Skip to main content
  • Research note
  • Open access
  • Published:

Increase of CaV3 channel activity induced by HVA β1b-subunit is not mediated by a physical interaction

Abstract

Objective

Low voltage-activated (LVA) calcium channels are crucial for regulating oscillatory behavior in several types of neurons and other excitable cells. LVA channels dysfunction has been implicated in epilepsy, neuropathic pain, cancer, among other diseases. Unlike for High Voltage-Activated (HVA) channels, voltage-dependence and kinetics of currents carried by recombinant LVA, i.e., CaV3 channels, are quite similar to those observed in native currents. Therefore, whether these channels are regulated by HVA auxiliary subunits, remain controversial. Here, we used the α1-subunits of CaV3.1, CaV3.2, and CaV3.3 channels, together with HVA auxiliary β-subunits to perform electrophysiological, confocal microscopy and immunoprecipitation experiments, in order to further explore this possibility.

Results

Functional expression of CaV3 channels is up-regulated by all four β-subunits, although most consistent effects were observed with the β1b-subunit. The biophysical properties of CaV3 channels were not modified by any β-subunit. Furthermore, although β1b-subunits increased colocalization of GFP-tagged CaV3 channels and the plasma membrane of HEK-293 cells, western blots analysis revealed the absence of physical interaction between CaV3.3 and β1b-subunits as no co-immunoprecipitation was observed. These results provide solid evidence that the up-regulation of LVA channels in the presence of HVA-β1b subunit is not mediated by a high affinity interaction between both proteins.

Introduction

Voltage-gated calcium (CaV) channels play a crucial role in cell Ca2+ influx, which in turn influences several cell functions as cellular excitability, muscle contraction, hormone and neurotransmitter secretion, and gene expression [1]. The CaV channels family is classified in low- and high-voltage activated (LVA and HVA) channels based on their activation threshold [2]. The conduction pore of these channels is formed by the α1-subunit, a four homologous domains (I–IV) single protein [3]. Auxiliary subunits, named β, α2δ and γ, modulate the activity of HVA channels [4, 5]. In particular, β-subunits, modulate HVA channels by increasing their surface expression [6,7,8,9], and modifying the voltage-dependence and current kinetics [reviewed by Refs. 4, 6]. α1- and β-subunits interaction takes place through the AID (alpha interaction domain) motif, localized at the intracellular link between domain I and II of α-1-subunits, and the alpha-binding pocket (ABP) site of β-subunits; this is a high-affinity interaction ranging from 2 to 54 nM [10,11,12]. In addition, low-affinity interaction sites at the amino and carboxy termini of HVA channels have also been implicated [13,14,15,16]. In contrast, it has been suggested that LVA channels (also known as T-type or CaV3 channels), are not modulated by HVA auxiliary subunits [17,18,19]. LVA channels are responsible for the low-threshold Ca2+ spikes in several central nervous system neurons [20], and they have also been implicated in pathophysiological conditions such as epilepsy, neuropathic pain, neuropsychiatric disorders and cancer [21,22,23,24]. Native LVA Ca2+ currents show an electrophysiological behavior quite similar to that of recombinant channels expressed without auxiliary subunits [25, 26], and their subunit composition is unknown due to the in silico strategy they were cloned with [25, 27,28,29]. Nevertheless, some studies support the notion that auxiliary subunits [30,31,32,33] might regulate CaV3 channels. More recently, a low-affinity association between synthetic peptides of CaV3.3 I-II loop and HVA β-subunits has been suggested [34]. Here, we addressed whether full-length CaV3 channel proteins and HVA β-subunits interact physically; our results provide experimental evidence that β-subunits up-regulate current density and the number of CaV3 channels in the plasma membrane by a mechanism that does not involve strong physical interactions between them, but rather they might have a low-affinity interaction.

Main text

Methods

Cell culture and transfection

HEK-293 cells were grown in DMEM/F12 mixture supplemented with 10% FBS, 100 U/ml penicillin, and 100 μg/ml streptomycin at 37 °C with 5% CO2. Transient transfections were performed with JetPei (polyplus-transfection) in 35-mm dishes, according to manufacturer’s protocol. Transfections were done with 1.5 µg of main α-subunit coding for CaV1.2 (GenBank accession #AY728090), CaV3.1 (#AF190860), CaV3.2 (#AF051946), CaV3.3 (#AF393329), or NaV1.6 (#NM_019266); 1.5 µg of the auxiliary β-subunit coding for CaVβ1a (M25817), CaVβ1b (X61394), CaVβ2a (M80545), CaVβ3 (X64300) and CaVβ4 (L02315); and 0.2 µg of GFP. Cells were dissociated 24-72 h after transfection and plated on coverslips for electrophysiological experiments.

Electrophysiology

Whole-cell Ca2+ (and Na+) currents were recorded at room temperature (21–23 °C) with the patch-clamp technique following the methods of Sanchez-Sandoval et al. [35]. External solutions composition was as follows (in mM): for LVA channels, 5 CaCl2 and 175 TEA-Cl; for HVA channels, 10 BaCl2 and 152 TEA-Cl; and for sodium channels, 158 NaCl, 2 CaCl2, and 2 MgCl2. Borosilicate glass pipettes (WPI Inc.) with resistances of 2–3 MΩ were filled with an internal solution containing (in mM): 130 CsCl, 10 EGTA, 2 CaCl2, 1 MgCl2, 4 Mg-ATP, and 0.3 Tris-GTP, for CaV channels; or with 106 CsCl, 30 NaCl, 1 CaCl2, 1 MgCl2 and 10 EGTA, for NaV channels. All solutions contained also 10 HEPES and were adjusted to pH 7.3 with TEA-OH, NaOH or CsOH, accordingly. Current recordings were analyzed using Clampfit software (Molecular Devices). Quantitative results are given as the mean ± standard error (SEM).

Construction of α1 and β-subunits of CaV channels tagged with fluorescent proteins

Fusion proteins were constructed by introducing restriction enzyme sites through the PCR mutagenesis technique. To generate CaV3.1-GFP and CaV3.2-GFP channels, the respective α1-subunits were cloned at the 3′ end of GFP in the multiple cloning site of pEGFP-C1 vector. The CaV3.3 channel with the GFP fused in the N-terminal and tagged with the hemagglutinin (HA) epitope in the extracellular S1–S2 of domain I [36], was modified to delete the HA epitope but conserving the GFP (CaV3.3-GFP). The CaV1.2 channel and the β1b-subunit were N-terminal tagged with the GFP and the BFP, by using the nucleotide sequence of the pRSET/GFP and pRSET/BFP vectors, respectively (Invitrogen). All constructions were verified by automated sequencing.

Confocal microscopy

For subcellular localization of GFP-tagged α1-subunits, HEK-293 cells were transfected with the corresponding α1-subunits alone or with the BFP-β1b subunit in 12 well-plates using 1.25 µg of each DNA subunit and PEI (Santa Cruz Biotechnology) as transfection reagent. Forty-eight hours after transfection the cells were cultured in 25-mm diameter coverslips for 12 h. Plasma membrane of HEK-293 cells was stained with FM4-64 (Invitrogen). Images were collected with an Olympus Fv10i confocal microscope equipped with a UPLSAPO 60×/1.35 oil immersion objective and using the following filters: for GFP-tagged channels the exciting wavelength was 489 nm; for β1b-BFP was 405 nm, and for FM4-64 was 559 nm. The confocal acquisition window was set to 512 × 512 pixels which allowed to acquire one image every 9 s for each fluorophore. Pearson colocalization coefficients where calculated with Imaris 8.2 software (Bitplane).

FRET measurements by sensitized emission

The Förster resonance energy transfer (FRET) measurements between the α1-subunits from CaV channels and the β1b subunit were obtained using the sensitized emission (SE) protocol. Briefly, FRET was obtained by measuring the acceptor emission resulting from donor excitation. To avoid overestimation of FRET first we evaluated the bleed-trough between both fluorescence channels for donor and acceptor. After subtracting the bleed-through from the donor emission we calculate FRET by using the following equation:

$$ nF = F^{{ex_{D} , em_{A} }} - \alpha F^{{ex_{A} ,em_{A} }} - \beta F^{{ex_{D} , em_{D} }} $$

Using an excitation wavelength that excites only the donor, the emission for the acceptor (\( F^{{ex_{D} , em_{A} }} \)) and donor (\( F^{{ex_{D} , em_{D} }} \)) channels are obtained. Next, fluorescence is measured in the acceptor channel (\( F^{{ex_{A} , em_{A} }} \)) at an excitation wavelength that only excites the acceptor. The amount of donor bleed-through into the acceptor channel is determined by a donor-only measurement, which provides the calibration constant \( \beta = {{F_{D}^{{ex_{D} , em_{A} }} } \mathord{\left/ {\vphantom {{F_{D}^{{ex_{D} , em_{A} }} } {F_{D}^{{ex_{D} , em_{D} }} }}} \right. \kern-0pt} {F_{D}^{{ex_{D} , em_{D} }} }} \) By measuring only the acceptor channel we obtained the constant \( \alpha = {{F_{A}^{{ex_{D} , em_{A} }} } \mathord{\left/ {\vphantom {{F_{A}^{{ex_{D} , em_{A} }} } {F_{A}^{{ex_{A} , em_{A} }} }}} \right. \kern-0pt} {F_{A}^{{ex_{A} , em_{A} }} }} \), for complete calibration procedures refer to [37]. FRET analysis was performed by using ImageJ with the Fret Analyzer plugin.

Co-immunoprecipitation and western blot

Total protein from transfected HEK-293 cells was extracted 48 h post-transfection using RIPA buffer supplemented with complete protease inhibitor (Roche). Co-immunoprecipitation (Co-IP) was performed by mixing total protein extracts from β1b-HA transfected cells with those of CaV1.2-GFP or CaV3.3-GFP, and incubated overnight at 4 °C in the presence of Anti-HA Affinity Matrix (Roche). Briefly, 50 μl of this matrix were washed and mixed with 1 mg of total protein from CaV1.2-GFP or CaV3.3-GFP transfected cells, and 1 mg of total protein from β1b-HA transfected cells. Then, beads were rinsed three times with RIPA buffer and proteins were eluted with Laemmli sample buffer, boiled at 95 °C for 3 min and analyzed by western blotting. The specific antibodies were used as follows: rat anti-HA (1:5000; Roche) for the β1b-HA subunit; and rabbit anti-GFP (1:5000; Santa Cruz Biotechnology) to identify CaV1.2-GFP and CaV3.3-GFP channels. Secondary antibodies were both raised in goat against rat and rabbit IgG-HRP (1:10,000; Santa Cruz Biotechnology). For control experiments, 15 µg of total protein from cell lysates of transfected cells were used in the immunoblots, as well as 10 µl (from a total of 500 µl) of the Co-IP supernatants. For loading control, a homemade monoclonal antibody against human β-actin was used (donated by Dr. Manuel Hernandez, CINVESTAV, Mexico).

Results

CaV3 channels current density increases in the presence of β1b

CaV3.1, CaV3.2 and CaV3.3 channels were co-expressed with each of the five β-subunits (β1a, β1b, β2a, β3 or β4) in HEK-293 cells, and whole-cell Ca2+ currents were analyzed by patch-clamp recordings. Figure 1a–c shows representative Ca2+ currents recorded at − 30 mV from each of the CaV3 channels in the absence (blue traces) and presence of the β1b-subunit (red traces). In all cases, current amplitudes were larger when β1b was co-transfected with the α1-subunits. The rise in current amplitudes was observed in the whole range of potentials that induced inward currents, without significant changes in the voltage-dependence of activation (Fig. 1d–f). Although all β-subunits increased current density of at least one of the CaV3 channels, only β1b-subunit was able to induce significant increments in CaV3.1 and CaV3.3 channels; however, the amplitude of CaV3.2 currents were not statistically different from the control (Fig. 1g–i). On average, β1b-subunit promoted increments of 63 ± 18, 57 ± 28, and 81 ± 16% in current density of HEK-293-cells transfected with CaV3.1, CaV3.2 and CaV3.3, respectively. Except for a discrete, but significant shift (4 mV) to more negative potentials in the V50 of voltage-dependence of activation of CaV3.3 currents, there were no additional changes in activation or inactivation channel gating, neither in current kinetics of activation, inactivation or recovery of inactivation due to the presence of the β1b-subunit (see Additional file 1). Like a positive control, we co-transfected the CaV1.2 channel and the β1b-subunit; as expected, the β1b-subunit induced a drastic (4-fold) increase in the current density; whereas no significant effect was observed when co-transfected with a voltage-gated sodium channel (negative control, see Additional file 2). Thus, the co-transfection of β1b-subunit with CaV3 channels induces specific and significant increases in current density without affecting the biophysical properties. Previous studies have shown similar effects [30, 31], although our electrophysiological data for CaV3.3 are the first to clearly show the effect of β-subunits.

Fig. 1
figure 1

The β1b-subunit increases the current density in all three LVA channels. ac Representative whole-cell currents recorded at -30 mV from HEK-293 cells transfected with each of the CaV3 channels (CaV3.1, CaV3.2 or CaV3.3) alone or together with the β1b-subunit. Patch-clamp experiments were performed by using a HP of − 100 mV and 5 mM CaCl2 as charge carrier. df Current–voltage (IV) relationships for the same channels as in ac. For IV plots data were obtained from currents evocated from − 80 mV to + 80 mV in 10 mV steps; current amplitudes were normalized by cell capacitance to obtain current density values. Smooth lines are fits to data with a modified Boltzmann function (see Experimental Procedures). The corresponding parameters are shown in Additional file 1. gi Current density (mean ± SEM) at − 30 mV calculated for HEK-293 cells transfected with the indicated CaV3 channels alone or together with β1a, β1b, β2a, β3 or β4 subunits. Only β1b increases current density significantly when transfected with any of the CaV3 channels. Data were normalized to the current density values obtained when CaV3 channels were transfected alone. *Statistical significance when using ANOVA followed by Dunnett’s multiple comparison test (P < 0.05). The number of studied cells varied from 5 to 29

CaV3 channels cell surface localization is increased when co-expressed with β1b-subunit

To determine whether the observed increments in current density of CaV3 channels co-expressed with β1b-subunits were the result of increases channel presence at the plasma membrane of HEK-293 cells, we investigated first the plasma membrane localization of all CaV3 channels by confocal microscopy in the presence or absence of β1b-subunit. For this purpose, we used GFP-tagged CaV1.2 and CaV3 channels, and BFP-tagged β1b-subunit (β1b-BFP). As a reference, plasma membrane was stained with the fluorescent marker FM4-64. Transient transfection of HEK-293 cells with CaV1.2 or CaV3 channels showed a fluorescence signal mainly restricted to the plasma membrane and some intracellular membranes (Fig. 2a; -β1b rows). Interestingly, when α1-subunits of CaV1.2 and CaV3 channels were co-transfected with β1b-BFP the colocalization of all CaV channels with the plasma membrane marker increased drastically (Fig. 2a, merge and colocalization columns). On average, the Pearson’s coefficient of colocalization between CaV channels and plasma membrane increased more than sixfold in the presence of β1b-subunit. These results suggest that β1b promotes the trafficking of CaV1.2 (HVA) and CaV3 (LVA) α1-subunits to the plasma membrane, which is consistent with our electrophysiological results shown in Fig. 1. Then, to determine if the increased trafficking of Cav3 to the plasma membrane was the result of a direct physical interaction with the β1b-subunit, we conducted FRET measurements between the α1-subunits from all CaV channels and the β1b-subunit by using the sensitized emission (SE) protocol [37]. As previously reported [38, 39], β1b showed a wide distribution throughout the cytoplasm and the nucleus (Fig. 2b, channel/β1b panels). When CaV channels were co-expressed with the β1b-subunit, a clear FRET signal was observed mainly in the plasma membrane of co-transfected cells with CaV1.2 channel (Fig. 2b, CaV1.2 FRET index panels). Most interestingly, no significant FRET was observed for Cav3 channels. Among LVA channels only the CaV3.3 showed a weak FRET signal (Fig. 2b, CaV3.3 FRET index panels). Thus, FRET analysis suggests that CaV1.2 α1-subunit and β1b-subunit are within the range distance of 10–100 nm, but only CaV3.3 channels and β1b- seem to be within the same distance.

Fig. 2
figure 2

The β1b-subunit increases the amount of CaV channels at the plasma membrane. a Co-localization analysis of HEK-293 cells expressing the α1-subunit of CaV1.2, CaV3.1, CaV3.2 and CaV3.3 fused to GFP, in the absence (-β1b) and presence (+β1b) of the β1b-subunit. Representative confocal microscopy images of cells expressing the respective CaV shown in green (left panels) and the plasma membrane marker FM4-64 (shown in red, middle panel). The co-localization between the CaV α1-subunit and the FM4-64 (in yellow) is shown in Merge panels. The plots to the right show the total pixels of colocalization between the green channel (CaV) and the red channel (FM-464), with the corresponding Pearson’s correlation coefficient (R) for each experimental condition. b The FRET between the GFP in each CaV channel and the blue fluorescent protein in the β1b-subunit. Notice that FRET is observed only between CaV1.2 (HVA channel) and β1b, but not with the other 3 CaV channels (CaV3.1, CaV3.2 and CaV3.3). High and low FRET was calculated pixel-by-pixel and the image shows in pseudo color FRET intensities. The number of cells analyzed for both colocalization and FRET panels were as follows: CaV1.2, 65; CaV3.1, 35; CaV3.2, 48; and CaV3.3, 32. The cells were obtained from 4 to 13 independent experiments. Scale bar: 10 μm

The CaV3.3 channel does not interact physically with the β1b-subunit

The potential physical interaction between CaV3.3 channels and β1b-subunits was further investigated with co-immunoprecipitation (Co-IPs) and western blot assays. An expected band of about 75-kDa was clearly detected in total protein extracts from HEK-293 cells transfected with β1b-HA, and from immunoprecipitations of these extracts with an anti-HA affinity matrix (Fig. 3, middle panel), but it was totally absent in extracts of transfected cells with either CaV1.2 or CaV3 channels (Fig. 3, lanes 5 and 6, middle panel), demonstrating the specificity of the HA antibody. Additionally, a ~ 250 kDa band was revealed with the anti-GFP antibody in extracts of HEK-293 cells transfected with the CaV1.2-GFP or CaV3.3-GFP channels (Fig. 3, lanes 5 and 6, upper panel). Furthermore, the CaV1.2-GFP/β1b-HA protein complex was co-immunoprecipitated with the anti-HA affinity matrix and the CaV1.2-GFP was detected as a 250-kDa band using anti-GFP (Fig. 3, lane 1, upper panel). On the contrary, the same procedure did not show any co-immunoprecipitation when CaV3.3-GFP channels were used instead of the HVA channel (Fig. 3, lane 2, upper panel). The supernatants obtained from Co-IPs samples loaded in lanes 1 and 2 showed considerable amounts of CaV1.2 (lane 7) and CaV3.3 (lane 8) channels, indicating that the lack of GFP-immunoreactivity signal in lane 2 was due to the absence of a strong physical interaction between the CaV3.3 channels and the β1b-subunit, rather than a shortage of the protein in the sample. Interestingly, when the Co-IPs were processed less exhaustively, by washing the beads only once instead of three times as done for those in lanes 1 and 2 of Fig. 3 (upper panel), the immunodetection with the GFP antibody revealed a band corresponding to the CaV3.3 channels, and an at least 3-times stronger signal for the β1b-HA-subunit in the same IPs (Fig. 3, lane 3), as well as the presence of actin (Fig. 3, lane 3, lower panel), indicating the importance of correct washing procedures. Altogether, these results suggest that the β1b-subunit does not interact with CaV3.3 channels as strongly (high-affinity) as with HVA CaV1.2 channels, on the contrary, such interaction, if any, is rather weak (low-affinity).

Fig. 3
figure 3

CaV3.3 channels and β1b-subunit do not coimmunoprecipitate. Western blot of IP of CaV1.2-GFP with β1b-HA (positive control, lane 1) and CaV3.3-GFP with β1b-HA (lanes 2 and 3). IPs in lanes 1 and 2 were washed three times with lysis buffer after overnight incubation with the indicated proteins, whereas the IP from lane 3 was washed only once. The immunoblot was probed with an antibody against GFP (upper panel), then striped and probed with an anti-HA antibody (middle panel), then striped again and probed with a homemade anti-actin antibody (lower panel). Lanes 4-6 were loaded with 15 µg of total protein from lysates of HEK-293 cells transfected with β1b-HA, CaV1.2-GFP and CaV3.3-GFP, respectively. Finally, lanes 7 and 8 were loaded with 10 µl of the IP supernatant from lanes 1 and 2, accordingly. Notice the importance of exhaustive washing procedures, since incomplete washing of the beads could lead to false positives results. As can be observed, CaV1.2 channels coimmunoprecipitate with β1b-HA (lane 1, upper panel), whereas CaV3.3 channels do not (lane 2, upper panel). Representative figures of three independent experiments

Discussion

LVA calcium channels display unique functional properties that support critical cell functions in a variety of tissues, and their dysfunction is associated with pathological consequences, such as epilepsy and spinocerebellar ataxia [40,41,42]. In addition, LVA channel activity is regulated by different cellular mechanisms involving the action of neurotransmitters and hormones [43,44,45]; and during cellular process like differentiation and proliferation [21, 46]; however, regulation by HVA channels accessory subunits is still a field of controversy. Several reports have shown that LVA calcium channels are not regulated by these auxiliary subunits [17,18,19, 47]. One of these reports showed that depletion of β-subunits in nodosus ganglion neurons had no significant changes in LVA calcium currents [47]. Nevertheless, these cells do not express the β1b-subunit, which according to our data, induces the most important changes in LVA channels activity. Thus, the results reported by [47] could be due to the absence of β1b-subunit expression in such neurons. In contrast, recent evidence suggest that LVA calcium channels are modulated by HVA accessory subunits [30, 31, 34]. By using in vitro immunoassays, Bae and coworkers [34] suggested a low-affinity interaction between β-subunits and synthetic peptides of CaV3.3 channels containing the equivalent AID sequence (30-residues peptides). Here, we show that co-transfection of CaV3 channels with different β-subunits lead to an increase in current density, an effect that was more consistent and robust with CaV3.3 channels, and to a lesser extent in CaV3.1 and CaV3.2. These effects were limited to current density as biophysical properties of channels were not affected, as previously reported by others [19, 31].

By using full-length α1-subunits of LVA and CaV1.2 channels, we observed a robust increase in colocalization of these α1-subunits with the plasma membrane in the presence of the β1b-subunit. In addition, FRET studies between CaV1.2 channels and β1b-subunit confirm a physical interaction between both proteins. However, FRET was practically absent for CaV3 channels and β1b-subunit. These observations were further confirmed by co-immunoprecipitation experiments where only CaV1.2 channel protein co-immunoprecipitated with the β1b-subunit, suggesting that LVA CaV3.3 channel is not close enough to the β1b-subunit to have a strong physical interaction as the one display by the HVA CaV1.2 channel (Fig. 3).

Thus, the electrophysiological regulation of CaV3 channels by β-subunits shown in Fig. 1 could not be explained by a strong physical interaction between these proteins, but by the increment of cell surface localization of these channels. The lack of a strong interaction between CaV3.3 and the β1b-subunit could be explained by the absence of the AID motif, which has been widely proven to mediate the physical interaction between HVA channels and β-subunits [48,49,50]. However, non-AID motif interactions between the CaV3 α1-subunits and β-subunits cannot be ruled out. In fact, this possibility is supported by the observation that synthetic CaV2.1-AID peptide did not alter the binding of the CaV3.3-AID peptide to β or β4, suggesting that the β-subunit ABP do not play a role in binding to the CaV3.3-AID [34]. Additional evidence leading to the possibility of multiple interaction sites between HVA α1-subunits and β-subunits include structural studies as well [51, 52]. Because we used the whole proteins for our co-immunoprecipitation experiments, identifying the precise amino acids involved in the interaction is not possible.

In summary, we have found that LVA calcium channels are regulated by the β1b-subunit by increasing membrane channel protein and current density in HEK-293 cells. Nevertheless, low or null FRET signal suggests a weak or null physical interaction between both proteins, which in turn could explain the increment in CaV3 channel membrane density. It is noteworthy that the weakness of this interaction might be the main reason for the discrete, and sometimes, totally absent regulation of the LVA channels expression and biophysical properties.

Limitations

Our results show the lack of physical interaction between full-length CaV3.3 channels and β1b-subunits, it remains to be explored this issue for the other HVA β-subunits.

Abbreviations

HEK-293 cells:

human embryonic kidney cells

DMEM:

Dulbecco’s Modified Eagle’s Medium

FBS:

fetal bovine serum

GFP:

green fluorescent protein

TEA:

tetra-ethyl-ammonium

PCR:

polymerase chain reaction

BFP:

blue fluorescent protein

FRET:

Förster resonance energy transfer

Co-IP:

co-immunoprecipitation

References

  1. Berridge MJ. Unlocking the secrets of cell signaling. Annu Rev Physiol. 2005;67:1–21.

    Article  CAS  Google Scholar 

  2. Hille B. Ion channels of excitable membranes. 3rd ed. Sunderland: Sinauer; 2001. p. 814.

    Google Scholar 

  3. Catterall WA. Voltage-gated calcium channels. Cold Spring Harb Persp Biol. 2011;3(8):a003947.

    Google Scholar 

  4. Simms BA, Zamponi GW. Neuronal voltage-gated calcium channels: structure, function, and dysfunction. Neuron. 2014;82(1):24–45.

    Article  CAS  Google Scholar 

  5. Arikkath J, Campbell KP. Auxiliary subunits: essential components of the voltage-gated calcium channel complex. Curr Opin Neurobiol. 2003;13(3):298–307.

    Article  CAS  Google Scholar 

  6. Jones LP, Wei SK, Yue DT. Mechanism of auxiliary subunit modulation of neuronal alpha(1E) calcium channels. J Gen Physiol. 1998;112(2):125–43.

    Article  CAS  Google Scholar 

  7. Josephson IR, Varadi G. The beta subunit increases Ca2+ currents and gating charge movements of human cardiac L-type Ca2+ channels. Biophys J. 1996;70(3):1285–93.

    Article  CAS  Google Scholar 

  8. Yasuda T, Chen L, Barr W, McRory JE, Lewis RJ, Adams DJ, et al. Auxiliary subunit regulation of high-voltage activated calcium channels expressed in mammalian cells. Eur J Neurosci. 2004;20(1):1–13.

    Article  Google Scholar 

  9. Kamp TJ, PerezGarcia MT, Marban E. Enhancement of ionic current and charge movement by coexpression of calcium channel beta(1A) subunit with alpha(1C) subunit in a human embryonic kidney cell line. J Physiol London. 1996;492(1):89–96.

    Article  CAS  Google Scholar 

  10. DeWaard M, Scott VES, Pragnell M, Campbell KP. Identification of critical amino acids involved in alpha(1)-beta interaction in voltage-dependent Ca2+ channels. FEBS Lett. 1996;380(3):272–6.

    Article  CAS  Google Scholar 

  11. Dewaard M, Witcher DR, Pragnell M, Liu HY, Campbell KP. Properties of the alpha(1)-beta anchoring site in voltage-dependent Ca2+ channels. J Biol Chem. 1995;270(20):12056–64.

    Article  CAS  Google Scholar 

  12. Canti C, Davies A, Berrow NS, Butcher AJ, Page KM, Dolphin AC. Evidence for two concentration-dependent processes for beta-subunit effects on alpha 1B calcium channels. Biophys J. 2001;81(3):1439–51.

    Article  CAS  Google Scholar 

  13. Cornet V, Bichet D, Sandoz G, Marty I, Brocard J, Bourinet E, et al. Multiple determinants in voltage-dependent P/Q calcium channels control their retention in the endoplasmic reticulum. Eur J Neurosci. 2002;16(5):883–95.

    Article  Google Scholar 

  14. Gao TY, Bunemann M, Gerhardstein BL, Ma H, Hosey MM. Role of the C terminus of the alpha(1C) (Ca(v)1.2) subunit in membrane targeting of cardiac L-type calcium channels. J Biol Chem. 2000;275(33):25436–44.

    Article  CAS  Google Scholar 

  15. Stephens GJ, Page KM, Bogdanov Y, Dolphin AC. The alpha 1B Ca(2+) channel amino terminus contributes determinants for beta subunit-mediated voltage-dependent inactivation properties. J Physiol London. 2000;525(2):377–90.

    Article  CAS  Google Scholar 

  16. Walker D, Bichet D, Campbell KP, De Waard M. A beta(4) isoform-specific interaction site in the carboxyl-terminal region of the voltage-dependent Ca2+ channel alpha(1A) subunit. J Biol Chem. 1998;273(4):2361–7.

    Article  CAS  Google Scholar 

  17. Lacinova L, Klugbauer N, Hofmann F. Absence of modulation of the expressed calcium channel alpha1G subunit by alpha2delta subunits. J Physiol. 1999;516(Pt 3):639–45.

    Article  CAS  Google Scholar 

  18. Leuranguer V, Bourinet E, Lory P, Nargeot J. Antisense depletion of beta-subunits fails to affect T-type calcium channels properties in a neuroblastoma cell line. Neuropharmacology. 1998;37(6):701–8.

    Article  CAS  Google Scholar 

  19. Arias JM, Murbartian J, Vitko I, Lee JH, Perez-Reyes E. Transfer of beta subunit regulation from high to low voltage-gated Ca2+ channels. FEBS Lett. 2005;579(18):3907–12.

    Article  CAS  Google Scholar 

  20. Cheong E, Shin HS. T-type Ca2+ channels in normal and abnormal brain functions. Physiol Rev. 2013;93(3):961–92.

    Article  CAS  Google Scholar 

  21. Weaver EM, Zamora FJ, Puplampu-Dove YA, Kiessu E, Hearne JL, Martin-Caraballo M. Regulation of T-type calcium channel expression by sodium butyrate in prostate cancer cells. Eur J Pharmacol. 2015;749:20–31.

    Article  CAS  Google Scholar 

  22. Nelson MT, Todorovic SM, Perez-Reyes E. The role of T-type calcium channels in epilepsy and pain. Curr Pharm Des. 2006;12(18):2189–97.

    Article  CAS  Google Scholar 

  23. Gangarossa G, Laffray S, Bourinet E, Valjent E. T-type calcium channel Ca(v)3.2 deficient mice show elevated anxiety, impaired memory and reduced sensitivity to psychostimulants. Front Behav Neurosci. 2014;8:92.

    Article  Google Scholar 

  24. Uslaner JM, Smith SM, Huszar SL, Pachmerhiwala R, Hinchliffe RM, Vardigan JD, et al. T-type calcium channel antagonism produces antipsychotic-like effects and reduces stimulant-induced glutamate release in the nucleus accumbens of rats. Neuropharmacology. 2012;62(3):1413–21.

    Article  CAS  Google Scholar 

  25. Perez-Reyes E, Cribbs LL, Daud A, Lacerda AE, Barclay J, Williamson MP, et al. Molecular characterization of a neuronal low-voltage-activated T-type calcium channel. Nature. 1998;391(6670):896–900.

    Article  CAS  Google Scholar 

  26. Klockner U, Lee JH, Cribbs LL, Daud A, Hescheler J, Pereverzev A, et al. Comparison of the Ca2+ currents induced by expression of three cloned alpha 1 subunits, alpha 1G, alpha 1H and alpha 1I, of low-voltage-activated T-type Ca2+ channels. Eur J Neurosci. 1999;11(12):4171–8.

    Article  CAS  Google Scholar 

  27. Cribbs LL, Lee JH, Yang J, Satin J, Zhang Y, Daud A, et al. Cloning and characterization of alpha 1H from human heart, a member of the T-type Ca(2+) channel gene family. Circ Res. 1998;83(1):103–9.

    Article  CAS  Google Scholar 

  28. Gomora JC, Murbartian J, Arias JM, Lee JH, Perez-Reyes E. Cloning and expression of the human T-type channel Ca(v) 3.3: insights into prepulse facilitation. Biophys J. 2002;83(1):229–41.

    Article  CAS  Google Scholar 

  29. Lee JH, Daud AN, Cribbs LL, Lacerda AE, Pereverzev A, Klockner U, et al. Cloning and expression of a novel member of the low voltage-activated T-type calcium channel family. J Neurosci. 1999;19(6):1912–21.

    Article  CAS  Google Scholar 

  30. Dolphin AC, Wyatt CN, Richards J, Beattie RE, Craig P, Lee JH, et al. The effect of alpha 2-delta and other accessory subunits on expression and properties cf the calcium channel alpha 1G. J Physiol London. 1999;519(1):35–45.

    Article  CAS  Google Scholar 

  31. Dubel SJ, Altier C, Chaumont S, Lory P, Bourinet E, Nargeot J. Plasma membrane expression of T-type calcium channel alpha(1) subunits is modulated by high voltage-activated auxiliary subunits. J Biol Chem. 2004;279(28):29263–9.

    Article  CAS  Google Scholar 

  32. Thompson WR, Majid AS, Czymmek KJ, Ruff AL, Garcia J, Duncan RL, et al. Association of the alpha(2)delta(1) subunit with Ca(v)3.2 enhances membrane expression and regulates mechanically induced ATP release in MLO-Y4 osteocytes. J Bone Mineral Res. 2011;26(9):2125–39.

    Article  CAS  Google Scholar 

  33. Hansen JP, Chen RS, Larsen JK, Chu PJ, Janes DM, Weis KE, et al. Calcium channel gamma 6 subunits are unique modulators of low voltage-activated (Cav3.1) calcium current. J Mol Cell Cardiol. 2004;37(6):1147–58.

    Article  CAS  Google Scholar 

  34. Bae J, Suh EJ, Lee C. Interaction of T-type calcium channel Ca(v)3.3 with the beta-subunit. Mol Cells. 2010;30(3):185–91.

    Article  CAS  Google Scholar 

  35. Sanchez-Sandoval AL, Carrillo ZH, Velasquez CED, Delgadillo DM, Rivera HM, Gomora JC. Contribution of S4 segments and S4–S5 linkers to the low-voltage activation properties of T-type Ca(v)3.3 channels. PLoS ONE. 2018;13(2):e0193490.

    Article  Google Scholar 

  36. Baumgart JP, Vitko I, Bidaud I, Kondratskyi A, Lory P, Perez-Reyes E. I–II loop structural determinants in the gating and surface expression of low voltage-activated calcium channels. PLoS One. 2008;3(8):e2976.

    Article  Google Scholar 

  37. Zeug A, Woehler A, Neher E, Ponimaskin EG. Quantitative intensity-based FRET approaches-A comparative snapshot. Biophys J. 2012;103(9):1821–7.

    Article  CAS  Google Scholar 

  38. Altier C, Garcia-Caballero A, Simms B, You HT, Chen LN, Walcher J, et al. The Cav beta subunit prevents RFP2-mediated ubiquitination and proteasomal degradation of L-type channels. Nat Neurosci. 2011;14(2):U173–252.

    Article  Google Scholar 

  39. Waithe D, Ferron L, Page KM, Chaggar K, Dolphin AC. Beta-subunits promote the expression of Ca(V)2.2 channels by reducing their proteasomal degradation. J Biol Chem. 2011;286(11):9598–611.

    Article  CAS  Google Scholar 

  40. Gomora JC, Daud AN, Weiergraber M, Perez-Reyes E. Block of cloned human, T-type calcium channels by succinimide antiepileptic drugs. Mol Pharmacol. 2001;60(5):1121–32.

    Article  CAS  Google Scholar 

  41. Shin HS. T-type Ca2+ channels and absence epilepsy. Cell Calcium. 2006;40(2):191–6.

    Article  CAS  Google Scholar 

  42. Morino H, Matsuda Y, Muguruma K, Miyamoto R, Ohsawa R, Ohtake T, et al. A mutation in the low voltage-gated calcium channel CACNA1G alters the physiological properties of the channel, causing spinocerebellar ataxia. Mol Brain. 2015;8:89.

    Article  Google Scholar 

  43. Chemin J, Traboulsie A, Lory P. Molecular pathways underlying the modulation of T-type calcium channels by neurotransmitters and hormones. Cell Calcium. 2006;40(2):121–34.

    Article  CAS  Google Scholar 

  44. Yu HJ, Seo JB, Jung SR, Koh DS, Hille B. Noradrenaline upregulates T-type calcium channels in rat pinealocytes. J Physiol-London. 2015;593(4):887–904.

    Article  CAS  Google Scholar 

  45. Zhang Y, Jiang XH, Snutch TP, Tao J. Modulation of low-voltage-activated T-type Ca2+ channels. Bba-Biomembranes. 2013;1828(7):1550–9.

    Article  CAS  Google Scholar 

  46. Lory P, Bidaud I, Chemin J. T-type calcium channels in differentiation and proliferation. Cell Calcium. 2006;40(2):135–46.

    Article  CAS  Google Scholar 

  47. Lambert RC, Maulet Y, Mouton J, Beattie R, Volsen S, De Waard M, et al. T-type Ca2+ current properties are not modified by Ca2+ channel beta subunit depletion in nodosus ganglion neurons. J Neurosci. 1997;17(17):6621–8.

    Article  CAS  Google Scholar 

  48. Van Petegem F, Clark KA, Chatelain FC, Minor DL. Structure of a complex between a voltage-gated calcium channel beta-subunit and an alpha-subunit domain. Nature. 2004;429(6992):671–5.

    Article  Google Scholar 

  49. Chen YH, Li MH, Zhang Y, He LL, Yamada Y, Fitzmaurice A, et al. Structural basis of the alpha(1)-beta subunit interaction of voltage-gated Ca2+ channels. Nature. 2004;429(6992):675–80.

    Article  CAS  Google Scholar 

  50. Berrou L, Klein H, Bernatchez G, Parent L. A specific tryptophan in the I–II linker is a key determinant of beta-subunit binding and modulation in Ca(V)2.3 calcium channels. Biophys J. 2002;83(3):1429–42.

    Article  CAS  Google Scholar 

  51. Maltez JM, Nunziato DA, Kim J, Pitt GS. Essential Ca(V)beta modulatory properties are AID-independent. Nat Struct Mol Biol. 2005;12(4):372–7.

    Article  CAS  Google Scholar 

  52. Dresviannikov AV, Page KM, Leroy J, Pratt WS, Dolphin AC. Determinants of the voltage dependence of G protein modulation within calcium channel beta subunits. Pflugers Arch. 2009;457(4):743–56.

    Article  CAS  Google Scholar 

Download references

Authors’ contributions

RAT conducted most of the experiments, analyzed the results, and wrote the original version of the manuscript. ALSS and BECR contributed with co-immunoprecipitation and western blot experiments. MJRP and LV performed and analyzed confocal experiments. JCG conceived the idea for the project, analyzed results and wrote the manuscript with RAT. All authors read and approved the final manuscript.

Acknowledgements

The CaV3.3-GFP-HA construct, as well as the three human clones of CaV3 channels were originally donated by Dr. Edward Perez-Reyes (University of Virginia); CaV1.2 clone was provided by Dr. Ricardo Felix (Cinvestav-Mexico); β1b plasmid was a gift from Dr. T. Snutch (University of British Columbia). The excellent technical assistance of Drs. Zazil Herrera-Carrillo, Clara E. Diaz-Velasquez, and Dulce M. Delgadillo-Alvarez is also gratefully acknowledged. We also thank Laura Ongay, Minerva Mora and Guadalupe Codiz from Unidad de Biología Molecular at Instituto de Fisiología Celular, UNAM, for technical support.

Competing interests

The authors declare that they have no competing interests.

Availability of data and materials

All data generated or analyzed during this study are included in this published article and its additional files.

Consent for publication

Not applicable.

Ethics approval and consent to participate

Not applicable.

Funding

This work was supported by grants from CONACYT-México (167790-B) and PAPIIT-DGAPA-UNAM (IN206917) to JCG. Rogelio Arteaga-Tlecuitl is a doctoral student from Programa de Doctorado en Ciencias Biomédicas, Universidad Nacional Autónoma de México (UNAM) and received fellowship 229977 from CONACYT. The funding bodies had no role in the design of the study and collection, analysis, and interpretation of data and in writing the manuscript.

Publisher’s Note

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

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Juan Carlos Gomora.

Additional files

Additional file 1.

Effects of β1b subunit in the biophysical properties of CaV3 channels. Table containing the biophysical properties of CaV3 channels in the absence and the presence of β1b subunit.

Additional file 2.

Modulation by the β1b subunit is specific on CaV channels. Electrophysiological recordings and I-V relationship for HVA CaV1.2 and NaV1.6 channels in the absence and the presence of the β1b subunit.

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

Arteaga-Tlecuitl, R., Sanchez-Sandoval, A.L., Ramirez-Cordero, B.E. et al. Increase of CaV3 channel activity induced by HVA β1b-subunit is not mediated by a physical interaction. BMC Res Notes 11, 810 (2018). https://doi.org/10.1186/s13104-018-3917-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13104-018-3917-1

Keywords