Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Brief Communication
  • Published:

Hunger or thirst state uncertainty is resolved by outcome evaluation in medial prefrontal cortex to guide decision-making

Abstract

Physiological need states direct decision-making toward re-establishing homeostasis. Using a two-alternative forced choice task for mice that models elements of human decisions, we found that varying hunger and thirst states caused need-inappropriate choices, such as food seeking when thirsty. These results show limits on interoceptive knowledge of hunger and thirst states to guide decision-making. Instead, need states were identified after food and water consumption by outcome evaluation, which depended on the medial prefrontal cortex.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Robust discrimination between hydrated food and water.
Fig. 2: Decision-making about hunger or thirst.
Fig. 3: The mPFC is required for evaluating action–outcome relationships to inform decision-making when need states are uncertain.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Code availability

The code used to collect and analyze the data in this study is available upon reasonable request.

References

  1. Stevenson, R. J., Mahmut, M. & Rooney, K. Individual differences in the interoceptive states of hunger, fullness and thirst. Appetite 95, 44–57 (2015).

    Article  Google Scholar 

  2. Mattes, R. D. Hunger and thirst: issues in measurement and prediction of eating and drinking. Physiol. Behav. 100, 22–32 (2010).

    Article  CAS  Google Scholar 

  3. Betley, J. N. et al. Neurons for hunger and thirst transmit a negative-valence teaching signal. Nature 521, 180–185 (2015).

    Article  CAS  Google Scholar 

  4. Dennis, E. A. et al. Water consumption increases weight loss during a hypocaloric diet intervention in middle-aged and older adults. Obesity 18, 300–307 (2010).

    Article  Google Scholar 

  5. Kennedy, P. J. & Shapiro, M. L. Retrieving memories via internal context requires the hippocampus. J. Neurosci. 24, 6979–6985 (2004).

    Article  CAS  Google Scholar 

  6. Hull, C. L. Differential habituation to internal stimuli in the albino rat. J. Comp. Psychol. 16, 255–273 (1933).

    Article  Google Scholar 

  7. Watts, A. G. Dehydration-associated anorexia: development and rapid reversal. Physiol. Behav. 65, 871–878 (1999).

  8. Ramachandran, R. & Pearce, J. M. Pavlovian analysis of interactions between hunger and thirst. J. Exp. Psychol. Anim. Behav. Process. 13, 182–192 (1987).

    Article  CAS  Google Scholar 

  9. Kendler, H. H. & Levine, S. Studies of the effect of change of drive. I. From hunger to thirst in a T-maze. J. Exp. Psychol. 41, 429–436 (1951).

    Article  CAS  Google Scholar 

  10. Tataranni, P. A. et al. Neuroanatomical correlates of hunger and satiation in humans using positron emission tomography. Proc. Natl Acad. Sci. USA 96, 4569–4574 (1999).

    Article  CAS  Google Scholar 

  11. de Araujo, I. E., Kringelbach, M. L., Rolls, E. T. & McGlone, F. Human cortical responses to water in the mouth, and the effects of thirst. J. Neurophysiol. 90, 1865–1876 (2003).

    Article  Google Scholar 

  12. Andersson, B. & Larsson, S. Water and food intake and the inhibitory effect of amphetamine on drinking and eating before and after ‘prefrontal lobotomy’ in dogs. Acta Physiol. Scand. 38, 22–30 (1956).

    Article  CAS  Google Scholar 

  13. Land, B. B. et al. Medial prefrontal D1 dopamine neurons control food intake. Nat. Neurosci. 17, 248–253 (2014).

    Article  CAS  Google Scholar 

  14. Nakayama, H., Ibanez-tallon, I. & Heintz, N. Cell-type specific contribution of medial prefrontal neurons to flexible behaviors. J. Neurosci. 38, 4490–4505 (2018).

    Article  CAS  Google Scholar 

  15. Yiin, Y. M., Ackroff, K. & Sclafani, A. Flavor preferences conditioned by intragastric nutrient infusions in food restricted and free-feeding rats. Physiol. Behav. 84, 217–231 (2005).

    Article  CAS  Google Scholar 

  16. Balleine, B. W. in Neurobiology of Sensation and Reward (ed Gottfried, J. A.) (CRC Press/Taylor & Francis, 2011).

  17. Balleine, B. Instrumental performance following a shift in primary motivation depends on incentive learning. J. Exp. Psychol. Anim. Behav. Process. 18, 236–250 (1992).

    Article  CAS  Google Scholar 

  18. Corbit, L. H. & Balleine, B. W. The role of prelimbic cortex in instrumental conditioning. Behav. Brain Res. 146, 145–157 (2003).

    Article  Google Scholar 

  19. Shipman, M. L., Trask, S., Bouton, M. E. & Green, J. T. Inactivation of prelimbic and infralimbic cortex respectively affects minimally-trained and extensively-trained goal-directed actions. Neurobiol. Learn. Mem. 155, 164–172 (2018).

    Article  Google Scholar 

  20. Killcross, S. & Coutureau, E. Coordination of actions and habits in the medial prefrontal cortex of rats. Cereb. Cortex 13, 400–408 (2003).

    Article  Google Scholar 

  21. Zhao, S. et al. Cell type-specific channelrhodopsin-2 transgenic mice for optogenetic dissection of neural circuitry function. Nat. Methods 8, 745–752 (2011).

    Article  CAS  Google Scholar 

  22. Guo, ZengcaiV. et al. Flow of cortical activity underlying a tactile decision in mice. Neuron 81, 179–194 (2014).

    Article  CAS  Google Scholar 

  23. Horst, N. K. & Laubach, M. Reward-related activity in the medial prefrontal cortex is driven by consumption. Front. Neurosci. 7, 56 (2013).

    Article  Google Scholar 

  24. Jun, J. J. et al. Fully integrated silicon probes for high-density recording of neural activity. Nature 551, 232–236 (2017).

    Article  CAS  Google Scholar 

  25. Pachitariu, M., Steinmetz, N. A., Kadir, S. N., Carandini, M. & Harris, K. D. Fast and accurate spike sorting of high-channel count probes with KiloSort. Adv. Neural Inf. Process. Syst. 29, 4448–4456 (2016).

    Google Scholar 

  26. Rossant, C. et al. Spike sorting for large, dense electrode arrays. Nat. Neurosci. 19, 634–641 (2016).

    Article  CAS  Google Scholar 

  27. Gallistel, C. R., Fairhurst, S. & Balsam, P. The learning curve: implications of a quantitative analysis. Proc. Natl Acad. Sci. USA 101, 13124–13131 (2004).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This research was funded by the Howard Hughes Medical Institute. We thank K. Svoboda’s lab for assistance with neuropixel recordings and S. Michaels and A. Hu for histology. We thank M. Rose, M. McManus, R. Gattoni, S. Erwin, C. Morrow, A. Zeladonis and C. Lopez for mouse breeding and procedures. We also thank S. Lindo, R. Gattoni and A. Kozlosky for training control animals and A. Hantman, J. Dudman, U. Heberlein, A. Hermundstad and R. Egnor for comments on the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

A.-K.E. and S.M.S. initiated the project, designed experiments, analyzed the data and prepared the manuscript. A.-K.E., J.C. and J.A. designed, engineered and tested the behavioral apparatus. A.-K.E. and T.K. performed experiments and collected data. S.C. performed electrophysiological recordings. M.P. and S.C. analyzed electrophysiological data.

Corresponding author

Correspondence to Scott M. Sternson.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Kate Wassum and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Recordings of consummatory responses to hydrated food and water.

a, Number of mPFC and M2 neurons recorded in hunger and thirst. b, Proportion of response types from all recorded mPFC neurons (1180). Need-state-appropriate and need-state-inappropriate selective neurons comprise 64% of recorded mPFC neurons. Other: neurons that respond to hydrated food and water but are not selective. c-d, Population mean firing rates of all recorded neurons (c) and only M2 neurons (d) that prefer hydrated food (left) or water (right) in hunger (upper row) or thirst (lower row). Responses are aligned to spout extension (dashed line). e-f, Mean firing rate of mPFC neurons (e) and all recorded neurons (f) for food-selective (left) and water-selective (right) neurons during hunger (top) and thirst (bottom) in trials with lower and higher lick rates (darker and lighter colors, respectively). Trials were sorted by number of licks and split into two equal portions (see Methods). Mean of licks in each subgroup of trials is shown in the insets. The mean firing rate for each trial-subgroup in hunger and thirst is shown for food- and water-preferring neurons. The response magnitudes and the response differences between food- and water-preferring neurons are similar in subgroups of trials with lower or higher numbers of licks, consistent with prior reports23 [ref: Takenouchi, K. et al. Emotional and behavioral correlates of the anterior cingulate cortex during associative learning in rats. Neuroscience 93, 1271–1287 (1999)]. g, Decoding accuracy of neural responses to food and water in hunger and thirst using all recorded neurons. Thick lines represent mean. Error bars represent SEM.

Extended Data Fig. 2 Characterization of prefrontal cortex silencing by optogenetic activation of inhibitory interneurons.

a, Experimental timeline of consumption and optogenetic silencing during pre-consumption period (red) or consumption period (orange). Grey bar indicates consumption window. ITI: inter-trial interval. b, Proportion of mPFC neurons activated (yellow), unmodulated (grey), or inhibited (yellow or red) by optogenetic stimulation of VGAT neurons during pre-consumption (red, upper panel) or consumption period (orange, lower panel) in hunger (left panel) or thirst (right panel) c, Firing rate of example mPFC neurons in hunger (upper panel) and thirst (lower panel) without VGAT neuron photostimulation (photoinhibition) (left, aligned to spout in), photoinhibition during pre-consumption period (middle, aligned to laser onset) and consumption period (right, aligned to laser onset). Grey bar indicates consumption period. d, Population responses in inhibited mPFC neurons to photoinhibition during the pre-consumption period (left, aligned to spout in) or consumption period (right, aligned to first lick) in hunger (upper panel) and thirst (lower panel). Responses are shown for preferred and non-preferred outcome with and without stimulation. Insets expand initial VGAT neuron photostimulation period, scale bar: 0.2 s. e, Mean firing rate for mPFC neurons activated by optogenetic VGAT neuron stimulation during pre-consumption period (left, aligned to spout in) and consumption period (right, aligned to first lick). Insets expand the end of the VGAT neuron photostimulation period aligned to stimulation offset (dashed line), scale bar: 0.2 s. f, Mean firing rate for all recorded neurons with VGAT neuron photostimulation during pre-consumption period (upper panel, aligned to spout in) and consumption period (lower panel, aligned to first lick) for activated and inhibited neurons. Insets expand the end of the VGAT neuron photostimulation period aligned to stimulation offset (dashed line), scale bar: 0.2 s. Grey bar indicates consumption period. g, For successive optogenetic inhibition trials, no effect of cumulative optogenetic inhibition on firing rate for inhibited neurons during pre-consumption (left) and consumption (right) period for laser trials (see Methods). h, In freely moving mice, lick-triggered consumption of food and water without (grey) or paired with lick-triggered mPFC silencing (red) in hunger (n=12) and thirst (n=9). Thick lines represent mean. Error bars represent SEM. ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 3 Behavioral apparatus and training.

a, Diagram of behavioral apparatus. b, (Left) Individual sessions (green: hunger, blue: thirst) required to reach training criteria for lever presses in constant need state (grey shading) and alternating need state (brown shading). (Right) Mean training sessions in constant need state (grey, n=9 mice in hunger, n=9 mice in thirst) and after need state was switched (brown, n=27 mice). Box: IQR, red horizontal lines: median, whiskers: closest data points>1.5*IQR. c, Total lever presses for water (blue) and food (green) in constant need state (n = 7). d, Breakpoint ratio in hunger (green) and thirst (blue) for constant need state animals. e, Examples of training performance of need state switching animals with low (top) or high (bottom) food bias in initial training. Testing after each need state switch was separated by 3–4 days. Sessions in the same need state were either consecutive or every other day. Thick lines represent mean. Error bars represent SEM. ***P<0.001; ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 4 mPFC is not required for constant need state decision-making.

a-c, Effect of mPFC silencing during pre-choice phase in constant need state animals on error rate and preference index (a), reaction time of correct presses (b) and lick count (c) in hunger and thirst with (red) or without (grey) mPFC silencing. d-f, Effect of mPFC silencing during the outcome evaluation period in constant need state animals on error rate and preference index (d), reaction time of correct presses (e) and lick count (f) in hunger and thirst with (orange) or without (grey) mPFC silencing (regular n = 13 mice, stimulation n = 9 mice each during hunger and thirst). g, Anatomical location of optical fiber placement for constant need state animals. Gray circles indicate tip of optical fiber. Cg1: cingulate cortex area 1, Cg2: cingulate cortex area 2, DP: dorsal peduncular cortex, D3V: dorsal 3rd ventricle, fmi: forceps minor of the corpus callosum, IL: infralimbic cortex, M1: primary motor cortex, M2: secondary motor cortex, MO: medial orbital cortex, PrL: prelimbic cortex, VO: ventral orbital cortex. Thick lines represent mean. Error bars represent SEM. ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 5 Breakpoint analysis for food and water reward in hunger and thirst.

a, Cumulative presses for food (green) and water (blue) for one example animal throughout one breakpoint session in hunger (left) and thirst (right) b, Total lever presses for water (blue) and food (green) in a group of mice switching between hunger and thirst (n = 7). c, Breakpoint ratio in hunger (green) and thirst (blue) for switching need state animals. d, Post-reinforcer pausing in seconds as a measure of motivation during the early, middle (mid), or late part of the breakpoint session for switching animals. e, Total lever presses for water (blue) and food (green) in switching and constant need state conditions with pre-exposure to food and water in home cage. f, Breakpoint ratio in hunger (green) and thirst (blue) for switching and constant need state animals with pre-exposure to food and water in home cage (n = 7 mice for each group in hunger and thirst). g, Breakpoint ratio for switching and constant animals in regular or pre-exposure conditions separately for hunger (left) and thirst (right). Thick lines represent mean. Error bars represent SEM. *P<0.05; ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 6 Food bias and learning rates for decision-making in thirst and hunger.

a, Error rates for decisions in hunger and thirst with need state switching (n = 27) for mice with intermediate experience on the task under need state-switching conditions as well as for mice during constant need state (n = 18). b, Error rates during early decision-making sessions under need state-switching conditions just after the task had been learned (left) and after extensive experience with task in late sessions (right) (early n = 27 mice, late n = 22 mice). c, Transition trial in hunger and thirst when correct performance exceeded 80% correct during early, intermediate, and late sessions shows lag in correct performance in thirst due to initial food-seeking bias. Boxplot’s central mark indicates the median, bottom and top edges of the box indicate the 25th and 75th percentiles, respectively. The whiskers extend to the most extreme data points not considered outliers, and the outliers are plotted individually using the ‘+’ symbol. d, Example of Weibull fit27 to correct responses for one session in hunger (left) and thirst (right). e, Cumulative density function for the three Weibull fit parameters (n = 27 mice). (Left) offset a (*P<0.05), (middle) onset Latency L, (right) shape/steepness of function S. f, Comparison of initial food bias during regular switching (green, n=22) and after mice were held in constant thirst and switched back to hunger (black, n=7) matched for equal number of training sessions (see Methods for more details). g, To test if mice determine which lever gives the greatest reward for their current need in each session, we reversed the lever contingencies (‘lever reversal’, n = 14). In both hunger and thirst, the initial performance (first block of 10 trials) was opposite for the reversed contingencies with mice pressing the lever previously associated with the need-appropriate outcome. Boxplot’s central mark indicates the median, bottom and top edges of the box indicate the 25th and 75th percentiles, respectively. The whiskers extend to the most extreme data points not considered outliers, and the outliers are plotted individually using the ‘+’ symbol. Error bars represent SEM. *P<0.05; **P<0.01; ***P<0.001; ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 7 Body weight and decision-performance.

Mean (n = 17 mice) preference index (black) and body weight (red) over multiple behavioral sessions switching animals between food restriction (hunger, green shading) and water restriction (thirst, blue shading). Error bars represent SEM.

Extended Data Fig. 8 mPFC silencing during pre-choice and outcome evaluation.

a, Error rates (left) and reaction times (right) in the pre-choice stimulation period (n = 17 mice). b, Error rates (left) and reaction times (right) in the outcome evaluation period (n = 17 mice). c, Conditioned place preference test (n=12). Black horizontal lines: median, box: interquartile (25th-75th percentile) range (IQR), whiskers: closest data points>1.5*IQR. d, Venn Diagram showing animals affected by mPFC silencing during the pre-choice (n = 13 mice) and outcome evaluation period (n = 12 mice), with one animal being unaffected in both periods. Mice affected by optogenetic stimulation (responders) showed an effect on performance greater than 2 standard deviations above or below the mean error rate of 3 independent non-stimulation sessions. The same mice were used in both hunger and thirst conditions and thus differences in sensitivity to mPFC silencing in hunger and thirst are measured across the same animals. e, Number of animals showing an effect in pre-choice and outcome evaluation stimulation depending on unilateral vs. bilateral fiber implant. f, Anatomical location of non-responding animals during outcome evaluation stimulation. Thick lines represent mean. Error bars represent SEM. *P<0.05; **P<0.01; ***P<0.001; ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 9 Laser stimulation control experiments using C57BL/6J mice with bilateral fiber placement.

a, Within-session learning curve in hunger (green) and thirst (blue) during need state switching with no laser stimulation (n=16). b, Within-session learning curve in hunger (green) and thirst (blue) during laser stimulation on all trials during the outcome evaluation phase. c, Within-session learning curve in hunger (green) and thirst (blue) during laser stimulation during trials 25–50 and 75–100 during the pre-choice phase. d, Preference Index in hunger and thirst in session with no laser stimulation (grey), with laser stimulation during the outcome evaluation phase (orange), and with laser stimulation during pre-choice phase (red). e, Error rate in hunger and thirst in session with no laser stimulation (grey), with laser stimulation during the outcome evaluation phase (orange), and with laser stimulation during pre-choice phase (red). f, Comparison of reaction times in all three conditions in hunger and thirst for correct trials (left) and error trials (right). g, Bilateral fiber placement of C57BL/6J mice (n = 16). Thick lines represent mean. Error bars represent SEM. *P<0.05; **P<0.01; ***P<0.001; ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Extended Data Fig. 10 mPFC controls evaluative decision-making in both hunger and thirst.

a, Reaction time and licks per trial in thirst during regular need state switching (blue) and with mPFC silencing during outcome evaluation after holding thirst state constant for several sessions (orange) (n = 6 mice). b, Reaction time and licks per trial in hunger after being in constant thirst for several sessions without (green) and with mPFC silencing during outcome evaluation (orange) (n = 7 mice). Thick lines represent mean. Error bars represent SEM. *P<0.05; **P<0.01; ***P<0.001; ns, P>0.05. Detailed information about the exact test statistics, sidedness, and values are provided in Supplementary Table 1.

Supplementary information

Supplementary Information

Supplementary Fig. 1 and Supplementary Table 1.

Reporting Summary

Supplementary Video 1

Example session of a thirsty animal showing lever-press performance in trials before, during and after optogenetic pre-choice mPFC silencing. After a variable inter-trial interval, each trial begins with the presentation of an auditory cue (1 s), which is presented followed by levers extension. In mPFC silencing trials, laser stimulation starts with the onset of cue presentation and lasts until the animal presses a lever or until the response window ends (5 s).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Eiselt, AK., Chen, S., Chen, J. et al. Hunger or thirst state uncertainty is resolved by outcome evaluation in medial prefrontal cortex to guide decision-making. Nat Neurosci 24, 907–912 (2021). https://doi.org/10.1038/s41593-021-00850-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-021-00850-4

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing