Next Article in Journal
Validation of a Novel Cuproptosis–Related Prognostic Gene Marker and Differential Expression Associated with Lung Adenocarcinoma
Previous Article in Journal
The Role of p16/Ki67 Dual Staining in Cervical Cancer Screening
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Presynaptic Purinergic Modulation of the Rat Neuro-Muscular Transmission

by
Adel E. Khairullin
1,2,
Sergey N. Grishin
3 and
Ayrat U. Ziganshin
4,*
1
Department of Biochemistry, Kazan State Medical University, 420012 Kazan, Russia
2
Research Laboratory of Mechanobiology, Kazan Federal University, 420008 Kazan, Russia
3
Department of Medicinal Physics, Kazan State Medical University, 420012 Kazan, Russia
4
Department of Pharmacology, Kazan State Medical University, 420012 Kazan, Russia
*
Author to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2023, 45(10), 8492-8501; https://doi.org/10.3390/cimb45100535
Submission received: 29 August 2023 / Revised: 8 October 2023 / Accepted: 18 October 2023 / Published: 19 October 2023
(This article belongs to the Special Issue Molecular Insights into Skeletal Muscle Homeostasis and Metabolism)

Abstract

:
ATP, being a well-known universal high-energy compound, plays an important role as a signaling molecule and together with its metabolite adenosine they both attenuate the release of acetylcholine in the neuro-muscular synapse acting through membrane P2 and P1 receptors, respectively. In this work, using a mechanomyographic method, we analyzed the presynaptic mechanisms by which ATP and adenosine can modulate the transduction in the rat m. soleus and m. extensor digitorum longus. N-ethylmaleimide, a G-protein antagonist, prevents the modulating effects of both ATP and adenosine. The action of ATP is abolished by chelerythrin, a specific phospholipase C inhibitor, while the inhibitory effect of adenosine is slightly increased by Rp-cAMPS, an inhibitor of protein kinase A, and by nitrendipine, a blocker of L-type Ca2+ channels. The addition of DPCPX, an A1 receptor antagonist, fully prevents the inhibitory action of adenosine in both muscles. Our data indicate that the inhibitory action of ATP involves metabotropic P2Y receptors and is mediated by phospholipase C dependent processes in rat motor neuron terminals. We suggest that the presynaptic effect of adenosine consists of negative and positive actions. The negative action occurs by stimulation of adenosine A1 receptors while the positive action is associated with the stimulation of adenosine A2A receptors, activation of protein kinase A and opening of L-type calcium channels. The combined mechanism of the modulating action of ATP and adenosine provides fine tuning of the synapse to fast changing conditions in the skeletal muscles.

1. Introduction

The synaptic modulatory function of adenosine-5′-triphosphate (ATP) was discovered at the end of the last century [1,2], and initially it was assumed to be only a precursor of synaptically active adenosine [3]. Subsequently, it was shown that ATP in the synaptic cleft acts independently according to the principle of negative feedback through the ending of motor neurons in the synapses of amphibian phasic muscles [2,3,4]. Released jointly from synaptic vesicles together with the main mediator acetylcholine (ACh) from motor nerve endings, ATP inhibits ACh secretion and muscle contractions [1,2,3,4].
ATP inhibits quantum [1,2,3,4] and non-quantum [5] secretion of ACh from presynaptic endings of motor neurons, acting on P2 receptors, while the end product of its decay, adenosine, implements its action through adenosine receptors [6,7]. Identification of specific subtypes involved adenosine and P2 receptors, and their pharmacological characteristics are the initial stages of studying the effector mechanisms of synaptic regulation by purines of the function of various types of skeletal muscles.
The presynaptic inhibitory effect of adenosine on the amplitude of the evoked phasic endplate currents was established in lake frog muscles [8]. Under similar conditions, the inhibitory presynaptic effect of ATP was demonstrated [9]. At the same time, both pre- and postsynaptic potentiating effects of ATP were established on 2-week-old Xenopus tadpoles and in cell cultures of myocytes [10]. Much less is known about action of purines on the muscles of warm-blooded animals. Experiments on synapses in the rat diaphragm showed a presynaptic potentiating effect of activation of P2 receptors leading to the facilitation of the release of acetylcholine [11], and later the inhibitory effect of adenosine and ATP on end plate potentials was demonstrated at the same object [12]. It is clear that further evaluation of the effects of these purines on other skeletal muscles of warm-blooded animals is needed.
In view of all this, our goal was to investigate the presynaptic transduction mechanisms of ATP and adenosine in the neuromuscular synapse of rat fast and slow skeletal muscles.

2. Materials and Methods

The animal study protocol was approved by the Ethics Committee of Kazan State Medical University (protocol code 106/4, from 14 December 2021).
The studies were carried out on the neuromuscular preparations of 38 white Wistar male laboratory rats weighing 140–180 g (4–5 months old), which were kept in groups of 3–5 individuals with unlimited access to water and food. The animals were divided into 8 groups: “Suramin”—6 rats, “8-SPT”—5 rats, “DPCPX”—5 rats, “N-ethylmaleimide”—5 rats, “Rp-cAMP”—5 rats, “Nitrendipine”—5 rats, “Chelerythrin”—4 rats and “d-tubocurarine”—3 rats. Animals were anesthetized by intraperitoneal administration of sodium ethaminal at a dose of 40 mg/kg, bled and m. soleus and m. EDL were isolated from both hind limbs.
The experimental procedures were made as described earlier in [13]. Briefly, the isolated muscles were fixed vertically tying one end motionless and attaching the other to a mechanical activity sensor and immersed in 10 mL baths filled with Krebs solution of the following composition (in mM): NaCl 118.0, KCl 4.75, CaCl2 2.5, NaHCO3 24.8, KH2PO4 1.18, MgSO4 7H2O 1.18, glucose 11, pH = 7.4, t = 37 ± 0.5 °C. The thermostat maintains the set temperature. Muscles were stretched with an initial load of 1 g, then left alone for 30 min to adapt to environmental conditions.
Electrical field stimulation (EFS) was performed by stimulating the nerve stump, which was placed in an original design suction electrode. For stimulation, a Digitimer MultiStim D330 stimulator (Digitimer Ltd, Hertfordshire, UK) was used. Muscle contractions were induced by stimulation with rectangular pulses with a frequency of 0.1 Hz, a duration of 0.5 ms, and an amplitude of 10 V for 2 min. Contraction parameters were recorded using a Linton FSG-01 isometric mechanical activity sensor (Linton, UK), and the analog signal was digitized and processed using a Biopack MP100WSW data acquisition system (BIOPAC Systems, Inc., Goleta, CA, USA). The average value of the parameters of all contractions received within 2 min (12 responses) was treated as one result.
After a 30 min initial control, stimulation of the muscle was performed twice with an interval of 5 min, and after ensuring the stability of contractile reactions, experimental procedures were started. All further contractile reactions were calculated as a percentage relative to these initial results.
ATP or adenosine (100 µM) was added to the bath with Krebs solution and the contractile responses of the muscles were evaluated after 7 min. The tissue was then incubated with 8-(p-sulfophenyl)-theophylline (8-SPT), 1,3-dipropyl-8-cyclopentylxanthine (DPCPX), suramin, N-ethylmaleimide, 6-(6-aminopurin-9-yl)-2-hydroxy-2-sulfanylidene-4a,6,7,7a-tetrahydro-4H-furo[3,2-d][1,3,2]dioxaphosphinin-7-ol;N,N-diethylethanamine (Rp-cAMPS), chelerythrin, nitrendipine, or d-tubocurarine for 15 min, followed by the addition of ATP or adenosine, and contractile responses were recorded again (Figure 1).
In a separate series of experiments on 3 male rats, the mechanical activity of the muscle was assessed in the presence of d-tubocurarine (a nicotinic cholinoceptor blocker), which was added to baths at a concentration of 10 µM 20 min before stimulation. In another control experiment, muscle preparations were stimulated using the same parameters (rectangular monophasic electrical impulses with a frequency of 0.1 Hz, a duration of 0.5 ms) but with a voltage of 100 V.
Mechanomiographic experiments on rats m. soleus and m. EDL were analyzed using the SPSS Statistic program. The compliance of the obtained data with the normal distribution was checked using the Kolmogorov’s test. The arithmetic mean of the analyzed parameters and the standard error were calculated. The statistical significance of the observed changes was assessed using Student’s t-test for independent and pairwise conjugate samples. Differences were considered significant at p < 0.05.

3. Results

3.1. The Effect of Purines on the Contraction Parameters of Rat m. soleus and m. EDL

ATP at a concentration of 100 μM inhibited contractions of the soleus muscle induced by EFS to 70.6 ± 5.2% of the initial values (n = 18, p < 0.05). Preincubation with the P2 receptor antagonist suramin (100 μM) abolished the effect of ATP on the contraction amplitude (101.9 ± 7.0%, n = 12, Figure 2 and Figure 3).
Adenosine (100 μM) reduced the amplitude of contractions of the soleus muscle induced by electric field stimulation to 70.3 ± 6.4% (n = 16, p < 0.05). Additionally, 8-SPT (100 μM), a non-selective adenosine receptor antagonist, almost completely inhibited the effect of adenosine on soleus muscle contractions (94.5 ± 6.0%, n = 9, Figure 3).
In the presence of ATP (100 μM), the contraction force of m. EDL was 86.2 ± 3.9% (n = 18), and these data are significantly different from both the initial contractions and the ATP effect in the soleus muscle (Figure 4 and Figure 5).
The inhibitory action of ATP was abolished by suramin (100 μM); in its presence, ATP acted only up to 97.5 ± 8.1% (n = 11, p > 0.05) of the initial control contractions.
Adenosine (100 μM) inhibited the contraction force of m. EDL to 71.3 ± 6.2% (n = 16, p < 0.05). Additionally, 8-SPT (100 μM) almost completely reversed the effect of adenosine on soleus muscle contractions (95.8 ± 7.9%; n = 9, Figure 5).
DPCPX (0.1 μM), a selective A1 receptor antagonist, abolished the effects of adenosine on both muscles; the contraction force of the soleus muscle became 106.2 ± 5.1% (n = 8, p > 0.05) and m. EDL—109.3 ± 6.4% (n = 8, p > 0.05) of the initial control (Figure 3 and Figure 5).
To evaluate whether the EFS parameters used in our experiments induced contractions through nerve stimulation rather than direct muscle stimulation, we tested effects of d-tubocurarine (10 μM), a nicotinic cholinoceptor blocker. We found that d-tubocurarine eliminated all contractile reactions of rat m. soleus and m. EDL induced by EFS with standard parameters (see Methods). An increase in voltage from 10 V to 100 V led to contractile responses of both m. soleus and m. EDL, which were not inhibited by d-tubocurarine; the presence of ATP, adenosine, suramin or 8-SPT did not affect these contractions.

3.2. Identification of Intracellular Mechanisms Involved in the Effects of Adenosine and ATP

It is known that P2Y receptors and all adenosine receptors belong to the big family of G-protein coupled receptors and some types of these receptors are sensitive to a G-protein blocker N-ethylmaleimide [14].
N-ethylmaleimide at a concentration of 10 μM did not significantly change the force of contraction of both muscles tested. Thus, the force of contraction was 105.2 ± 4.3% (n = 8; p > 0.05) and 103.7 ± 5.3% (n = 8; p > 0.05) of the initial control values, respectively, in m. EDL and m. soleus (Figure 6 and Figure 7).
In the presence of N-ethylmaleimide (10 μM), the inhibitory actions of ATP and adenosine were completely abolished in both muscles. Thus, the strengths of contraction in the presence of ATP (100 μM) were 102.7 ± 8.0% and 99.2 ± 7.1% (n = 8, p < 0.05), respectively, in m. EDL and m. soleus. Similarly, in the presence of adenosine, the contractile responses were 103.4 ± 7.7% in m. EDL and 104.6 ± 6.2% (n = 8; p > 0.05) in m. soleus (Figure 6 and Figure 7).
These data suggest that the effects of both purines are mediated by a G protein sensitive to N-ethylmaleimide. As such, the involvement of P2Y receptors in the inhibitory action of ATP is confirmed, since only this family of ATP receptors is metabotropic, G protein-mediated.
Possible ways of transmitting the activating signal from the GTP protein are enzyme proteins, the activation of which leads to the formation of various second messengers.
Incubation of the muscle preparations with adenosine-3′,5′-cyclomonophosphoro-thioate (50 µM, Rp-cAMP), a protein kinase A blocker, slightly increased the inhibitory action of adenosine (100 µM), which was 64.7 ± 5.8% (n = 8; p < 0.05) on soleus muscle and 67.2 ± 5.4% (n = 8; p < 0.05) on m. EDL relative to initial control values (Figure 6 and Figure 7).
In the presence of both DPCPX (0.1 μM) and Rp-cAMP (50 µM), the effect of adenosine on the contraction force was no longer observed, being 100.3 ± 3.4% (n = 8; p < 0.05) in m. soleus and 98.7 ± 2.9% (n = 8; p < 0.05) in m. EDL (Figure 6 and Figure 7).
With incubation of muscle preparations with nitrendipine (5 μM), an L-type calcium channel blocker, the inhibitory effect of adenosine on both muscles was not significantly changed. Thus, contractions of m. soleus became 65.2 ± 6.1% (n = 8, p < 0.05), while m. EDL—66.8 ± 3.9% (n = 8, p < 0.05).
Next, we found out whether the modulating effect of ATP is realized by the intracellular adenylate cyclase metabolic pathway.
In the presence of Rp-cAMP (50 µM), there was no significant change in the inhibitory effects of ATP (100 µM) on both muscles. This suggests that the modulating effects of ATP are mediated independently of protein kinase A.
Chelerythrin, a protein kinase C inhibitor, at a concentration of 5 μM did not significantly change the amplitude of contraction of the studied muscles. Thus, the contraction force of m. soleus was 106.8 ± 5.9% (n = 7; p > 0.05), m. EDL—104.7 ± 6.2% (n = 7; p > 0.05) of the control (Figure 6 and Figure 7).
In the presence of chelerythrin (5 μM), ATP (100 μM) had no effect on the contraction force m. EDL (97.2 ± 5.9%, n = 8; p > 0.05) as well as m. soleus (109.7 ± 7.6%, n = 8; p > 0.05).

4. Discussion

As is known, neuromodulation by adenine nucleotides occurs either directly, through the activation of ionotropic P2X/metabotropic P2Y receptors [6,7], or indirectly, after their extracellular breakdown to adenosine and subsequent activation of P1 receptors, which include four different subtypes (A1, A2A, A2B, and A3) [15]. Nerve endings, skeletal muscle fibers, and Schwann cells express different subtypes of these purine receptors [16,17,18]. However, the physiological significance of ATP and its metabolites in modulating neuromuscular transmission is far from universally recognized, mainly because studies have been conducted under different experimental conditions in different preparations from a variety of animals. It happened that no attempts were even made to separate the P2 receptor- from adenosine receptor-mediated effects [19].
In this paper, we tried to reveal the mechanisms of the modulation effects of both ATP and adenosine in the synapses of the fast and slow muscles of the rat calf. Initially, we carried out experiments with the application of endogenous ATP. In our experiments, we found that ATP inhibits the contractile activity of the slow soleus muscle of the rat and, to a lesser extent, the fast type of muscle—the extensor digitorum longus. These two types of muscles have certain differences. For example, it was found that they have different isoforms of the sarcoplasmic reticulum calcium pump [20], and that the amount of Ca2+ in the sarcoplasmic reticulum of slow muscles is greater than that of fast muscles [21]. The mechanism of the inhibitory effect of ATP may be associated with a decrease in the release of Ca2+ from the sarcoplasmic reticulum, as was previously shown for ADP in the slow fibers of the rat soleus muscle [22]. Suramin, a non-selective P2 receptor blocker, abolished this action of ATP. Thus, it can be assumed that ATP exerts its inhibitory effect on rat phasic muscle contractions by influencing P2Y receptors. The participation of presynaptic P2Y receptors in the inhibitory effect of ATP on neuromuscular transmission in the frog sartorius was previously shown [10]. Another confirmation of this assumption is that α,β-methylene-ATP, which predominantly acts on P2X receptors, had no effect on contractions of either fast or slow phase rat skeletal muscles [23].
We also showed that the inhibitory effect of ATP on the quantum yield in amphibian synapses is mediated by P2Y12 receptor-associated Gi/o proteins sensitive to pertussis toxin [23]. In the experimental data presented in this article, N-ethylmaleimide, a G-protein antagonist, prevents the modulating effect of the purines used, once again confirming the involvement of metabotropic receptors in the inhibitory effect of ATP.
Activation of metabotropic P2Y receptors can trigger an extensive mechanism, which, using the neuromuscular transition in amphibians as an example, includes the production of reactive oxygen species (H2O2) and prostaglandin E2, which reduces the supply of Ca2+ to the frog motor nerve endings [23]. But initially, the activating impulse from the P2Y receptor must be transmitted through the G protein to the primary link of intracellular second messengers, phospholipase or cyclase. In the next series of experiments, we tested the effect of ATP against the background of chelerythrin, a protein kinase C blocker, specific for the intracellular metabolic branch of phospholipase C. It turned out that chelerythrin eliminates the inhibitory effect of ATP on the preparation of both fast and slow muscles. In warm-blooded animals, the involvement of the intracellular phospholipase C metabolic pathway has previously been shown only in diaphragm preparations [23].
The final metabolite of ATP, adenosine, acts through A1, A2A, A2B, or A3 receptors. In rodents, A1 and A2A receptors show high affinity for adenosine, while A2B and A3 are low affinity receptors [15]. The A1 and A3 receptors are associated with the Gi/o protein, while the A2A and A2B are preferably associated with the Gs protein. Thus, activation of A2A and A2B typically results in stimulation of adenylate cyclase, resulting in an increase in intracellular 3′-5′-cyclic adenosine monophosphate (cAMP) and subsequent activation of protein kinase A (PKA). Conversely, the activation of A1 and A3 leads to suppression of adenylate cyclase activity, thus reducing cAMP levels and PKA activation [24]. Remarkably, these receptors can also bind to non-canonical pathways, thus regulating intracellular Ca2+ and/or K+ concentrations by activating Gq protein-coupled mechanisms; others involve direct interference with ionic currents across the plasma membrane through the βγ subunits of G proteins [25,26].
In our experiments, we used N-ethylmaleimide, a Gi/o inhibitor, which completely eliminated the inhibitory effect of adenosine. This may mean that the inhibitory effect is mediated by activation of the A1 or A3 receptor. Continuing the analysis further, we found that in the presence of the A1 receptor antagonist DPCPX, the inhibitory effect of adenosine was eliminated.
Previously, it was demonstrated by electrophysiological and neurochemical methods that A1-inhibitory and A2-excitatory receptors coexist at the rat phrenic nerve terminals to fine-tune the regulation of ACh exocytosis [27]. The experimental data of our group showed that adenosine activates both A1 and A2A in non-respiratory skeletal muscle, since the potentiating effect of activation of A2A receptors appeared against the background of DPCPX. This specificity, which is usually observed in the diaphragmatic and central synapses [28,29,30,31,32], may be associated with the proximity of ecto-5′-nucleotidase/CD73 and A2A, which contributes to the activation of this subtype of receptors after the formation of adenosine during the breakdown of AMP [28,30].
The A2A-mediated facilitation of ACh release from motor nerve endings is thought to be mainly caused by the normally “calm” high-efficiency L-type calcium channels (Cav1) that provide Ca2+ influx. When the operation of P/Q-type calcium channels (Cav2.1) prevails, L-type channels are desensitized, for example, during the high-frequency excitation of neurons [32,33,34]. In our experiments, in the presence of L-type calcium channel blockers, the facilitating effect of adenosine disappeared.
The mechanism of subsequent activation of the A2A motor terminal depends on the formation of cAMP by adenylate cyclase [35,36], which leads to stimulation of PKA activity and subsequent Ca2+ influx through L-type calcium channels [37,38]. Characteristically, in the results of the experiments presented in this article, the facilitating effect of adenosine continued to be observed against the background of the PKA inhibitor.
Interestingly, in the frog myoneural junction, A1 activation inhibits ACh release through a Ca2+-independent mechanism, which suggests direct interference with proteins involved in vesicular exocytosis [23]. A decrease in A2A receptor function seems to be a hallmark of aging and some neuromuscular disorders [15].
Summing up, in this paper we demonstrated that the two purinergic pathways for modulating the induced ACh exocytosis skillfully complement each other (see Figure 8). It is known that the inhibitory effect of ATP in the synaptic cleft is more often observed at low frequencies of neuronal stimulation (≤1 Hz) [15]. This suggests that it is adenosine that, by activating both presynaptic inhibitory A1 and facilitating A2A receptors, depending on the paradigm of the situation, leads the adjusting effects of purines [39,40].

5. Conclusions

Thus, we have studied the effects of both ATP and adenosine in the rat fast and slow skeletal muscles. Our data indicate that, in the motor neuron terminals of both studied muscles, the transmission of an activating impulse was mediated by G-protein-coupled ATP receptors (P2Y family) to PLC. We suggest that the presynaptic effect of adenosine consists of two mechanisms: negative and positive, which involve both A1 and A2A adenosine receptors. The combined mechanism of the modulating action of ATP and adenosine provides fine tuning of the synapse to fast-changing conditions in the skeletal muscles. The findings reveal potential targets for pharmacological intervention. However, the situation is not yet completely clear, and the purine signaling involved in the above-mentioned effects requires further comprehensive studies.

Author Contributions

Conceptualization, A.E.K. and S.N.G.; methodology, A.E.K.; software, A.E.K.; validation, A.E.K.; formal analysis, A.U.Z.; investigation, A.E.K.; resources, A.U.Z.; data curation, S.N.G.; writing—original draft preparation, A.E.K. and S.N.G.; writing—review and editing, A.U.Z.; visualization, A.E.K.; supervision, A.U.Z.; project administration, A.U.Z.; funding acquisition, A.U.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Kazan Federal University Strategic Academic Leadership Program (PRIORITY-2030).

Institutional Review Board Statement

The animal study protocol was approved by the Ethics Committee of Kazan State Medical University (protocol code 106/4, from 14 December 2021).

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

8-SPT8-(p-sulfophenyl)-theophylline
ACadenylate cyclase.
ACHacetylcholine
ATPadenosine-5′-triphosphate
cAMP3′-5′-cyclic adenosine monophosphate,
DAGdiacylglycerol,
DPCPX1,3-dipropyl-8-cyclopentylxanthine
EFSelectrical field stimulation
m. EDLextensor digitorum longus;
PKAprotein kinase A,
PKCprotein kinase C,
PLCphospholipase C,
Rp-cAMPS6-(6-aminopurin-9-yl)-2-hydroxy-2-sulfanylidene-4a,6,7,7a-tetrahydro-4H-furo[3,2-d][1,3,2]dioxaphosphinin-7-ol;N,N-diethylethanamine

References

  1. Silinsky, E.M.; Redman, R.S. Synchronous release of ATP and neurotransmitter within milliseconds of a motor nerve impulse in the frog. J. Physiol. 1996, 492, 815–822. [Google Scholar] [CrossRef] [PubMed]
  2. Giniatullin, R.A.; Sokolova, E.M. Modulating role of ATP in the neuromuscular junction. Russ. J. Physiol. 1998, 84, 1132–1138. [Google Scholar]
  3. Ziganshin, A.U.; Kamaliev, R.R.; Grishin, S.N.; Ziganshina, L.E.; Zefirov, A.L.; Burnstock, G. The influence of hypothermia on P2 receptor-mediated responses of frog skeletal muscle. Eur. J. Pharmacol. 2005, 509, 187–193. [Google Scholar] [CrossRef] [PubMed]
  4. Giniatullin, R.A.; Sokolova, E.M. ATP and adenosine inhibit transmitter release at the frog neuromuscular junction through distinct presynaptic receptors. Br. J. Pharmacol. 1998, 124, 839–844. [Google Scholar] [CrossRef]
  5. Galkin, A.V.; Giniatullin, R.A.; Mukhtarov, M.R.; Svandova, I.; Grishin, S.N.; Vyskocil, F. ATP but not adenosine inhibits nonquantal acetylcholine release at the mouse neuromuscular junction. Eur. J. Neurosci. 2001, 13, 2047–2053. [Google Scholar] [CrossRef]
  6. Alexander, S.P.H.; Christopoulos, A.; Davenport, A.P.; Kelly, E.; Mathie, A.; Peters, J.A.; Veale, E.L.; Armstrong, J.F.; Faccenda, E.; Harding, S.D.; et al. The concise guide to pharmacology 2019/20: G protein-coupled receptors. Br. J. Pharmacol. 2019, 176, 21–141. [Google Scholar] [CrossRef]
  7. Alexander, S.P.H.; Mathie, A.; Peters, J.A.; Veale, E.L.; Striessnig, J.; Kelly, E.; Armstrong, J.F.; Faccenda, E.; Harding, S.D.; Pawson, A.J.; et al. The concise guide to pharmacology 2019/20: Ion Channels. Br. J. Pharmacol. 2019, 176, 142–228. [Google Scholar] [CrossRef]
  8. Ribeiro, J.A.; Sebastião, A.M. On the role, inactivation and origin of endogenous adenosine at the frog neuromuscular junction. J. Physiol. 1987, 384, 571–585. [Google Scholar] [CrossRef]
  9. Sokolova, E.; Grishin, S.; Shakirzyanova, A.; Talantova, M.; Giniatullin, R. Distinct receptors and different transduction mechanisms for ATP and adenosine at the frog motor nerve endings. Eur. J. Neurosci. 2003, 18, 1254–1264. [Google Scholar] [CrossRef]
  10. Fu, W.M. Regulatory role of ATP at developing neuromuscular junctions. Prog. Neurobiol. 1995, 47, 31–44. [Google Scholar] [CrossRef]
  11. Salgado, A.I.; Cunha, R.A.; Ribeiro, J.A. Facilitation by P(2) receptor activation of acetylcholine release from rat motor nerve terminals: Interaction with presynaptic nicotinic receptors. Brain Res. 2000, 877, 245–250. [Google Scholar] [CrossRef]
  12. González Sanabria, J.; Hurtado Paso, M.; Frontera, T.; Losavio, A. Effect of endogenous purines on electrically evoked ACh release at the mouse neuromuscular junction. J. Neurosci. Res. 2022, 100, 1933–1950. [Google Scholar] [CrossRef] [PubMed]
  13. Ziganshin, A.U.; Khairullin, A.E.; Teplov, A.Y.; Gabdrakhmanov, A.I.; Ziganshina, L.E.; Hoyle, C.H.V.; Ziganshin, B.A.; Grishin, S.N. The effects of ATP on the contractions of rat and mouse fast skeletal muscle. Muscle Nerve. 2019, 59, 509–516. [Google Scholar] [CrossRef] [PubMed]
  14. Shapiro, M.S.; Wollmuth, L.P.; Hille, B. Modulation of Ca2+ channels by PTX-sensitive G-proteins is blocked by N-ethylmaleimide in rat sympathetic neurons. J. Neurosci. 1994, 14, 7109–7116. [Google Scholar] [CrossRef]
  15. Sousa-Soares, C.; Noronha-Matos, J.B.; Correia-de-Sá, P. Purinergic Tuning of the Tripartite Neuromuscular Synapse. Mol. Neurobiol. 2023, 60, 4084–4104. [Google Scholar] [CrossRef] [PubMed]
  16. Jimsheleishvili, S.; Marwaha, K.; Sherman, A.I. Physiology, Neuromuscular Transmission. StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2019; Available online: https://pubmed.ncbi.nlm.nih.gov/31082177/ (accessed on 28 August 2023).
  17. Mukund, K.; Subramaniam, S. Skeletal muscle: A review of molecular structure and function, in health and disease. Wiley Interdiscip. Rev. Syst. Biol. Med. 2020, 12, 1462. [Google Scholar] [CrossRef] [PubMed]
  18. Pasnoor, M.; Dimachkie, M.M. Approach to Muscle and Neuromuscular Junction Disorders. Continuum 2019, 25, 1536–1563. [Google Scholar] [CrossRef]
  19. Ribeiro, J.A.; Cunha, R.A.; Correia-de-Sá, P.; Sebastião, A.M. Purinergic regulation of acetylcholine release. Prog. Brain. Res. 1996, 109, 231–241. [Google Scholar] [CrossRef]
  20. Vizi, E.S.; Nitahara, K.; Sato, K.; Sperlágh, B. Stimulation-dependent release, breakdown, and action of endogenous ATP in mouse hemidiaphragm preparation: The possible role of ATP in neuromuscular transmission. J. Auton. Nerv. Syst. 2000, 81, 278–284. [Google Scholar] [CrossRef]
  21. Fryer, M.W.; Stephenson, D.G. Total and sarcoplasmic reticulum calcium contents of skinned fibres from rat skeletal muscle. J. Physiol. 1996, 493, 357–370. [Google Scholar] [CrossRef]
  22. Macdonald, W.A.; Stephenson, D.G. Effect of ADP on slow-twitch muscle fibres of the rat: Implications for muscle fatigue. J. Physiol. 2006, 573, 187–198. [Google Scholar] [CrossRef] [PubMed]
  23. Ziganshin, A.U.; Khairullin, A.E.; Hoyle, C.H.V.; Grishin, S.N. Modulatory Roles of ATP and Adenosine in Cholinergic Neuromuscular Transmission. Int. J. Mol. Sci. 2020, 21, 6423. [Google Scholar] [CrossRef] [PubMed]
  24. Fredholm, B.B.; Ijzerman, A.P.; Jacobson, K.A.; Linden, J.; Müller, C.E. International Union of Basic and Clinical Pharmacology. LXXXI. Nomenclature and classifcation of adenosine receptors—An update. Pharmacol. Rev. 2011, 63, 1–34. [Google Scholar] [CrossRef]
  25. Sebastião, A.M.; Cristóvão-Ferreira, S.; Ribeiro, J.A. Downstream Pathways of Adenosine. In Adenosine; Masino, S., Boison, D., Eds.; Springer: New York, NY, USA, 2013. [Google Scholar] [CrossRef]
  26. Jacobson, K.A.; Gao, Z.G. Adenosine receptors as therapeutic targets. Nat. Rev. Drug. Discov. 2006, 5, 247–264. [Google Scholar] [CrossRef] [PubMed]
  27. Correia-de-Sá, P.; Sebastião, A.M.; Ribeiro, J.A. Inhibitory and excitatory effects of adenosine receptor agonists on evoked transmitter release from phrenic nerve ending of the rat. Br. J. Pharmacol. 1991, 103, 1614–1620. [Google Scholar] [CrossRef] [PubMed]
  28. Barros-Barbosa, A.R.; Ferreirinha, F.; Oliveira, Â.; Mendes, M.; Lobo, M.G.; Santos, A.; Rangel, R.; Pelletier, J.; Sévigny, J.; Cordeiro, J.M.; et al. Adenosine A2A receptor and ecto-5’-nucleotidase/CD73 are upregulated in hippocampal astrocytes of human patients with mesial temporal lobe epilepsy (MTLE). Purinergic Signal. 2016, 12, 719–734. [Google Scholar] [CrossRef]
  29. Vieira, C.; Magalhães-Cardoso, M.T.; Ferreirinha, F.; Silva, I.; Dias, A.S.; Pelletier, J.; Sévigny, J.; Correia-de-Sá, P. Feed-forward inhibition of CD73 and upregulation of adenosine deaminase contribute to the loss of adenosine neuromodulation in postinfammatory ileitis. Mediat. Infamm. 2014, 2014, 254640. [Google Scholar] [CrossRef]
  30. Augusto, E.; Matos, M.; Sévigny, J.; El-Tayeb, A.; Bynoe, M.S.; Müller, C.E.; Cunha, R.A.; Chen, J.F. Ecto-5’-nucleotidase (CD73)- mediated formation of adenosine is critical for the striatal adenosine A2A receptor functions. J. Neurosci. 2013, 33, 11390–11399. [Google Scholar] [CrossRef]
  31. Duarte-Araújo, M.; Nascimento, C.; Timóteo, M.A.; Magalhães-Cardoso, M.T.; Correia-de-Sá, P. Dual efects of adenosine on acetylcholine release from myenteric motoneurons are mediated by junctional facilitatory A2A and extrajunctional inhibitory A1 receptors. Br. J. Pharmacol. 2004, 141, 925–934. [Google Scholar] [CrossRef]
  32. Correia-de-Sá, P.; Timóteo, M.A.; Ribeiro, J.A. A(2A) adenosine receptor facilitation of neuromuscular transmission: Infuence of stimulus paradigm on calcium mobilization. J. Neurochem. 2000, 74, 2462–2469. [Google Scholar] [CrossRef]
  33. Noronha-Matos, J.B.; Morais, T.; Trigo, D.; Timóteo, M.A.; Magalhães Cardoso, M.T.; Oliveira, L.; Correia-de-Sá, P. Tetanic failure due to decreased endogenous adenosine A2A tonus operating neuronal Cav1 (L-type) infux in myasthenia gravis. J. Neurochem. 2011, 117, 797–811. [Google Scholar] [CrossRef]
  34. Oliveira, L.; Timóteo, M.A.; Correia-de-Sá, P. Tetanic depression is overcome by tonic adenosine A2A receptor facilitation of L-type Ca2+ infux into rat motor nerve terminals. J. Physiol. 2004, 560, 157–168. [Google Scholar] [CrossRef] [PubMed]
  35. Correia-de-Sá, P.; Ribeiro, J.A. Evidence that the presynaptic A2A-adenosine receptor of the rat motor nerve endings is positively coupled to adenylate cyclase. Naunyn-Schmiedeberg’s Arch. Pharmacol. 1994, 350, 514–522. [Google Scholar] [CrossRef]
  36. Correia-de-Sá, P.; Ribeiro, J.A. Facilitation of [3H]-ACh release by forskolin depends on A2-adenosine receptor activation. Neurosci. Lett. 1993, 151, 21–24. [Google Scholar] [CrossRef] [PubMed]
  37. Oliveira, L.; Correia-de-Sá, P. Protein kinase A and Cav1 (L-type) channels are common targets to facilitatory adenosine A2A and muscarinic M1 receptors on rat motoneurons. Neurosignals 2005, 14, 262–272. [Google Scholar] [CrossRef]
  38. Tarasova, E.O.; Miteva, A.S.; Gaidukov, A.E.; Balezina, O.P. The role of adenosine receptors and L-type calcium channels in the regulation of the mediator secretion in mouse motor synapses. Biochem. Suppl. Ser. A Membr. Cell. Biol. 2015, 9, 318–328. [Google Scholar] [CrossRef]
  39. Oliveira, L.; Timóteo, M.A.; Correia-de-Sá, P. Modulation by adenosine of both muscarinic M1-facilitation and M2-inhibition of [3H]-acetylcholine release from the rat motor nerve terminals. Eur. J. Neurosci. 2002, 15, 1728–1736. [Google Scholar] [CrossRef] [PubMed]
  40. Correia-de-Sá, P.; Timóteo, M.A.; Ribeiro, J.A. Presynaptic A1 inhibitory/A2A facilitatory adenosine receptor activation balance depends on motor nerve stimulation paradigm at the rat hemidiaphragm. J. Neurophysiol. 1996, 76, 3910–3919. [Google Scholar] [CrossRef]
Figure 1. Time line of the experiments.
Figure 1. Time line of the experiments.
Cimb 45 00535 g001
Figure 2. Original traces of contractile responses of the isolated rat m. soleus evoked by electrical field stimulation (0.1 Hz, 0.5 ms, 10 V) in control (a) and in the presence of ATP (100 μM, (b)), suramin (100 μM, (c)), a combination of ATP and suramin (both at 100 μM, (d)).
Figure 2. Original traces of contractile responses of the isolated rat m. soleus evoked by electrical field stimulation (0.1 Hz, 0.5 ms, 10 V) in control (a) and in the presence of ATP (100 μM, (b)), suramin (100 μM, (c)), a combination of ATP and suramin (both at 100 μM, (d)).
Cimb 45 00535 g002
Figure 3. Contractions of rat m. soleus evoked by electrical field stimulation in the presence of ATP (100 μM), suramin (100 μM), adenosine (Ado, 100 μM), 8-sulphofenyltheophyllin (8-SPT, 100 μM), DPCPX (0.1 μM). Data shown as mean and standard error, n = 8–16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Figure 3. Contractions of rat m. soleus evoked by electrical field stimulation in the presence of ATP (100 μM), suramin (100 μM), adenosine (Ado, 100 μM), 8-sulphofenyltheophyllin (8-SPT, 100 μM), DPCPX (0.1 μM). Data shown as mean and standard error, n = 8–16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Cimb 45 00535 g003
Figure 4. Original traces of contractile responses of the isolated rat m. soleus evoked by electrical field stimulation (0.1 Hz, 0.5 ms, 10 V) in the control (a) and in the presence of adenosine (100 μM, (b)), 8-sulfophenyltheophylline (100 μM, (c)), a combination of adenosine and 8-sulfophenyltheophylline (both at 100 μM, (d)).
Figure 4. Original traces of contractile responses of the isolated rat m. soleus evoked by electrical field stimulation (0.1 Hz, 0.5 ms, 10 V) in the control (a) and in the presence of adenosine (100 μM, (b)), 8-sulfophenyltheophylline (100 μM, (c)), a combination of adenosine and 8-sulfophenyltheophylline (both at 100 μM, (d)).
Cimb 45 00535 g004
Figure 5. Contractions of rat m. EDL evoked by electrical field stimulation in the presence of ATP (100 μM), suramin (100 μM), adenosine (Ado, 100 μM), 8-sulphofenyltheophyllin (8-SPT, 100 μM), DPCPX (0.1 μM). Data shown as mean and standard error, n = 8—16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Figure 5. Contractions of rat m. EDL evoked by electrical field stimulation in the presence of ATP (100 μM), suramin (100 μM), adenosine (Ado, 100 μM), 8-sulphofenyltheophyllin (8-SPT, 100 μM), DPCPX (0.1 μM). Data shown as mean and standard error, n = 8—16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Cimb 45 00535 g005
Figure 6. Contractions of rat m. soleus evoked by electrical field stimulation in the presence of ATP (100 μM), adenosine (Ado, 100 μM), N-ethylmaleimide (10 μM), Rp-cAMP (50 µM), DPCPX (0.1 μM), nitrendipine (5 μM), chelerythrin (5 μM) or their combination. Data shown as mean and standard error, n = 8–16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Figure 6. Contractions of rat m. soleus evoked by electrical field stimulation in the presence of ATP (100 μM), adenosine (Ado, 100 μM), N-ethylmaleimide (10 μM), Rp-cAMP (50 µM), DPCPX (0.1 μM), nitrendipine (5 μM), chelerythrin (5 μM) or their combination. Data shown as mean and standard error, n = 8–16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Cimb 45 00535 g006
Figure 7. Contractions of rat m. EDL evoked by electrical field stimulation in the presence of ATP (100 μM), adenosine (Ado, 100 μM), N-ethylmaleimide (10 μM), Rp-cAMP (50 µM), DPCPX (0.1 μM), nitrendipine (5 μM), chelerythrin (5 μM) or their combination. Data shown as mean and standard error, n = 8–16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Figure 7. Contractions of rat m. EDL evoked by electrical field stimulation in the presence of ATP (100 μM), adenosine (Ado, 100 μM), N-ethylmaleimide (10 μM), Rp-cAMP (50 µM), DPCPX (0.1 μM), nitrendipine (5 μM), chelerythrin (5 μM) or their combination. Data shown as mean and standard error, n = 8–16. *—p < 0.05 vs. controls, #—p < 0.05 vs. ATP or Ado alone.
Cimb 45 00535 g007
Figure 8. Proposed scheme of purinergic regulation of myoneural transmission of skeletal muscles in warm-blooded animals and the direction of action of the pharmacological agents used in the study. The end of the motor neuron is shown. A synaptic vesicle is presented—a vesicle filled, in addition to the mediator ACh (acetylcholine), ATP. Showing exocytosis of the contents of the vesicles, hydrolysis of ATP to adenosine in the synaptic cleft; activation by adenosine of A1—and A2A—receptors, and ATP—P2Y—receptors, and their regulation of exocytosis according to the feedback principle through presynaptic intracellular mechanisms. Designations: PKC—protein kinase C, DAG—diacylglycerol, PLC—phospholipase C, PKA—protein kinase A, cAMP—3′-5′-cyclic adenosine monophosphate, AC—adenylate cyclase.
Figure 8. Proposed scheme of purinergic regulation of myoneural transmission of skeletal muscles in warm-blooded animals and the direction of action of the pharmacological agents used in the study. The end of the motor neuron is shown. A synaptic vesicle is presented—a vesicle filled, in addition to the mediator ACh (acetylcholine), ATP. Showing exocytosis of the contents of the vesicles, hydrolysis of ATP to adenosine in the synaptic cleft; activation by adenosine of A1—and A2A—receptors, and ATP—P2Y—receptors, and their regulation of exocytosis according to the feedback principle through presynaptic intracellular mechanisms. Designations: PKC—protein kinase C, DAG—diacylglycerol, PLC—phospholipase C, PKA—protein kinase A, cAMP—3′-5′-cyclic adenosine monophosphate, AC—adenylate cyclase.
Cimb 45 00535 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khairullin, A.E.; Grishin, S.N.; Ziganshin, A.U. Presynaptic Purinergic Modulation of the Rat Neuro-Muscular Transmission. Curr. Issues Mol. Biol. 2023, 45, 8492-8501. https://doi.org/10.3390/cimb45100535

AMA Style

Khairullin AE, Grishin SN, Ziganshin AU. Presynaptic Purinergic Modulation of the Rat Neuro-Muscular Transmission. Current Issues in Molecular Biology. 2023; 45(10):8492-8501. https://doi.org/10.3390/cimb45100535

Chicago/Turabian Style

Khairullin, Adel E., Sergey N. Grishin, and Ayrat U. Ziganshin. 2023. "Presynaptic Purinergic Modulation of the Rat Neuro-Muscular Transmission" Current Issues in Molecular Biology 45, no. 10: 8492-8501. https://doi.org/10.3390/cimb45100535

Article Metrics

Back to TopTop