Two-photon single-cell optogenetic control of neuronal activity by sculpted light

Communicated by Charles V. Shank, Howard Hughes Medical Institute, Janelia Farm Research Campus, Ashburn, VA 20147, May 14, 2010 (received for review April 6, 2010)
June 11, 2010
107 (26) 11981-11986

Abstract

Recent advances in optogenetic techniques have generated new tools for controlling neuronal activity, with a wide range of neuroscience applications. The most commonly used approach has been the optical activation of the light-gated ion channel channelrhodopsin-2 (ChR2). However, targeted single-cell-level optogenetic activation with temporal precessions comparable to the spike timing remained challenging. Here we report fast (≤1 ms), selective, and targeted control of neuronal activity with single-cell resolution in hippocampal slices. Using temporally focused laser pulses (TEFO) for which the axial beam profile can be controlled independently of its lateral distribution, large numbers of channels on individual neurons can be excited simultaneously, leading to strong (up to 15 mV) and fast (≤1 ms) depolarizations. Furthermore, we demonstrated selective activation of cellular compartments, such as dendrites and large presynaptic terminals, at depths up to 150 μm. The demonstrated spatiotemporal resolution and the selectivity provided by TEFO allow manipulation of neuronal activity, with a large number of applications in studies of neuronal microcircuit function in vitro and in vivo.
Artificial stimulation and inhibition of neuronal activity has applications ranging from fundamental neurobiology questions (14) to potential clinical treatment of neuropsychiatric disorders (5, 6). Historically, this task has been achieved mainly by electrical stimulation. Yet the recent development of genetic techniques for sensitizing neurons to optical stimulation (712) and silencing (1316) have, for the first time, provided cell type–specific control of neuronal activity, which has been successfully used to address significant biological questions (13, 1719). In the most widely used approach, genetically expressed light-gated ion channels, such as channelrhodopsin-2 (ChR2) (20, 21), or ion pumps, such as halorhodopsin or archaerhodopsin-3 (22, 23), and optical methods for triggering their function are combined to control neuronal activity. However, although optical activation of ChR2 has been suitable to induce population activity in a large number of neurons, directed, efficient, and fast stimulation of single cells has not been feasible. The main reason for that is the low channel conductance of ChR2 (21, 24). To achieve sufficient depolarizations, a large number of channels have to be activated nearly simultaneously, which typically extend over an area of tens of square micrometers. Although conventional one-photon excitation satisfies this requirement, light scattering and the extended axial beam parameter of the excitation area lead typically to inevitable activation of neurons in an untargeted fashion. Some success in addressing these issues has recently been reported (25, 26) using a two-photon scanning (27) approach. Although these studies (25, 26) could show high spatial resolution, the necessary time to activate a large surface area sufficient to fire action potentials (APs) was ≈30 ms. Thus, neither the one-photon nor the two-photon scanning excitation provides the necessary combination of high spatial selectivity and the ability to stimulate a membrane area that is large enough to produce simultaneous and significant rapid depolarizations on a single neuron.

Results

ChR2 Activation by Two-Photon Temporal Focusing.

We took a different approach for addressing the shortcomings of both the one-photon and the two-photon method in a fundamentally unique way and have demonstrated reliable single neuron–specific activation with temporal resolution (≤1 ms) in hippocampal slices. This was done by effectively decoupling the lateral (i.e., the waist size) and the axial (i.e., the Rayleigh length) beam parameters for two-photon absorption, which are usually coupled for a Gaussian beam, and thereby sculpting the spatial light distribution at the focal plane. For a pulsed laser source, the decoupling can effectively be achieved by using the spectrum of the pulse to control its two-photon absorption probability in the axial direction. This has been demonstrated in the technique of temporal focusing (28, 29), which has also found applications in widefield two-photon imaging (30) and in 3D multilayer super-resolution microscopy (31). Two-photon temporal focusing (TEFO) can be experimentally realized by first spatially broadening a femtosecond optical pulse using a diffraction grating. The spot on the grating is then imaged onto the sample plane using a telescope that consists of the microscope objective and an additional lens. This configuration leads to an axial geometry in which the pulse is broadened everywhere in the sample except at the image plane, where the spatially separated paths of the different frequency components in the pulse meet again and the pulse reaches its minimum width. (Fig. S1 A and B). The two-photon excitation probability in TEFO is inversely proportional to the pulse width squared, which leads to a depth of field ≈2 orders of magnitude shorter (31) than the widefield one-photon epifluorescence technique. This allows a sectioning capability close to a confocal or two-photon microscope (27), while providing simultaneous excitation of an area that is ≈3 orders of magnitude larger than the diffraction-limited spot (Fig. S1C). Therefore, TEFO can be used for simultaneous excitation of a thin disk-like volume in a biological sample. In the present work we took advantage of this fact and demonstrated that TEFO can be used for targeted single-cell excitation of ChR2 with high spatiotemporal resolution in mouse and rat hippocampal slices.
After confirming the above parameters of the excitation volume for our temporally focused beam (Figs. S1B and S2), we investigated the activation of ChR2 expressed in cultured human embryonic kidney (HEK293) (Fig. 1A) cells and measured ChR2-mediated currents with whole-cell voltage clamp recordings. Fast (≤2 ms rise time) and large (≤1 nA) inward currents mediated by ChR2 were detected using pulses between 1 and 100 ms (Fig. 1 B and C). We aimed to systematically identify the optimal parameters for two-photon temporal focusing excitation and investigated the activation of ChR2 as a function of laser power, wavelength, and beam size (Fig. 1 D–I). Keeping a constant beam diameter of 3.8 μm and pulse duration of 100 ms for a range of wavelengths between 800 nm and 960 nm, we first measured the peak amplitude and rise time of the inward current as a function of laser power (Fig. 1 D–F). At each wavelength for lower powers we observed a nonlinear dependence of ChR2 activation on power, which is the signature of the two-photon effect (27). This quadratic dependence on power is not evident at first sight over the studied range of powers, because it is overshadowed by other parameters of the channel kinetics, such as saturation and desensitization. However, under our experimental conditions it could clearly be seen for the power range between 0 and 50 mW (Fig. 1D, Inset). At high powers the relationship showed near-asymptotic behavior, likely indicating the saturation of channel activation (26). These experiments revealed the maximum two-photon activation efficiency to be at 880 nm, which is slightly blue shifted compared with the double of the one-photon absorption peak (Fig. 1F). Once the optimum wavelength was identified, we examined the dependence of the response on the beam size. Using 100-ms-long pulses at 880 nm we varied the power at three different spot sizes (6 μm, 10 μm, and 14 μm). The peak inward current increased with power for all beam sizes (Fig. 1G); however, by further reducing the spot size to 4 μm we could show that the highest efficiency of activation (as shown by the larger peak inward current and faster rise time; Fig. 1H) could be achieved with the ≈6-μm spot (Fig. 1I). This fact reflects the intrinsic tradeoff between the quadratic dependence of the probability for opening a channel on the intensity, which increases with decreasing spot size, and the total number of channels on the illuminated surface area, which decreases with the spot size. The induced peak current (or the total charge influx) is proportional to the intensity squared and the area of illumination. Therefore, for a constant photon number (or a constant power) the peak current should be inversely proportional to the square of the pot size (Fig. 1G, Inset). All cells were tested for activation of ChR2 with blue light (488 nm), which revealed comparable biophysical parameters to TEFO excitation (Fig. S3 A–E).
Fig. 1.
Characterization of the biophysical properties of TEFO activation of ChR2 in HEK293 cells. (A) Superimposed fluorescent and differential interference contrast (DIC) image of a HEK293 cell during recording. (B) Representative traces of induced inward current by one-photon excitation of ChR2 (blue), TEFO excitation of ChR2-expressing (red), and of control uninfected cell (black), using 100-ms-long pulses. (C) Representative traces of TEFO-activated ChR2-mediated inward current with different pulse durations. (D and E) Peak current amplitude (D) and rise time (E) at different wavelengths measured as a function of average power at the sample for a ≈3.8-μm spot and 100-ms duration (n = 22 cells). Inset in D shows the nonlinear dependence of peak inward current on the power at lower power levels, which is the result of two-photon excitation of ChR2. (F) Wavelength-dependency of TEFO excitation of ChR2 (n = 16 cells). Inset in F emphasizes the differences in activation efficiency of 800-nm and 880-nm light. (G and H) Peak current (G) and rise time (H) as a function of spot size at 880-nm wavelength and 100-ms duration (n = 13 cells). Inset in G shows the induced peak current with the TEFO for a constant photon number (constant power at 200 mW) for different spot sizes. Maximum amplitude and fastest rise time could be achieved with a spot size of ≈5–6 μm (I). Dots and error bars represent means ± SEM.

Temporal Resolution of TEFO.

These results motivated us to investigate the potential of this method as a tool with high spatiotemporal resolution for neuronal circuit mapping and other neuroscience studies. We turned to hippocampal slices and used different types of ChR2-expressing neurons: CA1 pyramidal cells (CA1PCs) in Thy1-ChR2-YFP transgenic mice (32) and parvalbumin-positive interneurons (PV-INs) in PV-Cre mice (33) that were infected with a Cre-dependent ChR2-GFP containing adeno-associated virus (ChR2-GFP-AAV) (34) (Materials and Methods and SI Materials and Methods). First we aimed to determine the efficacy and basic biophysical characteristics of our technique to explore the optimal parameters for different types of applications. Somatic current-clamp recordings were performed in ChR2-expressing neurons while the soma was illuminated by a temporally focused spot of ≈5-μm diameter with 1-, 2-, 5-, 10-, and 100-ms pulse duration at 880 nm with ≈460-mW power at the sample (Fig. 2A). Similar to HEK293 cells, in both types of neurons fast (≈1–5 ms) and large (≈5–15 mV) ChR2-mediated depolarizations could be evoked using TEFO excitation (Fig. 2A). At a holding potential of −80 mV (to prevent generation of APs), the ChR2-induced membrane depolarization continuously increased with pulse duration and reached a saturation value of ≈15–20 mV for pulses longer than ≈10 ms (Fig. 2B). In these experiments we used a static beam path and moved the specimen to position the temporally focused spot for targeting a certain area. However, to produce fast spatiotemporal activation patterns, it would be advantageous to be able to move the excitation spot on a submillisecond time scale to any desired number of target points. To achieve this we integrated the temporal focusing beam into a microscope equipped with a pair of laser scanning mirrors (Materials and Methods and SI Materials and Methods). We tested this technical advance by recording from CA1PCs and subiculum pyramidal cells (SubPCs) in Thy1-ChR2-YFP mice and scanning variable spots in the slice on and around the somata and dendrites of the patched cells. As expected, ChR2-mediated responses were detected only when the excitation spot was placed precisely on the soma or the dendrite, but no response was evoked when the beam was placed away from the recorded cell (Fig. 2C). Using very short pulsed illuminations (0.1 ms) with short interpulse intervals (0.1 ms) at multiple points positioned on the soma, we could effectively increase the activated surface area and decrease the necessary time to achieve the desired depolarization level to evoke APs within a submillisecond time scale (Fig. 2D). (Further advantages of the multiple-spot activation along single dendrites are discussed in Fig. S4).
Fig. 2.
Temporal resolution of TEFO activation of ChR2 in hippocampal neurons of acute brain slices. (A) Representative traces and summary graph (B) (n = 5) of depolarization evoked by 1- to 100-ms-long TEFO pulses at the somata of hippocampal neurons. (A, Inset) Representative response to 488-nm light. (C) Single focal section of a ChR2-expressing CA1PC loaded with Alexa594 in hippocampal slice from Thy1-ChR2-YFP mouse, with numbered TEFO illumination spots indicated by red dots. Bottom traces show membrane potential responses to 2-ms-long stimulations at the spots indicated in C. Note that somatic stimulation fires the cell reliably, whereas dendritic stimulation results in smaller depolarization, and stimulation at a spot outside of the cell area has no effect. (D) CA1PC soma with five two-photon temporal focusing spots indicated by dots. (Lower) Response to 1-ms-long TEFO excitation at a single spot (black) as well as to 0.1-ms-long two-photon temporal focusing illuminations at three to five spots with 0.1-ms intervals. Individual excitations are also shown (Lower, interval = 300 ms). Note that scale bar for the uppermost trace differs from that related to the four lower traces. Depolarization was more effective with multiple short illuminations than that obtained with a single 1-ms pulse, despite the shorter total excitation time.

Spatial Resolution of TEFO.

Precise spatial resolution of excitation was indicated by the elimination of ChR2-mediated depolarization when the beam was moved a few microns away from the dendrite (Fig. 3 A and B). Given that dendrites provide a spatially confined and a much smaller excitable region (≈1–2 μm diameter dendritic compartments) than the soma, they were ideal for precise measurement of the spatial resolution of our TEFO technique. We used thin apical dendrites (50–400 μm from the soma, in 50–160 μm depth) and quantitatively compared the spatial resolution of our techniques with that acquired with 488-nm light. We displaced the temporally focused spot relative to the dendrite in 3D while recording voltage signals at the soma (Fig. 3 A–D). When the TEFO excitation spot was moved laterally (x-direction) or axially (z-direction), the peak depolarization dropped off sharply to 50% of the peak response after ≈10 μm and was almost eliminated at ≈20 μm distance [Fig. 3 B, C, and D (red trace)]. However, in comparison, responses induced by 488-nm light only dropped off by ≈50% at ≈50 μm away from the dendrite (x-direction) [Fig. 3B (blue trace)] and remained almost unaltered at 100 μm above or below the dendrite (z-direction) (Fig. 3 C and D). This result demonstrates a 5-fold improvement in the lateral excitation precession and at least a 10-fold improvement in the axial excitation precession. The axial resolution could be further improved by decreasing the applied power by 50% (Fig. 3C). This is a strong indication, as also discussed in a recent publication (26), that saturation of ChR2 in the center of the target plane leads to an additional contribution of the out-of-focus regions to the evoked response.
Fig. 3.
Spatial resolution of TEFO activation of ChR2. (A) Stack image of a CA1PC loaded with Alexa594, showing location of laterally moved TEFO activation spots. (B) Lateral localization of response induced by 488-nm light (blue, n = 4) and TEFO (red, n = 3). (C) Axial localization of response induced by 488-nm light (blue) and TEFO with 50% (≈130 mW, red) and 100% (≈260 mW, black) power (dashed lines indicate Gaussian fit). (Inset) Individual responses at the focal plane of the dendrite (color-coded respectively). (D) Representative responses to 488-nm (blue) and TEFO-evoked (red) responses at different distances above and below the dendrite.

Depth Penetration.

We next quantified how the efficacy of TEFO excitation depended on the depth in the tissue. The maximum depolarization evoked by single 1-ms pulses, placed on somata located at variable depths in the slice, showed a linear attenuation with depth (Fig. 4A). However, as demonstrated above, the total induced depolarization could be increased by using the multispot temporal focusing excitation approach. As a result, when a multispot temporal focusing excitation pattern with pulsed illumination durations of 0.1–0.2 ms and interpulse intervals of 0.1 ms was used, the efficiency of the excitation was increased. As shown in Fig. 4A, this led for the studied cells to a more than 2-fold increase of depth of excitation for the same total illumination duration. This dependence of the spatiotemporal patterning of excitation was used to induce depolarizations that were strong enough to evoke APs even at the apical tuft at >150 μm depth, where depolarization induced by a single short-pulse illumination was barely detectable (Fig. 4B).
Fig. 4.
Dependence of induced depolarization on tissue depth. (A) Dependence of induced depolarization at the somata of CA1PCs from Thy1-ChR2-YFP mice on tissue depth and spatiotemporal patterning of TEFO excitation for three different pulsed illumination durations and interpulse times. Red, single 1-ms pulses (n = 6, fitted linearly); open black dot and squares (total n = 3), single submillisecond pulses; filled circle and black dots (total n = 3), 5× 0.1-ms spatiotemporally patterned excitation. Note that repeating the 0.1-ms pulses five times on the same cells (filled black dot and squares, total excitation time ≤1 ms) yields a higher total depolarization compared with the single-spot 1-ms excitation. This can be used to induce bigger depolarizations at larger tissue depths. (B) Ten pulses (0.1 ms long) repeated three times with 0.1-ms intervals at the indicated locations in the apical tuft (distance from soma ≈400 μm, depth in slice ≈150 μm) evoked sufficient depolarization to evoke somatic AP (upper trace), whereas responses to single activations were almost undetectable (bottom traces).

Induction of Suprathreshhold Activity.

In current clamp, at the resting membrane potential (−58 to −67 mV), we set out to determine the parameters necessary to induce the minimal depolarization that was sufficient to reach AP threshold in a given cell of all three neuron types included in our study (Fig. 5A). We used either one spot with variable pulse durations (PV-INs and two of six CA1PCs; total t = 1–5 ms), or five spots with submillisecond pulses (four of six CA1PC, and all SubPC, total t = 0.1–0.5 ms × 5). In all cells tested (n = 13), total pulse duration of ≤5 ms evoked suprathreshold depolarization and AP firing (PV-INs: 3.0 ± 1.2 ms, n = 3; CA1PCs: 2.5 ± 0.8 ms, n = 6; SubPCs: 1.3 ± 0.1 ms, n = 4; Fig. 5A). We next tested the ability of our method to induce high-frequency firing in PCs and PV-INs. We found that all cell types usually fired APs at the first few stimulations in the train. Subsequently, PCs responded to the train of stimuli only intermittently (Fig. 5 B–D). Increasing pulse duration partially overcame AP failures and improved firing fidelity (Fig. 5C). This reduced firing probability is the consequence of the spatial confinement of the TEFO excitation leading to an inactivation of ChR2 at higher frequencies, an effect previously also observed for blue light excitation (10, 32, 35, 36). Only recently a new genetic manipulation of ChR2 (12) could overcome the ChR2 inactivation problem at higher frequencies of stimulation.
Fig. 5.
Fast and reliable AP responses by TEFO ChR2 excitation. (A) ChR2-mediated depolarization efficiently evoked APs in CA1 PV-INs (n = 3) and pyramidal cells (PC) in CA1 (n = 6) and subiculum (n = 4). ChR2-induced depolarization was measured by hyperpolarizing the soma as necessary to prevent AP firing. Graphs indicate the difference between resting membrane potential and AP threshold (green circles), amplitude of ChR2-induced suprathreshold depolarization (red squares), and the corresponding pulse duration (black squares) for each cell type. (B) Action potentials evoked in ChR2-expressing PV-IN by 1-ms-long pulses at 20–100 Hz. Insets: Subthreshold responses at hyperpolarized holding potential. Similar results were obtained in n = 3 cells. (C and D) Representative responses of a CA1PC (C, n = 7) and a subiculum PC (D, n = 6) to 20-, 50-, or 100-Hz stimulation with different pulse durations using single or multiple spots with different illumination time.

Large Boutons Activation by TEFO.

Our early data showing the lack of effect of glutamate receptor blockers on ChR2-evoked depolarization in Thy1-ChR2-YFP mice (Materials and Methods and Fig. 2) suggested that excitatory axons of ChR2-expressing principal cells, innervating the recorded cell, are not activated by two-photon temporal focusing within our standard parameter range. Further investigations of excitatory axonal activation in the CA1 region confirmed this fact (Fig. S5). Although we could not evoke synaptic responses from Schaffer collaterals, we hypothesized that certain types of terminals/axons with more favorable properties could be excited by TEFO. Synaptic boutons of PV-INs are possible candidates, because they are relatively large and they form multiple contacts on target cells (37), thus these could provide more surface area to the temporally focused two-photon illumination. To investigate that, we recorded from (ChR2-negative) CA1PCs that were innervated by ChR2-expressing PV-INs (mostly basket, axo-axonic, and bistratified cells) in AAV-ChR2-GFP virus-infected PV-Cre mice (Materials and Methods and SI Materials and Methods) in the presence of ionotropic glutamate receptor antagonists. The temporally focused beam spot was first placed around the soma and proximal dendrites of the recorded CA1PC, overlapping with the location of ChR2-expressing boutons provided by PV-INs. Our experimental conditions were optimized to detect GABAergic postsynaptic hyperpolarization (holding potential, −60 to −65 mV; reversal potential of chloride, −80 mV). Strikingly, using pulse durations of 1–2 ms, we could evoke inhibitory synaptic responses of 0.5- to 4.0-mV amplitude that were completely eliminated by a specific GABAA receptor antagonist (SR-95531, 10 μM, n = 3) (Fig. 6A). Increasing the pulse duration did not lead to any further increase of the inhibitory postsynaptic potential (IPSP) amplitude, suggesting that local synapses making contact with the postsynaptic soma were activated in an all-or-none fashion (Fig. 6B). One-photon excitation at 488 nm resulted in ≈6–9-mV hyperpolarization (Fig. 6 A and B), presumably because the longer axial size of the beam and scattering lead to activation of significantly more synapses. To examine whether activation of boutons elicited APs in the axon, we tried to evoke postsynaptic responses at different spatial locations of the activation beam. We marked 5 spots around the recorded CA1PC soma (Fig. 6 C–E, Soma), 10 spots in stratum pyramidale (Fig. 6 C–E, Pyr), 10 spots in proximal stratum radiatum (Fig. 6 C–E, Rad), and finally 5 spots on a ChR2-expressing interneuron soma (Fig. 6 C–E, IN). In all of the locations, synchronous activation of the spots evoked IPSPs in the postsynaptic cell, and the responses were eliminated by application of tetrodotoxin (TTX, 0.5 μM; Fig. 6D). Notably, the latency of the responses varied depending on the distance and activation time of the excitation spot (Fig. 6E), further indicating that axonal APs were elicited and propagated to evoke the synaptic responses. These results altogether demonstrate that the TEFO technique is efficient to directly activate the large (≈2–3 μm diameter) synaptic boutons (or population of boutons) arising from PV-INs (37).
Fig. 6.
Large inhibitory synaptic terminals are activated by TEFO ChR2 excitation. (A) Representative traces of optically evoked hyperpolarizing postsynaptic potentials in a CA1PC innervated by ChR2-expressing PV-INs (red, TEFO excitation; blue, 488-nm excitation) by targeting the soma. SR-95531 (10 μM) blocked responses with both types of stimulations (black traces, representative for n = 3 cells). (B) Peak hyperpolarization evoked by TEFO (red) did not depend on pulse duration (n = 6). Excitation at 488 nm (blue) evoked larger responses than TEFO excitation (n = 5). Dots and error bars represent means ± SEM. (C) Stack image showing a recorded CA1PC and indicating locations of series of TEFO illumination spots at the soma (5 spots), stratum pyramidale (Pyr; 10 spots), stratum radiatum (Rad; 10 spots), and at the soma of a PV-IN (IN; 5 spots). (D) Representative traces showing responses to illumination at the locations indicated in C, using 100-ms stimulus intervals (green), 0.1-ms stimulus intervals (red), and 0.1-ms stimulus intervals in the presence of 0.5 μm TTX (black). (E) Overlaid traces evoked by stimulation at the soma, in stratum pyramidale (Pyr), and at the soma of the PV-IN (IN). Note the differences in response latency and the stepwise rise of the response evoked in stratum pyramidale.

Discussion

Although the use of genetically expressed ChR2 in different types of neurons has become widely popular in recent years, one-photon excitation does not allow for spatial precision of neuronal activation using this method. Recognizing this caveat, recent studies reported the possibility of two-photon excitation of ChR2 using a scanning approach in cultured neurons (26). Although the spatial resolution indeed dramatically improved with two-photon excitation, the long scanning time necessary for sufficient channel activation and consequent depolarization significantly compromises the temporal precision of neuronal activation (≈30 ms for evoking an AP). The TEFO technique, which does not require scanning, provides reliable activation of cells within ≈1–3 ms and thereby provides ≈15 times faster activation than the scanning approach. This improvement practically implies that only the two-photon temporal focusing approach allows quasi-synchronous activation of neurons and their different cellular compartments in future studies. The combination of TEFO with a conventional dual galvanometer–based scanning system, as described in our study, makes repositioning of the excitation spot fast (<0.2 ms to any point in a 100-μm field) and convenient. The precise spatial and temporal control of firing activity of individual or a desired number of individually selected cells, especially when combined with selective expression of ChR2 in particular cell populations, opens up the possibility (among others) for detailed, high-throughput studies of connectivity and dynamics of small- and large-scale neuronal networks and assessment of the functional consequences of their activation both in vitro and in vivo.
Although the current TEFO technique can reliably stimulate a few APs, further advances of the technique, such as 3D light sculpting and development of ChR2 mutants with higher conductance or less desensitization (35), are expected to enable precise induction of longer AP trains in a wide range of frequencies. Beside the reliable generation of APs at the soma, an important aspect of our study is the demonstration of the possibility for spatially and temporally precise activation of certain cellular compartments, such as dendrites and large synaptic boutons. This feature can be exploited in many applications, for example for studying dendritic properties and integration without or combined with uncaging methods (38), or for detailed investigation of the spatiotemporal characteristics and role of inhibition by genetically labeled interneuron types. On the other hand, the lack of activation of excitatory synaptic terminals, such as Schaffer collaterals, shows that different types of synaptic boutons, presumably having diverse anatomical and biophysical properties, can exhibit different levels of excitability by the TEFO technique.
The lack of axonal activation, at least under these experimental conditions and within the explored wide parameter range, can be crucial in studies in which target neurons lie within a densely intermingled network of thin ChR2-expressing axons with small boutons, and therefore makes the TEFO technique a powerful tool for studies such as circuit mapping. On the other hand, the ability to specifically activate axons with TEFO could be useful, for example, for mapping long-range connections in the brain. Further technical developments, such as combination of our current method with holographic techniques (39), are being considered to allow targeted and specific stimulation of axonal compartments with sculpted light. In summary, our method of two-photon excitation of ChR2 by temporal focusing, especially when combined with genetic and electrophysiological techniques, has broad potential to contribute to our understanding of the cellular and network basis of neuronal functions.

Materials and Methods

Optical Setup.

Please refer to SI Materials and Methods for a detailed description. Briefly, a tunable 140-fs pulsed laser was used as the source for temporal focusing. After passing through a variable zoom telescope the beam was diffracted by a diffraction grating. The spot on the grating was imaged via a telescope consisting of an achromatic lens and the microscope objective onto the specimen plane. For the multispot excitation experiments the temporally focused beam was integrated into the scan head of a commercial two-photon scanning microscope.

Virus Preparation and Injection.

Cre recombinase–dependent virus was assembled, harvested, and purified as previously described (40) and injected into the hippocampal CA1 region of parvalbumin-Cre (33) mice and hippocampal CA3 region of Sprague-Dawley rats.

Hippocampal Slice Preparation.

Transverse slices were prepared as previously described (41), according to methods approved by the Janelia Farm Institutional Animal Care and Use Committee.

Acknowledgments

We thank J. Magee for supporting this project; A. Losonczy, J. Makara, and K. Svoboda for numerous insightful discussions, experimental suggestions, and comments on the manuscript; A. Losonczy for testing cell type-specific viruses, for sample preparation, and for help with the analysis; and D. O’Connor and G. Paez for sample preparation.

Supporting Information

Supporting Information (PDF)
Supporting Information

References

1
VS Sohal, F Zhang, O Yizhar, K Deisseroth, Parvalbumin neurons and gamma rhythms enhance cortical circuit performance. Nature 459, 698–702 (2009).
2
JA Cardin, et al., Driving fast-spiking cells induces gamma rhythm and controls sensory responses. Nature 459, 663–667 (2009).
3
HC Tsai, et al., Phasic firing in dopaminergic neurons is sufficient for behavioral conditioning. Science 324, 1080–1084 (2009).
4
L Luo, EM Callaway, K Svoboda, Genetic dissection of neural circuits. Neuron 57, 634–660 (2008).
5
V Gradinaru, M Mogri, KR Thompson, JM Henderson, K Deisseroth, Optical deconstruction of parkinsonian neural circuitry. Science 324, 354–359 (2009).
6
AR Adamantidis, F Zhang, AM Aravanis, K Deisseroth, L de Lecea, Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature 450, 420–424 (2007).
7
BV Zemelman, GA Lee, M Ng, G Miesenböck, Selective photostimulation of genetically chARGed neurons. Neuron 33, 15–22 (2002).
8
BV Zemelman, N Nesnas, GA Lee, G Miesenbock, Photochemical gating of heterologous ion channels: Remote control over genetically designated populations of neurons. Proc Natl Acad Sci USA 100, 1352–1357 (2003).
9
M Banghart, K Borges, E Isacoff, D Trauner, RH Kramer, Light-activated ion channels for remote control of neuronal firing. Nat Neurosci 7, 1381–1386 (2004).
10
ES Boyden, F Zhang, E Bamberg, G Nagel, K Deisseroth, Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci 8, 1263–1268 (2005).
11
X Li, et al., Fast noninvasive activation and inhibition of neural and network activity by vertebrate rhodopsin and green algae channelrhodopsin. Proc Natl Acad Sci USA 102, 17816–17821 (2005).
12
LA Gunaydin, et al., Ultrafast optogenetic control. Nat Neurosci 13, 387–392 (2010).
13
HAE Lechner, ES Lein, EM Callaway, A genetic method for selective and quickly reversible silencing of Mammalian neurons. J Neurosci 22, 5287–5290 (2002).
14
EM Slimko, S McKinney, DJ Anderson, N Davidson, HA Lester, Selective electrical silencing of mammalian neurons in vitro by the use of invertebrate ligand-gated chloride channels. J Neurosci 22, 7373–7379 (2002).
15
AY Karpova, DGR Tervo, NW Gray, K Svoboda, Rapid and reversible chemical inactivation of synaptic transmission in genetically targeted neurons. Neuron 48, 727–735 (2005).
16
X Han, ES Boyden, Multiple-color optical activation, silencing, and desynchronization of neural activity, with single-spike temporal resolution. PLoS ONE 2, e299 (2007).
17
D Huber, et al., Sparse optical microstimulation in barrel cortex drives learned behaviour in freely moving mice. Nature 451, 61–64 (2008).
18
L Petreanu, D Huber, A Sobczyk, K Svoboda, Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nat Neurosci 10, 663–668 (2007).
19
V Varga, et al., Fast synaptic subcortical control of hippocampal circuits. Science 326, 449–453 (2009).
20
G Nagel, et al., Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc Natl Acad Sci USA 100, 13940–13945 (2003).
21
G Nagel, et al., Channelrhodopsins: Directly light-gated cation channels. Biochem Soc Trans 33, 863–866 (2005).
22
JK Lanyi, Halorhodopsin: A light-driven chloride ion pump. Annu Rev Biophys Biophys Chem 15, 11–28 (1986).
23
BY Chow, et al., High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature 463, 98–102 (2010).
24
K Feldbauer, et al., Channelrhodopsin-2 is a leaky proton pump. Proc Natl Acad Sci USA 106, 12317–12322 (2009).
25
P Schoenenberger, A Grunditz, T Rose, TG Oertner, Optimizing the spatial resolution of Channelrhodopsin-2 activation. Brain Cell Biol 36, 119–127 (2008).
26
JP Rickgauer, DW Tank, Two-photon excitation of channelrhodopsin-2 at saturation. Proc Natl Acad Sci USA 106, 15025–15030 (2009).
27
W Denk, JH Strickler, WW Webb, Two-photon laser scanning fluorescence microscopy. Science 248, 73–76 (1990).
28
GH Zhu, J van Howe, M Durst, W Zipfel, C Xu, Simultaneous spatial and temporal focusing of femtosecond pulses. Opt Express 13, 2153–2159 (2005).
29
D Oron, E Tal, Y Silberberg, Scanningless depth-resolved microscopy. Opt Express 13, 1468–1476 (2005).
30
E Tal, D Oron, Y Silberberg, Improved depth resolution in video-rate line-scanning multiphoton microscopy using temporal focusing. Opt Lett 30, 1686–1688 (2005).
31
A Vaziri, J Tang, H Shroff, CV Shank, Multilayer three-dimensional super resolution imaging of thick biological samples. Proc Natl Acad Sci USA 105, 20221–20226 (2008).
32
BR Arenkiel, et al., In vivo light-induced activation of neural circuitry in transgenic mice expressing channelrhodopsin-2. Neuron 54, 205–218 (2007).
33
S Hippenmeyer, et al., A developmental switch in the response of DRG neurons to ETS transcription factor signaling. PLoS Biol 3, e159 (2005).
34
D Atasoy, Y Aponte, HH Su, SM Sternson, A FLEX switch targets Channelrhodopsin-2 to multiple cell types for imaging and long-range circuit mapping. J Neurosci 28, 7025–7030 (2008).
35
JY Lin, MZ Lin, P Steinbach, RY Tsien, Characterization of engineered channelrhodopsin variants with improved properties and kinetics. Biophys J 96, 1803–1814 (2009).
36
H Wang, et al., High-speed mapping of synaptic connectivity using photostimulation in Channelrhodopsin-2 transgenic mice. Proc Natl Acad Sci USA 104, 8143–8148 (2007).
37
SR Cobb, et al., Synaptic effects of identified interneurons innervating both interneurons and pyramidal cells in the rat hippocampus. Neuroscience 79, 629–648 (1997).
38
C Lutz, et al., Holographic photolysis of caged neurotransmitters. Nat Methods 5, 821–827 (2008).
39
E Papagiakoumou, V de Sars, D Oron, V Emiliani, Patterned two-photon illumination by spatiotemporal shaping of ultrashort pulses. Opt Express 16, 22039–22047 (2008).
40
JC Grieger, VW Choi, RJ Samulski, Production and characterization of adeno-associated viral vectors. Nat Protoc 1, 1412–1428 (2006).
41
A Losonczy, JK Makara, JC Magee, Compartmentalized dendritic plasticity and input feature storage in neurons. Nature 452, 436–441 (2008).

Information & Authors

Information

Published in

The cover image for PNAS Vol.107; No.26
Proceedings of the National Academy of Sciences
Vol. 107 | No. 26
June 29, 2010
PubMed: 20543137

Classifications

Submission history

Published online: June 11, 2010
Published in issue: June 29, 2010

Keywords

  1. high-resolution neuronal stimulation
  2. channelrhodopsin
  3. temporal focusing
  4. circuit mapping
  5. electrophysiology

Acknowledgments

We thank J. Magee for supporting this project; A. Losonczy, J. Makara, and K. Svoboda for numerous insightful discussions, experimental suggestions, and comments on the manuscript; A. Losonczy for testing cell type-specific viruses, for sample preparation, and for help with the analysis; and D. O’Connor and G. Paez for sample preparation.

Authors

Affiliations

Bertalan K. Andrasfalvy
Howard Hughes Medical Institute, Janelia Farm Research Campus, Ashburn, VA 20147
Boris V. Zemelman
Howard Hughes Medical Institute, Janelia Farm Research Campus, Ashburn, VA 20147
Jianyong Tang
Howard Hughes Medical Institute, Janelia Farm Research Campus, Ashburn, VA 20147
Alipasha Vaziri1 [email protected]
Howard Hughes Medical Institute, Janelia Farm Research Campus, Ashburn, VA 20147

Notes

1
To whom correspondence should be addressed. E-mail: [email protected].
Author contributions: B.K.A. designed and performed all electrophysiological experiments; B.V.Z. prepared plasmids, designed AAV-ChR2-sfGFP and Cre recombinase-dependent rAAV-FLEX-rev ChR2-sfGFP viruses, and helped with the manuscript; J.T. developed software for data collection; A.V. designed and led research, built experimental setup, and performed optical characterization experiments; B.K.A. and A.V. analyzed data; and B.K.A. and A.V. wrote the paper.

Competing Interests

The authors declare no conflict of interest.

Metrics & Citations

Metrics

Note: The article usage is presented with a three- to four-day delay and will update daily once available. Due to ths delay, usage data will not appear immediately following publication. Citation information is sourced from Crossref Cited-by service.


Citation statements




Altmetrics

Citations

Export the article citation data by selecting a format from the list below and clicking Export.

Cited by

    Loading...

    View Options

    View options

    PDF format

    Download this article as a PDF file

    DOWNLOAD PDF

    Login options

    Check if you have access through your login credentials or your institution to get full access on this article.

    Personal login Institutional Login

    Recommend to a librarian

    Recommend PNAS to a Librarian

    Purchase options

    Purchase this article to access the full text.

    Single Article Purchase

    Two-photon single-cell optogenetic control of neuronal activity by sculpted light
    Proceedings of the National Academy of Sciences
    • Vol. 107
    • No. 26
    • pp. 11651-12058

    Figures

    Tables

    Media

    Share

    Share

    Share article link

    Share on social media