Abstract
The Ca2+-activated TRPM5 channel plays essential roles in taste perception and insulin secretion. However, the mechanism by which Ca2+ regulates TRPM5 activity remains elusive. We report cryo-EM structures of the zebrafish TRPM5 in an apo closed state, a Ca2+-bound open state, and an antagonist-bound inhibited state. We defined two novel ligand binding sites: a Ca2+ site (CaICD) in the intracellular domain (ICD), and an antagonist site in the transmembrane domain (TMD). The CaICD site is unique to TRPM5 and has two roles: modulating the voltage dependence and promoting Ca2+ binding to the CaTMD site that is conserved throughout TRPM channels. Conformational changes initialized from both Ca2+ sites cooperatively open the ion-conducting pore. The antagonist NDNA wedges into the space between the S1-S4 domain and pore domain, stabilizing the TMD in an apo-like closed state. Our results lay the foundation for understanding the voltage-dependent TRPM channels and developing new therapeutic agents.
Introduction
Taste perception is one of the fundamental chemosensations in mammals, detecting the availability and the quality of food by converting the signal from tastants into electrical signals that the brain can interpret. Highly expressed in type II taste bud cells, the TRPM5 channel has been considered a key player in sensing sweet, umami, and bitter stimuli1,2. TRPM5 is activated upon the elevation of cytoplasmic Ca2+ concentration that is caused by the binding of tastants to the taste receptors3. Activated TRPM5 then depolarizes the membrane and causes the CALHM1 channel to release the neurotransmitter ATP, which binds to the downstream P2X receptors that trigger the action potential of gustatory neurons, thus conveying taste information to the brain3,4. TRPM5 also participates in other physiological processes in diverse cell types in a similar manner. For example, it is involved in insulin secretion by pancreatic beta cells5,6 and in the immune response of tuft cells7. Thus, TRPM5 has broad implications for metabolic syndromes and immune disorders, and it is a potential drug target for the treatment of metabolic disorders such as obesity and type 2 diabetes8.
The transient receptor potential superfamily, melastatin subfamily (TRPM) consists of eight family members (TRPM1-8) that have diverse functional properties9. While most TRP family members are nonselective cation channels that are permeable to Na+ and Ca2+, TRPM5 and TRPM4 are the only two that are monovalent cation–selective and impermeable to Ca2+ (Ref 10–12). Moreover, TRPM5 and TRPM4 share substantial sequence similarity, and both are activated by intracellular Ca2+ in a voltage- and temperature-dependent manner; therefore, they have been classified as close homologs13. However, their biophysical properties vary with respect to Ca2+ sensitivity and ligand specificity, and the molecular basis for these differences is unknown14. For instance, TRPM5 is roughly 20-fold more sensitive to Ca2+ than TRPM414. TRPM4 is inhibited by ATP, and its voltage dependence is modulated by decavanadate, while TRPM5 is insensitive to these ligands14,15. By contrast, TRPM5, but not TRPM4, is modulated by the sweetener stevioside16. TRPM5 has also been an attractive pharmaceutical target for treating metabolic syndromes. For example, a family of small-molecule TRPM5 inhibitors, including N’-(3,4-dimethoxybenzylidene)-2-(naphthalen-1-yl)acetohydrazide (NDNA), has been invented. Of note, these inhibitors have been claimed to treat type II diabetes by enhancing insulin release and GLP-1 release17, which is contrary to the result of TRPM5 KO mice experiments6,18. Recently, structures of the voltage-dependent TRPM4 and TRPM8 channels have been reported19–24, but none have been captured in an active open state, preventing a detailed understanding of their gating mechanisms. To understand the molecular mechanisms by which Ca2+ activates and antagonist NDNA inhibits TRPM5, we performed electrophysiological and structural studies on TRPM5.
Results
Channel function, overall structures, and the ion-conducting pore.
The zebrafish TRPM5 is highly sensitive to Ca2+, and 1 μM Ca2+ elicited a robust outward rectifying current in an excised inside-out patch (Fig. 1a, Supplementary Fig. 6a). Interestingly, the outward rectification of the TRPM5 currents apparently depends on the Ca2+ concentration. That is, at low Ca2+ concentration, the activation of TRPM5 requires membrane depolarization, whereas, at high Ca2+ concentration, TRPM5 becomes markedly less voltage-dependent, having a nearly linear current–voltage relation (Fig. 1a–c). This unique property is conserved in human TRPM5 but not in its closest homolog TRPM4 (Extended Data Fig. 1a, g, j, p, s–u)20. Moreover, this change in voltage dependence was previously observed with native TRPM5 currents in taste cells25. Our data imply that besides being an agonist, Ca2+ may also serve as a modulator to tune the voltage dependence of TRPM5.
Figure 1: Zebrafish TRPM5 current.
a, Representative traces of Ca2+-activated currents from membrane patches excised from tsA201 cells overexpressing zebrafish TRPM5 recorded in the inside-out patch-clamp configuration. Ca2+ of 1, 30, 100, and 1000 μM were superfused and voltage clamps were imposed from +200 mV to −200 mV with a final tail current pulse at −140 mV. Background currents were subtracted by interleaved measurements with a calcium-free solution. b, c, Current amplitudes (at 50 ms) of experiments in (a) were plotted as a function of clamp voltage for unnormalized (b) and +200 mV normalized current values (c). Replicates consists of: 1 μM Ca2+ [n = 12 patches], 30 μM [n = 6], 100 μM [n = 4], 1000 μM [n = 3] from 12 transfections. Error bars represent SEM. Tail current analysis was also performed (see Supplementary Figure 6). Source data for b and c are available online.
We solved the structures of zebrafish TRPM5 in both glyco-diosgenin (GDN) and lipid nanodiscs (Extended Data Figs. 2–3; Table 1; Supplementary Figs. 1–4); the structures were virtually indistinguishable (Supplementary Figs. 4g). We focused on the TRPM5 in GDN because it produced cryo-EM maps of higher resolution. The apo, Ca2+-bound, and (Ca2+, NDNA)-bound structures were determined in the presence of 1 mM EDTA, 5 mM Ca2+, and 0.5 mM NDNA plus 5 mM Ca2+ and had estimated resolutions of 2.9, 2.3, and 2.8 Å, respectively (Extended Data Figs. 2b, 3b, 3e, 3i and Table 1). The maps were of excellent quality (Fig. 2a; Extended Data Fig. 4), which allowed us to de novo model nearly the entire protein (Fig. 2b), to identify two bound Ca2+ ions and an NDNA molecule in each subunit (Fig. 2c), two water molecules coordinating the Ca2+ in the transmembrane domain (TMD) and two water molecule in the pore loop region (Extended Data Fig. 5a–c), and to unambiguously define the channel gate and the selectivity filter (Extended Data Fig. 5c–g).
Table 1.
Cryo-EM data collection, refinement and validation statistics
| Apo–TRPM5 (EMD-23740, PDB 7MBP) | Ca2+–TRPM5 (EMD-23741, PDB 7MBQ) | Apo–TRPM5 (nanodisc) (EMD-23742) | Ca2+–TRPM5 (nanodisc) (EMD-23743) | Apo–TRPM5(E337A) (EMD-23746, PDB 7MBT) | Ca2+–TRPM5(E337A) (EMD-23747, PDB 7MBU) | (Ca2+, NDNA)–TRPM5 (EMD-23748, PDB 7MBV) | Apo–TRPM5(6μM Ca2+) (EMD-23744, PDB 7MBR) | Ca2+–TRPM5(6μM Ca2+) (EMD-23745, PDB 7MBS) | |
|---|---|---|---|---|---|---|---|---|---|
|
| |||||||||
| Data collection and processing | |||||||||
| Magnification | 130,000 | 105,000 | 130,000 | 130,000 | 105,000 | 105,000 | 105,000 | 130,000 | 130,000 |
| Voltage (kV) | 300 | 300 | 300 | 300 | 300 | 300 | 300 | 300 | 300 |
| Electron exposure (e−/Å2) | 49.6 | 47 | 49.6 | 49.6 | 47 | 47 | 47 | 49.6 | 49.6 |
| Defocus range (μm) | −0.9 – −1.9 | −0.9 – −1.9 | −0.9 – −1.9 | −0.9 – −1.9 | −0.9 – −1.9 | −0.9 – −1.9 | −0.9 – −1.9 | −0.9 – −1.9 | 1–2 |
| Pixel size (Å) | 1.042 | 0.826 | 1.042 | 1.042 | 0.826 | 0.826 | 0.826 | 1.042 | 1.00 |
| Symmetry imposed | C4 | C4 | C4 | C4 | C1 for single subunit; C4 for tetramer | C1 for single subunit; C4 for tetramer | C4 | C4 | C4 |
| Initial particle images (no.) | 750k | ~2,300k | ~203k | ~794k | 1,600k | 1,600k | ~1,400k | ~637k | ~637k |
| Final particle images (no.) | 84k | 292k | 13k | 44k | 208k | 139k | 109k | 32k | 23k |
| Map resolution (Å) | 2.84 | 2.34 | 3.6 | 3.06 | 2.92 | 2.97 | 2.83 | 3.51 | 3.47 |
| FSC threshold | 0.143 | 0.143 | 0.143 | 0.143 | 0.143 | 0.143 | 0.143 | 0.143 | 0.143 |
| Map resolution range (Å) | 2.84–246.2 | 2.34–246.2 | 3.6–246.2 | 3.06–246.2 | 2.92–246.2 | 2.97–246.2 | 2.83–246.2 | 3.51–246.2 | 3.47–246.2 |
| Refinement | |||||||||
| Initial model used (PDB code) | de novo | de novo | de novo | de novo | de novo | 7MBP | 7MBQ | ||
| Model resolution (Å) | 3.14 | 2.51 | 3.24 | 3.22 | 3.28 | 4.14 | 4.11 | ||
| FSC threshold | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 | ||
| Model resolution range (Å) | 2.84–246.2 | 2.34–246.2 | 2.92–246.2 | 2.97–246.2 | 2.83–246.2 | 3.51–246.2 | 3.47–246.2 | ||
| Map sharpening B factor (Å2) | −70.22 | −43.99 | −70.89 | −70.72 | −75.36 | −117.99 | −104.77 | ||
| Model composition | |||||||||
| Nonhydrogen atoms | 30,710 | 30,618 | 30,762 | 30,782 | 30,806 | 30,710 | 30,618 | ||
| Protein residues | 3,984 | 3,984 | 39,84 | 3,984 | 3,984 | 3,984 | 3,984 | ||
| Ligands | 12 | 20 | 12 | 16 | 24 | 12 | 20 | ||
| B factors (Å2) | |||||||||
| Protein | NA | NA | NA | NA | NA | NA | NA | ||
| Ligand | NA | NA | NA | NA | NA | NA | NA | ||
| R.m.s. deviations | |||||||||
| Bond lengths (Å) | 0.186 | 0.19 | 0.189 | 0.190 | 0.286 | 0.186 | 0.19 | ||
| Bond angles (°) | 1.039 | 1.053 | 1.041 | 1.080 | 1.46 | 1.039 | 1.053 | ||
| Validation | |||||||||
| MolProbity score | 1.51 | 1.61 | 1.72 | 1.85 | 1.45 | 1.51 | 1.61 | ||
| Clashscore | 6.32 | 6.23 | 9.33 | 11.6 | 6.90 | 6.32 | 6.23 | ||
| Poor rotamers (%) | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | ||
| Ramachandran plot | |||||||||
| Favored (%) | 97.06 | 96.10 | 96.55 | 95.99 | 97.69 | 97.06 | 96.10 | ||
| Allowed (%) | 2.94 | 3.90 | 3.45 | 4.01 | 2.31 | 2.94 | 3.90 | ||
| Disallowed (%) | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | 0.00 | ||
Figure 2: The overall architecture.
a, The cryo-EM map of Ca2+–TRPM5 viewed parallel to the membrane. One subunit is highlighted in red. The unsharpened reconstruction is shown as transparent envelope in. b, The atomic model of Ca2+–TRPM5 condition in the same view as (a). The Ca2+ ions are shown as green spheres. c, The atomic model of (Ca2+, NDNA)–TRPM5 viewed parallel to the membrane. One subunit is highlight in cyan. The NDNA molecule is colored in orange. d–f, The ion-conducting pore in apo–TRPM5 (d), Ca2+–TRPM5 (e), and (Ca2+, NDNA)–TRPM5 (f) viewed parallel to the membrane. Purple, green, and red spheres define radii of >2.3, 1.2–2.3, and <1.2 Å, respectively. The pore region (shown in cartoon), residues (shown in sticks) forming the gate, and the selectivity filter in two subunits are depicted. Lower right panel: the channel gate viewed from the intracellular side; the distance between the Cα atoms of adjacent I966 residues is labeled. Upper right box: a cartoon representing two subunits of the apo state. The unoccupied/occupied Ca2+ sites are shown as unfilled/filled circles, respectively. The membrane area is shown as a gray background. g, Plot of pore radius along the pore axis.
The tetrameric TRPM5 is assembled with a TMD formed by six transmembrane helices and four characteristic intracellular melastatin homology regions (MHR1/2 and MHR3/4) (Extended Data Fig. 2d–e). Despite being the closest homolog to TRPM4, TRPM5 has a distinct monomeric structure, with the MHR1/2 domain tilting toward the TMD, resulting in a more compact tetrameric assembly and a different intersubunit interface (Extended Data Fig. 6a). Besides the conserved Ca2+ site in the TMD (CaTMD), as observed in the structures of TRPM2, TRPM4, and TRPM8, we observed a novel Ca2+ binding site (CaICD) in the intracellular cytosolic domain at the interface between MHR1/2 and MHR3/4 domains (Fig. 2b).
The most remarkable difference between the apo and Ca2+-bound structures was at the ion-conducting pore (Fig. 2d–e, g). The intracellular half of the pore, which constitutes the channel gate, is restricted by I966 in the apo structure, giving a smallest radius of 0.8 Å (Fig. 2d, g); this represents an apo (resting) closed state (apo–TRPM5). By contrast, the Ca2+-bound structure has an enlarged pore with a smallest radius of 2.7 Å (Fig. 2e, g), which allows the passage of partially dehydrated monovalent cations, thus representing an agonist-bound active open state (Ca2+–TRPM5).
The extracellular half of the pore is confined by the pore loop, which shows little conformational change during channel gating and is generally considered to be responsible for ionic selectivity (Fig. 2d–e; Extended Data Fig. 6b). Here, we identified two ordered water molecules in each subunit (Extended Data Fig. 5c, e–f). Notably, the water molecules form a tight oxygen ring along with the backbone oxygen atoms of G905, constituting the narrowest site in the pore loop region (Fig. 2d). Here, the pore radius is approximately 2.5 Å, about the Na+–O distance in a six-coordinate hydrated sodium (2.4 Å)26. We suggest this oxygen ring acts as the selectivity filter, and the four water molecules provide a favorable hydration layer for sodium ions to permeate27. A similar selectivity filter likely also exists in TRPM4, given by the conserved sequences (Supplementary Fig. 5) and structures of their pore loop (Extended Data Fig. 6b). The replacement of Q977 (equivalent to Q906 in zebrafish TRPM5) by a glutamate gives human TRPM4 a moderate permeability to Ca2+ (Ref 28). A glutamate at this position may attract Ca2+ ions by creating a binding site near the selectivity filter29. In addition, the glutamate might no longer form the same hydrogen bonding network, resulting in an altered conformation and size of the selectivity filter.
In the presence of NDNA and Ca2+, we observe a well-defined density in each subunit wedging into a cleft between the S1-S4 domain and the pore domain of the TMD (Fig. 2c). This density unambiguously hews to the shape of an NDNA molecule (Extended Data Fig. 7a–b). Of note, this binding site is known as the “vanilloid binding site” in the structures of TRPV and TRPA channels30–32, but has not been reported in members of the TRPM family. Despite Ca2+ binding to both CaTMD and CaICD, the TMD of (Ca2+, NDNA)–TRPM5 shows an apo-like conformation with a closed pore (Fig. 2c, f–g). We thus define it as an antagonist-bound inhibited state.
Two calcium binding sites.
The CaTMD site is located within the S1-S4 domain and is surrounded by four amino acids and two water molecules in an octahedral geometry (Fig. 3a). The key residues coordinating the CaTMD are absolutely conserved across the TRPM family members (Fig. 3c upper panel). Their replacement by an alanine largely abolished Ca2+-invoked currents (Extended Data Fig. 8), which indicates that binding of Ca2+ to CaTMD is indispensable for the activation of zebrafish TRPM5. The integrity of CaTMD site is also important for the activity of rat TRPM5 and other Ca2+-dependent TRPM channels33,34. We propose that the octahedral geometry of the CaTMD site is likely conserved among Ca2+-sensitive TRPM channels, in light of the high sequence conservation of the CaTMD site and the similar spatial organization of the Ca2+-coordinating residues (Fig. 3c; Extended Data Fig. 6c).
Figure 3: Ca2+ binding sites.
a, b, The CaTMD (a) and CaICD (b) sites. Ca2+ is shown as a green sphere. The coordinating residues and water molecules are shown in sticks and spheres, respectively. Polar interactions are indicated by yellow bars. c, Sequence alignment of the CaTMD (top) and CaICD site (bottom) among zebrafish TRPM5 (drM5), human TRPM5 (hsM5), human TRPM4 (hsM4), human TRPM2 (hsM2), and human TRPM8 (hsM8). The Ca2+ coordinating residues in zebrafish TRPM5 are indicated by asterisks, and conserved coordinating residues are in red. The residue numbers are according to zebrafish TRPM5 (UniProtID: S5UH55). Sequence segments are separated by vertical bars. d, Remodeling of the CaICD site upon Ca2+ binding. Apo–TRPM5 and Ca2+–TRPM5 are in blue and red, respectively. Black arrows indicate the movement of E212 and D336. The MHR1/2 domains in apo–TRPM5 and Ca2+–TRPM5 are represented by blue and red surfaces, respectively, showing the movement of MHR1/2. e, Voltage-clamped current amplitudes were measured (as performed in Fig. 1) from inside-out patches pulled from tsA cells overexpressing E337A mutant channels. The number of patches are (1 μM Ca2+ [n = 7 patches], 30 μM [n = 7], 100 μM [n = 5], 1000 μM [n = 5] from 8 transfections). Error bars represent SEM. See Extended Data Fig. 1b for representative traces. f, The superimposition of CaICD site in TRPM5 (red) and TRPM4 (yellow, PDBID: 6BQR)21, by aligning the helix α11 and its equivalent in human TRPM4 (residues 396–403). The coordinating residues are shown as sticks. Equivalent structural elements and residues in TRPM4 are labeled with a prime symbol. The orientation of helix α12 and its equivalent (α12’) in human TRPM4 (residues 372–386) are indicated by colored 3D arrows. The differences between α11 and α11’, and between E337 and E396’, are indicated by black arrows. Source data for e are available online.
The newly defined CaICD site is at the interface between the MHR1/2 and MHR3/4 domains (Fig. 2b), in a negatively charged pocket. Part of the pocket is formed by a twisted helical segment, α12, of the MHR3 domain (Fig. 3b). The unique conformation of α12 allows the side chains of D336 and E337 on one side of the twist, and backbone oxygen of D333 on the other side, to face toward each other, thus creating a binding pocket to accommodate Ca2+.
To investigate the mechanism underlying CaICD binding, we carried out structural comparisons and electrophysiological experiments. The overlay of Ca2+–TRPM5 and apo–TRPM5 structures revealed that E337, C324, and the backbone oxygen of D333 form a rigid notch surrounding the binding site and show little movement upon Ca2+ binding; by contrast, D336 and E212 act as a flexible cap to enclose the CaICD site (Fig. 3d). Specifically, E212 from MHR1/2 approaches the CaICD site by moving approximately 3 Å, thus pulling the MHR1/2 domain closer to the MHR3 domain; meanwhile, the side-chain of D336 flips approximately 90° toward the Ca2+. We propose that E337, the only negatively charged residue in the rigid notch, plays a key role in the binding of Ca2+. Indeed, replacement of E337 with an alanine (E337A) rendered TRPM5 voltage-sensitive even at high Ca2+ concentration (up to 1000 μM), distinct from the wild-type TRPM5, which becomes nearly voltage-independent at high Ca2+ concentration (Figs. 1a–c, 3e; Extended Data Fig. 1b, k, s–u; Supplementary Fig. 6a–b). Mutations of other coordinating residues to alanine only moderately altered the voltage sensitivity (Extended Data Fig. 1c–f, l–o, s–u; Supplementary Fig. 6c–f).
The CaICD site is unique for TRPM5 because the residues coordinating CaICD are conserved among TRPM5 orthologues, but not in other TRPM channels, except for TRPM4 (Fig. 3c lower panel). To understand why a site like the TRPM5 CaICD was not observed in the published TRPM4 structures19–22 despite the conserved sequence, we compared their intracellular domains (ICDs). We found that the key residues in TRPM4 are not close enough to each other to form a binding site because of two major structural differences. First, the structural element in TRPM4, which corresponds to the twisted helical segment α12 in TRPM5, is an intact α-helix, so that E396 (equivalent to E337 in TRPM5) cannot face the backbone oxygen of A392 (equivalent to D333 in TRPM5) (Fig. 3f). Second, the interface of MHR1/2 and MHR3, where the CaICD site is located, has a markedly different arrangement than in TRPM5, manifested by different angles between helices α11 (on MHR1/2) and α12 (on MHR3) (Fig. 3f).
Antagonist binding site.
The antagonist NDNA is highly potent and inhibits Ca2+- induced TRPM5 currents with an IC50 of approximately 2.4 nM (Fig. 4a–b). The molecular structure of NDNA consists of two rings, a naphtalen and a dimethoxybenzylidene, which are linked by an acetohydrazide group (Extended Data Fig. 7a). In the (Ca2+, NDNA)–TRPM5 structure, the NDNA molecule is located at the interface between the S1-S4 domain and the pore domain (S5 and S6), near the CaTMD site (Fig. 4c). The two rings of NDNA are perpendicular to each other, forming a wedge shape. The naphtalen ring forms the base of the wedge, bracing on the S3 helix; the dimethoxybenzylidene ring forms the tip that penetrates through the cleft between S4 and S5, pressing against the pore domain of the adjacent subunit (Fig. 4d, e). The interaction between NDNA and TRPM5 is further enhanced by a hydrogen bond between the acetohydrazide linker and E853 on S5, and the proximity of the same linker to W793 on S3 (Fig. 4e). Within the binding site, while most residues preserve their conformations in the apo state, the side chain of W869 flips to accommodate NDNA, forming a hydrogen bond with one of the methoxyl moieties on the dimethoxybenzylidene ring (Extended Data Fig. 7c).
Figure 4: Effect and binding site of antagonist NDNA.
a, Voltage-clamped (+200 mV to −200 mV) calcium activated whole-cell currents from tsA201 cells over-expressing zebrafish WT TRPM5 were suppressed upon super-fusion of 10 μM NDNA. b, IC50 of NDNA, 2.4 nM, was determined by plotting (I+200 mV, NDNA / I+200 mV, bath) using various NDNA concentrations (1 fM, 10 pM, 100 pM, 1 nM, 100 nM, 0.5 μM, 10 μM). Concentration is log (M). Each point represents the mean, and bars indicate SEM. Number of cells is indicated in brackets. From non-linear fitting, the Hill Slope is −0.5, and the 95% CI is 0.5 – 23 nM. c, The pore domain of (Ca2+, NDNA)–TRPM5 viewed from the extracellular side. The four bound NDNA molecule is shown in orange. Transmembrane helices surrounding a copy of NDNA is labeled. Prime symbol indicates the adjacent subunit. CaTMD is shown in green sphere. d and e, Two close-up views for the detailed interactions mediated by the NDNA. One TRPM5 subunit is colored in cyan, whereas the adjacent subunit is colored in light cyan. Polar interactions between NDNA and residues are indicated by black lines. g, h, Ratio of whole-cell current amplitudes in the presence and absence of NDNA (10 μM) for various TRPM5 mutants. Each point represents a single cell and bars denote mean value. The number of cells analyzed were: tsA (n = 3 cells, 2 transfections), WT (n = 4, 4 transfections), C796A (n = 3 cells, 3 transfections), I849A (n = 3 cells, 1 transfection), E853A (n = 5 cells, 4 transfections), I836A (n = 5 cells, 5 transfections), W793A (n = 3 cells, 2 transfections), V852A (n = 3 cells, 2 transfections), L833A (n = 6 cells, 4 transfections), W869A (n = 3 cells, 2 transfections). Source data for b, f and g are available online.
NDNA binds to a region critical for channel gating. To investigate how alterations in the binding site affect the response of TRPM5 to NDNA, we performed mutagenesis and electrophysiology studies. NDNA failed to suppress the current of mutants E853A, I836A, and W793A, but inhibited the currents of C796A and I849A (Fig. 4f–g; Extended Data Fig. 9). Interestingly, NDNA seemed to potentiate the current of mutants L833A and W793A (Fig. 4f–g; Extended Data Fig. 9). While the mechanism by which the NDNA potentiated these mutants is unclear, the specific NDNA binding mode appears to be critical, and even subtle variations could lead to opposite effects on channel function. This is reminiscent of previous studies of TRPV and TRPA1 channels in which a number of small-molecule compounds, including both agonists and antagonists, bind to a similar location as NDNA30–32,35. Mutants L833A, W869A, and V852A showed hardly any basal currents precluding evaluation of NDNA inhibition (Extended Data Fig. 9). Overall, most of the mutants altered the sensitivity of TRPM5 to NDNA, thus supporting that the binding site is critical for NDNA-mediated channel inhibition.
In the (Ca2+, NDNA)–TRPM5 structure, although both CaTMD and CaICD sites are occupied by Ca2+, we observed major differences relative to the Ca2+–TRPM5 open state. First, the ICD showed the same trend of motion relative to the apo–TRPM5 structure but to a lesser extent (Extended Data Fig. 7d). Second, the S1-S4 domain retained apo-like conformation and did not show marked conformational changes as observed in the Ca2+-bound open state, resulting in a different coordination of CaTMD, with Q771 on the S2 helix not involved in the binding (Extended Data Fig. 7e, g). Lastly, the pore domain is closed, similar to the apo state (Extended Data Fig. 7h–i; Fig. 2f–g). Together, our data suggest that NDNA inhibits Ca2+-induced TRPM5 activation in a non-competitive manner. Despite Ca2+ binding, NDNA limits the movement of the S1-S4 domain and the pore domain, thus stabilizing the TMD in an apo-like closed state.
The two roles of the CaICD site.
The alanine mutants of the key residues in the CaICD site had the same Ca2+-induced channel activation as the wild type, but they shifted the voltage dependence toward a positive membrane potential to different degrees, with E337A having the strongest phenotype (Fig. 3e; Extended Data Fig. 1; Supplementary Fig. 6). E337A remained voltage-dependent regardless of Ca2+ concentration, which was in sharp contrast to the wild type, in which current changed from voltage-dependent at low Ca2+ concentration to nearly voltage-independent at high Ca2+ concentration. We further looked into the same mutant on human TRPM5 (E351A) and observed the same phenotype (Extended Data Fig. 1g, h, p, q, s–u).
These results indicate that CaICD modulates the voltage dependence of TRPM5. To ground our interpretations of the electrophysiological data, we determined the structure of E337A in the presence of 5 mM Ca2+ at 2.9 Å resolution (Fig. 5a; Extended Data Fig. 10; Supplementary Fig. 7). As expected, the ICD showed a wild type apo-like conformation and the CaICD site was unoccupied (Fig. 5b, c). This supports the idea that the replacement of E337 by an alanine indeed impaired Ca2+ binding to the CaICD site and that the altered voltage dependence of E337A was caused by abolished Ca2+ binding to the CaICD site.
Figure 5: The structures of the CaICD-deficient mutant E337A.
a, The upper panel shows the consensus map obtained from the Ca2+–TRPM5(E337A) data. The cartoons in the lower panel represent the two conformations that have distinct occupancies at the CaTMD site, obtained by single subunit analysis of the same data: apo–TRPM5(E337A) in magenta and Ca2+–TRPM5(E337A) in cyan. The unoccupied Ca2+ sites are shown as unfilled circles; occupied Ca2+ sites are shown as green circles. The cell membrane is represented in gray. b–e, the CaICD site (b) and CaTMD site (d) in apo–TRPM5(E337A), and the CaICD site (c), and CaTMD site (e) in Ca2+–TRPM5(E337A). The cryo-EM densities are shown in black mesh. The Ca2+ density is shown as a green sphere. Unoccupied sites are indicated by a dashed gray circle. f, The superimpositions of the S1-S4 domain of the apo–TRPM5(E337A) (magenta) and the Ca2+–TRPM5(E337A) (cyan) structures. g, The superimpositions of the S1-S4 domain of the apo–TRPM5(E337A) (magenta) and the apo–TRPM5 (blue) structures. h, The superimpositions of the S1-S4 domain of the Ca2+–TRPM5(E337A) (cyan) and the Ca2+–TRPM5 (red) structures. i, Pore radius plot along the pore axis. j, k, Remodeling of the CaICD site upon Ca2+ binding in TRPM5 (j) and TRPM4 (k). Apo–TRPM5 and Ca2+–TRPM5 are in blue and red, respectively (j). Apo–TRPM4 (PDBID: 6BQR) and Ca2+–TRPM4 (PDBID: 6BQV) are in gray and yellow, respectively21. (k). The Cα atoms of key residues and the Ca2+ are shown as spheres. Shown in parentheses are the distances of the root-mean-square-deviation (RMSD) between S2 (residues 767–772 in TRPM5 and 827–832 in TRPM4) and S3 (residues 793–798 in TRPM5 and 864–869 in TRPM4), and the distances of the Cα movements of E994 in TRPM5 and E1068 in TRPM4.
Although our data analysis workflow involved a focused classification of the TMD (Supplementary Fig. 7), part of the TMD is still not unambiguously defined, and the densities for CaTMD were markedly weaker than those in the structure of Ca2+-bound wild-type TRPM5, indicating the structural heterogeneity of the TMD. We, therefore, performed structural analysis on a single subunit and obtained two conformations that had the same ICD but distinct S1-S4 domains in the TMD (Fig. 5). One had an empty CaTMD, termed apo–TRPM5(E337A) (Fig. 5b, d), and the other had an occupied CaTMD, termed Ca2+–TRPM5(E337A) (Fig. 5c, e). This suggests that an unoccupied CaICD site lowers the binding affinity of Ca2+ for the CaTMD site. The cooperativity between these two Ca2+ binding sites agrees with the observation that upon activation with 1 μM Ca2+, the current amplitudes of the E337A mutant at a clamp of +200 mV were substantially smaller (75%) than those of the wild type (Extended Data Fig. 1t).
Within the TMD, a closed ion-conducting pore was observed in the Ca2+–TRPM5(E337A) structure (Fig. 5i). This further supports the role of CaICD as a voltage-modulating site, as its absence renders TRPM5(E337A) inactive due to the lack of membrane depolarization under the conditions for structural determination (Fig. 3e). Interestingly, the cytosolic vestibule, the part underneath the channel gate in Ca2+–TRPM5(E337A), is similar to that of Ca2+–TRPM5 (Fig. 5i), suggesting that the pore in Ca2+–TRPM5(E337A) may represent an intermediate state prior to channel opening.
The cooperativity between CaTMD and CaICD sites implies that CaICD might be physiologically relevant. To test this hypothesis, we determined the structure of wild-type TRPM5 in the presence of 6 μM Ca2+, a concentration similar to the Ca2+ EC50 of TRPM5 channels excised from native taste receptor cells (8 μM at −80 mV)25. Interestingly, this condition yielded both the apo conformation and the Ca2+-bound open conformation (Extended Data Fig. 3d–g; Supplementary Fig. 3b). The ratio of protein particles belonging to the apo and open conformations, respectively, is 1.4:1. This data thus quantitatively correlates agonist-induced conformational changes to the EC50 determined by excised patch recordings. Furthermore, CaTMD and CaICD are clearly occupied in the open state, which indicates that Ca2+ at EC50 concentration binds equally well to both sites, thus supporting the physiological relevance of CaICD (Extended Data Fig. 3g).
To understand why the occupation of the CaICD site is required for a high-affinity CaTMD site in TRPM5 but not in TRPM4, we compared the conformational changes in their CaTMD sites upon binding of Ca2+. In TRPM5, the helices S2 and S3—containing the coordinating residues and the TRP helix, which are key elements in transducing signals from the ICD to TMD—undergo substantial movement (Fig. 5j). By contrast, in TRPM4, the CaTMD site and the TRP helix mostly showed only minor sidechain rearrangement (Fig. 5k). This difference suggests that the CaTMD site in TRPM4 may be primed for Ca2+ binding, whereas in TRPM5, a high-affinity CaTMD site likely requires extensive rearrangement of the S2 and S3, assisted by Ca2+ binding to the CaICD site.
Taken together, our data suggest the CaICD site is physiologically relevant and has two important roles. First, it acts as a voltage modulator that shifts the voltage dependence toward negative potential, reminiscent of the effect of decavanadate and PIP2 on TRPM415,36. Second, it promotes Ca2+ binding to the CaTMD site and facilitates channel activation.
Signal transduction from CaICD to the TMD.
The structural comparison of TRPM5 in the absence versus the presence of Ca2+ showed conformational rearrangement throughout the protein, with individual domains mostly showing rigid body movement (Fig. 6a). To understand the respective contributions of the two Ca2+ binding sites and how they cooperatively open the channel, we traced the conformational changes from the ICD and the S1-S4 domain to the ion-conducting pore.
Figure 6: The signal transduction from ICD to TMD.
a, The superimposition of apo–TRPM5 (blue) and Ca2+–TRPM5 (red) structures by aligning the coiled-coil poles in the C-terminal domain (CTD), viewed parallel to the membrane. One subunit is also shown in cartoon representation. Ca2+ is shown as green spheres. b, The superimposition of the MHR1-4 domains of apo–TRPM5 (blue) and Ca2+–TRPM5 (red) by aligning the MHR3/4 domain, viewed parallel to the membrane. The rotation of the MHR1/2 relative to the MHR3/4 domain upon Ca2+ binding is indicated. The surfaces are outlined in blue for apo–TRPM5 and filled with red for Ca2+–TRPM5. c, Superimposition of the MHR3/4 domain and the CTD rib and pole helices of apo–TRPM5 (blue) and Ca2+–TRPM5 (red) by aligning the CTD coiled-coil poles, viewed from the intracellular side. One subunit in both structures are shown in blue (apo–TRPM5) or in red (Ca2+–TRPM5). The rotation of the rib helices is indicated. d, The superimposition of the ICD–TMD interface of apo–TRPM5 (blue) and Ca2+–TRPM5 (red) by aligning the CTD coiled-coil poles (not shown), viewed from the intracellular side. The rotations of helices squareN and squareC are indicated. The green circle highlights the location of the intersubunit interface and ICD–TMD interface, as detailed in panels (e, f). e, f, The conformational rearrangement at the intersubunit interface and the ICD–TMD interface in apo–TRPM5 (e) to Ca2+–TRPM5 (f), viewed parallel to the membrane. Prime symbol indicates residues or structural elements from the adjacent subunit. Interactions are shown in yellow bars. The single headed arrows indicate the movement of the square and TRP helices. The double headed arrow indicates the angle between S6 and TRP. The positions of the intersubunit interface are shown by green circles.
Clamped by two lobes, i.e., the MHR1/2 and MHR3/4 domains, the CaICD site underwent an opening of 5.8° upon binding of Ca2+ (Fig. 6b). As a result, the rib helix, which penetrates into the interface between the MHR3/4 domain and the adjacent MHR1/2 domain, showed a clockwise rotation of 9.4° as viewed from the intracellular side (Fig. 6c). Meanwhile, just above the rib helix, four helices that form a square shape also rotated clockwise (Fig. 6d). Because these “square” helices mediate the contact between adjacent subunits, their rotation altered the intersubunit interface. Specifically, in the absence of Ca2+, the N- and C-termini of adjacent square helices are in close contact (Fig. 6e). Upon binding of Ca2+, this interface is disrupted as the adjacent termini rotate away from each other, resulting in a new interface between the N-terminus of the square helix and helix α28, where R552 and D610 form a salt bridge (Fig. 6f). Notably, the square helix in TRPM5 is broken into two short segments in the middle where E560 interacts with the N-terminus of helix α27 (Fig. 6e, f), a feature that is unique to TRPM5. By contrast, all the other TRPM channels have (or are predicted to have) a continuous square helix (Extended Data Fig. 6d–e).
We speculate that the square helix plays a role in the signal transduction from the CaICD site to the TMD because it links the remodeling of the intersubunit interface upon Ca2+ binding to the TRP helix—a key element involved in channel gating—through helices α27 and α28. Indeed, the replacement of E560 by an alanine, which presumably weakens the interaction between the square helix and helix α27, led to slower channel activation and deactivation kinetics (Extended Data Fig. 1i, r). Moreover, the R578Q polymorphism in human TRPM5 (which corresponds to position 561 on the square helix in zebrafish TRPM5) has been associated with obesity-related metabolic syndrome37.
Channel opening by synergistic action of CaICD and CaTMD.
Accompanying the conformational changes of the ICD induced by Ca2+ binding, the TRP helix is pushed toward the TMD where the CaTMD binding site is located (Fig. 7a). Here, surrounded by the TRP helix, S4-S5 linker, S2, and S3, is the region where the conformational changes induced by Ca2+ binding to the CaICD and CaTMD sites meet, and where complex remodeling occurs (Fig. 7a, b). We performed a detailed structural analysis, and propose a mechanism by which the motion of the TRP helix promotes Ca2+ binding at the CaTMD site, ultimately leading to channel opening.
Figure 7: The channel opening.
a, The superimposition of the TMD of a single subunit in apo–TRPM5 (blue) and Ca2+–TRPM5 (red) by aligning their S1-S4 domain, viewed parallel to the membrane. The center-of-mass movement of the pore domain is indicated. b, The superimposition of the pore domain in apo–TRPM5 (blue) and Ca2+–TRPM5 (red) by aligning their S1-S4 domain, viewed from the intracellular side. The pore domain of Ca2+–TRPM5 is shown in surface representation and the S5-S6 domain of one subunit from each structure is shown as a cartoon. The relative movements of helices S5 and S6 are indicated. c, d, Close-ups of the circled area in (a), viewed from the intracellular side. The remodeling of the CaTMD site from apo–TRPM5 (c) to Ca2+–TRPM5 (d). The movements of S2, S3, S4, and TRP helices are indicated by arrows. Interactions are shown in yellow bars. e, f, Close-ups of the boxed area in (a). W984 on the TRP helix switches its interaction partner from P847 and G846 in apo–TRPM5 (e) to I841 in Ca2+–TRPM5 (f). Interactions are shown in yellow bars. The movement of W984 is indicated. The contact area between the TRP helix and the S4-S5 linker is highlighted in grey. The segment between I836 and I841 turns into a 310-helix in Ca2+–TRPM5. The inset shows the view along the axis of the S4 helix. g, A cartoon scheme of the activation and inhibition mechanism of TRPM5. Conformational changes initialized from both Ca2+ sites cooperatively open the ion-conducting pore. The antagonist NDNA wedges into the space between the S1-S4 domain and pore domain, stabilizing the TMD in an apo-like closed state. The movements of individual structural elements are indicated by arrows.
In the absence of Ca2+, i.e., when the CaICD site is unoccupied, the CaTMD site is in a configuration that is difficult to access by Ca2+ because two crucial residues, E768 and D797, are locked by R834 in a triangular hydrogen bond network (Fig. 7c). This conformation is stabilized by the interaction between helices S3 and S4, with W793 and H837 stacking with each other (Fig. 7c). This explains why only a subset of particles from the CaICD site-deficient mutant E337A showed an occupied CaTMD site even at high (5 mM) Ca2+ concentration (Supplementary Fig. 7).
When the CaICD site is occupied, the TRP helix tilts toward the TMD, leading to three consequences. First, it pushes the S4 helix away from S2 and S3 (Fig. 7c), allowing E768 and D797 to be readily released from R834 to coordinate Ca2+ together with Q771, N794, and two water molecules (Fig. 7d). Second, E994 on the TRP helix approaches the CaTMD site to coordinate one of the two water molecules, thus helping the CaTMD site bind Ca2+ (Fig. 7d). Third, Y995 on the TRP helix accommodates the flipped H837 by forming a hydrogen bond, thereby assisting in the decoupling of S4 from S3 by breaking the π-stacking between H837 and W793 (Fig. 7d). The decoupling of S4 from S3 is important, because it allows a relative movement between the S4-S5 linker and W984, a residue on the TRP helix that is absolutely conserved in the TRP superfamily and is crucial for channel gating38–40. As a result, the W984 switches its interaction partners from P847 and G846 on the N-terminus of S5 to the backbone oxygen of I841on S4, forcing the last turn of the S4 helix and part of the S4-S5 linker to stretch into a 310-helix (Fig. 7e–f). The movement of W984 breaks the major interaction between the TRP helix and S5, eventually enabling the pore domain (helices S5 and S6) to relocate. Indeed, the structural comparison between the apo and open states of TRPM5 showed a movement of the pore domain by half an α-helical turn toward the extracellular side with an outward expansion, thus opening the ion-conducting pore (Fig. 7a–b).
Discussion
Our TRPM5 structures define two Ca2+ binding sites, CaICD and CaTMD, which are 70 Å apart. Comparison of the structures in the apo and open states illustrates a molecular mechanism by which Ca2+ binding to the two locations synergistically leads to a complex conformational rearrangement at the interface between the ICD and the TMD (Fig. 7g). This rearrangement eventually gives rise to the decoupling between the TRP helix and the S5 helix and the opening of the ion-conducting pore. We conclude that CaTMD functions as an orthosteric binding site for channel activation, while CaICD modulates the voltage dependence and the accessibility of CaTMD. The molecular basis by which the CaICD site modulates the voltage dependence is still unclear; that will require the identification of the voltage sensor(s) and its working mechanism.
We also defined a novel antagonist binding site in the TMD and elaborated a non-competitive inhibition mechanism by which the antagonist NDNA stabilizes the ion-conducting pore in an apo-like closed conformation (Fig. 7g). Because NDNA is a potent TRPM5-selective antagonist, our work not only will facilitate the characterization of TRPM5 currents in many physiological processes, but is also important for the ongoing development of drugs targeting TRPM5.
TRPM4 and TRPM5 share substantial sequence similarity (55.7% between human TRPM4 and TRPM5, 58.1% between zebrafish TRPM4 and TRPM5), and both depolarize the cell membrane by sensing cytosolic Ca2+, but they are involved in different physiological processes. The unique CaICD site endows TRPM5 with a complex gating and modulation mechanism by Ca2+, which may link to the physiological roles of TRPM5 in taste signaling and the Ca2+ oscillation during insulin secretion by the pancreatic beta cells5,6. We have elaborated a Ca2+-induced gating mechanism of a voltage-sensitive TRPM channel, which differs from that of the voltage-insensitive TRPM234,41–43. Our study highlights the important role of the ICD as a ligand-sensing domain in TRPM channels and lays a solid foundation for the development of novel therapeutic drugs that distinguish between TRPM4 and TRPM5.
Methods
TRPM5 expression and purification
Genes encoding full-length human and zebrafish TRPM5 (UniProtKB accession numbers Q9NZQ8, and S5UH5, respectively) were synthesized by Bio Basic and were sub-cloned into a pEG BacMam vector with an His8 tag, GFP, and a thrombin cleavage site at the N terminus44. Site-directed mutagenesis is performed by using QuikChange II Site-directed mutagenesis (Qiagen) or Q5 Site-Directed Mutagenesis (NEB) protocol and confirmed via Sanger sequencing (Eurofins). For baculovirus production, each TRPM5 ortholog in a BacMam vector is transformed into DH10Bac cells, followed by P1 and P2 baculovirus generated in Sf9 cells. P2 viruses (8%) were used to infect tsA201 cells grown in Freestyle 293 Expression Medium in suspension culture (ThermoFisher). Infected cells were incubated for an initial 12 h at 37 °C before 10 mM sodium butyrate was added. Cells were then moved to a 30 °C incubator and allowed to grow for another 60 h with vigorous shaking. At 72 h post-infection, cells were harvested by centrifugation at 5000 rpm, 4 °C for 30 min. Cell pellets were washed with buffer containing 150 mM NaCl and 20 mM Tris pH 8.0 (TBS buffer) and stored at −80 °C.
Cell pellets from 200 ml culture were thawed on ice and resuspended in TBS buffer containing 1 mM PMSF, 0.8 μM aprotinin, 2 μg mL−1 leupeptin, 2 mM pepstatin A (which are all protease inhibitors) plus 1% GDN detergent (Anatrace). Protein was extracted from the membrane by whole-cell solubilization for 1 h at 4 °C with rotation. The solubilized protein was incubated with 2 mL TALON cobalt metal-affinity resin (Takara Bio) for 1 h. The TALON resin was then washed with 20 ml TBS buffer supplemented with 0.02% GDN and 15 mM imidazole. Protein was eluted with TBS buffer supplemented with 0.02% GDN and 250 mM imidazole. The eluent was concentrated to 500 μL and further purified by size-exclusion chromatography in TBS buffer containing 0.02% GDN. Peak fractions containing TRPM5 were pooled and concentrated to 4–5 mg/mL for grid freezing.
For nanodisc reconstitution, the eluent after immobilized metal affinity chromatography was mixed with MSP2N2 and soybean lipid extract at a molar ratio of 1:1:200 (TRPM5:MSP2N2:lipid). Three rounds of Bio-Beads (BIO-RAD) incubation at 4 °C was performed to facilitate nanodisc reconstitution. The Bio-Beads were then removed, and the sample was concentrated to 500 μL using an Amicon 100 kDa concentrator (MilliporeSigma). Size-exclusion chromatography was done in TBS buffer to further purify TRPM5-nanodisc complex. Peak fractions of TRPM5-nanodisc were collected and concentrated to 5 mg/mL for freezing grid.
EM sample preparation and data acquisition
Freshly purified TRPM5 protein in GDN detergent was mixed with 1 mM EDTA (apo–TRPM5 and apo–TRPM5(E337A)), 5 mM Ca2+ (Ca2+–TRPM5 and Ca2+–TRPM5(E337A)) or 6 μM Ca2+ before grid preparation. For TRPM5-nanodisc sample, we added 0.05 mM digitonin to improve particle distribution on the grid. The apo–TRPM5(nanodisc) condition contains 1 mM EDTA, and the Ca2+–TRPM5(nanodisc) contains 1 mM Ca2+ and 0.5 mM steviol (Sigma). After mixing with the designated additives, a 2.5 μL aliquot of the sample was applied to a glow-discharged Quantifoil holey carbon grid (gold, 1.2/1.3 μm size/hole space, 300 mesh or gold, 2/1 μm size/hole space, 300 mesh), blotted for 1.5 s at 100% humidity using a Vitrobot Mark III, and then plunge-frozen in liquid ethane cooled by liquid nitrogen. The grids were loaded into a FEI Titan Krios transmission electron microscope operating at 300 kV with a nominal magnification of 130,000× and an energy filter (20 eV slit width). The apo–TRPM5, Ca2+–TRPM5(6 μM)(GDN), apo–TRPM5(nanodisc), and Ca2+–TRPM5(nanodisc) dataset was recorded by a Gatan K2 Summit direct electron detector in super-resolution mode with a binned pixel size of 0.521 Å. Each K2 movie was dose-fractionated to 40 frames for 8 s with a total dose of 49.6 e−/Å2. The Ca2+–TRPM5, Ca2+–TRPM5(E337A), NDNA/Ca2+–TRPM5 datasets were collected by a K3 direct electron detector in super-resolution mode with a binned pixel size of 0.413 Å (K3). Each K3 movie was dose-fractionated to 75 frames for 1.5 s with a total dose of 47 e−/Å2. The automated image acquisition was facilitated using SerialEM45. The nominal defocus range was set from −0.9 μm to −1.9 μm.
Cryo-EM data analysis procedure
The detailed workflow of data processing procedure is summarized in Supplementary Fig. 2–4. In general, the raw super-resolution tif movie files for each dataset were motion-corrected and 2x binned using MotionCor2 v1.1.0 or RELION 3.0 or RELION3.1 (Ref 46,47). The per-micrograph defocus values were estimated using Gctf 1.06 or ctffind 4.1.10 (Ref 48,49). Particle picking was performed using gautomatch v0.56 (https://www2.mrc-lmb.cam.ac.uk/research/locally-developed-software/zhang-software/) or topaz v0.2.4 (Ref 50). Junk particles were removed by 2D classification and heterogeneous refinement using CryoSPARC (v0.65 or v2.0.9 or v3.0.0)51. Selected good particles were then used to generate an initial 3D model by ab initio reconstruction followed by homogeneous refinement with C4 symmetry51. Multiple rounds of CTF refinement and Bayesian polishing were performed in relion to further improve the map resolution52.
At this stage, conformational heterogeneity was observed in the transmembrane domain (TMD) of the consensus refinement, indicating significant flexibility is present for TRPM5, especially in the extracellular pore loop area. To further improve the map quality, we performed focused classification by subtracting TMD signals from the particles53. After TMD-focused classification, we focused on the map with the highest nominal resolution and well-defined extracellular region for atomic model building.
For the Ca2+–TRPM5(E337A) dataset, conformational heterogeneity is still present in the transmembrane domain after TMD-focused classification. To overcome this issue, we performed symmetry expansion to the best particle set obtained from the TMD-focused classification and subtracted the single-subunit signals. Focused classification was conducted at the single-subunit level followed by 3D refinement. Two distinct conformations of the single subunit are identified for the Ca2+–TRPM5(E337A) dataset. The two conformation differ by the Ca2+ occupancy in the transmembrane domain, i.e., apo–TRPM5(E337A) and Ca2+–TRPM5(E337A). The single subunit maps were used for model building. To obtain tetrameric map for apo–TRPM5(E337A) and Ca2+–TRPM5(E337A), we further identified the homotetrameric TRPM5 particles that were solely composed of particles from each of the single subunit class. Although homo-tetrameric particles obtained after this procedure were very less, the refinement map still allowed us to generate a tetrameric model based on the single subunit map (see the model building section).
For the (Ca2+, NDNA)–TRPM5 dataset, conformational heterogeneity is observed in the ICD after TMD-focused classification. We then performed symmetry expansion (C4) and subtracted the ICD for each single subunit of TRPM5 particles. Subsequent 3D classification allowed us to obtain a homogeneous set of single TRPM5 subunit. We then identified homo-tetramer of TRPM5 that are consists of the homogeneous TRPM5 single subunit and refined the structure. This allowed us to obtain a map with better defined ICD to assist model building, despite with slightly worse nominal resolution compared to the consensus refinement before ICD classification.
For all dataset, the Gold standard Fourier shell correlation (FSC) 0.143 criteria were used to provide the map resolution estimate54. The cryo-EM maps were visualized using UCSF ChimeraX55. PDB structures are visualized using UCSF ChimeraX or PyMOL55,56.
Model building
The atomic model for apo–TRPM5 was built into the cryo-EM density manually using Coot v0.89 and subjected to real-space refinement in Phenix57,58. The apo–TRPM5 model contained residues 16–429, 446–473, 489–653, 698–1020, and 1027–1092. One GDN molecule lacking one of the two maltose groups (GDP), and one diosgenin molecule (DIO) were modeled into the lipid- or detergent-like densities for each chain. The geometrical restraints for DIO, GDP and NDNA were generated using the Grade Web Server (http://grade.globalphasing.org). Glycosylation at N921 was modeled as N-acetyl-beta-D-glucosamine (NAG). The Ca2+–TRPM5 open models were built by first docking the apo-TRPM5 closed model into the corresponding cryo-EM map density and adjusted manually in Coot. Two Ca2+ atoms were added to the TMD and ICD Ca2+ binding sites of the Ca2+–TRPM5 model. The Ca2+–TRPM5 open model was of sufficient resolution to allow us further place two ordered water molecules in the TMD Ca2+ binding site, and two water molecules in the selectivity filter for each chain. The apo–TRPM5(E337A) and Ca2+–TRPM5(E337A) model were first built based on the cryo-EM maps of the single-subunit refinement result. The single-subunit model was then rigid-body-fitted into the tetramer cryo-EM maps reconstructed from homo-tetrameric TRPM5 particles. We did not build atomic model for Apo–TRPM5(nanodisc) and Ca2+–TRPM5(nanodisc) dataset because these maps are identical to the corresponding maps from the GDN detergent conditions (Extended Data Fig. 2j). It is worth mentioning that although we included 0.5 mM steviol when preparing the Ca2+–TRPM5(nanodisc) grid, we are not able to identify density that corresponds to the steviol molecule.
Electrophysiology
In the inside-out patch clamp configuration, voltage-clamped membrane currents were measured from tsA201 cells overexpressing plasmids encoding N-terminal GFP tagged WT and mutant TRPM5 channels from zebrafish and human. Following 1 d post-transfection with Lipofectamine 2000, cells were trypsinized and replated onto poly-L-lysine-coated (Sigma) glass coverslips. After cell adherence, the coverslips were transferred to a low-volume recording chamber with a pH 7.4 bath solution containing (in mM) 150 NaCl, 3 KCl, 10 HEPES, 2 CaCl2, 1 MgCl2, and 12 mannitol. Cells with fluorescence at the plasma membrane were patched with pipettes containing a pH 7.4 solution of (in mM) 150 NaCl, 10 HEPES, and 5 EGTA. Upon tight-seal formation, the bath solution was superfused with the calcium-free EGTA solution. Following excision, patches were exposed using a manifold to super-fused bath solutions containing various free calcium concentrations. For preparing 1, 30, 100, or 1000 μM of free calcium, 4.46, 5.01, 5.1, or 6 mM CaCl2 was added to a pH 7.4 solution of 150 mM NaCl, 10 mM HEPES, 5 mM EGTA. Free calcium concentrations were calculated with https://somapp.ucdmc.ucdavis.edu/pharmacology/bers/maxchelator/CaEGTA-TS.htm. At room temperature (21–23 °C), patches from a holding voltage of 0 mV were clamped (Clampex 11.0.3, Multiclamp 700 B) using 50-ms steps from +200 mV to −200 mV (intracellular side relative to extracellular) with a final tail pulse at −140 mV. Electrical signals were digitized at 10 kHz and filtered at 2 kHz. Typically, measurements of TRPM5 activation by individual bath solutions containing various calcium concentration were interleaved with measurements where the bath solution was superfused with calcium-free EGTA solution. Using offline analysis (ClampFit 11.0.3) the currents in the absence of calcium were then subtracted from currents measured in the presence of calcium to acquire specific calcium-activated currents. Current amplitudes were measured at the end of the pulse. For normalizing current, the clamp at + 200 mV was chosen. Whole-cell measurements were performed in tsA201 cells following 1 d transfection of zebrafish TRPM5 CaTMD mutant channels. Patch pipettes were filled with a 1 μM free calcium concentration solution (pH 7.4) composed of (in mM): 150 NaCl, 1 MgCl2, 10 Hepes, 5 EGTA, 4.45 CaCl2. The bath solution (pH 7.4) contained (in mM) 150 NaCl, 3 KCl, 10 Hepes, 2 CaCl2, 1 MgCl2, and 12 mannitol. Voltage clamps (50 ms steps from +200 to −200 mV) were imposed approximately 1 minute after the whole cell configuration was acquired. Analysis was performed with GraphPad.
The NDNA patch clamp experiments (Fig 4; Extended Data Fig. 9) were performed in the same way as experiments for CaTMD mutants (Extended Data Fig. 8). For determining the IC50 of NDNA, whole cell current analysis was performed where TRPM5 currents were evoked with 1 μM calcium in the patch pipette (as described above). Upon whole cell acquisition, currents were first measured in bath solution and then re-measured 30–60 s following super-fusion of bath solution containing various NDNA concentrations (1 fM, 10 pM, 100 pM, 1 nM, 100 nM, 0.5 μM, 10 μM). NDNA was stored at 50 mM (DMSO) and serially diluted using bath solution. For each cell measured, only one concentration of NDNA was tested. Inhibition kinetics was monitored using a step protocol (+100 mV) and typically complete (steady-state current) within a minute. Inhibited current was plotted as a function of NDNA concentration and fitted using Prism software (inhibitor versus response, variable slope equation).
Preparation of N’-(3,4-dimethoxybenzylidene)-2-(naphthalen-1-yl)acetohydrazide (compound NDNA)
N’-(3,4-dimethoxybenzylidene)-2-(naphthalen-1-yl)acetohydrazide (NDNA) was synthesized according to the US Patent US8193168 (Ref17) (Supplementary Fig. 8a). Briefly, A solution of commercial available ethyl 2-(naphthalen-1-yl)acetate (compound 1) (20 g, 93.34 mmol, 1 eq), NH2NH2.H2O (9.54 g, 186.69 mmol, 9.26 mL, 2 eq) in EtOH (100 mL) was stirred at 80°C for 16 h. TLC (Petroleum ether/Ethyl acetate = 2/1) showed that most of the compound 1 (Rf = 0.5) was consumed and a new spot (Rf = 0.05) was given. The reaction mixture was concentrated under vacuum to give white solid. The white solid was triturated with Petroleum ether/Ethyl acetate = 4:1 (100 mL) for 10 min. The mixture was filtered, and the filter cake was dried under vacuum to give the tittle compound 2 (12.8 g, 59.58 mmol, 63.8% yield, 93.2% purity) as a white solid. 1H NMR (400 MHz, DMSO-d6, see Supplementary Fig. 8b upper panel) δ ppm 9.34 (s, 1H), 8.13 (d, J = 8.4 Hz, 1H), 7.92 (d, J = 7.6 Hz, 1H), 7.81 (d, J = 5.6 Hz, 1H), 7.54–7.51 (m, 2H), 7.45–7.44 (m, 2H), 4.25 (s, 2H), 3.84 (s, 2H). LCMS 0–60% ACN–H2O, ESI + APCI: Rt = 0.767 min, m/z = 201.1 (M+H)+.
A solution of 2-(naphthalen-1-yl)acetohydrazide (compound 2) (10.8 g, 50.27 mmol, 1 eq) and 3,4-dimethoxybenzaldehyde (compound 3) (8.35 g, 50.27 mmol, 1 eq) in EtOH (50 mL) was stirred at 80°C for 30 min. A white solid separated out. TLC (Petroleum ether/Ethyl acetate = 1/1) showed that compound 2 (Rf = 0.1) was consumed and a major spot (Rf = 0.3) formed. The reaction mixture was cooled to 20°C. 100 mL EtOH was added to above solution and stirred for 10 min. The mixture was filtered and the filter cake was dried under vacuum to give NDNA (9.20 g, 26.33 mmol, 52.3% yield, 99.7% purity) as a white solid. 1H NMR (400 MHz, DMSO-d6, see Supplementary Fig. 8b lower panel) δ ppm 11.06–11.38 (m, 1H), 8.18 (m, 1.4H), 7.96 (m, 1.6H), 7.93 (m, 1H), 7.55–7.48 (m, 4H), 7.31–7.30 (m, 1H), 7.18 (m, 1H), 7.01–6.99 (m, 1H), 4.53–4.23 (m, 2H), 3.80–3.65 (m, 6H). LCMS 5–95% ACN–H2O, ESI + APCI Rt = 0.826 min, m/z = 349.0 (M+H)+.
Data availability
The cryo-EM density map and coordinates and atomic models were deposited in the EMDB (Electron Microscopy Data Bank) and the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCS-PDB), respectively, under the accession numbers EMD-23740 and PDB 7MBP (apo–TRPM5), EMD-23741 and PDB 7MBQ (Ca2+–TRPM5), EMD-23742 (apo–TRPM5(nanodisc)), EMD-23743 (Ca2+–TRPM5(nanodisc)), EMD-23744 and PDB 7MBR (apo–TRPM5(6μM Ca2+)), EMD-23745 and PDB 7MBS (Ca2+–TRPM5(6μM Ca2+)), EMD-23746 and PDB 7MBT (apo–TRPM5(E337A) single subunit and tetramer), EMD-23747 and PDB 7MBU (Ca2+–TRPM5(E337A) single subunit and tetramer), and EMD-23748 and PDB 7MBV ((Ca2+, NDNA)–TRPM5). Source data are provided with this paper.
Extended Data
Extended Data Figure 1: Patch-clamp analysis of CaICD mutants.
Representative traces of inside-out voltage-clamp measurements (+200 mV to −200 mV) from tsA201 cells overexpressing human TRPM4 (hsTRPM4), human TRPM5 (hsTRPM5), and zebrafish TRPM5 (drTRPM5) channels. Patches were stimulated with 1, 30, 100, or 1000 μM Ca2+. The number of patches analyzed: a, hsTRPM4(WT): 1 μM Ca2+ [4], 100 μM [4], 1000 μM [4] from 3 transfections; b, drTRPM5(E337A): 1 μM Ca2+ [6], 30 μM [6], 100 μM [4], 1000 μM [4] from 8 transfections; c, drTRPM5(C324A): 1 μM Ca2+ [6], 30 μM [6], 100 μM [6], 1000 μM [6] from 4 transfections; d, drTRPM5(D333A): 1 μM Ca2+ [5], 30 μM [4], 100 μM [5], 1000 μM [3] from 3 transfections; e, drTRPM5(E212A): 1 μM Ca2+ [4], 30 μM [3], 100 μM [4], 1000 μM [3] from 3 transfections; f, drTRPM5(D336A): 1 μM Ca2+ [3], 30 μM [3], 100 μM [3], 1000 μM [3] from 2 transfections; g, hsTRPM5(WT): 1 μM Ca2+ [5], 30 μM [5] from 3 transfections; h, hsTRPM5(E351A): 1 μM Ca2+ [3], 30 μM [3] from 1 transfection; and i, hsTRPM5(E560A): 1 μM Ca2+ [5], 30 μM [3], 100 μM [3], 1000 μM [3] from 5 transfections. j–r, Mean current (50 ms) of experiments were plotted versus indicated voltage. The +200 mV clamp was chosen for normalization. Horizontal bars represent SEM. The current-voltage relation plots of drTRPM5(E337A) are identical to those presented in Fig. 3e. s–u, Individual patch clamp measurements I−200mV / I+200mV, I+200mV, I−200mV, of experiments are shown as individual points, where bars represent mean values. v, w, Currents of drTRPM5 WT and CaICD mutants plotted as a function of Ca2+ concentration for voltage clamps of −200 mV and +200 mV. Symbols represent mean current and horizontal bar is SEM.
Extended Data Figure 2: Cyro-EM analysis of apo–TRPM5.
a, The representative 2D class average of apo–TRPM5. b, The Fourier shell correlation (FSC) curves for the apo–TRPM5. The cryo-EM map FSC is shown in black and the model vs. map FSC is shown in red. The map resolution was determined by the gold-standard FSC at 0.143 criterion, whereas the model vs. map resolution was determined by a threshold of 0.5. c, The angular distribution of particles that gave rise to the apo–TRPM5 cryo-EM map reconstruction. d, A schematic domain organization of a single TRPM5 subunit. Secondary structures and important domains are labeled. e, The atomic model of a single TRPM5 subunit in cartoon representation. The domains are colored as in (d). The left and right panels are two different views of the same subunit rotated 180° along the central axis.
Extended Data Figure 3: Cyro-EM analysis of TRPM5 in the presence of different concentrations of Ca2+, or in the presence of Ca2+ and NDNA.
a and d, The representative 2D class average the 5mM Ca2+ dataset (a) and 6 μM Ca2+ dataset (d), respectively. b and e, The Fourier shell correlation (FSC) curves for the 5mM Ca2+ dataset (b) and 6 μM Ca2+ dataset (e). The cryo-EM map FSC is shown in black and the model vs. map FSC is shown in red. The map resolution is determined by the gold-standard FSC at 0.143 criterion, whereas the model vs. map resolution is determined by a threshold of 0.5. c and f, The angular distribution of particles that give rise to the cryo-EM map reconstruction for 5mM Ca2+ dataset (c) and 6 μM Ca2+ dataset (f). g, The close up view of the CaTMD and CaICD of the 6 μM Ca2+ dataset. From left to right, CaTMD of apo–TRPM5(6 μM Ca2+), CaTMD of Ca2+–TRPM5(6 μM Ca2+), CaICD of apo–TRPM5(6 μM Ca2+), and CaICD of Ca2+–TRPM5(6 μM Ca2+). The cryo-EM densities are shown in mesh representation. The expected Ca2+ density is indicated by a circle. h, The representative 2D class average the (Ca2+, NDNA)–TRPM5 dataset. i, The Fourier shell correlation (FSC) curves for the (Ca2+, NDNA)–TRPM5. The cryo-EM map FSC is shown in black and the model vs. map FSC is shown in red. The map resolution was determined by the gold-standard FSC at 0.143 criterion, whereas the model vs. map resolution was determined by a threshold of 0.5. j, The angular distribution of particles that gave rise to the (Ca2+, NDNA)–TRPM5 cryo-EM map reconstruction.
Extended Data Figure 4: Local resolution estimation of TRPM5 structures and representative densities.
a-d, The local resolution estimation for apo–TRPM5(GDN) (a), Ca2+–TRPM5(GDN) (b), Ca2+–TRPM5(E337A)(GDN) consensus (c), (Ca2+, NDNA)–TRPM5(GDN) (d). For each map, a side view, a top-down view of the TMD from the extracellular side, and a focused side view of the S6 and pore helix are shown. The color bar unit is in Ångstroms. e, Representative densities from Ca2+–TRPM5(GDN) map. For the GDN density, one maltose group of the molecule is not resolved in the cryo-EM density map.
Extended Data Figure 5: The gate and the selectivity filter of TRPM5.
a, Cryo-EM densities of the CaICD site, contoured at 0.018. b, Cryo-EM densities of the CaTMD site, contoured at 0.022. c, Cryo-EM densities of the water molecule and residues in the selectivity filter, contoured at 0.023. Hydrogen bonds are shown as solid yellow lines. The “lower” water molecule is surrounded by the sidechain of Q906 and the backbone oxygen atoms of F904 and G905, forming three hydrogen bonds. d, Cryo-EM densities of I966, which forms the channel gate, contoured at 0.03. e, The selectivity filter formed by two layers of ordered water molecules (blue spheres) and backbone oxygen atoms (pink spheres) of G905. f and g, The two hydration layers the selectivity filter viewed from the extracellular side. Upper layer in (f) and lower layer in (g).
Extended Data Figure 6: Comparison of TRPM5 with other TRPM channels.
a, A structural comparison between Ca2+–TRPM5 and TRPM4 (PDBID: 6BQV)21. A single subunit is in color and shown as a cartoon. The TRPM5 channel is more compact, but wider than, the TRPM4 channel. b, An overlap of the selectivity filter of Ca2+–TRPM5 (red) and TRPM4 (yellow). c, A comparison of the CaTMD site for the available TRPM members. From left to right, drTRPM5, hsTRPM4 (6BQV), drTRPM2 (6DRJ), nvTRPM2 (6CO7), hsTRPM2 (6PUS), and pmTRPM8 (6O77)21,23,34,41,42; Shown in parentheses are the PDBIDs. d, Comparison of the “square” helices in TRPM2 (6PUO), TRPM4 (6BQR), TRPM5, TRPM7 (5ZX5), TRPM8 (6O6A); Shown in parentheses are the PDBIDs21,23,42,44. Only TRPM5 has a broken square helix. e, A sequence alignment of the square helix across different TRPM5 orthologs and TRPM family members. Red indicates that the α-helix is observed in structures. For TRPM1, TRPM3, and TRPM6, in which no structures are currently available, the helical annotation is based on the secondary structure prediction from PSIPRED server45.
Extended Data Figure 7: Comparison of (Ca2+, NDNA)–TRPM5 with apo–TRPM5 and Ca2+–TRPM5.
a, the chemical structure of N’-(3,4-dimethoxybenzylidene)-2-(naphthalen-1-yl)acetohydrazide (NDNA). b, Two close-up views of the cryo-EM densities of NDNA molecule. The surrounding protein structural element is shown in cartoon representation. c, Comparison of the NDNA binding site between (Ca2+, NDNA)–TRPM5 and apo–TRPM5 structures. The W869 is flipped in the (Ca2+, NDNA)–TRPM5 structure (cyan) compared to that in apo–TRPM5 structure (blue). d, Overlay of (Ca2+, NDNA)–TRPM5 (cyan) with apo–TRPM5 (blue) and Ca2+–TRPM5 (red) structures view from the intracellular side. One subunit is shown in cartoon representation and the other three subunits are in surface representation. The ICD of (Ca2+, NDNA)–TRPM5 adopts an intermediate state compared to the apo–TRPM5 and Ca2+–TRPM5 structures. e, The superimposition of the S1-S4 domain between (Ca2+, NDNA)–TRPM5 (cyan) and apo–TRPM5 (blue). f, The superimposition of the S1-S4 domain between (Ca2+, NDNA)–TRPM5 (cyan) and Ca2+–TRPM5 (red). g, A close-up view of the CaTMD site in (Ca2+, NDNA)–TRPM5 structure. The Q771 moved away from CaTMD. h and i, An overlay of the pore domain between (Ca2+, NDNA)–TRPM5 (cyan) with apo–TRPM5 (blue) (h) and Ca2+–TRPM5 (red) (i) structures viewed from the extracellular side.
Extended Data Figure 8: Patch-clamp electrophysiology experiments of CaTMD mutants of zebrafish TRPM5.
a, Representative whole-cell current traces of tsA overexpressing WT and CaTMD mutant TRPM5 channels. Clamps were imposed from +200 mV to −200 mV. The number of cells measured were tsA201 [n = 4 cells], TRPM5(WT) [5], TRPM5(E768A) [5], TRPM5(Q771A) [4], TRPM5(N794A) [4], TRPM5(D797A) [4], and TRPM5(E994A) [4] from 2–3 transfections. b, Mean current amplitudes of experiments in (a) were measured at 50 ms and plotted as a function of clamp voltage. Horizontal bars represent SEM. c, Individual measurements at clamps of +200 mV (I+200mV) and −200 mV (I−200mV) of experiments in (a) are shown as individual points, with bars representing mean values.
Extended Data Figure 9: Patch-clamp electrophysiology experiments on the NDNA binding site mutants.
a–j, Current voltage relations of whole-cell measurements in tsA cells over-expressing WT and mutant TRPM5 channels. Symbols represents mean current and horizontal bars are SEM. k–n, Individual measurements (of experiments in a–j) for clamps of +100 mV and −100 mV. Each point represents an individual cell and bars represent mean current. Cells were first measured (clamps from −100 mV to +100 mV) in bath solution and then re-measured following bath perfusion of 10 μM NDNA. See the legend of Fig 4 for the number of cells used.
Extended Data Figure 10: Ca2+–TRPM5(E337A) in GDN detergent.
a, The representative 2D class average of Ca2+–TRPM5(E337A). b, The FSC curve for the consensus map of Ca2+–TRPM5(E337A). The map resolution was determined by the gold-standard FSC at 0.143 criterion. c, The angular distribution of particles that give rise to the consensus map of Ca2+–TRPM5(E337A). d, The FSC curve for apo–TRPM5(E337A) (left) and Ca2+–TRPM5(E337A) (right). For each panel, the cryo-EM map FSC curve is shown in black and the model vs. map FSC is shown in red. The map resolution was determined by the gold-standard FSC at 0.143 criterion, whereas the model vs. map resolution was determined by a threshold of 0.5.
Supplementary Material
Acknowledgements
We thank G. Zhao and X. Meng for the support with data collection at the David Van Andel Advanced Cryo-Electron Microscopy Suite. We appreciate the high-performance computing team of VAI for computational support. We thank D. Nadziejka and M. Martin for technical editing. W.L. is supported by National Institutes of Health (NIH) grants (R56HL144929, R01HL153219, and R01NS112363). J.D. is supported by a McKnight Scholar Award, a Klingenstein-Simon Scholar Award, a Sloan Research Fellowship in neuroscience, a Pew Scholar in the Biomedical Sciences award, and NIH grant (R01NS111031). Z.R. is supported by an American Heart Association postdoctoral fellowship (20POST35120556).
Footnotes
Competing Interests Statement
The authors declare no conflicts of interest.
References
- 1.Pérez CA et al. A transient receptor potential channel expressed in taste receptor cells. Nat. Neurosci 5, 1169–1176 (2002). [DOI] [PubMed] [Google Scholar]
- 2.Zhang Y et al. Coding of sweet, bitter, and umami tastes: different receptor cells sharing similar signaling pathways. Cell 112, 293–301 (2003). [DOI] [PubMed] [Google Scholar]
- 3.Roper SD Signal transduction and information processing in mammalian taste buds. Pflugers Arch. 454, 759–776 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Taruno A et al. CALHM1 ion channel mediates purinergic neurotransmission of sweet, bitter and umami tastes. Nature 495, 223–226 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Brixel LR et al. TRPM5 regulates glucose-stimulated insulin secretion. Pflugers Arch. 460, 69–76 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Colsoul B et al. Loss of high-frequency glucose-induced Ca2+ oscillations in pancreatic islets correlates with impaired glucose tolerance in Trpm5−/− mice. Proc. Natl. Acad. Sci. U. S. A 107, 5208–5213 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Howitt MR et al. Tuft cells, taste-chemosensory cells, orchestrate parasite type 2 immunity in the gut. Science (80-. ). 351, 1329 LP–1333 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Vennekens R, Mesuere M & Philippaert K TRPM5 in the battle against diabetes and obesity. Acta Physiol. 222, e12949 (2018). [DOI] [PubMed] [Google Scholar]
- 9.Nilius B & Owsianik G The transient receptor potential family of ion channels. Genome Biol. 12, 218 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Owsianik G, Talavera K, Voets T & Nilius B Permeation and selectivity of TRP channels. Annu. Rev. Physiol 68, 685–717 (2006). [DOI] [PubMed] [Google Scholar]
- 11.Hofmann T et al. TRPM5 is a voltage-modulated and Ca(2+)-activated monovalent selective cation channel. Curr. Biol 100, 15166–15171 (2003). [DOI] [PubMed] [Google Scholar]
- 12.Prawitt D et al. TRPM5 is a transient Ca2+-activated cation channel responding to rapid changes in [Ca2+]i. Proc. Natl. Acad. Sci. U. S. A 100, 15166–15171 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Liman ER The Ca(2+)-Activated TRP Channels: TRPM4 and TRPM5. in (eds. Liedtke WB & Heller S) (2007). [PubMed] [Google Scholar]
- 14.Ullrich ND et al. Comparison of functional properties of the Ca2+-activated cation channels TRPM4 and TRPM5 from mice. Cell Calcium 37, 267–278 (2005). [DOI] [PubMed] [Google Scholar]
- 15.Nilius B, Prenen J, Janssens A, Voets T & Droogmans G Decavanadate modulates gating of TRPM4 cation channels. J. Physiol 560, 753–765 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Philippaert K et al. Steviol glycosides enhance pancreatic beta-cell function and taste sensation by potentiation of TRPM5 channel activity. Nat. Commun 8, 14733 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Bryant R et al. Use of a trpm5 inhibitor to regulate insulin and glp-1 release. (2008).
- 18.Shigeto M et al. GLP-1 stimulates insulin secretion by PKC-dependent TRPM4 and TRPM5 activation. J. Clin. Invest 125, 4714–4728 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Guo J et al. Structures of the calcium-activated, non-selective cation channel TRPM4. Nature 552, 205–209 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Winkler PA, Huang Y, Sun W, Du J & Lü W Electron cryo-microscopy structure of a human TRPM4 channel. Nature 552, 200–204 (2017). [DOI] [PubMed] [Google Scholar]
- 21.Autzen HE et al. Structure of the human TRPM4 ion channel in a lipid nanodisc. Science 359, 228–232 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Duan J et al. Structure of full-length human TRPM4. Proc. Natl. Acad. Sci. U. S. A 115, 2377–2382 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Diver MM, Cheng Y & Julius D Structural insights into TRPM8 inhibition and desensitization. Science 365, 1434–1440 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Yin Y et al. Structural basis of cooling agent and lipid sensing by the cold-activated TRPM8 channel. Science 363, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Zhang Z, Zhao Z, Margolskee R & Liman E The Transduction Channel TRPM5 Is Gated by Intracellular Calcium in Taste Cells. J. Neurosci 27, 5777 LP–5786 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Mähler J & Persson I A study of the hydration of the alkali metal ions in aqueous solution. Inorg. Chem 51, 425–438 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Huffer KE, Aleksandrova AA, Jara-Oseguera A, Forrest LR & Swartz KJ Global alignment and assessment of TRP channel transmembrane domain structures to explore functional mechanisms. Elife 9, (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Nilius B et al. The selectivity filter of the cation channel TRPM4. J. Biol. Chem 280, 22899–22906 (2005). [DOI] [PubMed] [Google Scholar]
- 29.Tang L et al. Structural basis for Ca2+ selectivity of a voltage-gated calcium channel. Nature 505, 56–61 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Gao Y, Cao E, Julius D & Cheng Y TRPV1 structures in nanodiscs reveal mechanisms of ligand and lipid action. Nature 534, 347–351 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Liu C et al. A Non-covalent Ligand Reveals Biased Agonism of the TRPA1 Ion Channel. Neuron 109, 273–284.e4 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Nadezhdin KD et al. Extracellular cap domain is an essential component of the TRPV1 gating mechanism. Nat. Commun 12, 2154 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Yamaguchi S, Tanimoto A, Iwasa S & Otsuguro K-I TRPM4 and TRPM5 Channels Share Crucial Amino Acid Residues for Ca(2+) Sensitivity but Not Significance of PI(4,5)P(2). Int. J. Mol. Sci 20, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Zhang Z, Tóth B, Szollosi A, Chen J & Csanády L Structure of a TRPM2 channel in complex with Ca(2+) explains unique gating regulation. Elife 7, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Hughes TET et al. Structural basis of TRPV5 channel inhibition by econazole revealed by cryo-EM. Nat. Struct. Mol. Biol 25, 53–60 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Nilius B et al. The Ca2+-activated cation channel TRPM4 is regulated by phosphatidylinositol 4,5-biphosphate. EMBO J. 25, 467–478 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Tabur S et al. Role of the transient receptor potential (TRP) channel gene expressions and TRP melastatin (TRPM) channel gene polymorphisms in obesity-related metabolic syndrome. Eur. Rev. Med. Pharmacol. Sci 19, 1388–1397 (2015). [PubMed] [Google Scholar]
- 38.Valente P et al. Identification of molecular determinants of channel gating in the transient receptor potential box of vanilloid receptor I. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 22, 3298–3309 (2008). [DOI] [PubMed] [Google Scholar]
- 39.Gregorio-Teruel L, Valente P, González-Ros JM, Fernández-Ballester G & Ferrer-Montiel A Mutation of I696 and W697 in the TRP box of vanilloid receptor subtype I modulates allosteric channel activation. J. Gen. Physiol 143, 361–375 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Teng J, Loukin SH, Anishkin A & Kung C L596–W733 bond between the start of the S4–S5 linker and the TRP box stabilizes the closed state of TRPV4 channel. Proc. Natl. Acad. Sci 112, 3386 LP–3391 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Huang Y, Winkler PA, Sun W, Lü W & Du J Architecture of the TRPM2 channel and its activation mechanism by ADP-ribose and calcium. Nature 562, 145–149 (2018). [DOI] [PubMed] [Google Scholar]
- 42.Huang Y, Roth B, Lü W & Du J Ligand recognition and gating mechanism through three ligand-binding sites of human TRPM2 channel. Elife 8, (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Wang L et al. Structures and gating mechanism of human TRPM2. Science 362, (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Goehring A et al. Screening and large-scale expression of membrane proteins in mammalian cells for structural studies. Nat. Protoc 9, 2574–2585 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Mastronarde DN Automated electron microscope tomography using robust prediction of specimen movements. J. Struct. Biol 152, 36–51 (2005). [DOI] [PubMed] [Google Scholar]
- 46.Zheng SQ et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nat. Methods 14, 331–332 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Scheres SHW RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol 180, 519–530 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Zhang K Gctf: Real-time CTF determination and correction. J. Struct. Biol 193, 1–12 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Rohou A & Grigorieff N CTFFIND4: Fast and accurate defocus estimation from electron micrographs. J. Struct. Biol 192, 216–221 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Bepler T et al. Positive-unlabeled convolutional neural networks for particle picking in cryo-electron micrographs. Nat. Methods 16, 1153–1160 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Punjani A, Rubinstein JL, Fleet DJ & Brubaker MA cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296 (2017). [DOI] [PubMed] [Google Scholar]
- 52.Zivanov J, Nakane T & Scheres SHW A Bayesian approach to beam-induced motion correction in cryo-EM single-particle analysis. IUCrJ 6, 5–17 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Bai X, Rajendra E, Yang G, Shi Y & Scheres SHW Sampling the conformational space of the catalytic subunit of human γ-secretase. Elife 4, (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Rosenthal PB & Henderson R Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol 333, 721–745 (2003). [DOI] [PubMed] [Google Scholar]
- 55.Goddard TD et al. UCSF ChimeraX: Meeting modern challenges in visualization and analysis. Protein Sci. 27, 14–25 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Schrödinger L The {PyMOL} Molecular Graphics System, Version~2.3. (2015). [Google Scholar]
- 57.Emsley P & Cowtan K Coot: model-building tools for molecular graphics. Acta Crystallogr. D. Biol. Crystallogr 60, 2126–2132 (2004). [DOI] [PubMed] [Google Scholar]
- 58.Afonine PV et al. New tools for the analysis and validation of cryo-EM maps and atomic models. Acta Crystallogr. Sect. D, Struct. Biol 74, 814–840 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The cryo-EM density map and coordinates and atomic models were deposited in the EMDB (Electron Microscopy Data Bank) and the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCS-PDB), respectively, under the accession numbers EMD-23740 and PDB 7MBP (apo–TRPM5), EMD-23741 and PDB 7MBQ (Ca2+–TRPM5), EMD-23742 (apo–TRPM5(nanodisc)), EMD-23743 (Ca2+–TRPM5(nanodisc)), EMD-23744 and PDB 7MBR (apo–TRPM5(6μM Ca2+)), EMD-23745 and PDB 7MBS (Ca2+–TRPM5(6μM Ca2+)), EMD-23746 and PDB 7MBT (apo–TRPM5(E337A) single subunit and tetramer), EMD-23747 and PDB 7MBU (Ca2+–TRPM5(E337A) single subunit and tetramer), and EMD-23748 and PDB 7MBV ((Ca2+, NDNA)–TRPM5). Source data are provided with this paper.

















