in immature neurons (Ramoa & McCormick, Reference Ramoa and McCormick1994b ; Macleod et al., Reference Macleod, Turner and Edgar1997; Pirchio et al., Reference Pirchio, Turner, Williams, Asprodini and Crunelli1997), as well as enhanced conduction through NMDA receptors at negative potentials because of a greater contribution of both NR2B as well as NR2C/D subunits at young synapses (Ramoa & McCormick, Reference Ramoa and McCormick1994a ; Liu & Chen, Reference Liu and Chen2008). These receptors exhibit slower decay kinetics and a lower sensitivity to magnesium block (compare between ages in Fig. 3), and their contribution declines over development in an activity-regulated manner (Ramoa & Prusky, Reference Ramoa and Prusky1997; Chen & Regehr, Reference Chen and Regehr2000; Liu & Chen, Reference Liu and Chen2008).

AMPAR current amplitudes at the immature retinogeniculate synapse are much smaller than later in life (Figs. 2A and 3B). In fact, a substantial fraction (22%) of immature RGC inputs are “silent” (lacking detectable AMPAR currents; Isaac et al., Reference Isaac, Nicoll and Malenka1995; Liao et al., Reference Liao, Hessler and Malinow1995; Chen & Regehr, Reference Chen and Regehr2000). Transmission before eye opening therefore seems to rely almost entirely on NMDAR transmission, with AMPARs influencing the latency to spike (Liu & Chen, Reference Liu and Chen2008). Finally, immature neurons also exhibit calcium plateau potentials (Jaubert-Miazza et al., Reference Jaubert-Miazza, Green, Lo, Bui, Mills and Guido2005; Lo et al., Reference Lo, Ziburkus and Guido2013), and more depolarized resting membrane potentials that are closer to firing threshold (Ramoa & McCormick, Reference Ramoa and McCormick1994b ; Macleod et al., Reference Macleod, Turner and Edgar1997; Pirchio et al., Reference Pirchio, Turner, Williams, Asprodini and Crunelli1997), increasing the efficacy of individual inputs.

In summary, the developing retinogeniculate synapse exhibits numerous adaptations that permit it to integrate and transfer visual signals to cortex even while it undergoes dramatic synaptic rearrangement. As these signals arise from the summation of multiple weak convergent RGC inputs, the computations that the immature dLGN performs, and therefore its role in visual processing, is substantially different from that of the mature dLGN. In addition to a role in conveying visual information, retinogeniculate transmission is important for cortical map formation (Huberman et al., Reference Huberman, Manu, Koch, Susman, Lutz, Ullian, Baccus and Barres2008b ; Cang & Feldheim, Reference Cang and Feldheim2013; Owens et al., Reference Owens, Feldheim, Stryker and Triplett2015), though its precise computational role is not yet understood.

Retinogeniculate connectivity

Developmental refinement of retinogeniculate connectivity

Retinogeniculate refinement is thought to lead to the maturation of receptive field properties in the dLGN. Before eye opening, RGC axons from the two eyes segregate into eye-specific layers: segments of the axon arbor that occupy the inappropriate layer are pruned, while the appropriately positioned portion of the arbor becomes more elaborate in a number of species (Robson, Reference Robson1981; Mason, Reference Mason1982; Sretavan & Shatz, Reference Sretavan and Shatz1984, Reference Sretavan and Shatz1986; Campbell & Shatz, Reference Campbell and Shatz1992; Garraghty & Sur, Reference Garraghty and Sur1993; Dhande et al., Reference Dhande, Hua, Guh, Yeh, Bhatt, Zhang, Ruthazer, Feller and Crair2011; Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). In vitro studies of the dLGN in rodents show that each TC neuron receives weak inputs from more than a dozen RGCs. Some of these inputs are subsequently pruned while others strengthen to become dominant drivers of postsynaptic activity (reviewed in Guido, Reference Guido2008; Huberman et al., Reference Huberman, Feller and Chapman2008a ; Hong & Chen, Reference Hong and Chen2011; Thompson et al., Reference Thompson, Gribizis, Chen and Crair2017). This refinement occurs over several weeks following eye opening in mice. The earliest phases depend on spontaneous input from the retina, while visual experience maintains the mature configuration and modifies connectivity via feedback from the cortex during a critical period (Hooks & Chen, Reference Hooks and Chen2006; Thompson et al., Reference Thompson, Picard, Min, Fagiolini and Chen2016).

The robust refinement of retinogeniculate connectivity demonstrated in vitro in mice corresponds temporally to the developmental transformation of initially broad, irregularly-shaped or temporally imprecise receptive fields to smaller, sharper, or temporally precise ones that closely match the receptive field of the dominant retinal inputs in vivo (cat: Wiesel & Hubel, Reference Wiesel and Hubel1963; Daniels et al., Reference Daniels, Pettigrew and Norman1978; Tootle & Friedlander, Reference Tootle and Friedlander1989; Gary-Bobo et al., Reference Gary-Bobo, Przybyslawski and Saillour1995; Cai et al., Reference Cai, DeAngelis and Freeman1997; Ferret: Tavazoie & Reid, Reference Tavazoie and Reid2000, Akerman et al., Reference Akerman, Smyth and Thompson2002, Davis et al., Reference Davis, Chapman and Cheng2015; primate: Blakemore & Vital-durand, Reference Blakemore and Vital-durand1985; mouse in vitro: Chen & Regehr, Reference Chen and Regehr2000; Jaubert-Miazza et al., Reference Jaubert-Miazza, Green, Lo, Bui, Mills and Guido2005). Surprisingly, there is a disconnect between structure and function—studies in cats, mice, and primates fail to show large-scale pruning of the axon arbor during this later window of development (Sur et al., Reference Sur, Weller and Sherman1984, Reference Sur, Esguerra, Garraghty, Kritzer and Sherman1987; Lachica & Casagrande, Reference Lachica and Casagrande1988; Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). A recent study that examined individually reconstructed axon arbors of a subtype of mouse RGCs, the BD-RGC (ON–OFF direction-selective RGC, Kim et al., Reference Kim, Zhang, Meister and Sanes2010), found that their size and branching complexity remain stable in the 2–3 weeks following eye opening. Instead, during the period of robust functional refinement, changes occur in bouton size and distribution along the arbor structure (Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). Before eye-opening, boutons are distributed broadly along the terminal arbor in mice, but gradually form tight clusters over the later window of development (for examples of mouse and cat RGC axon bouton clustering, see Fig. 1B and 1C). This development suggests that an immature axon makes transient contacts with a large number of potential postsynaptic targets, but redistributes its inputs onto a few targets during the period of activity-dependent refinement. Final pruning of the arbor skeleton, however, does not occur until well after the end of the geniculate and cortical critical periods (Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). These findings are consistent with observations in cat and primate studies of the complexity of mature RGC axon morphology (Fig. 1A), with multiple segments that can branch off the primary axon within the optic tract (Sur & Sherman, Reference Sur and Sherman1982; Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987; Sur et al., Reference Sur, Esguerra, Garraghty, Kritzer and Sherman1987; Garraghty et al., Reference Garraghty, Shatz, Sretavan and Sur1988; Dhande et al., Reference Dhande, Hua, Guh, Yeh, Bhatt, Zhang, Ruthazer, Feller and Crair2011; Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014), although the arbors may be more restricted in the primate dLGN (Glees & Le Gros Clark, Reference Glees and Le Gros Clark1941; Lachica & Casagrande, Reference Lachica and Casagrande1988; Michael, Reference Michael1988; Conley & Fitzpatrick, Reference Conley and Fitzpatrick1989). In cat, one RGC axon arbor spans the territory of far more postsynaptic neurons than it contacts (Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987). Reduction of the X-RGC axon arbor occurs between 4 and 12 weeks postnatal, after the peak of ocular dominance plasticity (Hubel & Wiesel, Reference Hubel and Wiesel1970; Sur et al., Reference Sur, Weller and Sherman1984).

Therefore, while retinogeniculate development yields a circuit with appreciable functional specificity, the anatomical correlates of this process suggests latent complexity in the mature system. The breadth of the RGC axon arbor, which may impart the potential to synapse onto new TC partners even in the adult, together with short-term plasticity mechanisms that modulate the efficacy of existing contacts, provide the scaffold for dynamic computation beyond the relay of retinal firing patterns to the cortex (Alonso et al., Reference Alonso, Yeh, Weng and Stoelzel2006; Martinez et al., Reference Martinez, Molano-Mazón, Wang, Sommer and Hirsch2014; Usrey & Alitto, Reference Usrey and Alitto2015).

Convergence at the retinogeniculate synapse

Retinogeniculate convergence (and divergence) add complexity to visual processing in the dLGN. The simplest circuit, where 1 RGC contacts 1 TC neuron, is most consistent with the concept of a thalamic “relay” (Glees & Le Gros Clark, Reference Glees and Le Gros Clark1941; Sherman & Guillery, Reference Sherman and Guillery1996). More complex circuits with converging RGC inputs and/or diverging single RGC axons onto multiple target TC neurons increase the likelihood of the emergence of novel visual features or receptive field properties (Dan et al., Reference Dan, Alonso, Usrey and Reid1998; Alonso et al., Reference Alonso, Yeh, Weng and Stoelzel2006; Koepsell et al., Reference Koepsell, Wang, Vaingankar, Wei, Wang, Rathbun, Usrey, Hirsch and Sommer2009; Usrey & Alitto, Reference Usrey and Alitto2015; Sherman, Reference Sherman2016; Weyand, Reference Weyand2016). For these reasons, studies quantifying connectivity, and in particular, the degree of retinogeniculate convergence, is an active area of research.

A tour-de-force serial electron microscopy (EM) reconstruction of the synaptic contacts of one branch of an X-type retinal axon in the cat dLGN demonstrated that a RGC axon makes connections selectively rather than randomly. The reconstructed portion of the axon (reproduced in Fig. 1A) innervated three X-cells and one Y-cell, and its inputs accounted for as much as 33, 49, and 100% of total innervation to the X-cells, and as little as <6% to the Y-cell. This study concluded simultaneously that a TC neuron can receive inputs from multiple RGCs (convergence), and that some of those inputs can also contact other TC neurons (divergence/multiplexing; Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987). However, the single X-cell that received all of its inputs from the labeled axon remains the best-recognized result, serving as exemplary anatomical evidence for low retinogeniculate convergence. In contrast, Robson (Reference Robson1993) estimated that cat Y-cells receive upwards of 10 inputs per cell, suggesting that convergence varies depending on cell type.

Recent studies using new anatomical methods in p30 mice, however, came to a conclusion that counters the general view of low convergence. Morgan and colleagues used an approach that combines serial section EM with circuit tracing, to identify the presynaptic RGC axons that connect to reconstructed postsynaptic TC neurons in the dLGN. They observed at least 40 RGC axon segments contacting one of these TC neurons (Morgan et al., Reference Morgan, Berger, Wetzel and Lichtman2016). Many axons also promiscuously diverged to innervate numerous other TC neurons. Hammer and colleagues reached a number closer to 10 inputs per cell from observations of bouton clustering of multicolor fluorescently labeled RGC axons (Brainbow labeling) in the mouse LGN (Hammer et al., Reference Hammer, Monavarfeshani, Lemon, Su and Fox2015). However, these studies were not able to trace the axon segments to the primary axon, raising the possibility that they were overestimating the number of inputs to a given TC neuron. Overcoming this limitation, and despite low efficiency of rabies tracing, Rompani and colleagues showed that 1–36 RGCs innervate monocular neurons, and up to 91 RGCs from both eyes converge onto binocularly innervated neurons in the mouse dLGN. The three anatomical studies made no distinctions between X- and Y-cells, but the rabies tracing identified three different patterns of convergence in the binocular dLGN region (Rompani et al., Reference Rompani, Mullner, Wanner, Roth, Yonehara, Zhang and Roska2017). Importantly, these studies cannot determine whether all the identified contacts are functional; many convergent inputs could be nonfunctional remnants of refinement, as final pruning of the axon arbor occurs between p30 and p60 (Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). Nonetheless, these independent studies using disparate anatomical methods demonstrate that tens of RGCs may converge onto mature mouse TC neurons.

To date, most functional studies have yielded a more conservative estimate of convergence than the ultrastructural literature. In the rodent slice prep, estimates of the number of afferent inputs obtained by varying the intensity of optic tract stimulation yield numbers ranging from 1 to 5 (Chen & Regehr, Reference Chen and Regehr2000; Jaubert-Miazza et al., Reference Jaubert-Miazza, Green, Lo, Bui, Mills and Guido2005; Ziburkus & Guido, Reference Ziburkus and Guido2006; Hooks & Chen, Reference Hooks and Chen2007; Chung et al., Reference Chung, Clarke, Wang, Stafford, Sher, Chakraborty, Joung, Foo, Thompson, Chen, Smith and Barres2013; Lee et al., Reference Lee, Brott, Kirkby, Adelson, Cheng, Feller, Datwani and Shatz2014; Dilger et al., Reference Dilger, Krahe, Morhardt, Seabrook, Shin and Guido2015). However, this approach likely underestimates convergence due to the severing of axons in slice. It also averages across the population of TC neurons accessible with this method, with no distinction between TC neuron subtypes with different input convergence that may exist in the mouse dLGN (Krahe et al., Reference Krahe, El-Danaf, Dilger, Henderson and Guido2011).

Estimates of retinogeniculate connectivity from carnivores and nonhuman primates have largely utilized recordings that assess the correlation of spiking activity of RGC-TC neuron pairs in vivo (Cleland et al., Reference Cleland, Dubin and Levick1971; Levick et al., Reference Levick, Cleland and Dubin1972; Kaplan & Shapley, Reference Kaplan and Shapley1984; Mastronarde, Reference Mastronarde1987, Reference Mastronarde1992; Usrey et al., Reference Usrey, Reppas and Reid1998; Rowe & Fischer, Reference Rowe and Fischer2001; Carandini et al., Reference Carandini, Horton and Sincich2007; Sincich et al., Reference Sincich, Adams, Economides and Horton2007; Rathbun et al., Reference Rathbun, Warland and Usrey2010). Many of these experiments show that a geniculate X-cell (cat) or M or P cell (primate) receives at least one dominant input that reliably drives EPSCs preceding all or most of a TC neuron’s spikes (Cleland et al., Reference Cleland, Dubin and Levick1971; Cleland & Lee, Reference Cleland and Lee1985; Soodak et al., Reference Soodak, Shapley and Kaplan1987; Sincich et al., Reference Sincich, Adams, Economides and Horton2007). Others, however (especially those focusing on Y cells in cats), show that the contribution from individual RGCs exhibits greater variability, and the activity of a single retinal input rarely accounts for the entirety of the activity of its TC neuron partner (Hubel & Wiesel, Reference Hubel and Wiesel1961; Cleland & Levick, Reference Cleland and Levick1971; Cleland et al., Reference Cleland, Dubin and Levick1971; Levick et al., Reference Levick, Cleland and Dubin1972; Mastronarde, Reference Mastronarde1992). Interestingly, one study using paired recordings across both X- and Y-cells yielded examples of RGCs that drove as few as ∼1% to as many as 82% of a TC neuron’s action potentials (Usrey et al., Reference Usrey, Reppas and Reid1999); similar results later emerged in the Y pathway (Yeh et al., Reference Yeh, Stoelzel, Weng and Alonso2009; Rathbun et al., Reference Rathbun, Alitto, Warland and Usrey2016; considered in detail in; Weyand, Reference Weyand2016). Furthermore, several studies corroborate anatomical observations of divergence, such that neurons with most closely matching receptive fields exhibit the greatest correlation among their firing patterns (Alonso et al., Reference Alonso, Usrey and Reid1996; Usrey et al., Reference Usrey, Reppas and Reid1998).

The variability in the contribution of a single RGC to postsynaptic spiking in cat dLGN is consistent with anatomical studies, if the number of contacts between a single RGC axon and TC cell relates to the functional strength of the individual input (Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987). The findings in cat are also consistent with in vitro functional data from mice. Even in adult mice (p60), the distribution of single RGC input amplitudes ranges from a tens of pA to several nA in strength (Fig. 4D; Hooks & Chen, Reference Hooks and Chen2008; Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014; Thompson et al., Reference Thompson, Picard, Min, Fagiolini and Chen2016). The fact that weak convergent inputs persist into adulthood, in both cats and mice, suggest that they have relevance for retinogeniculate function (Alonso et al., Reference Alonso, Usrey and Reid1996; Dan et al., Reference Dan, Alonso, Usrey and Reid1998; Usrey et al., Reference Usrey, Reppas and Reid1999).

Taken together, the retinogeniculate circuit exhibits organization that is set up to actively tune, select, or elaborate information that is being conveyed from the retina to the cortex, suggesting that the dLGN participates in complex processing of visual information (Sherman, Reference Sherman2016). While similarities across species support this view, insight into the function of dLGN should come from further elucidation of the differences between mice, cat and primates.

Retinogeniculate plasticity

In addition to short-term plasticity, the retinogeniculate circuit exhibits long-term plasticity of synaptic weights. The finding that both weak and strong inputs innervate mature TC neurons and contribute to their spiking activity highlights the possibility that experience-dependent plasticity of RGC input strength and number relies on the balancing of synaptic weights in the adult circuit (Thompson et al., Reference Thompson, Picard, Min, Fagiolini and Chen2016). The retinogeniculate connectivity map remodels to experience during development, and may also do so in mature animals. Depriving juvenile mice of visual experience for a week starting at p20 (late dark rearing) disrupts retinogeniculate connectivity, decreasing the amplitude of the average retinal input and increasing the overall number of RGC inputs onto TC neurons (Hooks & Chen, Reference Hooks and Chen2006, Reference Hooks and Chen2008; Narushima et al., Reference Narushima, Uchigashima, Yagasaki, Harada, Nagumo, Uesaka, Hashimoto, Aiba, Watanabe, Miyata and Kano2016). This manipulation also reduces the clustering of RGC axon boutons without significantly altering the size of the arbor or the number of boutons (Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014). Together, these observations suggest that while connectivity between axons and targets is selective, the large size of the arbor builds flexibility into the system: dramatic change in visual experience, such as late dark rearing can resculpt connectivity by rearranging boutons and adjusting input strength (Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014) without investing into remodeling the entire axonal arbor. Because excess branches of RGC arbors do not prune down until at least p60 in mice, retinogeniculate connectivity may exhibit substantial plasticity until at least this age (Hong et al., Reference Hong, Park, Litvina, Morales, Sanes and Chen2014).

A well-established mechanism for altering synaptic strength in response to activity or experience is through changes in AMPAR content of the post-synaptic density (Huganir & Nicoll, Reference Huganir and Nicoll2013). Several studies link the regulation of AMPAR trafficking and function to modulation of the strength of juvenile retinogeniculate synapses. The retinogeniculate synapse is one among several synapses that recruit GluA1-containing AMPARs in response to sensory stimulation (Clem & Barth, Reference Clem and Barth2006; Kielland et al., Reference Kielland, Bochorishvili, Corson, Zhang, Rosin, Heggelund and Zhu2009; Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014), suggesting that GluA1-dependent AMPAR-driven forms of transmission play an important role in the development and plasticity at this synapse (Fig. 4A–4C; activity-dependent changes in AMPAR subunit composition reviewed in Cull-Candy et al., Reference Cull-Candy, Kelly and Farrant2006; Liu & Zukin, Reference Liu and Zukin2007; Lee et al., Reference Lee, Brott, Kirkby, Adelson, Cheng, Feller, Datwani and Shatz2014). AMPAR subunit content is sensitive to visual experience: mice subjected to late dark rearing exhibited a decrease in AMPAR current rectification, a measure of the fraction of calcium-permeable to calcium-impermeable AMPARs in the postsynaptic density. In contrast, mice that never had any visual experience (chronically dark reared from birth) exhibited normal rectification (Fig. 4C; Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014). Experience-dependent changes in AMPAR content and function at the retinogeniculate synapse rely in part on stargazin, a transmembrane AMPA regulatory protein that modifies the trafficking and channel kinetics of AMPARs (Straub & Tomita, Reference Straub and Tomita2012). Indeed, late dark rearing increases the expression and phosphorylation of stargazin, which can in turn regulate the composition of postsynaptic AMPARs in both a Hebbian (Tomita et al., Reference Tomita, Stein, Stocker, Nicoll and Bredt2005) or homeostatic manner (Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014). Finally, changes in AMPAR expression also mediate the role of MHC class I molecule H2-Db in retinogeniculate developmental refinement (Lee et al., Reference Lee, Brott, Kirkby, Adelson, Cheng, Feller, Datwani and Shatz2014). H2-Db is one of a series of immune-related molecules that shape retinogeniculate development (Shatz, Reference Shatz2009; Schafer & Stevens, Reference Schafer and Stevens2010). Mice lacking H2-Db expression exhibit an increase in calcium-permeable AMPARs at retinogeniculate synapses, corresponding to a deficit in LTD. Together, these studies bespeak a critical role of AMPAR regulation in retinogeniculate synaptic plasticity, which may persist into adulthood.

Both Hebbian and homeostatic mechanisms of synaptic plasticity have been shown to alter retinogeniculate synaptic strength during development (Mooney et al., Reference Mooney, Madison and Shatz1993; Butts et al., Reference Butts, Kanold and Shatz2007; Ziburkus et al., Reference Ziburkus, Dilger, Lo and Guido2009; Krahe & Guido, Reference Krahe and Guido2011; Lin et al., Reference Lin, Kang and Chen2014; Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014). High frequency stimulation of the optic tract, or low frequency stimulation coincident with postsynaptic depolarization in slices from immature ferret dLGN result in long-term enhancement of the EPSC, with contribution from NMDAR activation (Mooney et al., Reference Mooney, Madison and Shatz1993). However, in rat, the same high-frequency stimulus results in long-term depression in dLGN explants before eye-opening, but long-term potentiation later in development (Ziburkus et al., Reference Ziburkus, Dilger, Lo and Guido2009). Finally, plasticity rules based on burst timing have been identified before eye-opening (Butts et al., Reference Butts, Kanold and Shatz2007 in rat), but these plasticity rules have not been examined in more mature slices. On the other hand, the contribution of homeostatic plasticity in retinogeniculate plasticity has been suggested through studies involving monocular deprivation, chronic dark rearing, manipulation of stargazin, and deletion of Mecp2 (a transcription factor necessary for homeostatic scaling up in the visual cortex; Blackman et al., Reference Blackman, Djukic, Nelson and Turrigiano2012; Noutel et al., Reference Noutel, Hong, Leu, Kang and Chen2011; Krahe & Guido, Reference Krahe and Guido2011; Lin et al., Reference Lin, Kang and Chen2014; Louros et al., Reference Louros, Hooks, Litvina, Carvalho and Chen2014). Similar paradigms likely also drive synaptic plasticity at the fully mature retinogeniculate synapse.

Recently described instances of rapid plasticity across species could also engage Hebbian or homeostatic mechanisms at the retinogeniculate synapse and cause a shift in the strength of individual retinogeniculate inputs (Moore et al., Reference Moore, Kiley, Sun and Usrey2011; Aguila et al., Reference Aguila, Cudeiro and Rivadulla2017). In fact, dLGN neurons readily adapt to changes in visual input. For example, pharmacologic blockade of On-center RGC activity in adult cats rapidly uncovers Off-center responses in dLGN neurons previously exhibiting On-center responses, instead of silencing them (Moore et al., Reference Moore, Kiley, Sun and Usrey2011). Further, acute suppression of cortical feedback in awake monkeys shifted the receptive field position of a subset of TC neurons (Aguila et al., Reference Aguila, Cudeiro and Rivadulla2017). Finally, rabbit geniculate neurons exhibit bidirectional sensory adaptation that improves signal detection (Stoelzel et al., Reference Stoelzel, Huff, Bereshpolova, Alonso, Swadlow, Zhuang, Hei, Alonso and Swadlow2015). While modulatory mechanisms could shape the response of the TC neuron to rapid changes in upstream inputs, the specificity that TC neuron responses exhibit in these studies indicates instead a role for rapid shifts in the synaptic efficacy of retinogeniculate connections. These changes occur too rapidly to rely on structural rearrangements of retinogeniculate connections, their timecourse is consistent with possible unsilencing or strengthening of functionally silent or weak subthreshold inputs via the insertion of postsynaptic receptors at dormant synaptic sites. Of course, local inhibitory circuits may also contribute to rapid shifts in RGC input efficacy (Fisher et al., Reference Fisher, Alitto and Usrey2017). Changes that occur over days rather than hours may also recruit the redistribution of presynaptic boutons along the broad RGC axon arbor. The capacity of the diverse types of retinogeniculate synaptic contacts for functional plasticity remains unexplored.

The expression of a variety of mechanisms for modification of synaptic weights is layered on top of a potentially densely interconnected network that is evident in considerable retinogeniculate divergence and convergence. Combined with the observation that both weak and strong inputs innervate mature TC neurons, these plasticity mechanism may endow the dLGN with a role in visual learning on multiple time scales (Ramos et al., Reference Ramos, Schwartz and Roy1976; Albrecht et al., Reference Albrecht, Davidowa and Gabriel1990).

Conclusion

Recent work is revealing new complexities of retinogeniculate transmission and circuit organization that further expand the potential role of the dLGN in visual processing beyond its classic attributes (Steriade et al., Reference Steriade, Jones and McCormick1997; Sherman & Guillery, Reference Sherman and Guillery2001). The morphological diversity of synaptic motifs and complex connectivity patterns, combined with short- and long-term plasticity mechanisms of the retinogeniculate circuit demonstrate that the retinogeniculate synapse makes substantial and dynamic contributions to the processing of visual information. Weak or non-dominant retinogeniculate inputs in the mature dLGN, which have repeatedly been dismissed as insignificant, as errors of development or leftovers from developmental plasticity with no functional relevance (Sur et al., Reference Sur, Weller and Sherman1984; Garraghty et al., Reference Garraghty, Salinger and Macavoy1985; Hamos et al., Reference Hamos, Van Horn, Raczkowski and Sherman1987), likely enhance extraction of visual features in the geniculate and visual cortex, and serve as strategic reserves of plasticity. Moreover, the convergence of potentially heterogeneous RGC inputs onto single geniculate neurons could give rise to new receptive field features such as orientation selectivity in mouse dLGN, as recently proposed (Stafford & Huberman, Reference Stafford and Huberman2017). Finally, the interplay between the strength and short-term plasticity properties of RGC inputs in the context of convergence and divergence adds to the richness of the dLGN circuitry.

Much is still left to understand about the extent and underlying basis of plasticity in the dLGN. However, the idea that the mature geniculate system can utilize activity-dependent plasticity mechanisms to fine-tune the contribution of its individual inputs in response to novel visual challenges, experiences, changes in modulatory state, or retinal degeneration appears to be rapidly gaining experimental support. Future studies are needed to clarify the differences in the number of inputs that converge onto TC neurons between species, because the results may correlate with the degree by which new receptive field features emerge at the level of dLGN. Elucidation of whether and how weak inputs contribute to visual processing in different species will also uncover the richness of thalamic function.

The continually expanding toolbox for interrogating diversity in neuronal circuits is already uncovering nuances in the contribution of different RGC types to TC neuron function (Storchi et al., Reference Storchi, Milosavljevic, Eleftheriou, Martial, Orlowska-Feuer, Bedford, Brown, Montemurro, Petersen and Lucas2015; Denman et al., Reference Denman, Siegle, Koch, Reid and Blanche2016). Advances in methods for labeling, activating, and measuring the activity of different neuronal populations that have been deployed extensively in mice may also reveal underappreciated subtleties of retinogeniculate transmission in other species (Scholl et al., Reference Scholl, Tan, Corey and Priebe2013; Zaltsman et al., Reference Zaltsman, Heimel and Van Hooser2015; Zeater et al., Reference Zeater, Cheong, Solomon, Dreher and Martin2015; Suresh et al., Reference Suresh, Ciftcio lu, Wang, Lala, Ding, Smith, Sommer and Hirsch2016). Shifting models of the organization of the visual system that take into account the important nuances of retinogeniculate functional organization and plasticity are certain to provide new models of visual system development and function.

Acknowledgment

The authors thank C.-E. Stephany, G. Rankin and A.D. Thompson for their critical reading and helpful discussions on the manuscript. This work was supported by the NIH R01EY013613 and NIH U54 HD090255 to CC and the Edward R. and Anne G. Lefler Center Predoctoral Fellowship to EYL.

References

Ackman, J.B., Burbridge, T.J. & Crair, M.C. (2012). Retinal waves coordinate patterned activity throughout the developing visual system. Nature 490, 219225.Google Scholar
Aguila, J., Cudeiro, F.J. & Rivadulla, C. (2017). Suppression of V1 feedback produces a shift in the topographic representation of receptive fields of LGN cells by unmasking latent retinal drives. Cerebral Cortex 27, 33313345.Google Scholar
Akerman, C.J., Smyth, D. & Thompson, I.D. (2002). Visual experience before eye-opening and the development of the retinogeniculate pathway. Neuron 36, 869879.Google Scholar
Albrecht, D., Davidowa, H. & Gabriel, H. (1990). Conditioning-related changes of unit activity in the dorsal lateral geniculate nucleus of urethane-anaesthetized rats. Brain Research Bulletin 25, 5563.Google Scholar
Alitto, H.J., Moore, B.D., Rathbun, D.L. & Martin Usrey, W. (2011). A comparison of visual responses in the lateral geniculate nucleus of alert and anaesthetized macaque monkeys. The Journal of Physiology 589, 8799.CrossRefGoogle ScholarPubMed
Alonso, J.M., Usrey, W.M. & Reid, R.C. (1996). Precisely correlated firing in cells of the lateral geniculate nucleus. Nature 383, 815819.Google Scholar
Alonso, J.M., Yeh, C.I., Weng, C. & Stoelzel, C.R. (2006). Retinogeniculate connections: A balancing act between connection specificity and receptive field diversity. Progress in Brain Research 154, 313.Google Scholar
Augustinaite, S. & Heggelund, P. (2007). Changes in firing pattern of lateral geniculate neurons caused by membrane potential dependent modulation of retinal input through NMDA receptors. The Journal of Physiology 582, 297315.CrossRefGoogle ScholarPubMed
Becker, N., Wierenga, C.J., Fonseca, R., Bonhoeffer, T. & Nägerl, U.V. (2008). LTD induction causes morphological changes of presynaptic boutons and reduces their contacts with spines. Neuron 60, 590597.Google Scholar
Benowitz, L.I., He, Z. & Goldberg, J.L. (2017). Reaching the brain: Advances in optic nerve regeneration. Experimental Neurology 287, 365373.Google Scholar
Bickford, M.E., Slusarczyk, A., Dilger, E.K., Krahe, T.E., Kucuk, C. & Guido, W. (2010). Synaptic development of the mouse dorsal lateral geniculate nucleus. Journal of Comparative Neurology 518, 622635.CrossRefGoogle ScholarPubMed
Bishop, P., Burke, W. & Davis, R. (1959). Activation of single lateral geniculate cells by stimulation of either optic nerve. Science 130, 506507.Google Scholar
Blackman, M.P., Djukic, B., Nelson, S.B. & Turrigiano, G.G. (2012). A critical and cell-autonomous role for MeCP2 in synaptic scaling up. The Journal of Neuroscience 32, 1352913536.Google Scholar
Blakemore, C. & Vital-durand, F. (1985). Organization and post-natal development of the monkey’s lateral geniculate nucleus. Journal of Physiology 380, 453491.CrossRefGoogle Scholar
Blitz, D.M., Foster, K.A. & Regehr, W.G. (2004). Short-term synaptic plasticity: A comparison of two synapses. Nature Reviews Neuroscience 5, 630640.Google Scholar
Blitz, D.M. & Regehr, W.G. (2003). Retinogeniculate synaptic properties controlling spike number and timing in relay neurons. Journal of Neurophysiology 90, 24382450.Google Scholar
Borst, J.G. & Sakmann, B. (1996). Calcium influx and transmitter release in a fast CNS synapse. Nature 383, 431434.Google Scholar
Budisantoso, T., Matsui, K., Kamasawa, N., Fukazawa, Y. & Shigemoto, R. (2012). Mechanisms underlying signal filtering at a multisynapse contact. The Journal of Neuroscience 32, 23572376.Google Scholar
Butts, D.A., Kanold, P.O. & Shatz, C.J. (2007). A burst-based ‘Hebbian’ learning rule at retinogeniculate synapses links retinal waves to activity-dependent refinement. PLoS Biology 5, e61.Google Scholar
Butts, D.A. & Rokhsar, D.S. (2001). The information content of spontaneous retinal waves. Journal of Neuroscience 21, 961973.Google Scholar
Cai, D., DeAngelis, G.C. & Freeman, R.D. (1997). Spatiotemporal receptive field organization in the lateral geniculate nucleus of cats and kittens. Journal of Neurophysiology 78, 10451061.Google Scholar
Campbell, G. & Shatz, C.J. (1992). Synapses formed by identified retinogeniculate axons during the segregation of eye input. The Journal of Neuroscience 12, 18471858.Google Scholar
Cang, J. & Feldheim, D.A. (2013). Developmental mechanisms of topographic map formation and alignment. Annual Review of Neuroscience 36, 5177.Google Scholar
Carandini, M., Horton, J.C. & Sincich, L.C. (2007). Thalamic filtering of retinal spike trains by postsynaptic summation. Journal of Vision 7, 20.120.11.Google Scholar
Carden, W.B., Datskovskaia, A., Guido, W., Godwin, D.W. & Bickford, M.E. (2000). Development of the cholinergic, nitrergic, and GABAergic innervation of the cat dorsal lateral geniculate nucleus. The Journal of Comparative Neurology 418, 6580.Google Scholar
Chen, C., Blitz, D.M. & Regehr, W.G. (2002). Contributions of receptor desensitization and saturation to plasticity at the retinogeniculate synapse. Neuron 33, 779788.Google Scholar
Chen, C. & Regehr, W.G. (2000). Developmental remodeling of the retinogeniculate synapse. Neuron 28, 955966.Google Scholar
Chen, C. & Regehr, W.G. (2003). Presynaptic modulation of the retinogeniculate synapse. The Journal of Neuroscience 23, 31303135.Google Scholar
Cheong, S.K.K., Tailby, C., Solomon, S.G. & Martin, P.R. (2013). Cortical-like receptive fields in the lateral geniculate nucleus of marmoset monkeys. Journal of Neuroscience 33, 68646876.CrossRefGoogle ScholarPubMed
Chung, W-S., Clarke, L.E., Wang, G.X., Stafford, B.K., Sher, A., Chakraborty, C., Joung, J., Foo, L.C., Thompson, A., Chen, C., Smith, S.J. & Barres, B.A. (2013). Astrocytes mediate synapse elimination through MEGF10 and MERTK pathways. Nature 504, 394400.Google Scholar
Cleland, B.G., Dubin, M.W. & Levick, W.R. (1971). Sustained and transient neurones in the cat’s retina and laterla geniculate nucleus. Journal of Physiology 217, 473496.Google Scholar
Cleland, B.G. & Lee, B.B. (1985). A comparison of visual responses of cat lateral geniculate nucleus neurones with those of ganglion cells afferent to them. Journal of Physiology 369, 249268.Google Scholar
Cleland, B.G. & Levick, W.R. (1971). Simultaneous recording of input and output of lateral geniculate neurones. Nature 231, 191192.Google Scholar
Clem, R.L. & Barth, A. (2006). Pathway-specific trafficking of native AMPARs by in vivo experience. Neuron 49, 663670.Google Scholar
Clements, J., Lester, R., Tong, G., Jahr, C. & Westbrook, G. (1992). The time course of glutamate in the synaptic cleft. Science 258, 14981501.Google Scholar
Conley, M. & Fitzpatrick, D. (1989). Morphology of retinogeniculate axons in the macaque. Visual Neuroscience 2, 287296.Google Scholar
Cull-Candy, S.G., Kelly, L. & Farrant, M. (2006). Regulation of Ca2+-permeable AMPA receptors: Synaptic plasticity and beyond. Current Opinion in Neurobiology 16, 288297.Google Scholar
Dan, Y., Alonso, J.M., Usrey, W.M. & Reid, R.C. (1998). Coding of visual information by precisely correlated spikes in the lateral geniculate nucleus. Nature Neuroscience 1, 501507.CrossRefGoogle ScholarPubMed
Danbolt, N.C. (2001). Glutamate uptake. Progress in Neurobiology 65, 1105.Google Scholar
Daniels, J.D., Norman, J.L. & Pettigrew, J.D. (1977). Biases for oriented moving bars in lateral geniculate nucleus neurons of normal and stripe-reared cats. Experimental Brain Research 29, 155–72.Google Scholar
Daniels, J.D., Pettigrew, J.D. & Norman, J.L. (1978). Development of single-neuron responses in kitten’s lateral geniculate nucleus. Journal of Neurophysiology 41, 13731393.Google Scholar
Davis, Z.W., Chapman, B. & Cheng, J.L. (2015). Increasing spontaneous retinal activity before eye opening accelerates the development of geniculate receptive fields. Journal of Neuroscience 35, 1461214623.Google Scholar
Denman, D., Siegle, J.H., Koch, C., Reid, R.C. & Blanche, T.J. (2016). Spatial organization of chromatic pathways in the mouse dorsal lateral geniculate nucleus. Journal of Neuroscience 37, 11021116.Google Scholar
Dhande, O.S., Hua, E.W., Guh, E., Yeh, J., Bhatt, S., Zhang, Y., Ruthazer, E.S., Feller, M.B. & Crair, M.C. (2011). Development of single retinofugal axon arbors in normal and β2 knock-out mice. The Journal of Neuroscience 31, 33843399.Google Scholar
Diamond, J.S. & Jahr, C.E. (1997). Transporters buffer synaptically released glutamate on a submillisecond time scale. The Journal of Neuroscience 17, 46724687.Google Scholar
Dilger, E.K., Krahe, T.E., Morhardt, D.R., Seabrook, T.A., Shin, H-S. & Guido, W. (2015). Absence of plateau potentials in dLGN cells leads to a breakdown in retinogeniculate refinement. The Journal of Neuroscience 35, 36523662.CrossRefGoogle ScholarPubMed
El-Danaf, R.N., Krahe, T.E., Dilger, E.K., Bickford, M.E., Fox, M. & Guido, W. (2015). Developmental remodeling of relay cells in the dorsal lateral geniculate nucleus in the absence of retinal input. Neural Development 10, 19.Google Scholar
Famiglietti, E.V. & Peters, A. (1972). The synaptic glomerulus and the intrinsic neuron in the dorsal lateral geniculate nucleus of the cat. The Journal of Comparative Neurology 144, 285333.Google Scholar
Felch, D.L. & Van Hooser, S.D. (2012). Molecular compartmentalization of lateral geniculate nucleus in the gray squirrel (Sciurus carolinensis). Frontiers in Neuroanatomy 6, 12.Google Scholar
Feller, M.B. (1999). Spontaneous Correlated Activity in Developing Neural Circuits. Neuron 22, 653656.Google Scholar
Feller, M.B. (2009). Retinal waves are likely to instruct the formation of eye-specific retinogeniculate projections. Neural Development 4, 24.Google Scholar
Fisher, T.G., Alitto, H.J. & Usrey, W.M. (2017). Retinal and non-retinal contributions to extraclassical surround suppression in the lateral geniculate nucleus. Journal of Neuroscience 37, 226235.Google Scholar
Freche, D., Pannasch, U., Rouach, N. & Holcman, D. (2011). Synapse geometry and receptor dynamics modulate synaptic strength. PLoS ONE 6, e25122.Google Scholar
Fukuda, Y., Hsiao, C.F., Watanabe, M. & Ito, H. (1984). Morphological correlates of physiologically identified Y-, X-, and W-cells in cat retina. Journal of Neurophysiology 52, 9991013.CrossRefGoogle Scholar
Garraghty, P.E., Salinger, W.L. & Macavoy, M.G. (1985). The development of cell size in the dorsal lateral geniculate nucleus of monocularly paralyzed cats. Brain Research 353, 99106.Google Scholar
Garraghty, P.E., Shatz, C.J., Sretavan, D.W. & Sur, M. (1988). Axon arbors of X and Y retinal ganglion cells are differentially affected by prenatal disruption of binocular inputs. Proceedings of the National Academy of Sciences 85, 73617365.CrossRefGoogle ScholarPubMed
Garraghty, P.E. & Sur, M. (1993). Competitive interactions influencing the development of retinal axonal arbors in cat lateral geniculate nucleus. Physiological Reviews 73, 529545.Google Scholar
Gary-Bobo, E., Przybyslawski, J. & Saillour, P. (1995). Experience-dependent maturation of the spatial and temporal characteristics of the cell receptive fields in the kitten visual cortex. Neuroscience Letters 189, 147150.Google Scholar
Glees, P. & Le Gros Clark, W.E. (1941). The termination of optic fibres in the lateral geniculate body of the monkey. The Journal of Anatomy 75, 295307.Google Scholar
Golding, B., Pouchelon, G., Bellone, C., Murthy, S., Di Nardo, A.A., Govindan, S., Ogawa, M., Shimogori, T., Lüscher, C., Dayer, A. & Jabaudon, D. (2014). Retinal input directs the recruitment of inhibitory interneurons into thalamic visual circuits. Neuron 81, 10571069.Google Scholar
Grant, E., Hoerder-Suabedissen, A. & Molnar, Z. (2016). The regulation of corticofugal fiber targeting by retinal inputs. Cerebral Cortex 26, 13361348.Google Scholar
Graydon, C.W., Cho, S., Diamond, J.S., Kachar, B., von Gersdorff, H. & Grimes, W.N. (2014). Specialized postsynaptic morphology enhances neurotransmitter dilution and high-frequency signaling at an auditory synapse. The Journal of Neuroscience 34, 83588372.CrossRefGoogle ScholarPubMed
Guido, W. (2008). Refinement of the retinogeniculate pathway. The Journal of Physiology 586, 43574362.Google Scholar
Guillery, R. & Sherman, S.M. (2002). Thalamic relay functions and their role in corticocortical communication: Generalizations from the visual system. Neuron 33, 163175.Google Scholar
Hammer, S., Monavarfeshani, A., Lemon, T., Su, J. & Fox, M.A. (2015). Multiple retinal axons converge onto relay cells in the adult mouse thalamus. Cell Reports 12, 15751583.CrossRefGoogle ScholarPubMed
Hamos, J.E., Van Horn, S.C., Raczkowski, D. & Sherman, S.M. (1987). Synaptic circuits involving an individual retinogeniculate axon in the cat. The Journal of Comparative Neurology 259, 165192.Google Scholar
Hamos, J.E., Van Horn, S.C., Raczkowski, D., Uhlrich, D.J. & Sherman, S.M. (1985). Synaptic connectivity of a local circuit neurone in lateral geniculate nucleus of the cat. Nature 317, 618621.Google Scholar
Hanganu, I.L., Ben-Ari, Y. & Khazipov, R.R. (2006). Retinal waves trigger spindle bursts in the neonatal rat visual cortex. The Journal of Neuroscience 26, 67286736.Google Scholar
Hauser, J.L., Edson, E.B., Hooks, B.M. & Chen, C. (2013). Metabotropic glutamate receptors and glutamate transporters shape transmission at the developing retinogeniculate synapse. Journal of Neurophysiology 109, 113123.CrossRefGoogle ScholarPubMed
Hauser, J.L., Liu, X., Litvina, E.Y. & Chen, C. (2014). Prolonged synaptic currents increase relay neuron firing at the developing retinogeniculate synapse. Journal of Neurophysiology 112, 17141728.Google Scholar
Hei, X., Stoelzel, C.R., Zhuang, J., Bereshpolova, Y., Huff, J.M., Alonso, J.M. & Swadlow, H.A. (2014). Directional selective neurons in the awake LGN: Response properties and modulation by brain state. Journal of Neurophysiology 112, 362373.CrossRefGoogle ScholarPubMed
Hong, Y.K. & Chen, C. (2011). Wiring and rewiring of the retinogeniculate synapse. Current Opinion in Neurobiology 21, 228237.Google Scholar
Hong, Y.K., Park, S., Litvina, E.Y., Morales, J., Sanes, J.R. & Chen, C. (2014). Refinement of the retinogeniculate synapse by bouton clustering. Neuron 84, 332339.Google Scholar
Hooks, B.M. & Chen, C. (2006). Distinct roles for spontaneous and visual activity in remodeling of the retinogeniculate synapse. Neuron 52, 281291.Google Scholar
Hooks, B.M. & Chen, C. (2007). Critical periods in the visual system: Changing views for a model of experience-dependent plasticity. Neuron 56, 312326.Google Scholar
Hooks, B.M. & Chen, C. (2008). Vision triggers an experience-dependent sensitive period at the retinogeniculate synapse. The Journal of Neuroscience 28, 48074817.Google Scholar
Howarth, M., Walmsley, L. & Brown, T.M. (2014). Report binocular integration in the mouse lateral geniculate nuclei. Current Biology 24, 12411247.Google Scholar
Hubel, D.H. & Wiesel, T.N. (1961). Integrative action in the cat’s lateral geniculate body. The Journal of Physiology 155, 385398.Google Scholar
Hubel, D.H. & Wiesel, T.N. (1970). The period of susceptibility to the physiological effects of unilateral eye closure in kittens. The Journal of Physiology 206, 419436.Google Scholar
Huberman, A.D., Feller, M.B. & Chapman, B. (2008a). Mechanisms underlying development of visual maps and receptive fields. Annual Review of Neuroscience 31, 479509.Google Scholar
Huberman, A.D., Manu, M., Koch, S.M., Susman, M.W., Lutz, A.B., Ullian, E.M., Baccus, S.A. & Barres, B.A. (2008b). Architecture and activity-mediated refinement of axonal projections from a mosaic of genetically identified retinal ganglion cells. Neuron 59, 425438.Google Scholar
Huganir, R.L. & Nicoll, R.A. (2013). AMPARs and synaptic plasticity: The last 25 years. Neuron 80, 704717.CrossRefGoogle ScholarPubMed
Isaac, J.T.R., Nicoll, R.A. & Malenka, R.C. (1995). Evidence for silent synapses: Implications for the expression of LTP. Neuron 15, 427434.Google Scholar
Jacobs, E.C., Campagnoni, C., Kampf, K., Reyes, S.D., Kalra, V., Handley, V., Xie, Y-Y., Hong-Hu, Y., Spreur, V., Fisher, R.S. & Campagnoni, A.T. (2007). Visualization of corticofugal projections during early cortical development in a τ-GFP-transgenic mouse. European Journal of Neuroscience 25, 1730.Google Scholar
Jaubert-Miazza, L., Green, E., Lo, F.-S., Bui, K., Mills, J. & Guido, W. (2005). Structural and functional composition of the developing retinogeniculate pathway in the mouse. Visual Neuroscience 22, 661676.Google Scholar
Jones, E.G. & Powell, T.P.S. (1969). Electron microscopy of synaptic glomruli in the thalamic relay nuclei in cat. Proceedings of the Royal Society 172, 153171.Google Scholar
Kano, M. & Hashimoto, K. (2009). Synapse elimination in the central nervous system. Current Opinion in Neurobiology 19, 154161.CrossRefGoogle ScholarPubMed
Kaplan, E. (2014). The M, P and K pathways of the primate visual system revisited Ehud Kaplan. In The New Visual Neuroscience, pp. 215226. MIT Press.Google Scholar
Kaplan, E. & Shapley, R. (1984). The origin of the S (slow) potential in the mammalian lateral geniculate nucleus. Experimental Brain Research 55, 111116.Google Scholar
Kastner, S., Schneider, K.A. & Wunderlich, K. (2006). Beyond a relay nucleus: neuroimaging views on the human LGN. Progress in Brain Research 155, 125143.Google Scholar
Katz, L.C. & Shatz, C.J. (1996). Synaptic activity and the construction of cortical circuits. Science 274, 11331138.CrossRefGoogle ScholarPubMed
Kielland, A., Bochorishvili, G., Corson, J., Zhang, L., Rosin, D.L., Heggelund, P. & Zhu, J.J. (2009). Activity patterns govern synapse-specific AMPA receptor trafficking between deliverable and synaptic pools. Neuron 62, 84101.Google Scholar
Kielland, A. & Heggelund, P. (2002). AMPA and NMDA currents show different short-term depression in the dorsal lateral geniculate nucleus of the rat. The Journal of Physiology 542, 99106.CrossRefGoogle ScholarPubMed
Kim, I-J., Zhang, Y., Meister, M. & Sanes, J.R. (2010). Laminar restriction of retinal ganglion cell dendrites and axons: Subtype-specific developmental patterns revealed with transgenic markers. The Journal of Neuroscience 30, 14521462.Google Scholar
Koch, C. (1985). Understanding the intrinsic circuitry of the cat’s lateral geniculate nucleus: Electrical properties of the spine-triad arrangement. Proceedings of the Royal Society of London Series B 225, 365390.Google ScholarPubMed
Koepsell, K., Wang, X., Vaingankar, V., Wei, Y., Wang, Q., Rathbun, D.L., Usrey, W.M., Hirsch, J.A. & Sommer, F.T. (2009). Retinal oscillations carry visual information to cortex. Frontiers in Systems Neuroscience 3, 4.CrossRefGoogle ScholarPubMed
Krahe, T.E., El-Danaf, R.N., Dilger, E.K., Henderson, S.C. & Guido, W. (2011). Morphologically distinct classes of relay cells exhibit regional preferences in the dorsal lateral geniculate nucleus of the mouse. Journal of Neuroscience 31, 1743717448.Google Scholar
Krahe, T.E. & Guido, W. (2011). Homeostatic plasticity in the visual thalamus by monocular deprivation. The Journal of Neuroscience 31, 68426849.Google Scholar
Kumar, S.S., Bacci, A., Kharazia, V. & Huguenard, J.R. (2002). A developmental switch of AMPA receptor subunits in neocortical pyramidal neurons. The Journal of Neuroscience 22, 30053015.Google Scholar
Kwon, Y.H., Esguerra, M. & Sur, M. (1991). NMDA and non-NMDA receptors mediate visual responses of neurons in the cat’s lateral geniculate nucleus. Journal of Neurophysiology 66, 414428.Google Scholar
Lachica, E.A. & Casagrande, V.A. (1988). Development of primate retinogeniculate axons. Visual Neuroscience 1, 103123.CrossRefGoogle Scholar
Lam, Y-W. & Sherman, S.M. (2013). Activation of both group I and group II metabotropic glutamatergic receptors suppress retinogeniculate transmission. Neuroscience 242, 7884.Google Scholar
Lee, B.B., Creutzfeldt, O.D. & Elepfandt, A. (1979). The responses of magno- and parvocellular cells of the monkey’s lateral geniculate body to moving stimuli. Experimental Brain Research 35, 547557.Google Scholar
Lee, H., Brott, B.K., Kirkby, L.A., Adelson, J.D., Cheng, S., Feller, M.B., Datwani, A. & Shatz, C.J. (2014). Synapse elimination and learning rules co-regulated by MHC class I H2-Db. Nature 509, 195200.Google Scholar
Levick, A.W.R., Oyster, C.W. & Takahashi, E. (2010). Rabbit lateral geniculate nucleus: Sharpener of directional information rabbit lateral geniculate nucleus: Sharpener of directional information. Science 165, 712714.Google Scholar
Levick, W.R., Cleland, B.G. & Dubin, M.W. (1972). Lateral geniculate neurons of cat: Retinal inputs and physiology. Investigative Ophthalmology 11, 302311.Google Scholar
Levick, W.R. & Thibos, L.N.L.N. (1980). Orientation bias of cat retinal ganglion cell. Nature 286, 389390.Google Scholar
Levick, W.R. & Williams, W.O. (1964). Maintained activity of lateral geniculate neurones in darkness. The Journal of Physiology 170, 582597.Google Scholar
Levine, M.W. & Cleland, B.G. (2001). An analysis of the effect of retinal ganglion cell impulses upon the firing probability of neurons in the dorsal lateral geniculate nucleus of the cat. Brain Research 902, 244254.Google Scholar
Liao, D., Hessler, N.A. & Malinow, R. (1995). Activation of postsynaptically silent synapses during pairing-induced LTP in CA1 region of hippocampal slice. Nature 375, 400404.Google Scholar
Lin, D.J.-P., Kang, E. & Chen, C. (2014). Changes in input strength and number are driven by distinct mechanisms at the retinogeniculate synapse. Journal of Neurophysiology 112, 942950.Google Scholar
Liu, S.J. & Zukin, R.S. (2007). Ca2+-permeable AMPA receptors in synaptic plasticity and neuronal death. Trends in Neurosciences 30, 126134.Google Scholar
Liu, X. & Chen, C. (2008). Different roles for AMPA and NMDA receptors in transmission at the immature retinogeniculate synapse. Journal of Neurophysiology 99, 629643.Google Scholar
Lo, F., Ziburkus, J. & Guido, W. (2013). Synaptic mechanisms regulating the activation of a Ca2+-mediated plateau potential in developing relay cells of the LGN. Journal of Neurophysiology 87, 11751185.Google Scholar
Louros, S.R., Hooks, B.M., Litvina, L., Carvalho, A.L. & Chen, C. (2014). A role for stargazin in experience-dependent plasticity. Cell Reports 7, 16141625.Google Scholar
Lund, R.D. & Cunningham, T.J. (1972). Aspects of synaptic and laminar organization of the mammalian lateral geniculate body. Investigative ophthalmology 11, 291302.Google Scholar
Luo, L. & O’Leary, D.D.M. (2005). Axon retraction and degeneration in development and disease. Annual Review of Neuroscience 28, 127156.Google Scholar
Macleod, N., Turner, C. & Edgar, J. (1997). Properties of developing lateral geniculate neurones in the mouse. International Journal of Developmental Neuroscience 15, 205224.CrossRefGoogle ScholarPubMed
Marshel, J.H., Kaye, A.P., Nauhaus, I. & Callaway, E.M. (2012). Anterior–posterior direction opponency in the superficial mouse lateral geniculate nucleus. Neuron 76, 713720.Google Scholar
Martinez, L.M., Molano-Mazón, M., Wang, X., Sommer, F.T. & Hirsch, J.A. (2014). Statistical wiring of thalamic receptive fields optimizes spatial sampling of the retinal image. Neuron 81, 943956.Google Scholar
Mason, C.A. (1982). Development of terminal arbors of retino-geniculate axons in the kitten–I. Light microscopical observations. Neuroscience 7, 541559.Google Scholar
Mastronarde, D.N. (1987). Two classes of single-input X-cells in cat lateral geniculate nucleus. II. Retinal inputs and the generation of receptive-field properties. Journal of Neurophysiology 57, 381413.Google Scholar
Mastronarde, D.N. (1992). Nonlagged relay cells and interneurons in the cat lateral geniculate nucleus: Receptive-field properties and retinal inputs. Visual Neuroscience 8, 407441.Google Scholar
Matthews, G. & Fuchs, P. (2010). The diverse roles of ribbon synapses in sensory neurotransmission. Nature Reviews Neuroscience 11, 812822.Google Scholar
McCormick, D.A. & Bal, T. (1994). Sensory gating mechanisms of the thalamus. Current Opinion in Neurobiology 4, 550555.Google Scholar
Michael, C.R. (1988). Retinal afferent arborization patterns, dendritic field orientations, and the segregation of function in the lateral geniculate nucleus of the monkey. Proceedings of the National Academy of Sciences 85, 49144918.Google Scholar
Montero, V.M. (1991). A quantitative study of synpatic contacts on interneurons and relay cells of the cat lateral geniculate nucleus. Experimental Brain Research 86, 2570270.Google Scholar
Montero, V.M. & Wenthold, R.J. (1989). Quantitative immunogold analysis reveals high glutamate levels in retinal and cortical synaptic terminals in the lateral geniculate nucleus of the macaque. Neuroscience 31, 639647.Google Scholar
Mooney, R., Madison, D.V. & Shatz, C.J. (1993). Enhancement of transmission at the developing retinogeniculate synapse. Neuron 10, 815825.Google Scholar
Mooney, R., Penn, A.A., Gallego, R. & Shatz, C.J. (1996). Thalamic relay of spontaneous retinal activity prior to vision. Neuron 17, 863874.Google Scholar
Moore, B.D., Kiley, C.W., Sun, C. & Usrey, W.M. (2011). Rapid plasticity of visual responses in the adult lateral geniculate nucleus. Neuron 71, 812819.Google Scholar
Morgan, J.L., Berger, D.R., Wetzel, A.W. & Lichtman, J.W. (2016). The fuzzy logic of network connectivity in mouse visual thalamus. Cell 165, 192206.Google Scholar
Murphy, G.J. & du Lac, S. (2001). Postnatal development of spike generation in rat medial vestibular nucleus neurons. Journal of Neurophysiology 85, 18991906.Google Scholar
Narushima, M., Uchigashima, M., Yagasaki, Y., Harada, T., Nagumo, Y., Uesaka, N., Hashimoto, K., Aiba, A., Watanabe, M., Miyata, M. & Kano, M. (2016). The metabotropic glutamate receptor subtype 1 mediates experience-dependent maintenance of mature synaptic connectivity in the visual thalamus. Neuron 91, 10971109.Google Scholar
Nirenberg, S. & Meister, M. (1997). The light response of retinal ganglion cells is truncated by a displaced amacrine circuit. Neuron 18, 637650.Google Scholar
Noutel, J., Hong, Y.K., Leu, B., Kang, E. & Chen, C. (2011). Experience-dependent retinogeniculate synapse remodeling is abnormal in MeCP2-deficient mice. Neuron 70, 3542.Google Scholar
Owens, M.T., Feldheim, D.A., Stryker, M.P. & Triplett, J.W. (2015). Stochastic interaction between neural activity and molecular cues in the formation of topographic maps. Neuron 87, 12611273.Google Scholar
Pirchio, M., Turner, J.P., Williams, S.R., Asprodini, E. & Crunelli, V. (1997). Postnatal development of membrane properties and delta oscillations in thalamocortical neurons of the cat dorsal lateral geniculate nucleus. Journal of Neuroscience 17, 54285444.Google Scholar
Piscopo, D.M., El-Danaf, R.N., Huberman, A.D. & Niell, C. (2013). Diverse visual features encoded in mouse lateral geniculate nucleus. Journal of Neuroscience 33, 46424656.Google Scholar
Rafols, J.A. & Valverde, F. (1973). The structure of the dorsal lateral geniculate nucleus in the mouse. A Golgi and electron microscopic study. The Journal of Comparative Neurology 150, 303331.Google Scholar
Ramoa, A.S. & McCormick, D.A. (1994a). Enhanced activation of NMDA receptor responses at the immature retinogeniculate synapse. The Journal of Neuroscience 14, 20982105.Google Scholar
Ramoa, A.S. & McCormick, D.A. (1994b). Developmental changes in electrophysiological properties of LGNd neurons during reorganization of retinogeniculate connections. The Journal of Neuroscience 14, 20892097.Google Scholar
Ramoa, A.S. & Prusky, G. (1997). Retinal activity regulates developmental switches in functional properties and ifenprodil sensitivity of NMDA receptors in the lateral geniculate nucleus. Developmental Brain Research 101, 165175.Google Scholar
Ramos, A., Schwartz, E.L. & Roy, E. (1976). Stable and plastic unit discharge patterns during behavioral generalization. Science 192, 393396.Google Scholar
Rathbun, D.L., Alitto, H.J., Warland, D.K. & Usrey, W.M. (2016). Stimulus contrast and retinogeniculate signal processing. Frontiers in Neural Circuits 10, 110.Google Scholar
Rathbun, D.L., Alitto, H.J. & Weyand, T.T.G. (2007). Interspike interval analysis of retinal ganglion cell receptive fields. Journal of Neurophysiology 98, 911919.Google Scholar
Rathbun, D.L., Warland, D.K. & Usrey, W.M. (2010). Spike timing and information transmission at retinogeniculate synapses. The Journal of Neuroscience 30, 1355813566.CrossRefGoogle ScholarPubMed
Rimmele, T.S. & Rosenberg, P.A. (2016). Neurochemistry international GLT-1: The elusive presynaptic glutamate transporter. Neurochemistry International 98, 1928.Google Scholar
Robson, J.A. (1981). Abnormal axonal growth in the dorsal lateral geniculate nucleus of the cat. The Journal of Comparative Neurology 195, 453476.Google Scholar
Robson, J.A. (1993). Qualitative and quantitative analyses of the patterns of retinal input to neurons in the dorsal lateral geniculate nucleus of the cat. The Journal of Comparative Neurology 334, 324336.Google Scholar
Robson, J.A. & Mason, C.A. (1979). The synaptic organization of terminals traced from individual labeled retino-geniculate axons in the cat. Neuroscience 4, 99111.Google Scholar
Rompani, S.B., Mullner, F.E., Wanner, A., Roth, C.N., Yonehara, K., Zhang, C. & Roska, B. (2017). Different modes of visual integration in the lateral geniculate nucleus revealed by single-cell-initiated transsynaptic tracing. Neuron 93, 767776.Google Scholar
Rowe, M.H. & Fischer, Q. (2001). Dynamic properties of retino-geniculate synapses in the cat. Visual Neuroscience 18, 219231.Google Scholar
Rozov, A., Zilberter, Y., Wollmuth, L.P. & Burnashev, N. (1998). Facilitation of currents through rat Ca2+-permeable AMPA receptor channels by activity-dependent relief from polyamine block. Journal of Physiology 511, 361377.Google Scholar
Sanderson, K.J., Bishop, P. & Darian-Smith, I. (1971). The properties of the binocular receptive fields of lateral geniculate neurons. Experimental Brain Research 13, 178207.Google Scholar
Schafer, D.P. & Stevens, B. (2010). Synapse elimination during development and disease: Immune molecules take centre stage. Biochemical Society Transactions 38, 476481.Google Scholar
Scholl, B., Tan, A.Y.Y., Corey, J. & Priebe, N.J. (2013). Emergence of orientation selectivity in the Mammalian visual pathway. The Journal of Neuroscience 33, 1061610624.Google Scholar
Seabrook, T.A., Eleftheriou, C.G., Krahe, T.E., Fox, M.A. & Guido, W. (2013). Retinal input regulates the timing of corticogeniculate innervation. The Journal of Neuroscience 33, 1008510097.Google Scholar
Seeburg, D.P., Liu, X. & Chen, C. (2004). Frequency-dependent modulation of retinogeniculate transmission by serotonin. The Journal of Neuroscience 24, 1095010962.Google Scholar
Sengpiel, F. & Kind, P.C. (2002). The role of activity in development of the visual system. Current Biology 12, 818826.Google Scholar
Shatz, C.J. (2009). MHC class I: An unexpected role in neuronal plasticity. Neuron 64, 4045.Google Scholar
Shatz, C.J. & Kirkwood, P.A. (1984). Prenatal development of functional connections in the cat’s retinogeniculate pathway. The Journal of Neuroscience 4, 13781397.Google Scholar
Sherman, S.M. (2004). Interneurons and triadic circuitry of the thalamus. Trends in Neurosciences 27, 670675.Google Scholar
Sherman, S.M. (2005). Thalamic relays and cortical functioning. Progress in Brain Research 149, 107126.Google Scholar
Sherman, S.M. (2007). The thalamus is more than just a relay. Current Opinion in Neurobiology 17, 417422.Google Scholar
Sherman, S.M. (2016). Thalamus plays a central role in ongoing cortical functioning. Nature Neuroscience 16, 533541.Google Scholar
Sherman, S.M. & Guillery, R. (1996). Functional organization of thalamocortical relays. Journal of Neurophysiology 76, 13671395.Google Scholar
Sherman, S.M. & Guillery, R. (2002). The role of the thalamus in the flow of information to the cortex. Philosophical Transactions of the Royal Society of London Series B. Biological Sciences 357, 16951708.Google Scholar
Sherman, S.M. & Guillery, R.W. (2001). Exploring the Thalamus. Elsevier.Google Scholar
Shou, T.D. & Leventhal, A.G. (1989). Organized arrangement of orientation-sensitive relay cells in the cat’s dorsal lateral geniculate nucleus. Journal of Neuroscience 9, 42874302.Google Scholar
Sillito, A.M., Murphy, P.C., Salt, T.E. & Moody, C.I. (1990). Dependence of retinogeniculate transmission in cat on NMDA receptors. Journal of Neurophysiology 63, 347355.Google Scholar
Sincich, L.C., Adams, D.L., Economides, J.R. & Horton, J.C. (2007). Transmission of spike trains at the retinogeniculate synapse. The Journal of Neuroscience 27, 26832692.Google Scholar
Sincich, L.C., Horton, J.C. & Sharpee, T. (2009). Preserving information in neural transmission. Journal of Neuroscience 29, 62076216.Google Scholar
Smith, E.L., Chino, Y., Ridder, W.H., Kitagawa, K. & Langston, A. (1990). Orientation bias of neurons in the lateral geniculate nucleus of macaque monkeys. Visual Neuroscience 5, 525545.Google Scholar
Soodak, R.E., Shapley, R.M. & Kaplan, E. (1987). Linear mechanism of orientation tuning in the retina and lateral geniculate nucleus of the cat. Journal of Neurophysiology 58, 267275.Google Scholar
Soto, D., Coombs, I.D., Kelly, L., Farrant, M. & Cull-Candy, S.G. (2007). Stargazin attenuates intracellular polyamine block of calcium-permeable AMPA receptors. Nature Neuroscience 10, 12601267.Google Scholar
Sretavan, D.W. & Shatz, C.J. (1984). Prenatal development of individual retinogeniculate axons during the period of segregation. Nature 308, 845848.Google Scholar
Sretavan, D.W. & Shatz, C.J. (1986). Prenatal development of retinal ganglion cell axons: Segregation into eye-specific layers within the cat’s lateral geniculate nucleus. The Journal of Neuroscience 6, 234251.Google Scholar
Sriram, B., Meier, P.M. & Reinagel, P. (2016). Temporal and spatial tuning of dorsal lateral geniculate nucleus neurons in unanesthetized rats. Journal of Neurophysiology 115, 26582671.Google Scholar
Stafford, B.K. & Huberman, A.D. (2017). Signal integration in thalamus: Labeled lines go cross-eyed and blurry. Neuron 93, 717720.CrossRefGoogle ScholarPubMed
Stephan, A.H., Barres, B.A. & Stevens, B. (2012). The complement system: An unexpected role in synaptic pruning during development and disease. Annual Review of Neuroscience 35, 369389.Google Scholar
Steriade, M., Jones, E.G. & McCormick, D.A. (1997). Thalamus. Oxford: Elsevier.Google Scholar
Stoelzel, C.R., Huff, J.M., Bereshpolova, Y., Alonso, J.M., Swadlow, H.A., Zhuang, J., Hei, X., Alonso, J.M. & Swadlow, H.A. (2015). Hour-long adaptation in the awake early visual system. Journal of Neurophysiology 114, 11721182.Google Scholar
Storchi, R., Milosavljevic, N., Eleftheriou, C.G., Martial, F.P., Orlowska-Feuer, P., Bedford, R.A., Brown, T.M., Montemurro, M.A., Petersen, R.S. & Lucas, R.J. (2015). Melanopsin-driven increases in maintained activity enhance thalamic visual response reliability across a simulated dawn. Proceedings of the National Academy of Sciences 112, E5734E5743.Google Scholar
Straub, C. & Tomita, S. (2012). The regulation of glutamate receptor trafficking and function by TARPs and other transmembrane auxiliary subunits. Current Opinion in Neurobiology 22, 488495.Google Scholar
Sur, M., Esguerra, M., Garraghty, P.E., Kritzer, M.F. & Sherman, S.M. (1987). Morphology of physiologically identified retinogeniculate X- and Y-axons in the cat. Journal of Neurophysiology 58, 132.Google Scholar
Sur, M. & Sherman, S.M. (1982). Retinogeniculate terminations in cats: Morphological differences between X and Y cell axons. Science 218, 389.Google Scholar
Sur, M., Weller, R.E. & Sherman, S.M. (1984). Development of X and Y-cell retinogeniculate terminations in kittens. Nature 310, 246249.Google Scholar
Suresh, V., Ciftcio lu, U.M., Wang, X., Lala, B.M., Ding, K.R., Smith, W.A., Sommer, F.T. & Hirsch, J.A. (2016). Synaptic contributions to receptive field structure and response properties in the rodent lateral geniculate nucleus of the thalamus. Journal of Neuroscience 36, 1094910963.Google Scholar
Tarusawa, E., Matsui, K., Budisantoso, T., Molnár, E., Watanabe, M., Matsui, M., Fukazawa, Y. & Shigemoto, R. (2009). Input-specific intrasynaptic arrangements of ionotropic glutamate receptors and their impact on postsynaptic responses. The Journal of Neuroscience 29, 1289612908.Google Scholar
Taschenberger, H. & von Gersdorff, H. (2000). Fine-tuning an auditory synapse for speed and fidelity: Developmental changes in presynaptic waveform, EPSC kinetics, and synaptic plasticity. The Journal of Neuroscience 20, 91629173.Google Scholar
Taschenberger, H., Leão, R.M., Rowland, K.C., Spirou, G.A. & von Gersdorff, H. (2002). Optimizing synaptic architecture and efficiency for high-frequency transmission. Neuron 36, 11271143.Google Scholar
Tavazoie, S.F. & Reid, R.C. (2000). Diverse receptive fields in the lateral geniculate nucleus during thalamocortical development. Nature Neuroscience 3, 608616.Google Scholar
Thomas, C.G., Tian, H. & Diamond, J.S. (2011). The relative roles of diffusion and uptake in clearing synaptically released glutamate change during early postnatal development. The Journal of Neuroscience 31, 47434754.Google Scholar
Thompson, A., Gribizis, A., Chen, C. & Crair, M.C. (2017). Activity-dependent development of visual receptive fields. Current Opinion in Neurobiology 42, 136143.Google Scholar
Thompson, A.D., Picard, N., Min, L., Fagiolini, M. & Chen, C. (2016). Cortical feedback regulates feedforward retinogeniculate refinement. Neuron 91, 10211033.Google Scholar
Thompson, K.G., Zhou, Y. & Leventhal, A.G. (1994). Direction-sensitive X and Y cells within the a laminae of the cat’s LGNd. Visual Neuroscience 11, 927938.Google Scholar
Tomita, S., Stein, V., Stocker, T.J., Nicoll, R.A. & Bredt, D.S. (2005). Bidirectional synaptic plasticity regulated by phosphorylation of stargazin-like TARPs. Neuron 45, 269277.Google Scholar
Toni, N., Buchs, P.A., Nikonenko, I., Bron, C.R. & Muller, D. (1999). LTP promotes formation of multiple spine synapses between a single axon terminal and a dendrite. Nature 402, 421425.Google Scholar
Tootle, J.S. & Friedlander, M.J. (1989). Postnatal development of the spatial contrast sensitivity X- and Y-cells in the kitten retinogeniculate pathway. The Journal of Neuroscience 9, 13241340.Google Scholar
Traynelis, S.F., Wollmuth, L.P., McBain, C.J., Menniti, F.S., Vance, K.M., Ogden, K.K., Hansen, K.B., Yuan, H., Myers, S.J. & Dingledine, R. (2010). Glutamate receptor ion channels: Structure, regulation, and function. Pharmacological Reviews 62, 405496.Google Scholar
Turner, J.P., Leresche, N., Guyon, A., Soltesz, I. & Crunelli, V. (1994). Sensory input and burst firing output of rat and cat thalamocortical cells: The role of NMDA and non-NMDA receptors. The Journal of Physiology 480, 281295.Google Scholar
Tzingounis, A.V. & Wadiche, J.I. (2007). Glutamate transporters: Confining runaway excitation by shaping synaptic transmission. Nature Reviews Neuroscience 8, 935947.Google Scholar
Usrey, W.M. & Alitto, H.J. (2015). Visual functions of the thalamus. Annual Review of Vision Science 1, 351371.Google Scholar
Usrey, W.M., Reppas, J.B. & Reid, R.C. (1998). Paired-spike interactions and synaptic efficacy of retinal inputs to the thalamus. Nature 395, 384387.Google Scholar
Usrey, W.M., Reppas, J.B. & Reid, R.C. (1999). Specificity and strength of retinogeniculate connections. Journal of Neurophysiology 82, 35273540.Google Scholar
Van Horn, S.C., Erişir, A. & Sherman, S.M. (2000). Relative distribution of synapses in the A-laminae of the lateral geniculate nucleus of the cat. Journal of Comparative Neurology 416, 509520.Google Scholar
Vidyasagar, T.R. & Urbas, J. V. (1982). Orientation sensitivity of cat LGN neurones with and without inputs from visual cortical areas 17 and 18. Experimental Brain Research 46, 157169.Google Scholar
von Krosigk, M., Bal, T. & McCormick, D. (1993). Cellular mechanisms of a synchronized oscillation in the thalamus. Science 261, 361364.Google Scholar
Wang, X., Hirsch, J.A. & Sommer, F.T. (2010). Recoding of sensory information across the retinothalamic synapse. The Journal of Neuroscience 30, 1356713577.Google Scholar
Wang, X., Wei, Y., Vaingankar, V., Wang, Q., Koepsell, K., Sommer, F.T. & Hirsch, J.A. (2007). Feedforward excitation and inhibition evoke dual modes of firing in the cat’s visual thalamus during naturalistic viewing. Neuron 55, 465478.Google Scholar
Weyand, T.G. (2007). Retinogeniculate transmission in wakefulness. Journal of Neurophysiology 98, 769785.Google Scholar
Weyand, T.G. (2016). The multifunctional lateral geniculate nucleus. Reviews in the Neurosciences 27, 135157.Google Scholar
Wiesel, T.N. & Hubel, D.H. (1963). Effects of visual deprivation on morphology and physiology of cells in the cat’s lateral geniculate body. Journal of Neurophysiology 26, 978993.Google Scholar
Wilson, J.R., Friedlander, J.M. & Sherman, S.M. (1984). Fine structural of identified X- and Y-cells morphology in the cat’s lateral geniculate nucleus. Proceedings of the Royal Society of London Series B 221, 411436.Google Scholar
Wilson, P.D., Rowe, M.H. & Stone, J. (1976). Properties of relay cells in cat’s lateral geniculate nucleus: A comparison of W-cells with X- and Y-cells. Journal of Neurophysiology 39, 11931209.Google Scholar
Winfield, D.A., Hiorns, R.W. & Powell, T.P.S. (1980). A quantitative electron-microscopical study of the postnatal development of the lateral geniculate nucleus in normal kittens and in kittens with eyelid suture. Proceedings of the Royal Society 210, 211234.Google Scholar
Wong, R.O. (1999). Retinal waves and visual system development. Annual Review of Neuroscience 22, 2947.Google Scholar
Wong, R.O., Meister, M. & Shatz, C.J. (1993). Transient period of correlated bursting activity during development of the mammalian retina. Neuron 11, 923938.Google Scholar
Yang, Y.C., Hu, C.C., Huang, C.S. & Chou, P.Y. (2014). Thalamic synaptic transmission of sensory information modulated by synergistic interaction of adenosine and serotonin. Journal of Neurochemistry 128, 852863.Google Scholar
Yeh, C-I., Stoelzel, C.R., Weng, C. & Alonso, J.M. (2009). Functional consequences of neuronal divergence within the retinogeniculate pathway. Journal of Neurophysiology 101, 21662185.Google Scholar
Yeow, M.B.L. & Peterson, E.H. (1991). Active zone organization and vesicle content scale with bouton size at a vertebrate central synapse. Journal of Comparative Neurology 307, 475486.Google Scholar
Zaltsman, J.B., Heimel, J.A. & Van Hooser, S.D. (2015). Weak orientation and direction selectivity in lateral geniculate nucleus representing central vision in the gray squirrel Sciurus carolinensis . Journal of Neurophysiology 113, 29872997.Google Scholar
Zeater, N., Cheong, S.K.K., Solomon, S.G.G., Dreher, B. & Martin, P.R. (2015). Binocular visual responses in the primate lateral geniculate nucleus. Current Biology 25, 16.Google Scholar
Zhang, W. & Linden, D.J. (2009). Neuromodulation at single presynaptic boutons of cerebellar parallel fibers is determined by bouton size and basal action potential-evoked Ca transient amplitude. Journal of Neuroscience 29, 1558615594.Google Scholar
Zhao, X., Chen, H., Liu, X. & Cang, J. (2013). Orientation-selective responses in the mouse lateral geniculate nucleus. The Journal of Neuroscience 33, 1275112763.Google Scholar
Ziburkus, J., Dilger, E.K., Lo, F-S. & Guido, W. (2009). LTD and LTP at the developing retinogeniculate synapse. Journal of Neurophysiology 102, 30823090.Google Scholar
Ziburkus, J. & Guido, W. (2006). Loss of binocular responses and reduced retinal convergence during the period of retinogeniculate axon segregation. Journal of Neurophysiology 96, 27752784.Google Scholar
Ziburkus, J., Lo, F-S. & Guido, W. (2003). Nature of inhibitory postsynaptic activity in developing relay cells of the lateral geniculate nucleus. Journal of Neurophysiology 90, 10631070.Google Scholar
Figure 0

Fig. 1. Synaptic structure shapes retinogeniculate transmission. (A) Tracing of an HRP-filled X-RGC arbor in the cat dLGN shows the location and morphology of a single branch (red box) of the X-RGC arbor used for EM reconstruction. This branch of the axon contacts 4 TC neurons out of 40 available neurons in the territory of the arbor. The remainder of the axon was not reconstructed, and likely contacts several other TC neurons. Bottom inset shows the location of the axonal arbor in the context of the cat LGN. Figure modified from Hamos et al. (1987). Unmarked scale bar = 100 μm. (B, C) Reconstructed arbors of single RGC axons showing distribution of presynaptic boutons into dense clusters in the LGN of (B) an adult cat and (C) a p20 mouse. Note the clustering of boutons along the arbor. Image in B is modified from Robson et al. (1993), showing a segment of a RGC axon; Image in C is from Hong et al. (2014), showing a BD-RGC axon. Scales bars are 100 μm. (D) A 3D reconstruction of a TC neuron dendrite and sites of contact between two neighboring RGC boutons from Budisantoso et al. (2012). In the top image, the dendrite and its appendages are depicted in blue, whereas pink and red sites label the postsynaptic densities of the two axons. In the bottom image, the structure of the terminals of two axons has been added. Spillover can occur between these two nearby terminals. (E) Evidence of spillover-mediated responses to the stimulation of a single RGC axon before eye opening. Two different synaptic responses were observed in response to single retinal fiber stimulation. Shown are recordings from TC neurons in whole cell voltage clamp at −70 mV in a dLGN slice in the presence of the NMDAR blocker, 20 µM CPP. On the left is an example of a retinogeniculate AMPAR EPSC with characteristic rapid rise time and decay kinetics (black trace). On the right is an atypical AMPAR EPSC response notable for significantly slower rise time and decay kinetics (black trace). The two types of EPSCs differ in their sensitivity to the low-affinity AMPAR antagonist, γ-DGG. Low affinity antagonists can be used to assess the relative concentration of glutamate in the synaptic cleft (Clements et al., 1992; Diamond & Jahr, 1997). As γ-DGG competes with glutamate for binding to AMPAR, its efficacy of inhibition decreases with increasing glutamate concentration. γ-DGG has only a small effect on the amplitude of the fast EPSC, but dramatically reduces the amplitude of the slow EPSC (overlaid gray traces), consistent with lower peak glutamate concentration in the synaptic cleft of the slow EPSC. Because the EPSCs are evoked by minimal stimulation, the rapid EPSC represents a direct input from a single RGC axon that forms a direct synapse onto the voltage-clamped relay neuron, whereas the slow EPSC corresponds to the activation of a RGC axon that does not directly synapse onto the voltage-clamped neuron. Modified from Hauser et al. (2014). All figures reprinted with permission.

Figure 1

Fig. 2. Contributions of retinogeniculate short-term plasticity. (A) Representative traces of AMPAR and NMDAR mediated currents recorded before eye opening (left) and in a mature mouse (right) in response to the stimulation of the optic tract. Whole-cell voltage clamp recordings were performed with bicuculline to block GABAA-receptor mediated currents. At −70 mV holding potential, AMPARs mediate the fast activating and decaying current. AMPAR and NMDAR currents both contribute to the EPSCs recorded at +40 mV with AMPARs contributing to the rapid rise and the NMDAR currents contributing to the slow decay of the EPSC. The average amplitude of AMPAR currents increases over development. (B) 5-CT-mediated activation of serotonin receptors alters retinogeniculate short-term plasticity. Experiments were performed in retinogeniculate slices from mature mice. Top and bottom traces overlay pairs of retinogeniculate EPSCs evoked with varying ISI before (top) and after (bottom) the application of 5-CT to active 5HT-1 receptors expressed in presynaptic retinogeniculate boutons. Application of 5-CT reduces the amplitude of the first EPSC and relieves short-term depression, increasing the amplitude of the second EPSC preferentially at short interstimulus interval. (C) Physiologically relevant stimulation frequencies preferentially diminish the contribution of AMPARs to relay neuron firing. Current clamp recordings of action potential firing in response to trains of optic tract stimulation in the presence of AMPAR (NBQX) or NMDAR (CPP) antagonists. Holding potential −50 mV. Blockade of AMPARs alters the latency to first spike but only minimally reduces the overall number of spikes. In contrast, blockade of NMDARs abolished EPSC summation toward action potential firing; only the first stimulus evokes an action potential, reflecting the contribution of AMPARs that rapidly desensitize after the first pulse. Therefore, NMDAR currents can sustain action potential generation without AMPAR contribution. Adapted from (A) Chen and Regehr (2000), B) Liu and Chen (2008) and (C) Augustinaite and Heggelund (2007). All figures reprinted with permission.

Figure 2

Fig. 3. Contribution of NMDAR-currents to retinogeniculate transmission over development. (A) NMDAR EPSCs recorded in the presence of the AMPAR blocker, NBQX, at +40 and −55 mV holding potentials in a p10 (left) and a p29 (right) retinogeniculate slice. Normalized traces are shown. Note the acceleration in NMDAR current decay time over development. (B) Example EPSCs recorded in young (top) and mature (bottom) TC neuron in slice before (left) and during (right) the application of NBQX. Holding potential, −55 mV. (C) NMDAR currents contribute more to the total retinogeniculate charge transfer at p9–11 than p26–32; however, even at the mature synapse, NMDARs contribute nearly half of the total charge transfer. Figure adapted from Liu and Chen (2008). All figures reprinted with permission.

Figure 3

Fig. 4. Substrates for retinogeniculate plasticity. (A) Overlaid AMPAR current traces recorded from different holding potentials to assess the current voltage (IV) relationship. Currents in the presence of CPP to block NMDAR currents and with spermine in the internal solution to examine the degree of IV rectification. Calcium-permeable AMPARs exhibit a rectifying IV relationship. Traces were recorded at 20 mV increments from −60 to +60 mV holding potentials. Left-example obtained before eye opening; right, example from a mature slice. From Hauser et al. (2014). (B) Change in the average AMPAR EPSC IV relationship over development. Rectification of IV currents increases significantly from p9–11 to maturity, indicating a gradual increase in the contribution of CP-AMPARs to AMPAR-mediated currents. Modified from Hauser et al. (2014). Red: p9–11; blue: p15–16; black: p27–32. (C) Changes in AMPAR subunit composition in response to visual experience. The effect of visual deprivation from p20 (late-dark rear, LDR) or dark rearing from birth (chronic dark reared, CDR) on the AMPAR EPSC IV relationship. Rectification of AMPAR currents is reduced in LDR but not in chronically dark reared (CDR) mice when compared to normally reared mice (light rear, LR) mice. P = 0.03. Recordings performed at p27–32. Modified from Louros et al. (2014). (D) Comparison of the distribution of amplitudes of single fiber RGC inputs in juvenile (p27–34) and adult (p60+) mice show the persistence of weak (small-amplitude) inputs with age. Modified from Hong et al. (2014). All figures reprinted with permission.