You are on page 1of 15

Article

A Central Role for Mixed Acetylcholine/GABA


Transmission in Direction Coding in the Retina
Highlights Authors
d ACh initiates responses to motion in natural scenes/low- Santhosh Sethuramanujam,
contrast stimuli Amanda J. McLaughlin,
Geoffery deRosenroll, Alex Hoggarth,
d Non-linear ACh-NMDA interactions amplify low-contrast David J. Schwab,
responses Gautam B. Awatramani
d Spatiotemporally distinct profiles of GABA and ACh underlie
low-contrast DS
Correspondence
gautam@uvic.ca
d Optogenetic stimulation of the SAC network in isolation
drives direction selectivity In Brief
Sethuramanujam et al. demonstrate that
the network of retinal GABAergic/
cholinergic starburst amacrine cells can
differentially transmit excitatory and
inhibitory signals and stimulate
directional responses in downstream
ganglion cells. These results demonstrate
for the first time a physiological role for
mixed transmission in the directional-
selective retinal circuit.

Sethuramanujam et al., 2016, Neuron 90, 1243–1256


June 15, 2016 ª 2016 Elsevier Inc.
http://dx.doi.org/10.1016/j.neuron.2016.04.041
Neuron

Article

A Central Role for Mixed Acetylcholine/GABA


Transmission in Direction Coding in the Retina
Santhosh Sethuramanujam,1 Amanda J. McLaughlin,1 Geoffery deRosenroll,1 Alex Hoggarth,1 David J. Schwab,2
and Gautam B. Awatramani1,*
1Department of Biology, University of Victoria, Victoria, BC V8W 3N5, Canada
2Department of Physics and Astronomy, Northwestern University, Evanston, IL 60208, USA
*Correspondence: gautam@uvic.ca
http://dx.doi.org/10.1016/j.neuron.2016.04.041

SUMMARY Elegant optical stimulation and paired-recording techniques


have revealed that nicotinic ACh receptor (nAChR)-mediated
A surprisingly large number of neurons throughout inputs to DSGCs are large and reliable and exhibit millisecond-
the brain are endowed with the ability to co-release scale kinetics (Lee et al., 2010; Yonehara et al., 2011). Pharma-
both a fast excitatory and inhibitory transmitter. cological studies investigating the role of ACh in mediating the
The computational benefits of dual transmitter DSGC’s light response, however, have provided inconsistent re-
release, however, remain poorly understood. Here, sults. Some studies suggest that ACh may contribute up to 50%
of the DSGC’s light response (Ariel and Daw, 1982; Grzywacz
we address the role of co-transmission of acetylcho-
et al., 1998; Kittila and Massey, 1997; Lipin et al., 2015; Weng
line (ACh) and GABA from starburst amacrine cells
et al., 2005), while others indicate a weak or negligible contribu-
(SACs) to direction-selective ganglion cells (DSGCs). tion of these receptors (Fried et al., 2005; Lee et al., 2010; Park
Using a combination of pharmacology, optoge- et al., 2014). Regardless of the extent of nAChR contribution,
netics, and linear regression methods, we estimated blocking these receptors does not appear to compromise the
the spatiotemporal profiles of GABA, ACh, and gluta- ability of DSGCs to compute direction. To better understand
mate receptor-mediated synaptic activity in DSGCs the role of nAChRs in mediating directional selectivity of DSGCs,
evoked by motion. We found that ACh initiates re- here we used pharmacology and a complementary linear regres-
sponses to motion in natural scenes or under low- sion method (Manookin et al., 2010; Taylor and Vaney, 2002) to
contrast conditions. In contrast, classical glutama- examine the contribution of nAChRs during physiologically intact
tergic pathways play a secondary role, amplifying conditions. Regression analysis was based on the distinct bio-
physical signatures of the individual synaptic receptors medi-
cholinergic responses via NMDA receptor activation.
ating the DSGC’s light response and provided for the first
Furthermore, under these conditions, the network
time a simultaneous estimate of the spatiotemporal profiles of
of SACs differentially transmits ACh and GABA to GABA, ACh, and glutamate signaling at dendrites of DSGCs.
DSGCs in a directional manner. Thus, mixed trans- Interestingly, we found that during natural/low-contrast stimu-
mission plays a central role in shaping directional lation, SACs serve as the primary source for both excitation and
responses of DSGCs. inhibition to DSGCs. Under these conditions, glutamatergic exci-
tation played only a secondary role in amplifying the response to
INTRODUCTION cholinergic input, acting mainly on NMDA receptors (NMDARs)
and not AMPA receptors (AMPARs). In response to moving stim-
In the mammalian retina, direction-selective (DS) responses are uli, SAC GABA and ACh were differentially transmitted to DSGCs
first observed in the radiating dendrites of GABAergic/cholin- in a manner that drives directional spiking responses (Figure 1B).
ergic starburst amacrine cells (SACs; Euler et al., 2002; Yonehara Using optogenetics, we directly demonstrated that the SAC
et al., 2013). SAC dendrites transmit information to postsynaptic network alone can trigger directional responses in DSGCs in
direction-selective ganglion cells (DSGCs) and endow them with the absence of glutamate input. Together, these results lead
the ability to code direction (Vlasits et al., 2014; Yoshida et al., us to posit a novel intermediary role for the SAC network in
2001). A specific asymmetric dendro-dendritic wiring connecting conveying directional information to ganglion cells, relying pre-
SACs to DSGCs allows SACs to mediate a ‘‘null’’ GABAergic in- dominantly on mixed cholinergic/GABAergic transmission.
hibition (Briggman et al., 2011; reviewed by Borst and Euler,
2011; Demb, 2007; Vaney et al., 2012). DSGCs are thought to RESULTS
compute direction by integrating asymmetric SAC inhibition
with a non-directional bipolar cell glutamatergic excitation Directional Information to ON-OFF DSGCs Is Routed
(Yonehara et al., 2013; Park et al., 2014; Figure 1A). In addition, through the SAC Network
SACs also release acetylcholine (ACh) (Massey and Neal, The first piece of evidence that SACs (rather than bipolar cells)
1979; Massey and Redburn, 1983), but the role of cholinergic initiate activity in DSGCs came from examining the effects of
excitation is not well understood. blocking nAChRs on DSGC responses to a natural movie

Neuron 90, 1243–1256, June 15, 2016 ª 2016 Elsevier Inc. 1243
Figure 1. Cholinergic Inputs Play a Central
Role in Initiating Responses in DSGCs
during Natural/Low-Contrast Stimulation
Conditions
(A) A schematic of the standard model for direction
selectivity in which DSGCs compute direction
by comparing glutamatergic excitation mediated
by bipolar cells (Glu; red) with GABAergic inhibition
from null-side SACs (null SAC shown in green).
Arrow indicates the preferred direction of the
DSGC.
(B) Diagram of the proposed model in which the
SAC network generates direction selectivity. SAC
dendrites are optimally stimulated by centrifugal
motion (soma-to-dendrite; Euler et al., 2002)
and therefore preferred-side SACs (blue) drive
preferred-direction cholinergic excitation, and null-
side SACs (green) provide null-direction GABAergic
inhibition (and also ACh). As only null-side SACs
make strong synaptic connections with DSGCs
(Briggman et al., 2011), ACh transmission may
occur through paracrine mechanisms (dotted blue
arrows), while GABAergic inhibition acts via con-
ventional synaptic mechanisms (solid green arrow).
The secondary contribution of bipolar cells medi-
ated by NMDARs has not been shown for simplicity.
(C) Spike raster plot in response to the repeated
presentation of a 5 s natural movie clip (100 trials,
presented at 0.1 Hz; a uniform gray background
was presented during the intervals). The spiking
in response to natural stimuli were DS (see Fig-
ure S1). The gray trace (top) shows the changes in
light intensity over the DSGC receptive field. Trials
in the presence of an nAChR antagonist (100 mM
HEX) are indicated in red. The instantaneous
spike rates (estimated through convolution with a
Gaussian kernel, s = 25 ms) averaged over seven
trials are shown below for control (black) and in
nAChR antagonists (red).
(D) The average total number of spikes generated/
trial under different conditions (n = 7; mean ± SEM is
plotted; *p < 0. 001) demonstrates that the potent
block by nAChR antagonists is completely reversible.
(E) The polar plot illustrates the peak ON spike rate evoked in eight directions (three trials) in control (black) and in the presence of an nAChR antagonist (red). The
DSI is indicated by the solid radial line. Example spike trains for both preferred (0 ) and null (180 ) motion in control and in the presence of nAChR antagonist are
illustrated around the polar plot.
(F) Similar to (E) but using low-contrast stimuli (C50).
(G) Plot of the average peak ON spike rate for responses evoked in the preferred direction under conditions shown in (E) and (F) (n = 8; mean ± SEM is plotted;
*p < 0. 001). For extended analysis of DS properties and quantification of the OFF spike rates, see Figure S2.

(Figures 1C and 1D). Natural movies were obtained using a CCD 1982; Fried et al., 2005; Grzywacz et al., 1998; Kittila and Mas-
camera mounted on a mouse’s head while it freely roamed in its sey, 1997; Lee et al., 2010; Lipin et al., 2015; Park et al., 2014;
cage (Froudarakis et al., 2014). Spike recordings were made Weng et al., 2005), possibly due to the differences in the visual
from ON-OFF DSGCs in an isolated whole-mount retina, onto stimuli used. One major difference was that the equivalent
which this movie was projected. DSGCs responded to certain contrast of the movie used here (as estimated by Tadmor and
moving features in the movie (Figure 1C), and rotating the Tolhurst, 2000) is significantly lower than that of high-contrast
movie revealed that indeed these responses were DS (Figure S1, moving bars, spots, or gratings that have been typically used
available online). Interestingly, the application of an nAChR in previous studies. Hence, we next tested the role of nAChRs
antagonist (100 mM hexamethonium; HEX; or 50 mM curare) in mediating responses to moving spot stimuli of varying
strongly and reversibly suppressed DSGC responses, indicating contrast.
a critical role for cholinergic excitation in generating responses Indeed, saturating ‘‘high’’ contrast spots (C100) moving on
to motion in the movie (Figures 1C and 1D). This strong suppres- a featureless background evoked stronger spiking responses
sion of DSGC responses contrasts with the weaker effect of in DSGCs compared to natural stimulation (the maximum peak
nAChR antagonists reported in the literature (Ariel and Daw, firing rate was 206 ± 14 Hz, n = 8, compared to 95 ± 10 Hz,

1244 Neuron 90, 1243–1256, June 15, 2016


Figure 2. AMPAR- and nAChR-Mediated
Synaptic Inputs to DSGCs Are Activated at
Distinct Contrasts
(A) EPSCs measured in a voltage-clamped DSGC
mediated by AMPARs and nAChRs in control
conditions (top) and in the presence of an nAChR
antagonist (AMPAR only; middle). The contribution
of nAChRs (bottom) was estimated by taking the
difference (control  nAChR antagonists).
(B) The peak EPSC amplitude plotted as a function
of contrast for the different conditions shown in
(A) (n = 7 DSGCs; mean ± SEM is plotted) was fit
by the Naka-Rushton equation (see Experimental
Procedures). The vertical dashed line indicates
the contrast required to evoke a half-maximal
response (C50) measured in control conditions.
Responses were normalized to the peak amplitude
measured in control.
(C) A plot of the fractional nAChR (blue)- and
AMPAR-mediated currents (red) as a function of
contrast.

n = 7 under natural stimulation conditions). Consistent with decreased responses to stimuli below C50, but only reduced
the previous literature, we found that the application of maximal responses by 49% ± 9% (n = 7, p < 0.005; Figure 2B).
nAChR antagonists reduced the DSGC peak spike rate to In addition, the C50 was significantly larger in the presence of an-
the leading edge of full contrast moving spots (250 mm in diam- tagonists (C50 = 24% ± 3% in control; C50 = 41% ± 3% in the
eter, 1 mm/s; Figures 1E and 1G) but did not compromise its presence of nAChR antagonist; n = 7; p < 0.005; Figure 2B). This
ability to code direction (Figures 1E and S2). Similar effects suggested that the direct AMPAR-mediated inputs from bipolar
were observed on the trailing edge response (OFF response), cells contribute less to responses evoked by low- compared to
and measuring spike rate or total spike number did not signifi- high-contrast stimuli. Conversely, when the contrast response
cantly impact our results (Figure S2). function of the cholinergic input to DSGCs was estimated by
When ‘‘low’’ contrast spots that produced 50% of the maximal subtracting the responses measured in nAChR antagonists
response (C50; 110 ± 12 Hz, n = 8) were used, application of from those measured under control conditions (Figure 2A), it
nAChR antagonists completely blocked the spiking response became apparent that cholinergic inputs dominate in the low-
(Figures 1F and 1G). This suggested that cholinergic inputs play contrast regime, accounting for >80% of the total excitatory in-
a central role in mediating DSGC responses in this regime. puts for contrast levels below C50 (Figure 2C).
Thus, it appears that simple low-contrast stimuli are similar to Previous studies have shown that voltage-dependent
natural scenes in their sensitivity to cholinergic antagonists. For NMDARs play a major role in mediating the DSGC’s light re-
simplicity, in the rest of this study we utilized low-contrast moving sponses (Kittila and Massey, 1997; Lee et al., 2010; Poleg-Pol-
stimuli to probe the properties of the DS circuit. Importantly, low- sky and Diamond, 2016), raising the possibility that glutamate
contrast stimuli drive a reliable DS response in DSGCs, indicating signals were mediated by these receptors in the low-contrast
that cholinergic/GABAergic inputs are balanced in such a way regime. To test this hypothesis, synaptic responses evoked
that DS responses are maintained, despite being weakly stimu- by low-contrast stimuli (C50) were measured at different holding
lated (Figures 1F and S2). potentials in the presence of an nAChR antagonist (Figure 3A).
Indeed, synaptic responses measured under these conditions
The Role of nAChRs, AMPARs, and NMDARs in Contrast were highly non-linear, as depicted in the current-voltage (I-V)
Encoding plot (Figure 3C), indicating the involvement of NMDARs. Subse-
To better understand the organization of excitatory inputs to quent addition of an NMDAR antagonist decreased the current
DSGCs, we used voltage-clamp methods to assess the roles amplitudes and linearized the I-V, shifting the reversal potential
of AMPARs, nAChRs (Figure 2), and NMDARs (Figure 3). Surpris- close to 40 mV (Figures 3B and 3D). This indicated that
ingly, we found that these receptors were not activated in parallel the residual component was made up of weak GABAA recep-
when stimulus contrast was varied, giving rise to several unex- tor (GABAR)- and AMPAR-mediated conductances (GABAR
pected consequences, as described below. 2.1 nS; AMPAR 0.6 nS). In contrast, the I-V of the currents
First, we recorded contrast response functions for AMPAR/ blocked by D-AP5 was ‘‘J’’ shaped, characteristic of NMDARs.
nAChR-mediated excitatory postsynaptic currents (EPSCs) in The maximal conductance (12.5 nS at +40 mV; Figure 3E) was
DSGCs (Figures 2A and 2B; VHOLD 60 mV, near ECl; large compared to other ganglion cell types (Buldyrev et al.,
D-AP5 was included in the bath to block NMDARs), which 2012; Manookin et al., 2010). Despite the large NMDA conduc-
were fit using the Naka-Rushton equation (Naka and Rushton, tance, the absence of any spiking activity in the same condi-
1966; Figure 2B). Applying an nAChR antagonist substantially tions in which NMDAR currents were measured indicates that

Neuron 90, 1243–1256, June 15, 2016 1245


Figure 3. NMDAR-Mediated Activity at Low
Contrast Relies on SAC ACh
(A and B) An example of a DSGC in which a half-
maximal contrast response is blocked by HEX
(inset). In the continued presence of HEX, voltage
clamping the same cell reveals the presence of
‘‘silent’’ synaptic conductances (A), which were
partially blocked by the NMDAR antagonist, D-AP5
(50 mM; B).
(C–E) The mean current measured in the boxed
region in (A) and (B) are plotted against holding
potential in (C) and (D), respectively. The I-V rela-
tionship of responses blocked by D-AP5 is shown
in (E), had the characteristics of NMDAR currents,
and could be fit by a sigmoidal function (gray
dashed line). The I-Vs were constructed from re-
sponses averaged over five cells (mean ± SEM is
plotted).
(F) Spiking responses of a DSGC to preferred and
null direction for stimuli eliciting half-maximal (C50)
and maximal (C100) responses, recorded in control
(black) and in the presence of the NMDAR antag-
onist (gray).
(G) Plots of the average peak spike rate measured
in DSGCs evoked by high- and low-contrast stimuli
in control or in the presence of D-AP5 (n = 8;
mean ± SEM is plotted; *p < 0.001).

to be critical for the generation of direc-


tion selectivity (Figure 8).

Relative Contribution of nACh,


GABA, and AMPARs to Preferred
Motion Revealed by Linear
Regression Analysis
So far, our pharmacological analysis sug-
gests that nAChRs mediate the dominant
NMDARs on their own do not drive spikes, i.e., they are excitatory inputs at low contrast and are therefore necessary to
‘‘silent.’’ generate directional responses observed under these condi-
In the context of nAChR-mediated activity, however, we tions. However, the potential for indirect presynaptic network
found that NMDARs play an important role in amplifying low- (Dmitrieva et al., 2007; Elgueta et al., 2015; Strang et al., 2015)
contrast responses (Figures 3F and 3G). At C50, responses and/or off-target effects of nAChR antagonists make it difficult
were reduced (20% to 80%; average suppression 55% ± 8% to unambiguously interpret these results. To estimate the relative
of control, n = 8; Figure 3F), in contrast to the nAChRs, which contribution of AMPARs, GABARs, and nAChRs to the DSGC’s
completely blocked responses at this contrast level (Figure 1G). synaptic response without application of antagonists for these
The remaining spikes observed in the presence of D-AP5 at receptors (which could disturb the dynamics of the circuit), we
C50 were completely blocked by the additional application of developed a linear regression model. The only antagonist
an nAChR antagonist, confirming that they were mediated by included, to simplify this experiment, was the well-characterized
ACh alone (data not shown). The finding that both nAChR NMDAR antagonist D-AP5, which is not critical for DS computa-
and NMDAR antagonists strongly affect the response to low- tion (Figure 8). The linear regression analysis relies on the distinct
contrast stimuli suggests a significant non-linear interaction ‘‘basis functions’’ of individual receptors, which is described by
between these receptors. Furthermore, as nAChRs, but not their I-V relationships (Figure 4). The composite synaptic light-
NMDARs, could independently drive spikes in DSGCs, it sug- evoked currents measured over a range of holding potentials
gests that SAC release of ACh initiates responses in DSGCs, were modeled as a linear weighted sum of the individual receptor
while bipolar cell glutamate augments these responses via components (nAChRs, GABARs, and AMPARs).
NMDARs. Thus, in DSGCs, nAChRs appear to usurp the role To measure the voltage-dependent properties of synaptic
of functionally absent AMPARs in recruiting NMDARs, consti- nAChRs and GABARs in DSGCs, we used an optogenetic strat-
tuting a novel mechanism whereby SACs can control excitation egy to stimulate SACs and measured monosynaptic responses
to DSGCs. However, as the effects mediated by NMDARs in DSGCs in pharmacological isolation. In transgenic mice in
appear secondary to cholinergic excitation, they are unlikely which ChR2 is selectively expressed in SACs (Chat-Cre::Ai32;

1246 Neuron 90, 1243–1256, June 15, 2016


Figure 4. Multiple Types of Synaptic Re-
ceptors Expressed by DSGCs Can Be
Distinguished by Their Voltage-Dependent
Characteristics
(A) A transgenic mouse line with ChR2 expression
specific to SACs (Figure S3) was used to directly
stimulate nAChR currents measured at a series of
holding potentials (indicated on right). The stimulus
was a 100 ms intense blue flash. Photoreceptors
and GABARs were pharmacologically blocked in
this experiment.
(B) Normalized peak current responses were
plotted against holding potential (mean ± SEM is
plotted). The I-V was fit by the Woodhull equation
(solid blue line; see Experimental Procedures).
(C) Similar to (A), SACs expressing ChR2 were
stimulated in conditions that isolated GABAR-
mediated activity.
(D) The GABA I-V was fit by an exponential function
(solid green line).
(E) AMPAR currents were pharmacologically iso-
lated in the presence of a cocktail of NMDAR,
GABAR, and nAChR antagonists. These responses
were driven by photoreceptors.
(F) The AMPA I-V was well described by a linear fit
(solid red line).

AMPAR-mediated currents were linear


(Figures 4E and 4F). Thus, each of the syn-
aptic receptors functionally expressed
by DSGCs has a distinct biophysical
signature, making it feasible to distinguish
their relative contributions to light-evoked
responses under physiological conditions.
To estimate the relative contributions
of the different receptors to the total
response evoked by low-contrast mov-
ing stimuli, preferred direction responses
were measured in DSGCs over a range
of holding potentials (Figure 5A). To
SAC-ChR2; Figure S3), cholinergic synaptic currents were ensure adequate voltage clamp, the superior coding population
measured in a cocktail of glutamate agonists/antagonists that of electrically coupled DSGCs (Trenholm et al., 2013a) was
is commonly used to block photoreceptor-driven responses excluded. The composite light-evoked synaptic I-V relationship
(50 mM D-AP5 + 50 mM L-AP4 + 10 mM CNQX), as well as was modeled by a weighted linear sum of the basis functions
the added presence of a GABAA receptor antagonist (10 mM for AMPARs, GABARs, and nAChRs, which yielded an estimate
SR-95531). nAChR-mediated currents were inwardly rectifying of a fractional conductance for each receptor type (Figure 5B).
(Haghighi and Cooper, 1998) and were well described by the Interestingly, examining the I-V relationship of the early phase
Woodhull equation (Figures 4A and 4B; note that spermine was of the synaptic response revealed inwardly rectifying currents,
included in the intracellular solution to maintain rectification; see indicating they were largely mediated by nAChRs (compare Fig-
Experimental Procedures). Similarly, SAC-ChR2 was used to ure 5B, top, and Figure 4B). A dominant cholinergic component
stimulate GABAR-mediated synaptic responses in a cocktail in the early phase was consistently observed across six cells
that blocked photoreceptor input as well as nAChRs (100 mM (gACh = 87% ± 6%, gGABA = 5% ± 2%, and gAMPA = 8% ± 5%;
HEX). Isolated GABAR-mediated currents were found to be where gACh, gGABA, and gAMPA indicate the conductance for
slightly outwardly rectifying (Figures 4C and 4D), and their I-V rela- each receptor at 60 mV; Figure 5C). Although the I-V relation-
tionship was fit by a single exponential equation, as previously ship appeared less rectified during the peak synaptic response,
described (Manookin et al., 2010). To measure the AMPAR basis it remained best estimated by a combination of nAChR and
function, photoreceptor-driven responses were measured in GABAR components with minimal contribution of AMPARs (Fig-
DSGCs in the presence of NMDAR, GABAR, and nAChR antago- ures 5B and 5C; gACh = 70% ± 3%, gGABA = 18% ± 3%, and
nists. In contrast to GABAR- and nAChR-mediated responses, gAMPA = 12% ± 5%; n = 6).

Neuron 90, 1243–1256, June 15, 2016 1247


Figure 5. Cholinergic SAC Inputs Dominate DSGC Responses to Low-Contrast Preferred Motion under Near-Physiological Conditions
(A) Photoreceptor-mediated synaptic responses in a DSGC to low-contrast preferred motion shown at four different holding potentials (indicated on the right;
mean ± SEM of three trials is plotted in black and gray, respectively). Responses are measured in the presence of an NMDAR antagonist (50 mM AP5). The inset
shows the active circuit components under these conditions.
(B) The I-V relationship computed from the current averaged over the two boxed regions in (A) (open circles, top; closed circles, bottom). The average current was
modeled by a weighted linear sum of the basis functions of each receptor as estimated in Figure 4 (black solid line; see Experimental Procedures). The computed
conductances (gAch, gGABA, and gAMPA) at 60 mV are indicated in each graph. Responses were dominated by nAChRs (blue lines).
(C) The average I-V relationship of the early and peak responses (open circles, top; closed circles, bottom) measured in six cells (mean ± SEM). The average
estimated contributions of the ACh, GABA, and AMPARs are indicated in blue, green, and red, respectively.

To validate the linear regression analysis, we compared the that linear regression provides a reliable estimate of the relative
performance of three different models (ACh + AMPA + GABA, contribution of different synaptic receptors underlying DSGC re-
ACh + GABA, or AMPA + GABA) under different experimental sponses to preferred direction motion.
conditions that systematically biased the contribution of the indi-
vidual receptor components. For low-contrast responses, the Spatiotemporal Profiles of ACh and GABA Underlying
errors associated with the fitting procedure (root mean square Low-Contrast Direction Selectivity
error; RMS error) are similar for the three-component model Previous studies have demonstrated SAC-mediated inhibition,
and the ACh + GABA model. The AMPA + GABA model, how- but not excitation, to be spatially asymmetric (Lee et al., 2010;
ever, performed relatively poorly. Together, these results confirm Yonehara et al., 2011). However, these studies relied on artificial
the lack of detectable AMPAR-mediated glutamatergic input optical or electrical methods to stimulate SACs, not capturing
in this regime (Figures 6A–6E). On the other hand, responses their dynamic behavior in the context of the network. Further-
evoked by high-contrast stimuli were best fit with the three- more, these experiments were performed in conditions in which
component model, consistent with the idea that nAChRs, network activity was pharmacologically blocked, casting doubt
AMPARs, and GABARs contribute to the synaptic responses on the physiological relevance of ACh (Taylor and Smith,
(Figures 6F–6J). Both of the two-receptor component models 2012). Thus, the temporal dynamics of ACh and GABA evoked
performed poorly in the high-contrast regime. Furthermore, by moving stimuli remain to be investigated. As we found basis
AMPARs and nAChRs appear to contribute equally to the synap- functions for the nAChRs, GABARs, and AMPARs did not signif-
tic response (Figure 6G), consistent with the pharmacology (Fig- icantly change over the duration of the synaptic response (Fig-
ure 2B). Finally, when we measured responses to high-contrast ure S4), we could reiterate the linear regression procedure
stimuli in the presence of nAChR antagonists, we found throughout the entire light response at 1 ms intervals (Figure 7B).
AMPA + GABA were sufficient to model the synaptic responses This provided a direct estimate of the temporal dynamics of
(Figures 6K–6O). The ACh component usually observed in con- nAChR-, GABAR-, and AMPAR-mediated synaptic responses
trol conditions was completely abolished by nAChR antagonists evoked by moving stimuli (Figure 7A; see Figure S5 for compar-
(n = 6; p < 0. 001, t test). Under these conditions, adding an ACh isons of data to model predictions).
component did not improve the fit, and the ACh + GABA model As noted previously, the contribution of AMPARs under these
provided a poor estimate of the response. Thus, we conclude conditions was minimal (Figures 7B and 7C). Importantly, under

1248 Neuron 90, 1243–1256, June 15, 2016


Figure 6. Experimentally Varying nAChR or AMPAR Activity Verifies the Fidelity of Multicomponent Linear Regression Models
(A) A schematic of the circuit involved under low-contrast motion where direction selectivity is driven by ACh and GABA (the contributions of NMDARs are omitted
for clarity; all conductance measurements were performed in the presence of an NMDAR antagonist [50 mM AP5]).
(B–D) The I-V of the peak phase of preferred direction response (mean ± SEM is plotted) for low-contrast motion shown in Figure 5C (bottom) was fit to a three-
component linear model (ACh + AMPA + GABA, B) or to two receptor component models (ACh + GABA,C; or AMPA + GABA, D). The residuals between the data
and their respective fits are shown below, shaded in gray. Similar results were obtained when the error analysis was performed for the duration of the light
response (Figure S4).
(E) A comparison of the average root mean square (RMS) error across six cells is plotted for the three models. Note that the AMPA + GABA model performs poorly
compared to the other two models (*p < 0. 001).
(F–J) Similar to (A)–(E), except high-contrast stimuli are used to increase the contribution of AMPARs (see Figure 2B). Under these conditions (F), the three-
component model (G) performs significantly better than the two-component models (H and I), as shown in (J) (*p < 0. 01; n = 5).
(K–O) Similar to (F)–(J) but with nAChRs blocked pharmacologically, as depicted in (K). Under these conditions, the ACh + GABA model (M) performs relatively
poorly compared to the ACh + GABA + AMPA (L) or AMPA + GABA models (N), as summarized in (O) (n = 5; *p < 0. 001).

low-contrast stimulation conditions, ACh and GABA signals ity of 1 mm/s). Interestingly, as inhibition increased, cholinergic
differed in several respects during motion in the preferred or excitation ceased, consistent with the idea that cholinergic
null directions. First, during preferred motion, excitation arose release was strongly controlled by GABAergic inhibition (Fried
a few milliseconds before inhibition (40 ± 13 ms delay between et al., 2005). The temporal delay between excitation and inhibi-
peak gACh and gGABA; n = 6 DSGCs; Figures 7B and 7F), which tion results in a net excitation in the preferred direction in a similar
corresponds to a spatial lag of 40 mm (given the stimulus veloc- time window in which spiking was observed in DSGCs (Figures

Neuron 90, 1243–1256, June 15, 2016 1249


Figure 7. Differential Transmission of ACh and GABA Mediates Direction Selectivity in Ganglion Cells
(A) A schematic illustrating the circuit mechanism involved in conveying low-contrast direction information to DSGCs (photoreceptor and bipolar cells driving
SACs are not shown for clarity), as described in Figure 1B.
(B) Spiking responses in a DSGC evoked by low-contrast stimuli moving in the preferred (P) or null (N) directions (as indicated; top panel). The temporal evolution
of GABA (green traces)-, ACh (blue traces)-, and AMPAR-mediated conductances (red) estimated at 60 mV in the same DSGCs using linear regression methods
(see text for details). Subtracting the GABA from the ACh conductance (ACh  GABA, bottom traces) provides an indication of net excitation (yellow box) or
inhibition.
(C–G) A comparison of the synaptic response properties evoked by low-contrast stimuli across six DSGCs. Plots of the peak ACh and AMPA conductances
evoked by null versus preferred stimuli (C), null versus preferred peak GABA conductances (D), net excitatory (peak ACh  GABA) conductances evoked by null
versus preferred motion (E), time to peak of the GABA versus ACh conductances evoked by preferred motion (F), and time to peak of the GABA versus ACh
conductance evoked by null motion (G).

7B and 7E). Second, during null motion, GABAergic inhibition terns of activity. However, convincing evidence that SACs alone
was significantly stronger than that observed during preferred could generate direction selectivity through mixed transmission
motion (Taylor and Vaney, 2002; Figures 7B and 7D), while was obtained when we measured responses to moving stimuli
cholinergic excitation was approximately equal in both direc- evoked by either photoreceptors or ChR2-expressing SACs in
tions (Figures 7B and 7C), indicating a differential transmission the same DSGC (Figures 8A–8D). In these experiments, after
of ACh and GABA signals to DSGCs. In the null direction, both characterizing the tuning properties of the DSGCs under normal
GABAergic inhibition and cholinergic excitation had a similar stimulation conditions (Figure 8B), photoreceptor output was
time course (Figures 7B and 7G), but since the GABA conduc- blocked using a standard cocktail of glutamate receptor antag-
tance was significantly larger than the cholinergic conductance, onists/agonist (Figure 8C). The stimulus intensity was then
the net activity was inhibitory (Figure 7B; bottom trace). While increased by 5 log units in order to stimulate ChR2 (Figure 8D).
directional inhibition and spatiotemporal offset inhibition/excita- Remarkably, we found that stimulation of SAC-ChR2 rein-
tion have been previously demonstrated to be key factors in stated the DS spiking responses after photoreceptor blockade
shaping direction selectivity (Taylor and Vaney, 2002), what is (Figures 8B–8D). Importantly, each cell’s preferred direction
surprisingly revealed here is that these features can arise from measured in control conditions remained unchanged when
the SAC network alone under physiological conditions, without measured under SAC-ChR2 stimulation (Figures 8E and 8F),
direct bipolar cell input to DSGCs. and SAC-ChR2-mediated direction selectivity was observed in
DSGCs coding a variety of directions (Figure 8F), indicating
Optogenetic Stimulation of SACs Evokes DS Responses that this mechanism was not specific for a particular DSGC
In general, results from optogenetic experiments are difficult to subtype. The observed direction-selective index (DSI; see
interpret because these stimulation techniques are not usually Experimental Procedures) during SAC-ChR2 stimulation was
designed to drive physiologically relevant spatiotemporal pat- increased compared to control (Figure 8G), likely due to a

1250 Neuron 90, 1243–1256, June 15, 2016


Figure 8. The SAC Network Alone Can Drive Direction Selectivity in Ganglion Cells
(A) A circuit diagram indicating that all synapses were pharmacologically blocked except GABA and ACh SAC to DSGC synapses. In these experiments, SACs
were directly stimulated with ChR2.
(B–D) Directional responses in the SAC-ChR2 retina under control (B) or in the presence of a cocktail of agents that block photoreceptor synapses (C). In the
continued presence of blockers, directional responses are recovered in the DSGC when stimulus intensity was increased by 5 orders of magnitude to drive
SAC-ChR2 (D). Stimulus intensity is indicated as R*/s on the right.
(E) Polar plots of spiking measured in (B) and (D).
(F) The direction encoded by the optogenetically stimulated SAC network (SAC-ChR2) is plotted against the direction encoded during stimulation of intact
photoreceptors (PR; n = 14 DSGCs). The dashed line is a reference line with a slope of 1.
(G) DSI computed under the two stimulation conditions are plotted against each other (for the same set of DSGCs in F).
(H) SAC-ChR2-mediated direction-selective responses in DSGCs (top panel) were abolished in the presence of the GABAR blocker (10 mM SR-95331; n = 5;
middle panel). Responses were completely blocked by the subsequent application of nAChR antagonists (n = 5). Figure S6 shows the temporal kinetics of the
nAChR and GABAR conductances generating this DS response.
(I–K) The mean (± SEM) number of spikes/peak spike rate (I), relative direction (0 indicates the same direction as photoreceptor-drive responses, J), and DSI (K),
evoked in DSGCs by optogenetic stimulation of SACs with bright spots moving at different velocities in the preferred direction.

Neuron 90, 1243–1256, June 15, 2016 1251


threshold effect (the ‘‘iceberg effect’’; Carandini and Ferster, amy et al., 2015) or electrical synapses (Kolb and Famiglietti,
2000) associated with the weaker responses observed in this 1974; Vaney, 1991; Xin and Bloomfield, 1997), the finding
condition (control, 43 ± 6 spikes; SAC-ChR2, 10 ± 1 spikes; that a single population of amacrine cells can provide feature-
n = 14, p < 0. 001). Varying the velocity of the stimulus showed specific information to a ganglion cell through direct excitation
that the ChR2-stimulated responses were optimally evoked and inhibition is unprecedented.
at 1 mm/s (Figures 8I–8K), similar to photoreceptor-mediated
responses (Trenholm et al., 2013b). In contrast, ChR2 stimula- A Novel Synaptic Organization of the Excitatory
tion did not drive responses to slower stimuli (0.4 mm/s), where Synapses in the DS Circuit
photoreceptor responses are apparent (Trenholm et al., 2013b). The synaptic organization of excitatory synapses in the DS cir-
However, whether these tuning properties reflect the intrinsic cuit appears to be specialized in a number of aspects, promoting
properties of SACs or those of ChR2 is not clear. space- and time-dependent non-linear interactions. Most strik-
Finally, blocking GABARs abolished the directionally selective ingly, AMPAR-mediated synaptic activity in DSGCs is absent
spiking, suggesting that cholinergic inputs on their own were not at low contrast and only emerges at the higher-contrast levels.
DS. The subsequent addition (or addition alone; n = 5; data This is unusual, as in most ganglion cells the AMPAR conduc-
not shown) of an nAChR antagonist blocked all spiking activity tance is the dominant excitatory conductance (Diamond and
(Figure 8H), confirming that there was no residual photoreceptor Copenhagen, 1993; Sagdullaev et al., 2006; Sethuramanujam
activity and that DS responses were indeed driven by SACs. Mea- and Slaughter, 2015; Zhang and Diamond, 2009), and AMPARs
surements of SAC-ChR2-evoked synaptic conductances indi- and NMDARs are usually activated in parallel when contrast
cated that direction selectivity was driven by a combination of (1) (Buldyrev et al., 2012; Diamond and Copenhagen, 1995; Manoo-
symmetrical cholinergic excitation, (2) directional GABAergic inhi- kin et al., 2010) or temporal frequency (Stafford et al., 2014)
bition, and (3) directional timing differences between excitation is varied. Whether different types of bipolar cells differentially
and inhibition, all similar to physiological conditions (Figure S6). use NMDARs and AMPARs, or whether these receptors are
A combination of results obtained from natural and artificial stimu- expressed at the same synapses and the system relies on the
lation of the SAC network highlights the importance of this network distinct affinities of these receptors (Patneau and Mayer, 1990)
in the generation of direction selectivity in ganglion cells. to signal different levels of glutamate release evoked by varying
In summary, results from pharmacology, linear regression, and contrasts remains to be investigated.
optogenetic experiments demonstrate that under low-contrast In contrast to the expression of ‘‘silent’’ NMDARs described
conditions (1) cholinergic inputs robustly drive DSGC spiking here, previous studies have demonstrated that bipolar cells can
activity (with the help of NMDARs), and (2) ACh and GABA drive spiking in DSGCs solely using NMDARs (Cohen and Miller,
are differentially transmitted to DSGCs, shaping their DS spiking 1995; Kittila and Massey, 1997; Weng et al., 2005). Similar
properties. Thus, mixed ACh/GABA transmission mediated by NMDAR-mediated spiking has also been observed in other types
the SAC network plays a more important role in conveying DS in- of ganglion cells (Buldyrev et al., 2012; Diamond and Copenha-
formation than previously envisioned. gen, 1993). However, this apparent difference can be easily
reconciled by considering the specific pharmacological condi-
DISCUSSION tions used to isolate NMDAR-mediated activity. Specifically, pre-
vious studies have used AMPAR antagonists to isolate NMDARs,
An Intermediary Stage for Processing Directional which also block feedforward inhibition. Without inhibition, intact
Information NMDAR-mediated activity is likely enhanced (Poleg-Polsky and
It is widely assumed that DSGCs compute motion by integrating Diamond, 2016). In contrast, we assessed NMDAR synaptic
excitation mediated by bipolar cells and inhibition mediated by functions in the presence of nAChR antagonists, leaving feedfor-
SACs (Figure 1A), leaving a less well-defined role for ACh (Borst ward inhibition intact. In these conditions, NMDAR-only spiking
and Euler, 2011; Demb, 2007; Taylor and Smith, 2012; Vaney was not observed despite its strong potential to mediate robust
et al., 2012). Here we posit that fast cholinergic and GABAergic excitation (gNMDA 10 nS at +40 mV). Thus, we conclude that
inputs from the SAC network are key factors in computing DS in the physiological context of low-contrast stimulation, NMDAR
responses in ganglion cells, while excitatory bipolar cell inputs inputs are silent, as widely observed in other parts of the brain
play a secondary role in amplifying DS responses through (Kerchner and Nicoll, 2008). Thus, our results suggest a funda-
NMDAR activation. mentally different role for NMDARs than previously envisioned
The importance of ACh in the DS computation was revealed (Cohen and Miller, 1995; Kittila and Massey, 1997; Weng et al.,
using pharmacology, as well as an in-depth analysis of the 2005), where DSGCs only respond to glutamate during coinci-
synaptic conductances mediating the DSGC’s light response. dent cholinergic activity initiated by SACs. This requirement
Linear regression analysis supports the pharmacological results ensures that SAC inputs control the timing of the DSGC excitatory
and further demonstrates that GABA and ACh are differentially responses, a key factor in determining direction selectivity (Taylor
transmitted to the DSGC. In addition, the re-instatement of direc- and Vaney, 2002; Lee et al., 2010; Fried et al., 2005).
tion selectivity in ganglion cells after photoreceptor blockade
in SAC-ChR2 retinas demonstrates the intrinsic capacity of the Differential Transmission of GABA and ACh Underlies
SAC network to differentially transmit ACh and GABA signals Direction Selectivity in Ganglion Cells
to DSGCs. While a variety of amacrine cells are known to excite More than 30 years ago, Ariel and Daw (1982) envisioned that
ganglion cells via direct chemical (Kim et al., 2015; Krishnasw- cholinergic excitation and GABAergic inhibition drove direction

1252 Neuron 90, 1243–1256, June 15, 2016


selectivity in ganglion cells, based on simple pharmacology. GABA by SACs could lead to distinct spatiotemporal patterns
Here, we evaluate for the first time the temporal dynamics of of activity at the level of the DSGC. Alternatively, a heteroge-
GABAR- and nAChR-mediated synaptic responses evoked by neous set of ACh and GABA SAC-DSGC synapses with distinct
low-contrast moving stimuli and found GABA and ACh to be symmetric and asymmetric wiring patterns could also explain
differentially transmitted to the DSGC, supporting this model. why ACh, but not GABA, is isotropic. Experimental verification
Specifically, in the preferred direction, SACs mediate cholinergic of the mechanisms underlying isotropic ACh signaling awaits
excitation that preceded the onset of inhibition. SAC ACh also the development of a new generation of genetically encoded
gates the bipolar cell glutamate signals by recruiting NMDARs, optical sensors that can directly probe ACh dynamics at the level
producing an extremely robust, well-timed excitation. In the of single synapses.
null direction, SACs co-release ACh and GABA on a similar time-
scale, generating a net GABAergic inhibition (note NMDARs are Technical Limitations of Voltage-Clamp Techniques
not recruited under these conditions). The functional relevance Applied to Mouse DSGCs
of the mixed transmission we’ve described can be easily under- The original study characterizing synaptic conductances in
stood in the context of the present framework of the DS circuit DSGCs in rabbit retina assumed that a majority of the inputs
(Borst and Euler, 2011; Demb, 2007; Taylor and Smith, 2012; Va- were mediated by bipolar cells, based on the linear AMPAR-like
ney et al., 2012), described below. I-V relationship of the measured synaptic responses (Taylor and
Anatomical studies have shown that SACs with somas located Vaney, 2002). However, since NMDAR- and nAChR- mediated
on the null side of the DSGC’s receptive field (null-side SACs, currents have complementary voltage-dependent relationships,
where null side refers to the side from which null stimuli enter together they would tend to linearize the total response and
the DSGC’s receptive field) make ten times the number of syn- make it difficult to tease apart the three receptor components
aptic contacts with the DSGC compared to preferred-side SAC (ACh, NMDA, and AMPA) known to drive DGSCs. Here, in mouse
dendrites (Briggman et al., 2011). Direct electrical/optogenetic DSGCs, using a large range of holding potentials and by blocking
stimulation of SACs in pharmacological isolation indicates a NMDARs, we simplify the analysis and clearly reveal the
similar asymmetrical spatial distribution of GABA inhibition (Fried separate contributions of linear AMPARs and rectifying nAChRs
et al., 2002; Lee et al., 2010; Wei et al., 2011; Yonehara et al., in the composite synaptic response using regression analysis.
2011). During motion, the intrinsic DS preferences of SAC Our results suggest that at high contrasts, both SACs and
dendrites (Euler et al., 2002) and asymmetries of the wiring/ bipolar cells contribute equally to DGSC responses, while at
functional connectivity, along with network interactions (Kostadi- low contrast, cholinergic inputs dominate (Figure 6).
nov and Sanes, 2015; Lee et al., 2010; Münch and Werblin, It should be noted that when comparing cholinergic conduc-
2006), lead to the formation of an anisotropic postsynaptic tances during null and preferred motion, it is possible that cholin-
GABAergic inhibition. Our results are consistent with these ergic excitation in the null direction is underestimated due to
well-established synaptic mechanisms of the DS circuit. How- space-clamp errors caused by the large inhibitory conductance
ever, the finding that direction selectivity in ganglion cells can evoked during null motion (Poleg-Polsky and Diamond, 2011),
be driven in the absence of bipolar cell inputs using optogenetics and whether cholinergic excitation is directional (Figure 7C;
brings into question the significance of more recent models in Lee et al., 2010; Pei et al., 2015) remains to be confirmed using
which distinct bipolar cells with different kinetics are proposed alternative methods (Park et al., 2014). Importantly, voltage-
to play a key role in the DS computation (Greene et al., 2016; clamp errors do not appear to strongly bias the estimation of
Kim et al., 2014). nAChR or AMPAR contributions to the low-contrast preferred
The spatial constraints arising from anatomical connectivity motion response, as the I-V relationships of the synaptic re-
that give rise to anisotropic GABAergic inhibition appear to be sponses changed in a predictable manner using a variety
relaxed for SAC excitation. ACh excitation appears isotropic of experimental manipulations (Figure 6). Thus, the major
either when driven artificially (Ariel and Daw, 1982; Grzywacz conclusion that cholinergic excitation is a prominent component
and Amthor, 1993; Kittila and Massey, 1997; Lee et al., 2010; of DSGC responses under low-contrast conditions is well
Yonehara et al., 2011) or when driven with moving stimuli as substantiated.
we have shown here (Figure 7; but see Lee et al., 2010; Pei
et al., 2015). As suggested in previous studies (Briggman et al., Conclusions
2011), isotropic excitation could be an outcome of the paracrine Mixed neurons releasing both fast excitatory/inhibitory transmit-
nature of ACh transmission. The dense plexus of SAC dendrites ters are found throughout the CNS, including the lateral/medial
releasing ACh (with each point containing overlapping dendrites habenula (Root et al., 2014; Shabel et al., 2014), auditory brain-
originating from 30–60 SACs; Keeley et al., 2007) and the diffuse stem (Gillespie et al., 2005), CA3 region of the hippocampus (Gil-
expression of acetylcholinesterase (Nichols and Koelle, 1968) lespie et al., 2005; Gutiérrez et al., 2003), and basal forebrain
together promote paracrine transmission of ACh in the retina cholinergic system (Saunders et al., 2015), as well as in the retina
(Ariel and Daw, 1982; Ford et al., 2012; Schmidt et al., 1987). (Brecha et al., 1988; Vaney and Young, 1988). Modulating the
Paracrine transmission would allow DSGCs to pool cholinergic balance between inhibition and excitation can provide mixed
signals arising from many dendrites orientated in different direc- neurons with an effective means of controlling their net output
tions. In contrast to ACh, the clearance of GABA from the synap- (Shabel et al., 2014). However, we have found that co-release
tic cleft relies on strong uptake mechanisms that confine its of mixed transmitters does not necessarily lead to a simple
action to the synapse. In this way, co-release of ACh and cancellation. By virtue of the divergent transmitter release and

Neuron 90, 1243–1256, June 15, 2016 1253


clearance mechanisms, together with the distinct biophysical exponential equation accounting for its weak rectification (Manookin et al.,
properties of postsynaptic receptors, mixed neurons can finely 2010),
control the spatiotemporal profiles of neural activity. This may fGABA = eaV + b;
represent a unique way to generate exquisitely timed inhibition
where a = 0.0105 and b = 0.4602 are constants. Responses mediated by
and excitation that is important for high-fidelity computations AMPARs were linear, resulting in fAMPA = 1. The reported receptor conduc-
(Cafaro and Rieke, 2010), even under conditions of high network tances (gACh, gGABA, and gAMPA) were estimated near the resting membrane
variability. In the case of DSGCs, a specific wiring of SACs to potential 60 mV. Population data have been expressed as mean ± SEM.
DSGCs (Briggman et al., 2011) allows inhibition and excitation Student’s t test was used to compare values under different conditions and
to be separated in space and time despite GABA and ACh orig- the differences were considered significant when p % 0.05.
inating from the same source, providing a unique signaling
mechanism for conferring upon downstream ganglion cells the SUPPLEMENTAL INFORMATION
ability to code direction using minimal hardware.
Supplemental Information includes Supplemental Experimental Procedures
and six figures and can be found with this article online at http://dx.doi.org/
EXPERIMENTAL PROCEDURES 10.1016/j.neuron.2016.04.041.

All procedures were performed in accordance with the Canadian Council on AUTHOR CONTRIBUTIONS
Animal Care and approved by the University of Victoria’s Animal Care Commit-
tee. Experiments were performed using adult Hb9eGFP (RRID: MGI_109160),
S.S. designed and performed most experiments, analyzed results, and
or ChatCre (RRID: MGI_5475195) crossed with reporter mice, Ai9 (RRID:
contributed to writing the paper. A.J.M. and A.H. collected data that contrib-
MGI_3809523) or Ai32 (RRID: MGI_5013789). Retinae were extracted from
uted to Figures 3F and 3G and Figures 1E and 1F, respectively. G.D. and
mouse eyes under dim red light and flat mounted on a piece of filter paper
D.J.S. helped design the conductance analysis/natural stimuli. G.B.A. de-
with a pre-cut window, through which visual stimuli were presented. Superior
signed experiments and wrote the paper.
coding GFP+ DSGCs were identified in the Hb9eGFP retina using two-photon
laser-scanning microscopy techniques (Trenholm et al., 2013a), while other
ACKNOWLEDGMENTS
types of GFP DSGCs were identified by their medium-sized elliptical somata
and by their DS responses. Light stimuli, produced using a digital light projec-
We thank Kerry Delaney, Bob Chow, and Craig Brown for their useful discus-
tor (Hitachi Cpx1, refresh rate 75 Hz), were focused onto the outer segments of
sions, and Laura Hanson and Rishi Vasandani for providing helpful technical
the photoreceptors using the sub-stage condenser. The background lumi-
support. This work was supported by operating grants from the Foundation
nance, measured with a calibrated spectrophotometer (Ocean Optics), was
Fighting Blindness (Canada), Brain Canada, and Canadian Institutes for Health
set to 10 photoisomerisations/s (R*/sec). Visual stimuli created in the Matlab
Research (CIHR MOP 142301) awarded to G.B.A.
environment (Psychtoolbox) were of positive contrasts, ranging between 15%
and 1,000% (Weber contrast). In some experiments, 50–100 ms light flashes
Received: November 2, 2015
using a blue LED (470 nm) were used to drive ChR2 (10 mW/cm2 or 108
Revised: March 9, 2016
R*/sec). All experiments were performed at 35 C–37 C.
Accepted: April 18, 2016
Light-evoked synaptic currents were recorded at membrane potentials be-
Published: May 26, 2016
tween 70 mV and +20 mV, stepped in 5–10 mV increments, and were fit to
the following equation using Matlab’s built-in Levenberg-Marquardt algorithm:
REFERENCES
IðVÞ = gACh  fACh ðVÞ  ðV  ECat Þ + gGABA  fGABA ðVÞ  ðV  ECl Þ + gAMPA
 fAMPA ðV  ECat Þ; Ariel, M., and Daw, N.W. (1982). Pharmacological analysis of directionally
sensitive rabbit retinal ganglion cells. J. Physiol. 324, 161–185.
where I is the instantaneous total current at the membrane voltage V,
Borst, A., and Euler, T. (2011). Seeing things in motion: models, circuits, and
Ecat the cationic reversal potential = 0 mV, and ECl the chloride reversal
mechanisms. Neuron 71, 974–994.
potential = 56mV. fACh, fGABA, and fAMPA are basis functions for these recep-
tors, normalized to 1 at 60 mV (described below). gACh, gAMPA, and gGABA Brecha, N., Johnson, D., Peichl, L., and Wässle, H. (1988). Cholinergic ama-
denote the conductance mediated by nAChR, AMPAR, and GABARs, respec- crine cells of the rabbit retina contain glutamate decarboxylase and gamma-
tively, estimated by the fitting procedure. The fitting was performed at a time aminobutyrate immunoreactivity. Proc. Natl. Acad. Sci. USA 85, 6187–6191.
interval of 1 ms (Figure 7) or over a 30–50 ms period (Figures 5 and 6). Briggman, K.L., Helmstaedter, M., and Denk, W. (2011). Wiring specificity in
The Woodhull equation, which well describes the voltage-dependent inward the direction-selectivity circuit of the retina. Nature 471, 183–188.
rectification of nAChRs (Haghighi and Cooper, 1998), was used to define its Buldyrev, I., Puthussery, T., and Taylor, W.R. (2012). Synaptic pathways that
basis function (fACh), shape the excitatory drive in an OFF retinal ganglion cell. J. Neurophysiol.
Gmax 107, 1795–1807.
fACh ðVÞ =  
Cafaro, J., and Rieke, F. (2010). Noise correlations improve response fidelity
1 + ½S
kd
and stimulus encoding. Nature 468, 964–967.

 Carandini, M., and Ferster, D. (2000). Membrane potential and firing rate in cat
VzdF
kd = kdð0Þ  e RT ; primary visual cortex. J. Neurosci. 20, 470–484.
Cohen, E.D., and Miller, R.F. (1995). Quinoxalines block the mechanism of
where Gmax is the maximum conductance, [S] is the internal concentration of
directional selectivity in ganglion cells of the rabbit retina. Proc. Natl. Acad.
spermine (100 mM), kd is the dissociation constant at a given voltage V, kd (0)
Sci. USA 92, 1127–1131.
is the dissociation constant at 0 mV estimated as 10.66 mM, z is the valence
of spermine assumed to be 3.8 at physiological pH (Haghighi and Cooper, Demb, J.B. (2007). Cellular mechanisms for direction selectivity in the retina.
1998), d is the fraction of the membrane electrical field sensed by spermine Neuron 55, 179–186.
from the intracellular side and was estimated to be 48.4%, F is the Faraday Diamond, J.S., and Copenhagen, D.R. (1993). The contribution of NMDA and
constant, and R and T are the universal gas constant and absolute tempera- non-NMDA receptors to the light-evoked input-output characteristics of retinal
ture, respectively. The GABAR basis function (fGABA) was described by an ganglion cells. Neuron 11, 725–738.

1254 Neuron 90, 1243–1256, June 15, 2016


Diamond, J.S., and Copenhagen, D.R. (1995). The relationship between light- Lee, S., Kim, K., and Zhou, Z.J. (2010). Role of ACh-GABA cotransmission in
evoked synaptic excitation and spiking behaviour of salamander retinal detecting image motion and motion direction. Neuron 68, 1159–1172.
ganglion cells. J. Physiol. 487, 711–725. Lipin, M.Y., Taylor, W.R., and Smith, R.G. (2015). Inhibitory input to the direc-
Dmitrieva, N.A., Strang, C.E., and Keyser, K.T. (2007). Expression of alpha 7 tion-selective ganglion cell is saturated at low contrast. J. Neurophysiol. 114,
nicotinic acetylcholine receptors by bipolar, amacrine, and ganglion cells of 927–941.
the rabbit retina. J. Histochem. Cytochem. 55, 461–476. Manookin, M.B., Weick, M., Stafford, B.K., and Demb, J.B. (2010). NMDA
Elgueta, C., Vielma, A.H., Palacios, A.G., and Schmachtenberg, O. (2015). receptor contributions to visual contrast coding. Neuron 67, 280–293.
Acetylcholine induces GABA release onto rod bipolar cells through hetero- Massey, S.C., and Neal, M.J. (1979). The light evoked release of acetylcholine
meric nicotinic receptors expressed in A17 amacrine cells. Front. Cell. from the rabbit retina iN vivo and its inhibition by gamma-aminobutyric acid.
Neurosci. 9, 6. J. Neurochem. 32, 1327–1329.
Euler, T., Detwiler, P.B., and Denk, W. (2002). Directionally selective calcium Massey, S.C., and Redburn, D.A. (1983). The cholinergic amacrine cells of
signals in dendrites of starburst amacrine cells. Nature 418, 845–852. rabbit retina receive on and off input: an analysis of [3H]-ACh release using
Ford, K.J., Félix, A.L., and Feller, M.B. (2012). Cellular mechanisms underlying 2-amino-4-phosphonobutyric acid (APB) and chloride free medium. Vision
spatiotemporal features of cholinergic retinal waves. J. Neurosci. 32, 850–863. Res. 23, 1615–1620.
Fried, S.I., Münch, T.A., and Werblin, F.S. (2002). Mechanisms and circuitry Münch, T.A., and Werblin, F.S. (2006). Symmetric interactions within a homo-
underlying directional selectivity in the retina. Nature 420, 411–414. geneous starburst cell network can lead to robust asymmetries in dendrites of
starburst amacrine cells. J. Neurophysiol. 96, 471–477.
Fried, S.I., Münch, T.A., and Werblin, F.S. (2005). Directional selectivity is
formed at multiple levels by laterally offset inhibition in the rabbit retina. Naka, K.I., and Rushton, W.A. (1966). S-potentials from luminosity units in the
Neuron 46, 117–127. retina of fish (Cyprinidae). J. Physiol. 185, 587–599.

Froudarakis, E., Berens, P., Ecker, A.S., Cotton, R.J., Sinz, F.H., Yatsenko, D., Nichols, C.W., and Koelle, G.B. (1968). Comparison of the localization of
Saggau, P., Bethge, M., and Tolias, A.S. (2014). Population code in mouse V1 acetylcholinesterase and non-specific cholinesterase activities in mammalian
facilitates readout of natural scenes through increased sparseness. Nat. and avian retinas. J. Comp. Neurol. 133, 1–16.
Neurosci. 17, 851–857. Park, S.J., Kim, I.J., Looger, L.L., Demb, J.B., and Borghuis, B.G. (2014).
Gillespie, D.C., Kim, G., and Kandler, K. (2005). Inhibitory synapses in the Excitatory synaptic inputs to mouse on-off direction-selective retinal ganglion
developing auditory system are glutamatergic. Nat. Neurosci. 8, 332–338. cells lack direction tuning. J. Neurosci. 34, 3976–3981.

Greene, M.J., Kim, J.S., and Seung, H.S.; EyeWirers (2016). Analogous Patneau, D.K., and Mayer, M.L. (1990). Structure-activity relationships for
convergence of sustained and transient inputs in parallel on and off pathways amino acid transmitter candidates acting at N-methyl-D-aspartate and quis-
for retinal motion computation. Cell Rep. 14, 1892–1900. qualate receptors. J. Neurosci. 10, 2385–2399.

Grzywacz, N.M., and Amthor, F.R. (1993). Facilitation in ON-OFF directionally Pei, Z., Chen, Q., Koren, D., Giammarinaro, B., Acaron Ledesma, H., and Wei,
selective ganglion cells of the rabbit retina. J. Neurophysiol. 69, 2188–2199. W. (2015). Conditional knock-out of vesicular GABA transporter gene from
starburst amacrine cells reveals the contributions of multiple synaptic mech-
Grzywacz, N.M., Amthor, F.R., and Merwine, D.K. (1998). Necessity of acetyl- anisms underlying direction selectivity in the retina. J. Neurosci. 35, 13219–
choline for retinal directionally selective responses to drifting gratings in rabbit. 13232.
J. Physiol. 512, 575–581.
Poleg-Polsky, A., and Diamond, J.S. (2011). Imperfect space clamp permits
Gutiérrez, R., Romo-Parra, H., Maqueda, J., Vivar, C., Ramı̀rez, M., Morales, electrotonic interactions between inhibitory and excitatory synaptic conduc-
M.A., and Lamas, M. (2003). Plasticity of the GABAergic phenotype of the ‘‘glu- tances, distorting voltage clamp recordings. PLoS ONE 6, e19463.
tamatergic’’ granule cells of the rat dentate gyrus. J. Neurosci. 23, 5594–5598.
Poleg-Polsky, A., and Diamond, J.S. (2016). NMDA receptors multiplicatively
Haghighi, A.P., and Cooper, E. (1998). Neuronal nicotinic acetylcholine recep- scale visual signals and enhance directional motion discrimination in retinal
tors are blocked by intracellular spermine in a voltage-dependent manner. ganglion cells. Neuron 89, 1277–1290.
J. Neurosci. 18, 4050–4062.
Root, D.H., Mejias-Aponte, C.A., Qi, J., and Morales, M. (2014). Role of gluta-
Keeley, P.W., Whitney, I.E., Raven, M.A., and Reese, B.E. (2007). Dendritic matergic projections from ventral tegmental area to lateral habenula in aver-
spread and functional coverage of starburst amacrine cells. J. Comp. sive conditioning. J. Neurosci. 34, 13906–13910.
Neurol. 505, 539–546.
Sagdullaev, B.T., McCall, M.A., and Lukasiewicz, P.D. (2006). Presynaptic
Kerchner, G.A., and Nicoll, R.A. (2008). Silent synapses and the emergence of inhibition modulates spillover, creating distinct dynamic response ranges of
a postsynaptic mechanism for LTP. Nat. Rev. Neurosci. 9, 813–825. sensory output. Neuron 50, 923–935.
Kim, J.S., Greene, M.J., Zlateski, A., Lee, K., Richardson, M., Turaga, S.C., Saunders, A., Granger, A.J., and Sabatini, B.L. (2015). Corelease of acetylcho-
Purcaro, M., Balkam, M., Robinson, A., Behabadi, B.F., et al.; EyeWirers line and GABA from cholinergic forebrain neurons. eLife 4, http://dx.doi.org/
(2014). Space-time wiring specificity supports direction selectivity in the retina. 10.7554/eLife.06412.
Nature 509, 331–336.
Schmidt, M., Humphrey, M.F., and Wässle, H. (1987). Action and localization of
Kim, T., Soto, F., and Kerschensteiner, D. (2015). An excitatory amacrine cell acetylcholine in the cat retina. J. Neurophysiol. 58, 997–1015.
detects object motion and provides feature-selective input to ganglion cells
Sethuramanujam, S., and Slaughter, M.M. (2015). Properties of a glutamater-
in the mouse retina. eLife 4, http://dx.doi.org/10.7554/eLife.08025.
gic synapse controlling information output from retinal bipolar cells. PLoS ONE
Kittila, C.A., and Massey, S.C. (1997). Pharmacology of directionally selective 10, e0129133.
ganglion cells in the rabbit retina. J. Neurophysiol. 77, 675–689. Shabel, S.J., Proulx, C.D., Piriz, J., and Malinow, R. (2014). Mood regulation.
Kolb, H., and Famiglietti, E.V. (1974). Rod and cone pathways in the inner GABA/glutamate co-release controls habenula output and is modified by an-
plexiform layer of cat retina. Science 186, 47–49. tidepressant treatment. Science 345, 1494–1498.
Kostadinov, D., and Sanes, J.R. (2015). Protocadherin-dependent dendritic Stafford, B.K., Manookin, M.B., Singer, J.H., and Demb, J.B. (2014). NMDA
self-avoidance regulates neural connectivity and circuit function. eLife 4, and AMPA receptors contribute similarly to temporal processing in mamma-
http://dx.doi.org/10.7554/eLife.08964. lian retinal ganglion cells. J. Physiol. 592, 4877–4889.
Krishnaswamy, A., Yamagata, M., Duan, X., Hong, Y.K., and Sanes, J.R. Strang, C.E., Long, Y., Gavrikov, K.E., Amthor, F.R., and Keyser, K.T. (2015).
(2015). Sidekick 2 directs formation of a retinal circuit that detects differential Nicotinic and muscarinic acetylcholine receptors shape ganglion cell response
motion. Nature 524, 466–470. properties. J. Neurophysiol. 113, 203–217.

Neuron 90, 1243–1256, June 15, 2016 1255


Tadmor, Y., and Tolhurst, D.J. (2000). Calculating the contrasts that retinal Wei, W., Hamby, A.M., Zhou, K., and Feller, M.B. (2011). Development
ganglion cells and LGN neurones encounter in natural scenes. Vision Res. of asymmetric inhibition underlying direction selectivity in the retina. Nature
40, 3145–3157. 469, 402–406.
Taylor, W.R., and Smith, R.G. (2012). The role of starburst amacrine cells in
Weng, S., Sun, W., and He, S. (2005). Identification of ON-OFF direction-selec-
visual signal processing. Vis. Neurosci. 29, 73–81.
tive ganglion cells in the mouse retina. J. Physiol. 562, 915–923.
Taylor, W.R., and Vaney, D.I. (2002). Diverse synaptic mechanisms generate
direction selectivity in the rabbit retina. J. Neurosci. 22, 7712–7720. Xin, D., and Bloomfield, S.A. (1997). Tracer coupling pattern of amacrine and
ganglion cells in the rabbit retina. J. Comp. Neurol. 383, 512–528.
Trenholm, S., McLaughlin, A.J., Schwab, D.J., and Awatramani, G.B. (2013a).
Dynamic tuning of electrical and chemical synaptic transmission in a network Yonehara, K., Balint, K., Noda, M., Nagel, G., Bamberg, E., and Roska, B.
of motion coding retinal neurons. J. Neurosci. 33, 14927–14938. (2011). Spatially asymmetric reorganization of inhibition establishes a mo-
Trenholm, S., Schwab, D.J., Balasubramanian, V., and Awatramani, G.B. tion-sensitive circuit. Nature 469, 407–410.
(2013b). Lag normalization in an electrically coupled neural network. Nat.
Yonehara, K., Farrow, K., Ghanem, A., Hillier, D., Balint, K., Teixeira, M.,
Neurosci. 16, 154–156.
Juttner, J., Noda, M., Neve, R.L., Conzelmann, K.K., et al. (2013). The first
Vaney, D.I. (1991). Many diverse types of retinal neurons show tracer coupling
stage of cardinal direction selectivity is localized to the dendrites of retinal gan-
when injected with biocytin or Neurobiotin. Neurosci. Lett. 125, 187–190.
glion cells. Neuron 79, 1078–1085.
Vaney, D.I., and Young, H.M. (1988). GABA-like immunoreactivity in cholin-
ergic amacrine cells of the rabbit retina. Brain Res. 438, 369–373. Yoshida, K., Watanabe, D., Ishikane, H., Tachibana, M., Pastan, I., and
Nakanishi, S. (2001). A key role of starburst amacrine cells in originating
Vaney, D.I., Sivyer, B., and Taylor, W.R. (2012). Direction selectivity in
retinal directional selectivity and optokinetic eye movement. Neuron 30,
the retina: symmetry and asymmetry in structure and function. Nat. Rev.
771–780.
Neurosci. 13, 194–208.
Vlasits, A.L., Bos, R., Morrie, R.D., Fortuny, C., Flannery, J.G., Feller, M.B., and Zhang, J., and Diamond, J.S. (2009). Subunit- and pathway-specific localiza-
Rivlin-Etzion, M. (2014). Visual stimulation switches the polarity of excitatory tion of NMDA receptors and scaffolding proteins at ganglion cell synapses in
input to starburst amacrine cells. Neuron 83, 1172–1184. rat retina. J. Neurosci. 29, 4274–4286.

1256 Neuron 90, 1243–1256, June 15, 2016

You might also like