Skip to main content

Cholecystokinin type B receptor-mediated inhibition of A-type K+ channels enhances sensory neuronal excitability through the phosphatidylinositol 3-kinase and c-Src-dependent JNK pathway

Abstract

Background

Cholecystokinin (CCK) is implicated in the regulation of nociceptive sensitivity of primary afferent neurons. Nevertheless, the underlying cellular and molecular mechanisms remain unknown.

Methods

Using patch clamp recording, western blot analysis, immunofluorescent labelling, enzyme-linked immunosorbent assays, adenovirus-mediated shRNA knockdown and animal behaviour tests, we studied the effects of CCK-8 on the sensory neuronal excitability and peripheral pain sensitivity mediated by A-type K+ channels.

Results

CCK-8 reversibly and concentration-dependently decreased A-type K+ channel (IA) in small-sized dorsal root ganglion (DRG) neurons through the activation of CCK type B receptor (CCK-BR), while the sustained delayed rectifier K+ current was unaffected. The intracellular subunit of CCK-BR coimmunoprecipitated with Gαo. Blocking G-protein signaling with pertussis toxin or by the intracellular application of anti-Gβ antibody reversed the inhibitory effects of CCK-8. Antagonism of phosphatidylinositol 3-kinase (PI3K) but not of its common downstream target Akts abolished the CCK-BR-mediated IA response. CCK-8 application significantly activated JNK mitogen-activated protein kinase. Antagonism of either JNK or c-Src prevented the CCK-BR-mediated IA decrease, whereas c-Src inhibition attenuated the CCK-8-induced p-JNK activation. Application of CCK-8 enhanced the action potential firing rate of DRG neurons and elicited mechanical and thermal pain hypersensitivity in mice. These effects were mediated by CCK-BR and were occluded by IA blockade.

Conclusion

Our findings indicate that CCK-8 attenuated IA through CCK-BR that is coupled to the Gβγ-dependent PI3K and c-Src-mediated JNK pathways, thereby enhancing the sensory neuronal excitability in DRG neurons and peripheral pain sensitivity in mice.

Background

Cholecystokinin (CCK), a gastrointestinal polypeptide hormone existing in a variety of amino acid chain lengths, has been isolated from the central nervous system and peripheral tissues [1]. Two types of functional membrane receptors, cholecystokinin A receptor (CCK-AR), located mainly on pancreatic acinar cells, and CCK-BR, mostly in the stomach and nervous system tissues, have been identified as the endogenous receptors of CCK [2]. CCK, acting through its receptors, is found to be involved in the regulation of a variety of physiological functions in the nervous system, including metabolic, neurotrophic and modulatory actions [3]. Additionally, in vitro experiments have suggested that CCK might regulate the sensitivity of nociceptive sensory neurons [4], where CCK-BRs were abundantly expressed [5, 6]. It has been established that the elevated level of CCK mRNA in the dorsal root ganglia (DRG) induced by peripheral nerve injury sensitizes and excites primary afferent sensory neurons, leading to pain hypersensitivity [7], with the application of CCK inducing pronociceptive effects [8]. Evidence also suggests that CCK is a potential trigger for increased visceral sensitivity in healthy subjects [9] as well as in irritable bowel syndrome patients [10, 11]. Moreover, antagonism of CCK receptors can effectively reverse burn-induced mechanical allodynia [5], and the deletion of the CCK-BR gene attenuates the symptoms of mechanical allodynia in neuropathic pain [12]. So far, however, the mechanisms underlying the CCK-mediated hyperalgesia still remain unclear.

Changes in neuronal excitability of peripheral sensory neurons might directly regulate symptoms of pain, such as allodynia, hyperalgesia and spontaneous pain [13]. Voltage-gated K+ channels (Kv) are one of the major classes of ion channels responsible for driving neuronal excitability in both the central and peripheral nervous system [14] and in whole-cell patch clamp recordings are separated into two major categories: a large transient component characteristic for fast-inactivating A-type channels and a sustained delayed-rectifying channels, which respectively mediate IA and IDR currents [15, 16]. A-type channels are sensitive to millimolar concentrations of 4-aminopyridine (4-AP) [17], and they play pivotal roles in the control of the electrical properties and excitability of nociceptive neurons [14, 15]. Recent evidence has suggested that A-type channels have been implicated in pain plasticity and neuropathic conditions [17], which begin with the aberrant firing of action potential bursts in damaged neuronal tissue. For example, peripheral nerve injury results in the reduction of channel expression, thereby decreasing IA, and enhancing the sensory neuronal excitability and pain sensitivity [18, 19]. Manipulation of IA is, therefore, predicted to affect neuronal excitability and useful for pain treatment.

In the current study, we examined the regulation of CCK-8 on IA in small-sized (< 30 μm in soma diameter) DRG neurons to determine whether A-type K+ channels mediate the nociceptive actions of CCK-BR.

Materials and methods

Animals

Adult ICR mice (male, 6–8 weeks of age) were purchased from the Experimental Animal Center of Soochow University. Mice were maintained in specific-pathogen-free facilities on a 12-h light-dark cycle at a room temperature of 22 ± 1 °C, and housed in cages with access to food and water ad libitum. All animal studies were conducted in accordance with the National Institutes of Health’s Guidelines for Animal Care and Use and approved by the Animal Care and Use Committee of Soochow University.

Isolation of DRG neurons

DRG neurons were obtained from lumber L4–6 segments in adult ICR mice (male, 6–8 weeks of age) using enzymatic dissociation procedure as described previously [20, 21]. Dissociated DRG neurons were plated onto Matrigel-coated glass coverslips and maintained in incubators until recording. We sorted DRG neurons into groups based on the soma diameter distribution of small (< 30 μm in soma diameter), medium (30–40 μm in soma diameter), and large-sized (> 40 μm in soma diameter) DRG neurons [22], and in patch-clamp experiments we only recorded the small-sized neurons. Cells were subjected to whole-cell recordings within 24–48 h of plating.

Electrophysiological recordings

Whole-cell patch-clamp recordings were performed at room temperature (23 ± 1 °C). Recording electrodes (World Precision Instruments, USA) pulled from borosilicate glass microcapillary tubes (Sutter Instruments) had resistances from 3 to 5 MΩ when filled with internal solution. Recordings were made using a MultiClamp 700B patch-clamp amplifier (Molecular Devices, USA). Whole-cell currents were low-pass filtered at 2–5 kHz. The values of cell capacitance and series resistance were taken directly from readings of the amplifier after electronic subtraction of the capacitive transients. In voltage-clamp mode, 80% of the series resistance was electronically compensated. A P/6 protocol was used for on-line leak subtractions. The external solution used for Kv current recordings contained (mM): KCl (5), choline-Cl (150), MgCl2 (1), CaCl2 (0.03), glucose (10) and HEPES (10) adjusted to pH 7.4 with KOH, 310 mOsm. The internal solution used for Kv current recordings contained (mM): KCl (140), Na-GTP (0.3), Mg-ATP (3), 0.5 CaCl2, MgCl2 (1), EGTA (5) and HEPES (10) adjusted to pH 7.4 with KOH, 295 mOsm. The external solution used for Cav current recordings contained the following (mM): BaCl2 (5), tetraethylammonium chloride (TEA-Cl) (140), CsCl (5), HEPES (10), MgCl2 (0.5) and glucose (5.5) adjusted to pH 7.35 with TEA-OH, 310 mOsm. The internal solution used for Cav current recordings contained the following (mM): CsCl (110), Mg-ATP (4), EGTA (10), Na-GTP (0.3) and HEPES (25) adjusted to pH 7.4 with CsOH, 295 mOsm. To separate T-type channel currents, we applied the L-type channel blocker nifedipine (5 μM) and the N- and P/Q-type channel blocker ω-conotoxin MVIIC (0.2 μM) in the external solution. The external solution used for both Nav current and current-clamp recordings contained the following (mM): KCl (2), MgCl2 (2), CaCl2 (2), NaCl (128), HEPES (25) and glucose (30) adjusted to pH 7.4 with NaOH, 305 mOsm. The internal solution contained the following (in mM): NaCl (10), KCl (110), Na-GTP (0.3), Mg-ATP (4), EGTA (2) and HEPES (25) adjusted to pH 7.3 with KOH, 295 mOsm. Path pipettes had resistance from 2 to 3 MΩ for internal dialysis with compounds. Monoclonal antibodies raised against Gαo (anti-Go-Ab, Santa Cruz Biotechnology) or Gαi (anti-Gi-Ab, Santa Cruz Biotechnology) were diluted to 1: 100 times from a stock solution (200 μg/ml). This antibody used in the present study has been previously shown to specifically recognize the G-protein Gαo but not Gαi subunit [20, 23]. For the whole cell experiments, the antibody was diluted into the intracellular solution and loaded into the pipette. In patch-clamp recording in which cells were dialyzed with inhibitors or activators, protocols were initiated at least 5 min after breaking the membrane patch.

Immunoblotting

Immunoblot analysis was conducted as described in our previous studies [23]. Briefly, equal amounts of proteins (25 μg) were separated by 10% SDS–PAGE and electroblotted onto PVDF membranes (Merck Millipore, Germany). Blotted proteins were probed with the following primary antibodies: goat anti-CCK-AR (1: 500, Abcam), goat anti-CCK-BR (1: 500, Abcam), rabbit anti-phospho-Akt (1:2000, Cell Signaling Technology), rabbit anti-Akt (1:1000, Cell Signaling Technology), rabbit anti-phospho-p38 MAPK (1:1000, Cell Signaling Technology), rabbit anti-p38 MAPK (1:1000, Cell Signaling Technology), rabbit anti-phospho-ERK1/2 (1:2000, Cell Signaling Technology), rabbit anti-ERK1/2 (1:1000, Cell Signaling Technology), rabbit anti-phospho-SAPK/JNK (1:1000, Cell Signaling Technology), rabbit anti-SAPK/JNK (1:1000, Cell Signaling Technology), rabbit anti-phospho-Src (pTyr418, 1:2000, Abcam), rabbit anti-Src (1: 1000, Abcam) and rabbit anti-GAPDH (1:5000, Abcam). After extensive washing in TBST, membranes were incubated with horseradish peroxidase-conjugated anti-rabbit or anti-goat IgG (1: 8000, Multi Sciences). Chemiluminescent signals were generated using a SuperSignal West Pico trial kit (Pierce) and detected using ChemiDoc XRS System (Bio-Rad Laboratories). The software Quantity One (Bio-Rad Laboratories) was used for background subtraction and for quantification of immunoblotting data.

Co-immunoprecipitation (co-IP)

Total cellular proteins were extracted using homogenization buffer (Tris 20 mM, pH 7.4, NaCl 150 mM, EDTA 1 mM, 0.5% Triton, and DTT 1 mM) supplemented with a cocktail of protease inhibitors. After centrifugation at 21,000 g for 30 min at 4 °C, the supernatant was saved and protein concentration was measured by BCA protein assay (Beyotime, Shanghai, China). Extract containing 500 μg of protein was incubated at 4 °C for 3 h with 3 μg of goat polyclonal antibody against CCK-BR (1: 500, Abcam). Protein A-Sepharose beads (Amersham Biosciences) were added to the samples and gently shaken for 4 h at 4 °C. Beads were then rinsed and removed in lysis buffer. The pellet was boiled with 4 × Laemmli sample buffer and separated by SDS-PAGE. Immunoreactive proteins on membranes were developed as described above.

Immunofluorescence staining

Immunohistochemistry was performed as previously described [20, 23]. Tissue samples were sectioned into 15 μm thin slices using a cryostat (CM 1950; Leica, San Jose, USA). The sections were blocked with 5% normal goat serum in PBS, plus 0.2% Triton X-100 for 1 h at room temperature then incubated overnight at 4 °C with primary antibody against CCK-BR (goat, 1: 500, Abcam), antibody against NF-200 (mouse, 1:1000, Abcam), or antibody against CGRP (mouse, 1:500, Abcam). Sections were washed three times with PBS at room temperature, followed by Cy3-conjugated donkey anti-goat IgG (1:500, BBI Life Science), FITC-conjugated donkey anti-mouse IgG (1:200, BBI Life Science) or FITC-IB4 (1:200, Sigma) in PBS at room temperature for 2 h. After sections were washed three times with PBS at room temperature, images were captured with a fluorescence microscope (Nikon 104c, Japan). Negative controls, omitting each primary antibody, were used in each case, and no significant staining was observed in these samples (data not shown).

PKA activity assay

PKA activity in homogenates was determined by enzyme-linked immuno sorbent assay (ELISA, Promega), according to the manufacture’s instructions. Briefly, the cells were pretreated with either vehicle or KT-5720 for 30 min, followed by treatment with either vehicle (0.1% DMSO), or forskolin for 15 min. The cells were washed with ice-cold phosphate-buffered saline (PBS), placed on ice, and incubated with 200 μl lysis buffer. After a 10-min incubation on ice, the cells were transferred to microcentrifuge tubes. Cell lysates were centrifuged for 15 min, and aliquots of the supernatants containing 0.2 μg of protein were assayed for PKA activity. The activity is expressed as RLU− 1 (relative light units)/amount of protein.

PI3K activity assay

Cells were stimulated with or without CCK-8 (100 nM) for 15 min. After stimulation, PI3K activity in homogenates was determined with a PI3-Kinase HTRF™ Assay kit (Millipore Corporation, Bedford, MA), using 20 μg of protein for each sample, as stated in the manufacturer’s protocol. HTRF was then measured with an excitation wavelength of 335 nm and emission wave length of 620 and 665 nm with a spectrofluorometer (Tecan, Infinite M1000, Salzburg, Austria).

Behavioral test

Behavioral testing was conducted in an appropriately lighted, quiet room, always during the light cycle between 9:00 AM and 4:00 PM in a series and by the same experimenter. The operator who performed the behavioral tests was blinded to all treatments. Animals were allowed to acclimate to a testing room for at least 30 min before performing the assessments. Mechanical sensitivity was determined on paw withdrawal to manual application of graded von Frey hairs (0.02–2.56 g; Stoelting) to the plantar surface as described previously. Thermal sensitivity was tested using a commercially available paw thermal stimulation system (IITC Life Sciences), and are expressed as paw-withdrawal latency (PWL) and tail-flick latency. Animals were gently dropped into an acrylic box with a metal floor that was preheated to a certain temperature. The values of PWL were calculated using a timer that was started when the animal is released onto the preheated plate and stopped at the moment of withdrawal, shaking, or licking of either hind paw. The cutoff latency was set to prevent tissue damage. All animals were tested once for each temperature per session in a random sequence. All drugs or vehicle were injected subcutaneously into the plantar surface of one hind paw in a volume of 10 μl. The pH of the solutions was adjusted at 7.4 to prevent skin irritation.

Materials

All drugs were purchased from Sigma (MO, USA) unless otherwise indicated. Stock solutions of 4-aminopyridine (4-AP), pertussis toxin (PTX), cholera toxin (CTX), PMA (Tocris Bioscience, Ellisville, MO) and ω-conotoxin MVIIC (Tocris Bioscience, Ellisville, MO) were prepared in distilled deionized water. Z941 was a kind gift from Dr. Terrance P. Snutch (University of British Columbia, Vancouver, British Columbia, Canada). Stock solutions of cholecystokinin-8 (Tocris Bioscience, Ellisville, MO), LY294002, CCK-4, nifedipine, forskolin, gallein, wortmannin, KT5720 (RD system), devazepide, LY225910, GW5823, BC264, SP600125, SB203580, anisomycin, PP2, PP3, Akt inhibitor III, U0126, and GF109203X were prepared in dimethylsulfoxide (DMSO). The final concentration of DMSO in the bath solution was estimated to be less than 0.01%, and this compound had no functional effects on IA (not shown).

Data analysis

In electrophysiological experiments, data acquisition and analysis were performed with Clampfit 10.2 (Axon Instruments) and/or GraphPad Prism 5.0 software (Prism Software). The amplitude of IA was measured at the peak. The data plots were fit by the Boltzmann equation for the activation and inactivation curves as described previously [24]. All data are presented as means ± SEM. Statistical significance between two groups was determined using a paired or unpaired Student’s t two-tailed test. Comparisons of multiple groups against a pooled control were tested using one-way analysis of variance (ANOVA) followed by a Bonferroni’s post-test. Differences in values over time among groups were done using two-way ANOVA. The criterion for significance in all analyses was considered as p < 0.05.

Results

CCK-8 selectively decreased I A in DRG neurons

The studies in vitro of nociceptive processing usually examined different subtypes of peripheral sensory neurons [25, 26]. In the present study, we limited patch-clamp recordings to small-sized DRG neurons (< 30 μm in soma diameter) as these neurons are primarily involved in nociceptive signaling [19, 26]. Two main types of outward voltage-gated K+ channel (Kv) currents have been characterized in these nociceptive neurons — the transient A-type K+ channel currents (IA) and the sustained and delayed-rectifier K+ channel currents (IDR) [15, 16]. We first isolated these two kinetically different whole-cell currents. A total outward current exhibiting a rapidly inactivating and a more sustained component was elicited by a depolarizing pulse from the holding potential of − 80 mV to + 40 mV (Fig. 1a). Biophysical separation of a delayed-rectifier current (IDR) was obtained by a depolarizing prepulse to − 10 mV, which inactivated the transient channels. IA was then isolated by subtracting IDR from the total current (Fig. 1a). Addition of 5 mM 4-aminopyridine (4-AP) inhibited the remaining outward current by 87.1 ± 5.3% (n = 6, Fig. 1b), further confirming the effective isolation of IA.

Fig. 1
figure 1

CCK-8 selectively decreased IA. a, isolation of IA in mouse DRG neurons. IA was isolated by digital subtraction of current traces in which the transient outward K+ current had been inactivated by a brief delay at − 10 mV (100 ms, see Insets) from corresponding current traces without such a delay. This two-step voltage protocol was used for the isolation of IA indicated in all subsequent experiments. b, effects of 5 mM 4-AP on the current density of IA (n = 6). The current density was calculated as the ratio of peak current to membrane capacity (pA/pF). c, d time course of changes in IA (c) or IDR (d) peak amplitude mediated by 100 nM CCK-8. Insets show the exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. e, bar graph showing that 100 nM CCK-8 selectively decreased the current density of IA indicated in panels B (n = 8) and C (n = 10) respectively. f a concentration-response curve for CCK-8 is displayed. Mean values on concentration-response curves were fitted to the sigmoidal Hill equation: PD ([CCK-8]) = PDmax/ (1 + (EC50/[CCK-8])n), where PDmax is the maximal percent decrease of peak IA, EC50 is the concentration that produces half-maximum effect occurs and n is the Hill coefficient. Cell numbers at each concentration were shown in round brackets. g, h, CCK-8 did not significantly alter the steady-state activation curve of IA (n = 9, g), but shifted the steady-state inactivation curve of IA leftward (n = 12, f). i, summary data showing the effects of 100 nM CCK-8 on Vhalf of the activation and inactivation curves. Voltage-dependent activation was measured with voltage commands ranging from − 70 to + 70 mV (400 ms, in 10 mV increment). Steady-state voltage-dependent inactivation of IA was determined by varying a 150-ms conditioning prepulse from − 120 to + 20 mV followed by a 500-ms voltage step pulse to + 40 mV. *p < 0.05, **p < 0.01 and ***p < 0.001 vs. control

Application of 100 nM CCK-8 to small-sized DRG neurons significantly decreased IA by 30.9 ± 3.7% (n = 8, Figs. 1c and e), while IDR was not effectively affected (decreased by 1.2 ± 0.9%, n = 10, Figs. 1d and e). The amplitude of IA partially recovered after CCK-8 washout (Fig. 1c). The CCK-8 effect on IA was concentration-dependent (Fig. 1f). The half-maximal inhibitory concentration (IC50) calculated from a sigmoidal Hill equation [23, 24] observed at 47.3 nM (Fig. 1f). Further, we examined whether CCK-8 would alter the biophysical properties of IA. While no significant changes were observed in the activation properties of IA (Vhalf from 5.8 ± 1.6 mV to 6.5 ± 2.5 mV, n = 9, Figs. 1h and i), CCK-8 shifted the steady-state inactivation curve to the hyperpolarized level by 7.8 mV (Vhalf from − 55.9 ± 3.9 mV to − 63.7 ± 2.8 mV, n = 12, Figs. 1g and i). These findings reveal that the CCK-8-induced reduction in IA is mainly contributed by an increased proportion of channels retained in the inactivated state.

The CCK-BR mediated the CCK-8–induced I A decrease

The CCK-AR and CCK-BR have been identified as the endogenous receptors for CCK-8 [27]. To determine which one is involved in the CCK-8–induced IA reduction, we first examined the protein profile and subcellular expression of these receptors in mouse DRGs. Immunoblot analysis revealed that only CCK-BR (predicted size of 80 kDa), but not CCK-AR (predicted size of 95 kDa), were endogenously expressed (Fig. 2a). Protein lysates prepared from the gallbladder of the same mice were used as a positive control. Small unmyelinated sensory neurons have been classified into isolectin B4 (IB4)-positive (non-peptidergic) subset and peptidergic (IB4 negative) subset expressing calcitonin gene-related peptide (CGRP), while large neurons in myelinated A-fibers express neurofilament 200 (NF200). We analyzed the CCK-BR expression in DRGs subsets by coimmunostaining of CCK-BR with the mentioned markers. CCK-BRs were found to be heavily colocalized with IB4 and CGRP, and less with neurofilament-200 (NF-200), a marker for myelinated A-fibers (Fig. 2b). Next, we determined the participation of CCK-BR in the effect of CCK-8 on IA. While the CCK-8 mediated reduction of currents was not affected by the presence of 1 μM of the CCK-AR antagonist devazepide (decreased by 29.3 ± 3.7%, n = 10, Figs. 2c and e), such effect was completely abolished in the presence of 1 μM of the CCK-BR antagonist LY225910 (decreased by 4.7 ± 0.9%, n = 10, Figs. 2d and e). This evidence undoubtably indicates that the CCK-BR, but not the CCK-AR, is involved in the CCK-8-induced IA reductions. Furthermore, application of 0.1 μM BC-264, a selective CCK-BR agonist, significantly decreased IA by 30.9 ± 2.7% (n = 7, Fig. 2f), while the selective CCK-AR agonist GW5823 (5 μM) elicited no such effect (decreased by 3.1 ± 1.2%, n = 9, Fig. 2f). Since it is known that both CCK-4 and CCK-8 are active forms of CCKs in the nervous system [28, 29], we also test whether application of another selective CCK-BR agonist CCK-4 affects IA in small-sized DRG neurons. Indeed, CCK-4 at 300 nM significantly decreased IA by 32.2 ± 4.9% (n = 7). These findings further support the conclusion that the CCK agonist driven IA decrease was mediated specifically by CCK-BR in small-sized DRG neurons.

Fig. 2
figure 2

CCK-8 decreased IA via activation of CCK-BR. a, western blot analysis demonstrated the expression of CCK-AR and CCK-BR protein in mouse DRGs. GAPDH is shown as loading control. Representative western blots are shown from at least three independent experiments. b, co-localization of CCK-BR (Red) with three markers (Green) (CGRP, IB4, and NF-200) in naïve mouse DRGs. Arrows show the co-localization. Scale bars: 30 μm. c, d, time course of changes in IA amplitude mediated by 100 nM CCK-8 in the presence of either devazepide (1 μM, n = 10, c) or LY225910 (1 μM, n = 8, d). Insets show the exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. e, bar graph showing that treatment of DRG neurons with LY225910, instead of devazepide, abrogated the CCK-8-induced IA response. f, bar graph showing that application of BC264 (0.1 μM, n = 7) or CCK-4 (300 nM, n = 7), but not GW5823 (5 μM, n = 9), significantly decrease IA in small DRG neurons. *p < 0.05 and **p < 0.01 vs. control

The CCK-BR-mediated IA decrease requires the βγ subunits (Gβγ) of go-protein

CCK-BR coupled to heterotrimeric G-proteins, which are key transducers to control numerous cellular processes [3]. We next examined actions of different subtypes of G-proteins in the CCK-BR-mediated IA modulation. Inactivation of Gs by pre-treating DRG neurons with cholera toxin (CTX, 500 ng/ml) had no significant effects on the CCK-8-induced IA decrease (decreased by 32.3 ± 3.6%, n = 11, Fig. 3a and c). Contrastingly, pre-treating cells with pertussis toxin (PTX, 200 ng/ml) to inactivate Gi/o abrogated the CCK-8-induced response (decreased by 3.9 ± 2.2%, n = 7, Fig. 3b and c). This CCK-BR-induced PTX-sensitive, but CTX-insensitive decrease in IA, indicated the involvement of Gi/o, but not Gs in the signaling cascade. Further, dialysis of cells with an antibody specifically against Gαo (2 μg/ml) blocked the effect of CCK-8 on IA reduction (decreased by 1.5 ± 2.9%, n = 8, Fig. 3d), whereas a Gαi-specific antibody (2 μg/ml) had no such effect (decreased by 27.6 ± 4.7%, n = 9, Fig. 3d). Together, these findings suggest that Gαo-protein mediates the response to CCK-8. Moreover, we found that endogenous Gαo (Fig. 3e), but not Gαi (Fig. 3f), was co-immunoprecipitated with an antibody against CCK-BR from DRG tissues, indicating that the CCK-BR and the Gαo subunit form a complex in situ. Further, intracellular application of a Gβ-specific antibody abrogated the CCK-8-induced IA reduction (decreased by 3.3 ± 3.5%, n = 9; Figs. 3g and h), while its denatured form did not elicit such effects (decreased by 32.5 ± 1.3%, n = 8; Fig. 3h). Similar results were obtained with a specific Gβγ inhibitor, gallein. Pretreatment of cells with gallein (10 μM) completely abolished the CCK-8-induced IA reduction (decreased by 2.7 ± 0.9%, n = 9; Fig. 3h). Thus, the Gβγ subunit of Go-protein is also required for the CCK-BR-mediated IA reductions.

Fig. 3
figure 3

The CCK-BR-mediated IA decrease requires the βγ subunits (Gβγ) of Go protein. a, b, time courses showing the effect of 100 nM CCK-8 on the peak amplitude of IA in cells pretreated with CTX (500 ng/ml, 16 h pretreatment, n = 11, a) and PTX (200 ng/ml, 16 h pretreatment, n = 7, b), respectively. Insets indicate the exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. c, summary of results showing the effect of CCK-8 (100 nM) on IA in CTX- or PTX-pretreated DRG neurons. d, bar graph indicating the effect of CCK-8 (100 nM) on IA in cells dialyzed with an antibody specific against Gαi (anti-Gi-Ab, 2 μg/ml, n = 8) and anti-Go-Ab (2 μg/ml, n = 9), respectively. e, f, the association of Gαo with the CCK-BR. Co-IP studies reveal that Gαo (e), but not Gαi (f), interacts with CCK-BR in mouse DRG extracts. Representative western blots are shown from at least three independent experiments. g, time course showing the effect of 100 nM CCK-8 on IA in the presence of anti-Gβ-Ab (intracellular application, 2 μg/ml). Insets show exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. h, bar graph showing that intracellular application of gallein (10 μM, n = 9) or anti-Gβ-Ab (n = 9), but not the denatured form of anti-Gβ-Ab (n = 8), abolished the CCK-8-induced IA decrease. *p < 0.05, **p < 0.01 and ***p < 0.001 vs. control

The CCK-BR–mediated I A decrease requires PI3K, but independently of Akt

In view of the fact that protein kinase C (PKC) has been shown to act downstream of Gβγ [30] and regulate IA superficial dorsal horn neurons [31], we pre-incubated cells with PKC inhibitors and found that pretreating with GF109203X (1 μM) did not affect the CCK-8-induced IA response (decreased by 31.8 ± 4.3%, n = 9, Fig. 4a) while pre-incubation of cells with GF109203X substantially blocked the PKC activator PMA (phorbol 12-myristate 13-acetate)-induced IA reduction (5 μM, n = 9, Fig. 4b). Previous studies have highlighted the critical role of PI3K/Akt cascades in Gβγ–mediated responses [32]. Thus, we investigated whether the inhibitory effect of CCK-8 on IA was PI3K/Akt-dependent. We found that CCK-8 application significantly induced PI3K activation and that pretreating cells with the PI3K inhibitor LY294002 at 20 μM abolished this effect (Fig. 4c). Consistently, pre-treatment with the PI3K inhibitor LY294002 (20 μM) (decreased by 1.9 ± 2.7%, n = 8, Fig. 4d and f) or wortmannin (1 μM) (decreased by 2.7 ± 3.5%, n = 10, Figs. 4e and f) also prevented the CCK-8 effects on IA, indicating the involvement of PI3K in the CCK-BR–mediated IA decrease. Further, we examined whether the CCK-8 action is also mediated by Akt, a major downstream target of the PI3Ks [33]. We measured the Akt activity in DRG cells and found that 100 nM CCK-8 significantly increased the phosphorylated Akt (p-Akt) level, while the total Akt (t-Akt) remained unchanged (Fig. 4g). This effect was abrogated by the Akt inhibitor III (10 μM, Fig. 4g). To further determine the involvement of Akt in the modulation of IA by CCK-8, cells were pretreated with Akt inhibitor III prior to CCK-8 application. Interestingly, in the presence of 10 μM Akt inhibitor III, CCK-8 at 100 nM still induced a significant decrease in IA (decreased by 28.5 ± 3.8%, n = 9, Fig. 4h), revealing that the CCK-8–induced IA decrease was mediated by PI3K, but independently of Akt.

Fig. 4
figure 4

The CCK-BR-mediated IA decrease requires PI3K, but not Akt. a, time course of changes in IA amplitude (left) and summary data (right) showing the effects of 100 nM CCK-8 on IA in the presence of GF109203X (1 μM for 30 min, n = 9). Insets show the representative current traces. The numbers on the plot indicate which points were used for sample traces. b, representative traces (left) and bar graph (right) indicating the effect of 5 μM PMA on IA in the absence (n = 5) or presence (n = 9) of 1 μM GF109203X. c, pretreating cells with 20 μM LY294002 prevented the CCK-8-induced increase in PI3K activity. The experiments were conducted in triplicate and yielded with similar results. d, e, time course showing the effect of CCK-8 (100 nM) on IA in the presence of LY294002 (20 μM for 30 min, d) or wortmannin (1 μM for 30 min, e). Insets show exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. f, bar graph showing the effects of CCK-8 on IA in the presence of LY294002 (n = 8) or wortmannin (n = 10) indicated in panels d and e respectively. g, CCK-8 induced a significant increase in the phosphorylated Akt (p-Akt) in DRG cells. This effect was abrogated by pretreating cells with LY294002 (20 μM for 30 min) or the Akt inhibitor III (10 μM for 30 min). h, left: representative traces and time course indicating the inhibitory effects of 100 nM CCK-8 on IA in the presence of Akt inhibitor III (10 μM). Insets show exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. Right: bar graph showing that treatment of DRG neurons with Akt inhibitor III (n = 9) had no effect on the CCK-8–induced IA response. **p < 0.01 and ***p < 0.001 vs. control

CCK-8 attenuates I A through c-Src-dependent JNK pathway

Mitogen-activated protein kinases (MAPKs), composing a family of protein kinases that play pivotal roles in mediating pain sensitivity [34], were shown to regulate neuronal IA [31]. Thus, it was of interest to the current study to examine whether the MAPK cascades would be involved in the CCK-BR-induced response. Immunoblot analysis indicated that the exposure of DRG cells to CCK-8 (100 nM) significantly increased the expression of phosphorylated c-Jun N-terminal kinase (p-JNK), while the protein levels of total JNK (t-JNK), p-p38 as well as p-ERK activity remained unchanged (Fig. 5a). Blockade of CCK-BR with LY225910 (1 μM), as well as pretreating cells with the PI3K antagonist LY294002 (20 μM), eliminated the CCK-8–induced JNK activation (Fig. 5b). These findings reveal that the PI3K-mediated JNK signaling was involved in the CCK-8–induced effects. Next, we pretreated cells with 10 μM SP600125, a specific JNK inhibitor, and found that SP600125 abrogated the IA decrease induced by CCK-8 (decreased by 0.7 ± 1.9%, n = 7, Figs. 5c and d). Contrastingly, U0126 (1 μM), a MAPK/ERK (MEK) inhibitor, as well as the p-38 inhibitor SB203580 (10 μM), elicited no such effects (U0126: decreased by 28.3 ± 5.5%, n = 9; SB203580: decreased by 29.3 ± 3.8%, n = 11; Fig. 5d). As complementary support, the application of the JNK agonist anisomycin (25 ng/ml) to DRG neurons induced a significant decrease in IA amplitude (decreased by 32.7 ± 3.9%, n = 11, Fig. 5e).

Fig. 5
figure 5

CCK-8 attenuates IA through Src-dependent JNK pathway. a CCK-8 increased the protein levels of phosphorylated JNK (p-JNK), with no significant changes in p-p38 and p-ERK. Representative western blots are shown from at least three independent experiments. b activation of JNK by CCK-8 is via CCK-BR and requires PI3K activity. Pretreatment of cells with LY225910 (1 μM) or LY294002 (20 μM) abolished CCK-8–induced JNK phosphorylation. Representative western blots are shown from at least three independent experiments. c, time course showing the effect of 100 nM CCK-8 on IA in the presence of SP600125. Insets show representative current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. d bar graph showing the effects of CCK-8 (100 nM) on IA in the presence of SP600125 (10 μM, n = 7), U0126 (1 μM, n = 9), and SB203580 (10 μM, n = 11), respectively. e Time course (left panel) and summary of results (right panel) indicating that application of anisomycin (25 ng/ml, n = 11) significantly decreased peak IA amplitude in small DRG neurons. f time course (left panel) and summary of results (right panel) indicating the effects of 100 nM CCK-8 on IA in the presence of KT-5720 (1 μM for 30 min, n = 10). Insets show exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. g bar graph showing the effects of 20 μM forskolin on PKA activity in DRG cells pretreated with KT-5720 (1 μM). The experiments were conducted in triplicate and yielded with similar results. h the level of phosphorylated Src (pTyr418, p-Src) increased following treatment with CCK-8 (100 nM). This effect was abolished by the PI3K inhibitor LY294002 (20 μM for 30 min). Representative western blots are shown from at least three independent experiments. i CCK-8-induced JNK activation is blocked by the Src inhibitor PP2 (10 μM). PP2 or its inactive structure analog PP3 (10 μM) was pre-administered for 30 min before CCK-8 addition. Representative western blots are shown from at least three independent experiments. j time course indicating the effects of CCK-8 on IA in the presence of 10 μM PP2. Insets show exemplary current traces. The Arabic numerals indicate the relative points utilized for exemplary current traces. k bar graph indicating that application of PP2 (10 μM for 30 min, n = 12), but not PP3 (10 μM for 30 min, n = 8), abolished the CCK-8-induced IA decrease. * p < 0.05 and ** p < 0.01 vs. control; # p < 0.05 vs. vehicle; & p < 0.05 vs. forskolin without KT-5720

Protein kinase A (PKA) was involved in the regulation of Kv channel currents [24, 31] and has been suggested to mediate the crosstalk between the PI3K and MAPK pathways [35]. Thus, we asked whether the JNK-dependent regulation of IA by CCK-8 required PKA. Pretreatment of cells with the PKA inhibitor, KT-5720 (1 μM), had no significant effect on the CCK-8–induced IA decrease (decreased by 29.1 ± 3.2%, n = 10, Fig. 5f), while the administration of KT-5720 (1 μM) blocked 20 μM forskolin-induced PKA activation (Fig. 5g), indicating a PKA-independent mechanism involved in CCK-BR-mediated response. Src kinase (Src) has been demonstrated to activate JNK pathways [36, 37]. We assayed the cellular Src activity in DRG cells treated with CCK-8. Figure 5h illustrates that phosphorylated Src (pTyr418, p-Src) increased following treatment with CCK-8 (100 nM), whereas the protein levels of total Src (t-Src) was unchanged (Fig. 5h). This effect was abolished by the PI3K inhibitor LY294002 (20 μM, Fig. 5h). Moreover, cells were treated with the Src-specific inhibitor PP2 (10 μM) prior to CCK-8 exposure, and the JNK activation was monitored. Pretreatment of the cells with the Src inhibitor PP2 (10 μM) completely abolished CCK-8-induced JNK activation, while the inactively structural analog PP3 (10 μM) elicited no such effect (Fig. 5i). Consistent with this, pretreatment of cells with PP2 (10 μM) (decreased by 0.6 ± 2.7%, n = 12, Fig. 5j and k), but not PP3 (decreased by 26.9 ± 3.1%, n = 8, Fig. 5k), completely abolished the CCK-8-induced IA decrease. Collectively, these results suggest that Src, but not PKA, mediated the signaling between PI3K and JNK in the CCK-BR-mediated IA response.

Activation of CCK-BRs induces DRG neuronal hyperexcitability

Kv exerts pivotal effects in modulating neuronal excitability in peripheral sensory neurons [38]. To determine the functional roles of the CCK-BR-mediated IA response, we determined whether the membrane excitability of DRG neurons would be affected by CCK-8. Bath application of CCK-8 (100 nM) had no significant effects on the whole-cell currents of Nav (n = 11, Fig. 6a) and the high voltage-activated (HVA) calcium channel currents (n = 8, Fig. 6b) in small DRG neurons, whereas CCK-8 increased LVA (T-type) channel currents by 8.3% (n = 9, Fig. 6c). Using an external solution including Z941 (10 μM) to block T-type channels, we found that 100 nM CCK-8 significantly increased action potential (AP) firing in response to 1-s current injection (by 56.6 ± 2.9% compared to control, n = 17, Fig. 6d and e). After washout, the firing rate was partially restored (Fig. 6e). Additionally, CCK-8 (100 nM) significantly shortened the first spike latency (Fig. 6f) and decreased AP threshold (n = 17, Fig. 6g). The other membrane properties of neuronal excitability, including resting membrane potential, were not significantly changed by 100 nM CCK-8 (not shown). Pretreating neurons with LY225910 (1 μM) abrogated the CCK-8-induced increase in AP firing rate, indicating the CCK-BR involvement (n = 12, Fig. 6h). To further verify the CCK-BR-induced neuronal hyperexcitability through IA decrease, 4-AP was applied prior to CCK-8. Pre-treatment of DRG neurons with 5 mM 4-AP abrogated the neuronal hyperexcitability induced by 100 nM CCK-8 (n = 12, Fig. 6j and k), indicating that the CCK-BR-mediated IA decrease subsequently induced neuronal hyperexcitability in small DRG neurons.

Fig. 6
figure 6

CCK-8 induces neuronal hyperexcitability. a-c exemplary current traces (left) and bar graph (right) indicating the effect of 100 nM CCK-8 on Nav currents (INa, n = 11, A), high-voltage-activated Cav currents (IBa of HVGCC, n = 8, b), or low-voltage-activated Cav currents (IBa of LVGCC, n = 9, c), respectively. Either Nav currents or IBa of HVGCC were elicited by a test pulse to 0 mV from a holding potential of − 60 mV. A stepped voltage protocol from − 110 to − 40 mV with a holding potential of − 110 mV was applied to elicit IBa of LVGCC. d, e exemplary traces (d) and summary of results (e, n = 17) indicating the effect of CCK-8 (100 nM) on action potential firing rate. Representative traces were recorded when small-sized DRG neurons were subjected to 130 pA current injections. f, g CCK-8 at 100 nM significantly shortened first spike latency (f) and decreased the AP threshold (g) in small-sized DRG neurons (n = 17). h, i representative traces (h) and summary of results (i) showing that pretreatment of cells with LY225910 (1 μM, n = 12) abolished the 100 nM CCK-8-induced the increase in firing rate. Representative traces were recorded when small-sized DRG neurons were subjected to 130 pA current injections. j, k exemplary current traces (j) and summary of results (k) indicating that application of 4-AP at 5 mM abrogated the 100 nM CCK-8-induced neuronal hyperexcitability (n = 12). Representative traces were recorded when small-sized DRG neurons were subjected to 80 pA current injections. *p < 0.05 and **p < 0.01 vs. control

Involvement of A-type channels encoding IA in CCK-induced pain hypersensitivity

Further, we determined whether CCK-8 would affect in vivo pain sensitivity in animals. Intraplantar injection of CCK-8 (50 ng) markedly increased pain sensitivity to both mechanical and heating stimuli (Figs. 7a and b). The CCK-8-induced pain hypersensitivity to mechanical or heating stimulation was completely abrogated by intraplantar pretreatment of the CCK-BR antagonist LY225910 (0.5 μg, Figs. 7c and d), but not by the CCK-AR antagonist devazepide (1 μg, Figs. 7c and d). Moreover, intraplantar pretreatment with 4-AP (25 nmol) induced a significant increase in mechanical and heat sensitivity as compared with animals received a saline injection (Figs. 7e and f). Sensitivity assessed after intraplantar injection of CCK-8 showed that CCK-8 did not induce any additive effects to that of 4-AP on mechanical (Fig. 7e) and thermal (Fig. 7f) pain sensitivity, strongly suggesting that CCK-8 and 4-AP likely target molecules in the same cellular signaling pathway in vivo. Collectively, these findings reveal that A-type channels encoding IA contribute to the CCK-BR–mediated acute pain hypersensitivity.

Fig. 7
figure 7

Involvement of peripheral CCK-BR activation in pain hypersensitivity. a,b intraplantar injection (i.p.l.) of CCK-8 (50 ng) induced mechanical (a) and heat (b) pain hypersensitivity. *p < 0.05, **p < 0.01, CCK-8 injection vs. vehicle. c, d pretreatment of LY225910 (0.5 μg, i.p.l., c), but not devezepide (1 μg, i.p.l., d), significantly attenuated CCK-8-induced mechanical hypersensitivity and thermal hyperalgesia. e, f intraplantar pre-injection of 4-AP (25 nmol) occluded the thermal and mechanical hypersensitivity induced by CCK-8 (50 ng). *p < 0.05, **p < 0.01, CCK-8 injection vs. vehicle; + p < 0.05 and ++ p < 0.01, CCK-8 injection vs. CCK-8 + devazepide (Dev.) at 45 min; #p < 0.05 and ##p < 0.01, 4-AP injection vs. vehicle at 45 min. N = at least 8 mice for all animal behavior experiments

Discussion

The present study provides mechanistic data describing a novel functional role of CCK-8 in modulating transient IA in small-sized DRG neurons, without any concurrent effect on IDR. Based on our findings, we propose a signaling cascade model in which CCK-8–stimulated PI3K recruits the Src-dependent JNK to suppress IA. This attenuation of IA induced by CCK-8 application is mediated by the stimulation of CCK-BR and leads to sensory neuronal hyperexcitability and pain hypersensitivity in mice (see Fig. 8).

Fig. 8
figure 8

Schematic shows the regulation of CCK-BR on IA and the involvement of CCK-8/CCK-BR in pain sensitivity. CCK-8 stimulates the Go-protein coupled CCK-BR and thereafter releases the βγ subunits (Gβγ). The released Gβγ subsequently activates PI3K, which decreases the IA and induces neuronal hyperexcitability and pain hypersensitivity. PI3K catalyzes the conversion of PtdIns(4,5)P2 (PIP2) to PtdIns(3,4,5)P3 (PIP3), which serves as a second messenger that helps to activate Akt. However, neither PKA/PKC/Akt nor the direct binding of Gβγ with A-type channels contributes to the CCK-BR-mediated IA response. In mouse DRG neurons, PI3K signaling may activate Src, which then phosphorylates JNK to modulate IA. Whether the activated p-JNK would phosphorylates Kv channels encoding IA or in turn stimulated intermediate molecules still needs further examined

The PKC family of isozymes mediates IA responses in a cell-type and tissue-specific manner. For instance, activation of group I metabotropic glutamate receptors led to an inhibition of IA through a PKC-dependent mechanism in striatal cholinergic interneurons, while in large aspiny neurons activation of PKCα increases IA. Interestingly, in murine proximal colonic myocytes, the PKC-independent regulation of IA has also been reported [39]. In this study, the CCK-8-induced IA decrease was independent of PKC and was mediated by PI3K, through the JNK-dependent signaling. These results are supported by previous studies that Kv currents including IA recorded from trigeminal ganglion neurons and pancreatic β cells decreased in response to PI3K pathway activation [21, 40]. Interestingly, the activation of PI3K has also been reported to increase IA in cultured rat cerebellar granule cells [41]. In addition, PI3K-induced activation of Kv4.3 channels through glucocorticoid-inducible kinase-1 (SGK1) was also reported [42]. Although these discrepancies require further clarification, the regulatory effects of PI3K would be variable in tissues/cell types expressing different A-type channel subunits. Another appropriate alternative hypothesis is that the stimulatory PI3K can also phosphorylate an intermediate protein that in turn down-regulates IA in small-sized DRG neurons. Furthermore, different splice variants of KChIP auxiliary subunits of IA channels can engender different, even opposing, modulation of Kv4 channel currents [43].

A known target of Gβγ is PI3K [33]. In contrast to many other common Gβγ-dependent PI3K signal transduction events, the CCK-8–induced PI3K dependent IA attenuation isnot mediated by Akt, as demonstrated by the specific inhibition of Akt with pharmacological agents. Interestingly, previous studies have shown that Akt both negatively and positively regulates Kv4 [44, 45], which forms one of the major components mediating IA. For example, Akt down-regulates the activity of Kv4 channels in cultured cerebellar granule cells of rats [45]; in the same neurons, a different study demonstrates that enhanced Akt activity is required for IA amplification and Kv4.2 induction [44]. This Akt-dependent stimulation of Kv4 also occurs in the arcuate nucleus [46]. Thus, it appears that Akt differentially regulates the activity of Kv4 channels in a tissue-specific manner. In our study, the CCK-BR-mediated IA response was found to be independent of Akt; therefore, we went on to investigate what mediating PI3K signals to suppress IA in DRG neurons. Considerable in vivo and in vitro studies indicate that ERK plays pivotal roles in neuropathic pain [34, 47]. Phosphorylated ERK is elevated in DRG cells following peripheral nerve injury [48]. Intrathecal application of ERK inhibitors reduces the pain behavior associated with nerve injury [49]. Moreover, one of the most convincing evidence comes from the direct phosphorylation of the pore-forming channel subunit of Kv4.2 by ERK [50] that determine a downregulation of IA in superficial dorsal horn neurons [31]. Contrastingly, antagonism of ERK completely abrogated IA increase induced by dopamine in lateral pyloric neurons [51]. However, we found that the CCK-8-induced decrease of IA was unlikely induced by ERK phosphorylation, because the CCK-8 application did not change the ERK activity in DRG cells, whereas the levels of p-JNK were significantly increased. Moreover, the MAPK/ERK inhibitor did not affect the CCK-8–induced IA response. Our findings suggested that PI3K stimulated JNK in DRG neurons and that this signaling is essential for the CCK-BR–mediated IA response. Our results showed that 1) application of the JNK inhibitor SP600125, but not the p38 MAPK inhibitor SB203580 or the MAPK/ERK (MEK) inhibitor U016, blocked the CCK-8–induced IA decrease and 2) antagonism of PI3K blocked CCK-BR–mediated JNK activation. Consistent with these findings, the increased activity of JNK in ventricular myocytes markedly decreased the amplitude of transient outward K+ current density [52]. These observations are in line with an earlier study showing a C-reactive protein (CRP)-induced modulation of intracellular JNK and interactions with voltage-activated K+ channels [53].

Up till now, it is still relatively unclear how PI3K activates JNK. It has been established that that PI3K may stimulate PKA, subsequently activating the downstream MAPK pathway [35]. In the present study, activation of CCK-BR did not influence the PKA activity in DRG cells, indicating some other mechanisms, but not of PKA, mediate the crosstalk between PI3K and JNK signaling. Src kinases are downstream of PI3K and can facilitate JNK activity [37], suggesting possible crosstalk between PI3K and JNK signaling. In support of this observation, the current study demonstrated that the Src kinase inhibitor PP2 blocked the CCK-8–induced JNK activation. The blockade of PI3K also abolished the CCK-BR-mediated increase in Src activity, indicating that PI3K may modulate the JNK pathway through c-Src. Therefore, it is likely that CCK-8–activated PI3K recruits Src to up-regulate JNK activity, and thereby regulating CCK-BR–mediated IA response in DRG neurons.

IA, encoded by A-type K+ channels, is important determinants of both the delay of spike onset (first spike latency) and the decrease in the firing frequency [17]. Acute decreases in IA in sensory neurons cause robust increases in neuronal excitability, [38] and may increase the responsiveness to nociceptive stimulation and contribute to mechanical hypersensitivity and thermal hyperalgesia [54]. Genetic studies have firmly established a prominent role for A-type channels in amplifying nociceptive signals in the periphery and in contributing to central sensitization in the spinal dorsal horn [17, 19, 55]. Further, recent evidence has suggested that modulation of peripheral A-type channels influences somatic and visceral nociceptive inputs and thus an increase of A-type channel currents results in significant anti-nociception in a variety of animal neuropathic pain models [19]. In the current study, consistently with the CCK-8-induced IA decrease, activation of CCK-BR led to increased excitability in DRG neurons with increased spike frequency and shortened first-spike latency, both of which are major parameters determining the timing of neurotransmitter release, and hence pain transmission [56]. In addition, acute mechanical hypersensitivity and thermal hyperalgesia mediated by CCK-BR can be occluded by the A-type K+ channel blockade. As such, our findings are supportive of the reasonable assumption that nociceptive actions of CCK-BR are mediated, at least in part, through the JNK-dependent reduction of IA. Our present results are, indeed, in accordance with previous studies that CCK-8 might induce pro-nociceptive actions [8, 9]. Intrathecal inhibition of JNK, a key modulator IA in the present study, has been found to attenuate the CCI-induced mechanical allodynia and thermal hyperalgesia in rats [57]. Following spinal nerve ligation (SNL), phosphorylated JNK in small-sized DRG neurons have been found to be greatly increased [58] and the intrathecal infusion of JNK inhibitor can reverse mechanical but not thermal hypersensitivity [59].

Conclusions

Collectively, this study found that CCK-8 decreases IA through the Gβγ-mediated PI3K/Src/JNK pathway. This mechanism occurred via CCK-BR and mediated the neuronal hyperexcitability in peripheral sensory neurons and pain hypersensitivity in mice. Modulation of IA by CCK-8 in peripheral sensory neurons is of particular interest. The identification of CCK-BR-mediated molecular mechanisms contributing to pain hypersensitization may offer insights into opportunities for analgesic pharmacotherapy.

Availability of data and materials

All data and materials generated in this study are available upon request.

Abbreviations

4-AP:

4-aminopyridine

CCK-8:

Cholecystokinin-8

CCK-BR:

CCK type B receptor

CTX:

Cholera toxin

ERK:

Extracellular regulated protein kinases

GDP-β-S:

Guanosine-5′-O-(2-thiodiphosphate)

GPCR:

G-protein coupled receptor

Gβγ :

G-protein βγ subunit

I A :

A-type K+ current

I DR :

Sustained delayed rectifier K+ current

JNK:

c-Jun N-terminal kinase

MAPK:

Mitogen-activated protein kinase

PI3K:

Phosphatidylinositol 3-kinase

PKA:

Protein kinase A;

PKC:

Protein kinase C

PTX:

Pertussis toxin

References

  1. Crawley JN, Corwin RL. Biological actions of cholecystokinin. Peptides. 1994;15:731–55.

    Article  CAS  Google Scholar 

  2. Dufresne M, Seva C, Fourmy D. Cholecystokinin and gastrin receptors. Physiol Rev. 2006;86:805–47.

    Article  CAS  Google Scholar 

  3. Noble F, Wank SA, Crawley JN, Bradwejn J, Seroogy KB, Hamon M, Roques BP. International Union of Pharmacology. XXI. Structure, distribution, and functions of cholecystokinin receptors. Pharmacol Rev. 1999;51:745–81.

    CAS  PubMed  Google Scholar 

  4. Juarez-Rojop IE, Granados-Soto V, Diaz-Zagoya JC, Flores-Murrieta FJ, Torres-Lopez JE. Involvement of cholecystokinin in peripheral nociceptive sensitization during diabetes in rats as revealed by the formalin response. Pain. 2006;122:118–25.

    Article  CAS  Google Scholar 

  5. Yin K, Deuis JR, Lewis RJ, Vetter I. Transcriptomic and behavioural characterisation of a mouse model of burn pain identify the cholecystokinin 2 receptor as an analgesic target. Mol Pain. 2016;12.

    Article  Google Scholar 

  6. Yang Y, Li Q, He QH, Han JS, Su L, Wan Y. Heteromerization of mu-opioid receptor and cholecystokinin B receptor through the third transmembrane domain of the mu-opioid receptor contributes to the anti-opioid effects of cholecystokinin octapeptide. Exp Mol Med. 2018;50:64.

    Article  Google Scholar 

  7. Xu XJ, Puke MJ, Verge VM, Wiesenfeld-Hallin Z, Hughes J, Hokfelt T. Up-regulation of cholecystokinin in primary sensory neurons is associated with morphine insensitivity in experimental neuropathic pain in the rat. Neurosci Lett. 1993;152:129–32.

    Article  CAS  Google Scholar 

  8. Cao B, Zhang X, Yan N, Chen S, Li Y. Cholecystokinin enhances visceral pain-related affective memory via vagal afferent pathway in rats. Mol Brain. 2012;5:19.

    Article  CAS  Google Scholar 

  9. Sabate JM, Gorbatchef C, Flourie B, Jian R, Coffin B. Cholecystokinin octapeptide increases rectal sensitivity to pain in healthy subjects. Neurogastroenterol Motil. 2002;14:689–95.

    Article  Google Scholar 

  10. van der Schaar PJ, van Hoboken E, Ludidi S, Masclee AA. Effect of cholecystokinin on rectal motor and sensory function in patients with irritable bowel syndrome and healthy controls. Color Dis. 2013;15:e29–34.

    Article  Google Scholar 

  11. Wang EM, Li WT, Yan XJ, Chen X, Liu Q, Feng CC, Cao ZJ, Fang JY, Chen SL. Vagal afferent-dependent cholecystokinin modulation of visceral pain requires central amygdala NMDA-NR2B receptors in rats. Neurogastroenterol Motil. 2015;27:1333–43.

    Article  CAS  Google Scholar 

  12. Kurrikoff K, Koks S, Matsui T, Bourin M, Arend A, Aunapuu M, Vasar E. Deletion of the CCK2 receptor gene reduces mechanical sensitivity and abolishes the development of hyperalgesia in mononeuropathic mice. Eur J Neurosci. 2004;20:1577–86.

    Article  Google Scholar 

  13. Julius D, Basbaum AI. Molecular mechanisms of nociception. Nature. 2001;413:203–10.

    Article  CAS  Google Scholar 

  14. Waxman SG, Zamponi GW. Regulating excitability of peripheral afferents: emerging ion channel targets. Nat Neurosci. 2014;17:153–63.

    Article  CAS  Google Scholar 

  15. Rasband MN, Park EW, Vanderah TW, Lai J, Porreca F, Trimmer JS. Distinct potassium channels on pain-sensing neurons. Proc Natl Acad Sci U S A. 2001;98:13373–8.

    Article  CAS  Google Scholar 

  16. Winkelman DL, Beck CL, Ypey DL, O'Leary ME. Inhibition of the A-type K+ channels of dorsal root ganglion neurons by the long-duration anesthetic butamben. J Pharmacol Exp Ther. 2005;314:1177–86.

    Article  CAS  Google Scholar 

  17. Hu HJ, Carrasquillo Y, Karim F, Jung WE, Nerbonne JM, Schwarz TL, RWt G. The kv4.2 potassium channel subunit is required for pain plasticity. Neuron. 2006;50:89–100.

    Article  CAS  Google Scholar 

  18. Kim DS, Choi JO, Rim HD, Cho HJ. Downregulation of voltage-gated potassium channel alpha gene expression in dorsal root ganglia following chronic constriction injury of the rat sciatic nerve. Brain Res Mol Brain Res. 2002;105:146–52.

    Article  CAS  Google Scholar 

  19. Duan KZ, Xu Q, Zhang XM, Zhao ZQ, Mei YA, Zhang YQ. Targeting A-type K(+) channels in primary sensory neurons for bone cancer pain in a rat model. Pain. 2012;153:562–74.

    Article  CAS  Google Scholar 

  20. Zhang Y, Ji H, Wang J, Sun Y, Qian Z, Jiang X, Snutch TP, Tao J. Melatonin-mediated inhibition of Cav3.2 T-type ca(2+) channels induces sensory neuronal hypoexcitability through the novel protein kinase C-eta isoform. J Pineal Res. 2018;64:e12476.

    Article  Google Scholar 

  21. Wang H, Qin J, Gong S, Feng B, Zhang Y, Tao J. Insulin-like growth factor-1 receptor-mediated inhibition of A-type K(+) current induces sensory neuronal hyperexcitability through the phosphatidylinositol 3-kinase and extracellular signal-regulated kinase 1/2 pathways, independently of Akt. Endocrinology. 2014;155:168–79.

    Article  Google Scholar 

  22. Andre S, Boukhaddaoui H, Campo B, Al-Jumaily M, Mayeux V, Greuet D, Valmier J, Scamps F. Axotomy-induced expression of calcium-activated chloride current in subpopulations of mouse dorsal root ganglion neurons. J Neurophysiol. 2003;90:3764–73.

    Article  CAS  Google Scholar 

  23. Zhang Y, Qin W, Qian Z, Liu X, Wang H, Gong S, Sun YG, Snutch TP, Jiang X, Tao J. Peripheral pain is enhanced by insulin-like growth factor 1 through a G protein-mediated stimulation of T-type calcium channels. Sci Signal. 2014;7:ra94.

    Article  Google Scholar 

  24. Zhang Y, Jiang D, Jiang X, Wang F, Tao J. Neuromedin U type 1 receptor stimulation of A-type K+ current requires the betagamma subunits of go protein, protein kinase a, and extracellular signal-regulated kinase 1/2 (ERK1/2) in sensory neurons. J Biol Chem. 2012;287:18562–72.

    Article  CAS  Google Scholar 

  25. Nelson MT, Woo J, Kang HW, Vitko I, Barrett PQ, Perez-Reyes E, Lee JH, Shin HS, Todorovic SM. Reducing agents sensitize C-type nociceptors by relieving high-affinity zinc inhibition of T-type calcium channels. J Neurosci. 2007;27:8250–60.

    Article  CAS  Google Scholar 

  26. Todorovic SM, Jevtovic-Todorovic V, Meyenburg A, Mennerick S, Perez-Reyes E, Romano C, Olney JW, Zorumski CF. Redox modulation of T-type calcium channels in rat peripheral nociceptors. Neuron. 2001;31:75–85.

    Article  CAS  Google Scholar 

  27. Moran TH, Robinson PH, Goldrich MS, McHugh PR. Two brain cholecystokinin receptors: implications for behavioral actions. Brain Res. 1986;362:175–9.

    Article  CAS  Google Scholar 

  28. Pavlasevic S, Bednar I, Qureshi GA, Sodersten P. Brain cholecystokinin tetrapeptide levels are increased in a rat model of anxiety. Neuroreport. 1993;5:225–8.

    Article  CAS  Google Scholar 

  29. Qureshi GA, Bednar I, Min Q, Sodersten P, Silberring J, Nyberg F, Thornwall M. Quantitation and identification of two cholecystokinin peptides, CCK-4 and CCK-8s, in rat brain by HPLC and fast atom bombardment mass spectrometry. Biomed Chromatogr. 1993;7:251–5.

    Article  CAS  Google Scholar 

  30. Zhang Y, Jiang D, Zhang J, Wang F, Jiang X, Tao J. Activation of neuromedin U type 1 receptor inhibits L-type Ca2+ channel currents via phosphatidylinositol 3-kinase-dependent protein kinase C epsilon pathway in mouse hippocampal neurons. Cell Signal. 2010;22:1660–8.

    Article  CAS  Google Scholar 

  31. Hu HJ, Glauner KS. Gereau RWt: ERK integrates PKA and PKC signaling in superficial dorsal horn neurons. I. Modulation of A-type K+ currents. J Neurophysiol. 2003;90:1671–9.

    Article  CAS  Google Scholar 

  32. Brock C, Schaefer M, Reusch HP, Czupalla C, Michalke M, Spicher K, Schultz G, Nurnberg B. Roles of G beta gamma in membrane recruitment and activation of p110 gamma/p101 phosphoinositide 3-kinase gamma. J Cell Biol. 2003;160:89–99.

    Article  CAS  Google Scholar 

  33. Naguib A. Following the trail of lipids: signals initiated by PI3K function at multiple cellular membranes. Sci Signal. 2016;9:re4.

    Article  Google Scholar 

  34. Ji RR, RWt G, Malcangio M, Strichartz GR. MAP kinase and pain. Brain Res Rev. 2009;60:135–48.

    Article  CAS  Google Scholar 

  35. Subramaniam S, Shahani N, Strelau J, Laliberte C, Brandt R, Kaplan D, Unsicker K. Insulin-like growth factor 1 inhibits extracellular signal-regulated kinase to promote neuronal survival via the phosphatidylinositol 3-kinase/protein kinase a/c-Raf pathway. J Neurosci. 2005;25:2838–52.

    Article  CAS  Google Scholar 

  36. Dhanasekaran DN, Reddy EP. JNK signaling in apoptosis. Oncogene. 2008;27:6245–51.

    Article  CAS  Google Scholar 

  37. Bast A, Fischer K, Erttmann SF, Walther R. Induction of peroxiredoxin I gene expression by LPS involves the Src/PI3K/JNK signalling pathway. Biochim Biophys Acta. 2010;1799:402–10.

    Article  CAS  Google Scholar 

  38. Vydyanathan A, Wu ZZ, Chen SR, Pan HL. A-type voltage-gated K+ currents influence firing properties of isolectin B4-positive but not isolectin B4-negative primary sensory neurons. J Neurophysiol. 2005;93:3401–9.

    Article  CAS  Google Scholar 

  39. Choi S, Parajuli SP, Lim GH, Kim JH, Yeum CH, Yoon PJ, Jun JY. Imipramine inhibits A-type delayed rectifier and ATP-sensitive K+ currents independent of G-protein and protein kinase C in murine proximal colonic myocytes. Arch Pharm Res. 2006;29:998–1005.

    Article  CAS  Google Scholar 

  40. Xu Y, Chiamvimonvat N, Vazquez AE, Akunuru S, Ratner N, Yamoah EN. Gene-targeted deletion of neurofibromin enhances the expression of a transient outward K+ current in Schwann cells: a protein kinase A-mediated mechanism. J Neurosci. 2002;22:9194–202.

    Article  CAS  Google Scholar 

  41. Wu RL, Butler DM, Barish ME. Potassium current development and its linkage to membrane expansion during growth of cultured embryonic mouse hippocampal neurons: sensitivity to inhibitors of phosphatidylinositol 3-kinase and other protein kinases. J Neurosci. 1998;18:6261–78.

    Article  CAS  Google Scholar 

  42. Lang F, Shumilina E. Regulation of ion channels by the serum- and glucocorticoid-inducible kinase SGK1. FASEB J. 2013;27:3–12.

    Article  CAS  Google Scholar 

  43. Van Hoorick D, Raes A, Keysers W, Mayeur E, Snyders DJ. Differential modulation of Kv4 kinetics by KCHIP1 splice variants. Mol Cell Neurosci. 2003;24:357–66.

    Article  Google Scholar 

  44. Yao JJ, Gao XF, Chow CW, Zhan XQ, Hu CL, Mei YA. Neuritin activates insulin receptor pathway to up-regulate Kv4.2-mediated transient outward K+ current in rat cerebellar granule neurons. J Biol Chem. 2012;287:41534–45.

    Article  CAS  Google Scholar 

  45. Pieri M, Amadoro G, Carunchio I, Ciotti MT, Quaresima S, Florenzano F, Calissano P, Possenti R, Zona C, Severini C. SP protects cerebellar granule cells against beta-amyloid-induced apoptosis by down-regulation and reduced activity of Kv4 potassium channels. Neuropharmacology. 2010;58:268–76.

    Article  CAS  Google Scholar 

  46. Roepke TA, Malyala A, Bosch MA, Kelly MJ, Ronnekleiv OK. Estrogen regulation of genes important for K+ channel signaling in the arcuate nucleus. Endocrinology. 2007;148:4937–51.

    Article  CAS  Google Scholar 

  47. Ma W, Quirion R. The ERK/MAPK pathway, as a target for the treatment of neuropathic pain. Expert Opin Ther Targets. 2005;9:699–713.

    Article  CAS  Google Scholar 

  48. Sanna MD, Stark H, Lucarini L, Ghelardini C, Masini E, Galeotti N. Histamine H4 receptor activation alleviates neuropathic pain through differential regulation of ERK, JNK and P38 MAPK phosphorylation. Pain. 2015.

  49. Popiolek-Barczyk K, Makuch W, Rojewska E, Pilat D, Mika J. Inhibition of intracellular signaling pathways NF-kappaB and MEK1/2 attenuates neuropathic pain development and enhances morphine analgesia. Pharmacol Rep. 2014;66:845–51.

    Article  CAS  Google Scholar 

  50. Schrader LA, Birnbaum SG, Nadin BM, Ren Y, Bui D, Anderson AE, Sweatt JD. ERK/MAPK regulates the Kv4.2 potassium channel by direct phosphorylation of the pore-forming subunit. Am J Physiol Cell Physiol. 2006;290:C852–61.

    Article  CAS  Google Scholar 

  51. Rodgers EW, Krenz WD, Jiang X, Li L, Baro DJ. Dopaminergic tone regulates transient potassium current maximal conductance through a translational mechanism requiring D1Rs, cAMP/PKA, Erk and mTOR. BMC Neurosci. 2013;14:143.

    Article  Google Scholar 

  52. Tang K, Li X, Zheng MQ, Rozanski GJ. Role of apoptosis signal-regulating kinase-1-c-Jun NH2-terminal kinase-p38 signaling in voltage-gated K+ channel remodeling of the failing heart: regulation by thioredoxin. Antioxid Redox Signal. 2011;14:25–35.

    Article  CAS  Google Scholar 

  53. Xie Y, Mai JT, Wang F, Lin YQ, Yuan WL, Luo NS, Fang MC, Wang JF, Chen YX. Effects of C-reactive protein on K(+) channel interaction protein 2 in cardiomyocytes. Am J Transl Res. 2015;7:922–31.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Takeda M, Tsuboi Y, Kitagawa J, Nakagawa K, Iwata K, Matsumoto S. Potassium channels as a potential therapeutic target for trigeminal neuropathic and inflammatory pain. Mol Pain. 2011;7:5.

    Article  Google Scholar 

  55. Chien LY, Cheng JK, Chu D, Cheng CF, Tsaur ML. Reduced expression of A-type potassium channels in primary sensory neurons induces mechanical hypersensitivity. J Neurosci. 2007;27:9855–65.

    Article  CAS  Google Scholar 

  56. Tolhurst DJ, Smyth D, Thompson ID. The sparseness of neuronal responses in ferret primary visual cortex. J Neurosci. 2009;29:2355–70.

    Article  CAS  Google Scholar 

  57. Li J, Zhao PP, Hao T, Wang D, Wang Y, Zhu YZ, Wu YQ, Zhou CH. Urotensin II inhibitor eases neuropathic pain by suppressing the JNK/NF-kappaB pathway. J Endocrinol. 2017;232:165–74.

    Article  CAS  Google Scholar 

  58. Zhuang ZY, Wen YR, Zhang DR, Borsello T, Bonny C, Strichartz GR, Decosterd I, Ji RR. A peptide c-Jun N-terminal kinase (JNK) inhibitor blocks mechanical allodynia after spinal nerve ligation: respective roles of JNK activation in primary sensory neurons and spinal astrocytes for neuropathic pain development and maintenance. J Neurosci. 2006;26:3551–60.

    Article  CAS  Google Scholar 

  59. Obata K, Yamanaka H, Kobayashi K, Dai Y, Mizushima T, Katsura H, Fukuoka T, Tokunaga A, Noguchi K. Role of mitogen-activated protein kinase activation in injured and intact primary afferent neurons for mechanical and heat hypersensitivity after spinal nerve ligation. J Neurosci. 2004;24:10211–22.

    Article  CAS  Google Scholar 

Download references

Acknowledgments

We would like to give our thanks to Mrs. Shan Gong and Dr. Yue Wang from the Department of Anatomy for their technical assistance. We also appreciate Dr. Xianmin Yu for his invaluable comments.

Funding

This study was supported by the National Natural Science Foundation of China (No. 81873731, No. 81771187, No. 81622014, No. 81671080, No. 81571063, No. 31800879), Qing-Lan Project of Jiangsu Province (J. Tao), the Six Talent Peak Project of Jiangsu Province (No. JY-065), Jiangsu Key Laboratory of Neuropsychiatric Diseases (BM2013003), and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Author information

Authors and Affiliations

Authors

Contributions

The study was designed by SY, XJ and JT. Data were collected and analysed by SY, XZ, ZC, YW, YS, DJ and YZ. The manuscript was drafted by SY, YZ and JT. All authors revised the article critically for intellectual content. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Jin Tao.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

All authors read and are consent for the publication of the manuscript.

Competing interests

The authors declare that they have no competing interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yu, S., Zhang, Y., Zhao, X. et al. Cholecystokinin type B receptor-mediated inhibition of A-type K+ channels enhances sensory neuronal excitability through the phosphatidylinositol 3-kinase and c-Src-dependent JNK pathway. Cell Commun Signal 17, 68 (2019). https://doi.org/10.1186/s12964-019-0385-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12964-019-0385-8

Keywords