Figure 7 - uploaded by Pavel Adamek
Content may be subject to copyright.
Overexpression of TMEM100 in articular afferents induces secondary hyperalgesia in the hind paw but no knee joint pain. (A), Cartoon depicting the experimental approach (left), example image of a DRG cross section from a mouse that had received intraarticular AAV-PHP.S-TMEM100-Ires-dsRed (middle) and quantification of the total number of dsRed + neurons (right). Bars represent means ± SEM and values from individual DRGs are shown as black dots. (B), dsRed fluorescence in the tibial nerve distal to the knee (left) and the saphenous (right) nerve proximal to the knee which contains the medial articular nerve. (C), Comparison of the time courses of changes in stand time (left) and leg swing speed (right) of WT mice that intraarticularly received AAV-PHP.S-dsRed control virus (white circles) and AAV-PHP-S-TMEM100-Ires-dsRed (orange circles). Ratios at different time points were compared using multiple Mann-Whitney tests. (ns, P>0.05; *, P<0.05; N-numbers are provided in the graph legend). (D), Time courses of changes in mechanical paw withdrawal thresholds of AAV-PHP.S-dsRed (white circles) and AAV-PHP-S-TMEM100-Ires-dsRed (blue circles) injected WT mice (left) as well as comparison of the response rates at 14 dpi -i.e. % paw withdrawals in response to five successive stimulations with the indicated von Frey filaments (right). Mean paw withdrawal thresholds at different time points and response rates to different von Frey filaments were compared using Mann-Whitney test (***, P<0.001; number of tested animals is the same in both panels and is indicated in the graph legend). (E), Example traces of mechanicallyevoked action potentials recorded from cutaneous C-fiber nociceptors in the tibial nerve from control mice (top, AAV-PHP.S-dsRed) and from mice that overexpress TMEM100 in articular afferents (bottom, AAV-PHP.S-TMEM100-Ires-dsRed). (F), Comparison of the firing rates evoked by ramp-and-hold stimuli that exerted the indicated force to the receptive fields. Symbols represent means ± SEM numbers of action potentials, which were compared using multiple Mann-Whitney tests (*, P<0.05). (G), Comparison of the proportions of C-fiber nociceptors that respond to mechanical stimulation with the indicated von Frey filaments. The proportions were compared pairwise using the Chi-square test (ns, not significant; *, P<0.05). (H), Cartoon depicting the proposed mechanism underlying the induction of secondary hyperalgesia. Detailed statistical information and raw data is provided in supplemental Data S1.

Overexpression of TMEM100 in articular afferents induces secondary hyperalgesia in the hind paw but no knee joint pain. (A), Cartoon depicting the experimental approach (left), example image of a DRG cross section from a mouse that had received intraarticular AAV-PHP.S-TMEM100-Ires-dsRed (middle) and quantification of the total number of dsRed + neurons (right). Bars represent means ± SEM and values from individual DRGs are shown as black dots. (B), dsRed fluorescence in the tibial nerve distal to the knee (left) and the saphenous (right) nerve proximal to the knee which contains the medial articular nerve. (C), Comparison of the time courses of changes in stand time (left) and leg swing speed (right) of WT mice that intraarticularly received AAV-PHP.S-dsRed control virus (white circles) and AAV-PHP-S-TMEM100-Ires-dsRed (orange circles). Ratios at different time points were compared using multiple Mann-Whitney tests. (ns, P>0.05; *, P<0.05; N-numbers are provided in the graph legend). (D), Time courses of changes in mechanical paw withdrawal thresholds of AAV-PHP.S-dsRed (white circles) and AAV-PHP-S-TMEM100-Ires-dsRed (blue circles) injected WT mice (left) as well as comparison of the response rates at 14 dpi -i.e. % paw withdrawals in response to five successive stimulations with the indicated von Frey filaments (right). Mean paw withdrawal thresholds at different time points and response rates to different von Frey filaments were compared using Mann-Whitney test (***, P<0.001; number of tested animals is the same in both panels and is indicated in the graph legend). (E), Example traces of mechanicallyevoked action potentials recorded from cutaneous C-fiber nociceptors in the tibial nerve from control mice (top, AAV-PHP.S-dsRed) and from mice that overexpress TMEM100 in articular afferents (bottom, AAV-PHP.S-TMEM100-Ires-dsRed). (F), Comparison of the firing rates evoked by ramp-and-hold stimuli that exerted the indicated force to the receptive fields. Symbols represent means ± SEM numbers of action potentials, which were compared using multiple Mann-Whitney tests (*, P<0.05). (G), Comparison of the proportions of C-fiber nociceptors that respond to mechanical stimulation with the indicated von Frey filaments. The proportions were compared pairwise using the Chi-square test (ns, not significant; *, P<0.05). (H), Cartoon depicting the proposed mechanism underlying the induction of secondary hyperalgesia. Detailed statistical information and raw data is provided in supplemental Data S1.

Source publication
Preprint
Full-text available
Silent nociceptors are sensory afferents that are insensitive to noxious mechanical stimuli under normal conditions but become sensitized to such stimuli during inflammation. Using RNA-sequencing and quantitative RT-PCR we demonstrate that inflammation selectively upregulates the expression of the transmembrane protein TMEM100 in silent nociceptors...

Contexts in source publication

Context 1
... https://doi.org/10. 1101/2022 this end, we selectively overexpressed TMEM100 in knee joint afferents by intraarticular injection of an AAV-PHP.S-TMEM100-Ires-dsRed virus (30 µl, 1.5*10 11 vg; Figure 7A). Four days after intraarticular AAV-PHP.S-TMEM100-Ires-dsRed administration, we observed prominent dsRed fluorescence in a total of 339 ± 7 neurons in ipsilateral L3 and L4 DRG. ...
Context 2
... importantly, we observed numerous dsRed expressing nerve fibers in the saphenous nerve proximal to the knee, which includes the medial articular nerve that supplies the knee joint and in which silent nociceptors had first been described ( Schaible and Schmidt, 1988). Importantly, we observed hardly any dsRed + fibers in the tibial nerve distal to the knee, which contains cutaneous afferents that supply the plantar surface of the hind paw ( Figure 7B). Hence, intraarticularly administered AAV-PHP.S-TMEM100-Ires-dsRed causes selective overexpression of TMEM100 in knee joint but not in skin afferents. ...
Context 3
... intraarticularly administered AAV-PHP.S-TMEM100-Ires-dsRed causes selective overexpression of TMEM100 in knee joint but not in skin afferents. Interestingly, TMEM100-overexpressing mice exhibited normal gait, indicating that un-silencing of knee joint MIAs does not trigger knee joint pain ( Figure 7C). Strikingly, however, these mice developed profound mechanical hyperalgesia in the ipsilateral hind paw five days post AAV injection, which persisted until the end of the observation period (21 dpi; Figure 7D). ...
Context 4
... TMEM100-overexpressing mice exhibited normal gait, indicating that un-silencing of knee joint MIAs does not trigger knee joint pain ( Figure 7C). Strikingly, however, these mice developed profound mechanical hyperalgesia in the ipsilateral hind paw five days post AAV injection, which persisted until the end of the observation period (21 dpi; Figure 7D). Thus, the mechanical paw withdrawal thresholds decreased from 0.92 ± 0.066 g (-1 dpi) to 0.373 ± 0.035 g (14 dpi). ...
Context 5
... accordance with the behavioral outcome of TMEM100 over-expression in articular afferents, single-unit action potential recordings from a tibial nerve-glabrous skin preparation showed that cutaneous C-fiber nociceptors fire about twice as many action potentials in response to a given stimulus and have significantly reduced mechanical activation thresholds in mice that overexpress TMEM100 in articular afferents compared to mice that had received a control virus ( Figure 7E-G). Similar to mice that received intraarticular CFA, the mechanical activation thresholds of cutaneous Aδ-fiber nociceptors were slightly reduced and the action potential firing rate in response to suprathreshold stimuli was significantly increased ( Figure S4). ...
Context 6
... summary, our data shows that TMEM100 overexpression-induced un-silencing of mechanically insensitive articular afferents is sufficient to trigger mechanical hyperalgesia in remote skin regions ( Figure 7D). Together with the observation that TMEM100 is specifically up-regulated in MIAs during CFA-induced inflammation and that knock-out of TMEM100 exclusively abolishes long-lasting secondary hyperalgesia, these results suggest that sensory input from unsilenced MIAs is the . ...
Context 7
... https://doi.org/10. 1101/2022 main trigger for the induction of secondary hyperalgesia ( Figure 7H). While we have not examined the well-established contribution of central sensitization, our data reveals a previously unrecognized contribution of peripheral sensitization to secondary hyperalgesia. ...
Context 8
... we show that blocking the un-silencing of articular MIAs by knocking out TMEM100 prevents the development of long-lasting secondary hyperalgesia in remote skin regions, but does not alter pain at the actual site of CFAinduced inflammation (Figure 4). Moreover, TMEM100 overexpression-induced unsilencing of articular MIAs in the absence of inflammation or injury, induces secondary hyperalgesia but not pain hypersensitivity in the knee joint ( Figure 7C and D). Finally, our skin-nerve recordings demonstrate that secondary hyperalgesia, in addition to the previously described central sensitization, is partly driven by peripheral sensitization of cutaneous C-fiber nociceptors (Figure 6 and 7E-G). ...
Context 9
... however, seems unlikely, considering that TMEM100 expression is exclusively upregulated in MIAs in CFA-induced monoarthritis ( Figure 3C) and that mice lacking TMEM100 only show functional deficits in MIAs but not in other articular nociceptors (Figure 3 and 5). It should further be noted, that we observed off-target expression of TMEM100 in a few fibers in the tibial nerve distal to the knee, which supplies the plantar surface of the hind paw ( Figure 7B). Cutaneous C-fiber nociceptors that detect von Frey stimuli express MRGPRD ( Cavanaugh et al., 2009) and require PIEZO2 and TRPA1 for normal mechanosensitivity ( Garrison and Stucky, 2014;Murthy et al., 2018), but do not express TRPV1 (Cavanaugh et al., 2009;Zeisel et al., 2018). ...
Context 10
... have not explicitly tested if un-silencing of MIAs also triggers central sensitization, which is thought to be the major cause of secondary hyperalgesia. Yet, considering that the AAV-PHP.S-TMEM100 induced reduction of paw withdrawal thresholds ( Figure 7D) was larger than the reduction of the mechanical activation thresholds of individual cutaneous C-and Aδ-fiber nociceptors ( Figure 7G and Figure . CC-BY-NC-ND 4.0 International license made available under a (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. ...
Context 11
... have not explicitly tested if un-silencing of MIAs also triggers central sensitization, which is thought to be the major cause of secondary hyperalgesia. Yet, considering that the AAV-PHP.S-TMEM100 induced reduction of paw withdrawal thresholds ( Figure 7D) was larger than the reduction of the mechanical activation thresholds of individual cutaneous C-and Aδ-fiber nociceptors ( Figure 7G and Figure . CC-BY-NC-ND 4.0 International license made available under a (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. ...
Context 12
... our data support a mechanistic model of secondary hyperalgesia in which inflammation-induced upregulation of TMEM100 un-silences MIAs, which subsequently triggers central sensitization and -via a yet unknown central mechanism -peripheral sensitization of cutaneous nociceptors, which eventually leads to secondary hyperalgesia in skin regions remote from the site of inflammation ( Figure 7H). By demonstrating that primary and secondary hyperalgesia are triggered by separate subclasses of primary sensory afferents and considering that MIAs constitute almost fifty percent of all nociceptors in viscera and deep somatic tissues, our study provides an invaluable framework for future studies that aim at deciphering the contribution of different afferent subtypes to other clinically relevant forms of pain and to develop new strategies for preventing the chronification of pain after injury and inflammation. ...
Context 13
... /2022 Supplemental Figure S4, Sensitization of cutaneous Aδ-fiber nociceptors by TMEM100 overexpression in articular afferents, related to Figure 7 (A), Example traces of mechanically-evoked action potentials recorded from cutaneous Aδ-fiber nociceptors in the tibial nerve from control mice (top, AAV-PHP.S-dsRed) and from mice that overexpress TMEM100 in articular afferents (bottom, AAV-PHP.S-TMEM100-Ires-dsRed). (B), Comparison of the firing rates evoked by ramp-and-hold stimuli that exerted the indicated force to the receptive fields. Symbols represent means ± SEM numbers of action potentials, which were compared using multiple Mann-Whitney tests (ns, not significant; *, P<0.05; **, P<0.01). ...

Similar publications

Article
Full-text available
Mechanically silent nociceptors are sensory afferents that are insensitive to noxious mechanical stimuli under normal conditions but become sensitized to such stimuli during inflammation. Using RNA-sequencing and quantitative RT-PCR we demonstrate that inflammation upregulates the expression of the transmembrane protein TMEM100 in silent nociceptor...

Citations

... Paw lifting and licking were defined as positive responses. The 50% withdrawal threshold (g) was determined by fitting the response rate vs. von Frey force curves with a Boltzmann sigmoid equation with constant bottom and top constraints equal to 0 and 100, respectively [34]. The integral of response frequency-von Frey force intensity (0.008 to 0.1 g) curves was calculated as area under the curve (AUC, A.U. = arbitrary unit). ...
Article
Full-text available
Drugs enhancing the availability of noradrenaline are gaining prominence in the therapy of chronic neuropathic pain. However, underlying mechanisms are not well understood, and research has thus far focused on α2-adrenergic receptors and neuronal excitability. Adrenergic receptors are also expressed on glial cells, but their roles toward antinociception are not well deciphered. This study addresses the contribution of β2-adrenergic receptors (β2-ARs) to the therapeutic modulation of neuropathic pain in mice. We report that selective activation of β2-ARs with Formoterol inhibits pro-inflammatory signaling in microglia ex vivo and nerve injury-induced structural remodeling and functional activation of microglia in vivo. Systemic delivery of Formoterol inhibits behaviors related to neuropathic pain, such as mechanical hypersensitivity, cold allodynia as well as the aversive component of pain, and reverses chronically established neuropathic pain. Using conditional gene targeting for microglia-specific deletion of β2-ARs, we demonstrate that the anti-allodynic effects of Formoterol are primarily mediated by microglia. Although Formoterol also reduces astrogliosis at late stages of neuropathic pain, these functions are unrelated to β2-AR signaling in microglia. Our results underline the value of developing microglial β2-AR agonists for relief from neuropathic pain and clarify mechanistic underpinnings.
Article
Full-text available
Introduction Pain is a common comorbidity in patients with hemophilia (PwH) due to hemophilic arthropathy. This study aims to explore pain sensitivity in PwH methodologically investigating in cuff pressure testing compared to algometer testing. Methods 37 PwH and 35 healthy control subjects (Con) enrolled in this study. Joint health status was assessed. Subjective pain was evaluated using numeric rating scales. Pain sensitivity was measured with pressure algometry and cuff pressure algometry. Pressure pain thresholds of the algometer (PPTa) were measured at knee, ankle joints, and forehead. Subsequently, thresholds of cuff pressure were measured at the left and right lower legs (PPTcuff). In both, lower values represent higher pain sensitivity. Results PwH exerted a worse joint health status than Con. Pain sensitivity was higher in PwH compared to Con as PPTa of the knee and ankle joints were lower in PwH. No difference was observed in PPTa at the forehead. Contrastingly, lower pain sensitivity was detected in PwH by higher PPTcuff values compared to Con in both legs. Conclusion While PPTa of the knee and ankle joints are lower in PwH, PPTcuff are higher in PwH compared to Con. This reveals a paradox situation, highlighting that PwH experience local, joint- and hemophilic arthropathy-related pain, whereas pain sensitivity of non-affected soft tissue structures is lower. The reasons explaining the PPTcuff results remain elusive but might be explained by coping strategies counteracting chronic joint pain, resulting in lower sensitivity at non-affected structures.