Main

Temperature has a profound impact on the life of all organisms on Earth1,2. To survive and adapt to the ever-changing environment, animals and humans have evolved a specialized somatosensory system to detect temperature changes in the environment, thereby adjusting their behavior and physiology accordingly1,2. A central player in this process is molecular thermosensors that act in peripheral somatosensory neurons to sense a wide range of thermal cues, such as cold, cool, warm and hot temperatures1,2. Research in the past three decades has identified thermosensors detecting cool, warm and hot temperatures1. For example, genetic studies in mice have uncovered a trio of TRP channels (TRPV1, TRPM3 and TRPA1) essential for sensing hot temperatures3, while TRPM2 is found to be required for detecting warm temperatures4,5. Several other TRP channels have also been suggested as thermosensors detecting warm and hot temperatures, such as TRPV2, TRPV3, TRPV4, TRPM4 and TRPM5, though it remains unclear whether they mediate warm and hot sensing in vivo1. In addition, the menthol receptor TRPM8 is essential for sensing cool temperatures6,7. However, little is known about the thermosensors specialized in detecting cold temperatures8. Though several candidates, particularly those thermosensitive TRP channels, have been proposed as cold sensors, their roles in mediating cold sensing in somatosensory neurons in vivo have not been validated1,8. As such, the molecular identities of cold sensors remain elusive, leaving a knowledge gap in our understanding of thermosensation.

Ionotropic glutamate receptors (iGluRs), including kainate, AMPA and NMDA receptors, are glutamate-gated cation channels9. These chemical-sensing receptors are primarily expressed in the brain where they sense and transmit glutamate signals to mediate synaptic transmission9. The kainate receptor GluK2 has emerged as a novel class of candidate cold sensor. This hypothesis is inspired by the observation that GLR-3, a Caenorhabditis elegans homolog of GluK2, mediates cold sensing in C. elegans10. Expression of GluK2 in heterologous systems confers cold sensitivity, indicating that GluK2 can function as a cold sensor in vitro10. The activation threshold of GluK2 is ~18 °C (ref. 10), which is substantially lower than that of TRPM8 (~23–28 °C)11,12,13. Surprisingly, GluK2 does not depend on its channel function for cold sensing but instead acts as a novel non-channel type of metabotropic cold sensor in heterologous systems, which transmits cold signals through G protein signaling10. This sets GluK2 apart from classical thermosensors such as the aforementioned thermosensitive TRP family channels, all of which are temperature-gated ion channels1,2,14. Besides the brain, GluK2 is expressed in dorsal root ganglion (DRG) neurons that directly sense temperature cues in the environment10; notably, the expression pattern of GluK2 in the DRG is largely nonoverlapping with those thermosensitive TRP channels15,16. Despite these interesting observations, it is not known whether GluK2 mediates cold sensing in DRG neurons in vivo. Nor is it clear whether GluK2 has any physiological roles in cold sensing at the organismal level.

Here, we interrogated the role of GluK2 in somatosensation in mice. We found that while GluK2 knockout (KO) mice respond normally to mechanical and heat stimuli, they exhibit a specific deficit in sensing cold rather than cool temperatures, and are also defective in cold nociception. This reveals a specific role of GluK2 in cold sensing. Selective deletion of GluK2 from DRG neurons led to similar cold-sensing phenotypes, suggesting that GluK2 functions in DRG neurons to mediate cold sensing. Functional recording of DRG neurons both in vitro and in vivo further supports a specific role of GluK2 in sensing cold rather than cool temperatures in these primary somatosensory neurons. Our results identify an unexpected function of GluK2—a kainate receptor best known to transmit chemical synaptic signals in the central nervous system—in sensing cold temperatures in the periphery.

Results

GluK2 KO mice respond normally to mechanical stimuli

Though GluK2 is best known for its function in transmitting synaptic glutamate signals in the brain, it is also expressed in the peripheral nervous system, particularly in DRG somatosensory neurons10,15,16, raising the possibility that GluK2 may have a role in somatosensation. To test this idea, we characterized somatosensory behaviors of GluK2 KO mice. We first tested how GluK2 KO mice respond to mechanical stimuli using several behavioral assays. We found that compared with control littermates, GluK2 KO mice showed normal responses in the von Frey assay and the brush assay which measure the animal’s sensitivity to punctate and dynamic mechanical stimuli17, respectively (Fig. 1a,b and Supplementary Fig. 1a,b). To test mechanical nociception, we performed the pinprick assay, but observed no defect in GluK2 KO mice (Fig. 1c and Supplementary Fig. 1c). Mice lacking TRPM8, the menthol receptor essential for cool sensing6,7,12,13, as well as GluK2/TRPM8 double knockout (dKO) mice, also exhibited no notable phenotype in all three mechanosensation assays (Fig. 1a–c and Supplementary Fig. 1a–c). All three types of KO mice showed normal locomotion in the rotarod test (Fig. 1d and Supplementary Fig. 1d). These results indicate that GluK2 is not required for mice to sense mechanical stimuli.

Fig. 1: GluK2 KO mice respond normally to mechanical and heat stimuli.
figure 1

a, GluK2 KO mice behave normally in the von Frey test measuring the mouse’s response to punctate mechanical stimuli. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 12, GluK2 KO n = 9, TRPM8 KO n = 10, GluK2/TRPM8 dKO n = 11. b, GluK2 KO mice behave normally in the brush test measuring the mouse’s response to dynamic mechanical stimuli. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 13, GluK2 KO n = 10, TRPM8 KO n = 13, GluK2/TRPM8 dKO n = 15. c, GluK2 KO mice behave normally in the pinprick test measuring the mouse’s response to noxious mechanical stimuli. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 13, GluK2 KO n = 10, TRPM8 KO n = 13, GluK2/TRPM8 dKO n = 15. d, GluK2 KO mice behave normally in the rotarod test measuring the mouse’s locomotion. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 10, GluK2 KO n = 9, TRPM8 KO n = 9, GluK2/TRPM8 dKO n = 12. e, GluK2 KO mice respond normally to heat in the two-temperature choice test. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 14, GluK2 KO n = 9, TRPM8 KO n = 12, GluK2/TRPM8 dKO n = 10. f, GluK2 KO mice show normal heat nociception in the hot plate assay. The latency of forepaw licking was quantified. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 13, GluK2 KO n = 10, TRPM8 KO n = 13, GluK2/TRPM8 dKO n = 15. g, GluK2 KO mice show normal heat nociception in the Hargreaves assay. The latency of paw withdrawal was quantified. Error bars, s.e.m. P values: all >0.05 (one-way ANOVA with Tukey’s test). Sample size: control n = 12, GluK2 KO n = 9, TRPM8 KO n = 9, GluK2/TRPM8 dKO n = 11. Note: all data are presented as mean ± s.e.m. Wild-type and heterozygous littermate mice were used as the controls in all panels. Both male and female mice were tested, and no significant difference between them was observed (Supplementary Fig. 1).

GluK2 KO mice respond normally to heat

We next checked the sensitivity of GluK2 KO mice to heat. We first conducted the two-temperature choice assay which assesses temperature preference and avoidance6. In this behavioral assay, animals were allowed to choose between their preferred temperature (30 °C) and a higher testing temperature varying from 30 °C to 55 °C, which covered the range of warm and hot temperatures (Fig. 1e). No defect was detected in GluK2 KO mice in this assay (Fig. 1e and Supplementary Fig. 1e), indicating that GluK2 KO mice possessed normal sensitivity to heat. To test heat nociception, we performed the hot plate assay and the Hargreaves assay, and found that GluK2 KO mice responded normally in both assays (Fig. 1f,g and Supplementary Fig. 1f,g), indicating that heat nociception is preserved in GluK2 KO mice. Similar results were obtained with TRPM8 KO and GluK2/TRPM8 dKO mice in all three heat-sensing tests (Fig. 1e–g and Supplementary Fig. 1e–g). It thus appears that GluK2 is not required for mice to sense heat.

GluK2 KO mice are defective in sensing cold temperatures

To test cold sensitivity, we performed the two-temperature choice assay in which mice were allowed to choose between their preferred temperature (30 °C) and a lower testing temperature varying from 30 °C to 5 °C, covering the range of cool (30–15 °C) and cold (<15 °C) temperatures6 (Fig. 2a–d). TRPM8 KO mice displayed a severe deficit in avoiding cool temperatures down to 15 °C (Fig. 2b and Supplementary Fig. 2a), supporting the notion that TRPM8 is a cool sensor. Notably, when the testing temperature reached the cold range (<15 °C), no significant difference was observed between TRPM8 KO mice and littermate control mice (Fig. 2b), consistent with previous reports that TRPM8 is not required for sensing cold temperatures6,7. By contrast, we observed no defect in GluK2 KO mice at all tested temperatures (Fig. 2b and Supplementary Fig. 2a). The lack of a phenotype in GluK2 KO mice in the cool temperature range is expected, as TRPM8 is the principle cool sensor with a relatively low activation threshold. This suggests that GluK2 does not contribute to cool sensing, which is consistent with its relatively high activation threshold10.

Fig. 2: GluK2 KO mice are defective in cold sensing.
figure 2

ad, GluK2/TRPM8 dKO and cKO mice are defective in sensing cold temperatures in the two-temperature choice assay. a, Schematic. b, GluK2/TRMP8 dKO mice are defective in cold sensing. Sample size: control n = 10, GluK2 KO n = 14, TRPM8 KO n = 13, GluK2/TRPM8 dKO n = 13. c, GluK2 cKOAvil/TRPM8 KO mice are defective in cold sensing. Sample size: control n = 10, GluK2 cKOAvil n = 6, TRPM8 KO n = 15, GluK2 cKOAvil/TRPM8 KO n = 17. d, GluK2 cKOPirt/TRPM8 KO mice are defective in cold sensing. Sample size: control n = 15, GluK2 cKOPirt n = 11, TRPM8 KO n = 18, GluK2 cKOPirt/TRPM8 KO n = 21. *P < 0.05, **P < 0.01 (one-way ANOVA with Tukey’s test, compared with control). #P < 0.05, ##P < 0.01 (one-way ANOVA with Tukey’s test, compared with TRPM8 KO). e, TRPM8 KO but not GluK2 KO mice are defective in sensing cool temperature (18 °C water droplet) in the paw withdrawal assay. Error bars, s.e.m. Sample size: control n = 15, GluK2 KO n = 9, TRPM8 KO n = 15, GluK2/TRPM8 dKO n = 14. **P = 6.9 × 10−6 (TRPM8 KO versus control), **P = 1.3 × 10−6 (GluK2/TRPM8 dKO versus control), one-way ANOVA with Tukey’s test. f, GluK2/TRPM8 dKO mice are defective in sensing cold temperature (0 °C water droplet). Sample size: control n = 15, GluK2 KO n = 9, TRPM8 KO n = 15, GluK2/TRPM8 KO n = 14. **P = 1.8 × 10−11 (GluK2/TRPM8 dKO versus control), one-way ANOVA with Tukey’s test. g, TRPM8 KO but not GluK2 cKOPirt mice are defective in cool sensing. Sample size: control n = 10, GluK2 cKOPirt n = 11, TRPM8 KO n = 15, GluK2 cKOPirt/TRPM8 KO n = 13. **P = 5.3 × 10−5 (TRPM8 KO versus control), **P = 4.8 × 10−5 (GluK2 cKOPirt/TRPM8 KO versus control), one-way ANOVA with Tukey’s test. h, GluK2 cKOPirt/TRPM8 KO mice are defective in cold sensing. Sample size: control n = 10, GluK2 cKOPirt n = 11, TRPM8 KO n = 15, GluK2 cKOPirt/TRPM8 dKO n = 13. **P = 1.4 × 10−6 (GluK2 cKOPirt/TRPM8 KO versus control), one-way ANOVA with Tukey’s test. Note: all error bars, s.e.m. All data are presented as mean ± s.e.m. Wild-type and heterozygous littermates were used as the controls in all panels. Both male and female mice were assayed, except in c where only males were tested because of a germline-leak concern in Avil-Cre females. No significant difference between males and females was observed (Supplementary Fig. 2).

The absence of a phenotype in GluK2 KO mice within the cold temperature range is not surprising, as TRPM8 is still functional in these mice. Though TRPM8 is classified as a cool sensor owing to its relatively low activation threshold, it retains activity at cold temperatures12,13, and is thus expected to sense cold temperatures in GluK2 KO mice. We therefore tested GluK2/TRPM8 dKO mice and found that indeed the dKO mice possessed a strong defect at cold temperatures in the two-temperature choice assay when compared with control littermates or TRPM8 KO mice (Fig. 2b), revealing an important role for GluK2 in mediating cold sensing in mice.

To provide additional evidence, we assessed the mouse’s sensitivity to cool and cold temperature stimuli by testing paw withdrawal responses evoked by low-temperature water droplets applied to their paw (Fig. 2e–h). Unlike the two-temperature choice assay that tests thermal preference and avoidance, this assay offers a more direct assessment of the mouse’s avoidance responses to temperature cues. As a control, we found that mice showed minimal or no response to water with temperature close to their paw’s skin surface temperature (27 °C) (Extended Data Fig. 1 and Supplementary Table 1). By contrast, mice responded robustly to cool temperature (18 °C) water mostly by lifting their paw (Fig. 2e, Supplementary Fig. 2c and Supplementary Table 1). Mice exhibited a more robust response to cold temperature (0 °C) water by lifting their paw followed by flinching and/or licking (Fig. 2f, Supplementary Fig. 2d and Supplementary Table 1). While TRPM8 KO mice displayed a severe deficit in response to cool temperature stimuli (Fig. 2e, Supplementary Fig. 2c and Supplementary Table 1), GluK2 KO mice responded normally (Fig. 2e, Supplementary Fig. 2c and Supplementary Table 1), further suggesting that GluK2 is not important for mice to sense cool temperatures. Strikingly, while neither GluK2 KO nor TRPM8 KO mice showed a significant defect in response to cold temperatures (Fig. 2f, Supplementary Fig. 2d and Supplementary Table 1), GluK2/TRPM8 dKO mice manifested a strong defect when challenged by such cold stimuli (Fig. 2f, Supplementary Fig. 2d and Supplementary Table 1), unveiling an important role for GluK2 in sensing cold temperatures. This set of data, together with those described above, uncovers a rather specific function of GluK2 in sensing cold but not cool, heat or mechanical stimuli.

GluK2 KO mice are defective in cold nociception

The observation that GluK2 is important for mice to sense cold temperatures prompted us to ask whether GluK2 contributes to cold nociception. As both the two-temperature choice assay and paw withdrawal assay are better suited for testing threshold-level thermal discrimination and avoidance rather than nociception18, we performed the cold-plate assay which is commonly used to examine cold nociception19. To do so, we assayed cold-induced repetitive paw-licking responses evoked by sustained exposure to a 0 °C cold plate19 (Fig. 3a). Consistent with previous reports6,7, TRPM8 KO mice did not show a defect in the cold-plate assay (Fig. 3b and Supplementary Fig. 3a), supporting the notion that the cool sensor TRPM8 does not play a major role in mediating cold nociception. By contrast, GluK2 KO mice displayed a strong defect in cold nociception (Fig. 3b and Supplementary Fig. 3a), and a similar defect was observed in GluK2/TRPM8 dKO mice (Fig. 3b and Supplementary Fig. 3a). Thus, GluK2 is not only important for mice to sense cold temperatures but also mediates cold nociception.

Fig. 3: GluK2 KO mice are defective in cold nociception.
figure 3

a, Schematic diagram of the cold-plate assay. b, GluK2 KO mice are defective in cold nociception as shown by the cold-plate assay. The latency of forepaw licking was quantified. Error bars, s.e.m. Sample size: control n = 15, GluK2 KO n = 13, TRPM8 KO n = 12, GluK2/TRPM8 dKO n = 10. **P = 2.7 × 10−5 (GluK2 KO versus control), **P = 0.0053 (GluK2/TRPM8 dKO versus control), one-way ANOVA with Tukey’s test. c, GluK2 cKOAvil mice are defective in cold nociception as shown by the cold-plate assay. An Avil-Cre line was used to selectively ablate GluK2 from DRG neurons. The latency of forepaw licking was quantified. Error bars, s.e.m. Sample size: control n = 15, GluK2 cKOAvil n = 10, TRPM8 KO n = 15, GluK2 cKOAvil/TRPM8 dKO n = 12. **P = 0.00022 (GluK2 cKOAvil versus control), **P = 0.0019 (GluK2 cKOAvil/TRPM8 dKO versus control), one-way ANOVA with Tukey’s test. d, GluK2 cKOPirt mice are defective in cold nociception as shown by the cold-plate assay. A Pirt-Cre line was used to selectively ablate GluK2 from DRG neurons. The latency of forepaw licking was quantified. Error bars, s.e.m. Sample size: control n = 23, GluK2 cKOPirt n = 15, TRPM8 KO n = 12, GluK2 cKOPirt/TRPM8 dKO n = 15. **P = 4.8 × 10−7 (GluK2 cKOPirt versus control), **P = 1.6 × 10−7 (GluK2 cKOPirt/TRPM8 dKO versus control), one-way ANOVA with Tukey’s test. Note: all data are presented as mean ± s.e.m. Wild-type and heterozygous littermate mice were used as the controls in all panels. Both male and female mice were assayed, except in c where only male mice were selected for testing because of a germline-leak concern in Avil-Cre female mice. No significant difference between males and females was observed in the behavioral assays (Supplementary Fig. 3).

Mick lacking GluK2 in DRG neurons are cold-sensing defective

In addition to the central nervous system, GluK2 is expressed in DRG neurons20. We thus wondered if the observed cold-sensing phenotypes in GluK2 KO mice resulted from a defect in DRG neurons. To address this question, we conditionally knocked out GluK2 from DRG neurons using two independent Cre lines: Avil-Cre and Pirt-Cre21,22. As a control, we first examined the heat and mechanical sensitivity of GluK2 conditional knockout (cKO) mice. We found that GluK2 cKOAvil and GluK2 cKOPirt mice, as well as GluK2 cKOAvil/TRPM8 KO and GluK2 cKOPirt/TRPM8 KO mice, all responded normally to heat and mechanical stimuli (Extended Data Figs. 2 and 3), a result similar to that obtained with whole-body KO mice. Notably, as was the case with whole-body KO mice, GluK2 cKOPirt/TRPM8 KO and GluK2 cKOAvil/TRPM8 KO mice manifested a strong defect in sensing cold temperatures in the two-temperature choice assay (Fig. 2c,d and Supplementary Fig. 2b). Similarly, a strong defect was found in GluK2 cKOPirt/TRPM8 KO mice in sensing cold temperatures in the paw withdrawal assay (Fig. 2g,h, Supplementary Fig. 2e,f and Supplementary Table 1). As observed in whole-body KO mice, both GluK2 cKOAvil and GluK2 cKOPirt mice were also defective in cold nociception in the cold-plate assay (Fig. 3c,d and Supplementary Fig. 3b). This set of data demonstrates that selective deletion of GluK2 from DRG neurons led to similar cold-sensing phenotypes, suggesting that GluK2 can function in the DRG to mediate cold sensing in mice.

GluK2 is important for DRG neurons to sense cold in vitro

Having identified an important role for GluK2 in mediating cold sensing at the behavioral level, we then sought to characterize the underlying neuronal mechanisms. To do so, we dissociated DRG neurons and recorded their responses to cooling by calcium imaging (Fig. 4a). As previously reported6, we identified two major types of cooling-activated DRG neurons. The first group can be activated by cool temperature stimuli (cooling to room temperature at 21 °C) (Fig. 4b and Extended Data Figs. 4 and 5), though deeper cooling to the cold range (10 °C) can also activate this type of DRG neurons (Fig. 4b and Extended Data Fig. 4a). These neurons were sensitive to menthol, suggesting that they are TRPM8+ neurons (Fig. 4b and Extended Data Fig. 4a). We thus named them cool-sensitive neurons. A second group of cooling-activated DRG neurons possessed a relatively higher activation threshold, as they were insensitive to cool temperatures and can only be activated by cold temperatures (Fig. 4b and Extended Data Fig. 4a). Unlike cool-sensitive neurons, these neurons were insensitive to menthol (Fig. 4b and Extended Data Fig. 4a). We thus named them cold-specific neurons. Both cool-sensitive and cold-specific neurons showed little or no sensitivity to AITC (Extended Data Fig. 6a,b), an agonist of TRPA1 that plays an important role in sensing hot temperatures in DRG neurons3. In TRPM8 KO mice, cool-sensitive neurons were largely absent (Fig. 4c,f and Extended Data Figs. 4 and 5), suggesting that their cool sensitivity is mediated by TRPM8. As expected, the population of cool-sensitive neurons remained intact in GluK2 KO mice (Fig. 4d,f and Extended Data Figs. 4 and 5), supporting the notion that GluK2 does not contribute to cool sensing. Notably, the cold-specific neuron population remained intact in TRPM8 KO mice (Fig. 4c,g and Extended Data Figs. 4 and 5), indicating that TRPM8 does not contribute to the cold sensitivity of cold-specific neurons. Strikingly, while the loss of GluK2 did not affect the population of cool-sensitive neurons (Fig. 4d,f and Extended Data Figs. 4 and 5), the cold-specific neuron population was greatly reduced in the DRG culture prepared from GluK2 KO mice (Fig. 4d,g and Extended Data Figs. 4 and 5), revealing a specific function for GluK2 in mediating cold rather than cool sensing in DRG neurons.

Fig. 4: Calcium imaging reveals that GluK2 is important for sensing cold rather than cool temperatures in dissociated DRG neurons.
figure 4

a, Schematic description of the assay. be, Sample traces of cool-sensitive and cold-specific DRG neurons. b, Control: DRG neurons isolated from wild-type littermate mice. c, TRPM8 KO. d, GluK2 KO. e, GluK2/TRPM8 dkO. Dissociated DRG neurons were recorded for their responses to temperature stimuli by calcium imaging. The temperature stimulus protocol is indicated on the top. Menthol: 100 μM. KCl: 50 mM. f, KO of GluK2 does not affect the cool-sensitive DRG neuron population, though these neurons are nearly eliminated by KO of TRPM8. Control: DRG neurons isolated from wild-type littermate mice. The numbers of responding neurons and total neurons assayed are listed. **P = 1 × 10−10 (TRPM8 KO versus control group), **P = 5.2 × 10−7 (GluK2/TRPM8 dKO versus control group), chi-squared test. g, KO of GluK2 but not TRPM8 greatly reduces the cold-specific DRG neuron population. Control: DRG neurons isolated from wild-type littermate mice. The numbers of responding neurons and total neurons assayed are listed. The number of mice used to isolate DRG neurons in f and g: control (31 mice), TRPM8 KO (19 mice), GluK2 KO (15 mice), GluK2/TRPM8 dKO (10 mice). **P = 4.7 × 10−7 (GluK2 KO versus control group), **P = 5.2 × 10−7 (GluK2/TRPM8 dKO versus control group), chi-squared test.

We also identified a small population (~3% of DRG neurons) of a third type of cooling-activated DRG neurons that were sensitive to menthol but unresponsive to cool temperature stimuli, though they can be activated by cold temperatures (Extended Data Fig. 7a). These menthol-sensitive neurons were not affected by GluK2 deletion but were nearly absent in TRPM8 KO mice (Extended Data Fig. 7a), indicating that they were also TRPM8+ neurons. A similar phenomenon has been reported previously23, but it is unclear exactly why these TRPM8+ neurons were insensitive to cool temperature stimuli; nevertheless, due to their small population size and independence of GluK2, we did not further characterize these neurons.

Notably, in GluK2/TRPM8 dKO mice, both the cool-sensitive and cold-specific neuron populations were greatly decreased (Fig. 4e–g and Extended Data Figs. 4 and 5). Thus, it appears that the majority of cooling-activated neurons in the DRG depend on GluK2 and TRPM8. Together, this set of experiments unveil a specific role for GluK2 in sensing cold rather than cool temperatures in dissociated DRG neurons.

To obtain additional evidence supporting a role for GluK2 in cold sensing, we recorded dissociated DRG neurons by whole-cell patch-clamp recording. To facilitate the identification of cold-specific neurons, we focused on patching DRG neurons isolated from TRPM8 KO mice, as these mice lack cool-sensitive neurons in the DRG. We found that cold stimuli triggered action potential (AP) firing in ~11% (14 of 130) of DRG neurons isolated from TRPM8 KO mice (Fig. 5a,c). Importantly, the percentage of such cold-activated DRG neurons was reduced to ~2.4% (2 of 84) in GluK2/TRPM8 dKO mice (Fig. 5b,c), uncovering an important role of GluK2 in mediating cold sensing in DRG neurons. As a control, we detected no notable difference in the intrinsic excitability and other electrical properties of DRG neurons between TRPM8 KO and GluK2/TRPM8 dKO mice (Extended Data Fig. 8), indicating that loss of GluK2 did not affect the overall health of DRG neurons. These recordings provide further evidence supporting a role for GluK2 in sensing cold temperatures in dissociated DRG neurons, unveiling a neuronal mechanism underlying the observed cold-sensing behavioral phenotypes in GluK2 KO mice.

Fig. 5: Patch-clamp recording reveals that GluK2 is important for cold sensing in dissociated DRG neurons.
figure 5

a, Sample traces of cold-evoked AP firing in DRG neurons isolated from TRPM8 KO mice. The numbers of responding and nonresponding neurons and total neurons recorded are indicated. b, Sample traces of cold-evoked AP firing in DRG neurons isolated from GluK2/TRPM8 dKO mice. The numbers of responding and nonresponding neurons and total neurons recorded are indicated. c, The percentage of cold-activated DRG neurons is greatly reduced in GluK2/TRPM8 dKO mice compared with TRPM8 KO mice. Bar graph summarizing the data in a and b. The numbers of responding neurons and total neurons recorded are listed. The numbers of mice used to isolate DRG neurons: TRPM8 KO (11 mice), GluK2/TRPM8 dKO (6 mice). Error bars, s.e.m. *P = 0.023 (GluK2/TRPM8 dKO versus TRPM8 KO group), chi-squared test.

GluK2 is important for DRG neurons to sense cold in vivo

To provide further evidence, we set out to record the activity of DRG neurons in response to cooling by in vivo calcium imaging using a protocol as previously described24,25,26. To do so, we expressed GCaMP6, a genetically encoded calcium sensor, in DRG neurons by intrathecal injection of AAV9-GCaMP6 viruses into mice (Fig. 6a). Similar to the case with dissociated DRG neurons cultured in vitro, we also detected two distinct classes of cooling-activated DRG neurons, cool-sensitive neurons and cold-specific neurons, when imaging the DRG in vivo (Fig. 6b,c and Extended Data Fig. 9a). Specifically, application of cool temperature stimuli (cooling to 21 °C) to the mouse paw was sufficient to activate cool-sensitive DRG neurons (Fig. 6c and Extended Data Fig. 9a). Similarly, deeper cooling to the cold range (10 °C) can also activate these neurons (Fig. 6c and Extended Data Fig. 9a). By contrast, cold-specific DRG neurons possessed a higher activation threshold, as they were unresponsive to cool temperatures and can only be turned on by cold temperature stimuli (Fig. 6c and Extended Data Fig. 9a). Knocking out TRPM8 nearly eliminated cool-sensitive neurons while leaving the population of cold-specific neurons unchanged (Fig. 6d,g,h and Extended Data Fig. 9), supporting the notion that TRPM8 is required for DRG neurons to sense cool but not cold temperatures. By contrast, GluK2 ablation led to a severe reduction in cold-specific neurons while having no significant effect on the cool-sensitive neuron population (Fig. 6e,g,h and Extended Data Fig. 9), uncovering a specific role for GluK2 in mediating cold sensing in DRG neurons in vivo. Lastly, dKO of GluK2 and TRPM8 greatly reduced the population of both cool-sensitive and cold-specific DRG neurons (Fig. 6f–h and Extended Data Fig. 9). These results provide a neuronal basis for the cold-sensing behavioral phenotypes observed in GluK2 KO mice.

Fig. 6: GluK2 is important for sensing cold rather than cool temperatures in DRG neurons in vivo.
figure 6

a, Schematic description of the assay. GCaMP6 was expressed in DRG neurons by intrathecal injection of AAV9-GCaMP6 viruses into the subarachnoid space near L5. L5 DRG neurons were imaged for responses to temperature stimuli applied to the hindpaw. b, Sample images of cool-sensitive and cold-specific DRG neurons. Shown are snapshot pseudocolor images of DRG neurons in response to cool and cold temperature stimuli. Pseudocolors refer to relative calcium levels in the neuron. Five mice were recorded. cf, Sample traces of cool-sensitive and cold-specific DRG neurons. c, Control: wild-type and heterozygous littermate mice. d, TRPM8 KO mice. e, GluK2 KO mice. f, GluK2/TRPM8 dKO mice. The temperature stimulus protocol is shown on the top. g, The cool-sensitive DRG neuron population is not affected in GluK2 KO mice but is nearly eliminated in TRPM8 KO and GluK2/TRPM8 dKO mice. Control: wild-type and heterozygous littermate mice. The numbers of responding neurons and total neurons assayed are listed. Five mice in each group. **P = 4.7 × 10−7 (TRPM8 KO versus control), **P = 2.5 × 10−6 (GluK2/TRPM8 dKO versus control group), chi-squared test. h, The cold-specific DRG neuron population is greatly reduced in GluK2 KO mice and GluK2/TRPM8 dKO mice but is not notably affected in TRPM8 KO mice. Control: wild-type and heterozygous littermate mice. The numbers of responding neurons and total neurons assayed are listed. Five mice in each group. **P = 0.0023 (GluK2 KO versus control group), **P = 0.00026 (GluK2/TRPM8 dKO versus control group), chi-squared test.

Discussion

Animals and humans are capable of sensing a wide range of temperature cues, spanning from cold, cool and warm to hot temperatures1,2. The identification of a large number of temperature-sensitive TRP channels as molecular thermosensors detecting cool, warm and hot temperatures has greatly advanced our understanding of cool, warm and hot sensing1. By contrast, much less is known about cold sensing, largely due to the lack of a molecular understanding of thermosensors specialized in sensing cold temperatures1,8. In the current study, by characterizing GluK2 KO mice, we provided multiple lines of evidence uncovering a key role of GluK2 in cold sensing. Our results demonstrate that GluK2 mediates cold sensing in mice, supporting GluK2 as a cold sensor.

The role of GluK2 in cold sensing appears rather specific, as GluK2 KO mice respond normally to mechanical and heat stimuli. GluK2 KO mice also lack a phenotype in cool sensing, which is further supported by DRG neuron recordings performed both in vitro and in vivo. A specific role of GluK2 in sensing cold but not cool temperatures is consistent with its relatively high activation threshold compared with that of the cool sensor TRPM8, indicating that GluK2 and TRPM8 can function independently of each other in the DRG. This is also consistent with the finding that GluK2 and TRPM8 transcripts are detected in distinct groups of DRG neurons (refs. 15,16 and Extended Data Fig. 10). Nevertheless, it remains possible that GluK2 and TRPM8 may function together in a small DRG population, as a low percentage of DRG neurons coexpress GluK2 and TRPM8 (refs. 15,16 and Extended Data Fig. 10). Though TRPM8 is classified as a cool sensor, it can also be activated by cold temperatures and participates in cold sensing along with GluK2. Whether and how GluK2 interacts with TRPM8 at the cellular and/or circuit levels to mediate cold sensing requires further investigation. Lastly, while GluK2 appears to be a major contributor to cold sensing in DRG neurons, we do not exclude the possibility that additional cold sensors may exist, given the presence of residual cold responses in GluK2/TRPM8 dKO mice. One possibility is that other proposed cold sensors, such as TRPA1, TRPC5, K2P and Orai channels8,27,28,29, might play a role; alternatively, an unknown cold sensor(s) might be involved.

One interesting observation is that GluK2 KO but not TRPM8 KO mice are defective in cold nociception. While the absence of a notable phenotype in cold nociception in TRPM8 KO mice is expected, the presence of such a deficit in GluK2 KO mice is rather surprising, considering that TRPM8 would be expected to sense cold in GluK2 KO mice. On the other hand, cold nociception (paw-licking behavior) represents a typical interoceptive self-caring response to body injury, which is distinct from those threshold-level cold avoidance/discrimination behaviors that are considered reflex responses and likely require distinct circuit mechanisms in the spinal and brain regions18. This might explain the observed cold nociception phenotype in GluK2 KO mice. Exposure to cold, particularly chronic cold, causes tissue damage and evokes pain30,31. Cold-induced pain represents a severe health condition30,31. Currently, there are no effective cures for cold pain30,31. Our study identifies GluK2 as a new drug target for developing therapeutics treating cold pain.

GluK2 is known to possess both ionotropic and metabotropic functions32,33. In the central nervous system, GluK2 primarily functions as an ionotropic glutamate-sensing chemoreceptor transmitting synaptic glutamate signals9,34. In the DRG, our data suggest that GluK2 can act as a cold-sensing thermoreceptor. Interestingly, GluK2 relies on its metabotropic rather than ionotropic function for cold sensing in vitro by transmitting cold signals via G protein signaling, particularly Gi/o signaling10. Specifically, the cold-sensing function of GluK2, when expressed in heterologous systems, is not affected by mutations that ablate its channel activity, but instead can be blocked by the Gi/o inhibitor pertussis toxin10. In dissociated DRG neurons, we found that the percentage of cold-specific neurons but not cool-sensitive neurons was greatly reduced by pertussis toxin treatment (Extended Data Fig. 6c–e), suggesting that a similar metabotropic mechanism may underlie the function of GluK2 in cold sensing in DRG neurons. Thus, GluK2 may represent a novel non-channel type of thermosensor. How GluK2 transmits cold signals via G protein signaling to regulate neuronal excitability is currently unknown. Presumably, opening/closing of a downstream transduction channel(s) might be involved. It should be noted that over 20% of DRG neurons express GluK2 transcripts (Extended Data Fig. 10c), which is considerably higher than that of cold-specific neurons in the DRG. One possibility is that some GluK2+ DRG neurons might lack certain key components that act downstream of GluK2 to transmit cold signals; alternatively, such neurons might express some genes that inhibit the cold-sensing function of GluK2. Future studies are needed to elucidate the detailed signaling pathway by which GluK2 transmits cold signals. Evolutionarily, it remains a mystery as to how an ionotropic chemosensor in the central nervous system is co-opted as a metabotropic thermosensor in the periphery. Notably, iGluRs form an ancient gene family present in nearly every organism including bacteria9. Perhaps, some iGluRs (for example, GluK2) are multifunctional, capable of sensing chemicals as well as other cues such as temperature. Notably, most thermosensitive TRP channels (for example, TRPV1, TRPM2, TRPM3 and TRPM8) are only found in vertebrates but are absent in invertebrates1,35,36. In addition, for those thermosensitive TRPs that are present in both vertebrate and invertebrate genomes (for example, TRPA1), their temperature-sensing function tends to diverge across phyla1,35,36; for example, TRPA1 is a heat sensor in insects and snakes, but acts as a cold sensor in nematodes37,38,39,40,41. Unlike thermosensitive TRP channels, homologs of GluK2 exist in both vertebrate and invertebrate species9, and, importantly, they all can function as cold sensors at least in vitro10. We thus propose that GluK2 represents an ancient class of evolutionarily conserved thermosensor.

Methods

Animals

All animal experiments were performed in accordance with protocols approved by the Institutional Animal Care and Use Committee at the University of Michigan and Johns Hopkins University following National Institutes of Health (NIH) guidelines. All mouse lines were in the C57BL/6J background. Both male and female mice were used for all experiments, with the exception of mice containing Advillin-Cre, in which case only male mice were selected for testing due to a germline-leak concern in female mice. Mice were group housed at room temperature with ad libitum access to standard lab mouse pellet food and water on a 12-h light/12-h dark cycle. GluK2 KO mice were kindly provided by Susumu Tomita’s laboratory at Yale University20. GluK2-floxed mice were kindly provided by Anis Contractor’s laboratory at Northwestern University42. Pirt-Cre mice were generated previously43. Advillin-Cre mice were generated by Fan Wang’s laboratory at Massachusetts Institute of Technology and were acquired from the Jackson Laboratory (no. 032536). TRPM8 KO mice were generated by David Julius’s laboratory at University of California San Francisco and were acquired from the Jackson Laboratory (no. 008198).

Behavioral tests

All mice at 8–12 weeks of age were measured in a blinded manner. Mice were acclimatized to the behavioral testing equipment for 2–3 d before each test. After the ‘habituation’ session, acute somatosensory measures were recorded as described previously17,44.

The rotarod assay

Mice were placed on the accelerating rotarod (IITC). The rotarod began at 4 r.p.m. and accelerated to 40 r.p.m. over 5 min. The time to fall was automatically recorded. The test was repeated three times at 20-min intervals and the average was calculated as rotarod latency.

Mechanical sensitivity tests

Mice were habituated in individual black plexiglass chambers with a mesh floor for 30 min for 3 consecutive days before the test. On the testing day, mechanical sensitivity was measured via three assays in the given order at 20-min intervals: (1) von Frey assay, (2) brush assay and (3) pinprick assay. (1) von Frey assay. The plantar surface of the mouse hindpaw was stimulated with calibrated von Frey monofilaments (0.008 g to 2 g, North Coast). The paw withdrawal threshold to the stimulation was determined by Dixon’s up–down method45. (2) Brush assay. The hindpaw plantar was stimulated by gently stroking with a paintbrush (4-0) in the direction from heel to toe. The stimulation was performed ten times. Walking away or occasionally brief paw lifting (~1 s or less) following each stimulation was considered as a positive response. The test was repeated ten times at 1-min intervals, and the percentage of positive response was calculated. (3) Pinprick assay. The hindpaw plantar was stimulated with a pin without penetrating into the skin. Hindpaw lifting/flinching responses were considered as a positive response following each stimulation. The stimulation was performed ten times at 1-min intervals, and the percentage of positive response was calculated.

The Hargreaves test

Mice were habituated for 30 min in plastic boxes (IITC) on the glass floor for 3 d before the test. On the testing day, the mouse hindpaw plantar was exposed to a beam of radiant heat through a Hargraves device (IITC). The latency of paw withdrawal was recorded. The heat stimulation was repeated three times at 1-min intervals for each mouse, and the mean of the withdrawal latencies was calculated. A cutoff time of 30 s was set to prevent potential tissue damage.

The hot-plate/cold-plate tests

Mice were habituated on a temperature plate (IITC) at 22 °C (room temperature) for 10–15 min on 2 consecutive days. On the testing day, the temperature plate was set at 54 °C (hot) or 0 °C (cold). Mouse responses to temperatures were recorded by a camcorder (Sony). After the test, we went through the videos and calculated the latency of forepaw licking. To avoid tissue injury, a cutoff time was set at 30 s and 180 s for assays at 54 °C and 0 °C, respectively.

The two-temperature choice assay

Each of the two temperature plates (Bioseb) was set to the desired temperature. Mice were usually first tested using the hot- and cold-plate assays and then subjected to the two-temperature choice test. Mice were placed between the two plates, and their movements were tracked for 5 min via a camera controlled by a tracking software (Bioseb). The time spent on each plate during recording was analyzed to assess their thermal preference. As a first step, we tested each mouse to determine if it displayed a strong preference to either plate independently of temperature. To do so, we set both plates to 30 °C to assess if mice spent roughly 50% of their time on each side; however, some mice did not due to variations intrinsic to this assay, and those that spent more than 70% of their time on one plate were not included in further testing. Room humidity was maintained at 25–30%, as humidity affects the assay especially when the plate was cooled to cold temperatures, likely due to moisture accumulation on the plate. The same cohort of mice was assayed at all the temperatures for each genotype.

The temperature-controlled water droplet stimulation test

Mice were habituated in individual black plexiglass chambers with a mesh floor for 30 min on 2 consecutive days. On the testing day, following 30 min of habituation, a water droplet (20–25-μl volume) at a set temperature was applied to the plantar surface of the mouse hindpaw using a 2.5-ml syringe (Becton Dickinson) connected to a trimmed 200-μl pipette tip (inner diameter of the flat opening: ~2.5 mm). It is important to let the water droplet attached to the plantar for at least one second to allow enough time to cool down the plantar surface; otherwise, the test would be considered unsuccessful and reperformed 3 min later. Caution should be exercised to avoid directly touching the plantar surface with the pipette tip. The temperature of both the syringe and the water inside was carefully controlled by incubating the whole syringe in a water bath with a set temperature verified by a thermometer. The stimulation was performed three times at 3-min intervals. Each application was video-recorded for the first 2 min for subsequent analysis. The behavior in response to the temperature-controlled water stimulation was evaluated and scored: no response or simply walking away scored 0, a brief paw lifting scored 1, a paw flinch scored 2, multiple flinches or a paw lick scored 3, and repetitive paw licking and/or paw guarding scored 4.

DRG dissociation and calcium imaging of dissociated DRG neurons

Acute dissociation of DRG neurons

DRG neurons were acutely dissociated as described previously44. Mice at the age of 25–35 days postnatal were euthanized with CO2. DRGs from T10 to L6 were collected in icy Ca2+/Mg2+-free Hanks’ buffered salt solution (pH 7.0, Invitrogen). Collected DRGs were digested with papain (1.5 mg ml−1, adjust pH to 7.0, Sigma) with 1 mg of l-Cys (Sigma) for 15–20 min at 37 °C. Digestion solution was then replaced with collagenase II and dispase II (6 mg ml−1, Sigma) and incubated for 20–28 min at 37 °C. Following digestion, DRGs were gently triturated 8–10 times with a polished pipette. The cells were purified using a 70-μm filter (Falcon) and were centrifuged for 5 min at 1,000 r.p.m. to separate the cells from the supernatant. The cells were further resuspended in a neurobasal medium supplemented with 2% B27 supplement, 1% GlutaMAX, nerve growth factor (100 ng ml−1) and glial cell line-derived neurotrophic factor (40 ng ml−1) and then were plated on 13-mm coverslips that had been precoated with laminin and poly-d-lysine overnight and cultured for 1 h at 37 °C in an incubator. The medium was then replaced with fresh medium containing 2.5% FBS (Gibco) and cultured for 6 h before calcium imaging experiments.

Calcium imaging of dissociated DRG neurons

As previously described44, DRG neurons were first incubated at 37 °C for 30 min in 4 μM Fura-2-acetoxymethyl ester with 0.2% Pluronic F127 (Thermo Fisher Scientific), and then rinsed with PBS buffer. DRG neurons were then incubated in standard extracellular solution (10 mM HEPES, 5 mM KCl, 140 mM NaCl, 2 mM MgCl2, 2 mM CaCl2 and 10 mM glucose, pH 7.4) for 15 min in an incubator. Calcium imaging was performed using a ×20 water-immersion objective with high transmission efficiency of ultraviolet light (Olympus IX73) under the excitation wavelengths 340 nm and 380 nm. A Hamamatsu Digital Camera C11440 was used for image acquisition, and images were processed with MetaFluor software (Molecular Devices). Menthol (100 μM) and KCl (50 mM) were dissolved in standard extracellular solution for use in the experiment. We used a bipolar In-line Cooler/Heater (SC-20 from Warner Instruments) to control the temperature of the recording solution. The temperature protocol was described in Fig. 4. Lastly, high KCl solution (50 mM) was perfused to DRG neurons to validate neuronal activity, and those neurons displaying a KCl response lower than a 0.5-fold increase in ratio (ΔR/R) were considered unhealthy and not analyzed. Responding neurons were defined by the fluorescence ratio increase ΔR/R ≥ 0.2.

Whole-cell patch-clamp recording of dissociated DRG neurons

DRG neurons were dissociated as described above and cultured overnight for whole-cell patch-clamp recording. MultiClamp 700B amplifier and Axon Digidata 1550 digitizer (Axon Instruments) were used. Microelectrodes were pulled using borosilicate capillaries bearing filament (with inner diameter 0.86 mm, outer diameter 1.5 mm; World Precision Instruments) with a P-1000 pipette puller (Sutter Instruments). The resistance of the microelectrodes was 3–5 MΩ filled with pipette solution. After gigaseal formation, the whole-cell configuration was formed, then cell membrane capacitance and series resistance were compensated. The data were sampled at 10 kHz and low pass-filtered at 2 kHz with pClamp10.4 (Axon Instruments) and were analyzed with Clampfit 10.2 (Axon Instruments). The pipette solution contains (in mM): potassium gluconate 130, KCl 5, MgCl2 2, EGTA 5, HEPES 10, ATP·Na2 5, CaCl2 1, and pH adjusted to 7.2 with KOH, 300–310 mOsm. The external solution contains (in mM): NaCl 140, KCl 5, CaCl2 2.5, MgCl2 1, HEPES 10, D-glucose 10, and pH adjusted to 7.2 with NaOH, 315–320 mOsm.

The intrinsic properties and excitability of DRG neurons were studied by injecting a family of current steps ranging from −20 to 100 pA in 10-pA increments with a duration of 1,000 ms. AP firing frequency was quantified as the number of APs per second. Rheobase was detected by the minimum current (in pA) required to evoke an AP. The voltage threshold was defined as the voltage (in mV) at which the AP initiated, and the threshold level was determined as the voltage point where the rapid increase in voltage occurred (dV/dt > 15 mV s−1). The peak amplitude of APs (in mV) and half-width (in ms) were measured with the first AP appeared through event statistics.

To characterize cold temperature-evoked AP induction in DRG neurons, gap-free membrane potentials were recorded under current-clamp mode for a total of 6 min. The temperature of the bath solution was maintained and controlled by a temperature controller (CL-100, Warner Instruments) from 25 °C (1 min) to 10 °C (4 min) and back to 25 °C.

In vivo calcium imaging of DRG neurons

Intrathecal administration of Adeno-associated virus

Mice were anesthetized with 1.5–2% isoflurane and then placed on a heating pad to maintain the temperature around 35 °C. An incision was made on the back skin above the lumbar region, and a small cut was made on the intervertebral membrane between Lumbar 5 (L5) and L6 vertebrae (near the L5) to insert a small catheter of 0.2-mm diameter. Then, 5 μl of Adeno-associated virus (AAV) (pAAV9.CAG.GCaMP6s. WPRE.SV40, 2.5 × 1013 gc ml−1, Addgene) was infused into the intrathecal space through a catheter.

DRG exposure surgery and in vivo calcium imaging

Mice were anesthetized with sodium pentobarbital (40–50 mg kg−1). A dorsal laminectomy was performed from the spinal level L4 to L6, and only the transverse process of L5 was removed to expose the underlying L5 DRG while the vertebral plate over the spinal cord and dura was left intact. The surgery and imaging were performed as previously described24. During the surgery and imaging, mice were placed on a heating pad with a feedback-based system to maintain body temperature at 35 ± 0.5 °C as monitored by a rectal probe (DC temperature controller, FHC). Mice were kept anesthetized with 1.5–2% isoflurane (Anesthesia system, Harvard Apparatus) and placed on a microscope stage with the spinal column stabilized on a holder to avoid movement. The microscope stage was fixed under a laser-scanning confocal microscope (Leica), equipped with a macro-based, large-objective and fast EM-CCD camera. The live images were acquired at typically four frames with 600 Hz in frame-scan mode per 3 s (at depths below the dura from 0 to 70 μm, with a ×5, 0.5 numerical aperture macro dry objective at 512 × 512 pixel resolution). GCaMP6 was used as a highly sensitive calcium indicator of calcium transients in DRG neurons. For green fluorescence, solid diode lasers (Leica) were used (excitation at 488 nm and emission at 500–549 nm).

The L5 whole DRG was placed at the focal plane, and images were recorded while applying thermal stimulation to the right hindpaw of the mouse. The hindpaw was stabilized in a custom-designed stimulation container, and water at the adaptation temperature was infused into the container at an average flow rate as previously described46. The water temperature was monitored and adjusted using a thermometer. Before each trial, images were recorded for 30–60 s without water stimulation to obtain baseline fluorescence. Then, the container was filled with water of different stimulation temperatures with the same average flow rate. Water-based stimulation was maintained for 60 s for each stimulation temperature. At the end of each trial, the viability and virus infection efficiency were confirmed by assessing calcium responses in DRG neurons triggered by hindpaw pinch.

Data analysis

The raw image stacks were imported into FIJI (NIH) for further analysis. The optical sections of sequential time points were realigned, and motion corrected using a cross-correlation-based image alignment plugin in FIJI to adjust for minor motion shifts during recording. The corrected four optical sections from sequential time points were processed with z projection to obtain the final whole DRG neuron image in time-lapse movies. The amplitudes of the calcium transient signal were calculated as ΔF/F, where ΔF/F = (Ft − F0)/F0. Ft was the maximum fluorescence intensity measured by calculating the average pixel values (peak intensity − background) of the chosen region of interest (ROI) for each frame recorded, and F0 was the average fluorescence intensity of each ROI in the baseline period recorded 10–20 frames before temperature stimulation. Responding neurons were defined by the fluorescence intensity increase ΔF/F ≥ 0.2.

RNAscope in situ hybridization and immunohistochemistry

Mice were euthanized with CO2 and perfused with PBS followed with 4% PFA (4 °C) in PBS. L3 to L5 DRGs were extracted and postfixed in 4% PFA for 2 h at 4 °C. DRGs were then incubated in 20% sucrose for dehydration. DRGs were sectioned into 14-μm slices using a Cryostat (Leica Microsystems) and mounted onto SuperFrost Plus microscope slides. The RNAscope system (Advanced Cell Diagnostics) was used according to the manufacturer’s protocol. The mounting slices were pretreated with hydrogen peroxide and protease IV provided by the RNAscope Multiplex Fluorescent Reagent v2 assay kit (Advanced Cell Diagnostics). Samples were hybridized with a GluK2 messenger RNA probe (GRIK2-O2-C1) customized with targeted sequences or TRMP8-C3/TRPA1-C2 probe for 2 h. Samples were further incubated with AMP1, AMP2 or AMP3 for 15–30 min, and HRP-C1, HRP-C2 or HRP-C3 for 15 min. Opal Dye 570 (1:1,500; cat. no. FP1488001KT) or TSA Plus Cy5 (1:1,500; cat. no. NEL745E001KT) was added, incubated for 30 min and HRP blocked for 15 min. All the reagents were incubated in a 40 °C hybridization oven. To quantify positive neuron populations, neurons with more than three positive signals (observed as puncta) within the periphery of the neuron (determined by the phase contrast image) were considered as positive. Following RNAscope in situ hybridization, immunostaining was performed using the methods described previously10. Primary antibodies were used at the following concentrations: rabbit anti-NF200 antibody (1:500, cat. no. N4142, Millipore Sigma), rabbit anti-CGRP antibody (1:500, cat. no. PC205L, Millipore Sigma), IB4-Alexa Fluor 647 (1:500, cat. no. I32450, Thermo Fisher), Guinea pig anti-TRPV1 (VR1) antibody (1:500, cat. no. ACC-030-GP, Alomone Labs). Nuclear counterstaining was performed with DAPI solution provided by the RNAscope Multiplex Fluorescent Reagent v2 assay kit for 5 min. DRG sections were then mounted and imaged by fluorescence microscopy (Leica DMi8 and Olympus FV3000). Data were quantified with LAS X Life Science Microscope software (Leica Microsystems).

Single-cell RNA sequencing analysis

Raw single-cell RNA sequencing data in the form of .h5 files were obtained from a previously published article15. These files contained preprocessed gene expression matrices and relevant metadata. Data preprocessing was initiated by loading the .h5 files into R using the rhdf5 tool. The initial quality control steps were performed, including filtering out low-quality cells based on criteria such as gene count (fewer than 200 genes) and mitochondrial gene content (greater than 5%). Subsequent steps included normalization, scaling and dimensionality reduction using principal component analysis and Uniform Manifold Approximation and Projection (UMAP), as implemented in the Seurat pipeline. The preprocessed data were subjected to clustering analysis using Seurat’s FindNeighbors and FindClusters functions, with a default clustering resolution parameter. Further data analysis and visualization were performed using R with the Seurat package and the ShineyCell package47. We mapped GluK2 mRNA (GRIK2) expression in known subsets of DRG neurons such as myelinated neurons (NF classes, LdhbHigh), TRPM8, TRPA1 and TRPV1 (ref. 15). To calculate the percentage of overlap between the GRIK2-expressing neurons and other sensory neuron markers, we used a single-cell RNA analysis tool, miCV48. We investigated the coexpression of GRIK2 in the LdhbHigh (normalized expression above 6), TRPM8, TRPV1 and TRPA1High (normalized expression above 4) populations. Expression thresholds for Ldhb and TRPA1 were defined to overcome the ambiguity caused by the noise created by cells showing low or very-low expression in false-positive cells.

Statistical analysis

Statistical methods, error bars (s.e.m.), P values and sample sizes (n number) are indicated in the figure legends. One-way analysis of variance (ANOVA) with Tukey’s test, Student’s t-tests and chi-squared tests were applied separately to analyze data, and are described in the figure legends. Both the one-way ANOVA and chi-squared tests used were two-sided. No data were excluded from statistical analysis. A P value less than 0.05 was considered statistically significant, *P < 0.05, **P < 0.01, #P < 0.05, ##P < 0.01. All data were presented as mean ± s.e.m. and analyzed with GraphPad Prism 8.0 software (GraphPad). Mice were randomly allocated to different experimental groups. Data collection and analysis were performed blind to the conditions of the experiments in behavioral tests and in vitro calcium imaging. No statistical methods were used to predetermine sample sizes, but our sample sizes are similar to those reported in previous publications17. Data distribution was assumed to be normal, but this was not formally tested.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.