Skip to main content
Erschienen in: Molecular Pain 1/2013

Open Access 01.12.2013 | Research

Differential modulation of nociceptive versus non-nociceptive synapses by endocannabinoids

verfasst von: Alexandra Higgins, Sharleen Yuan, Yanqing Wang, Brian D Burrell

Erschienen in: Molecular Pain | Ausgabe 1/2013

Abstract

Background

Although a number of clinical and preclinical studies have demonstrated analgesic effects of cannabinoid treatments, there are also instances when cannabinoids have had no effect or even exacerbated pain. The observed pro-nociceptive effects appear to be due to cannabinoid-induced disinhibition of afferent synaptic input to nociceptive circuits. To better understand how cannabinoid-mediated plasticity can have both pro- and anti-nociceptive effects, we examined the possibility that cannabinoids differentially modulate nociceptive vs. non-nociceptive synapses onto a shared postsynaptic target. These experiments were carried out in the central nervous system (CNS) of the medicinal leech, in which it is possible to intracellularly record from presynaptic nociceptive (N-cell) or pressure-sensitive (P-cell) neurons and their shared postsynaptic targets.

Results

The endocannabinoid 2-arachidonoyl glycerol (2AG) elicited significant long-lasting depression in nociceptive (N-cell) synapses. However, non-nociceptive (P-cell) synapses were potentiated following 2AG treatment. 2AG-induced potentiation of non-nociceptive synapses was blocked by the TRPV antagonist SB366791, suggesting involvement of the same TRPV-like receptor that has already been shown to mediate endocannabinoid-dependent depression in nociceptive inputs. Treatment with the GABA receptor antagonist bicuculline also blocked 2AG-induced potentiation, consistent with the idea that increased synaptic signaling was the result of endocannabinoid-mediated disinhibition. Interestingly, while bicuculline by itself increased non-nociceptive synaptic transmission, nociceptive synapses were depressed by this GABA receptor antagonist indicating that nociceptive synapses were actually excited by GABAergic input. Consistent with these observations, GABA application depolarized the nociceptive afferent and hyperpolarized the non-nociceptive afferent.

Conclusions

These findings show that endocannabinoids can differentially modulate nociceptive vs. non-nociceptive synapses and that GABAergic regulation of these synapses plays an important role in determining whether endocannabinoids have a potentiating or depressing effect.
Begleitmaterial
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​1744-8069-9-26) contains supplementary material, which is available to authorized users.
Alexandra Higgins, Sharleen Yuan contributed equally to this work.

Competing interests

The authors declare no competing interests.

Authors’ contributions

AH, SY and YW carried out the electrophysiological recordings and data analysis. BDB conceived of the study and participated in its design, coordination and data analysis. BDB and AH helped to draft the manuscript. All authors read and approved the final manuscript.

Background

Endogenous cannabinoids (endocannabinoids or eCBs), such as anandamide and 2-arachidonoyl glycerol (2AG), are lipid neurotransmitters found throughout the central nervous system of both vertebrates and invertebrates that modulate a number of behavioral processes including appetite, cognition, emotion, sensory processing and nociception [1, 2]. In many regions of the CNS, the primary physiological effect of eCBs is depression of synaptic transmission that can be short-term (milliseconds to seconds) or long-term (tens of minutes to hours) [3]. eCBs can depress both excitatory and inhibitory synapses, so that from a functional standpoint these neurotransmitters can be bi-directional modulators of neural circuits depending on whether there is depression of an excitatory pathway (decreasing circuit activity) or depression of an inhibitory pathway (increasing circuit activity via disinhibition). This is a significant consideration in terms of development of eCB-based pharmacotherapies because treatments focused on depression of excitatory synapses may produce undesired effects through depression of inhibitory synapses. The importance of this issue has become apparent in the potential use of cannabinoid-based treatments for chronic pain. In a recent study, cannabinoids were found to reduce GABAergic and glycinergic inhibitory signaling in the spinal cord that contributed to secondary hyperalgesia due to disinhibition of afferent synaptic input to spinal pain circuits [4]. However, it was not known whether this pro-algesic effect occurred at the nociceptive afferent synapses (Aδ and C fibers) or at non-nociceptive afferent synapses (Aβ fibers). Non-nociceptive afferents can have input to pain circuits in the spinal cord as a result of central sensitization and can potentially contribute to chronic pain conditions [5].
To address this issue, the CNS of the medicinal leech (Hirudo verbana) was used to test the effects of eCBs on nociceptive vs. non-nociceptive synapses. The leech is a useful model system in which to carry out these studies because it has a well-defined nervous system in which the identity of nociceptive and non-nociceptive neurons are known as well as their synaptic targets. The leech CNS consists of a chain of segmentally-arranged ganglia, each with its own set of sensory, motor and local and intersegmental interneurons. Similar to mechanosensation in mammals, the leech receives input from three distinct catergories of sensory neurons innervating the skin: rapidly-adapting touch cells (T-cells), slow-adapting pressure-sensitive neurons (P-cells) and two types of nociceptive neurons (N-cells), one that is a mechanical and the other a polymodal nociceptor that responds to mechanical, thermal and chemical stimuli [68]. The cell bodies for each type of somatosensory neuron are located in the segmental ganglia and can be easily identified based on their size, position within each ganglion (see Figure 1) and electrophysiological properties. Furthermore, a number of the postsynaptic targets of these afferents are known and in many cases are shared by all three types of sensory neurons [9]. Therefore, it is possible to selectively carry out dual intracellular recordings of both the pre- and postsynaptic neurons in these nociceptive vs. non-nociceptive synapses. This capacity to carry out detailed studies from identifiable neurons and their synapses has already been used to develop a detailed understanding the cellular signal mechanisms mediating eCB-dependent depression of nociceptive synapses in the leech [1012].
The leech CNS is known to contain the eCBs 2AG and anandamide with 2AG being the more abundant transmitter [14]. Furthermore, 2AG has been shown to mediate long-term depression (LTD) in leech nociceptive synapses that are remarkably similar to mammalian eCB-LTD in terms of cellular mechanisms [3, 10]. Like other protostomal invertebrates, the leech lacks orthologues to the vertebrate cannabinoid receptors CB1 and CB2 [2]. However, transient potential receptor vanilloid (TRPV) channels can also act as cannabinoid receptors [1518] and previous pharmacological studies from our laboratory indicate that a leech TRPV-like receptor mediates 2AG-induced depression in nociceptive synapses [10].
In the present study, the effects of 2AG on nociceptive synapses (those made by the N-cells) versus non-nociceptive synapses (those made by the P-cells) were compared in which both afferent types converged on the same postsynaptic target (the longitudinal or “L” motor neuron). 2AG elicited depression at nociceptive synapses, but potentiation at non-nociceptive synapses. This 2AG-mediated potentiation of the non-nociceptive synapses required functional GABA receptors, indicating eCB-induced increases in synaptic signaling were the result of synaptic disinhibition.

Results

Effects of 2AG on nociceptive vs. non-nociceptive synapses

For most of the experiments in this study, excitatory post-synaptic potentials (EPSPs) were recorded from synapses made by either the lateral (polymodal) nociceptive (N-) cell or one of the pressure-sensitive (P-) cells onto the L motor neuron (Figure 1). The L motor neuron innervates the longitudinal muscles and is active during the defensive withdrawal reflex, whole-body shortening [19]. This motor neuron receives monosynaptic, glutamatergic input from the both the P- and N-cells [10, 20]. Previous studies have shown that 2AG (60 μM for 15 mins) induced LTD in nociceptive synapses made by the lateral N-cell onto the L motor neuron (Figure 2A, B; data re-presented from [10]; t=6.45, p≤0.0001, n=5). In contrast, the non-nociceptive P-to-L synapse was actually potentiated 1 hr following bath application of 60–100 μM 2AG for 15 min (N=10) relative to control synapses in which 2AG application was omitted (Figure 2A, B; t=3.37, p≤0.005, n=10). Potentiation did not appear to be due to postsynaptic changes in intrinsic excitability given that input resistance (IR) in the L motor neuron was unchanged (post-test IR was 100.6±2.9% of pre-test levels in the 2AG-treated group and 103.2±3.7% in the vehicle control group). The effects of 60 and 100 μM 2AG on the P-to-L synapse were indistinguishable with 60 μM 2AG producing a 48.2±10.4% increase (n=4) and 100 μM producing a 56.1±22.1% (n=6) increase in EPSP amplitude (independent t-test p=0.79); therefore the data from these two concentrations were combined. N-to-L synapses treated with 100 μM 2AG still underwent depression (64.6±10.3% of initial EPSP amplitude; n=3) that was no different from the effects observed at 60 μM (63±4%; independent t-test p=0.87). These results indicate that the potentiating effect of 100 μM 2AG on the non-nociceptive synapse was not a consequence of the concentration of 2AG used.
To determine if these opposing effects of 2AG were a feature of other nociceptive and non-nociceptive synapses, N- and P-cell input to the anterior pagoda (AP) cell was also examined. Although the function of the AP cell is not known, the monosynaptic, glutamatergic P-to-AP synapses have been used to study a variety of forms of synaptic plasticity in the leech [2125]. However, the N-to-AP synapse had not been previously characterized and this was done prior to conducting the 2AG experiments. In general the N-to-AP synapse was quite similar to the N-to-L connection. N-to-AP was substantially reduced by CNQX (20 μM) indicating glutamatergic transmission (Additional file 1: Figure S1A). In experiments with high MgCl2 saline (15 mM), which blocked all chemical synaptic transmission [20], the majority of the N-to-AP EPSP was eliminated except for a small (<1 mV) early component that appeared to be a rectifying electrical synapse between the N-and AP-cell (Additional file 1: Figure S1B). This putative electrical EPSP was also observed in the CNQX experiments. Experiments with high divalent saline (18 mM MgCl2/15 mM CaCl2) eliminated later components of the N-to-AP EPSP indicating the presence of polysynaptic elements (Additional file 1: Figure S1C). While positive current associated with the N-cell action potential did propagate to the AP via electrical coupling, negative current injected into the N-cell did not spread to the AP (data not shown). Interestingly, evidence of electrical coupling was also observed in the AP-to-N direction. Injection of 500 msec current pulses showed that negative current, but not positive current, could spread from the AP to the N-cell (Additional file 1: Figure S1D).
In terms of eCB modulation, 2AG (100 μM) depressed N-to-AP synapses (n=9 and 5 respectively; t=2.91, p≤0.05), but potentiated P-to-AP synapses relative to controls (n=6 and 5 respectively; t=2.36, p≤0.05) identical to N- and P-cell inputs to the L motor neuron (Figure 2C). No obvious changes in postsynaptic IR were observed during the P-to-AP (2AG group=111.5±3.2%, control=106.8±5.6) or N-to-AP (2AG group=104.6±3.8%, control=97.7±6.2%) synapses.
To test whether 2AG-induced potentiation of the non-nociceptive P-cell synapses is mediated by a TRPV-like receptor, experiments were carried out using the selective TRPV1 antagonist SB366791 (10 μM), which blocks eCB-mediated LTD in leech nociceptive synapses [10]. When SB366791 was co-applied with 2AG, potentiation of the non-nociceptive P-to-L synapse was prevented (Figure 3A; n=6). Application of SB366791 alone had no effect on the P-to-L EPSP (N=5). One-way ANOVA of the 2AG+SB366791, SB366791, 2AG and vehicle control groups showed a significant treatment effect (F2,18=6.01, p≤0.05). Post-hoc analysis showed that the 2AG group was statistically different from the control group (p≤0.05), but that the 2AG+SB366791 and SB366791 groups were not. Postsynaptic IR was unchanged in the 2AG+SB366791 (110.1±4.6% of pre-test levels), SB366791-only (101.7±5.6%) and vehicle control groups (101±3.9%; n=6). These findings indicate that the difference in the effects of 2AG on nociceptive vs. non-nociceptive synapses is not due to activation of different endocannabinoid receptor.

Role of GABA receptors during 2AG-induced potentiation

eCBs are known to depress inhibitory synaptic transmission and eCB-induced disinhibition has been shown to increase synaptic signaling within spinal nociceptive circuits, although the identity of the presynaptic input(s) being enhanced was not known [4]. Therefore, we examined the potential role of GABA-A receptors during 2AG-induced potentiation of the non-nociceptive synapse in the leech. The leech CNS contains GABAergic neurons and the GABA-A receptor antagonist bicuculline has been shown to disinhibit other circuits that utilize P-cell input [2628]. To test whether functional GABA-A receptors are required for 2AG-mediated potentiation, ganglia were pre-treated with bicuculline (100 μM) followed by the pre-test measurements of the P-to-L EPSP, 15 min application of 2AG, washout for 60 mins and finally post-test EPSP measurements. The 2AG treatment that normally produces potentiation was blocked in the non-nociceptive P-to-L synapses that had undergone bicuculline+2AG treatment (Figure 3B; n=5). P-to-L synapses that were pre-treated with bicuculline, but subsequent 2AG treatment omitted, were unchanged between the pre- and post-tests (Figure 3B, n=5). One-way ANOVA of the 2AG+bicuculline, bicuculline, 2AG and vehicle control group showed a significant effect of treatment (F2,24=7.13, p≤0.005). Post-hoc analysis showed that the 2AG group was statistically different from the control group (p<0.01) and that the 2AG+bicuculline and bicuculline groups were not. No change in postsynaptic input resistance was observed in the 2AG+bicuculline (104±4.4%) or bicuculline only group (96±4.6%). These findings support the hypothesis that 2AG-induced potentiation of non-nociceptive synapses is mediated by disinhibition as a result of a decrease in GABAergic input.
If 2AG potentiates non-nociceptive synapses due to a decrease in GABAergic input, why are nociceptive synapses not similarly disinhibited? GABA has been observed to depolarize the lateral N-cells used in these experiments [29] so it is possible that GABA excites rather than inhibits the nociceptive synapse. To confirm this earlier finding and to determine what effect GABA has on the P-cell, intracellular recordings of the N- and P-cells were made in response to GABA delivered via puffer electrodes. GABA reliably elicited depolarization in the lateral N-cells, consistent with earlier findings, but produced hyperpolarization in the P-cells (Figure 4A). Both the N-cell (n=6) and P-cell (n=3) responses were significantly inhibited following 15 min bath-application of 100 μM bicuculline (Figure 4A; N-cell, paired t-test t=8.80, p≤0.001; P-cell, t=4.56, p≤0.05). No changes in N-cell (n=5) or P-cell (n=5) GABA response were observed in control experiments in which responses were recorded prior to and then 15 mins following treatment with normal saline (Figure 4A; N-cell, paired t-test t=1.65, p>0.05; P-cell, t=0.81, p>0.05).
Next, P-to-L and N-to-L EPSPs were recorded prior to and then during acute bicuculline treatment (100 μM for 10–15 mins). As expected, the P-to-L EPSP amplitude increased during bicuculline treatment (n=7) compared to control synapses (n=5); that is, disinhibition of the non-nociceptive synapse was observed (Figure 4B; t=2.71, p≤0.05). However, bicuculline treatment actually decreased the amplitude of the N-to-L EPSP (n=8) compared to control synapses (n=5) suggesting that GABA has a tonic excitatory effect on this nociceptive synapse (Figure 4B; t=2.35, p≤0.05). No change in postsynaptic input resistance was observed during either the P-to-L synapse (bicuculline=96.7±4.9, control=105.7±6.4%) or N-to-L synapse (bicuculline=107.7±2.2%, control=98.7±3%) experiments. These findings are consistent with the idea that this nociceptive synapse does not undergo 2AG-induced disinhibition, at least in part, because the presynaptic neuron is excited, not inhibited, by GABA.

Discussion

The endocannabinoid 2AG was found to differentially modulate nociceptive vs. non-nociceptive synapses in the leech, depressing the former and potentiating the latter. 2AG-induced potentiation of non-nociceptive synapses was blocked by the TRPV1 inhibitor SB366791, suggesting that the involvement of a TRPV-like receptor. In mammals, the TRPV1 receptor has been shown to be an important cannabinoid receptor that mediates synaptic depression in a variety of brain regions [1618, 30]. Protostomal invertebrates lack orthologues to the vertebrate CB1/CB2 receptors [2] and we have previously proposed that TRP channels, such as TRPV, may represent the earliest endocannabinoid receptors in the animal kingdom [10, 11]. One critical difference between the nociceptive and non-nociceptive synapses is that while N-cells appear to possess TRPV-like receptors, P-cells do not [8]. This would indicate that the effects of 2AG on P-cell synapses are indirect and mediated by an unknown, 2AG-sensitive neuron. It is unlikely that the postsynaptic L motor neuron is being directly modulated by 2AG because it too lacks TRPV-like receptors, although it does appear to be capable of synthesizing 2AG [10].
Potentiation of non-nociceptive synapses was blocked when the GABA-A receptor antagonist bicuculline was applied. Bicuculline by itself was found to enhance non-nociceptive synaptic transmission indicating that this synapse is tonically regulated by inhibitory GABAergic input, consistent with previous studies of P-cell signaling [26, 27]. From these results, we propose a model in which 2AG depresses inhibitory synaptic transmission from an unknown GABAergic neuron(s) onto P-cell synapses resulting in disinhibition of these non-nociceptive synapses (see Figure 5). When GABAergic input was reduced by bicuculline pre-treatment, the ability of 2AG to elicit potentiation via this proposed disinhibition was occluded. It should be noted that the specificity of both SB366791 and bicuculline as antagonists has not been directly tested on isolated leech versions of the TRPV and GABA receptors. Therefore, the possibility exists that the effects of both of these pharmacological agents are due to actions not associated with either of these receptors.
The fact that 2AG did not potentiate nociceptive synapses appears to be due to differences in how GABA affects nociceptive vs. non-nociceptive synapses (Figure 5). The same bicuculline treatment that disinhibited non-nociceptive synapses actually decreased N-cell synaptic transmission suggesting that GABA has an excitatory effect on these nociceptive synapses. This is supported by the observation that GABA applied to the soma depolarizes the lateral N-cell [29]. In the present study, the ability of GABA to depolarize the N-cell was replicated and extended by showing that this GABA-induced depolarization could be inhibited by bicuculline. Furthermore, direct application of GABA to the P-cell was found to elicit hyperpolarization that was also inhibited by bicuculline. The excitatory effects of GABA on the N-cell are likely the result of elevated intracellular Cl- concentration so that the Cl- equilibrium potential is depolarized relative to the resting potential. Thus, activation of ionotropic GABA receptors would cause a Cl- efflux, depolarizing the N-cell. It well-established that GABA can depolarize nociceptive afferents in the spinal cord as a result of elevated intracellular Cl- concentrations/Cl- equilibrium potential (ECl) [31]. GABA-mediated depolarization has been shown to increase synaptic transmission in both the brain and the spinal cord circuits [3235], although there are also instances when GABA-mediated depolarization inhibits synaptic transmission via current shunts or Na+ channel inactivation [31, 35].
It is not clear whether GABAergic modulation of the P-to-L synapse is being exerted on the presynaptic or postsynaptic neuron (or both). Leech motor neurons do receive inhibitory GABAergic input [26, 36], but no changes in postsynaptic input resistance were observed during any of the 2AG experiments, which would argue against a postsynaptic mechanism. Furthermore, the fact that N- and P-cell synapses onto the same postsynaptic target (L motor neuron or AP cell), but underwent opposite changes following 2AG treatment would also support a presynaptic mechanism for both eCB-induced potentiation and depression. Previous studies have been supportive of a presynaptic mechanism mediating 2AG-induced depression of nociceptive (N-to-L) synapse [10, 11]. Nevertheless, the possibility of localized, synapse-specific changes in the postsynaptic neuron cannot be eliminated and additional studies are required to further assess the role of pre- vs. postsynaptic modulation during eCB-mediated synaptic plasticity.
The results from this study address important issues regarding the modulatory effects of eCB at the microcircuit level. It is well-established that eCBs depress inhibitory (as well as excitatory) synaptic transmission and that such eCB-mediated disinhibition is likely to have important consequences in terms of neural circuit function [16, 3742]. For example, eCB-induced disinhibition has been observed to lower the threshold for initiating LTP [4345]. However, this is the first time, to our knowledge, that eCB-induced increases in evoked synaptic transmission via disinhibition have been directly observed. Furthermore, by using the well-described CNS of the leech it was possible to observe opposing synaptic effect of eCBs (depression vs. potentiation) on distinct afferent inputs that converge onto a shared postsynaptic target. This allowed us to address the issue as to why eCB-induced disinhibition does not lead to increased synaptic signaling in more neurons given the widespread nature of GABAergic regulation of synaptic transmission. We found evidence that while some synapses were sensitive to eCB-mediated disinhibition; others were “protected” from such modulation because they were not inhibited by GABA. In fact, GABA appeared to have an excitatory effect on these synapses based on results from experiments in this study and those from Sargeant et al. [29], presumably as a result of elevated intracellular Cl- levels. This would suggest that the Cl- gradient in neurons plays a critical in regulating the effects of eCBs on synaptic signaling. Such a mechanism has already been suggested by Christie and Mallet [46].
These findings may contribute to understanding why there is such conflicting evidence from both animal model and clinical studies regarding the efficacy of cannabinoid based analgesic therapies [4650]. In some clinical studies, cannabinoids were found to be ineffective and could even enhanced pain associated with evoked mechanical stimuli resulting in mechanical hyperalgesia or allodynia [4, 46, 51]. One explanation for these pro-nociceptive effects is that eCBs have been shown to depress inhibitory synapses in the spinal cord, thereby disinhibiting spinal pain circuits, which contributes to the development of secondary mechanical hyperalgesia [4]. It is possible that eCBs selectively enhance synaptic input from afferents that contribute to mechanical hyperalgesia or allodynia, such as Aβ fibers [5], but depress synaptic input from nociceptive afferents, such as Aδ or C fibers [50]. As stated above, these synaptic potentiation vs. depression effects of eCBs may be due to whether GABA inhibits or excites the relevant afferent [46]. Although the present study is focused on changes in afferent signaling, inhibitory and excitatory interneurons in the dorsal horn can also undergo shifts in the ECl that lead to GABA-elicited depolarization during neuropathic pain [52, 53]. Changes in GABAergic signaling onto these neurons may also impact how cannabinoid-based treatments affect these nociceptive neural circuits.

Conclusions

Nociception is a fundamental sensory process that exhibits considerable evolutionary conservation between vertebrates and invertebrates [54, 55]. The leech, in particular, provides a useful model system in which to study the basic physiological processes related to nociception. It possesses afferents innervating the skin that demonstrate clear nociceptive properties and share many features with vertebrate polymodal nociceptors including a high threshold for both mechanical and thermal (>40°C) stimuli as well as sensitivity to noxious chemical stimuli [8]. These nociceptors are clearly distinguishable from the lower threshold non-nociceptive mechanosensory neurons and it is possible to carry out detailed recordings from nociceptive vs. non-nociceptive synapses.
The findings from the present study suggest that the effectiveness of cannabinoid-based analgesic therapies is likely to depend on the type of nociception that is being experienced. As already stated, there is evidence from both preclinical and clinical studies that eCBs can have both pro- and anti-nociceptive effects [4, 46]. Interestingly, there is also evidence to suggest that central TRPV receptor activation can have opposing effects on spinal nociceptive circuits in rodents mediating both synaptic disinhibition that resulted in allodynia [56] and persistent depression of C fiber evoked EPSPs [5759]. Cannabinoid-based therapies may be appropriate for conditions that result from spontaneous activity from nociceptive afferents, such neuropathy-associated chronic pain [49, 60]. On the other hand, conditions that are dominated by mechanical hyperalgesia and/or allodynia may be insensitive to or even exacerbated by cannabinoid-based treatments due to the potential involvement of non-nociceptive afferents [5]. Additional studies using the leech CNS, at both the synaptic and behavioral level, may contribute to better understanding the pro- vs. anti-nociceptive effects of eCBs.

Methods

Animal preparation

Leeches (Hirudo verbana) were obtained from commercial suppliers (Leeches USA, Westbury, NY and Niagara Leeches, Cheyenne, WY) and maintained in artificial pond water (0.50g/L H2O Hirudo salt) on a 12 hour light/dark cycle at 18°C. Ganglia were dissected and pinned in a recording chamber with constant perfusion of normal leech saline (≈1.5 ml/min). All dissections and recordings were carried out in normal leech saline (110 mM NaCl, 4 mM KCl, 1.8 mM CaCl2, 1 mM MgCl2, 5 mM NaOH, and 10 mM HEPES, pH=7.4). Drugs were dissolved in leech saline from stock solutions and final concentrations were made just prior to respective experiments. The following drug was obtained from Tocris (Ellisville, MO): 2-arachidonoyl glycerol (2AG). Drugs obtained from Sigma-Aldrich (St. Louis, MO) included CNQX, dimethyl sulfoxide (DMSO), and bicuculline.

Electrophysiology

Techniques used in this study have been described in detail in [10]. Briefly, current clamp (bridge balanced) intracellular recordings were carried out using sharp glass microelectrodes (tip resistance 35–40 MΩ) made from borosilicate capillary tubing (1.0 mm OD, 0.75 mm ID; FHC, Bowdoinham, ME) using a horizontal puller (Sutter Instruments P-97; Novato, CA). Microelectrodes were filled with 3M potassium acetate. Manual micropositioners (Model 1480; Siskiyou Inc., Grants Pass, OR) were used to impale individual neurons during experiments. Current was delivered to electrodes using a multi-channel programmable stimulator (STG 1004; Multi-Channel Systems; Reutlingen, Germany) and the signal was recorded using a bridge amplifier (BA-1S; NPI, Tamm, Germany) and digitally converted for analysis (Axoscope; Molecular Devices, Sunnyvale, CA).
The presynaptic lateral nociceptive (N) and pressure (P) cells and the postsynaptic longitudinal (L) motor neuron and anterior pagoda (AP) cell were identified based on their position with the ganglion (Figure 1), size, and characteristic electrophysiological properties (size and shape of action potential). L motor neuron identification could be confirmed by recording from the electrically coupled contralateral L motor neurons and observing synchronous activity [61]. For experiments utilizing N-to-L and P-to-L synapse recordings, the ganglion was pinned dorsal side up so that the L motor neurons could be located on the dorsal side along with access to the lateral-most N- and P-cells. For N-to-AP and P-to-AP synapse recordings, the ganglion was pinned ventral side up. Following pre-test recordings of the excitatory postsynaptic potentials (EPSPs), the ganglion was superfused with 2AG for 15 minutes and then returned to normal saline. In vehicle control experiments, 2AG was replaced with saline containing 0.01% DMSO. After one hour, the EPSP was retested (post-test). Separate electrode impalements of the same presynaptic and postsynaptic neuron were made for pre- and post-test recordings. Chronic intracellular recordings of these neurons were not carried out because this results in progressive rundown of the EPSP within 10–15 mins most likely due to damage caused by movements of the tissue during the electrode impalement (there are muscle fibers and connective tissue present in the leech CNS). Input resistance was recorded at the pre- and post-test level and only consistent, stable recordings were included in the data analysis (see Results section). The peak EPSP amplitude was recorded every 10 seconds and calculated by averaging 5–10 EPSP (pre- or post-test) sweeps.
For N-to-AP characterization experiments, the ganglion was pinned ventral side up as the lateral N-cell and AP-cell were both located on the ventral side. Treatments for the set of characterization experiments (high Mg2+ saline, high divalent saline, and CNQX) were perfused for 10 to 15 minutes between pre-test and post-test recordings while controls were carried out in constant perfusion of normal leech saline. Coupling experiments were performed by injecting single 500 msec hyperpolarizing current pulses with increasing amounts of current into the AP- or N-cell and recording any subsequent response from the opposite cell.
The effects of GABA on the N- and P-cells were tested using a “puffer” electrode (a patch electrode with a resistance of ≈ 5 MΩ) connected to a picospritzer. The puffer electrode was positioned approximately 100 μm from the N or P soma to prevent movement artifact and a 200 msec pulse of GABA was applied at 10 psi. The membrane potential of the P-cell was maintained at −50 mV. The membrane potential of the N-cell was maintained at −60 mV in order to prevent the cell from firing when GABA was applied. The concentration of GABA within the puffer electrode was 1 M, consistent with similar studies in the leech by Sargent et al. [29]. This high concentration in the electrode was required due to the substantial amount of dilution that occurred by the time the expelled GABA reached the neuropil, which is relatively deep in the ganglion. That responses to GABA appear to arise from N- and P-cell processes in the neuropil is based on the delay between the GABA puff and the P- or N-cell response which ranged between 100–300 msec.
Post-test EPSP amplitudes and input resistance measurements were normalized relative to pre-test levels and presented as mean ± standard error. Statistical analyses using both independent and paired t-test or one-way analysis of variance (ANOVA) were performed to determine main effects with Newman-Keuls post-hoc tests to confirm the ANOVA results.

Acknowledgements

The authors thank the three anonymous reviewers whose comments significantly strengthened this manuscript. Supported by the National Science Foundation (IOS-1051734; BDB) and the National Institute of Neurological Disorders and Stroke (F31 NS074473; SY). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of Neurological Disorders And Stroke, the National Institutes of Health, and the National Science Foundation. Also supported by the University of South Dakota Council of Undergraduate Research and Creative Activities (CURCA) U.Discover Summer research fellowship (AH) and the Medical Student Summer Research program at the USD Sanford School of Medicine (AH).
Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://​creativecommons.​org/​licenses/​by/​2.​0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Competing interests

The authors declare no competing interests.

Authors’ contributions

AH, SY and YW carried out the electrophysiological recordings and data analysis. BDB conceived of the study and participated in its design, coordination and data analysis. BDB and AH helped to draft the manuscript. All authors read and approved the final manuscript.
Anhänge
Literatur
1.
Zurück zum Zitat Di Marzo V, Bifulco M, De Petrocellis L: The endocannabinoid system and its therapeutic exploitation. Nat Rev Drug Discov 2004, 3(9):771–784. 10.1038/nrd1495CrossRefPubMed Di Marzo V, Bifulco M, De Petrocellis L: The endocannabinoid system and its therapeutic exploitation. Nat Rev Drug Discov 2004, 3(9):771–784. 10.1038/nrd1495CrossRefPubMed
2.
Zurück zum Zitat Elphick MR: The evolution and comparative neurobiology of endocannabinoid signalling. Philos Trans R Soc Lond B Biol Sci 2012, 367(1607):3201–3215. 10.1098/rstb.2011.0394PubMedCentralCrossRefPubMed Elphick MR: The evolution and comparative neurobiology of endocannabinoid signalling. Philos Trans R Soc Lond B Biol Sci 2012, 367(1607):3201–3215. 10.1098/rstb.2011.0394PubMedCentralCrossRefPubMed
3.
Zurück zum Zitat Heifets BD, Castillo PE: Endocannabinoid signaling and long-term synaptic plasticity. Annu Rev Physiol 2009, 71: 283–306. 10.1146/annurev.physiol.010908.163149PubMedCentralCrossRefPubMed Heifets BD, Castillo PE: Endocannabinoid signaling and long-term synaptic plasticity. Annu Rev Physiol 2009, 71: 283–306. 10.1146/annurev.physiol.010908.163149PubMedCentralCrossRefPubMed
4.
Zurück zum Zitat Pernia-Andrade AJ, Kato A, Witschi R, Nyilas R, Katona I, Freund TF, Watanabe M, Filitz J, Koppert W, Schuttler J: Spinal endocannabinoids and CB1 receptors mediate C-fiber-induced heterosynaptic pain sensitization. Science 2009, 325(5941):760–764. 10.1126/science.1171870PubMedCentralCrossRefPubMed Pernia-Andrade AJ, Kato A, Witschi R, Nyilas R, Katona I, Freund TF, Watanabe M, Filitz J, Koppert W, Schuttler J: Spinal endocannabinoids and CB1 receptors mediate C-fiber-induced heterosynaptic pain sensitization. Science 2009, 325(5941):760–764. 10.1126/science.1171870PubMedCentralCrossRefPubMed
5.
Zurück zum Zitat Torsney C, MacDermott AB: Disinhibition opens the gate to pathological pain signaling in superficial neurokinin 1 receptor-expressing neurons in rat spinal cord. J Neurosci 2006, 26(6):1833–1843. 10.1523/JNEUROSCI.4584-05.2006CrossRefPubMed Torsney C, MacDermott AB: Disinhibition opens the gate to pathological pain signaling in superficial neurokinin 1 receptor-expressing neurons in rat spinal cord. J Neurosci 2006, 26(6):1833–1843. 10.1523/JNEUROSCI.4584-05.2006CrossRefPubMed
6.
Zurück zum Zitat Nicholls JG, Baylor DA: Specific modalities and receptive fields of sensory neurons in CNS of the leech. J Neurophysiol 1968, 31(5):740–756.PubMed Nicholls JG, Baylor DA: Specific modalities and receptive fields of sensory neurons in CNS of the leech. J Neurophysiol 1968, 31(5):740–756.PubMed
7.
Zurück zum Zitat Blackshaw SE, Nicholls JG, Parnas I: Physiological responses, receptive fields and terminal arborizations of nociceptive cells in the leech. J Physiol 1982, 326: 251–260.PubMedCentralCrossRefPubMed Blackshaw SE, Nicholls JG, Parnas I: Physiological responses, receptive fields and terminal arborizations of nociceptive cells in the leech. J Physiol 1982, 326: 251–260.PubMedCentralCrossRefPubMed
8.
Zurück zum Zitat Pastor J, Soria B, Belmonte C: Properties of the nociceptive neurons of the leech segmental ganglion. J Neurophysiol 1996, 75(6):2268–2279.PubMed Pastor J, Soria B, Belmonte C: Properties of the nociceptive neurons of the leech segmental ganglion. J Neurophysiol 1996, 75(6):2268–2279.PubMed
9.
Zurück zum Zitat Baylor DA, Nicholls JG: Chemical and electrical synaptic connexions between cutaneous mechanoreceptor neurones in the central nervous system of the leech. J Physiol 1969, 203(3):591–609.PubMedCentralCrossRefPubMed Baylor DA, Nicholls JG: Chemical and electrical synaptic connexions between cutaneous mechanoreceptor neurones in the central nervous system of the leech. J Physiol 1969, 203(3):591–609.PubMedCentralCrossRefPubMed
10.
Zurück zum Zitat Yuan S, Burrell BD: Endocannabinoid-dependent LTD in a nociceptive synapse requires activation of a presynaptic TRPV-like receptor. J Neurophysiol 2010, 104(5):2766–2777. 10.1152/jn.00491.2010PubMedCentralCrossRefPubMed Yuan S, Burrell BD: Endocannabinoid-dependent LTD in a nociceptive synapse requires activation of a presynaptic TRPV-like receptor. J Neurophysiol 2010, 104(5):2766–2777. 10.1152/jn.00491.2010PubMedCentralCrossRefPubMed
11.
Zurück zum Zitat Yuan S, Burrell BD: Long-term depression of nociceptive synapses by non-nociceptive afferent activity: role of endocannabinoids, Ca2+, and calcineurin. Brain Res 2012, 1460: 1–11.CrossRefPubMed Yuan S, Burrell BD: Long-term depression of nociceptive synapses by non-nociceptive afferent activity: role of endocannabinoids, Ca2+, and calcineurin. Brain Res 2012, 1460: 1–11.CrossRefPubMed
12.
Zurück zum Zitat Yuan S, Burrell BD: Endocannabinoid-dependent long-term depression in a nociceptive synapse requires coordinated presynaptic and postsynaptic transcription and translation. J Neurosci 2013, 33(10):4349–4358. 10.1523/JNEUROSCI.3922-12.2013PubMedCentralCrossRefPubMed Yuan S, Burrell BD: Endocannabinoid-dependent long-term depression in a nociceptive synapse requires coordinated presynaptic and postsynaptic transcription and translation. J Neurosci 2013, 33(10):4349–4358. 10.1523/JNEUROSCI.3922-12.2013PubMedCentralCrossRefPubMed
13.
Zurück zum Zitat Muller KJ, Nicholls JG, Stent GS: Neurobiology of the leech. Cold Spring Habbor: Cold Spring Harbor Laboratory Press; 1981. Muller KJ, Nicholls JG, Stent GS: Neurobiology of the leech. Cold Spring Habbor: Cold Spring Harbor Laboratory Press; 1981.
14.
Zurück zum Zitat Matias I, Bisogno T, Melck D, Vandenbulcke F, Verger-Bocquet M, De Petrocellis L, Sergheraert C, Breton C, Di MV, Salzet M: Evidence for an endocannabinoid system in the central nervous system of the leech Hirudo medicinalis. Brain Res MolBrain Res 2001, 87(2):145–159. 10.1016/S0169-328X(00)00290-4CrossRef Matias I, Bisogno T, Melck D, Vandenbulcke F, Verger-Bocquet M, De Petrocellis L, Sergheraert C, Breton C, Di MV, Salzet M: Evidence for an endocannabinoid system in the central nervous system of the leech Hirudo medicinalis. Brain Res MolBrain Res 2001, 87(2):145–159. 10.1016/S0169-328X(00)00290-4CrossRef
15.
Zurück zum Zitat Ross RA, Gibson TM, Brockie HC, Leslie M, Pashmi G, Craib SJ, Di Marzo V, Pertwee RG: Structure-activity relationship for the endogenous cannabinoid, anandamide, and certain of its analogues at vanilloid receptors in transfected cells and vas deferens. Br J Pharmacol 2001, 132(3):631–640. 10.1038/sj.bjp.0703850PubMedCentralCrossRefPubMed Ross RA, Gibson TM, Brockie HC, Leslie M, Pashmi G, Craib SJ, Di Marzo V, Pertwee RG: Structure-activity relationship for the endogenous cannabinoid, anandamide, and certain of its analogues at vanilloid receptors in transfected cells and vas deferens. Br J Pharmacol 2001, 132(3):631–640. 10.1038/sj.bjp.0703850PubMedCentralCrossRefPubMed
16.
Zurück zum Zitat Gibson HE, Edwards JG, Page RS, Van Hook MJ, Kauer JA: TRPV1 channels mediate long-term depression at synapses on hippocampal interneurons. Neuron 2008, 57(5):746–759. 10.1016/j.neuron.2007.12.027PubMedCentralCrossRefPubMed Gibson HE, Edwards JG, Page RS, Van Hook MJ, Kauer JA: TRPV1 channels mediate long-term depression at synapses on hippocampal interneurons. Neuron 2008, 57(5):746–759. 10.1016/j.neuron.2007.12.027PubMedCentralCrossRefPubMed
17.
Zurück zum Zitat Chavez AE, Chiu CQ, Castillo PE: TRPV1 activation by endogenous anandamide triggers postsynaptic long-term depression in dentate gyrus. Nat Neurosci 2010, 13(12):1511–1518. 10.1038/nn.2684PubMedCentralCrossRefPubMed Chavez AE, Chiu CQ, Castillo PE: TRPV1 activation by endogenous anandamide triggers postsynaptic long-term depression in dentate gyrus. Nat Neurosci 2010, 13(12):1511–1518. 10.1038/nn.2684PubMedCentralCrossRefPubMed
18.
Zurück zum Zitat Grueter BA, Brasnjo G, Malenka RC: Postsynaptic TRPV1 triggers cell type-specific long-term depression in the nucleus accumbens. Nat Neurosci 2010, 13(12):1519-U1107. 10.1038/nn.2685PubMedCentralCrossRefPubMed Grueter BA, Brasnjo G, Malenka RC: Postsynaptic TRPV1 triggers cell type-specific long-term depression in the nucleus accumbens. Nat Neurosci 2010, 13(12):1519-U1107. 10.1038/nn.2685PubMedCentralCrossRefPubMed
19.
Zurück zum Zitat Shaw BK, Kristan WB Jr: The whole-body shortening reflex of the medicinal leech: motor pattern, sensory basis, and interneuronal pathways. J Comp Physiol A 1995, 177(6):667–681.CrossRefPubMed Shaw BK, Kristan WB Jr: The whole-body shortening reflex of the medicinal leech: motor pattern, sensory basis, and interneuronal pathways. J Comp Physiol A 1995, 177(6):667–681.CrossRefPubMed
20.
Zurück zum Zitat Nicholls JG, Purves D: Monosynaptic chemical and electrical connexions between sensory and motor cells in the central nervous system of the leech. JPhysiol 1970, 209(3):647–667.CrossRef Nicholls JG, Purves D: Monosynaptic chemical and electrical connexions between sensory and motor cells in the central nervous system of the leech. JPhysiol 1970, 209(3):647–667.CrossRef
21.
Zurück zum Zitat Wessel R, Kristan WB Jr, Kleinfeld D: Supralinear summation of synaptic inputs by an invertebrate neuron: dendritic gain is mediated by an "inward rectifier" K(+) current. J Neurosci 1999, 19(14):5875–5888.PubMed Wessel R, Kristan WB Jr, Kleinfeld D: Supralinear summation of synaptic inputs by an invertebrate neuron: dendritic gain is mediated by an "inward rectifier" K(+) current. J Neurosci 1999, 19(14):5875–5888.PubMed
22.
Zurück zum Zitat Gaudry Q, Kristan WB Jr: Behavioral choice by presynaptic inhibition of tactile sensory terminals. Nat Neurosci 2009, 12(11):1450–1457. 10.1038/nn.2400CrossRefPubMed Gaudry Q, Kristan WB Jr: Behavioral choice by presynaptic inhibition of tactile sensory terminals. Nat Neurosci 2009, 12(11):1450–1457. 10.1038/nn.2400CrossRefPubMed
23.
Zurück zum Zitat Grey KB, Burrell BD: Co-induction of LTP and LTD and its regulation by protein kinases and phosphatases. J Neurophysiol 2010, 103(5):2737–2746. 10.1152/jn.01112.2009PubMedCentralCrossRefPubMed Grey KB, Burrell BD: Co-induction of LTP and LTD and its regulation by protein kinases and phosphatases. J Neurophysiol 2010, 103(5):2737–2746. 10.1152/jn.01112.2009PubMedCentralCrossRefPubMed
24.
Zurück zum Zitat Baccus SA: Synaptic facilitation by reflected action potentials: enhancement of transmission when nerve impulses reverse direction at axon branch points. Proc NatlAcadSciUSA 1998, 95(14):8345–8350. 10.1073/pnas.95.14.8345CrossRef Baccus SA: Synaptic facilitation by reflected action potentials: enhancement of transmission when nerve impulses reverse direction at axon branch points. Proc NatlAcadSciUSA 1998, 95(14):8345–8350. 10.1073/pnas.95.14.8345CrossRef
25.
Zurück zum Zitat Todd KL, Kristan WB Jr, French KA: Gap junction expression is required for normal chemical synapse formation. J Neurosci 2010, 30(45):15277–15285. 10.1523/JNEUROSCI.2331-10.2010PubMedCentralCrossRefPubMed Todd KL, Kristan WB Jr, French KA: Gap junction expression is required for normal chemical synapse formation. J Neurosci 2010, 30(45):15277–15285. 10.1523/JNEUROSCI.2331-10.2010PubMedCentralCrossRefPubMed
26.
Zurück zum Zitat Cline HT: Evidence for GABA as a neurotransmitter in the leech. J Neurosci 1986, 6(10):2848–2856.PubMed Cline HT: Evidence for GABA as a neurotransmitter in the leech. J Neurosci 1986, 6(10):2848–2856.PubMed
27.
Zurück zum Zitat Baca SM, Marin-Burgin A, Wagenaar DA, Kristan WB Jr: Widespread inhibition proportional to excitation controls the gain of a leech behavioral circuit. Neuron 2008, 57(2):276–289. 10.1016/j.neuron.2007.11.028PubMedCentralCrossRefPubMed Baca SM, Marin-Burgin A, Wagenaar DA, Kristan WB Jr: Widespread inhibition proportional to excitation controls the gain of a leech behavioral circuit. Neuron 2008, 57(2):276–289. 10.1016/j.neuron.2007.11.028PubMedCentralCrossRefPubMed
28.
Zurück zum Zitat Cline HT: 3H-GABA uptake selectively labels identifiable neurons in the leech central nervous system. J Comp Neurol 1983, 215(3):351–358. 10.1002/cne.902150309CrossRefPubMed Cline HT: 3H-GABA uptake selectively labels identifiable neurons in the leech central nervous system. J Comp Neurol 1983, 215(3):351–358. 10.1002/cne.902150309CrossRefPubMed
29.
Zurück zum Zitat Sargent PB, Yau KW, Nicholls JG: Extrasynaptic receptors on cell bodies of neurons in central nervous system of the leech. J Neurophysiol 1977, 40(2):446–452.PubMed Sargent PB, Yau KW, Nicholls JG: Extrasynaptic receptors on cell bodies of neurons in central nervous system of the leech. J Neurophysiol 1977, 40(2):446–452.PubMed
30.
Zurück zum Zitat Di Marzo V: The endocannabinoid system: its general strategy of action, tools for its pharmacological manipulation and potential therapeutic exploitation. Pharmacol Res: the official journal of the Italian Pharmacological Society 2009, 60(2):77–84. 10.1016/j.phrs.2009.02.010CrossRef Di Marzo V: The endocannabinoid system: its general strategy of action, tools for its pharmacological manipulation and potential therapeutic exploitation. Pharmacol Res: the official journal of the Italian Pharmacological Society 2009, 60(2):77–84. 10.1016/j.phrs.2009.02.010CrossRef
31.
Zurück zum Zitat Willis WD Jr: Dorsal root potentials and dorsal root reflexes: a double-edged sword. Exp Brain Res 1999, 124(4):395–421. 10.1007/s002210050637CrossRefPubMed Willis WD Jr: Dorsal root potentials and dorsal root reflexes: a double-edged sword. Exp Brain Res 1999, 124(4):395–421. 10.1007/s002210050637CrossRefPubMed
32.
Zurück zum Zitat Chavas J, Marty A: Coexistence of excitatory and inhibitory GABA synapses in the cerebellar interneuron network. J Neurosci 2003, 23(6):2019–2031.PubMed Chavas J, Marty A: Coexistence of excitatory and inhibitory GABA synapses in the cerebellar interneuron network. J Neurosci 2003, 23(6):2019–2031.PubMed
33.
Zurück zum Zitat Isomura Y, Sugimoto M, Fujiwara-Tsukamoto Y, Yamamoto-Muraki S, Yamada J, Fukuda A: Synaptically activated Cl- accumulation responsible for depolarizing GABAergic responses in mature hippocampal neurons. J Neurophysiol 2003, 90(4):2752–2756. 10.1152/jn.00142.2003CrossRefPubMed Isomura Y, Sugimoto M, Fujiwara-Tsukamoto Y, Yamamoto-Muraki S, Yamada J, Fukuda A: Synaptically activated Cl- accumulation responsible for depolarizing GABAergic responses in mature hippocampal neurons. J Neurophysiol 2003, 90(4):2752–2756. 10.1152/jn.00142.2003CrossRefPubMed
34.
Zurück zum Zitat Ruiz A, Fabian-Fine R, Scott R, Walker MC, Rusakov DA, Kullmann DM: GABAA receptors at hippocampal mossy fibers. Neuron 2003, 39(6):961–973. 10.1016/S0896-6273(03)00559-2PubMedCentralCrossRefPubMed Ruiz A, Fabian-Fine R, Scott R, Walker MC, Rusakov DA, Kullmann DM: GABAA receptors at hippocampal mossy fibers. Neuron 2003, 39(6):961–973. 10.1016/S0896-6273(03)00559-2PubMedCentralCrossRefPubMed
35.
Zurück zum Zitat Knabl J, Witschi R, Hosl K, Reinold H, Zeilhofer UB, Ahmadi S, Brockhaus J, Sergejeva M, Hess A, Brune K: Reversal of pathological pain through specific spinal GABAA receptor subtypes. Nature 2008, 451(7176):330–334. 10.1038/nature06493CrossRefPubMed Knabl J, Witschi R, Hosl K, Reinold H, Zeilhofer UB, Ahmadi S, Brockhaus J, Sergejeva M, Hess A, Brune K: Reversal of pathological pain through specific spinal GABAA receptor subtypes. Nature 2008, 451(7176):330–334. 10.1038/nature06493CrossRefPubMed
36.
Zurück zum Zitat Granzow B, Kristan WB Jr: Inhibitory connections between motor neurons modify a centrally generated motor pattern in the leech nervous system. Brain Res 1986, 369(1–2):321–325.CrossRefPubMed Granzow B, Kristan WB Jr: Inhibitory connections between motor neurons modify a centrally generated motor pattern in the leech nervous system. Brain Res 1986, 369(1–2):321–325.CrossRefPubMed
37.
Zurück zum Zitat Peterfi Z, Urban GM, Papp OI, Nemeth B, Monyer H, Szabo G, Erdelyi F, Mackie K, Freund TF, Hajos N: Endocannabinoid-mediated long-term depression of afferent excitatory synapses in hippocampal pyramidal cells and GABAergic interneurons. J Neurosci 2012, 32(41):14448–14463. 10.1523/JNEUROSCI.1676-12.2012PubMedCentralCrossRefPubMed Peterfi Z, Urban GM, Papp OI, Nemeth B, Monyer H, Szabo G, Erdelyi F, Mackie K, Freund TF, Hajos N: Endocannabinoid-mediated long-term depression of afferent excitatory synapses in hippocampal pyramidal cells and GABAergic interneurons. J Neurosci 2012, 32(41):14448–14463. 10.1523/JNEUROSCI.1676-12.2012PubMedCentralCrossRefPubMed
38.
Zurück zum Zitat Adermark L, Talani G, Lovinger DM: Endocannabinoid-dependent plasticity at GABAergic and glutamatergic synapses in the striatum is regulated by synaptic activity. Eur J Neurosci 2009, 29(1):32–41. 10.1111/j.1460-9568.2008.06551.xPubMedCentralCrossRefPubMed Adermark L, Talani G, Lovinger DM: Endocannabinoid-dependent plasticity at GABAergic and glutamatergic synapses in the striatum is regulated by synaptic activity. Eur J Neurosci 2009, 29(1):32–41. 10.1111/j.1460-9568.2008.06551.xPubMedCentralCrossRefPubMed
39.
Zurück zum Zitat Gerdeman GL, Ronesi J, Lovinger DM: Postsynaptic endocannabinoid release is critical to long-term depression in the striatum. Nat Neurosci 2002, 5(5):446–451.PubMed Gerdeman GL, Ronesi J, Lovinger DM: Postsynaptic endocannabinoid release is critical to long-term depression in the striatum. Nat Neurosci 2002, 5(5):446–451.PubMed
40.
Zurück zum Zitat Chevaleyre V, Castillo PE: Heterosynaptic LTD of hippocampal GABAergic synapses: a novel role of endocannabinoids in regulating excitability. Neuron 2003, 38(3):461–472. 10.1016/S0896-6273(03)00235-6CrossRefPubMed Chevaleyre V, Castillo PE: Heterosynaptic LTD of hippocampal GABAergic synapses: a novel role of endocannabinoids in regulating excitability. Neuron 2003, 38(3):461–472. 10.1016/S0896-6273(03)00235-6CrossRefPubMed
41.
Zurück zum Zitat Heifets BD, Chevaleyre V, Castillo PE: Interneuron activity controls endocannabinoid-mediated presynaptic plasticity through calcineurin. ProcNatlAcadSciUSA 2008, 105(29):10250–10255. 10.1073/pnas.0711880105CrossRef Heifets BD, Chevaleyre V, Castillo PE: Interneuron activity controls endocannabinoid-mediated presynaptic plasticity through calcineurin. ProcNatlAcadSciUSA 2008, 105(29):10250–10255. 10.1073/pnas.0711880105CrossRef
42.
Zurück zum Zitat Huang Y, Yasuda H, Sarihi A, Tsumoto T: Roles of endocannabinoids in heterosynaptic long-term depression of excitatory synaptic transmission in visual cortex of young mice. J Neurosci 2008, 28(28):7074–7083. 10.1523/JNEUROSCI.0899-08.2008CrossRefPubMed Huang Y, Yasuda H, Sarihi A, Tsumoto T: Roles of endocannabinoids in heterosynaptic long-term depression of excitatory synaptic transmission in visual cortex of young mice. J Neurosci 2008, 28(28):7074–7083. 10.1523/JNEUROSCI.0899-08.2008CrossRefPubMed
43.
Zurück zum Zitat Carlson G, Wang Y, Alger BE: Endocannabinoids facilitate the induction of LTP in the hippocampus. Nat Neurosci 2002, 5(8):723–724.PubMed Carlson G, Wang Y, Alger BE: Endocannabinoids facilitate the induction of LTP in the hippocampus. Nat Neurosci 2002, 5(8):723–724.PubMed
44.
Zurück zum Zitat Chevaleyre V, Castillo PE: Endocannabinoid-mediated metaplasticity in the hippocampus. Neuron 2004, 43(6):871–881. 10.1016/j.neuron.2004.08.036CrossRefPubMed Chevaleyre V, Castillo PE: Endocannabinoid-mediated metaplasticity in the hippocampus. Neuron 2004, 43(6):871–881. 10.1016/j.neuron.2004.08.036CrossRefPubMed
45.
Zurück zum Zitat Zhu PJ, Lovinger DM: Persistent synaptic activity produces long-lasting enhancement of endocannabinoid modulation and alters long-term synaptic plasticity. J Neurophysiol 2007, 97(6):4386–4389. 10.1152/jn.01228.2006CrossRefPubMed Zhu PJ, Lovinger DM: Persistent synaptic activity produces long-lasting enhancement of endocannabinoid modulation and alters long-term synaptic plasticity. J Neurophysiol 2007, 97(6):4386–4389. 10.1152/jn.01228.2006CrossRefPubMed
46.
Zurück zum Zitat Christie MJ, Mallet C: Endocannabinoids can open the pain gate. Sci Signal 2009, 2(88):pe57. 10.1126/scisignal.288pe57PubMed Christie MJ, Mallet C: Endocannabinoids can open the pain gate. Sci Signal 2009, 2(88):pe57. 10.1126/scisignal.288pe57PubMed
47.
Zurück zum Zitat Mbvundula EC, Rainsford KD, Bunning RA: Cannabinoids in pain and inflammation. Inflammopharmacology 2004, 12(2):99–114. 10.1163/1568560041352275CrossRefPubMed Mbvundula EC, Rainsford KD, Bunning RA: Cannabinoids in pain and inflammation. Inflammopharmacology 2004, 12(2):99–114. 10.1163/1568560041352275CrossRefPubMed
48.
Zurück zum Zitat Rahn EJ, Hohmann AG: Cannabinoids as pharmacotherapies for neuropathic pain: from the bench to the bedside. Neurotherapeutics 2009, 6(4):713–737. 10.1016/j.nurt.2009.08.002PubMedCentralCrossRefPubMed Rahn EJ, Hohmann AG: Cannabinoids as pharmacotherapies for neuropathic pain: from the bench to the bedside. Neurotherapeutics 2009, 6(4):713–737. 10.1016/j.nurt.2009.08.002PubMedCentralCrossRefPubMed
49.
Zurück zum Zitat Lynch ME, Campbell F: Cannabinoids for treatment of chronic non-cancer pain; a systematic review of randomized trials. Br J Clin Pharmacol 2011, 72(5):735–744. 10.1111/j.1365-2125.2011.03970.xPubMedCentralCrossRefPubMed Lynch ME, Campbell F: Cannabinoids for treatment of chronic non-cancer pain; a systematic review of randomized trials. Br J Clin Pharmacol 2011, 72(5):735–744. 10.1111/j.1365-2125.2011.03970.xPubMedCentralCrossRefPubMed
50.
Zurück zum Zitat Kato A, Punnakkal P, Pernia-Andrade AJ, von Schoultz C, Sharopov S, Nyilas R, Katona I, Zeilhofer HU: Endocannabinoid-dependent plasticity at spinal nociceptor synapses. J Physiol 2012, 590(Pt 19):4717–4733.PubMedCentralCrossRefPubMed Kato A, Punnakkal P, Pernia-Andrade AJ, von Schoultz C, Sharopov S, Nyilas R, Katona I, Zeilhofer HU: Endocannabinoid-dependent plasticity at spinal nociceptor synapses. J Physiol 2012, 590(Pt 19):4717–4733.PubMedCentralCrossRefPubMed
51.
Zurück zum Zitat Naef M, Curatolo M, Petersen-Felix S, Arendt-Nielsen L, Zbinden A, Brenneisen R: The analgesic effect of oral delta-9-tetrahydrocannabinol (THC), morphine, and a THC-morphine combination in healthy subjects under experimental pain conditions. Pain 2003, 105(1–2):79–88.CrossRefPubMed Naef M, Curatolo M, Petersen-Felix S, Arendt-Nielsen L, Zbinden A, Brenneisen R: The analgesic effect of oral delta-9-tetrahydrocannabinol (THC), morphine, and a THC-morphine combination in healthy subjects under experimental pain conditions. Pain 2003, 105(1–2):79–88.CrossRefPubMed
52.
Zurück zum Zitat Coull JA, Boudreau D, Bachand K, Prescott SA, Nault F, Sik A, De Koninck P, De Koninck Y: Trans-synaptic shift in anion gradient in spinal lamina I neurons as a mechanism of neuropathic pain. Nature 2003, 424(6951):938–942. 10.1038/nature01868CrossRefPubMed Coull JA, Boudreau D, Bachand K, Prescott SA, Nault F, Sik A, De Koninck P, De Koninck Y: Trans-synaptic shift in anion gradient in spinal lamina I neurons as a mechanism of neuropathic pain. Nature 2003, 424(6951):938–942. 10.1038/nature01868CrossRefPubMed
53.
Zurück zum Zitat Sandkuhler J: Models and mechanisms of hyperalgesia and allodynia. Physiol Rev 2009, 89(2):707–758. 10.1152/physrev.00025.2008CrossRefPubMed Sandkuhler J: Models and mechanisms of hyperalgesia and allodynia. Physiol Rev 2009, 89(2):707–758. 10.1152/physrev.00025.2008CrossRefPubMed
54.
Zurück zum Zitat Smith ES, Lewin GR: Nociceptors: a phylogenetic view. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 2009, 195(12):1089–1106. 10.1007/s00359-009-0482-zPubMedCentralCrossRefPubMed Smith ES, Lewin GR: Nociceptors: a phylogenetic view. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 2009, 195(12):1089–1106. 10.1007/s00359-009-0482-zPubMedCentralCrossRefPubMed
55.
Zurück zum Zitat Walters ET, Moroz LL: Molluscan memory of injury: evolutionary insights into chronic pain and neurological disorders. Brain Behav Evol 2009, 74(3):206–218. 10.1159/000258667PubMedCentralCrossRefPubMed Walters ET, Moroz LL: Molluscan memory of injury: evolutionary insights into chronic pain and neurological disorders. Brain Behav Evol 2009, 74(3):206–218. 10.1159/000258667PubMedCentralCrossRefPubMed
56.
Zurück zum Zitat Kim YH, Back SK, Davies AJ, Jeong H, Jo HJ, Chung G, Na HS, Bae YC, Kim SJ, Kim JS: TRPV1 in GABAergic interneurons mediates neuropathic mechanical allodynia and disinhibition of the nociceptive circuitry in the spinal cord. Neuron 2012, 74(4):640–647. 10.1016/j.neuron.2012.02.039CrossRefPubMed Kim YH, Back SK, Davies AJ, Jeong H, Jo HJ, Chung G, Na HS, Bae YC, Kim SJ, Kim JS: TRPV1 in GABAergic interneurons mediates neuropathic mechanical allodynia and disinhibition of the nociceptive circuitry in the spinal cord. Neuron 2012, 74(4):640–647. 10.1016/j.neuron.2012.02.039CrossRefPubMed
57.
Zurück zum Zitat Yang K, Kumamoto E, Furue H, Li YQ, Yoshimura M: Action of capsaicin on dorsal root-evoked synaptic transmission to substantia gelatinosa neurons in adult rat spinal cord slices. Brain Res 1999, 830(2):268–273. 10.1016/S0006-8993(99)01408-0CrossRefPubMed Yang K, Kumamoto E, Furue H, Li YQ, Yoshimura M: Action of capsaicin on dorsal root-evoked synaptic transmission to substantia gelatinosa neurons in adult rat spinal cord slices. Brain Res 1999, 830(2):268–273. 10.1016/S0006-8993(99)01408-0CrossRefPubMed
58.
Zurück zum Zitat Baccei ML, Bardoni R, Fitzgerald M: Development of nociceptive synaptic inputs to the neonatal rat dorsal horn: glutamate release by capsaicin and menthol. J Physiol 2003, 549(Pt 1):231–242.PubMedCentralCrossRefPubMed Baccei ML, Bardoni R, Fitzgerald M: Development of nociceptive synaptic inputs to the neonatal rat dorsal horn: glutamate release by capsaicin and menthol. J Physiol 2003, 549(Pt 1):231–242.PubMedCentralCrossRefPubMed
59.
Zurück zum Zitat Kusudo K, Ikeda H, Murase K: Depression of presynaptic excitation by the activation of vanilloid receptor 1 in the rat spinal dorsal horn revealed by optical imaging. Mol Pain 2006, 2: 8. 10.1186/1744-8069-2-8PubMedCentralCrossRefPubMed Kusudo K, Ikeda H, Murase K: Depression of presynaptic excitation by the activation of vanilloid receptor 1 in the rat spinal dorsal horn revealed by optical imaging. Mol Pain 2006, 2: 8. 10.1186/1744-8069-2-8PubMedCentralCrossRefPubMed
60.
Zurück zum Zitat Iskedjian M, Bereza B, Gordon A, Piwko C, Einarson TR: Meta-analysis of cannabis based treatments for neuropathic and multiple sclerosis-related pain. Curr Med Res Opin 2007, 23(1):17–24. 10.1185/030079906X158066CrossRefPubMed Iskedjian M, Bereza B, Gordon A, Piwko C, Einarson TR: Meta-analysis of cannabis based treatments for neuropathic and multiple sclerosis-related pain. Curr Med Res Opin 2007, 23(1):17–24. 10.1185/030079906X158066CrossRefPubMed
61.
Zurück zum Zitat Stuart AE: Physiological and morphological properties of motoneurones in the central nervous system of the leech. J Physiol 1970, 209(3):627–646.PubMedCentralCrossRefPubMed Stuart AE: Physiological and morphological properties of motoneurones in the central nervous system of the leech. J Physiol 1970, 209(3):627–646.PubMedCentralCrossRefPubMed
Metadaten
Titel
Differential modulation of nociceptive versus non-nociceptive synapses by endocannabinoids
verfasst von
Alexandra Higgins
Sharleen Yuan
Yanqing Wang
Brian D Burrell
Publikationsdatum
01.12.2013
Verlag
BioMed Central
Erschienen in
Molecular Pain / Ausgabe 1/2013
Elektronische ISSN: 1744-8069
DOI
https://doi.org/10.1186/1744-8069-9-26

Weitere Artikel der Ausgabe 1/2013

Molecular Pain 1/2013 Zur Ausgabe

Delir bei kritisch Kranken – Antipsychotika versus Placebo

16.05.2024 Delir Nachrichten

Um die Langzeitfolgen eines Delirs bei kritisch Kranken zu mildern, wird vielerorts auf eine Akuttherapie mit Antipsychotika gesetzt. Eine US-amerikanische Forschungsgruppe äußert jetzt erhebliche Vorbehalte gegen dieses Vorgehen. Denn es gibt neue Daten zum Langzeiteffekt von Haloperidol bzw. Ziprasidon versus Placebo.

Eingreifen von Umstehenden rettet vor Erstickungstod

15.05.2024 Fremdkörperaspiration Nachrichten

Wer sich an einem Essensrest verschluckt und um Luft ringt, benötigt vor allem rasche Hilfe. Dass Umstehende nur in jedem zweiten Erstickungsnotfall bereit waren, diese zu leisten, ist das ernüchternde Ergebnis einer Beobachtungsstudie aus Japan. Doch es gibt auch eine gute Nachricht.

Darf man die Behandlung eines Neonazis ablehnen?

08.05.2024 Gesellschaft Nachrichten

In einer Leseranfrage in der Zeitschrift Journal of the American Academy of Dermatology möchte ein anonymer Dermatologe bzw. eine anonyme Dermatologin wissen, ob er oder sie einen Patienten behandeln muss, der eine rassistische Tätowierung trägt.

Ein Drittel der jungen Ärztinnen und Ärzte erwägt abzuwandern

07.05.2024 Klinik aktuell Nachrichten

Extreme Arbeitsverdichtung und kaum Supervision: Dr. Andrea Martini, Sprecherin des Bündnisses Junge Ärztinnen und Ärzte (BJÄ) über den Frust des ärztlichen Nachwuchses und die Vorteile des Rucksack-Modells.

Update AINS

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.