Skip to main content
Advertisement
  • Loading metrics

Nicotine induces abnormal motor coupling through sensitization of a mechanosensory circuit in Caenorhabditis elegans

  • Yuting Liu ,

    Contributed equally to this work with: Yuting Liu, Leiru Huang, Ruipeng Wang

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Leiru Huang ,

    Contributed equally to this work with: Yuting Liu, Leiru Huang, Ruipeng Wang

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Ruipeng Wang ,

    Contributed equally to this work with: Yuting Liu, Leiru Huang, Ruipeng Wang

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Fukang Qi,

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Yiwen Liu,

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Qingyuan Chen,

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Morgane Mialon,

    Roles Resources

    Affiliation Melis, Universite Claude Bernard Lyon 1, CNRS UMR5284, INSERM U1314, Institut NeuroMyoGene—Faculte de Medecine et de Pharmacie, Lyon, France

  • Lili Chen,

    Roles Investigation

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

  • Berangere Pinan-Lucarre,

    Roles Resources, Writing – review & editing

    Affiliation Melis, Universite Claude Bernard Lyon 1, CNRS UMR5284, INSERM U1314, Institut NeuroMyoGene—Faculte de Medecine et de Pharmacie, Lyon, France

  • Shangbang Gao

    Roles Conceptualization, Funding acquisition, Investigation, Supervision, Validation, Writing – original draft, Writing – review & editing

    sgao@hust.edu.cn

    Affiliation Key Laboratory of Molecular Biophysics of the Ministry of Education, College of Life Science and Technology, Huazhong University of Science and Technology, Wuhan, China

Abstract

Nicotine exposure elicits diverse behavioral changes, yet the underlying neural pathways and molecular mechanisms remain incompletely understood. Here, we demonstrate that chronic nicotine exposure markedly increases both the initiation and duration of reversals in Caenorhabditis elegans. Strikingly, these reversals were tightly coupled with the rhythmic body contractions of the defecation motor program (DMP). Through pharmacological, genetic, in situ electrophysiological, and calcium imaging analyses, we show that nicotine enhances the activity of the AVA interneuron via selective upregulation of ACR-16, a nicotinic ACh receptor critical for nicotine-induced motor coupling. Ablation of touch receptor neurons (TRNs) or inhibition of TRNs-mediated mechanosensation completely abolished this coupling. Furthermore, optogenetic activation of TRNs in nicotine-treated animals evoked stronger AVA depolarization, and nicotine amplified gentle touch-evoked reversals. Together, these findings reveal a potential interoceptive effect of nicotine mediated by sensitization of the TRNs-AVA mechanosensory pathway, providing new insight into the neural and molecular basis of nicotine’s modulation of sensory-motor coupling.

Introduction

Smoking poses severe health risks, and nicotine—the primary psychoactive component of tobacco—exerts potent rewarding effects by upregulating nicotinic acetylcholine receptors (nAChRs) in dopaminergic neurons, particularly within brain regions such as the ventral tegmental area [13]. This receptor upregulation drives dopamine release, a key mechanism underlying nicotine addiction [46]. During withdrawal, the sudden absence of nicotine reduces dopamine release, contributing to a range of neurological and psychiatric symptoms [7]. Beyond dopamine dysregulation, additional mechanisms—including elevated corticotropin-releasing factor activity [8] and altered GABAergic and glutamatergic transmission [9,10]—also shape nicotine’s effects. These changes underlie withdrawal symptoms such as cravings and mood disturbances, which often lead to relapse and complicate smoking cessation.

Chronic heavy smokers, exposed to sustained high nicotine concentrations, frequently experience more severe withdrawal symptoms [11,12]. Recent evidence indicates that activation of upstream GABAergic neurons plays a crucial role in regulating dopamine release and driving nicotine-dependent behavioral states [1315]. Given the profound health consequences of smoking and incomplete understanding of the neural mechanisms underlying high-dose nicotine effects, it is essential to further dissect the neural circuits and molecular targets mediating these responses.

The nematode Caenorhabditis elegans displays well-defined behavioral responses to nicotine, including acute reactions, adaptation, and withdrawal-like symptoms [1622]. Similar to mammals, nicotine responses in C. elegans are dose-dependent [16,18]. At low concentrations (0.5–5 μM), worms exhibit acute nicotine responses and withdrawal symptoms upon cessation [16,18]. At moderate (0.1 mM) to high (1 mM) concentrations, nicotine alters locomotion speed [18], while at extremely high concentrations (>20 mM), C. elegans undergoes behavioral adaptation and eventual paralysis [18,20,22]. Low-dose nicotine responses require nAChRs, which are linked to TRP channels [16], and strikingly, these behaviors can be rescued in mutants by expressing the mammalian α4β2 nAChR, underscoring the evolutionary conservation of nicotine’s effects. In contrast, resistance to high-dose nicotine has been linked to genes such as tax-6, nra-1, and soc-1 [20]. Prolonged exposure to extremely high nicotine concentrations (30–31 mM) reduces nicotinic receptor abundance and suppresses egg-laying [22]. Despite these advances, the neural circuits and molecular mechanisms underlying high-dose nicotine-induced behavioral modulation remain poorly defined.

In this study, we report that high-concentration nicotine (1 mM) induces marked locomotor alterations in C. elegans, including frequent reversals and increased locomotion speed. Unexpectedly, these reversals are strongly coupled to the rhythmic body contractions of the defecation motor program (DMP), suggesting enhanced interoceptive sensitivity to internal mechanical state. High-dose nicotine selectively upregulates the α7-like nAChR ACR-16 [23] in the AVA command interneuron, which mediates the coupling between DMP and reversals. Mechanosensory input from touch receptor neurons (TRNs) is indispensable for this behavior: TRNs ablation or mechanosensation blockade completely abolished the nicotine-induced coupling. Moreover, nicotine enhanced TRN-to-AVA signal strength and prolonged gentle touch-evoked reversals. Together, these findings reveal that nicotine induces abnormal motor coupling via sensitization of a mechanosensory neural circuit in C. elegans, offering insight into how nicotine might trigger anxiety-like or aversive behaviors through altered sensory-motor integration.

Results

Nicotine increases reversal frequency in C. elegans

To investigate the behavioral effects of nicotine, we developed a brief chronic nicotine exposure assay in which young adult hermaphrodites were incubated with 1 mM nicotine for 3 h (Fig 1A; Materials and methods). Following exposure, animals were transferred to nicotine-free Nematode Growth Medium (NGM) agar plates seeded with a thin layer of Escherichia coli OP50 for behavioral analysis. Under these conditions, worms exhibited a robust nicotine-dependent behavioral response [24].

thumbnail
Fig 1. Nicotine increased the coupling of DMP and reversals.

(A) An assay diagram of response to nicotine exposure and withdrawal in Caenorhabditis elegans. Young adult hermaphrodites were used. (B) Representative 3-min locomotion trajectories of two animals on withdrawal plates without (− nic) and with (+ nic) nicotine exposure. (C) Coupling time diagrams for reversal and defecation motor program (DMP). Upper, representative speed traces labeled with the moment of DMP setup. Gray bar: DMP not couple with reversal; Orange bar: DMP couples with reversal. Bottom, raster plots of 10 consecutive DMP cycles that coupled with or without reversal in 10 animals. (D) Quantification of the DMP-reversal coupling proportion. Nicotine, but not glutamate, significantly enables the coupling of DMP to reversal. Two-way ANOVA was performed (interaction: F(1, 37) = 225.9, P < 0.0001). (E) 100 segmented DMP-associated movements, each shown for 30 s. The 0-s time point represents the initiation of the DMP and serves as the alignment reference. The speed of 10 s before and 20 s after DMP initiation were used. The figure below shows the average speed of the 100 segments. (F) The dose dependence of nicotine (3 h) for the coupling of DMP and reversal. (G) The time dependence of nicotine (1 mM) for the coupling of DMP and reversal. (H) The failure time dependence of nicotine. n ≥ 10 animals. ns, no significance, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.g001

Compared with nicotine-naïve controls (− nic), nicotine-exposed worms (+ nic) displayed disrupted forward locomotion, characterized by discontinuous crawling trajectories, markedly increased reversal frequency, and elevated locomotion speed (Fig 1B). The most prominent difference was the substantial increase in reversal frequency and movement velocity. Whereas spontaneous backward movements in control worms were brief and sporadic, nicotine-exposed animals exhibited reversals that were highly rhythmic, significantly prolonged, and executed at faster speeds (S1A Fig). Consequently, the proportion of time spent moving backward was markedly greater in the nicotine-treated group (S1B Fig). Quantitatively, reversal frequency increased nearly 3-fold (− nic, 3.3 ± 0.34/3 min, n = 22; + nic, 9.3 ± 0.46/3 min, n = 22, p < 0.001) (S1C Fig), reversal duration was extended more than 4-fold (− nic, 1 ± 0.08 s, n = 22; + nic, 4.7 ± 0.27 s, n = 22, p < 0.001) (S1D Fig), and reversal speed was significantly elevated (− nic, 0.09 ± 0.003 mm/s, n = 22; + nic, 0.12 ± 0.003 mm/s, n = 22, p < 0.001) (S1E Fig). Meanwhile, forward locomotion speed was also enhanced (− nic, 0.11 ± 0.003 mm/s, n = 22; + nic, 0.13 ± 0.004 mm/s, n = 22, p < 0.001) (S1F Fig). These high-frequency, prolonged, high-speed reversals resemble nicotine-induced anxious or aversive behaviors in mammals [25,26].

Nicotine-enhanced reversals are temporally coupled with the defecation motor program

Nicotine-induced behavior changes have been reported across multiple model organisms [26,27]. Here, we observed that nicotine-evoked high-frequency reversals in C. elegans exhibited a striking rhythmicity. These reversals occurred approximately every 45–50 s, prompting us to consider whether they might be linked to an intrinsic rhythmic process. We analyzed locomotion over 10 consecutive cycles of the DMP in wild-type animals. In controls, ~ 20% of DMP cycles were followed by short backward movements (Fig 1C and 1D and S1 Movie), consistent with previous reports of DMP-locomotion coupling [28]. By contrast, nicotine drove nearly 100% of DMP cycles to trigger reversals immediately following DMP onset, indicating robust DMP-reversal coupling (Fig 1C and 1D and S2 Movie). This suggested that nicotine’s effect arises from modulation of intrinsic rhythmic signaling, rather than simply producing general excitatory drive.

To distinguish between these possibilities, we examined the effect of glutamate—another neurotransmitter implicated in reward regulation [29] and nicotine dependence [10]—under identical conditions. Glutamate exposure did not alter the proportion of DMP-reversal coupling (Fig 1D). Furthermore, although nicotine affects gastrointestinal motility in mammals [30], the DMP cycle length in C. elegans remained unchanged after nicotine exposure (S2AS2C Fig). These results suggest that nicotine-induced motor coupling does not stem from altered gut rhythmicity or nonspecific excitation.

Interestingly, nicotine-enhanced couplings showed <5 s delays between DMP onset and reversal initiation, in contrast to previous observations in which reversals typically occurred ~10 s before DMP [28]. The DMP consists of three sequential phases: posterior body wall muscle contraction (pBoc), followed by anterior body wall muscle contraction (aBoc), and the final enteric muscle expulsion (Exp) [31] (S2A Fig). To determine whether nicotine alters the coupling temporal relationship between DMP phases and reversal, we categorized reversal onset times relative to these phases and found that nicotine-driven reversals occurred predominantly during pBoc (S2D Fig).

We next assessed coupling in mutants with phase-specific DMP defect. In egl-8 mutants (defective pBoc), nicotine-enhanced coupling was strongly reduced, whereas nlp-40 and aex-2 mutants (impaired Exp) [31,32] showed no significant change (S2E Fig). aex-3 mutants, defective in both aBoc and Exp [31,33], exhibited partial reduction, but less than egl-8. These results indicate that intact pBoc activity is critical for nicotine-enhanced motor coupling, consistent with nicotine amplifying an interoceptive response to body wall contraction rather than altering DMP timing. In agreement, the delay from DMP onset to reversal initiation was unchanged by nicotine (S2F Fig).

To more precisely quantify the timing, we aligned all DMP events to the moment of minimal body length (time 0) and extracted locomotion speed traces spanning 10 s before to 20 s after this point (Fig 1Ei). This alignment allowed us to precisely synchronize locomotion speed at the moment of DMP, facilitating a comprehensive analysis of the locomotion characteristics associated with 100 DMP events across 10 worms. By aligning the data in this way, we ensured that variations in DMP cycle timing did not obscure the underlying coupling patterns, thereby improving the accuracy and reliability of our analysis. We found that nicotine induced a pronounced backward movement immediately following DMP compared to controls (Fig 1Eii).

Nicotine-enhanced DMP-reversal coupling was both concentration and exposure-time-dependent. With a fixed 3-h exposure, coupling increased with nicotine concentration, yielding a half-maximal effective concentration (EC50) of 0.34 mM (Fig 1F). At a fixed saturated dose (1 mM), coupling increased with exposure duration, with a half-maximal effect at ~15.2 min (Fig 1G). Furthermore, similar to how nicotine induces a reversible aversive reaction in humans [34], nicotine-enhanced coupling was reversible: after transfer to nicotine-free plates, coupling gradually returned to baseline (~20%) with a half-life of ~1.2 h (Fig 1H).

Given that DMP is an intrinsic bodily rhythm, these findings suggest that nicotine acts as an interoceptive modulator, selectively enhancing sensory-motor linkage between internal rhythmic contractions and reversal behavior [3537]. This prompted us to investigate the molecular and circuit-level mechanisms underlying this phenomenon.

Nicotine-induced abnormal motor coupling requires the nicotinic acetylcholine receptor ACR-16

Previous studies have shown that nicotine-dependent behaviors require nAChRs [5,38], which mediate responses to the neurotransmitter acetylcholine (ACh) [39,40]. To test whether nAChRs mediate nicotine’s effects in our assay, we pharmacologically blocked these receptors using DHβE, a competitive antagonist of nAChRs [40]. DHβE treatment significantly suppressed nicotine-enhanced DMP-reversal coupling (Fig 2A). Consistently, other nicotine-induced motor behaviors—including increased reversal frequency (S3A Fig), prolonged reversal duration (S3B Fig), and enhanced reversal and forward speeds (S3C and S3D Fig)—were also inhibited. In contrast, a cocktail of glutamate receptor antagonists, DNQX and MK-801, which block ionotropic excitatory glutamate receptors [41,42], did not affect the coupling proportion (Fig 2B). These pharmacological results provide strong evidence that nicotine’s action in C. elegans depends on nAChRs rather than glutamate receptors.

thumbnail
Fig 2. Nicotine enhanced DMP-reversal coupling requires ACR-16 in AVA.

(A) Suppression of nicotine enhanced DMP-reversal coupling by nAChR antagonist DHβE (20 µM). Two-way ANOVA was performed (interaction: F(1, 39) = 275.6, P < 0.0001). (B) The GluR antagonists DNQX (500 µM) and MK-801 (50 µM) do not affect nicotine-enhanced DMP reverse coupling. Two-way ANOVA was performed (interaction: F(1, 36) = 1.619, P = 0.2114). (C) Dendrogram of Caenorhabditis elegans nAChRs. A fraction of nAChRs from Drosophila, mouse, and humans were included, such as mouse α7. (D) Genetic screening identified a specific nAChR gene, acr-16, disrupts nicotine-enhanced DMP-reversal coupling. Two-way ANOVA was performed (interaction: F(21, 437) = 12.95, P < 0.0001). (E) Expression pattern of acr-16. Up, endogenous acr-16 promoter driven GFP was observed in AVA command interneuron and body wall muscles. Scale bar, 20 µm. Bottom, the co-localization of Pacr-16::GFP and Ptwk-40s::RFP in AVA neuron. Scale bar, 10 μm. Worm orientation, A anterior, D dorsal. (F) Quantification of the rescue of DMP-reversal coupling by expression of ACR-16 in different tissues. AVA, AVE, etc. command interneurons, but not motor neurons or muscles, are the critical functional cells for ACR-16. Expression of mouse α7 nAChR restored the DMP-reversal coupling proportion in acr-16(ok789) mutants. Two-way ANOVA was performed (interaction: F(6, 129) = 34.81, P < 0.0001). (G) Overexpression of ACR-16 in wild-type N2 increases the DMP-reversal coupling proportion even without nicotine exposure. Two-way ANOVA was performed (interaction: F(1, 70) = 22.05, P < 0.0001). n ≥ 10 animals. ns, no significance, *** p < 0.001 by Two-way ANOVA analysis. # p < 0.05 by Student t test. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.g002

To further validate this, we performed a genetic screen focusing on nAChR mutants. The C. elegans genome encodes over 29 nAChR subunits [43], many with available loss-of-function alleles (Fig 2C) [16,44,45]. Among these, only acr-16 mutants completely lost the nicotine-enhanced DMP-reversal coupling, with nicotine failing to induce motor coupling (Fig 2D and S3 and S4 Movies). Other uncoordinated mutants such as unc-29, unc-38, and unc-63 showed reduced coupling proportions (S3E Fig), although their severe locomotion defects may nonspecifically affect the behavior [46,47]. In agreement with the pharmacological data, mutants lacking ionotropic glutamate receptors like glr-1, glr-2, glr-4, glr-5, nmr-1, and nmr-2, displayed normal DMP-reversal coupling (S3F Fig).

Together, these findings demonstrate that nicotine-induced abnormal motor coupling specifically requires the nAChR subunit ACR-16 in C. elegans, highlighting its pivotal role in mediating nicotine’s modulatory effects on sensory-motor integration.

ACR-16 acts in command interneurons to mediate nicotine-induced motor responses

acr-16 encodes an α7-like nAChR broadly expressed in both the nervous system and body wall muscles [44,45,48]. To assess how acr-16 contributes to nicotine-induced motor behavioral changes in C. elegans, we examined its expression pattern and functional site of action. Using a fluorescence reporter driven by the endogenous acr-16 promoter, we detected expression in both neuronal and muscular tissues (Fig 2E). Within the nervous system, acr-16 is predominantly expressed in AVA and AVB command interneurons and in a subset of motoneurons, all of which are critical for regulating motor behaviors [45,49,50]. In mammals, nicotine’s behavioral effects are primarily mediated by neuronal, rather than muscle, nAChRs [40]. To pinpoint the neuronal site of acr-16 action, we performed cell-specific rescue experiments in acr-16 mutants. Restoring ACR-16 expression in the backward command interneurons AVA and AVE using the nmr-1 promoter fully rescued the nicotine-enhanced DMP-reversal coupling (Fig 2F). In contrast, expression in body wall muscles (myo-3 promoter), backward motor neurons (unc-4 promoter), or forward command interneurons such as AVB and PVC (sra-11 promoter) failed to restore nicotine responsiveness. These results indicate that ACR-16 acts primarily in the backward command interneurons to mediate nicotine-induced motor coupling.

Interestingly, overexpression of ACR-16 in AVA not only rescued the nicotine response in acr-16 mutants but also elevated baseline DMP-reversal coupling in the absence of nicotine (Fig 2F). This suggests that the expression level of ACR-16 in AVA may set the threshold for coupling. To test this, we overexpressed ACR-16 in AVA using the short twk-40s promoter [51] in wild-type N2 worms. Indeed, even without nicotine, these animals exhibited increased baseline coupling, and nicotine exposure drove the coupling proportion to 100% (Fig 2G).

Together, these findings demonstrate that ACR-16 mediates nicotine’s behavioral effects primarily via backward command interneurons, and that the strength of this response is strongly dependent on ACR-16 expression levels in these neurons.

The Mouse nAChR α7 subunit can functionally substitute for worm ACR-16

Given the essential role of ACR-16 in nicotine-dependent behavioral response in C. elegans, we asked whether the mammalian homolog could serve a similar function. ACR-16 shares high sequence similarity with the vertebrate α7 subunit of nAChRs (Fig 2C) [44,45]. To test cross-species functional compatibility, we expressed the mouse α7 nAChR in acr-16 mutant worms under the control of the endogenous acr-16 promoter. Remarkably, mouse α7 expression restored nicotine-enhanced DMP-reversal coupling in the mutants (Fig 2F). However, unlike worm ACR-16, the mouse α7 receptor did not increase baseline coupling proportion in the absence of nicotine (Fig 2F). These findings suggest that α7/acr-16-dependent nicotine responses are conserved between mammals and C. elegans.

Nicotine enhances AVA calcium activity linked to the DMP

To establish the physiological role of AVA interneuron in nicotine’s interoceptive effects, we measured real-time AVA activity in freely moving animals using the genetically encoded calcium (Ca2+) indicator GCaMP6s, with wCherry as a ratiometric reference [52] (Materials and methods). These AVA::GCaMP transgenic animals responded to nicotine similarly to wild-type N2 worms in terms of reversal frequency, duration, and locomotion speed (wild-type, S4 Fig). In parallel, we used an independent infrared imaging system to monitor the rhythmic defecation behavior in real time [53] (Materials and methods). AVA Ca2+ activity was recorded across five consecutive DMP cycles to assess their synchrony.

Following nicotine exposure, AVA displayed significantly elevated Ca2+ oscillation compared to nicotine-free conditions (Fig 3A). Both the frequency and amplitude of Ca2+ transients were markedly increased in the wild-type transgenic worms (Fig 3A3D), indicating that nicotine enhances AVA excitability. This nicotine-induced AVA activation was strongly temporally coupled to DMP events. To precisely define this relationship, we aligned AVA Ca2+ traces to the shortest body length point of each DMP cycle, extracting activity from 10 s before to 20 s after this reference point (Fig 3A). This alignment allowed us to synchronize AVA activity with 100 DMP events across 20 worms. Under nicotine treatment, nearly every DMP event was accompanied by a robust AVA Ca2+ transient (Fig 3B and 3C), whereas in the absence of nicotine, DMP initiation elicited minimal AVA activity (Fig 3B). These results parallel the behavioral observation that nicotine enhances DMP-triggered reversals, suggesting that elevated AVA activity drives this coupling.

thumbnail
Fig 3. Nicotine enhances calcium activity in AVA.

(A) Left, A representative calcium (Ca2+) transient of AVA neurons over five consecutive DMP cycles from a freely behaving animal without nicotine treatment. Only one DMP triggers Ca2+ activity in AVA. Right, Nicotine significantly enhances spontaneous Ca2+ transient activity in AVA, with all five DMP cycles initiating Ca2+ activity in AVA. Green boxes and arrows indicate tight coupling between DMP and AVA Ca2+ transients, while gray boxes and arrows represent the absence of coupling. Dashed boxes highlight 30 s segments of AVA Ca2+ activity, with 10 s before and 20 s after DMP initiation. Arrows indicate the initiation moments of DMP. (B) The AVA Ca2+ activity from five consecutive DMP cycles, recorded from 20 individual worms, is shown for both nicotine-free (− nic) and nicotine-exposed (+ nic) conditions across different genotypes (n = 20 animals per genotype). The Ca2+ transients are aligned based on the DMP initiation moments (indicated by arrows and dashed lines). Numbers 1 through 5 correspond to the five consecutive DMP cycles. The average AVA Ca2+ response for each DMP is displayed below. (C) Quantitative analysis of the coupling percentage of AVA Ca2+ transient responses to DMP. Nicotine enhanced coupling proportion of DMP-AVA Ca2+ transients is dependent on ACR-16. Two-way ANOVA was performed (interaction: F(2, 114) = 83.67, P < 0.0001). (D) Quantification of the peak amplitude of AVA Ca2+ transients that coupled with DMP. Nicotine increased the amplitude of AVA Ca2+ transients, which is dependent on ACR-16. Two-way ANOVA was performed (interaction: F(2, 277) = 26.60, P < 0.0001). n = 20 animals in each genotype. ns, no significance, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.g003

Consistent with behavioral requirements, AVA Ca2+ activity was nearly absent in acr-16 mutants under nicotine exposure (Fig 3B3D). In these mutants, nicotine failed to increase either the response probability or the peak amplitude of AVA transients. Cell-specific rescue of ACR-16 in AVA using the short twk-40s promoter fully restored both the frequency and amplitude of DMP-evoked Ca2+ transients (Fig 3B3D). Notably, ACR-16 rescue also modestly increased baseline AVA activity in the absence of nicotine, mirroring the enhanced baseline DMP-reversal coupling seen in ACR-16-overexpressing worms. Furthermore, the ACR-16 dependence of nicotine-induced increases in reversal frequency, duration, and locomotion speed was confirmed in these rescue animals (S4 Fig). Together, these findings indicate that nicotine enhances motor coupling by elevating AVA activity in an ACR-16-dependent manner.

Nicotine upregulates functional expression of ACR-16

Having established that enhanced AVA activity is crucial for nicotine’s action and that this regulation depends on ACR-16, we next aimed to identify how ACR-16 mediates this activity. In mammals, chronic nicotine exposure is well known to up-regulate nAChRs, contributing to addiction and withdrawal responses [54,55]. Given the specific requirement of ACR-16 for both the behavioral and Ca2+ responses we observed, we hypothesized that nicotine might increase the functional expression of ACR-16 in AVA, thereby driving the enhanced coupling.

To test this, we generated a transgenic C. elegans strain expressing functional ACR-16 cDNA tagged with a fluorescent marker (SL2d::GFP) in AVA neuron in wild-type N2 worms. Quantification of GFP fluorescence intensity before and after nicotine treatment revealed a significant increase following nicotine exposure (S5A and 5B Fig). To avoid potential artifacts arising from overexpression or non-specific promoters, we employed CRISPR-Cas9 genome editing to generate an AVA-specific ACR-16::GFP knock-in strain (EN9002) that preserves endogenous expression levels [50,56,57]. In this strain, an spGFP11 tag was inserted into the acr-16 locus, spGFP1–10 was expressed specifically in AVA using a Cre-loxP-based promoter combination (Materials and methods). ACR-16::GFP signal was enriched in the AVA soma, with punctate labeling along the neurite (Fig 4A). Consistently, nicotine exposure significantly enhanced ACR-16::GFP fluorescence in both soma and neurite compartments (Fig 4A4C), indicating that nicotine upregulates ACR-16 expression in AVA.

thumbnail
Fig 4. Upregulation of ACR-16 is responsible for the nicotine-enhanced DMP-reversal coupling.

(A) Nicotine up-regulates ACR-16 expression in AVA. Using fluorescently-tagged ACR-16 based on split GFP reconstitution, we inserted three copies of the spGFP11 (split GFP11) sequence into the acr-16::aid::Scarlet locus via CRISPR/Cas9 (ACR-16-AID-Scarlet-spGFP11) and expressed the complementary spGFP1-10 (split GFP1-10) moiety in the AVA to generate AVA specific ACR-16::GFP transgenic line. Scale bar, 10 µm. (B, C) After nicotine exposure, the fluorescence intensity in both AVA soma and neurite are significantly increased. (D) Up, A schematic diagram of in situ whole-cell recording of Caenorhabditis elegans AVA under voltage clamping configuration (V Clamp). Bottom, Representative currents evoked by acetylcholine (1 mM) (I ACh) in different genotypes without (− nic) or with (+ nic) nicotine exposure. The purple arrows indicate the dosing time. AVA was hold at −60 mV. (E) Quantification of peak currents across different genotypes. Two-way ANOVA was performed (interaction: F(2, 42) = 4.105, P = 0.0235). n ≥ 6 animals. ns, no significance, * p < 0.05, ** p < 0.01, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.g004

The behavioral, Ca2+ imaging, and expression results all point to a potential mechanism in which nicotine upregulates ACR-16, thereby enhancing AVA activity and promoting reversals. To directly test this idea, we performed in situ whole-cell patch-clamp recordings from dissected AVA neurons to measure ACh-evoked currents (Fig 4D). In untreated wild-type animals, ACh (1 mM) elicited robust inward currents (Fig 4D), confirming the presence of functional AChRs in AVA. Strikingly, nicotine-pretreated animals displayed significantly larger ACh-evoked currents (Fig 4D and 4E). Application of the nAChR-selective antagonist DHβE strongly attenuated this nicotine-induced potentiation (S5C and S5D Fig). These results indicate that AVA not only expresses functional ionotropic AChRs [49], but that their expression may also be enhanced following nicotine exposure. In light of our previous findings identifying a specific ACR-16 dependence, we next examined whether ACR-16 was required for this potentiation. In acr-16 mutants, nicotine exposure failed to increase ACh-evoked currents (Fig 4D and 4E). Reintroducing ACR-16 in AVA restored the nicotine-induced enhancement (Fig 4D and 4E). These findings suggest that nicotine-enhanced ACh current in AVA is also mediated by ACR-16.

Notably, in nicotine-free conditions, acr-16 mutants exhibited smaller ACh currents than wild-type worms, and this deficit was rescued by AVA-specific ACR-16 expression—consistent with the reduced DMP-reversal coupling observed in these mutants in the absence of nicotine (Fig 4D and 4E). Importantly, neither nicotine exposure, acr-16 mutation, nor DHβE application significantly altered AVA resting membrane potential or voltage-dependent conductance (S6 Fig), suggesting that the enhanced Ca2+ activity arises from increased synaptic input rather than changes in intrinsic excitability.

Touch receptor neurons mediate nicotine-induced abnormal motor coupling

What signaling pathway links the DMP-reversal coupling with enhanced AVA Ca2+ activity? The C. elegans DMP involves rhythmic release of transmitters (e.g., H+) and peptides (e.g., NLP-40) from the intestine, which trigger strong contractions of body wall and enteric muscles [32,58]. One possibility is that these signaling molecules directly stimulate AVA, leading to enhanced activity. However, the receptors for H+ and NLP-40 are neither expressed nor functional in AVA [32,59], making direct activation unlikely. An alternative hypothesis is that the forceful muscle contractions during DMP generate mechanical stimuli that activate mechanosensory, proprioceptive, or nociceptive neurons, which in turn influence AVA activity. To test this, we systematically screened a range of sensory-deficient mutants. Worms lacking sensory receptors in nociceptive neurons (ASH, OLQ, ALA, IL1), including deg-1, tmc-1, trpa-1, and osm-9 [6062], displayed normal nicotine-enhanced DMP-reversal coupling (Fig 5A). Similarly, mutants deficient in major proprioceptive receptors (trp-1, trp-2, trp-4) expressed in DVA [16,63] or SMD neurons [63] also retained robust coupling in response to nicotine (Fig 5A). These results suggest that nociceptive and proprioceptive neurons are not required for the effect.

thumbnail
Fig 5. Exteroceptive TRNs are essential for nicotine enhanced DMP-reversal coupling.

(A) Suppression of nicotine enhanced DMP-reversal coupling in mec-2, mec-4, mec-6 mutation worms. Neuronal specific expression of MEC-6 restored nicotine enhanced DMP-reversal coupling in mec-6 mutants. The coupling proportion was unaffected by mutations in genes essential for harsh touch, proprioception, and nociceptive functions. Two-way ANOVA was performed (interaction: F(12, 217) = 19.72, P < 0.0001). (B) Up, Blue light irradiation of TRNs expressing GtACR2 completely blocked nicotine-enhanced DMP-reversal coupling. The numbers 1 through 3 refer to three consecutive DMP cycles. Gray bar: DMP not couple with reversal; Orange bar: DMP couples with reversal. Bottom, Quantification of the proportion in transgenic TRN::GtACR2 animals without and with blue light after nicotine exposure. (C) Up, A schematic diagram of ablation of TRNs by miniSOG. Bottom, TRNs injury abolished nicotine enhanced DMP-reversal coupling. Two-way ANOVA was performed (interaction: F(1, 25) = 87.52, P < 0.0001). n ≥ 5 animals. ns, no significance, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.g005

Strikingly, mutants lacking key components of the gentle-touch mechanosensory channel complex—mec-2, mec-4, or mec-6—exhibited a profound loss of nicotine sensitivity, such that nicotine exposure no longer induces significant DMP-reversal coupling (Fig 5A). These mec genes encode essential mechanosensory channel complexes in TRNs (ALM, AVM, PLM, and PVM) [64] and did not significantly affect the DMP cycle length (S7A Fig). Furthermore, analysis of mec-4(e1611) mutants showed no significant changes in reversal frequency under either on-food or off-food conditions (S7B Fig) [65], indicating no locomotory defects that would interfere with the analysis of motor coupling. In mec-6 mutants, reintroducing MEC-6 specifically into TRNs fully restored nicotine-enhanced coupling (Fig 5A). By contrast, other mechanosensory neurons, such as PVD and FLP, which mediate harsh-touch sensation and are affected by mec-10 mutations, showed no changes in coupling (Fig 5A). Together, these results indicate that mechanosensory input from TRNs is essential for nicotine’s effect on abnormal motor coupling.

To further confirm this, we performed optogenetic inhibition of TRNs using GtACR2, a light-gated inhibitory chloride channel [66,67]. Light illumination completely suppressed nicotine-enhanced DMP-reversal coupling (Fig 5B). Likewise, targeted ablation of TRNs using neuron-specific expression of miniSOG (mini singlet oxygen generator) [68] (S7C Fig)—which induces acute functional loss and subsequent neuronal death upon photoactivation—abolished the nicotine effect (Fig 5C). In contrast, ablation of the proprioceptive neuron DVA, previously reported to sense the body contraction [69], had no effect on the coupling (S7D Fig). These results demonstrate that TRNs are indispensable for mediating nicotine’s action.

Nicotine enhances the reversal sensitivity to gentle touch

TRNs are well-characterized mechanosensory neurons responsible for detecting exteroceptive gentle touch stimuli [70]. The finding that TRNs mediate nicotine-enhanced interoceptive response raises the question of whether nicotine might also affect exteroceptive sensitivity. To examine this, we applied gentle-touch stimulation to the anterior body using an eyelash probe [71]. Wild-type animals exhibited rapid and sustained reversals to touch, with an average reversal speed of 0.104 ± 0.007 mm/s and a reversal duration of 4.9 ± 0.4 s (Fig 6A6C). Remarkably, in nicotine-treated worms, the touch elicited a faster reversal speed (+ nic, 0.136 ± 0.006 mm/s) (Fig 6A and 6B), and a prolonged reversal duration (+ nic, 6.4 ± 0.5 s) (Fig 6A and 6C).

thumbnail
Fig 6. Nicotine facilitates the signal strength from TRNs to AVA.

(A) Upper, a brief diagram shows standard gentle touch evoked reversal for exteroceptive sensitivity examination. Bottom, raster plots showing individual speed responses of different genotypic animals to gentle touch on the anterior body, with and without nicotine exposure (− nic/ + nic) (n = 20 animals per group). Arrows denote gentle touch. (B, C) Quantification of the gentle touch evoked reversal speed (B) and duration (C). Two-way ANOVA was performed (interaction: F(2, 114) = 3.557, P = 0.0317 for (B); F(2, 114) = 2.581, P = 0.0801 for (C)). (D) Left, schematic diagram of in situ whole cell recording of dissected AVA neuron by optogenetically stimulating TRNs under current clamp configuration (I Clamp). Right, light-evoked excitatory post-synaptic potentials (EPSPs) under different pharmacological conditions in wild-type animals. (E) Quantification of the peak amplitude of the EPSPs. Nicotine significantly increased the signal strength in TRN-AVA circuit. ns, no significance, * p < 0.05, *** p < 0.001 by One-way ANOVA analysis (F(3, 20) = 23.87, P < 0.0001). Error bars, SEM. (F) TRNs evoked EPSPs in acr-16 mutants and rescue lines. (G) Quantification of the peak amplitude of the EPSPs in different genotypes. Nicotine increased EPSPs were reduced in acr-16 mutant animals and could be restored by expression ACR-16 back to AVA. Two-way ANOVA was performed (interaction: F(2, 30) = 3.067, P = 0.0614). n ≥ 6 animals. ns, no significance, * p < 0.05, ** p < 0.01, *** p < 0.001 by Two-way ANOVA analysis. # p < 0.05 by Student t test. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.g006

Consistently, when we examined the gentle touch-evoked reversal response in acr-16 mutants, nicotine exposure failed to enhance reversal speed (− nic, 0.096 ± 0.006 mm/s; + nic, 0.098 ± 0.005 mm/s) or duration (− nic, 4.3 ± 0.3 s; + nic, 4.6 ± 0.3 s) (Fig 6A6C). Restoring acr-16 expression in AVA rescued the nicotine-induced enhancement of reversal responses (Fig 6A6C). These results demonstrate that nicotine not only induces an anxiety/aversion-like motor coupling behavior but also heightens escape responses to mechanical stimulation via ACR-16 in AVA. Both behaviors are regulated by TRNs, highlighting the critical role of TRNs in integrating nicotine’s effects on abnormal sensory-motor coupling.

Nicotine enhances neural activity from TRNs to AVA via ACR-16

We next asked how TRNs regulate nicotine-modulated behaviors. Connectome analyses indicate that TRNs form functional neural connections with command interneurons to mediate escape responses [72]. To assess the impact of nicotine on the neural activity between TRNs and AVA, we combined optogenetics together with in vivo electrophysiological recordings [7375]. Light activation of TRNs (TRN::ChR2) with expressing the excitatory Channelrhodopsin-2 (ChR2) [76], evoked excitatory postsynaptic potentials (EPSPs) in AVA (Fig 6D). Nicotine exposure significantly increased EPSP amplitude (Fig 6D and 6E). Notably, this increase was blocked by the nicotinic receptor antagonist DHβE (Fig 6D and 6E). These results demonstrate that nicotine enhances the signal strength from TRNs to AVA.

To determine whether ACR-16 is required for nicotine-induced signal potentiation between TRNs and AVA, we analyzed the light-evoked EPSPs in acr-16 mutants. Nicotine failed to further enhance EPSPs in acr-16 mutants (Fig 6F and 6G). Reintroducing acr-16 into AVA restored nicotine-enhanced EPSP potentiation. These results denote that nicotine-potentiated EPSPs depend on ACR-16 (Fig 6F and 6G). Interestingly, in the absence of nicotine exposure, EPSPs were reduced in acr-16 mutants compared to wild-type animals (Fig 6G), suggesting that ACR-16 contributes intrinsically, at least in part, to basic neural signal from TRNs to AVA. This is consistent with our observation that ACh perfusion-evoked currents also require ACR-16.

In summary, our behavioral and in vivo electrophysiological data demonstrate that nicotine strengthens TRNs and AVA synaptic signal via ACR-16. This synaptic enhancement provides a neuronal integration mechanism for nicotine’s modulation of both interoceptive and exteroceptive responses.

Discussion

Nicotine exposure and withdrawal are known to induce a spectrum of emotional and sensory disturbances, yet the underlying mechanisms remain incompletely understood. In this study, we identify an unexpected neural pathway mediating high-dose nicotine-induced anxious/aversive-like reversal in C. elegans. We show that exteroceptive TRNs sense internal DMP, which in turn activate command interneurons to regulate the nicotine-induced behavioral response. Both TRNs mechanosensory function and the nicotine-dependent upregulation of the ACR-16 nAChR in AVA neurons are essential for this process (Fig 7). Moreover, nicotine potentiates signal strength between TRNs and AVA, revealing a striking form of sensory-motor coupling that links interoceptive and exteroceptive pathways at the circuit level. These findings broaden the current view of interoceptive processing, which is often considered to operate through dedicated, isolated ascending pathways.

thumbnail
Fig 7. Working model.

Nicotine significantly induces abnormal DMP-reversal coupling response in Caenorhabditis elegans. An unexpected mechanosensory neural pathway involving TRN-AVA circuit mediates this response. Key mechanisms include the nicotine-induced upregulation of ACR-16 nAChRs in AVA and nicotine-strengthened signal transmission within TRN-AVA circuit. TRNs may sense internal body contractions associated with the DMP, abnormally activating sensitized command interneurons to drive reversals. This finding underscores an unexpected functional integration induced by nicotine exposure.

https://doi.org/10.1371/journal.pbio.3003423.g007

Nicotine is well established to modulate internal bodily sensations such as heart rate, blood pressure, and respiratory depth in mammals [7779]. Here, we demonstrate that high-dose nicotine amplifies an apparent interoceptive signal in C. elegans by sensitizing the exteroceptive mechanosensory circuit. A similar nicotine-induced enhancement of mechanosensitivity has been reported in mice, where nicotine sensitizes cutaneous C-fiber sensory nerves [80]. While the contribution of such exteroceptive sensitization to mammalian interoception remains unclear, these parallels underscore a potentially conserved mechanism across species [81,82].

C. elegans has been reported to exhibit remarkable sensitivity to internal body states, such as muscle contractions [28,83], arousal levels [84], and intestinal activity [85,86]. Consistent with the earlier observations linking DMP to directed locomotion [28,83], our results suggest that somatic muscle contractions during the DMP can trigger reversals, and that nicotine markedly amplifies this coupling—from ~20% under baseline conditions to nearly 100% following exposure. The mechanism underlying nicotine-enhanced motor coupling appears to differ from spontaneous motor coupling. This is due to the following observations: (1) TRNs ablation does not reduce spontaneous DMP-reversal coupling; (2) spontaneous coupling persists in TRNs mechanosensory receptor-deficient mec-2, mec-4, and mec-6 mutants; and (3) in acr-16 mutants, spontaneous coupling remains present, albeit reduced, and is unaffected by nicotine. These distinctions indicate that nicotine does not simply amplify existing circuit output but instead recruits a distinct modulatory mechanism—one that could serve as a selective therapeutic target for nicotine-related behaviors without disrupting endogenous motor coupling.

At the molecular level, our data reveal that high-dose nicotine-induced coupling critically depends on ACR-16 nAChR upregulation in AVA neurons. This is distinct from low-dose nicotine effects, which involve both ACR-15 and ACR-16 [16]. While changes in other nAChR subtypes cannot be excluded [17,21,87,88], our electrophysiological and imaging results demonstrate that functional ACR-16 is indispensable. Notably, expression of mouse α7 nAChR fully restored the nicotine-induced coupling defect in acr-16 mutants, despite its inability to substitute for low-dose nicotine effects [16]. Given that α7 nAChRs mediate high-dose nicotine aversion in rodents [13,89], these findings suggest a conserved role for this receptor subtype in high-dose nicotine responses across phyla.

Our circuit analyses also raise mechanistic questions. Optogenetic activation of TRNs evoked EPSPs in AVA, which were enhanced by nicotine and blocked by DHβE, but not by glutamate receptor antagonists. Neither glutamate exposure, glutamate receptor antagonists, nor mutations in ionotropic glutamate receptors affected nicotine’s effects, implying that the observed excitatory EPSPs in the TRN-AVA circuit are not glutamatergic. Since TRNs are glutamatergic and lack expression of cholinergic markers such as cha-1 (choline acetyltransferase) and unc-17 (ACh transporter) [90], direct cholinergic transmission from TRNs to AVA seems also unlikely. A more plausible model is that the TRN-AVA functions indirectly, requiring additional ACh-releasing interneurons to relay TRN activation to AVA. Based on anatomical connectivity, we propose that AVD and AVE, both of which are cholinergic interneurons that receive strong TRNs inputs and project to AVA [91,92]. Cooperative activation of AVA, AVD, and AVE—well-known premotor interneurons for backward locomotion—could robustly drive nicotine-enhanced reversals. Additional or parallel mechanisms may also contribute, including neuropeptidergic modulation from TRNs [49,93].

Another open question is how TRNs detect DMP-associated muscle contractions. The spatial organization of TRNs suggests that posterior contractions should preferentially activate PLM and PVM, which typically promote forward movement, rather than the reversals we observed. This discrepancy implies the existence of alternative contractile signal transmission routes or recruitment of other cholinergic neurons downstream of TRNs to bias AVA activation. Resolving these circuit-level details will be key to fully understanding nicotine’s modulation of interoceptive–exteroceptive integration.

An important consideration in our study is that residual nicotine within the organism may constitute a lingering pharmacological state capable of directly influencing behavior. While we interpret these effects within a broad framework of interoception, nicotine may modulate motor coupling through multiple mechanisms—either via passive pharmacological action or by engaging specific internal sensors that shape interoceptive processing. Thus, nicotine-enhanced coupling likely reflects a combination of residual pharmacological effects and bona fide interoceptive sensing.

Finally, we find that nicotine upregulates ACR-16 expression in AVA under both heterologous (Ptwk-40s) and endogenous (Pacr-16, CRISPR knock-in) promoter control. The similarity of these effects suggests a complex regulatory process that may involve both transcriptional and post-transcriptional mechanisms [94,95]. How does nicotine up-regulate ACR-16 expression at the cellular level? Nicotine may act as a molecular chaperone, enhancing the assembly, maturation, and trafficking of ACR-16 subunits, as reported in both heterologous systems and native neurons [9698]. Another possibility is nicotine-mediated inhibition of receptor turnover, as observed for mammalian α4β2 receptors [99], suggesting that receptor accumulation may arise from reduced degradation. Determining the precise in vivo regulatory mechanism will be an important focus for future work.

Materials and methods

Maintenance and generation of C. elegans strains

All strains were maintained on standard NGM plates seeded with E. coli OP50, which were allowed to grow for at least 16 h prior to experimentation, and kept at 22 °C under standard conditions. Unless otherwise stated, “wild-type” refers to the Bristol N2 strain. For optogenetic experiments, strains were maintained in darkness on NGM plates supplemented with 0.5 mM all-trans retinal (ATR), with control groups reared on plates lacking ATR. Genetic mutants used to generate transgenic lines and compound mutants were obtained from the Caenorhabditis Genetics Center (CGC; http://www.cbs.umn.edu/CGC/) or the National Bio-Resources Project (NBRP; http://www.shigen.nig.ac.jp/c.elegans/indices.jsp).

Transgenic animals carrying non-integrated, extra-chromosomal arrays (gaaEx) were produced by co-injecting an injection marker along with one or more DNA constructs at 5–30 ng/µL. Stable integrated lines (gaaIs) were obtained from gaaEx strains via UV irradiation, followed by at least four rounds of outcrossing with N2. A complete list of strains used in this study is provided in S1 Table.

Molecular biology

All expression plasmids were generated using the Multisite Gateway system (Invitrogen, Thermo Fisher Scientific, Waltham, MA, USA). Three entry clones—containing the promoter, target gene, and fluorescent marker (designated as slot1, slot2, and slot3, respectively)—were recombined into the pDEST R4-R3 Vector II via an LR reaction [100]. Genomic DNA from wild-type C. elegans was used to amplify acr-16, twk-40s, nmr-1, mec-4, myo-3, sra-11, and unc-4. The promoter lengths of Pacr-16, Ptwk-40s, Pnmr-1, Pmec-4, Pmyo-3, Psra-11, and Punc-4 were 3.5, 1.5, 5.0, 1.5, 2.3, 2.0, and 3.0 kb, respectively.

Genes of interest—including endogenous proteins (e.g., ACR-16) and exogenous functional proteins (e.g., GCaMP6s, GtACR2, ChR2, and miniSOG)—were inserted into pDONR221 to generate slot2 constructs. Fluorescent proteins or the unc-54-3′UTR sequences were cloned into the pDONR-P2R-P3 vector to generate slot3 constructs. The Pacr-16::GFP reporter was generated using the Gateway recombination system (Invitrogen) in combination with In-Fusion cloning. To make the entry vectors, a 3.5 kb of sequence upstream of the acr-16 gene was PCR amplified from C. elegans N2 genomic DNA using gene-specific promoter primers containing attB1 and attB2 sequences. The gene-specific sequences for Pacr-16 are as follows:

  1. Forward: 5′-AGTTGCTCCCTCGAGACATTCACAACATATACGGTG-3′
  2. Reverse: 5′-CTTGCACGACTAGTGGAGTGAATGTACATCCG-3′

The amplified promoter fragment was inserted into a linearized Gateway entry vector (Entry Clone A) using In-Fusion cloning, generating an entry clone containing the acr-16 promoter flanked by attL1 and attL2 recombination sites. An LR recombination reaction was subsequently performed between the entry clone and a Gateway destination vector harboring GFP and the unc-54–3′UTR, resulting in the final expression construct, Pacr-16::GFP::unc-54-3′UTR.

To visualize endogenous ACR-16 expression, three tandem copies of the spGFP11 sequence were inserted into the acr-16::aid::Scarlet locus using CRISPR/Cas9, generating acr-16::aid::Scarlet::spGFP11. Complementary spGFP1-10 was expressed specifically in AVA neurons. Scarlet fluorescence, visible with NeuroSIL, labeled all ACR-16 clusters, while GFP fluorescence was reconstituted exclusively in AVA neurons expressing spGFP1-10. AVA-specific expression was achieved using a Cre-loxP recombination strategy with Pflp-18 and Pgpa-14 promoters.

The strain driven by the Ptwk-40s promoter expressed ACR-16 cDNA tagged with GFP, resulting in predominant expression of functional ACR-16 in AVA neurons, as confirmed by GFP fluorescence. The cDNA sequence was amplified via 2× Phanta Max Master Mix (Dye Plus) Taq DNA Polymerase (Vazyme Biotech Co.). Detailed plasmid and primers information are provided in S2 and S3 Tables.

Neuronal manipulation

To investigate the roles of TRNs and the proprioceptive neuron DVA in regulating DMP-reversal coupling, we expressed the genetically encoded photosensitizer miniSOG in these neurons driven by Pmec-4 and Pnlp-12 promoters, respectively. Neuronal ablation was achieved by exposing animals to 470 nm blue LED light (8.3 mW/cm2) for 45 min. To verify the specificity and efficiency of ablation, cytoplasmic wCherry was co-expressed with miniSOG under the same promoter, enabling visualization of target neurons. Ablations were performed at the L2–L3 larval stage, and wCherry fluorescence was examined 24 h later to confirm loss of target neurons.

For optogenetic manipulation of TRN neurons, we used the neuron-specific promoter Pmec-4 to drive the expression of the GtACR2 or ChR2 in TRN neurons. Neuronal activity was inhibited or activated using blue light during behavioral assays. In GtACR2 experiments, TRN inhibition was used to evaluate the contribution of these neurons to nicotine-induced modulation of DMP behavior. Worms were subjected to three consecutive DMP trials: in the first and third trials (no blue light), TRNs were active, and nicotine-treated worms exhibited reversals following DMP. In the second trial, blue light was applied to inhibit TRNs, which abolished reversals after the DMP, indicating that TRN activity is required for nicotine-enhanced DMP-reversal coupling.

Behavioral analysis

Behavioral assays were conducted at 22 °C. Late L4 stage or early young adults (12 h post L4 stage) hermaphrodites were used. 4.86 μL of nicotine (Shanghai PureOne Biotechnology, l-Nicotine, HPLC purity ≥ 98%) was added into 1 mL of an E. coli OP50 suspension to create a 30 mM nicotine-OP50 mixture after vortexing for 1 min. 100 μL of this mixture was then dropped and spread onto the surface of a standard 3 mL NGM agar plate, and the NGM was shaken to ensure that the mixed solution evenly covered the entire surface. These nicotine plates were left at 22 °C for 16 h to allow diffusion of nicotine into the agar and OP50 bacterial lawn formation [101]. The final nicotine concentration in these plates is approximately 1 mM. NGM plates were prepared both with and without nicotine, with DHβE added to a final concentration of 20 µM. DNQX (500 µM) and MK-801 (50 µM) were simultaneously included in both nicotine-present and nicotine-absent NGM plates to serve as the treatment conditions for C. elegans. Worms were transferred to NGM plates with OP50 containing nicotine and cultured for 3 h prior to the experiment.

Defecation behavior was measured on NGM plates under standard conditions, as previously described [31]. The nematodes’ defecation behavior was monitored for 10 consecutive cycles. The backward coupling rate was then calculated by dividing the number of backward events by the total number of defecations. The time window for assessing motor coupling was defined as the 5-s period following DMP initiation, during which reversals and calcium transients were recorded. Defecation phenotypes were scored under a modified stereomicroscope (Axio Zoom V16, Zeiss) equipped with a digital camera (acA2500-60 um, Basler).

For locomotion analysis, hermaphroditic worms, cultured under standard conditions, were transferred to imaging plates seeded with a thin layer of OP50 bacteria. One minute after the transfer, a 3-min video of the nematodes’ locomotion was recorded. The eyelash touch assay was performed on food. Post-imaging analysis was conducted using an in-house developed MATLAB script, with the centerline used for tracking. The speed analysis of each animal was performed by dividing the recorded images into 33 segments, and the midpoint of each segment was used to calculate the velocity [100]. All images were captured using a 12× objective at a frame rate of 10 Hz.

Confocal fluorescence microscopy

L4 stage transgenic animals expressing fluorescence markers were picked to imaging. Live worms were immobilized by 2.5 mM levamisole (Sigma-Aldrich) in M9 buffer. SGA741 gaaIs55 N2; [Ptwk-40s::ACR-16::GFP] fluorescence signals were captured using a Plan-Apochromatic 60× objective on a confocal microscope (FV3000, Olympus). EN9002 krSi81 [Pmyo-3::TIR1::bfp] I; bab535 [acr-16::aid::scarlet::spgfp11x3] V; krEx1404 [Pflp-18::lox::STOP::lox::spgfp1-10; Pgap-14::cre; Pmyo-2::mCherry] fluorescence signals were performed using an Andor Revolution XD laser confocal microscope system (Andor Technology plc., Springvale Business Park, Belfast, UK) based on a spinning-disk confocal scanning head CSU-X1 (Yokogawa Electric Corporation, Musashino-shi, Tokyo, Japan). The confocal system was mounted on an Olympus IX-71 inverted microscope (Olympus, Tokyo, Japan) and controlled by Andor IQ 1.91 software. The neurons and muscles labeled with GFP or RFP have been imaged with an excitation wavelength laser at 488 or 561 nm, respectively. The images were displayed using Image J 1.52o software (Wayne Rasband, National Institutes of Health, USA).

Calcium imaging

The free-tracking Ca2+ imaging setup consists of two integrated modules: a behavior tracking system and a fluorescence recording system. Worm behavior was imaged under dark-field illumination in the near-infrared range. Individual worms were picked from 6 cm NGM plates and placed on a THOPLABS SPW602 XY motorized stage for automated control. Imaging plates were prepared using the same agar composition as behavioral assay plates.

Behavioral tracking was performed using a 10× inverted objective (Olympus, Japan) and captured with a CCD camera (Point Gray Research, CM3-U3-13S2M, Canada). Custom real-time computer vision software kept the worm centered in the field of view by tracing the centerline of the worm. The fluorescence recording system records neuronal Ca2+ activity using an sCMOS digital camera (Hamamatsu, Japan) equipped with a 10× objective lens (Olympus, Japan). To simultaneously capture images of wCherry and GCaMP6 side by side, we utilized a two-channel imager (W-VIEW GEMINI, Japan). Red and green channel images were simultaneously recorded at 10 frames per second (fps) with exposure times of 10 and 60 ms, respectively, using HCImage software (Hamamatsu).

Neural activity was quantified as the normalized change in the fluorescence ratio between GCaMP6s and wCherry, calculated as: ΔR/R0 = (RR0)/R0, where R=IGCaMP6s/IwCherry, and R0 was defined as the minimum R value within the recording period. Intensities IGCaMP6s and IwCherry were extracted as the pixel intensity in the green and red channels, respectively, using custom MATLAB scripts. Analysis focused on AVA neuronal activity during freely moving worms performing the DMP.

A calcium response was defined as an increase in fluorescence intensity of ≥10% that occurred with a reversal behavior immediately following the pBoc phase of the DMP. Each such event was counted as one calcium response. We then calculated the calcium response percentage by dividing the number of these response events by five—the total number of DMP cycles analyzed per animal. This quantification yields discrete values (e.g., 0, 20%, 40%, …, 100%) corresponding to the number of DMP cycles that showed a coupled calcium response. To minimize signal drift and motion artifacts, we limited our analysis to five consecutive DMP cycles per worm.

Electrophysiology

Whole-cell patch-clamp recordings were performed following established protocols with minor modifications [102,103]. Briefly, 1–2-day-old hermaphrodite adults were immobilized on a Sylgard-coated (Sylgard 184, Dow Corning) coverslip using Histoacryl Blue adhesive (Braun) under a stereomicroscope (M50, Leica) and submerged in bath solution. Viscera were removed by gentle suction through a glass pipette, and the resulting cuticle flap was folded back and secured with WORMGLU adhesive (GluStitch) to expose the neuromuscular system.

Body wall muscle cells and neurons were patched using borosilicate glass pipettes (1B100F-4, World Precision Instruments) with tip resistances of 4–6 MΩ and 15–20 MΩ, respectively. Pipettes were pulled by micropipette puller P-1000 (Sutter) and fire-polished using an MF-830 microforge (Narishige). Recordings were obtained in the whole-cell configuration using an EPC9 amplifier (HEKA) controlled by Pulse software, with data processed using Igor Pro (WaveMetrics) and Clampfit 10 (Molecular Devices). Membrane currents were recorded at a holding potential of −60 mV, while membrane potentials were recorded at 0 pA. Signals were digitized at 10 kHz and low-pass filtered at 2.6 kHz.

The pipette solution contains (in mM): K-gluconate 115; KCl 25; CaCl2 0.1; MgCl2 5; BAPTA 1; HEPES 10; Na2ATP 5; Na2GTP 0.5; cAMP 0.5; cGMP 0.5, pH 7.2 with KOH, ∼320 mOsm. cAMP and cGMP were included to maintain the activity and longevity of the preparation. The bath solution consists of (in mM): NaCl 150; KCl 5; CaCl2 5; MgCl2 1; glucose 10; sucrose 5; HEPES 15, pH 7.3 with NaOH, ∼330 mOsm. All chemicals were purchased from Sigma unless otherwise specified. Experiments were conducted at 20–22 °C. To test AVA receptor function, ACh (1 mM in bath solution) was directly perfused onto patched neurons. For pharmacological blockade, recordings were performed in bath solution containing 20 μM DHβE. Optogenetic activation of TRNs during recordings was achieved by blue light illumination (470 nm, 13.75 mW/cm2).

Statistical analysis and display

Statistical analyses were conducted via Student t test, One-way ANOVA, or Two-way ANOVA analysis, two-sample Z test, as appropriate. Data are presented as mean ± SEM, and significance thresholds were set at p < 0.05 (* p < 0.05, ** p < 0.01, *** p < 0.001). Calcium imaging data were visualized as heat maps (MATLAB, MathWorks), time-course curves (Igor Pro, WaveMetrics), and scatter plots (GraphPad Prism 9, GraphPad Software), with each point representing a single worm. Additional graphing software included Clampfit (Molecular Devices), ImageJ (NIH), and MEGA 6.60 (for phylogenetic analyses). For behavioral and fluorescence imaging experiments, unless otherwise noted, each recording trace represented an independent animal.

Supporting information

S1 Fig. Nicotine increases reversal frequency and speed.

(A) Raster plots displaying individual locomotion speed for animals without (− nic) and with (+ nic) nicotine exposure (n = 22 animals per group). (B) Quantification of the proportion change of forward and reversal movements with nicotine. Nicotine significantly increased the proportion of reversal. *** p < 0.001 by Two-sample Z test. (C, D) Quantification of the reversal frequency and duration changes after nicotine exposure. (E, F) Distribution of instantaneous speed of reversal (E) and forward (F) locomotion. Nicotine leads to a drastic increase of speed in both reversal and forward locomotion. *** p < 0.001 by Student t test. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s001

(TIF)

S2 Fig. Nicotine does not alter Caenorhabditis elegans DMP cycle.

(A) A schematic diagram of C. elegans defecation motor program (DMP). DMP is initiated by posterior body contraction (pBoc), and followed by anterior body contraction (aBoc) after ~2 s relaxation phase and then enteric muscle contraction, leading to expulsion of the gut contents (Exp), and an intercycle of approximately 45 s. (B) Representative ethograms of consecutive 10 defecation cycles in wild-type worms before (− nic) and after (+ nic) the exposure of nicotine (1 mM). Each dot represents 1 s. ‘‘p’’ stands for pBoc and ‘‘x’’ indicates Exp. aBoc is omitted due to difficulties in observation. (C) Quantification of the DMP cycle with or without nicotine exposure. (D) The distribution of the delay time between reversal initiation and different DMP phases. The aBoc is based on estimation. (E) Quantification of the average coupling events in wild-type and respective mutants in 10 min. Two-way ANOVA was performed (interaction: F(4, 81) = 15.82, P < 0.0001). (F) Quantification of the average delay time, which was not affected by nicotine exposure. n ≥ 8 animals (Each dot represents a single DMP cycle). ns, no significance, * p < 0.05, ** p < 0.01, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s002

(TIF)

S3 Fig. Nicotine enhanced DMP-reversal coupling is independent on glutamate.

(A) Suppression of nicotine increased reversal frequency by DHβE (20 µM). Two-way ANOVA was performed (interaction: F(1, 36) = 34.69, P < 0.0001). (B) Suppression of nicotine increased reversal duration by DHβE (20 µM). Two-way ANOVA was performed (interaction: F(1, 154) = 5.442, P = 0.0210). (C, D) Distribution of instantaneous speed of reversal and forward locomotion. Nicotine leads to a drastic increase of speed in both reversal and forward locomotion (C). Velocity changes due to nicotine are all inhibited by DHβE (D). ns, no significance, ** p < 0.01, *** p < 0.001 by Student t test. Error bars, SEM. (E) Mutations causing uncoordinated locomotion in nAChR mutants, including unc-29, unc-38, unc-63, suppressed the nicotine exposure induced DMP-reversal coupling. Two-way ANOVA was performed (interaction: F(2, 29) = 1.219, P = 0.3101). (F) No significant suppression of nicotine enhanced DMP-reversal coupling was observed in glutamate receptors mutant worms. Two-way ANOVA was performed (interaction: F(5, 79) = 2.669, P = 0.0279). n ≥ 5 animals. ns, no significance, *** p < 0.001 by Two-way ANOVA. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s003

(TIF)

S4 Fig. Nicotine-enhanced reversal in hpIs580 was abolished in acr-16 mutants.

(A) Quantification shows the nicotine enhanced reversal frequency observed in hpIs580 (AVA::GCaMP) worms is abolished in acr-16 mutants. Two-way ANOVA was performed (interaction: F(2, 54) = 17.92, P < 0.0001). (B) Quantification shows that the nicotine enhanced reversal duration observed in hpIs580 worms is abolished in acr-16 mutants. Reversal frequency and duration were restored by expression ACR-16 back. Two-way ANOVA was performed (interaction: F(2, 274) = 1.870, P = 0.1561). ns, no significance, ** p < 0.01, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. (C, D) Distribution of instantaneous speed of reversal (C) and forward (D) locomotion. Nicotine causes a sharp increase in the rate of reverse and forward motion in hpIs580 worms, eliminated by acr-16 mutants. n = 10 animals. ns, no significance, ** p < 0.01, *** p < 0.001 by Student t test. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s004

(TIF)

S5 Fig. Nicotine increases the expression of ACR-16.

(A) Nicotine up-regulates ACR-16 expression in AVA. Using ACR-16 cDNA with a fluorescent marker (sl2d::GFP) specifically in AVA neurons. (B) After nicotine exposure, the fluorescence intensity in AVA soma is significantly increased. (C) Nicotine-induced increase in I ACh could be suppressed by DHβE (20 µM). (D) Quantification of the peak current with DHβE. n ≥ 6 animals. *** p < 0.001 by Student t test. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s005

(TIF)

S6 Fig. The membrane potential characteristics of AVA neurons are unaffected by nicotine and ACR-16.

(A) Resting membrane potentials (RMP) of all individual AVA neurons (gray lines) and the average RMP (black lines). (B) Quantification shows that the RMP of AVA exhibit no significantly change in different genotypes before and after nicotine exposure (1 mM, 3 h). Two-way ANOVA was performed (interaction: F(2, 42) = 0.3219, P = 0.7265). ns, no significance, by Two-way ANOVA analysis. Error bars, SEM. (C) Representative step currents in AVA neurons from different genotypes. The currents were evoked by voltage clamp from −80 mV to + 80 mV at 20 mV increment. (D) Quantification of the I-V curve of the voltage-dependent currents. n ≥ 6 animals. (E–H) DHβE (20 µM) had no effect on either the RMP (E, F) or the step currents (G, H) in AVA neurons. ns, no significance, by Student t test. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s006

(TIF)

S7 Fig. Ablation of DVA has no effect of nicotine’s action.

(A) DMP cycle in wild-type N2 and mec-2(e75), mec-4(e1611), and mec-6(e1342) mutants. Two-way ANOVA was performed (interaction: F(3, 712) = 1.238, P = 0.2948). (B) Quantification of the spontaneous reversal rate (# number of reversals per 3 min) in the presence (On food) and absence (Off food) of food, for wild-type (N2) and mec-4(e1611) mutants. Two-way ANOVA was performed (interaction: F(1, 36) = 0.08663, P = 0.7702). (C) Representative TRNs (PLM) before and after the ablation. (D) Ablation of DVA did not change the nicotine enhanced DMP-reversal coupling proportion. Two-way ANOVA was performed (interaction: F(1, 12) = 0.1765, P = 0.6818). ns, no significance, *** p < 0.001 by Two-way ANOVA analysis. Error bars, SEM. The data underlying this figure can be found in S1 Data.

https://doi.org/10.1371/journal.pbio.3003423.s007

(TIF)

S1 Table. All Caenorhabditis elegans strains used in this study.

Genotypes used in each figure.

https://doi.org/10.1371/journal.pbio.3003423.s008

(PDF)

S2 Table. The plasmids used in this study.

List of all plasmids used to generate transgenic lines.

https://doi.org/10.1371/journal.pbio.3003423.s009

(PDF)

S3 Table. The primers used in this study.

Primers utilized for PCR amplification to identify mutants and construct transgenic strains.

https://doi.org/10.1371/journal.pbio.3003423.s010

(PDF)

S1 Movie. Wild-type N2 movement without nicotine.

Wild-type N2 nematode movements during three consecutive DMP cycles without nicotine-exposure (N2, − nic).

https://doi.org/10.1371/journal.pbio.3003423.s011

(MP4)

S2 Movie. Wild-type N2 movement with nicotine-exposure.

Wild-type N2 nematode movements during three consecutive DMP cycles with nicotine-exposure (N2, + nic).

https://doi.org/10.1371/journal.pbio.3003423.s012

(MP4)

S3 Movie. acr-16 movement without nicotine.

acr-16 mutant movements during three consecutive DMP cycles without nicotine-exposure (acr-16, − nic).

https://doi.org/10.1371/journal.pbio.3003423.s013

(MP4)

S4 Movie. acr-16 movement with nicotine-exposure.

acr-16 mutant movements during three consecutive DMP cycles with nicotine-exposure (acr-16, + nic).

https://doi.org/10.1371/journal.pbio.3003423.s014

(MP4)

S1 Data. The data underlying the graphs shown in Figs 16 and S1S7.

https://doi.org/10.1371/journal.pbio.3003423.s015

(XLSX)

Acknowledgments

We thank Tianqi Xu and Quan Wen for their assistance with data analysis. We thank Mei Zhen, Shawn Xu, Quan Wen, Lijun Kang, Zhuo Du, Xiao Liu, Haijun Tu for reagents and strains. We thank Mei Zhen, Quan Wen, Shawn Xu, and Lijun Kang for their valuable comments and discussion. We thank the Caenorhabditis Genetics Center, which is funded by the NIH Office of Research Infrastructure Programs (P40 OD010440, https://orip.nih.gov/), for strains.

References

  1. 1. Mansvelder HD, Keath JR, McGehee DS. Synaptic mechanisms underlie nicotine-induced excitability of brain reward areas. Neuron. 2002;33(6):905–19. pmid:11906697
  2. 2. Volkow ND, Michaelides M, Baler R. The neuroscience of drug reward and addiction. Physiol Rev. 2019;99(4):2115–40. pmid:31507244
  3. 3. Henderson BJ, Wall TR, Henley BM, Kim CH, McKinney S, Lester HA. Menthol enhances nicotine reward-related behavior by potentiating nicotine-induced changes in nAChR Function, nAChR upregulation, and DA neuron excitability. Neuropsychopharmacology. 2017;42(12):2285–91. pmid:28401925
  4. 4. Baker TB, Piper ME, McCarthy DE, Majeskie MR, Fiore MC. Addiction motivation reformulated: an affective processing model of negative reinforcement. Psychol Rev. 2004;111(1):33–51. pmid:14756584
  5. 5. Wills L, Ables JL, Braunscheidel KM, Caligiuri SPB, Elayouby KS, Fillinger C, et al. Neurobiological mechanisms of nicotine reward and aversion. Pharmacol Rev. 2022;74(1):271–310. pmid:35017179
  6. 6. Laviolette SR, van der Kooy D. The neurobiology of nicotine addiction: bridging the gap from molecules to behaviour. Nat Rev Neurosci. 2004;5(1):55–65. pmid:14708004
  7. 7. Wise RA, Robble MA. Dopamine and addiction. Annu Rev Psychol. 2020;71:79–106. pmid:31905114
  8. 8. Zhao-Shea R, DeGroot SR, Liu L, Vallaster M, Pang X, Su Q, et al. Increased CRF signalling in a ventral tegmental area-interpeduncular nucleus-medial habenula circuit induces anxiety during nicotine withdrawal. Nat Commun. 2015;6:6770. pmid:25898242
  9. 9. Markou A. Review. Neurobiology of nicotine dependence. Philos Trans R Soc Lond B Biol Sci. 2008;363(1507):3159–68. pmid:18640919
  10. 10. Li X, Semenova S, D’Souza MS, Stoker AK, Markou A. Involvement of glutamatergic and GABAergic systems in nicotine dependence: implications for novel pharmacotherapies for smoking cessation. Neuropharmacology. 2014;76 Pt B(0 0):554–65. pmid:23752091
  11. 11. McNeill AD, West RJ, Jarvis M, Jackson P, Bryant A. Cigarette withdrawal symptoms in adolescent smokers. Psychopharmacology (Berl). 1986;90(4):533–6. pmid:3101108
  12. 12. Killen JD, Fortmann SP, Telch MJ, Newman B. Are heavy smokers different from light smokers? A comparison after 48 hours without cigarettes. JAMA. 1988;260(11):1581–5. pmid:3411738
  13. 13. Liu C, Tose AJ, Verharen JPH, Zhu Y, Tang LW, de Jong JW, et al. An inhibitory brainstem input to dopamine neurons encodes nicotine aversion. Neuron. 2022;110(18):3018–3035.e7. pmid:35921846
  14. 14. Wolfman SL, Gill DF, Bogdanic F, Long K, Al-Hasani R, McCall JG, et al. Nicotine aversion is mediated by GABAergic interpeduncular nucleus inputs to laterodorsal tegmentum. Nat Commun. 2018;9(1):2710. pmid:30006624
  15. 15. Grieder TE, Besson M, Maal-Bared G, Pons S, Maskos U, van der Kooy D. β2* nAChRs on VTA dopamine and GABA neurons separately mediate nicotine aversion and reward. Proc Natl Acad Sci U S A. 2019;116(51):25968–73. pmid:31776253
  16. 16. Feng Z, Li W, Ward A, Piggott BJ, Larkspur ER, Sternberg PW, et al. A C. elegans model of nicotine-dependent behavior: regulation by TRP-family channels. Cell. 2006;127(3):621–33. pmid:17081982
  17. 17. Sellings L, Pereira S, Qian C, Dixon-McDougall T, Nowak C, Zhao B, et al. Nicotine-motivated behavior in Caenorhabditis elegans requires the nicotinic acetylcholine receptor subunits acr-5 and acr-15. Eur J Neurosci. 2013;37(5):743–56. pmid:23351035
  18. 18. Sobkowiak R, Kowalski M, Lesicki A. Concentration- and time-dependent behavioral changes in Caenorhabditis elegans after exposure to nicotine. Pharmacol Biochem Behav. 2011;99(3):365–70. pmid:21624385
  19. 19. Wescott SA, Ronan EA, Xu XZS. Insulin signaling genes modulate nicotine-induced behavioral responses in Caenorhabditis elegans. Behav Pharmacol. 2016;27(1):44–9. pmid:26317299
  20. 20. Gottschalk A, Almedom RB, Schedletzky T, Anderson SD, Yates JR 3rd, Schafer WR. Identification and characterization of novel nicotinic receptor-associated proteins in Caenorhabditis elegans. EMBO J. 2005;24(14):2566–78. pmid:15990870
  21. 21. Polli JR, Dobbins DL, Kobet RA, Farwell MA, Zhang B, Lee M-H, et al. Drug-dependent behaviors and nicotinic acetylcholine receptor expressions in Caenorhabditis elegans following chronic nicotine exposure. Neurotoxicology. 2015;47:27–36. pmid:25530353
  22. 22. Waggoner LE, Dickinson KA, Poole DS, Tabuse Y, Miwa J, Schafer WR. Long-term nicotine adaptation in Caenorhabditis elegans involves PKC-dependent changes in nicotinic receptor abundance. J Neurosci. 2000;20(23):8802–11. pmid:11102488
  23. 23. Sattelle DB, Buckingham SD, Akamatsu M, Matsuda K, Pienaar IS, Jones AK, et al. Comparative pharmacology and computational modelling yield insights into allosteric modulation of human alpha7 nicotinic acetylcholine receptors. Biochem Pharmacol. 2009;78(7):836–43. pmid:19549506
  24. 24. McLaughlin I, Dani JA, De Biasi M. Nicotine withdrawal. Curr Top Behav Neurosci. 2015;24:99–123. pmid:25638335
  25. 25. Kenny PJ, Markou A. Neurobiology of the nicotine withdrawal syndrome. Pharmacol Biochem Behav. 2001;70(4):531–49. pmid:11796152
  26. 26. Jackson KJ, Muldoon PP, De Biasi M, Damaj MI. New mechanisms and perspectives in nicotine withdrawal. Neuropharmacology. 2015;96(Pt B):223–34. pmid:25433149
  27. 27. Chellian R, Behnood-Rod A, Bruijnzeel DM, Wilson R, Pandy V, Bruijnzeel AW. Rodent models for nicotine withdrawal. J Psychopharmacol. 2021;35(10):1169–87. pmid:33888006
  28. 28. Nagy S, Huang Y-C, Alkema MJ, Biron D. Caenorhabditis elegans exhibit a coupling between the defecation motor program and directed locomotion. Sci Rep. 2015;5:17174. pmid:26597056
  29. 29. McGovern DJ, Polter AM, Root DH. Neurochemical signaling of reward and aversion to ventral tegmental area glutamate neurons. J Neurosci. 2021;41(25):5471–86. pmid:34001626
  30. 30. Hajek P, Gillison F, McRobbie H. Stopping smoking can cause constipation. Addiction. 2003;98(11):1563–7. pmid:14616182
  31. 31. Thomas JH. Genetic analysis of defecation in Caenorhabditis elegans. Genetics. 1990;124(4):855–72. pmid:2323555
  32. 32. Wang H, Girskis K, Janssen T, Chan JP, Dasgupta K, Knowles JA, et al. Neuropeptide secreted from a pacemaker activates neurons to control a rhythmic behavior. Curr Biol. 2013;23(9):746–54. pmid:23583549
  33. 33. Iwasaki K, Toyonaga R. The Rab3 GDP/GTP exchange factor homolog AEX-3 has a dual function in synaptic transmission. EMBO J. 2000;19(17):4806–16. pmid:10970871
  34. 34. De Biasi M, Dani JA. Reward, addiction, withdrawal to nicotine. Annu Rev Neurosci. 2011;34:105–30. pmid:21438686
  35. 35. Sherrington CS. The integrative action of the nervous system. Yale University Press; 1906.
  36. 36. Craig ADB. How do you feel--now? The anterior insula and human awareness. Nat Rev Neurosci. 2009;10(1):59–70. pmid:19096369
  37. 37. Critchley HD, Harrison NA. Visceral influences on brain and behavior. Neuron. 2013;77(4):624–38. pmid:23439117
  38. 38. Picciotto MR, Kenny PJ. Mechanisms of nicotine addiction. Cold Spring Harb Perspect Med. 2021;11(5):a039610. pmid:32341069
  39. 39. Picciotto MR, Kenny PJ. Molecular mechanisms underlying behaviors related to nicotine addiction. Cold Spring Harb Perspect Med. 2013;3(1):a012112. pmid:23143843
  40. 40. Champtiaux N, Changeux J-P. Knockout and knockin mice to investigate the role of nicotinic receptors in the central nervous system. Prog Brain Res. 2004;145:235–51. pmid:14650919
  41. 41. Rose JK, Kaun KR, Chen SH, Rankin CH. GLR-1, a non-NMDA glutamate receptor homolog, is critical for long-term memory in Caenorhabditis elegans. J Neurosci. 2003;23(29):9595–9. pmid:14573539
  42. 42. Brockie PJ, Mellem JE, Hills T, Madsen DM, Maricq AV. The C. elegans glutamate receptor subunit NMR-1 is required for slow NMDA-activated currents that regulate reversal frequency during locomotion. Neuron. 2001;31(4):617–30. pmid:11545720
  43. 43. Jones AK, Sattelle DB. Diversity of insect nicotinic acetylcholine receptor subunits. Adv Exp Med Biol. 2010;683:25–43. pmid:20737786
  44. 44. Francis MM, Evans SP, Jensen M, Madsen DM, Mancuso J, Norman KR, et al. The Ror receptor tyrosine kinase CAM-1 is required for ACR-16-mediated synaptic transmission at the C. elegans neuromuscular junction. Neuron. 2005;46(4):581–94. pmid:15944127
  45. 45. Touroutine D, Fox RM, Von Stetina SE, Burdina A, Miller DM 3rd, Richmond JE. acr-16 encodes an essential subunit of the levamisole-resistant nicotinic receptor at the Caenorhabditis elegans neuromuscular junction. J Biol Chem. 2005;280(29):27013–21. pmid:15917232
  46. 46. Fleming JT, Squire MD, Barnes TM, Tornoe C, Matsuda K, Ahnn J, et al. Caenorhabditis elegans levamisole resistance genes lev-1, unc-29, and unc-38 encode functional nicotinic acetylcholine receptor subunits. J Neurosci. 1997;17(15):5843–57. pmid:9221782
  47. 47. Culetto E, Baylis HA, Richmond JE, Jones AK, Fleming JT, Squire MD, et al. The Caenorhabditis elegans unc-63 gene encodes a levamisole-sensitive nicotinic acetylcholine receptor alpha subunit. J Biol Chem. 2004;279(41):42476–83. pmid:15280391
  48. 48. Ballivet M, Alliod C, Bertrand S, Bertrand D. Nicotinic acetylcholine receptors in the nematode Caenorhabditis elegans. J Mol Biol. 1996;258(2):261–9. pmid:8627624
  49. 49. Taylor SR, Santpere G, Weinreb A, Barrett A, Reilly MB, Xu C, et al. Molecular topography of an entire nervous system. Cell. 2021;184(16):4329–4347.e23. pmid:34237253
  50. 50. Mialon M, Patrash L, Weinreb A, Özkan E, Bessereau J-L, Pinan-Lucarre B. A trans-synaptic IgLON adhesion molecular complex directly contacts and clusters a nicotinic receptor. bioRxiv. 2024. pmid:39314492
  51. 51. Meng J, Ahamed T, Yu B, Hung W, Ei Mouridi S, Wang Z, et al. A tonically active master neuron modulates mutually exclusive motor states at two timescales. Sci Adv. 2024;10(15):eadk0002. pmid:38598630
  52. 52. Nguyen JP, Shipley FB, Linder AN, Plummer GS, Liu M, Setru SU, et al. Whole-brain calcium imaging with cellular resolution in freely behaving Caenorhabditis elegans. Proc Natl Acad Sci U S A. 2016;113(8):E1074–81. pmid:26712014
  53. 53. Li Y, Chitturi J, Yu B, Zhang Y, Wu J, Ti P, et al. UBR-1 ubiquitin ligase regulates the balance between GABAergic and glutamatergic signaling. EMBO Rep. 2023;24(11):e57014. pmid:37811674
  54. 54. Buisson B, Bertrand D. Chronic exposure to nicotine upregulates the human (alpha)4((beta)2 nicotinic acetylcholine receptor function. J Neurosci. 2001;21(6):1819–29. pmid:11245666
  55. 55. Vallejo YF, Buisson B, Bertrand D, Green WN. Chronic nicotine exposure upregulates nicotinic receptors by a novel mechanism. J Neurosci. 2005;25(23):5563–72. pmid:15944384
  56. 56. He S, Cuentas-Condori A, Miller DM 3rd. NATF (Native and Tissue-Specific Fluorescence): a strategy for bright, tissue-specific GFP labeling of native proteins in Caenorhabditis elegans. Genetics. 2019;212(2):387–95. pmid:30952669
  57. 57. Schmitt C, Schultheis C, Pokala N, Husson SJ, Liewald JF, Bargmann CI, et al. Specific expression of channelrhodopsin-2 in single neurons of Caenorhabditis elegans. PLoS One. 2012;7(8):e43164. pmid:22952643
  58. 58. Beg AA, Ernstrom GG, Nix P, Davis MW, Jorgensen EM. Protons act as a transmitter for muscle contraction in C. elegans. Cell. 2008;132(1):149–60. pmid:18191228
  59. 59. Bender A, Woydziak ZR, Fu L, Branden M, Zhou Z, Ackley BD, et al. Novel acid-activated fluorophores reveal a dynamic wave of protons in the intestine of Caenorhabditis elegans. ACS Chem Biol. 2013;8(3):636–42. pmid:23256594
  60. 60. Schafer WR. Mechanosensory molecules and circuits in C. elegans. Pflugers Arch. 2015;467(1):39–48. pmid:25053538
  61. 61. Geffeney SL, Cueva JG, Glauser DA, Doll JC, Lee TH-C, Montoya M, et al. DEG/ENaC but not TRP channels are the major mechanoelectrical transduction channels in a C. elegans nociceptor. Neuron. 2011;71(5):845–57. pmid:21903078
  62. 62. Kindt KS, Viswanath V, Macpherson L, Quast K, Hu H, Patapoutian A, et al. Caenorhabditis elegans TRPA-1 functions in mechanosensation. Nat Neurosci. 2007;10(5):568–77. pmid:17450139
  63. 63. Yeon J, Kim J, Kim D-Y, Kim H, Kim J, Du EJ, et al. A sensory-motor neuron type mediates proprioceptive coordination of steering in C. elegans via two TRPC channels. PLoS Biol. 2018;16(6):e2004929. pmid:29883446
  64. 64. Chelur DS, Ernstrom GG, Goodman MB, Yao CA, Chen L, O’ Hagan R, et al. The mechanosensory protein MEC-6 is a subunit of the C. elegans touch-cell degenerin channel. Nature. 2002;420(6916):669–73. pmid:12478294
  65. 65. Staum M, Abraham A-C, Arbid R, Birari VS, Dominitz M, Rabinowitch I. Behavioral adjustment of C. elegans to mechanosensory loss requires intact mechanosensory neurons. PLoS Biol. 2024;22(7):e3002729. pmid:39024405
  66. 66. Mohammad F, Stewart JC, Ott S, Chlebikova K, Chua JY, Koh T-W, et al. Optogenetic inhibition of behavior with anion channelrhodopsins. Nat Methods. 2017;14(3):271–4. pmid:28114289
  67. 67. Mahn M, Gibor L, Patil P, Cohen-Kashi Malina K, Oring S, Printz Y, et al. High-efficiency optogenetic silencing with soma-targeted anion-conducting channelrhodopsins. Nat Commun. 2018;9(1):4125. pmid:30297821
  68. 68. Qi YB, Garren EJ, Shu X, Tsien RY, Jin Y. Photo-inducible cell ablation in Caenorhabditis elegans using the genetically encoded singlet oxygen generating protein miniSOG. Proc Natl Acad Sci U S A. 2012;109(19):7499–504. pmid:22532663
  69. 69. Li W, Feng Z, Sternberg PW, Xu XZS. A C. elegans stretch receptor neuron revealed by a mechanosensitive TRP channel homologue. Nature. 2006;440(7084):684–7. pmid:16572173
  70. 70. Bounoutas A, Chalfie M. Touch sensitivity in Caenorhabditis elegans. Pflugers Arch. 2007;454(5):691–702. pmid:17285303
  71. 71. Li W, Kang L, Piggott BJ, Feng Z, Xu XZS. The neural circuits and sensory channels mediating harsh touch sensation in Caenorhabditis elegans. Nat Commun. 2011;2:315. pmid:21587232
  72. 72. Pirri JK, Alkema MJ. The neuroethology of C. elegans escape. Curr Opin Neurobiol. 2012;22(2):187–93. pmid:22226513
  73. 73. Liewald JF, Brauner M, Stephens GJ, Bouhours M, Schultheis C, Zhen M, et al. Optogenetic analysis of synaptic function. Nat Methods. 2008;5(10):895–902. pmid:18794862
  74. 74. Lindsay TH, Thiele TR, Lockery SR. Optogenetic analysis of synaptic transmission in the central nervous system of the nematode Caenorhabditis elegans. Nat Commun. 2011;2:306. pmid:21556060
  75. 75. Zhang Y, Shi Y, Zeng K, Chen L, Gao S. Hierarchical competing inhibition circuits govern motor stability in C. elegans. Nat Commun. 2025;16(1):4405. pmid:40355468
  76. 76. Shipley FB, Clark CM, Alkema MJ, Leifer AM. Simultaneous optogenetic manipulation and calcium imaging in freely moving C. elegans. Front Neural Circuits. 2014;8:28. pmid:24715856
  77. 77. Tiwari RK, Sharma V, Pandey RK, Shukla SS. Nicotine addiction: neurobiology and mechanism. J Pharmacopuncture. 2020;23(1):1–7. pmid:32322429
  78. 78. Morris PB, Ference BA, Jahangir E, Feldman DN, Ryan JJ, Bahrami H, et al. Cardiovascular effects of exposure to cigarette smoke and electronic cigarettes: clinical perspectives from the Prevention of Cardiovascular Disease Section Leadership Council and Early Career Councils of the American College of Cardiology. J Am Coll Cardiol. 2015;66(12):1378–91. pmid:26383726
  79. 79. Haass M, Kübler W. Nicotine and sympathetic neurotransmission. Cardiovasc Drugs Ther. 1997;10(6):657–65. pmid:9110108
  80. 80. Bernardini N, Sauer SK, Haberberger R, Fischer MJ, Reeh PW. Excitatory nicotinic and desensitizing muscarinic (M2) effects on C-nociceptors in isolated rat skin. J Neurosci. 2001;21(9):3295–302. pmid:11312314
  81. 81. Hawkins JL, Denson JE, Miley DR, Durham PL. Nicotine stimulates expression of proteins implicated in peripheral and central sensitization. Neuroscience. 2015;290:115–25. pmid:25637801
  82. 82. Le Foll B, Goldberg SR. Effects of nicotine in experimental animals and humans: an update on addictive properties. Handb Exp Pharmacol. 2009;(192):335–67. pmid:19184655
  83. 83. Krieg M, Pidde A, Das R. Mechanosensitive body-brain interactions in Caenorhabditis elegans. Curr Opin Neurobiol. 2022;75:102574. pmid:35687973
  84. 84. Scheer E, Bargmann CI. Sensory neurons couple arousal and foraging decisions in Caenorhabditis elegans. Elife. 2023;12:RP88657. pmid:38149996
  85. 85. Boor SA, Meisel JD, Kim DH. Neuroendocrine gene expression coupling of interoceptive bacterial food cues to foraging behavior of C. elegans. Elife. 2024;12:RP91120. pmid:38231572
  86. 86. Park Y-J, Yeon J, Cho J, Kim D-Y, Bai X, Oh Y, et al. PIEZO acts in an intestinal valve to regulate swallowing in C. elegans. Nat Commun. 2024;15(1):10072. pmid:39567502
  87. 87. Rauthan M, Gong J, Liu J, Li Z, Wescott SA, Liu J, et al. MicroRNA regulation of nAChR expression and nicotine-dependent behavior in C. elegans. Cell Rep. 2017;21(6):1434–41. pmid:29117550
  88. 88. Smith MA Jr, Zhang Y, Polli JR, Wu H, Zhang B, Xiao P, et al. Impacts of chronic low-level nicotine exposure on Caenorhabditis elegans reproduction: identification of novel gene targets. Reprod Toxicol. 2013;40:69–75. pmid:23735997
  89. 89. Laviolette SR, van der Kooy D. The motivational valence of nicotine in the rat ventral tegmental area is switched from rewarding to aversive following blockade of the alpha7-subunit-containing nicotinic acetylcholine receptor. Psychopharmacology (Berl). 2003;166(3):306–13. pmid:12569428
  90. 90. Hobert O. The neuronal genome of Caenorhabditis elegans. WormBook. 2013:1–106. pmid:24081909
  91. 91. Chalfie M, Sulston JE, White JG, Southgate E, Thomson JN, Brenner S. The neural circuit for touch sensitivity in Caenorhabditis elegans. J Neurosci. 1985;5(4):956–64. pmid:3981252
  92. 92. Cook SJ, Jarrell TA, Brittin CA, Wang Y, Bloniarz AE, Yakovlev MA, et al. Whole-animal connectomes of both Caenorhabditis elegans sexes. Nature. 2019;571(7763):63–71. pmid:31270481
  93. 93. Sieburth D, Ch’ng Q, Dybbs M, Tavazoie M, Kennedy S, Wang D, et al. Systematic analysis of genes required for synapse structure and function. Nature. 2005;436(7050):510–7. pmid:16049479
  94. 94. Marks MJ, Pauly JR, Gross SD, Deneris ES, Hermans-Borgmeyer I, Heinemann SF, et al. Nicotine binding and nicotinic receptor subunit RNA after chronic nicotine treatment. J Neurosci. 1992;12(7):2765–84. pmid:1613557
  95. 95. Bencherif M, Fowler K, Lukas RJ, Lippiello PM. Mechanisms of up-regulation of neuronal nicotinic acetylcholine receptors in clonal cell lines and primary cultures of fetal rat brain. J Pharmacol Exp Ther. 1995;275(2):987–94. pmid:7473192
  96. 96. Nashmi R, Dickinson ME, McKinney S, Jareb M, Labarca C, Fraser SE, et al. Assembly of alpha4beta2 nicotinic acetylcholine receptors assessed with functional fluorescently labeled subunits: effects of localization, trafficking, and nicotine-induced upregulation in clonal mammalian cells and in cultured midbrain neurons. J Neurosci. 2003;23(37):11554–67. pmid:14684858
  97. 97. Sallette J, Pons S, Devillers-Thiery A, Soudant M, Prado de Carvalho L, Changeux J-P, et al. Nicotine upregulates its own receptors through enhanced intracellular maturation. Neuron. 2005;46(4):595–607. pmid:15944128
  98. 98. Kuryatov A, Luo J, Cooper J, Lindstrom J. Nicotine acts as a pharmacological chaperone to up-regulate human alpha4beta2 acetylcholine receptors. Mol Pharmacol. 2005;68(6):1839–51. pmid:16183856
  99. 99. Peng X, Gerzanich V, Anand R, Whiting PJ, Lindstrom J. Nicotine-induced increase in neuronal nicotinic receptors results from a decrease in the rate of receptor turnover. Mol Pharmacol. 1994;46(3):523–30. pmid:7935334
  100. 100. Chen L, Liu Y, Su P, Hung W, Li H, Wang Y, et al. Escape steering by cholecystokinin peptidergic signaling. Cell Rep. 2022;38(6):110330. pmid:35139370
  101. 101. Kudelska MM, Lewis A, Ng CT, Doyle DA, Holden-Dye L, O’Connor VM, et al. Investigation of feeding behaviour in C. elegans reveals distinct pharmacological and antibacterial effects of nicotine. Invert Neurosci. 2018;18(4):14. pmid:30406344
  102. 102. Gao S, Zhen M. Action potentials drive body wall muscle contractions in Caenorhabditis elegans. Proc Natl Acad Sci U S A. 2011;108(6):2557–62. pmid:21248227
  103. 103. Wang Y, Chow CH, Zhang Y, Huang M, Higazy R, Ramakrishnan N, et al. The exocytosis regulator complexin controls spontaneous synaptic vesicle release in a CAPS-dependent manner at C. elegans excitatory synapses. PLoS Biol. 2025;23(2):e3003023. pmid:39913617