Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Dec 1.
Published in final edited form as: Curr Opin Neurobiol. 2014 Jul 26;0:165–171. doi: 10.1016/j.conb.2014.07.016

What optogenetic stimulation is telling us (and failing to tell us) about fast neurotransmitters and neuromodulators in brain circuits for wake-sleep regulation

Elda Arrigoni 1, Clifford B Saper 1,*
PMCID: PMC4268002  NIHMSID: NIHMS613831  PMID: 25064179

Abstract

In the last eight years optogenetic tools have been widely used to identify functional synaptic connectivity between specific neuronal populations. Most of our knowledge comes from the photo-activation of channelrhodopsin-2 (ChR2) expressing inputs that release glutamate and GABA. More recent studies have been reporting releases of acetylcholine and biogenic amines but direct evidence for photo-evoked released of neuropetides is still limited particularly in brain slice studies. The high fidelity in the responses with photo-evoked amino-acid transmission is ideal for ChR2-assisted circuit mapping and this approach has been successfully used in different fields of neuroscience. Conversely, neuropeptides employ a slow mode of communication and might require higher frequency and prolonged stimulations to be released. These factors may have contributed to the apparent lack of success for optogenetic release of neuropetides. In addition, once released, neuropeptides often act on multiple sites and at various distances from the site of release resulting in a greater complexity of postsynaptic responses. Here, we focus on what optogenetics is telling us – and failing to tell us – about fast neurotransmitters and neuropeptides.


Optogenetic activation of specific neurons or projections has been rapidly adopted in many areas of neuroscience to study circuits of specific behaviors and to establish functional synaptic connectivity between specific neuronal populations [1,2]. Light pulses directed on the cell bodies of neurons expressing channelrhodopsin-2 (ChR2) produce a photocurrent that depolarizes and evokes neuronal firing [3] and trains of short light pulses can entrain neuronal firing at different ranges of frequencies depending on the membrane properties of the target cells [4]. Photostimulation directed at axons/terminals activates specific inputs and evokes neurotransmitter release. This approach has been used in brain slices to activate short and long-range projections even when the cell bodies that give rise to the targeted input are not contained in the recorded slices. Combining optogenetic targeting of the presynaptic elements with patch-clamp recordings of the postsynaptic neurons is a powerful approach to demonstrate functional synaptic connectivity between selective nodes of the neuronal circuit [5]. In whole-animal experiments fiber optics can be implanted to stimulate specific afferent inputs and to deconstruct neuronal circuits controlling specific behaviors [6]. The modalities of synaptic communication of amino-acid neurotransmitters and neuropeptides are very different [7-9] and probably respond differently to optogenetic activation. Other neurotransmitters such as of acetylcholine (through muscarinic receptors) and biogenic amines share some properties with amino-acids and some with neuropeptides [10-12] and are not discussed in this review. We focus on the comparison between fast-acting amino-acid and slow-acting neuropeptide transmissions and on the electrophysiological outcomes when they are optogenetically activated. We review here the stimulation patterns required for release, then compare the electrophysiological responses of amino-acids and neuropeptides, and finally review studies in which combined whole-animal and brain slice optogenetic approaches were used to map the neuronal circuitry for sleep and wake regulation.

Release of neuropeptides might require a different pattern of photostimulation than that used to release fast neurotransmitters

Although whole animal behavioral studies have reported that optogenetic stimulation can also evoke responses consistent with release of peptides [13-19], the dynamic of their release and the postsynaptic neurons on which they act remain unclear. These questions could be answered in a brain slice preparation. However, in brain slices, where photo-evoked release of amino-acid neurotransmitters has been demonstrated in many studies, there has been little evidence of release of the accompanied neuropeptides [17,20,21]. This might be due to the type of photostimulation paradigms that has been used in the whole-animal and the brain slice studies. While single brief (2-5 ms) light pulses are sufficient to evoke release of glutamate or GABA these protocols might be insufficient to release neuropeptides. Generally, the release of neuropeptides requires a higher firing frequency and longer duration of firing than the release of glutamate or other fast neurotransmitters [22-24]. This might be related to the location of the synaptic vesicles. While the clear vesicles containing glutamate or GABA are closely related to the voltage dependent Ca2+ channels in the active zone [25,26], the large dense core vesicles containing neuropeptides are located in the synaptic terminal more diffusely and away from the active zone [9,27]. Hence the release of neuropeptides requires intracellular Ca2+ to rise to higher levels to allow Ca2+ to reach the site where the neuropeptide-containing vesicles are stored. Overall the release of neuropeptides is less efficient and slower (occurring over 10 ms to seconds after stimulation) than the release of amino-acid transmitters (which occurs in less than one millisecond) [23,28]. Single action potentials can release fast neurotransmitters but can rarely release neuropeptides [27,29]. It is interesting that studies using electrical stimulation found that the ideal pattern for neuropeptide release is an initial brief high-frequency stimulation to quickly raise Ca2+ levels, followed by a tonic low frequency stimulation between 3 to 10 Hz [30]. This suggests that in addition to the frequency, the pattern and duration of the photostimulation might be important factors for releasing neuropeptides.

Amino-acids and neuropeptides are most often co-expressed and co-released [7,9]. Researchers have selectively blocked the release or the action of individual transmitters in order to isolate the functions of the different neurotransmitters involved. This can be achieved using specific antagonists [14,31] or knock-out mice [13,14]. In addition, because amino-acids and neuropeptides are probably released at different firing frequencies, one could use low frequency stimulation to drive the release of glutamate or GABA without neuropeptides and use higher frequency stimulations to drive the co-release of amino-acids and neuropeptides [13,32]. Furthermore, in behavioral studies, the inhibition of neuronal activity or specific synaptic inputs is perhaps even more informative than the activation of synaptic release. Indeed the inhibitory optogenetic tools such as the halorhodopsins (NpHRs), the archaerhodopsins (Arch or ArchT) and the chloride-conducting channelrhodopsin (ChloCs) are starting to be more widely used [33-36]. These tools are employed to silence neurons and their synaptic outputs in freely-behaving animals and they allow the testing of whether selective neuronal populations and synaptic inputs are necessary for specific behavioral functions [13,37-41].

Diversity in the electrophysiological responses to the release of fast neurotransmitters and neuropeptides

A second problem in recording peptidergic responses to optogenetic stimulation is the complexity and time scale of the responses that are seen. Fast neurotransmitters are mostly released from synaptic terminals whereas neuropeptides are released from terminals and from somato-dendritic varicosites [7,8]. Once released, amino-acid transmitters rapidly cross the synaptic cleft to activate receptors on the postynaptic membrane (point-to-point transmission) whereas neuropeptides diffuse in the extracellular space and they can act on multiple sites. Neuropeptides can activate G-protein coupled receptors (GPCRs) on postsynaptic target neurons or on synaptic terminals to modulate synaptic activity. The G-protein receptors can have myriad effects in addition, including modifying the kinetics of channels and intracellular signaling cascade responses to later incoming stimuli. Such changes can affect the membrane potential of the target neuron or its response to other stimuli subtly, but for a prolonged period of time. Neuropeptides have also been shown to act as retrograde signals regulating their own release (homosynaptic modulation) or the activity of other afferent inputs to their target neurons (heterosynaptic modulation) [23,29]. Technical aspects of brain slice recordings should also be taken into consideration. For example, the responses through G-protein receptors may be diminished by the wash-out of the cytoplasm as the patch electrode contents diffuse into the neuron. Additionally, in slice preparations, neurons are not receiving their normal patterned inputs, so it is impossible to determine how the neuropeptide might modulate the response to those inputs. Photo-evoked release of neuropeptides from ChR2 expressing neurons might therefore result in a complex array of electrophysiological responses, in many cases quite different from the responses mediated by optogenetic release of glutamate or GABA.

Another important difference between amino-acid and peptidergic transmission is the kinetics of the synaptic responses. For example, neuropeptide release is slower than the release of amino-acid, and peptides, neuropeptides diffuse to activate receptors that may be many microns away from the release site (volume transmission). In addition neuropeptides act on the postsynaptic target neurons by activating GPCRs that leads to a cascade of intracellular second messengers. This response is much slower than the response through the ionotropic receptors employed by amino-acid neurotransmitters. As a result the onset of the neuropeptide-mediated responses is much longer than for glutamate or GABA mediated postsynaptic events. Also the rise-time of amino-acid mediated postsynaptic currents is short allowing for precise measurement of postsynaptic response onset delay whereas peptide mediated currents have slow onset kinetics and it is often difficult to segregate single synaptic events. In addition, while for amino-acid-mediated postsynaptic responses, the delay and the dependence on action potentials (responses blocked by tetrodotoxin) are reliable parameters to distinguish between monosynaptic versus polysynaptic connections (4-6 ms delay is considered the threshold for monosynaptic responses) [42,43], these same properties cannot be used for neuropeptide transmission. In fact, because neuropeptides have such a long-range of diffusion that can vary among circuits, neuropeptide-mediated postsynaptic responses have long delays and low fidelity, impairing point to point mapping.

Another potential problem in studying optogenetic-mediated release of neuropeptides is the detection of the synaptic response. The current evoked from the activation of GPCRs can be small and can go undetected. On the other hand, the duration of neuropetide transmission is often much longer than amino-acid signals. Therefore optogenetic stimulation of terminals expressing glutamate or GABA together with neuropeptides is likely to produce an initial response mediated by the fast neurotransmitters, immediately followed by a longer lasting response mediated by the release of the neuropeptides (see Fig. 1).

Figure 1. Comparison of optogenetic releases of amino-acid neurotransmitters and neuropeptides.

Figure 1

Photostimulation at low frequency of ChR2 expressing axons/terminals releases only amino-acid neurotransmitters (left) but a higher frequency and prolonged photostimulation evokes release of both amino-acid and neuropeptides (right). Fast acting amino-acid transmitters activate ionotropic receptors on the postsynaptic neurons (1). Neuropeptides act slowly to activate postsynaptic G-protein-coupled receptors (GPCRs) on the postsynaptic neurons (2), on synaptic terminals to modulate synaptic activity (3, homosynaptic and 4, heterosynaptic modulations) and they can diffuse from the release site to act on distant postsynaptic targets (5). Neuropeptides can also be released from somato-dendritic varicosites and they can act as retrograde signals (6).

Optogenetic studies reveal neuronal circuits for sleep and wake regulation

Circuits that govern sleep and wake regulation were among the very first to be studied with optogenetic methods [14]. In addition, a combination of in vivo and in vitro measurements has given us important information regarding the sleep-wake neuronal circuitry, the conditions for the release of fast neurotransmitters and neuropeptides and the functional consequences of their release [32,44-46].

One example is represented by the optogenetic stimulation of the orexin (also called hypocretin) system. The orexin neurons form a cluster in the lateral hypothalamus [47,48] that promotes and maintains wakefulness; they are also involved in feeding behavior and reward processes [49-52]. In addition to orexin-A and -B, orexin neurons make and release glutamate [20,53,54], and they produce the inhibitory peptide dynorphin [55,56]. Orexin and glutamate can be localized at the same terminals but in different vesicles [53] and if co-released they should act synergistically to excite target neurons. Little is known about the conditions of release of dynorphin, although a recent electron microcopy study has reported that orexin and dynorphin are colocalized within the same synaptic vesicles [57]. Different targets of the orexin neurons may express orexin, glutamate, or κ-opioid receptors, or some combination of these. For example while orexin directly excites tuberomammillary nucleus (TMN) neurons and NPY neurons of the arcuate nucleus [58-60] dynorphin has no post-synaptic effects but reduces GABAergic synaptic input to these neurons [61]. Therefore in these cases co-release of orexin, glutamate and dynorphin would produce synergistic effects that increase activity in the target cells. A further level of complexity is that orexin, glutamate and dynorphin have effects that differ over time [62]. In addition, dynorphin could also be locally released from somato-dendritic varicosites and thus act as a retrograde signal to inhibit afferent inputs to the orexin neurons [63]. Experiments using optogenetic stimulation of orexin neurons and terminals are starting to unravel the dynamics of orexin and glutamate releases [32] and may soon reveal the functional role of dynorphin in these neurons.

Optogenetic activation of orexin neurons or their projections to the locus coeruleus (LC) using ChR2 in whole-animal studies has been shown to promote awakening [14,31]. This response was significantly reduced but not occluded in orexinligand knockout mice or when an orexin antagonist was given, suggesting that in addition to the orexin signaling another neurotransmitter, most likely glutamate, contributes to the arousal response [14,18]. The dynamics of co-release of orexin and glutamate are starting to be understood. We know for example that the optogenetic stimulation that promotes transitions to wakefulness is time- and frequency-dependent and requires trains of light pulses (5 Hz or greater) that last at least 10 seconds. These conditions might be more relevant for the release of orexin than for the release of glutamate. In fact single light pulses to orexin terminals expressing ChR2 were sufficient to release glutamate onto tuberomammillary nucleus (TMN) neurons in brain slices, but failed to evoke an orexin-mediated postsynaptic response [20]. TMN neurons express orexin receptor 2 and are excited by orexin [58,64]. Thus the lack of orexin-mediated response during in vitro optogenetic stimulation could be due to (i.) lack of orexin exocytosis; (ii.) release of orexin at levels not sufficient to evoke a detectable electrophysiological postsynaptic response; or (iii.) the stimulation parameters being inadequate to cause orexin release. The latter possibility was underscored by the recent demonstration that either long (> 30 seconds) trains of light pulses at 5 Hz or shorter trains (10 s) but at higher frequency (20 Hz) could evoke an orexin-mediated slow-onset inward postsynaptic current in TMN neurons [32]. Similar results were also reported with photostimulation of orexin axons and terminals in the LC where short trains (1 s) of light pulses at 10 Hz elicited a small orexin-mediated slow-onset depolarization of LC neurons [65]. Therefore it appears that particularly high frequency photostimulation is not a requirement for the release of orexin, which is consistent with evidence that orexin neurons fire at maximum frequency not higher than 10 Hz during attentive wakefulness [66] and up to 20 Hz during sensory stimulation [67], but hints that prolonged firing is perhaps key for orexin release.

This requirement may be physiologically relevant if brief firing of the orexin neurons during brief awakenings or during postural changes is insufficient to evoke the release of orexin, thus permitting an easy return to sleep. However if the orexin neurons remain active the release of orexin would begin promoting full arousal. In support of this hypothesis, research in orexin knockout mice suggests that orexin signaling is essential for sustaining long wake bouts but has little impact on brief arousals from sleep [68]. These findings are consistent with those showing that optogenetic release of glutamate and orexin in brain slices evokes a rapid glutamate-mediated postsynaptic current that rapidly wanes due to the depletion of the readily-releasable pool of glutamate containing vesicles [32]. Therefore while the orexin-mediated postsynaptic response has a delayed onset and slow kinetics compatible with the slow dynamics of neuropeptide release and the kinetics of GPCR-mediated currents, the orexin outlasts the glutamate-mediated response consistent with the long lasting effects of neuropeptides [32].

Three recent studies have shown that optogenetic activation of the melanin-concentrating hormone (MCH) neurons promotes sleep – both REM and non-REM sleeps in one study [69] and only REM sleep in the others [13,70]. MCH neurons are a cluster of neurons in the lateral hypothalamus that are involved in energy homeostasis [71]. MCH neurons are also active during REM sleep [72] but the direct evidence for their involvement in REM sleep regulation only came with the results from these optogenetic studies. In both studies 1 Hz stimulation was insufficient but 10-20 Hz was effective in prolonging REM sleep. This effect is maintained in MCH receptor knock-out mice suggesting that there is an as yet unknown receptor for MCH, or that the effect is due to release of another neurotransmitter, most likely GABA [13]. In addition optogenetic stimulation of MCH input to the TMN and to the medial septum but not to the dorsal raphe increased REM sleep as well [13]. In brain slices single brief light pulses directed to MCH cell bodies, axons and terminals evoked GABA release and short latency GABAA mediated inhibitory postsynaptic currents (IPSCs) in TMN neurons, but 10 second trains of stimulation at 20 Hz appeared to release MCH that had no direct effect on TMN neurons but increased GABAergic IPSC frequency in TMN neurons suggesting a presynaptic facilitation of the GABAergic input to the TMN [13].

We recently found in brain slices that photostimulation of TMN neurons and their axons and terminals in which we selectively expressed ChR2, evokes the release of histamine locally and distally in the ventrolateral preoptic nucleus (VLPO), which contains sleep-active neurons [73,74]. Within the TMN region photo-release of histamine, possibly from somato-dendritic vesicles, suppresses GABAergic inputs onto TMN neurons via H3 receptors. In the VLPO, photostimulation of the TMN input increased inhibitory synaptic input to the VLPO neurons. This response is indirect; it is mediated via H1 receptors through the activation of GABAergic interneurons that project to VLPO neurons. Both these effects were only observed when we used very long, slow trains of light pulse stimulation (1 Hz, for 2-8 minutes) whereas single light pulses failed to evoke histamine-mediated responses. The onset of the effects of photo-released histamine was slow (1-3 min) and persisted long after the stimulation was completed (10-15 min). Based on the onset delay, kinetics, and duration of the responses to photostimulation of the TMN neurons, histamine appears to act via volume transmission [75-77]. In addition, the action on synaptic inputs would suggest that once released, histamine might diffuse a short distance to activate H3 receptors in the TMN region and H1 receptors in VLPO, acting on presynaptic terminals of afferent inputs to TMN and VLPO neurons. Overall the release of histamine has more in common with the mechanisms of transmission of neuropeptides than with amino-acid neurotransmission. Importantly, this in vitro optogenetic study demonstrated synergistic mechanisms through which histamine may promote arousal by disinhibiting TMN neurons while inhibiting VLPO neurons and perhaps their projections to TMN neurons.

Conclusions

Studies employing ChR2-assisted circuit mapping in brain slices provide strong evidence for functional synaptic connectivity between selective neuronal populations. The majority of our current knowledge still comes from photo-evoked release of fast neurotransmitters whereas direct evidence for the release of neuropepdies is lagging. This scenario however might rapidly change, particularly since new promising optogenetic tools with large photocurrents and faster kinetics have been recently developed [34]. The advantage of these new opsins is their fast-off kinetics that prevent sustained depolarization between pulses allowing membrane repolarization in a similar fashion as during the firing of a burst of high frequency action potentials. These suggest great potential for investigating synaptic co-release. However, in the case of neuropeptide transmission, slow release, variability in the time delay of the postsynaptic responses and the complexity of the actions might still be a limitation for circuit mapping. The development of fast-kinetics optogenetic tools combined with the growing understanding of the dynamics of release and the actions of classic neurotrasmitters and neuromodulators is bound to broaden our functional mapping of the brain circuits involved in sleep and wakefulness as well as many other areas of neuroscience.

Highlights.

  • Optogenetic tools release fast neurotransmitters but maybe not peptides.

  • Neuropeptides use a slow mode of communication with complex postsynaptic responses.

  • ChR2-assisted circuit mapping for neuropeptide transmission have some limitations.

  • Optogenetic tools have been used to map the circuitry for sleep and wake regulation.

Acknowledgements

This study was supported by NIH grants: RO1NS061863, P01HL095491 and R21NS082854.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflict of interests: The authors declare no competing financial interests

References

  • 1.Zhang F, Gradinaru V, Adamantidis AR, Durand R, Airan RD, de Lecea L, Deisseroth K. Optogenetic interrogation of neural circuits: technology for probing mammalian brain structures. Nat Protoc. 2010;5:439–456. doi: 10.1038/nprot.2009.226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Scanziani M, Hausser M. Electrophysiology in the age of light. Nature. 2009;461:930–939. doi: 10.1038/nature08540. [DOI] [PubMed] [Google Scholar]
  • 3.Boyden ES, Zhang F, Bamberg E, Nagel G, Deisseroth K. Millisecond-timescale, genetically targeted optical control of neural activity. Nat Neurosci. 2005;8:1263–1268. doi: 10.1038/nn1525. [DOI] [PubMed] [Google Scholar]
  • 4.Herman AM, Huang L, Murphey DK, Garcia I, Arenkiel BR. Cell type-specific and time-dependent light exposure contribute to silencing in neurons expressing Channelrhodopsin-2. Elife. 2013;3:e01481. doi: 10.7554/eLife.01481. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Petreanu L, Huber D, Sobczyk A, Svoboda K. Channelrhodopsin-2-assisted circuit mapping of long-range callosal projections. Nat Neurosci. 2007;10:663–668. doi: 10.1038/nn1891. [DOI] [PubMed] [Google Scholar]
  • 6.Bernstein JG, Boyden ES. Optogenetic tools for analyzing the neural circuits of behavior. Trends Cogn Sci. 2011;15:592–600. doi: 10.1016/j.tics.2011.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Merighi A, Salio C, Ferrini F, Lossi L. Neuromodulatory function of neuropeptides in the normal CNS. J Chem Neuroanat. 2011;42:276–287. doi: 10.1016/j.jchemneu.2011.02.001. [DOI] [PubMed] [Google Scholar]
  • 8.Salio C, Lossi L, Ferrini F, Merighi A. Neuropeptides as synaptic transmitters. Cell Tissue Res. 2006;326:583–598. doi: 10.1007/s00441-006-0268-3. [DOI] [PubMed] [Google Scholar]
  • **9.van den Pol AN. Neuropeptide transmission in brain circuits. Neuron. 2012;76:98–115. doi: 10.1016/j.neuron.2012.09.014. [The author provides an important comparison between amino-acid and peptidergic transmissions. The paper is also a thorough review of the role of neuropeptides in the control of energy homeostasis, drug addiction, mood and motivation, sleep-wake states, and neuroendocrine regulation.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Ghijsen WE, Leenders AG. Differential signaling in presynaptic neurotransmitter release. Cell Mol Life Sci. 2005;62:937–954. doi: 10.1007/s00018-004-4525-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Merighi A. Costorage and coexistence of neuropeptides in the mammalian CNS. Prog Neurobiol. 2002;66:161–190. doi: 10.1016/s0301-0082(01)00031-4. [DOI] [PubMed] [Google Scholar]
  • 12.Vizi ES. Role of high-affinity receptors and membrane transporters in nonsynaptic communication and drug action in the central nervous system. Pharmacol Rev. 2000;52:63–89. [PubMed] [Google Scholar]
  • **13.Jego S, Glasgow SD, Herrera CG, Ekstrand M, Reed SJ, Boyce R, Friedman J, Burdakov D, Adamantidis AR. Optogenetic identification of a rapid eye movement sleep modulatory circuit in the hypothalamus. Nat Neurosci. 2013;16:1637–1643. doi: 10.1038/nn.3522. [This is the first direct evidence of the involvement of MCH neurons in REM sleep. The authors employ optogenetic activation and inhibition of MCH neurons to alter REM sleep. They also show that stimulation of MCH cell bodies or MCH terminals in the tuberomammillary nucleus and medial septum, but not in the dorsal raphe, prolongs REM sleep episodes.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Adamantidis AR, Zhang F, Aravanis AM, Deisseroth K, de Lecea L. Neural substrates of awakening probed with optogenetic control of hypocretin neurons. Nature. 2007;450:420–424. doi: 10.1038/nature06310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *15.Atasoy D, Betley JN, Su HH, Sternson SM. Deconstruction of a neural circuit for hunger. Nature. 2012;488:172–177. doi: 10.1038/nature11270. [Elegant study using in vivo and in vitro optogenetic approaches to map the circuit for food seeking and consumption. Importantly the authors use a combination of optogenetic activation of terminals with inhibition of cell bodies to eliminate any risk of back propagation of the excitatory stimulus to the cell bodies and from there to other synaptic outputs.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Walsh JJ, Friedman AK, Sun H, Heller EA, Ku SM, Juarez B, Burnham VL, Mazei-Robison MS, Ferguson D, Golden SA, et al. Stress and CRF gate neural activation of BDNF in the mesolimbic reward pathway. Nat Neurosci. 2014;17:27–29. doi: 10.1038/nn.3591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Pfeffer CK, Xue M, He M, Huang ZJ, Scanziani M. Inhibition of inhibition in visual cortex: the logic of connections between molecularly distinct interneurons. Nat Neurosci. 2013;16:1068–1076. doi: 10.1038/nn.3446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Carter ME, Brill J, Bonnavion P, Huguenard JR, Huerta R, de Lecea L. Mechanism for Hypocretin-mediated sleep-to-wake transitions. Proc Natl Acad Sci U S A. 2012;109:E2635–2644. doi: 10.1073/pnas.1202526109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kempadoo KA, Tourino C, Cho SL, Magnani F, Leinninger GM, Stuber GD, Zhang F, Myers MG, Deisseroth K, de Lecea L, et al. Hypothalamic neurotensin projections promote reward by enhancing glutamate transmission in the VTA. J Neurosci. 2013;33:7618–7626. doi: 10.1523/JNEUROSCI.2588-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Schone C, Cao ZF, Apergis-Schoute J, Adamantidis A, Sakurai T, Burdakov D. Optogenetic probing of fast glutamatergic transmission from hypocretin/orexin to histamine neurons in situ. J Neurosci. 2012;32:12437–12443. doi: 10.1523/JNEUROSCI.0706-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Carter ME, Soden ME, Zweifel LS, Palmiter RD. Genetic identification of a neural circuit that suppresses appetite. Nature. 2013;503:111–114. doi: 10.1038/nature12596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Bartfai T, Iverfeldt K, Fisone G, Serfozo P. Regulation of the release of coexisting neurotransmitters. Annu Rev Pharmacol Toxicol. 1988;28:285–310. doi: 10.1146/annurev.pa.28.040188.001441. [DOI] [PubMed] [Google Scholar]
  • 23.Tallent MK. Presynaptic inhibition of glutamate release by neuropeptides: use-dependent synaptic modification. Results Probl Cell Differ. 2008;44:177–200. doi: 10.1007/400_2007_037. [DOI] [PubMed] [Google Scholar]
  • 24.De Camilli P, Jahn R. Pathways to regulated exocytosis in neurons. Annu Rev Physiol. 1990;52:625–645. doi: 10.1146/annurev.ph.52.030190.003205. [DOI] [PubMed] [Google Scholar]
  • 25.Leenders AG, Scholten G, Wiegant VM, Da Silva FH, Ghijsen WE. Activity-dependent neurotransmitter release kinetics: correlation with changes in morphological distributions of small and large vesicles in central nerve terminals. Eur J Neurosci. 1999;11:4269–4277. doi: 10.1046/j.1460-9568.1999.00865.x. [DOI] [PubMed] [Google Scholar]
  • 26.Sudhof TC. The presynaptic active zone. Neuron. 2012;75:11–25. doi: 10.1016/j.neuron.2012.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bruns D, Jahn R. Real-time measurement of transmitter release from single synaptic vesicles. Nature. 1995;377:62–65. doi: 10.1038/377062a0. [DOI] [PubMed] [Google Scholar]
  • 28.Martin TF. Tuning exocytosis for speed: fast and slow modes. Biochim Biophys Acta. 2003;1641:157–165. doi: 10.1016/s0167-4889(03)00093-4. [DOI] [PubMed] [Google Scholar]
  • 29.Weisskopf MG, Zalutsky RA, Nicoll RA. The opioid peptide dynorphin mediates heterosynaptic depression of hippocampal mossy fibre synapses and modulates long-term potentiation. Nature. 1993;362:423–427. doi: 10.1038/362423a0. [DOI] [PubMed] [Google Scholar]
  • 30.Muschol M, Salzberg BM. Dependence of transient and residual calcium dynamics on action-potential patterning during neuropeptide secretion. J Neurosci. 2000;20:6773–6780. doi: 10.1523/JNEUROSCI.20-18-06773.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *31.Carter ME, Yizhar O, Chikahisa S, Nguyen H, Adamantidis A, Nishino S, Deisseroth K, de Lecea L. Tuning arousal with optogenetic modulation of locus coeruleus neurons. Nat Neurosci. 2010;13:1526–1533. doi: 10.1038/nn.2682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **32.Schone C, Apergis-Schoute J, Sakurai T, Adamantidis A, Burdakov D. Coreleased orexin and glutamate evoke nonredundant spike outputs and computations in histamine neurons. Cell Rep. 2014;7:697–704. doi: 10.1016/j.celrep.2014.03.055. [This is the first clear demonstration of optogenetic release of peptides in brain slices. It's also an elegant paper showing that the glutamate-dependent transmission is rapid but transient, whereas orexin-mediated response has a later onset but is longer lasting. No cross-modulation between glutamate and orexin transmission was found.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Wietek J, Wiegert JS, Adeishvili N, Schneider F, Watanabe H, Tsunoda SP, Vogt A, Elstner M, Oertner TG, Hegemann P. Conversion of channelrhodopsin into a light-gated chloride channel. Science. 2014;344:409–412. doi: 10.1126/science.1249375. [DOI] [PubMed] [Google Scholar]
  • **34.Mattis J, Tye KM, Ferenczi EA, Ramakrishnan C, O'Shea DJ, Prakash R, Gunaydin LA, Hyun M, Fenno LE, Gradinaru V, et al. Principles for applying optogenetic tools derived from direct comparative analysis of microbial opsins. Nat Methods. 2012;9:159–172. doi: 10.1038/nmeth.1808. [This is a comprehensive review covering depolarizing and hyperpolarizing opsins. It includes an extensive comparative analysis of the properties and performances of different optogenetic tools under matched experimental conditions.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Chow BY, Han X, Dobry AS, Qian X, Chuong AS, Li M, Henninger MA, Belfort GM, Lin Y, Monahan PE, et al. High-performance genetically targetable optical neural silencing by light-driven proton pumps. Nature. 2010;463:98–102. doi: 10.1038/nature08652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Hayashi S. Biophysics. Silencing neurons with light. Science. 2014;344:369–370. doi: 10.1126/science.1253616. [DOI] [PubMed] [Google Scholar]
  • 37.Stefanik MT, Kalivas PW. Optogenetic dissection of basolateral amygdala projections during cue-induced reinstatement of cocaine seeking. Front Behav Neurosci. 2013;7:213. doi: 10.3389/fnbeh.2013.00213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Tsunematsu T, Tabuchi S, Tanaka KF, Boyden ES, Tominaga M, Yamanaka A. Long-lasting silencing of orexin/hypocretin neurons using archaerhodopsin induces slow-wave sleep in mice. Behav Brain Res. 2013;255:64–74. doi: 10.1016/j.bbr.2013.05.021. [DOI] [PubMed] [Google Scholar]
  • 39.Berglind F, Ledri M, Sorensen AT, Nikitidou L, Melis M, Bielefeld P, Kirik D, Deisseroth K, Andersson M, Kokaia M. Optogenetic inhibition of chemically induced hypersynchronized bursting in mice. Neurobiol Dis. 2014;65:133–141. doi: 10.1016/j.nbd.2014.01.015. [DOI] [PubMed] [Google Scholar]
  • 40.Ilango A, Kesner AJ, Keller KL, Stuber GD, Bonci A, Ikemoto S. Similar roles of substantia nigra and ventral tegmental dopamine neurons in reward and aversion. J Neurosci. 2014;34:817–822. doi: 10.1523/JNEUROSCI.1703-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Stubblefield EA, Costabile JD, Felsen G. Optogenetic investigation of the role of the superior colliculus in orienting movements. Behav Brain Res. 2013;255:55–63. doi: 10.1016/j.bbr.2013.04.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Hull C, Adesnik H, Scanziani M. Neocortical disynaptic inhibition requires somatodendritic integration in interneurons. J Neurosci. 2009;29:8991–8995. doi: 10.1523/JNEUROSCI.5717-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Petreanu L, Mao T, Sternson SM, Svoboda K. The subcellular organization of neocortical excitatory connections. Nature. 2009;457:1142–1145. doi: 10.1038/nature07709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Adamantidis A, Carter MC, de Lecea L. Optogenetic deconstruction of sleep-wake circuitry in the brain. Front Mol Neurosci. 2010;2:31. doi: 10.3389/neuro.02.031.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *45.de Lecea L, Carter ME, Adamantidis A. Shining light on wakefulness and arousal. Biol Psychiatry. 2012;71:1046–1052. doi: 10.1016/j.biopsych.2012.01.032. [This is an overview of recent results involving optogenetic approaches to investigate the complex dynamics of neural circuits controlling arousal and arousal-related behaviors.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Carter ME, de Lecea L, Adamantidis A. Functional wiring of hypocretin and LC-NE neurons: implications for arousal. Front Behav Neurosci. 2013;7:43. doi: 10.3389/fnbeh.2013.00043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Sakurai T, Amemiya A, Ishii M, Matsuzaki I, Chemelli RM, Tanaka H, Williams SC, Richardson JA, Kozlowski GP, Wilson S, et al. Orexins and orexin receptors: a family of hypothalamic neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell. 1998;92:573–585. doi: 10.1016/s0092-8674(00)80949-6. [DOI] [PubMed] [Google Scholar]
  • 48.de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE, Fukuhara C, Battenberg EL, Gautvik VT, Bartlett FS, 2nd, et al. The hypocretins: hypothalamus-specific peptides with neuroexcitatory activity. Proc Natl Acad Sci U S A. 1998;95:322–327. doi: 10.1073/pnas.95.1.322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Li J, Hu Z, de Lecea L. The hypocretins/orexins: integrators of multiple physiological functions. Br J Pharmacol. 2014;171:332–350. doi: 10.1111/bph.12415. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Espana RA, Scammell TE. Sleep neurobiology from a clinical perspective. Sleep. 2011;34:845–858. doi: 10.5665/SLEEP.1112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Saper CB, Chou TC, Scammell TE. The sleep switch: hypothalamic control of sleep and wakefulness. Trends Neurosci. 2001;24:726–731. doi: 10.1016/s0166-2236(00)02002-6. [DOI] [PubMed] [Google Scholar]
  • 52.Tsujino N, Sakurai T. Orexin/hypocretin: a neuropeptide at the interface of sleep, energy homeostasis, and reward system. Pharmacol Rev. 2009;61:162–176. doi: 10.1124/pr.109.001321. [DOI] [PubMed] [Google Scholar]
  • 53.Torrealba F, Yanagisawa M, Saper CB. Colocalization of orexin a and glutamate immunoreactivity in axon terminals in the tuberomammillary nucleus in rats. Neuroscience. 2003;119:1033–1044. doi: 10.1016/s0306-4522(03)00238-0. [DOI] [PubMed] [Google Scholar]
  • 54.Abrahamson EE, Leak RK, Moore RY. The suprachiasmatic nucleus projects to posterior hypothalamic arousal systems. Neuroreport. 2001;12:435–440. doi: 10.1097/00001756-200102120-00048. [DOI] [PubMed] [Google Scholar]
  • 55.Chou TC, Lee CE, Lu J, Elmquist JK, Hara J, Willie JT, Beuckmann CT, Chemelli RM, Sakurai T, Yanagisawa M, et al. Orexin (hypocretin) neurons contain dynorphin. J Neurosci. 2001;21:RC168. doi: 10.1523/JNEUROSCI.21-19-j0003.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Crocker A, Espana RA, Papadopoulou M, Saper CB, Faraco J, Sakurai T, Honda M, Mignot E, Scammell TE. Concomitant loss of dynorphin, NARP, and orexin in narcolepsy. Neurology. 2005;65:1184–1188. doi: 10.1212/01.wnl.0000168173.71940.ab. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Muschamp JW, Hollander JA, Thompson JL, Voren G, Hassinger LC, Onvani S, Kamenecka TM, Borgland SL, Kenny PJ, Carlezon WA., Jr. Hypocretin (orexin) facilitates reward by attenuating the antireward effects of its cotransmitter dynorphin in ventral tegmental area. Proc Natl Acad Sci U S A. 111:E1648–1655. doi: 10.1073/pnas.1315542111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Eriksson KS, Sergeeva O, Brown RE, Haas HL. Orexin/hypocretin excites the histaminergic neurons of the tuberomammillary nucleus. J Neurosci. 2001;21:9273–9279. doi: 10.1523/JNEUROSCI.21-23-09273.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.van den Top M, Lee K, Whyment AD, Blanks AM, Spanswick D. Orexigen-sensitive NPY/AgRP pacemaker neurons in the hypothalamic arcuate nucleus. Nat Neurosci. 2004;7:493–494. doi: 10.1038/nn1226. [DOI] [PubMed] [Google Scholar]
  • 60.Acuna-Goycolea C, van den Pol AN. Peptide YY(3-36) inhibits both anorexigenic proopiomelanocortin and orexigenic neuropeptide Y neurons: implications for hypothalamic regulation of energy homeostasis. J Neurosci. 2005;25:10510–10519. doi: 10.1523/JNEUROSCI.2552-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Eriksson KS, Sergeeva OA, Selbach O, Haas HL. Orexin (hypocretin)/dynorphin neurons control GABAergic inputs to tuberomammillary neurons. Eur J Neurosci. 2004;19:1278–1284. doi: 10.1111/j.1460-9568.2004.03243.x. [DOI] [PubMed] [Google Scholar]
  • 62.Li Y, van den Pol AN. Differential target-dependent actions of coexpressed inhibitory dynorphin and excitatory hypocretin/orexin neuropeptides. J Neurosci. 2006;26:13037–13047. doi: 10.1523/JNEUROSCI.3380-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Iremonger KJ, Bains JS. Retrograde opioid signaling regulates glutamatergic transmission in the hypothalamus. J Neurosci. 2009;29:7349–7358. doi: 10.1523/JNEUROSCI.0381-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Marcus JN, Aschkenasi CJ, Lee CE, Chemelli RM, Saper CB, Yanagisawa M, Elmquist JK. Differential expression of orexin receptors 1 and 2 in the rat brain. J Comp Neurol. 2001;435:6–25. doi: 10.1002/cne.1190. [DOI] [PubMed] [Google Scholar]
  • 65.Sears RM, Fink AE, Wigestrand MB, Farb CR, de Lecea L, Ledoux JE. Orexin/hypocretin system modulates amygdala-dependent threat learning through the locus coeruleus. Proc Natl Acad Sci U S A. 2013;110:20260–20265. doi: 10.1073/pnas.1320325110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Lee MG, Hassani OK, Jones BE. Discharge of identified orexin/hypocretin neurons across the sleep-waking cycle. J Neurosci. 2005;25:6716–6720. doi: 10.1523/JNEUROSCI.1887-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Mileykovskiy BY, Kiyashchenko LI, Siegel JM. Behavioral correlates of activity in identified hypocretin/orexin neurons. Neuron. 2005;46:787–798. doi: 10.1016/j.neuron.2005.04.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Behn CG, Brown EN, Scammell TE, Kopell NJ. Mathematical model of network dynamics governing mouse sleep-wake behavior. J Neurophysiol. 2007;97:3828–3840. doi: 10.1152/jn.01184.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Konadhode RR, Pelluru D, Blanco-Centurion C, Zayachkivsky A, Liu M, Uhde T, Glen WB, Jr., van den Pol AN, Mulholland PJ, Shiromani PJ. Optogenetic stimulation of MCH neurons increases sleep. J Neurosci. 2013;33:10257–10263. doi: 10.1523/JNEUROSCI.1225-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Tsunematsu T, Ueno T, Tabuchi S, Inutsuka A, Tanaka KF, Hasuwa H, Kilduff TS, Terao A, Yamanaka A. Optogenetic manipulation of activity and temporally controlled cell-specific ablation reveal a role for MCH neurons in sleep/wake regulation. J Neurosci. 2014;34:6896–6909. doi: 10.1523/JNEUROSCI.5344-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Pissios P, Bradley RL, Maratos-Flier E. Expanding the scales: The multiple roles of MCH in regulating energy balance and other biological functions. Endocr Rev. 2006;27:606–620. doi: 10.1210/er.2006-0021. [DOI] [PubMed] [Google Scholar]
  • 72.Hassani OK, Lee MG, Jones BE. Melanin-concentrating hormone neurons discharge in a reciprocal manner to orexin neurons across the sleep-wake cycle. Proc Natl Acad Sci U S A. 2009;106:2418–2422. doi: 10.1073/pnas.0811400106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Takahashi K, Lin JS, Sakai K. Characterization and mapping of sleep-waking specific neurons in the basal forebrain and preoptic hypothalamus in mice. Neuroscience. 2009;161:269–292. doi: 10.1016/j.neuroscience.2009.02.075. [DOI] [PubMed] [Google Scholar]
  • *74.Williams RH, Chee MJ, Kroeger D, Ferrari LL, Maratos-Flier E, Scammell TE, Arrigoni E. Optogenetic-mediated release of histamine reveals distal and autoregulatory mechanisms for controlling arousal. J Neurosci. 2014;34:6023–6029. doi: 10.1523/JNEUROSCI.4838-13.2014. [This is the first evidence of optogenetic release of monoamines in brain slices. The authors show that the release of histamine has more in common with the mechanisms of transmission of neuropeptides than with amino-acid neurotransmission.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Vila A, Satoh H, Rangel C, Mills SL, Hoshi H, O'Brien J, Marshak DR, Macleish PR, Marshak DW. Histamine receptors of cones and horizontal cells in Old World monkey retinas. J Comp Neurol. 2012;520:528–543. doi: 10.1002/cne.22731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Freeman WJ. Ndn, volume transmission, and self-organization in brain dynamics. J Integr Neurosci. 2005;4:407–421. doi: 10.1142/s0219635205000963. [DOI] [PubMed] [Google Scholar]
  • 77.Wada H, Inagaki N, Yamatodani A, Watanabe T. Is the histaminergic neuron system a regulatory center for whole-brain activity? Trends Neurosci. 1991;14:415–418. doi: 10.1016/0166-2236(91)90034-r. [DOI] [PubMed] [Google Scholar]

RESOURCES