Skip to main content

Main menu

  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Collections
    • Podcast
  • ALERTS
  • FOR AUTHORS
    • Information for Authors
    • Fees
    • Journal Clubs
    • eLetters
    • Submit
    • Special Collections
  • EDITORIAL BOARD
    • Editorial Board
    • ECR Advisory Board
    • Journal Staff
  • ABOUT
    • Overview
    • Advertise
    • For the Media
    • Rights and Permissions
    • Privacy Policy
    • Feedback
    • Accessibility
  • SUBSCRIBE

User menu

  • Log out
  • Log in
  • My Cart

Search

  • Advanced search
Journal of Neuroscience
  • Log out
  • Log in
  • My Cart
Journal of Neuroscience

Advanced Search

Submit a Manuscript
  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Collections
    • Podcast
  • ALERTS
  • FOR AUTHORS
    • Information for Authors
    • Fees
    • Journal Clubs
    • eLetters
    • Submit
    • Special Collections
  • EDITORIAL BOARD
    • Editorial Board
    • ECR Advisory Board
    • Journal Staff
  • ABOUT
    • Overview
    • Advertise
    • For the Media
    • Rights and Permissions
    • Privacy Policy
    • Feedback
    • Accessibility
  • SUBSCRIBE
PreviousNext
Featured ArticleResearch Articles, Cellular/Molecular

Membrane Stretch Gates NMDA Receptors

Sophie Belin, Bruce A. Maki, James Catlin, Benjamin A. Rein and Gabriela K. Popescu
Journal of Neuroscience 20 July 2022, 42 (29) 5672-5680; https://doi.org/10.1523/JNEUROSCI.0350-22.2022
Sophie Belin
Department of Biochemistry, Jacobs School of Medicine and Biomedical Sciences, University at Buffalo, SUNY, Buffalo, New York 14214
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Bruce A. Maki
Department of Biochemistry, Jacobs School of Medicine and Biomedical Sciences, University at Buffalo, SUNY, Buffalo, New York 14214
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
James Catlin
Department of Biochemistry, Jacobs School of Medicine and Biomedical Sciences, University at Buffalo, SUNY, Buffalo, New York 14214
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Benjamin A. Rein
Department of Biochemistry, Jacobs School of Medicine and Biomedical Sciences, University at Buffalo, SUNY, Buffalo, New York 14214
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Gabriela K. Popescu
Department of Biochemistry, Jacobs School of Medicine and Biomedical Sciences, University at Buffalo, SUNY, Buffalo, New York 14214
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Gabriela K. Popescu
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

NMDARs are ionotropic glutamate receptors widely expressed in the CNS, where they mediate phenomena as diverse as neurotransmission, information processing, synaptogenesis, and cellular toxicity. They function as glutamate-gated Ca2+-permeable channels, which require glycine as coagonist, and can be modulated by many diffusible ligands and cellular cues, including mechanical stimuli. Previously, we found that, in cultured astrocytes, shear stress initiates NMDAR-mediated Ca2+ entry in the absence of added agonists, suggesting that more than being mechanosensitive, NMDARs may be mechanically activated. Here, we used controlled expression of rat recombinant receptors and noninvasive on-cell single-channel current recordings to show that mild membrane stretch can substitute for the neurotransmitter glutamate in gating NMDAR currents. Notably, stretch-activated currents maintained the hallmark features of the glutamate-gated currents, including glycine-requirement, large unitary conductance, high Ca2+ permeability, and voltage-dependent Mg2+ blockade. Further, we found that the stretch-gated current required the receptor's intracellular domain. Our results are consistent with the hypothesis that mechanical forces can gate endogenous NMDAR currents even in the absence of synaptic glutamate release, which has important implications for understanding mechanotransduction and the physiological and pathologic effects of mechanical forces on cells of the CNS.

SIGNIFICANCE STATEMENT We show that, in addition to enhancing currents elicited with low agonist concentrations, membrane stretch can gate NMDARs in the absence of the neurotransmitter glutamate. Stretch-gated currents have the principal hallmarks of the glutamate-gated currents, including requirement for glycine, large Na+ conductance, high Ca2+ permeability, and voltage-dependent Mg2+ block. Therefore, results suggest that mechanical forces can initiate cellular processes presently attributed to glutamatergic neurotransmission, such as synaptic plasticity and cytotoxicity. Given the ubiquitous presence of mechanical forces in the CNS, this discovery identifies NMDARs as possibly important mechanotransducers during development and across the lifespan, and during pathologic processes, such as those associated with traumatic brain injuries, shaken infant syndrome, and chronic traumatic encephalopathy.

  • ionotropic glutamate receptors
  • mechanotransduction
  • NMDARs
  • patch-clamp
  • signal transduction
  • single-molecule

Introduction

Cells of the CNS experience endogenous and environmental mechanical forces in vivo, and respond to osmotic and atmospheric pressure ex vivo (Tyler, 2012; Koser et al., 2016; Bliznyuk et al., 2020). Mechanical stimuli affect several neurophysiological processes, including neuronal firing, vesicle fusion, dendritic spine formation, and synaptic activity (Hill, 1950; Korkotian and Segal, 2001; Star et al., 2002; Kim et al., 2007; Ucar et al., 2021). However, the mechanism of mechanotransduction in the CNS remains poorly understood largely because of experimental, technological, and theoretical challenges unique to examining the effect of mechanical forces in biological tissues. Among these obstacles are the omnipresence of mechanical cues, their diverse 3D and dynamic actions, the variety of macromolecules that participate in mechanotransduction, and the multiplicity of mechanisms by which transducers sense and respond to mechanical stimuli (Cox et al., 2019; Le Roux et al., 2019; Kefauver et al., 2020).

On a millisecond timescale, mechanotransduction is mediated by mechanically activated and mechanically sensitive ion channels (Cox et al., 2019; Kefauver et al., 2020). Mechanically activated channels are membrane proteins dedicated to scanning the environment for mechanically encoded information; they represent the molecular basis for a wide array of mechanosensory processes, including hearing, touch, and proprioception; and are critical for normal development and adaptation throughout life (Walsh et al., 2015; Murthy et al., 2017). On the other hand, a large swath of ion channels whose primary physiological function is to respond to electrical and chemical signals, while not directly gated by mechanical stimuli, are mechanosensitive. These channels mediate much of the CNS mechanotransduction and are essential to how mechanical forces influence the normal development and functioning of the brain and spinal cord, and also how they initiate or aggravate acute and chronic neuropathologies (Tyler, 2012).

NMDARs are glutamate-gated channels with demonstrated mechanosensitivity (Johnson et al., 2019). NMDARs mediate excitatory transmission and plasticity in CNS and are critical for the normal physiology of excitatory synapses; moreover, their overactivation mediates glutamate excitotoxicity, which has been implicated as a causal factor in several neuropathologies. Ambient pressure, membrane stretch, and membrane lipid composition modulate their agonist-gated currents in native preparations, in heterologous systems, and in artificial lipid bilayers (Fagni et al., 1987; Miller et al., 1992; Nishikawa et al., 1994; Paoletti and Ascher, 1994; Casado and Ascher, 1998; Kloda et al., 2007). In addition to mechanosensitivity, we reported recently that shear stress, as applied by shear microfluidic flow onto cultured astrocytes, elicits NMDAR-mediated Ca2+ influx in the absence of glutamate, suggesting that mechanical stimuli per se can gate NMDAR currents (Maneshi et al., 2017). This observation has important implications for a potential role of NMDARs in mechanotransduction during the normal development and function of the CNS (Tyler, 2012; Goriely et al., 2015; Heuer and Toro, 2019); and also in severe neuropsychiatric pathologies, including those associated with acute traumatic brain and spinal cord injuries, chronic traumatic encephalopathy, shaken infant syndrome, and episodic edema or tumor growth (Bonnier et al., 2004; Shively et al., 2012; Sloley et al., 2021). Therefore, we undertook the work reported here to investigate our novel observation in more depth.

Given that shear force can elicit NMDAR-dependent Ca2+ fluxes in primary cultures of astrocytes in the absence of agonist (Maneshi et al., 2017), here we investigate more specifically the sensitivity of NMDAR currents to membrane stretch, using a recombinant system and cultured neurons, with single-channel current recordings. We found that, as with sheer stress in cultured astrocyte, gentle suction applied to a membrane patches elicited currents from recombinant NMDARs expressed in HEK cells in the absence of the neurotransmitter glutamate. Importantly, the stretch-gated current maintained the characteristic biophysical properties of the glutamate-gated current, including requirement for glycine, high unitary conductance, Ca2+ permeability, and voltage-dependent Mg2+ blockade. In addition, we found that the C-terminus of NMDARs is required to initiate stretch-induced currents.

Materials and Methods

Cells and receptor expression

HEK293 cells (American type Culture Collection number CRL-1573) were grown and maintained in DMEM supplemented with 10% FBS (Invitrogen) and 1% penicillin/streptomycin. Cells were grown to 80% confluency, and passages 24 and 31 were used for transfections. Cells were transfected transiently via the Ca2+-phosphate method using pcDNA3.1 (+) plasmids encoding rat GluN1-1a (P35439-1), rat GluN2A (Q00959), and GFP (P42212) in a 1:1:1 ratio. When indicated, plasmids encoding GluN1-1a and GluN2A were replaced by plasmids encoding CTD-truncated GluN1-a (GluN1-a 838stop) and CTD-truncated GluN2A (GluN2A 844stop), provided by Westbrook (Krupp et al., 1999, 2002). Alternatively, when indicated, the GluN2A-encoding plasmid was substituted with plasmids expressing rat GluN2B (Q00960), rat GluN2C (Q00961), or rat GluN2D (Q62645). Cells were incubated with the DNA mixture for 2 h, were washed twice with PBS, and incubated in growth medium supplemented with 2 mm MgCl2, to prevent excitotoxicity. They were used for electrophysiological recordings within 24 h.

Culture of dissociated hippocampal neurons

Low-density cultures of acutely dissociated hippocampal neurons were prepared from Sprague Dawley rat embryos (Envigo) of unknown sex, at embryonic day 18 (E18) with minor adjustments from previously described methods (Misonou and Trimmer, 2005; Borschel et al., 2012). Briefly, a pregnant rat was killed in a CO2 chamber and quickly decapitated, and the uterus was surgically removed. Embryos were decapitated, and the hippocampi were removed and placed in ice-cold dissecting solution containing HBSS supplemented with 4 mm sodium bicarbonate (Sigma), 10 mm HEPES (Sigma), and 1% penicillin/streptomycin (Corning). Cells were enzymatically dissociated with 0.25% trypsin (20 min at 37°C), and then gently triturated and filtered through a 40 mm strainer (BD Falcon). Dissociated cells were counted and plated at a density of 100,000 cell/cm2 onto glass coverslips precoated with poly-D-lysine (Corning) in plating media containing MEM (Invitrogen) supplemented with 10% FBS, 0.6% glucose (Sigma), 2 mm GlutaMAX (Invitrogen), 1 mm sodium pyruvate (Sigma), and 1% penicillin/streptomycin. Within a few hours, after cells have adhered to plates, the medium was gently replaced with Neurobasal A medium (Invitrogen) supplemented with B27 (Invitrogen) and 2 mm GlutaMAX. Three days after plating, the proliferation of non-neuronal cells was inhibited by including arabinofuranosylcytosine (5 µm, Sigma). Neurons were used for electrophysiological measurements between 7 and 30 DIV.

Electrophysiology

To maintain consistency in seal formation with minimal mechanical disruption to the patch, we used the following procedure. Before entering the bath, we applied slight positive pressure (5 mmHg) through the recording pipette with a high speed pressure-clamp system (HSPC-1, ALA Scientific) (McBride and Hamill, 1993, 1999). Electrical resistance through the pipette (20 ± 5 mΩ) was monitored by observing the amplitude of the current elicited by a test voltage-pulse. After contacting the cell, the positive pressure was released to 0 mmHg, and slight suction (−5 mmHg) was applied to initiate slow seal formation onto the cellular membrane, which was monitored as an increase in pipette resistance. Finally, after obtaining a high-resistance seal, we released the negative pressure and applied 100 mV to the patch to visualize the activity of NMDARs at 0 mmHg, as inward Na+ currents.

To examine the dependency of channel activity on the level of applied pressure, cells were bathed in PBS; after seal formation, we applied pressure in increments of 10 mmHg, and recorded activity for periods lasting ∼5 min for each pressure level, over the indicated range. When specified, a 5 min recovery step was recorded after relaxing the pressure to 0 mmHg. Channel activity was evaluated in cell-attached patches obtained with pipettes filled with the following (in mm): 150 NaCl, 2.5 KCl, 10 HEPES, 1 EDTA, pH 8.0 (NaOH) and the indicated agonists glutamate (1 mm), glycine (0.1 mm), or NMDA (0.1 mm), as previously described (Hamill et al., 1981; Maki et al., 2014). Solutions lacking agonists were prepared using double-distilled deionized ultrapure water (Fisher Scientific) to prevent contamination (Cummings and Popescu, 2015).

To examine the effect of pressure on the receptor's conductance, Ca2+ permeability, and voltage dependency of its Mg2+ blockade, cells were bathed in a high K+ bath solution as follows (in mm): 142 KCl, 5 NaCl, 1.8 CaCl2, 1.7 MgCl2, 10 HEPES, pH 7.2 (with KOH) to collapse the physiological membrane potential of HEK cells, which is ∼10 mV (Borschel et al., 2012). Pipette solution was as follows (in mm): 150 NaCl, 2.5 KCl, 10 HEPES, 1 EDTA, 10 tricine, pH 8.0, and glycine (0.1) and/or glutamate (1) as indicated. Ca2+ and Mg2+ were added as chloride salts and were buffered to the indicated free concentration according to MAXCHELATOR software. After seal formation, we applied sustained suction (−40 mmHg) and varied the applied voltage in 20 mV increments, each lasting 1 min, over the 100 to 20 mV range.

All current traces were filtered (10 kHz), amplified (Axopatch 200b), and then sampled (40 kHz) and stored as digital files using QuB software (Nicolai and Sachs, 2013).

Data analysis

Current traces were inspected visually offline and only recordings with low-noise and stable-baseline were selected for analyses. Traces were initially processed to correct for spurious noise events and minor baseline drifts (Maki et al., 2014). Corrected traces were idealized separately for each applied pressure within the QuB suite for kinetic analyses, with the SKM algorithm after applying a digital filter (12 kHz) (Qin, 2004). We estimated the open probability (nPo) in each trace according to the following relationship: nPo∑n=1Nn⋅Po(n)/N

Where Po is the open probability of each channel, n is the indeterminate number of channels in each patch, and N is the minimum number of channels in each patch, estimated as the number of overlapping unitary currents (simultaneous openings) observed in the condition producing maximal activity. Values for nPo were obtained by averaging activity in each 5 min segment, and were considered non-zero for a threshold of >1000 events.

Unitary channel conductance (γ) and reversal potential (Erev) were estimated from linear fits to the unitary current–voltage relationship measured over a 1 min period. Ca2+ permeability was estimated as a function of the measured Ca2+-induced shifts in Erev using the Lewis equation below (Lewis, 1979), with the experimental constant α = 25.4 mV. PCaPNa=[Na](eΔErevα−1)4[Ca]

Statistics

Results are given as the mean ± SEM of a minimum of three measurements per condition. Statistical analyses were performed using two-way ANOVA multiple comparisons and the Bonferroni correction, or unpaired Student's t test relative to controls measured at zero pressure, as indicated. Means were considered significantly different for p < 0.05.

Results

Membrane stretch substitutes for glutamate in gating NMDARs

NMDARs are tetrameric transmembrane proteins that assemble from three subfamilies of subunits: glycine-binding GluN1 and GluN3(A, B), and glutamate-binding GluN2(A-D). Functional NMDARs assemble as heterotetramers of two obligatory GluN1 subunits, which are widely expressed in cells of the CNS, and two of GluN2 and/or GluN3 subunits whose expression is regulated developmentally and regionally. Of the glutamate-binding GluN2 subunits, GluN2A predominates in adult animals and at mature synapses, whereas GluN2B is expressed mostly in juvenile animals and at immature synapses (Monyer et al., 1992; Goebel and Poosch, 1999; Paoletti et al., 2013).

To examine whether NMDARs are simply mechanically sensitive or whether they can be gated by mechanical forces in the absence of neurotransmission, we expressed rat recombinant GluN1/GluN2A receptors in HEK293 cells and recorded inward Na+ currents from cell-attached patches, while gently varying the pressure applied through the recording pipette in 10 mmHg increments over the −40 mmHg to 40 mmHg range. These pressures are typical for the activation of dedicated mechanotransducers, such as piezo channels (Coste et al., 2012; Kim et al., 2012). Observing NMDAR activity over long periods is necessary to reduce patch-to-patch variability because of modal gating, which for NMDARs occurs on a minutes time scale (Popescu and Auerbach, 2003; Borschel et al., 2012). Therefore, at each pressure level, we recorded 5 min of continuous activity.

When the recording pipette included supra-saturating levels of the neurotransmitter glutamate (1 mm; Kd, 3 µm) (Popescu et al., 2004) and the obligatory coagonist glycine (0.1 mm; Kd, 2.5 µm) (Cummings and Popescu, 2015), applying 100 mV through the pipette produced large inward unitary currents (8–10 pA) indicative of channel activation, at all levels of applied pressure tested (Fig. 1A, top traces). Often, overlapping openings were apparent, indicating that multiple active channels were trapped in the recorded patch. In these conditions, neither negative nor positive pressure altered channel activity. When glycine was omitted, we observed only minimal and sporadic currents (<1000 events per 5 min segment), regardless of whether glutamate was present or not, and applying either negative or positive pressure did not alter this low baseline activity (Fig. 1A, middle traces). However, when glycine was present, negative but not positive pressure gated substantial current in the absence of glutamate (Fig. 1A, bottom traces). The suction-gated current increased with increasing pressure in a consistent manner, although the magnitude of the effect varied. On average, −40 mmHg of hydrostatic pressure increased GluN1/GluN2A channel activity (nPo) from 0.10 ± 0.06 to 0.50 ± 0.18 (n = 6, p = 0.007) (Fig. 1A,B). This result demonstrates that suction alone can gate GluN1/GluN2A receptors; therefore, it is possible to open the NMDAR pore mechanically, in the absence of neurotransmission.

Figure 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 1.

Mild suction gates NMDARs in the presence of glycine. A, Current traces recorded from cell-attached patches expressing GluN1/GluN2A receptors with 100 mV applied through the recording pipette. Downward traces represent inward Na+ currents at the indicated pressure levels, in the presence (+) or absence (–) of glutamate (Glu, 1 mm) and/or glycine (Gly, 0.1 mm). B, Summary of response dependency on pressure level for each GluN2 subtype in the presence of glycine (0.1 mm) with no glutamate added. *p < 0.05; **p < 0.01; one-way ANOVA, with Bonferroni correction.

To ascertain whether pressure can gate currents from other members of the NMDAR family, we coexpressed GluN1 with GluN2B, GluN2C, or GluN2D subunits in HEK293 cells and recorded single-channel inward Na+ currents from cell-attached patches with pipettes containing glycine (0.1 mm) but not glutamate. As with the adult GluN1/GluN2A receptor, we observed a selective increase in channel activity with negative pressure, and no effect with positive pressure of similar magnitude (Fig. 1B). The application of negative pressure increased the nPo for GluN2B from 0.03 ± 0.02 at 0 mmHg to 0.43 ± 0.15 at −30 mmHg (n = 4, p = 0.03). We could not detect significant changes for GluN2C and for GluN2D channels, for which measured averages at 0 and –40 mmHg, were as follows: 0.010 ± 0.004 and 0.04 ± 0.01 (n = 5, p > 0.05), and 0.07 ± 0.03 and 0.18 ± 0.04 (n = 4, p > 0.05), respectively. This may reflect in part the well-documented high kinetic variability of GluN2B- and GluN2C-containing receptors (Amico-Ruvio and Popescu, 2010; Khatri et al., 2014), and the low open probability of GluN2D-containing receptors, which makes detection more challenging (Vance et al., 2013), and their much lower maximal open probabilities measured with glutamate: 0.16 ± 0.02 for GluN2B (Borschel et al., 2012), 0.032 ± 0.015 for GluN2C (Khatri et al., 2014), and 0.023 ± 0.001, for GluN2D (Vance et al., 2013).

Overall, these results support the hypothesis that mechanical forces, in addition to modulating the glutamate-gated current, can by themselves provide the energy necessary to shift the receptor's closed-to-open equilibrium and produce a detectable increase in open probability. We focused next on GluN1/GluN2A receptors, which generally produce more robust and reliable responses (Borschel et al., 2012).

Mindful of the many sources that can contribute to the variability of the observed changes, we aimed to reduce the incidence of confounding effects because of cellular processes over the long recording period necessary to cover the 80 mmHg range investigated with the protocol above. For this, we shortened the experiment by limiting observations to negative pressure, which allowed us to add a 5 min recovery step to test the reversibility of the pressure-dependent effect. As in the first set of experiments, with this shorter protocol, we found that negative pressure had no effect on channel activity in the absence of glycine, or in the presence of saturating concentrations of glycine and glutamate (Fig. 2; Table 1). However, when glutamate was omitted, −40 mmHg of pressure increased the observed current (nPo) from 0.03 ± 0.01 to 0.08 ± 0.05 (n = 4, p = 0.006), which represented 19% of the maximal glutamate-gated current in the same conditions (0.41 ± 0.07, n = 4). The ambient glutamate concentration at extrasynaptic sites in adult rat hippocampal slices is estimated at 25-80 nm (Herman and Jahr, 2007; Moldavski et al., 2020), which represents <10% of the synaptic concentration (∼1 mm) (Clements et al., 1992; Wadiche and Jahr, 2001; Budisantoso et al., 2013). Therefore, the level of activity we observed with mild stretch is on par with that reported for extrasynaptic receptors activated by synaptic glutamate spillover or by glutamate leak from injured neurons (Moldavski et al., 2020), and may be physiologically significant if the stretch-gated currents maintain the biophysical properties of glutamate-gated currents, especially their large unitary conductance, high Ca2+ permeability, and voltage-dependent Mg2+ block. Therefore, we next examined these biophysical properties of the stretch-gated current.

View this table:
  • View inline
  • View popup
Table 1.

Summary of activity recorded in cell-attached patchesa

Figure 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 2.

Negative hydrostatic pressure agonizes NMDARs. A, Effect of suction on GluN1/GluN2A receptors, with the indicated agonists (Glu 1 mm, Gly 0.1 mm) and 100 mV in the cell-attached recording pipette. *p < 0.05; **p < 0.01; one-way ANOVA with Bonferroni correction. B, Summary of the effects of suction on receptor activity. *p < 0.05 (unpaired Student's t test).

Biophysical properties of stretch-gated NMDAR currents

Within the larger family of glutamate-gated channels, NMDARs have characteristically large unitary conductance, high Ca2+ permeability, and voltage-dependent Mg2+ block (Hansen et al., 2018). These distinctive biophysical properties of the glutamate-gated current are essential for the many roles NMDARs play in health and disease (Iacobucci and Popescu, 2017; Hansen et al., 2018). To estimate the conductance and permeability properties of the stretch-gated receptors from cell-attached recordings, we bathed the cells in a high K+ solution to collapse the cellular transmembrane potential. After gentle seal formation, we applied −40 mmHg of pressure and recorded activity at several applied pipette potentials for 1 min periods (Fig. 3A). From these data, we measured unitary current amplitude at each voltage, and estimated the unitary conductance as the slope of the voltage–current relationship (Fig. 3B). Relative to the glutamate-gated Na+ currents, which had γNa = 81 ± 9 pS (n = 5), stretch-gated currents had similar unitary Na+ conductance, γNa = 87 ± 6 pS (n = 3, p > 0.5) (Fig. 3B). Therefore, stretch-gated currents retain the high unitary conductance characteristic of NMDARs.

Figure 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 3.

Biophysical properties of stretch-gated currents from recombinant NMDARs. A, On-cell patch-clamp current traces recorded from cells expressing GluN1/GluN2A receptors in response to saturating concentrations of Glu (1 mm) (left) and gentle stretch (−40 mmHg). B, Voltage dependency of unitary current amplitude, and summary of Ca2+-dependent reduction in unitary conductance. *p < 0.05; **p < 0.01; two-way ANOVA, with Bonferroni correction. C, Current traces recorded with external Mg2+ (10 µm) and summary of voltage-dependent reduction in mean open durations.

In physiological conditions, external Ca2+ permeates NMDARs and concurrently reduces channel conductance (voltage-independent block). To examine how external calcium affects stretch-gated currents, we measured single-channel current amplitudes of glutamate-gated and stretch-gated currents at several applied voltages, in the presence of external calcium. We found that 1.8 mm Ca2+ reduced the glutamate-gated conductance to γ = 61 ± 2 pS (p < 0.05) indicative of ∼25% current blockade (Fig. 3B), a value consistent with previous reports (Ascher and Nowak, 1988; Maki and Popescu, 2014). Similarly, 1.8 mm Ca2+ reduced the amplitude of stretch-gated currents to γ1.8 = 51 ± 6 pS (p < 0.01), and this reduction was not statistically different in magnitude from that observed for glutamate-evoked currents (p = 0.19, two-way ANOVA) (Fig. 3A, B).

From the same data, we constructed linear fits to the current–voltage relationships obtained in 0 and 1.8 mm Ca2+, to estimate reversal potentials for each condition. Relative to 0 Ca2+, in 1.8 mm Ca2+, the reversal potential of glutamate-gated currents shifted by 6 mV, indicative of a high relative Ca2+ permeability, PCa/PNa = 10.7, as reported previously (Wollmuth and Sakmann, 1998; Maki and Popescu, 2014). For stretch-activated currents, the measured shift in reversal potential was 16 mV, corresponding to twofold increase in permeability, PCa/PNa = 21, relative to glutamate-gated currents. Together, these measurements suggest that, in physiologic Ca2+ concentrations, membrane stretch gates NMDAR currents that maintain characteristic high unitary conductance, and voltage-independent Ca2+ block, and may have slightly stronger higher Ca2+ permeability relative to the glutamate-gated currents.

Last, we examined the sensitivity of the stretch-gated current to block by external Mg2+. We recorded on-cell single-channel currents from GluN1/GluN2A receptors at several applied voltages, with pipettes containing glycine (0.1 mm), Mg2+ (10 µm, Kd = 1 µm) (Premkumar and Auerbach, 1996), and either glutamate (1 mm) or sustained negative pressure (−40 mmHg) (Fig. 3C). At each voltage, we identified nonoverlapping bursts of activity and measured the channel mean open time as a measure of Mg2+ block. We found that glutamate-gated currents were sensitive to block by external Mg2+ in a voltage-dependent manner, such that the mean duration of openings decreased from 5.1 ± 2 ms at −20 mV, to 1.0 ± 0.2 ms at −60 mV, as reported previously (Nowak et al., 1984). For stretch-gated currents, we observed a similar shortening of open durations with hyperpolarization, from 4.0 ± 0.6 ms at −20 mV, to 1.0 ± 0.05 ms at −60 mV (p < 0.05, paired Student's t test), indicating similar sensitivity to voltage-dependent block (Fig. 3C) (Premkumar and Auerbach, 1996). At all examined voltages, the difference between the estimated mean open durations for glutamate-gated and stretch-gated currents was not statistically significant (p > 0.05, two-way ANOVA). Together, these results indicate that stretch-activated channels maintain the characteristic biophysical properties of glutamate-gated channels, including high conductance, large Ca2+ permeability, strong voltage-dependent Mg2+ block, and long openings.

Stretch-gated NMDA currents require the receptor's carboxyl terminal

Given the potentially significant physiological implications of Ca2+-rich currents gated by mechanical forces through NMDARs, it will be important to understand the mechanism by which these arise, and more specifically, to identify the allosteric network responsible for mechanotransduction. The existing literature on the mechanosensitivity of NMDARs suggests several mechanisms by which mechanical forces may facilitate the glutamate-gated current. These include a reduction of Mg2+ block (L. Zhang et al., 1996; Kloda et al., 2007; Mor and Grossman, 2010), perhaps transmitted through the transmembrane domain (Casado and Ascher, 1998), but also allosteric mechanisms that implicate the C-terminal domain (CTD) (Singh et al., 2012). For the stretch-gated current, our results exclude a mechanism mediated by changes in Mg2+ block. Therefore, we asked whether the CTD influences the receptor's mechanically elicited current.

We recorded single-channel currents from on-cell patches expressing receptors lacking the intracellular CTD (GluN1Δ838/GluN2AΔ844). We reported previously that relative to WT receptors (Po = 0.54 ± 0.04), glutamate-gated currents from these truncated receptors have lower but measurable open probabilities (ΔCTD, 0.08 ± 0.02, n = 8, p < 0.5) (Maki et al., 2012). Using the pressure protocol described here, with only glycine in the pipette and no external pressure, we observed low spontaneous activity from ΔCTD receptors (0.05 ± 0.01, n = 4), which was not different from WT GluN1/Glu2A (Fig. 4; Table 1). However, suction up to −40 mmHg did not increase the basal activity of truncated receptors (0.04 ± 0.01, n = 3) (Fig. 4; Table 1). This result suggests that the ΔCTD of GluN1/GluN2A receptors is necessary for their mechanical activation by mild suction. This observation may indicate that the CTD is necessary to transmit force from the cytoskeleton to the gate; alternatively, it may indicate that the energy provided by suction, transmitted by some other unknown mechanism, is enough to gate the channel only when the tethering of the CTD to intracellular structures endow the receptor under observation a certain threshold of rigidity.

Figure 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 4.

Mechanical activation of NMDARs by suction requires their intracellular CTD. Cell-attached Na+-current traces recorded from GluN1/GluN2A receptors lacking the intracellular CTD (ΔCTD) (left) and summary of results compared with WT receptors. *p < 0.05 (unpaired Student's t test).

Mechanical activation of neuronal NMDARs

Regardless of mechanism, our result that the intracellular domain is required for mechanical gating of currents from NMDARs suggests that the intracellular milieu in which NMDARs operate, and specifically the intracellular interactions mediated by their CTD will influence the effectiveness with which hydrostatic pressure will gate currents from glycine-bound receptors. In addition, lipid composition of membranes varies widely across cell type, development stage, and subcellular location, and can be a critical determinant of mechanotransduction (Perozo et al., 2002; Phillips et al., 2009). We therefore investigated the effectiveness of hydrostatic pressure to gate NMDARs in a neuronal environment.

We cultured primary rat hippocampal neurons (P7-P30), and recorded cell-attached currents with pipette solutions containing low concentrations of NMDA (0.1 mm; EC50, 90 µm) (Erreger et al., 2007) and glycine (0.1 mm) to identify currents mediated by endogenous NMDARs. We observed inward Na+ currents with large unitary amplitudes (8.9 ± 0.3 pA) consistent with NMDAR activation. Hydrostatic pressure (−40 mmHg) increased substantially the measured nPo from 0.13 ± 0.02 at rest to 0.40 ± 0.04 (n = 5, p < 0.0001, one-way ANOVA); this potentiation was fully reversible (Fig. 5B; Table 1) and mirrored results obtained with low NMDA and glycine from GluN1/GluN2A receptors in HEK293 cells (Fig. 5A; Table 1). In similar experiments, and with only glycine in the pipette, the average nPo measured in neurons was 0.04 ± 0.01 at rest, and 0.07 ± 0.02 (n = 4) with −40 mmHg, and these values were not statistically different (one-way ANOVA, with Bonferroni correction) (Fig. 5B; Table 1).

Figure 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 5.

Stretch gates native NMDARs. A, Suction potentiates currents elicited from recombinant GluN1/GluN2A receptors with low concentration of NMDA (0.1 mm). B, In neurons, suction potentiates NMDA-elicited currents and gates currents of similar unitary amplitude. **p < 0.01; ***p < 0.001; ****p < 0.0001; one-way ANOVA test with Bonferroni correction.

Together, these observations validate the results obtained with recombinant receptors in HEK cells and support our proposal that mild stretch can gate native NMDARs in the absence of neurotransmission, and likely potentiate responses elicited by low concentrations of glutamate (<0.1 mm), as may occur at extrasynaptic locations (Moldavski et al., 2020).

Discussion

Glutamate-gated NMDAR currents can be modulated by several types of mechanical perturbations, including those generated by changes in environmental pressure (Fagni et al., 1987; Mor and Grossman, 2006), membrane composition (Miller et al., 1992; Nishikawa et al., 1994; Casado and Ascher, 1998), osmotic and hydrostatic pressure (Paoletti and Ascher, 1994; LaPlaca and Thibault, 1998), and microfluidic sheer stress (Maneshi et al., 2017). Although the effects of mechanical stimulation on channel responses vary across stimulation procedure, receptor preparations, and experimental conditions, these results have established that NMDARs are mechanosensitive (Paoletti and Ascher, 1994). Here, we report that gentle suction can activate NMDARs in the absence of glutamate. This observation establishes that NMDARs, in addition to being mechanosensitive, can be activated mechanically, which is consequential to understanding the mechanobiology of the CNS. Before addressing this point of impact, we note several caveats.

As previously reported for mechano-sensitivity (Paoletti and Ascher, 1994), the mechano-activity we observed here was variable, despite taking a number of experimental precautions. Among these, we examined recombinant receptors residing in cell-attached membrane patches. This approach minimizes variability due the uncertain molecular composition of endogenous receptors; it maintains cellular integrity and a near-physiologic cellular environment; it allows precise control of the magnitude of the applied pressure with a high-speed pressure clamp; and provides a high-resolution single-molecule readout for receptor activity. Nonetheless, a direct correlation between the applied pressure and the receptor's microscopic properties is complicated by several uncontrollable variables. First, even when using pipettes of specified geometry, the area of the electrically accessible membrane patch (the dome delimited by the seal) varies from patch to patch and can change during a single recording because of membrane creep (Suchyna et al., 2009). Further, the tension experienced by receptors varies with their position within the patch, being highest at the apex and lowest near the perimeter (Bavi et al., 2014). Last, the size and mechanical properties of the cytosolic mass pulled within the pipette are not uniform across observations, and can detach from the bilayer on continuous mechanical stimulation. Such blebbing may produce additional inconsistencies in the magnitude of the force that reaches the receptor and can also modify the receptor's gating properties (Suchyna et al., 2009). Last, NMDARs display intrinsic gating heterogeneity because of modal gating (Popescu and Auerbach, 2003; Popescu, 2012; Vance et al., 2013), which is responsible for the characteristic biphasic decay in their macroscopic response (W. Zhang et al., 2008). As such, even in controlled experimental conditions, the measured equilibrium open probability of NMDARs varies considerably (Borschel et al., 2012; Vance et al., 2013). Moreover, receptor activity is sensitive to cellular factors that may vary from cell to cell and may change during extended recording periods (Chen and Huang, 1991; Cerne et al., 1993; Tong et al., 1995; Wyszynski et al., 1997). With these considerations in mind, the magnitude of the changes in activity we observed with gentle suction are consistent with substantial mechano-activation of NMDARs.

This observation is important for several reasons. Controlled NMDAR-mediated Ca2+ is required for the normal physiology of excitatory synapses, and mechanical forces may be important in initiating these processes during development and throughout life (Tyler, 2012). Alternatively, NMDAR Ca2+ can also initiate synaptic pruning, spine shrinkage, and neuronal death. NMDARs are expressed not only at postsynaptic, mechanically stable locations, but also in mechanically active or osmotically sensitive zones, such as growing axons or dendritic boutons, where local deformations in extracellular matrix, membrane tension or curvature, and intracellular cytoskeleton can impinge mechanically on receptors. Therefore, NMDARs operate in a mechanically rich landscape and, depending on their location, may experience differential mechanical forces. Our results show no effect of membrane stretch on currents elicited with maximally effective glutamate concentrations (Figs. 1A and 2; Table 1). Therefore, it is unlikely that this mechanism will influence synaptic transmission. However, the levels of mechano-activation we observed with gentle membrane stretch can have a significant impact on signal transduction by neuronal extrasynaptic NMDARs, or those expressed in glial cells. Additionally, NMDARs, of unknown function, have been identified at nontraditional sites, such as gastrointestinal, lung, and adrenal tissue during human development (Szabo et al., 2015); and in adult tissues, such as kidney (Leung et al., 2002), bone (Itzstein et al., 2001), myocytes (Seeber et al., 2004; Dong et al., 2021), colon (Del Valle-Pinero et al., 2007) and others, such as cancerous tissue (Yan et al., 2021). Therefore, the significance of the mechano-activity described here will vary with the site of NMDAR expression and their microenvironment.

For GluN2A-containing receptors, which is the most prevalent NMDAR isoform expressed in adult mammals, −40 mmHg of pressure produced currents that had 19% open probability, 80% unitary conductance, and 200% Ca2+ permeability, relative to the current produced by saturating glutamate (1 mm) in similar conditions (1.8 mm Ca2+). With the more sensitive protocol illustrated in Figure 2, the response did not appear to plateau with −40 mmHg; therefore, it is possible that stronger forces may elicit higher activity. Together with the observation reported here and previously (Paoletti and Ascher, 1994; Casado and Ascher, 1998) that gentle membrane stretch potentiates responses elicited with low concentrations of the GluN2-site agonist (glutamate or NMDA), the mechano-activity of NMDARs may represent an important physiologic mechanism, especially in development or at sites of dendritic growth and synaptic formation. Alternatively, inappropriate mechanical activation of extrasynaptic NMDARs, because of, for example, external mechanical forces experienced by brain or spinal cord, may initiate or aggravate apoptotic or necrotic cell injury through additional Ca2+ influx.

In some experimental paradigms, the mechanosensitivity of NMDARs reflects mechanically induced changes in the receptor's sensitivity to voltage-dependent Mg2+ block (L. Zhang et al., 1996; Kloda et al., 2007; Parnas et al., 2009; Cox et al., 2019). Our measurements were done in the absence of external Mg2+, and we were able to demonstrate similar voltage-dependent block for stretch-gated and glutamate-gated currents (Fig. 3C); therefore, we can definitively exclude this mechanism for the stretch-gated activity we examined here. Aside from modulating Mg2+ block, previous reports found mechanosensitivity to depend on the receptor's intracellular CTD (Singh et al., 2012; Bliznyuk et al., 2015). In our hands, the CTD of NMDARs was required for mechanical activation by gentle membrane stretch.

Given the modular makeup of NMDARs, and their complex interactions with extracellular matrix proteins, membrane proteins, and lipids, and with intracellular proteins and cytoskeletal components, it is likely that, depending on the type of stimulation, mechanical forces will impinge on separate receptor domains. For example, in the experiments reported by the Martinac group (Kloda et al., 2007), when mechanosensitivity was tested on purified recombinant NMDARs inserted in liposomal particles, it was reasonable to infer a force-from lipid transduction mechanism, given the absence of interacting proteins or cellular structures. However, when operating in their native environments, receptors are much more mechanically constrained and they can sense membrane deformation not only through direct interactions with lipid but also through their extracellular or intracellular domains. In addition, mechanical constraints imposed by interaction with cellular and/or intracellular structures, even if not serving as mechanical transducers, will limit the receptor's conformational freedom and thus affect their sensitivity to mechanical activation. Therefore, although the CTD appears necessary for mechanical gating of NMDARs, it remains to be determined whether this is a case of force-from filament mechanism, or the CTD simply stabilizes the molecule in mechano-sensitive conformations.

Therefore, although unknown at this time, it will be important to determine the mechanism by which mechanical forces gate NMDAR currents. Mechano-activation of NMDARs may be of importance at extrasynaptic and non-neuronal sites in the CNS, where it may contribute to fundamental processes, such as synapse formation, dendrite remodeling, and glial physiology, and outside of the nervous system in gastric and pulmonary development, and cardiac and bone remodeling. Conversely, this knowledge may help better understand, and therefore prevent or address, neuropsychiatric disorders, such as shaken infant syndrome, chronic traumatic encephalopathy, and other trauma-associated neuropathologies.

Footnotes

  • This work was supported by National Institutes of Health R21NS098385, R01NS052669, and R01NS097016 to G.K.P. We thank Eileen Kasperek for assistance with molecular biology and tissue culture; Richard Burke, Cheryl Movsesian, and Ayman Mustafa for sharing recordings; and Gary J. Iacobucci for assistance in developing the MATLAB code to analyze the SKM data.

  • The authors declare no competing financial interests.

  • Correspondence should be addressed to Gabriela K. Popescu at popescu{at}buffalo.edu

SfN exclusive license.

References

  1. ↵
    1. Amico-Ruvio SA,
    2. Popescu GK
    (2010) Stationary gating of GluN1/GluN2B receptors in intact membrane patches. Biophys J 98:1160–1169. doi:10.1016/j.bpj.2009.12.4276 pmid:20371315
    OpenUrlCrossRefPubMed
  2. ↵
    1. Ascher P,
    2. Nowak L
    (1988) The role of divalent cations in the N-methyl-D-aspartate responses of mouse central neurones in culture. J Physiol 399:247–266. doi:10.1113/jphysiol.1988.sp017078 pmid:2457089
    OpenUrlCrossRefPubMed
  3. ↵
    1. Bavi N,
    2. Nakayama Y,
    3. Bavi O,
    4. Cox CD,
    5. Qin QH,
    6. Martinac B
    (2014) Biophysical implications of lipid bilayer rheometry for mechanosensitive channels. Proc Natl Acad Sci USA 111:13864–13869. doi:10.1073/pnas.1409011111 pmid:25201991
    OpenUrlAbstract/FREE Full Text
  4. ↵
    1. Bliznyuk A,
    2. Aviner B,
    3. Golan H,
    4. Hollmann M,
    5. Grossman Y
    (2015) The N-methyl-D-aspartate receptor's neglected subunit - GluN1 matters under normal and hyperbaric conditions. Eur J Neurosci 42:2577–2584. doi:10.1111/ejn.13022 pmid:26202884
    OpenUrlCrossRefPubMed
  5. ↵
    1. Bliznyuk A,
    2. Hollmann M,
    3. Grossman Y
    (2020) The mechanism of NMDA receptor hyperexcitation in high pressure helium and hyperbaric oxygen. Front Physiol 11:1057. doi:10.3389/fphys.2020.01057 pmid:32982789
    OpenUrlCrossRefPubMed
  6. ↵
    1. Bonnier C,
    2. Mesples B,
    3. Gressens P
    (2004) Animal models of shaken baby syndrome: revisiting the pathophysiology of this devastating injury. Pediatr Rehabil 7:165–171. doi:10.1080/13638490410001703325 pmid:15204568
    OpenUrlCrossRefPubMed
  7. ↵
    1. Borschel WF,
    2. Myers JM,
    3. Kasperek EM,
    4. Smith TP,
    5. Graziane NM,
    6. Nowak LM,
    7. Popescu GK
    (2012) Gating reaction mechanism of neuronal NMDA receptors. J Neurophysiol 108:3105–3115. doi:10.1152/jn.00551.2012 pmid:22993263
    OpenUrlCrossRefPubMed
  8. ↵
    1. Budisantoso T,
    2. Harada H,
    3. Kamasawa N,
    4. Fukazawa Y,
    5. Shigemoto R,
    6. Matsui K
    (2013) Evaluation of glutamate concentration transient in the synaptic cleft of the rat calyx of Held. J Physiol 591:219–239. doi:10.1113/jphysiol.2012.241398
    OpenUrlCrossRefPubMed
  9. ↵
    1. Casado M,
    2. Ascher P
    (1998) Opposite modulation of NMDA receptors by lysophospholipids and arachidonic acid: common features with mechanosensitivity. J Physiol 513:317–330. doi:10.1111/j.1469-7793.1998.317bb.x
    OpenUrlCrossRefPubMed
  10. ↵
    1. Cerne R,
    2. Rusin KI,
    3. Randic M
    (1993) Enhancement of the N-methyl-D-aspartate response in spinal dorsal horn neurons by cAMP-dependent protein kinase. Neurosci Lett 161:124–128. doi:10.1016/0304-3940(93)90275-p pmid:8272253
    OpenUrlCrossRefPubMed
  11. ↵
    1. Chen L,
    2. Huang LY
    (1991) Sustained potentiation of NMDA receptor-mediated glutamate responses through activation of protein kinase C by a mu opioid. Neuron 7:319–326. doi:10.1016/0896-6273(91)90270-a pmid:1678615
    OpenUrlCrossRefPubMed
  12. ↵
    1. Clements JD,
    2. Lester RA,
    3. Tong G,
    4. Jahr CE,
    5. Westbrook GL
    (1992) The time course of glutamate in the synaptic cleft. Science 258:1498–1501. doi:10.1126/science.1359647 pmid:1359647
    OpenUrlAbstract/FREE Full Text
  13. ↵
    1. Coste B,
    2. Xiao B,
    3. Santos JS,
    4. Syeda R,
    5. Grandl J,
    6. Spencer KS,
    7. Kim SE,
    8. Schmidt M,
    9. Mathur J,
    10. Dubin AE,
    11. Montal M,
    12. Patapoutian A
    (2012) Piezo proteins are pore-forming subunits of mechanically activated channels. Nature 483:176–181. doi:10.1038/nature10812 pmid:22343900
    OpenUrlCrossRefPubMed
  14. ↵
    1. Cox CD,
    2. Bavi N,
    3. Martinac B
    (2019) Biophysical principles of ion-channel-mediated mechanosensory transduction. Cell Rep 29:1–12. doi:10.1016/j.celrep.2019.08.075 pmid:31577940
    OpenUrlCrossRefPubMed
  15. ↵
    1. Cummings KA,
    2. Popescu GK
    (2015) Glycine-dependent activation of NMDA receptors. J Gen Physiol 145:513–527. doi:10.1085/jgp.201411302 pmid:25964432
    OpenUrlAbstract/FREE Full Text
  16. ↵
    1. Del Valle-Pinero AY,
    2. Suckow SK,
    3. Zhou Q,
    4. Perez FM,
    5. Verne GN,
    6. Caudle RM
    (2007) Expression of the N-methyl-D-aspartate receptor NR1 splice variants and NR2 subunit subtypes in the rat colon. Neuroscience 147:164–173. doi:10.1016/j.neuroscience.2007.02.063 pmid:17509768
    OpenUrlCrossRefPubMed
  17. ↵
    1. Dong YN,
    2. Hsu FC,
    3. Koziol-White CJ,
    4. Stepanova V,
    5. Jude J,
    6. Gritsiuta A,
    7. Rue R,
    8. Mott R,
    9. Coulter DA,
    10. Panettieri RA Jr.,
    11. Krymskaya VP,
    12. Takano H,
    13. Goncharova EA,
    14. Goncharov DA,
    15. Cines DB,
    16. Lynch DR
    (2021) Functional NMDA receptors are expressed by human pulmonary artery smooth muscle cells. Sci Rep 11:8205. doi:10.1038/s41598-021-87667-0 pmid:33859248
    OpenUrlCrossRefPubMed
  18. ↵
    1. Erreger K,
    2. Geballe MT,
    3. Kristensen A,
    4. Chen PE,
    5. Hansen KB,
    6. Lee CJ,
    7. Yuan H,
    8. Le P,
    9. Lyuboslavsky PN,
    10. Micale N,
    11. Jorgensen L,
    12. Clausen RP,
    13. Wyllie DJ,
    14. Snyder JP,
    15. Traynelis SF
    (2007) Subunit-specific agonist activity at NR2A-, NR2B-, NR2C-, and NR2D-containing N-methyl-D-aspartate glutamate receptors. Mol Pharmacol 72:907–920. doi:10.1124/mol.107.037333 pmid:17622578
    OpenUrlAbstract/FREE Full Text
  19. ↵
    1. Fagni L,
    2. Zinebi F,
    3. Hugon M
    (1987) Helium pressure potentiates the NMDA and DHS-induced decreases of field potentials in the rat hippocampal slice preparation. Neurosci Lett 81:285–290. doi:10.1016/0304-3940(87)90397-1
    OpenUrlCrossRefPubMed
  20. ↵
    1. Goebel DJ,
    2. Poosch MS
    (1999) NMDA receptor subunit gene expression in the rat brain: a quantitative analysis of endogenous mRNA levels of NR1, NR2A, NR2B, NR2C, NR2D and NR3A. Brain Res Mol Brain Res 69:164–170. doi:10.1016/s0169-328x(99)00100-x pmid:10366738
    OpenUrlCrossRefPubMed
  21. ↵
    1. Goriely A,
    2. Geers MG,
    3. Holzapfel GA,
    4. Jayamohan J,
    5. Jerusalem A,
    6. Sivaloganathan S,
    7. Squier W,
    8. van Dommelen JA,
    9. Waters S,
    10. Kuhl E
    (2015) Mechanics of the brain: perspectives, challenges, and opportunities. Biomech Model Mechanobiol 14:931–965. doi:10.1007/s10237-015-0662-4 pmid:25716305
    OpenUrlCrossRefPubMed
  22. ↵
    1. Hamill OP,
    2. Marty A,
    3. Neher E,
    4. Sakmann B,
    5. Sigworth FJ
    (1981) Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflugers Arch 391:85–100. doi:10.1007/BF00656997 pmid:6270629
    OpenUrlCrossRefPubMed
  23. ↵
    1. Hansen KB,
    2. Yi F,
    3. Perszyk RE,
    4. Furukawa H,
    5. Wollmuth LP,
    6. Gibb AJ,
    7. Traynelis SF
    (2018) Structure, function, and allosteric modulation of NMDA receptors. J Gen Physiol 150:1081–1105. doi:10.1085/jgp.201812032 pmid:30037851
    OpenUrlAbstract/FREE Full Text
  24. ↵
    1. Herman MA,
    2. Jahr CE
    (2007) Extracellular glutamate concentration in hippocampal slice. J Neurosci 27:9736–9741. doi:10.1523/JNEUROSCI.3009-07.2007 pmid:17804634
    OpenUrlAbstract/FREE Full Text
  25. ↵
    1. Heuer K,
    2. Toro R
    (2019) Role of mechanical morphogenesis in the development and evolution of the neocortex. Phys Life Rev 31:233–239. doi:10.1016/j.plrev.2019.01.012 pmid:30738760
    OpenUrlCrossRefPubMed
  26. ↵
    1. Hill DK
    (1950) The volume change resulting from stimulation of a giant nerve fibre. J Physiol 111:304–327. doi:10.1113/jphysiol.1950.sp004481 pmid:14795441
    OpenUrlCrossRefPubMed
  27. ↵
    1. Iacobucci GJ,
    2. Popescu GK
    (2017) NMDA receptors: linking physiological output to biophysical operation. Nat Rev Neurosci 18:236–249. doi:10.1038/nrn.2017.24 pmid:28303017
    OpenUrlCrossRefPubMed
  28. ↵
    1. Itzstein C,
    2. Cheynel H,
    3. Burt-Pichat B,
    4. Merle B,
    5. Espinosa L,
    6. Delmas PD,
    7. Chenu C
    (2001) Molecular identification of NMDA glutamate receptors expressed in bone cells. J Cell Biochem 82:134–144. doi:10.1002/jcb.1114 pmid:11400170
    OpenUrlCrossRefPubMed
  29. ↵
    1. Johnson LR,
    2. Battle AR,
    3. Martinac B
    (2019) Remembering mechanosensitivity of NMDA receptors. Front Cell Neurosci 13:533. doi:10.3389/fncel.2019.00533 pmid:31866826
    OpenUrlCrossRefPubMed
  30. ↵
    1. Kefauver JM,
    2. Ward AB,
    3. Patapoutian A
    (2020) Discoveries in structure and physiology of mechanically activated ion channels. Nature 587:567–576. doi:10.1038/s41586-020-2933-1 pmid:33239794
    OpenUrlCrossRefPubMed
  31. ↵
    1. Khatri A,
    2. Burger PB,
    3. Swanger SA,
    4. Hansen KB,
    5. Zimmerman S,
    6. Karakas E,
    7. Liotta DC,
    8. Furukawa H,
    9. Snyder JP,
    10. Traynelis SF
    (2014) Structural determinants and mechanism of action of a GluN2C-selective NMDA receptor positive allosteric modulator. Mol Pharmacol 86:548–560. doi:10.1124/mol.114.094516 pmid:25205677
    OpenUrlAbstract/FREE Full Text
  32. ↵
    1. Kim GH,
    2. Kosterin P,
    3. Obaid AL,
    4. Salzberg BM
    (2007) A mechanical spike accompanies the action potential in mammalian nerve terminals. Biophys J 92:3122–3129. doi:10.1529/biophysj.106.103754 pmid:17307820
    OpenUrlCrossRefPubMed
  33. ↵
    1. Kim SE,
    2. Coste B,
    3. Chadha A,
    4. Cook B,
    5. Patapoutian A
    (2012) The role of Drosophila piezo in mechanical nociception. Nature 483:209–212. doi:10.1038/nature10801 pmid:22343891
    OpenUrlCrossRefPubMed
  34. ↵
    1. Kloda A,
    2. Lua L,
    3. Hall R,
    4. Adams DJ,
    5. Martinac B
    (2007) Liposome reconstitution and modulation of recombinant N-methyl-D-aspartate receptor channels by membrane stretch. Proc Natl Acad Sci USA 104:1540–1545. doi:10.1073/pnas.0609649104 pmid:17242368
    OpenUrlAbstract/FREE Full Text
  35. ↵
    1. Korkotian E,
    2. Segal M
    (2001) Spike-associated fast contraction of dendritic spines in cultured hippocampal neurons. Neuron 30:751–758. doi:10.1016/s0896-6273(01)00314-2 pmid:11430808
    OpenUrlCrossRefPubMed
  36. ↵
    1. Koser DE,
    2. Thompson AJ,
    3. Foster SK,
    4. Dwivedy A,
    5. Pillai EK,
    6. Sheridan GK,
    7. Svoboda H,
    8. Viana M,
    9. Costa LD,
    10. Guck J,
    11. Holt CE,
    12. Franze K
    (2016) Mechanosensing is critical for axon growth in the developing brain. Nat Neurosci 19:1592–1598. doi:10.1038/nn.4394 pmid:27643431
    OpenUrlCrossRefPubMed
  37. ↵
    1. Krupp JJ,
    2. Vissel B,
    3. Thomas CG,
    4. Heinemann SF,
    5. Westbrook GL
    (1999) Interactions of calmodulin and alpha-actinin with the NR1 subunit modulate Ca2+-dependent inactivation of NMDA receptors. J Neurosci 19:1165–1178. doi:10.1523/JNEUROSCI.19-04-01165.1999
    OpenUrlAbstract/FREE Full Text
  38. ↵
    1. Krupp JJ,
    2. Vissel B,
    3. Thomas CG,
    4. Heinemann SF,
    5. Westbrook GL
    (2002) Calcineurin acts via the C-terminus of NR2A to modulate desensitization of NMDA receptors. Neuropharmacology 42:593–602. doi:10.1016/s0028-3908(02)00031-x pmid:11985816
    OpenUrlCrossRefPubMed
  39. ↵
    1. LaPlaca MC,
    2. Thibault LE
    (1998) Dynamic mechanical deformation of neurons triggers an acute calcium response and cell injury involving the NMDA glutamate receptor. J Neurosci Res 52:220–229. doi:10.1002/(SICI)1097-4547(19980415)52:2<220::AID-JNR10>3.0.CO;2-B
    OpenUrlCrossRefPubMed
  40. ↵
    1. Le Roux AL,
    2. Quiroga X,
    3. Walani N,
    4. Arroyo M,
    5. Roca-Cusachs P
    (2019) The plasma membrane as a mechanochemical transducer. Philos Trans R Soc Lond B Biol Sci 374:20180221. doi:10.1098/rstb.2018.0221 pmid:31431176
    OpenUrlCrossRefPubMed
  41. ↵
    1. Leung JC,
    2. Travis BR,
    3. Verlander JW,
    4. Sandhu SK,
    5. Yang SG,
    6. Zea AH,
    7. Weiner ID,
    8. Silverstein DM
    (2002) Expression and developmental regulation of the NMDA receptor subunits in the kidney and cardiovascular system. Am J Physiol Regul Integr Comp Physiol 283:R964–R971. doi:10.1152/ajpregu.00629.2001 pmid:12228067
    OpenUrlCrossRefPubMed
  42. ↵
    1. Lewis CA
    (1979) Ion-concentration dependence of the reversal potential and the single channel conductance of ion channels at the frog neuromuscular junction. J Physiol 286:417–445. doi:10.1113/jphysiol.1979.sp012629 pmid:312319
    OpenUrlCrossRefPubMed
  43. ↵
    1. Maki BA,
    2. Popescu GK
    (2014) Extracellular Ca2+ ions reduce NMDA receptor conductance and gating. J Gen Physiol 144:379–392. doi:10.1085/jgp.201411244 pmid:25348411
    OpenUrlAbstract/FREE Full Text
  44. ↵
    1. Maki BA,
    2. Aman TK,
    3. Amico-Ruvio SA,
    4. Kussius CL,
    5. Popescu GK
    (2012) C-terminal domains of N-methyl-D-aspartic acid receptor modulate unitary channel conductance and gating. J Biol Chem 287:36071–36080. doi:10.1074/jbc.M112.390013 pmid:22948148
    OpenUrlAbstract/FREE Full Text
  45. ↵
    1. Maki BA,
    2. Cummings KA,
    3. Paganelli MA,
    4. Murthy SE,
    5. Popescu GK
    (2014) One-channel cell-attached patch-clamp recording. J Vis Exp 88:51629.
    OpenUrl
  46. ↵
    1. Maneshi MM,
    2. Maki B,
    3. Gnanasambandam R,
    4. Belin S,
    5. Popescu GK,
    6. Sachs F,
    7. Hua SZ
    (2017) Mechanical stress activates NMDA receptors in the absence of agonists. Sci Rep 7:39610. doi:10.1038/srep39610 pmid:28045032
    OpenUrlCrossRefPubMed
  47. ↵
    1. McBride DW Jr.,
    2. Hamill OP
    (1993) Pressure-clamp technique for measurement of the relaxation kinetics of mechanosensitive channels. Trends Neurosci 16:341–345. doi:10.1016/0166-2236(93)90089-5 pmid:7694402
    OpenUrlCrossRefPubMed
  48. ↵
    1. McBride DW Jr.,
    2. Hamill OP
    (1999) Simplified fast pressure-clamp technique for studying mechanically gated channels. Methods Enzymol 294:482–489. pmid:9916244
    OpenUrlCrossRefPubMed
  49. ↵
    1. Miller B,
    2. Sarantis M,
    3. Traynelis SF,
    4. Attwell D
    (1992) Potentiation of NMDA receptor currents by arachidonic acid. Nature 355:722–725. doi:10.1038/355722a0 pmid:1371330
    OpenUrlCrossRefPubMed
  50. ↵
    1. Misonou H,
    2. Trimmer JS
    (2005) A primary culture system for biochemical analyses of neuronal proteins. J Neurosci Methods 144:165–173. doi:10.1016/j.jneumeth.2004.11.007 pmid:15910974
    OpenUrlCrossRefPubMed
  51. ↵
    1. Moldavski A,
    2. Behr J,
    3. Bading H,
    4. Bengtson CP
    (2020) A novel method using ambient glutamate for the electrophysiological quantification of extrasynaptic NMDA receptor function in acute brain slices. J Physiol 598:633–650. doi:10.1113/JP278362 pmid:31876958
    OpenUrlCrossRefPubMed
  52. ↵
    1. Monyer H,
    2. Sprengel R,
    3. Schoepfer R,
    4. Herb A,
    5. Higuchi M,
    6. Lomeli H,
    7. Burnashev N,
    8. Sakmann B,
    9. Seeburg PH
    (1992) Heteromeric NMDA receptors: molecular and functional distinction of subtypes. Science 256:1217–1221. doi:10.1126/science.256.5060.1217 pmid:1350383
    OpenUrlAbstract/FREE Full Text
  53. ↵
    1. Mor A,
    2. Grossman Y
    (2006) Modulation of isolated N-methyl-D-aspartate receptor response under hyperbaric conditions. Eur J Neurosci 24:3453–3462. doi:10.1111/j.1460-9568.2006.05233.x pmid:17229094
    OpenUrlCrossRefPubMed
  54. ↵
    1. Mor A,
    2. Grossman Y
    (2010) The efficacy of physiological and pharmacological N-methyl-D-aspartate receptor block is greatly reduced under hyperbaric conditions. Neuroscience 169:1–7. doi:10.1016/j.neuroscience.2010.05.009 pmid:20457226
    OpenUrlCrossRefPubMed
  55. ↵
    1. Murthy SE,
    2. Dubin AE,
    3. Patapoutian A
    (2017) Piezos thrive under pressure: mechanically activated ion channels in health and disease. Nat Rev Mol Cell Biol 18:771–783. doi:10.1038/nrm.2017.92 pmid:28974772
    OpenUrlCrossRefPubMed
  56. ↵
    1. Nicolai C,
    2. Sachs F
    (2013) Solving ion channel kinetics with the QuB software. Biophys Rev Lett 08:191–211. doi:10.1142/S1793048013300053
    OpenUrlCrossRef
  57. ↵
    1. Nishikawa M,
    2. Kimura S,
    3. Akaike N
    (1994) Facilitatory effect of docosahexaenoic acid (DHA) on N-methyl-D-aspartate response in pyramidal neurones of rat cerebral cortex. J Physiol 475:83–93. doi:10.1113/jphysiol.1994.sp020051 pmid:7514666
    OpenUrlCrossRefPubMed
  58. ↵
    1. Nowak L,
    2. Bregestovski P,
    3. Ascher P,
    4. Herbet A,
    5. Prochiantz A
    (1984) Magnesium gates glutamate-activated channels in mouse central neurones. Nature 307:462–465. doi:10.1038/307462a0 pmid:6320006
    OpenUrlCrossRefPubMed
  59. ↵
    1. Paoletti P,
    2. Ascher P
    (1994) Mechanosensitivity of NMDA receptors in cultured mouse central neurons. Neuron 13:645–655. doi:10.1016/0896-6273(94)90032-9 pmid:7917295
    OpenUrlCrossRefPubMed
  60. ↵
    1. Paoletti P,
    2. Bellone C,
    3. Zhou Q
    (2013) NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nat Rev Neurosci 14:383–400. doi:10.1038/nrn3504 pmid:23686171
    OpenUrlCrossRefPubMed
  61. ↵
    1. Parnas M,
    2. Katz B,
    3. Lev S,
    4. Tzarfaty V,
    5. Dadon D,
    6. Gordon-Shaag A,
    7. Metzner H,
    8. Yaka R,
    9. Minke B
    (2009) Membrane lipid modulations remove divalent open channel block from TRP-like and NMDA channels. J Neurosci 29:2371–2383. doi:10.1523/JNEUROSCI.4280-08.2009 pmid:19244513
    OpenUrlAbstract/FREE Full Text
  62. ↵
    1. Perozo E,
    2. Kloda A,
    3. Cortes DM,
    4. Martinac B
    (2002) Physical principles underlying the transduction of bilayer deformation forces during mechanosensitive channel gating. Nat Struct Biol 9:696–703. doi:10.1038/nsb827 pmid:12172537
    OpenUrlCrossRefPubMed
  63. ↵
    1. Phillips R,
    2. Ursell T,
    3. Wiggins P,
    4. Sens P
    (2009) Emerging roles for lipids in shaping membrane-protein function. Nature 459:379–385. doi:10.1038/nature08147 pmid:19458714
    OpenUrlCrossRefPubMed
  64. ↵
    1. Popescu G
    (2012) Modes of glutamate receptor gating. J Physiol 590:73–91. doi:10.1113/jphysiol.2011.223750 pmid:22106181
    OpenUrlCrossRefPubMed
  65. ↵
    1. Popescu G,
    2. Auerbach A
    (2003) Modal gating of NMDA receptors and the shape of their synaptic response. Nat Neurosci 6:476–483. doi:10.1038/nn1044 pmid:12679783
    OpenUrlCrossRefPubMed
  66. ↵
    1. Popescu G,
    2. Robert A,
    3. Howe JR,
    4. Auerbach A
    (2004) Reaction mechanism determines NMDA receptor response to repetitive stimulation. Nature 430:790–793. doi:10.1038/nature02775 pmid:15306812
    OpenUrlCrossRefPubMed
  67. ↵
    1. Premkumar LS,
    2. Auerbach A
    (1996) Identification of a high affinity divalent cation binding site near the entrance of the NMDA receptor channel. Neuron 16:869–880. doi:10.1016/s0896-6273(00)80107-5 pmid:8608005
    OpenUrlCrossRefPubMed
  68. ↵
    1. Qin F
    (2004) Restoration of single-channel currents using the segmental k-means method based on hidden Markov modeling. Biophys J 86:1488–1501. doi:10.1016/S0006-3495(04)74217-4 pmid:14990476
    OpenUrlCrossRefPubMed
  69. ↵
    1. Seeber S,
    2. Humeny A,
    3. Herkert M,
    4. Rau T,
    5. Eschenhagen T,
    6. Becker CM
    (2004) Formation of molecular complexes by N-methyl-D-aspartate receptor subunit NR2B and ryanodine receptor 2 in neonatal rat myocard. J Biol Chem 279:21062–21068. doi:10.1074/jbc.M313009200 pmid:15010472
    OpenUrlAbstract/FREE Full Text
  70. ↵
    1. Shively S,
    2. Scher AI,
    3. Perl DP,
    4. Diaz-Arrastia R
    (2012) Dementia resulting from traumatic brain injury: what is the pathology? Arch Neurol 69:1245–1251. doi:10.1001/archneurol.2011.3747 pmid:22776913
    OpenUrlCrossRefPubMed
  71. ↵
    1. Singh P,
    2. Doshi S,
    3. Spaethling JM,
    4. Hockenberry AJ,
    5. Patel TP,
    6. Geddes-Klein DM,
    7. Lynch DR,
    8. Meaney DF
    (2012) N-methyl-D-aspartate receptor mechanosensitivity is governed by C terminus of NR2B subunit. J Biol Chem 287:4348–4359. doi:10.1074/jbc.M111.253740 pmid:22179603
    OpenUrlAbstract/FREE Full Text
  72. ↵
    1. Sloley SS,
    2. Main BS,
    3. Winston CN,
    4. Harvey AC,
    5. Kaganovich A,
    6. Korthas HT,
    7. Caccavano AP,
    8. Zapple DN,
    9. Wu JY,
    10. Partridge JG,
    11. Cookson MR,
    12. Vicini S,
    13. Burns MP
    (2021) High-frequency head impact causes chronic synaptic adaptation and long-term cognitive impairment in mice. Nat Commun 12:2613. doi:10.1038/s41467-021-22744-6 pmid:33972519
    OpenUrlCrossRefPubMed
  73. ↵
    1. Star EN,
    2. Kwiatkowski DJ,
    3. Murthy VN
    (2002) Rapid turnover of actin in dendritic spines and its regulation by activity. Nat Neurosci 5:239–246. doi:10.1038/nn811 pmid:11850630
    OpenUrlCrossRefPubMed
  74. ↵
    1. Suchyna TM,
    2. Markin VS,
    3. Sachs F
    (2009) Biophysics and structure of the patch and the gigaseal. Biophys J 97:738–747. doi:10.1016/j.bpj.2009.05.018 pmid:19651032
    OpenUrlCrossRefPubMed
  75. ↵
    1. Szabo L,
    2. Morey R,
    3. Palpant NJ,
    4. Wang PL,
    5. Afari N,
    6. Jiang C,
    7. Parast MM,
    8. Murry CE,
    9. Laurent LC,
    10. Salzman J
    (2015) Statistically based splicing detection reveals neural enrichment and tissue-specific induction of circular RNA during human fetal development. Genome Biol 16:126. doi:10.1186/s13059-015-0690-5 pmid:26076956
    OpenUrlCrossRefPubMed
  76. ↵
    1. Tong G,
    2. Shepherd D,
    3. Jahr CE
    (1995) Synaptic desensitization of NMDA receptors by calcineurin. Science 267:1510–1512. doi:10.1126/science.7878472 pmid:7878472
    OpenUrlAbstract/FREE Full Text
  77. ↵
    1. Tyler WJ
    (2012) The mechanobiology of brain function. Nat Rev Neurosci 13:867–878. doi:10.1038/nrn3383 pmid:23165263
    OpenUrlCrossRefPubMed
  78. ↵
    1. Ucar H,
    2. Watanabe S,
    3. Noguchi J,
    4. Morimoto Y,
    5. Iino Y,
    6. Yagishita S,
    7. Takahashi N,
    8. Kasai H
    (2021) Mechanical actions of dendritic-spine enlargement on presynaptic exocytosis. Nature 600:686–689. doi:10.1038/s41586-021-04125-7
    OpenUrlCrossRef
  79. ↵
    1. Vance KM,
    2. Hansen KB,
    3. Traynelis SF
    (2013) Modal gating of GluN1/GluN2D NMDA receptors. Neuropharmacology 71:184–190. doi:10.1016/j.neuropharm.2013.03.018
    OpenUrlCrossRefPubMed
  80. ↵
    1. Wadiche JI,
    2. Jahr CE
    (2001) Multivesicular release at climbing fiber-Purkinje cell synapses. Neuron 32:301–313. doi:10.1016/s0896-6273(01)00488-3 pmid:11683999
    OpenUrlCrossRefPubMed
  81. ↵
    1. Walsh CM,
    2. Bautista DM,
    3. Lumpkin EA
    (2015) Mammalian touch catches up. Curr Opin Neurobiol 34:133–139. doi:10.1016/j.conb.2015.05.003 pmid:26100741
    OpenUrlCrossRefPubMed
  82. ↵
    1. Wollmuth LP,
    2. Sakmann B
    (1998) Different mechanisms of Ca2+ transport in NMDA and Ca2+-permeable AMPA glutamate receptor channels. J Gen Physiol 112:623–636. doi:10.1085/jgp.112.5.623 pmid:9806970
    OpenUrlAbstract/FREE Full Text
  83. ↵
    1. Wyszynski M,
    2. Lin J,
    3. Rao A,
    4. Nigh E,
    5. Beggs AH,
    6. Craig AM,
    7. Sheng M
    (1997) Competitive binding of alpha-actinin and calmodulin to the NMDA receptor. Nature 385:439–442. doi:10.1038/385439a0 pmid:9009191
    OpenUrlCrossRefPubMed
  84. ↵
    1. Yan Z,
    2. Li P,
    3. Xue Y,
    4. Tian H,
    5. Zhou T,
    6. Zhang G
    (2021) Glutamate receptor, ionotropic, N-methyl D-aspartate-associated protein 1 promotes colorectal cancer cell proliferation and metastasis, and is negatively regulated by miR-296-3p. Mol Med Rep 24:700. doi:10.3892/mmr.2021.12339
    OpenUrlCrossRef
  85. ↵
    1. Zhang L,
    2. Rzigalinski BA,
    3. Ellis EF,
    4. Satin LS
    (1996) Reduction of voltage-dependent Mg2+ blockade of NMDA current in mechanically injured neurons. Science 274:1921–1923. doi:10.1126/science.274.5294.1921 pmid:8943207
    OpenUrlAbstract/FREE Full Text
  86. ↵
    1. Zhang W,
    2. Howe JR,
    3. Popescu GK
    (2008) Distinct gating modes determine the biphasic relaxation of NMDA receptor currents. Nat Neurosci 11:1373–1375. doi:10.1038/nn.2214 pmid:18953348
    OpenUrlCrossRefPubMed
Back to top

In this issue

The Journal of Neuroscience: 42 (29)
Journal of Neuroscience
Vol. 42, Issue 29
20 Jul 2022
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Masthead (PDF)
Email

Thank you for sharing this Journal of Neuroscience article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Membrane Stretch Gates NMDA Receptors
(Your Name) has forwarded a page to you from Journal of Neuroscience
(Your Name) thought you would be interested in this article in Journal of Neuroscience.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
View Full Page PDF
Citation Tools
Membrane Stretch Gates NMDA Receptors
Sophie Belin, Bruce A. Maki, James Catlin, Benjamin A. Rein, Gabriela K. Popescu
Journal of Neuroscience 20 July 2022, 42 (29) 5672-5680; DOI: 10.1523/JNEUROSCI.0350-22.2022

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Respond to this article
Request Permissions
Share
Membrane Stretch Gates NMDA Receptors
Sophie Belin, Bruce A. Maki, James Catlin, Benjamin A. Rein, Gabriela K. Popescu
Journal of Neuroscience 20 July 2022, 42 (29) 5672-5680; DOI: 10.1523/JNEUROSCI.0350-22.2022
Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Introduction
    • Materials and Methods
    • Results
    • Discussion
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Keywords

  • ionotropic glutamate receptors
  • mechanotransduction
  • NMDARs
  • patch-clamp
  • signal transduction
  • single-molecule

Responses to this article

Respond to this article

Jump to comment:

No eLetters have been published for this article.

Related Articles

Cited By...

More in this TOC Section

Research Articles

  • Chemogenetic disruption of monkey perirhinal neurons projecting to rostromedial caudate impairs associative learning
  • Specializations in amygdalar and hippocampal innervation of the primate nucleus accumbens shell
  • The Inattentional Rhythm in Audition
Show more Research Articles

Cellular/Molecular

  • Sex differences in histamine regulation of striatal dopamine
  • Time-Dependent Actions of Corticosterone on Infralimbic Cortex Pyramidal Neurons of Adult Male Rats
  • Intracellular Spermine Is a Key Player in GSG1L’s Regulation of Calcium-Permeable AMPAR Channel Conductance and Recovery from Desensitization
Show more Cellular/Molecular
  • Home
  • Alerts
  • Follow SFN on BlueSky
  • Visit Society for Neuroscience on Facebook
  • Follow Society for Neuroscience on Twitter
  • Follow Society for Neuroscience on LinkedIn
  • Visit Society for Neuroscience on Youtube
  • Follow our RSS feeds

Content

  • Early Release
  • Current Issue
  • Issue Archive
  • Collections

Information

  • For Authors
  • For Advertisers
  • For the Media
  • For Subscribers

About

  • About the Journal
  • Editorial Board
  • Privacy Notice
  • Contact
  • Accessibility
(JNeurosci logo)
(SfN logo)

Copyright © 2025 by the Society for Neuroscience.
JNeurosci Online ISSN: 1529-2401

The ideas and opinions expressed in JNeurosci do not necessarily reflect those of SfN or the JNeurosci Editorial Board. Publication of an advertisement or other product mention in JNeurosci should not be construed as an endorsement of the manufacturer’s claims. SfN does not assume any responsibility for any injury and/or damage to persons or property arising from or related to any use of any material contained in JNeurosci.