You are on page 1of 8

Molecular and Cellular Endocrinology 518 (2020) 110986

Contents lists available at ScienceDirect

Molecular and Cellular Endocrinology


journal homepage: www.elsevier.com/locate/mce

Thermoregulation in fish
Martin Haesemeyer
The Ohio State University College of Medicine, Department of Neuroscience, Columbus, OH, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Thermoregulation is critical for survival and animals therefore employ strategies to keep their body temperature
Thermoregulation within a physiological range. As ectotherms, fish exclusively rely on behavioral strategies for thermoregulation.
Fish Different species of fish seek out their specific optimal temperatures through thermal navigation by biasing
Zebrafish
behavioral output based on experienced environmental temperatures. Like other vertebrates, fish sense water
Behavior
Circuit
temperature using thermoreceptors in trigeminal and dorsal root ganglia neurons that innervate the skin. Recent
Ethology research in larval zebrafish has revealed how neural circuits subsequently transform this sensation of temper­
ature into thermoregulatory behaviors. Across fish species, thermoregulatory strategies rely on a modulation of
swim vigor based on current temperature and a modulation of turning based on temperature change. Interest­
ingly, temperature preferences are not fixed but depend on other environmental cues and internal states. The
following review is intended as an overview on the current knowledge as well as open questions in fish
thermoregulation.

1. Introduction recent success of the zebrafish model in informing clinically relevant


research through mutant analysis and drug screening (Cully, 2019),
Metabolic rate and function strongly depend on temperature. While studying temperature perception in fish could have translational po­
temperatures below critical values limit cellular metabolism and ner­ tential here as well.
vous system function, high temperatures can lead to protein instability In the following article I review current knowledge on thermoregu­
and cell death. Animals therefore evolved strategies to regulate their latory behaviors in fish as well as the neural circuits controlling these
body temperature. These include behavioral strategies, such as seeking behaviors. I will start with an ethological perspective, review sensory
out environments of preferred temperature, as well as autonomous systems and central processing, and lay out interesting questions related
strategies employed in endothermic animals, such as heat generation by to the flexibility of thermoregulation. Thermoregulatory strategies are
brown fat or muscle shivering. However, even animals that can regulate often conserved across phyla, and I will therefore touch on comparative
their own body temperature generally employ behavioral strategies as aspects as well. In the end I will provide a brief outlook on potentially
well to conserve energy and to reduce water expenditure (Crawshaw and fruitful future research directions.
Stitt, 1975; Tan and Knight, 2018). Also, while muscle shivering ther­
mogenesis is part of the autonomic regulation of body temperature in 2. The behavioral and neural basis of fish thermoregulation
endotherms, it essentially co-opts a behavioral process. Interestingly,
temperature setpoints are flexible and adjust to metabolic needs of the 2.1. Ethological perspective
animal as seen in the circadian fluctuations of core body temperature in
mammals (Refinetti, 1999). A particularly striking example of temper­ Fish are largely ectothermic animals, with the exception of some
ature setpoint changes is the induction of fever, a process conserved species of tuna and sharks which can conserve and redistribute meta­
across many animal groups including ectotherms (Nakamura, 2011; bolic heat to raise their core body temperature (Carey et al., 1971). This
Rakus et al., 2017). Furthermore, dysregulation of thermal perception means that unless fish live in very stable environments, their core body
and core body temperature regulation have been found to be associated temperature fluctuates with the temperature of the water they occupy.
with both psychiatric disorders such as Tourette’s syndrome (Kessler, Therefore, even though many fish species can tolerate a wide tempera­
2002) as well as pharmacological agents that are used to treat neuro­ ture range, they rely on navigational strategies to keep their body
psychiatric disorders (Löffler et al., 2008; Singer, 2010). Given the temperature near a metabolic and reproductive optimum (Kennedy and

E-mail address: haesemeyer.1@osu.edu.

https://doi.org/10.1016/j.mce.2020.110986
Received 7 December 2019; Received in revised form 7 August 2020; Accepted 9 August 2020
Available online 22 August 2020
0303-7207/© 2020 Elsevier B.V. All rights reserved.
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

Mihursky, 1967). streams into rivers. In the case of larval zebrafish, habitat temperatures
Given the vast differences in temperature across water habitats on in the wild fluctuate widely since they live in shallow pools and streams
earth, ranging from near freezing in arctic regions to 30 ◦ C and higher in which can be easily heated locally by sunlight (Engeszer et al., 2007;
the tropics, it is not surprising that preferred temperatures vary widely Roger, 1996). This means that fish likely evolved navigational strategies
across fish species (Coutant, 1977). This suggests a co-adaptation be­ that allow them to seek out preferred temperatures in their habitat at
tween metabolic strategies and temperature preferences. Indeed, there is which their metabolism functions optimally.
evidence for such co-adaptation. E.g., vertical temperature differences in
Lake Stechlin have allowed sympatric speciation in the common ven­
2.2. Behavioral strategies
dace. Here, C. albula and C. fontanae occupy the same pelagic feeding
zone but have slightly different vertical distribution with metabolic
To reach their preferred water temperature, fish need to employ
adaptations to different temperature optima (Ohlberger et al., 2008). In
navigational strategies that maximize their time in the preferred envi­
Atlantic cod on the other hand there is a segregation in hemoglobin
ronment while allowing them to avoid harmful conditions. Fish indeed
polymorphisms with a HbI2 allele dominating in northern populations
possess such behavioral strategies since they have been demonstrated to
and a HbI1 allele dominating in warmer areas (Petersen and Steffensen,
aggregate in fairly narrow temperature ranges within their habitat and
2003). In the laboratory, cod homozygous for the HbI2 allele prefer
these preferences are often stable across years (Gammon, 1998).
lower temperature water (~8 ◦ C) than cod homozygous for the HbI1
Furthermore, fish can quickly adapt to changing conditions. Atlantic cod
allele (~15 ◦ C) (Petersen and Steffensen, 2003). This change in prefer­
will migrate towards the shore when changes in wind direction push
ence is aligned with differences in oxygen affinity in the two hemoglobin
warmer surface water out into the open ocean (Rose and Leggett, 1988).
variants, where HbI2 has higher oxygen affinity at lower temperatures,
Another opportunity to study thermal navigation in the wild is offered
while HbI1 is better at binding oxygen at higher temperatures (Brix
by warm plumes from power plant cooling water effluents into lakes and
et al., 1998; Karpov and Novikov, 1980). This suggests that even within
streams (Fig. 1A). Studies conducted in Lake Monona revealed that
a species, as the metabolism adapts to the environment, behavioral
different species of fish aggregate in narrow temperature zones within or
preferences change as well. This however is not always the case, as
away from the warm outflow area of a power plant (Neill and Magnuson,
temperature preferenda recorded within the lab diverge from general
1974). These distributions were notably stable even when food abun­
habitat temperatures for some fish species suggesting that adaptation
dance was higher in the outflow area, indicating that fish will avoid too
isn’t necessarily absolute (Beitinger and Fitzpatrick, 1979).
warm waters even if this reduces their access to food (Neill and Mag­
Importantly, temperatures in natural habitats of fish fluctuate across
nuson, 1974). Similarly, trout in Lake Leydoux restrict their distribution
both space and time. These fluctuations arise due to changes in wind
to cool areas of the lake (12–16 ◦ C) while the water reaches tempera­
direction, vertical temperature differences or influxes of cold or warm
tures greater than 26 ◦ C near the shore. Trout are able to maintain this

Fig. 1. Thermoregulatory behaviors. A:


Heated outflow from power plants or in­
fluxes of cold streams offer the opportunity
to study how different fish species distribute
within a temperature gradient in their nat­
ural habitat. B: Artificially created tempera­
ture gradients within the lab allow studying
navigation behavior in great detail and in
the absence of other potentially confounding
cues. C: In such controlled settings detailed
behavioral kinematics (such as the distance
d traveled in a swim or the turn angle α) can
be extracted and modeled in dependence of
previous sensory experience such as the
resting temperature T0 before the swim or
the previously experienced change in tem­
perature ΔT).. D: Work across fish (and in
fact many species navigating sensory gradi­
ents) has revealed a general strategy which
consists of modulating movement vigor by
current temperature and turning by tem­
perature changes.

2
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

distribution in spite of swimming average distances larger than 100 m a by increasing the frequency of swim bout initiation and the speed of
day, indicating active avoidance of areas of warm water (Goyer et al., individual swim bouts. Furthermore, larval zebrafish increase the fre­
2014). Furthermore, models based on the migration of salmon suggest quency of turning when the change of temperature experienced during
that these fish might use water temperature profiles to direct their the last swim is unfavorable (Fig. 1D) (Haesemeyer et al., 2015, 2019;
long-range migration (Booker et al., 2008; Byron and Burke, 2014). This Robson, 2013). While increases in turning aren’t directional with
suggests that fish not only use temperature to navigate thermal profiles respect to the gradient, larval zebrafish tend to string together swims of
themselves but that they can also use such information to accomplish the same turn direction (Dunn et al., 2016) thereby allowing trajectory
geographical navigation in general. reversals which aides heat gradient navigation (Robson, 2013). Notably,
Observing thermal navigation in the wild is critical to establish that the observed behavioral rules can be used to simulate behavior. Such
fish actually face thermal challenges in their habitat and that they simulations approximate true gradient navigation to a large degree,
attempt to move to ideal temperatures under natural conditions. How­ indicating that these rules are sufficient to explain the observed re­
ever, studying thermal navigation in the laboratory is essential to go striction to preferred temperatures (Robson, 2013).
beyond establishing temperature preferences and investigate the un­ How do larval zebrafish process temperature stimuli in order to enact
derlying behavioral algorithms animals use to navigate. In particular, these behavioral rules of gradient navigation? To address this question,
laboratory experiments can disambiguate influences of cues learned to we adopted a classical neuroscience paradigm, in which randomly
be correlated with temperature through experience or evolution (“pre­ fluctuating stimuli are used to map the ideal stimulus driving a neuron,
dictive thermal navigation”) from navigational strategies that depend on to a behavioral task. Using a precisely targeted laser beam following a
temperature sensation in the current environment (“reactive thermal freely swimming larval zebrafish allowed delivering such randomly
navigation”) (Neill, 1979). And indeed, many fish species have been fluctuating temperature stimuli while recording behavioral responses.
shown to seek out preferred temperatures in laboratory settings. In an This paradigm revealed that larval zebrafish integrate thermal infor­
elegant series of experiments Neill and Magnuson allowed fish to control mation in a time window of 500 ms preceding a swim bout. Further­
the temperature of their tanks to determine preferred temperature (Neill more, analyzing the ideal stimulus driving swim bouts demonstrated
and Magnuson, 1974; Neill et al., 1972). Namely, whenever the fish that larval zebrafish behavior is both driven by absolute temperature
swam into one side of a two-tank system, water temperature was and the rate of change of temperature within this short time window, in
increased, while swimming into the other side would lead to decreases line with the observed behavioral strategy (Haesemeyer et al., 2015).
in water temperature (Neill et al., 1972). Fish could thereby control the Intriguingly, while the ideal stimulus was largely similar across different
tank water temperature. Intriguingly, the temperature fish chose to set behaviors such as straight swims and turns, comparing it across behavior
the system to would largely agree with the temperature they preferred in types indicated that rapid temperature decreases towards the preferred
their natural habitat (Neill and Magnuson, 1974). This suggests that fish temperature led to fast straight movements (Haesemeyer et al., 2015).
can indeed navigate thermal gradients to find their optimal temperature. This effect favors staying the course when moving in the direction of
However, the strategy developed by fish in this paradigm is likely a mix favorable temperature changes as predicted based on simulations for
of reactive and predictive modes. general fish temperature navigation (Neill, 1979). This difference in
These experiments clearly reveal that fish can seek out preferred processing likely also underlies the modulation of turning based on
temperatures, but how do they accomplish this task? Experiments swimming direction within the temperature gradient (Robson, 2013).
recording fish behavior in relation to water temperature revealed that in Behavioral thermoregulatory strategies are present across animal
changing thermal environments such as a heat gradient, fish tend to phyla as well as in bacteria (Paster and Ryu, 2008). A shared element in
display the lowest movement vigor close to their preferred temperature all these navigational strategies is the modulation of turning based on
with swim vigor increasing as temperature departs from preferred whether the animal is moving towards or away from the preferred
values (Ivlev, 1960; Sullivan, 1954). This points to a strategy of “diffu­ temperature. C. elegans will lengthen straight runs when moving toward
sional trapping” or “orthokinesis”. Such a strategy by itself will indeed the preferred temperature and shorten runs, thereby increasing turning,
lead to accumulation near the preferred temperature (since animals tend when moving away (Ryu and Samuel, 2002). Unlike fish, when
to rest there) and away from less ideal conditions (since animals tend to C. elegans are close to the preferred temperature, they will engage in a
vigorously move and thereby leave those areas), however it is not precise isothermal tracking behavior, effectively moving along a line of
effective in avoiding noxious conditions (Fraenkel and Gunn, 1961). By near equal temperature (Hedgecock and Russell, 1975). Both adult
itself it is therefore far from an ideal strategy for survival (Neill, 1979). (Sayeed and Benzer, 1996) and larval Drosophila (Luo et al., 2010)
However, computer simulations suggest that adding a strategy based on navigate temperature gradients. Much like C. elegans, Drosophila
increased turning when temperatures are worsening, in other words, modulate run durations based on whether they are moving away or
when the change in temperature is unfavorable, could explain thermal towards the preferred temperature (Luo et al., 2010). However, while
preferences observed in fish without risking increased exposure to turn direction at the termination of a run is random in C. elegans,
noxious conditions (Neill, 1979). Notably, these strategies are akin to Drosophila larvae will sample the gradient using headsweeps at the end
the biased random walk displayed during bacterial chemotaxis (Alt, of a run and subsequently bias the turn in a favorable direction (Luo
1980; Berg, 1985). et al., 2010). As in zebrafish, behavioral decisions in C. elegans and
To delineate the behavioral algorithm underlying thermal navigation Drosophila larvae rely on temporal integration of heat stimuli (Klein
in fish requires to observe behavior at high resolution in many in­ et al., 2015; Ryu and Samuel, 2002).
dividuals while precisely controlling the stimuli available to the animal. To enact these behavioral strategies, fish and other animals need to
Larval zebrafish readily navigate heat gradients in the laboratory (Gau sense environmental temperature, represent this information in the
et al., 2013; Haesemeyer et al., 2015; Robson, 2013) and behavioral brain and ultimately transform the sensory information into appropriate
experiments can be performed with high enough throughput to obtain behaviors.
quantitative measures of behavior (Fig. 1B–C). Furthermore, larval
zebrafish swim in discrete bouts (Budick and O’Malley, 2000) which 2.3. The thermosensory system
greatly aids in the analysis of behavioral strategies, as each initiation of a
swim bout can be viewed as a separate behavioral decision (Johnson Across vertebrates, environmental temperature is sensed by neurons
et al., 2020; Marques et al., 2018). Behavioral quantification during of the trigeminal (Fig. 2A) and dorsal root ganglia that innervate the
gradient navigation revealed that larval zebrafish utilize two major facial skin and the skin of the trunk and tail, respectively (Butler and
behavioral strategies during thermal navigation. They will increase their Hodos, 2005). There is also strong evidence that at least some of the
swim vigor the further they are from their preferred temperature, both molecules, namely thermo Trp (transient receptor potential) channels

3
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

Fig. 2. Temperature sensation and process­


ing. A: Neurons of the trigeminal ganglion,
situated between the eye and the ear in
vertebrates, detect skin temperature across
the face. B: TrpV1 is one member of the
thermo-Trp family that gets activated by
higher temperatures. It passes cations when
activated, and thereby depolarizes the sen­
sory neuron. C: Schematic of the tempera­
ture processing circuit identified in larval
zebrafish (note, the brain shown is that of an
adult zebrafish). Environmental temperature
is detected by trigeminal (and dorsal root)
ganglia neurons which are exclusively
responding to temperature level. Local cir­
cuits in the hindbrain subsequently extract
the rate of temperature change and ulti­
mately control behavior generation.

(Jordt et al., 2003), involved in temperature sensation are even more so far no cold receptor has been identified in fish. The cold receptor
ancient and shared across divergent phyla. Their individual role how­ TrpM8 which is present in mammals, birds and amphibians is absent
ever varies widely across species. from fish genomes (Gracheva and Bagriantsev, 2015; Kastenhuber et al.,
In Drosophila, TrpA1 functions as a heat sensor and is required for 2013; Xu et al., 2014).
thermotaxis behavior (Rosenzweig et al., 2005). Similarly, in pit vipers In zebrafish, the responses of trigeminal neurons to temperature
TrpA1 is required for the detection of infrared radiation (Gracheva et al., stimuli have been studied. TrpV1 expressing neurons respond to
2010). In mammals, the role of TrpA1 is controversial. While TrpA1 has warming in vivo (Gau et al., 2013) and overall temperature stimuli seem
been reported as a sensor of cold and noxious cold (Bandell et al., 2004; to be encoded in two distinct functional classes of cells in the caudal part
Story et al., 2003), recent evidence suggests that it does not play a role in of the trigeminal ganglion (Haesemeyer et al., 2018). Here, one class of
thermosensation but rather in chemosensation, most notably the cells gets activated by warmth, potentially corresponding to the TrpV1
detection of mustard oil (Bautista et al., 2006; Knowlton et al., 2010). expressing population, while another population is activated by cold or
Furthermore, ablating TrpA1 expressing neurons in mice does not affect inhibited by warmth. In both these functional classes, activity encodes
cold avoidance (Pogorzala et al., 2013). In zebrafish, TrpA1 is not sen­ the current temperature level (Haesemeyer et al., 2018). Since behav­
sitive to temperature at all but also exclusively functions as a chemo­ ioral studies have shown that larval zebrafish use information both
sensor (Prober et al., 2008). The medaka TrpA1 homolog, on the other about the temperature level and the rate of change (Haesemeyer et al.,
hand, activates upon heating like the Drosophila variant (Oda et al., 2015), this raises the question of how the sensed environmental tem­
2017). The TrpV1 channel is generally heat sensitive across species perature is processed by downstream circuits.
(Fig. 2B). In vitro, mammalian TrpV1 activates at temperatures above
40 ◦ C (Jordt et al., 2003). Zebrafish TrpV1 starts to activate at slightly 2.4. Central circuits and behavior generation
lower temperatures, covering a temperature range between 25 and 42 ◦ C
(Gau et al., 2013). However, the role of TrpV1 in vivo is less well After environmental temperature is detected in sensory neurons, the
established and at least in mice it might be involved in sensing innoc­ brain needs to process this information in order to generate adaptive
uous rather than noxious heat similar to the zebrafish activation range behaviors such as heat gradient navigation or shivering thermogenesis.
(Yarmolinsky et al., 2016). Furthermore, while ablation of TrpV1 In general, little is known about central circuits processing temperature.
expressing neurons leads to profound deficits in heat responses in mice However, in mammals different brain regions have been implicated in
(Pogorzala et al., 2013), knockouts of the TrpV1 gene have only modest temperature processing (Solinski and Hoon, 2019). This includes the
effects (Caterina et al., 2000; Yarmolinsky et al., 2016). Similarly in preoptic hypothalamus for autonomous body temperature regulation as
zebrafish, morpholino mediated knockdown of TrpV1 impairs but does well as brainstem and thalamic regions that control behavioral programs
not abolish heat avoidance (Gau et al., 2013). This suggests that the while the cortex seems to be dispensable for thermoregulation (Yahiro
neurons expressing TrpV1 might be an important substrate in temper­ et al., 2017). In the common carp an fMRI study delineated a sensori­
ature sensation but that there is a redundancy in receptors. Interestingly, motor pathway that controls escape in response to rapid cooling stimuli.

4
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

This pathway involves the trigeminal nerve as well as the cerebellum hypothalamus which plays a key role in temperature regulation in
(Van den Burg et al., 2006). However, while very useful to implicate mammals (Haesemeyer et al., 2018; Nakamura, 2011). Temperature
brain regions, the resolution of fMRI precludes a cellular level encoding in these brain regions might serve higher order processes such
understanding. as learning. Through operant conditioning goldfish can learn to control
Larval zebrafish, due to its transparency and small size, allows the water temperature in their bowl using lever presses (Rozin and
cellular resolution functional imaging of the entire brain. This approach Mayer, 1961) and larval zebrafish can bias tail flick direction to turn off
not only allows to implicate brain regions in thermosensation but can a noxious heat stimulus (Lin et al., 2019). Such tasks may well rely on
identify different functional cell classes. Importantly, while many brain temperature processing in different brain regions. Furthermore, tem­
regions might have cells responding to a given stimulus, having access to perature preference is not absolute but modulated by internal state and
all cells allows inferring the most likely flow of information through the other environmental factors. Therefore, heat encoding cells in other
brain. Combining functional calcium imaging with heat stimuli and brain regions might be important for setpoint changes and behavioral
behavioral recording in larval zebrafish allowed to delineate how in­ flexibility.
formation is processed in the brain during thermoregulatory behaviors
(Haesemeyer et al., 2018). Temperature sensitive cells in the caudal
domain of the trigeminal ganglion encode environmental temperature in 2.5. Regulation of thermal responses
two channels: One ON channel that is activated by warmth and one OFF
channel that is inhibited. Circuits in the lateral hindbrain subsequently Just as metabolic needs can change, temperature preferences in an­
extract information about the rate of change of temperature by effec­ imals are often flexible and can vary dependent on internal state as well
tively differentiating the trigeminal inputs (Fig. 2C). Here, one new cell as environmental contingencies (Fig. 3). Upon infection, pyrogenic
type encodes the rate of temperature increase while another cell type is messengers induce fever in mammals during which the core body tem­
activated as the temperature is decreasing. This information is then used perature is elevated. A critical mediator of this pathway is the preoptic
by hindbrain pre-motor cells that drive behavior such as swimming and area of the hypothalamus which integrates information about external
turning. In other words, as predicted by behavioral studies, larval and internal body temperature (Liedtke, 2017; Nakamura, 2011).
zebrafish has access to information about the rate of temperature However, fever responses are not limited to endotherms, but many
change, which in this case is extracted by circuits in the hindbrain ectotherm species including fish employ behavioral means to raise their
(Haesemeyer et al., 2015, 2018). Having access to cellular level infor­ body temperature upon infection (Boltana et al., 2013; Casterlin and
mation together with behavior allowed to generate a circuit model of Reynolds, 1977; Reynolds and Casterlin, 1979). And our own experience
temperature processing that can successfully predict neuronal activity as well as experimental evidence tells us that fever responses also
and behavior in response to novel stimuli (Haesemeyer et al., 2018). include a behavioral component in mammals (Crawshaw and Stitt,
Such cellular level realistic circuit models have the advantage that they 1975). Intriguingly, the zebrafish preoptic area contains temperature
make testable predictions about circuit architecture which can be used encoding cells (Haesemeyer et al., 2018) which may indicate a conser­
to design targeted future experiments. Namely, the precise biophysical vation of pathways involved in modulating preferred body temperature.
mechanism of how hindbrain neurons compute the rate of temperature Changes in temperature preference on short timescales are observed
change from trigeminal inputs is still unknown, requiring further study. in the sculpin. After feeding, juvenile sculpin seek out warmer
Temperature processing in both mouse and Drosophila shares a
splitting into heat and cold activated channels with zebrafish (Frank
et al., 2015; Gallio et al., 2011; Liu et al., 2015; Ran et al., 2016). In
zebrafish specifically, these responses act in concert to control behavior,
such that heat and cold activated sensory neurons form separate ON and
OFF channels with overlapping temperature responses (Haesemeyer
et al., 2018). Such a feature of splitting information into ON and OFF
channels is widespread across sensory modalities and a similar over­
lapping mode of warmth detection has recently been confirmed in
mouse thermosensation (Paricio-Montesinos et al., 2020). In the mouse
trigeminal ganglion, sensory neurons tile temperature space with
different neurons activating at different temperatures (Yarmolinsky
et al., 2016). Presumed second order neurons in the spinal cord display a
coding strategy whereby warmth activated neurons largely encode the
temperature level while another class of cells specifically responds to
decreases in temperature (Ran et al., 2016). Second order temperature
neurons in Drosophila encode temperature level or rate of change in both
ON and OFF channels similar to zebrafish hindbrain cells (Frank et al.,
2015; Liu et al., 2015). However, while there are no cells in zebrafish
that respond to both warming and cooling such cells with broader
specificity have been found in Drosophila (Frank et al., 2015; Liu et al.,
2015). Furthermore, while sensory neurons in zebrafish exclusively
encode the temperature level, some Drosophila sensory neurons already Fig. 3. Control of temperature preference by internal states and environmental
specifically encode heating or cooling instead of warmth and cold factors. On an abstract level, thermoregulation can be compared to an engi­
(Budelli et al., 2019). This is similar to the C. elegans AFD sensory neuron neered thermostat controlling the temperature of an apartment. As in a ther­
which encodes a mixture of warmth and warming (Clark et al., 2006, mostat, larval zebrafish presumably compare (Blue triangle) current
temperature conditions (Thermometer) with an internal desired setpoint (Dial)
2007; Mori and Ohshima, 1995). This argues for both conservation in
to control thermoregulatory behaviors. The generated behavior on the other
the representation of temperature in the brain across species as well as
hand will move the fish to a different place and thereby feeds back on the
species specific adaptations. temperature experience (Bottom arrow). Work across fish species indicates that
However, aside from the hindbrain circuits that likely control different environmental and internal conditions influence the desired temper­
behavior, temperature is represented in many brain regions in zebrafish ature setpoint, effectively “turning the dial” of the thermostat to a new
on diverse time scales. This includes forebrain regions and the preoptic desired setting.

5
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

temperatures by vertically migrating towards the surface. Here, water is temperature processing may be used to suggest targeted experiments to
on average 10 ◦ C warmer than at the bottom. This migration to warmer study thermoregulation (Haesemeyer et al., 2018, 2019). This would
water in fact significantly enhances digestion and growth rates in reduce the experimental burden and might therefore increase feasibility
sculpin indicating an adaptive link between feeding state and thermo­ of testing central processing across species.
regulatory behavior (Wurtsbaugh and Neverman, 1988). Such feeding Beyond modeling studies, developmental adaptations within species
state dependent changes in preferred temperature are not limited to the can give important insight into how circuits could potentially adapt to
sculpin but are also observed in other fish (Reynolds and Casterlin, different thermal habitats. When Xenopus tadpoles are grown at cold
1979) indicating a general synchronization between metabolic need and temperatures, activation of TrpM8 channels in the spinal cord leads to
thermoregulation. an increase in spontaneous motor neuron activity. This activity increase
Hypoxic conditions tend to reduce the preferred temperature in fish ultimately results in a larger motor neuron pool through a reduction in
(Schurmann et al., 1991; Petersen and Steffensen, 2003; Schurmann and apoptosis (Spencer et al., 2019). This adaptation allows the tadpoles to
Steffensen, 1992, 1994; Schurmann et al., 1991). This is adaptive since readily perform escape maneuvers at lower temperature. This suggests
the metabolism and therefore oxygen need decreases at lower temper­ that sensitization through increases in the number of circuit elements
atures. At the same time, the higher oxygen solubility in cold water could form an important part of behavioral adaptations across species as
means that oxygen exchange via the gills becomes more efficient at well.
lower temperatures (Schurmann et al., 1991). In a temperature gradient Another fruitful future research avenue is to try and understand the
it is expected that cooler areas of the gradient will have more dissolved interaction of internal states and temperature preference. Based on
oxygen than warmer zones. While this is likely adding to the adaptive behavioral observations discussed above this includes questions as to
potential of lowering the temperature preference it obfuscates whether how hypoxic conditions are sensed by fish and how states of hypoxia,
fish navigate a thermal or an oxygen gradient. However, even when fish infection and satiation are integrated with the thermosensory system to
can control the temperature in a tank system without the presence of a bring about changes in preferred temperature.
gradient under constant hypoxic conditions they will still seek out a
lower preferred temperature (Schurmann et al., 1991). This argues that Conflicts of interest
there is a true interaction between sensing hypoxic conditions and
lowering preferred temperature. The author declares that no competing interests exist.
How might these different setpoint changes be regulated? Different
peptidergic cell populations in the zebrafish hypothalamus are impli­ Acknowledgements
cated in heat avoidance behavior (Lovett-Barron et al., 2020; Madelaine
et al., 2017) and overexpression of specific peptides such as cart and cgrp I thank The Ohio State University Wexner Medical Center for funding
enhance behavioral temperature responses (Woods et al., 2014). This and the anonymous reviewers for comments and suggestions on the
hints at an endocrine regulation of thermoregulatory behaviors which manuscript.
could influence the temperature fish ultimately seek out in a heat
gradient. Such a regulation could for example occur by peptidergic References
modulation of synaptic gain in the hindbrain circuits that control ther­
moregulatory behavior such as the heat circuit identified in larval Alt, W., 1980. Biased random walk models for chemotaxis and related diffusion
approximations. J. Math. Biol. 9, 147–177.
zebrafish. Bandell, M., Story, G.M., Hwang, S.W., Viswanath, V., Eid, S.R., Petrus, M.J., Earley, T.J.,
Patapoutian, A., 2004. Noxious cold ion channel TRPA1 is activated by pungent
3. Outlook compounds and bradykinin. Neuron 41, 849–857.
Bautista, D.M., Jordt, S.-E., Nikai, T., Tsuruda, P.R., Read, A.J., Poblete, J., Yamoah, E.N.,
Basbaum, A.I., Julius, D., 2006. TRPA1 mediates the inflammatory actions of
Thermoregulation is clearly important to fish and they display environmental irritants and proalgesic agents. Cell 124, 1269–1282.
thermoregulatory behavior such as heat gradient navigation in order to Beitinger, T.L., Fitzpatrick, L.C., 1979. Physiological and ecological correlates of
preferred temperature in fish. Integr. Comp. Biol. 19, 319–329.
reach their optimal body temperature. Circuits controlling these be­ Berg, H.C., 1985. Physics of bacterial chemotaxis. In: Sensory Perception and
haviors are beginning to be unraveled in the larval zebrafish model Transduction in Aneural Organisms. Springer US), pp. 19–30.
system. However, much could be learned from comparative studies Boltana, S., Rey, S., Roher, N., Vargas, R., Huerta, M., Huntingford, F.A., Goetz, F.W.,
Moore, J., Garcia-Valtanen, P., Estepa, A., et al., 2013. Behavioural fever is a
given the vast differences in preferred temperatures between fish. At this synergic signal amplifying the innate immune response. Proc. Biol. Sci. 280,
point it is entirely unclear where these differences in temperature 20131381.
preference arise: Are receptor proteins tuned to different temperature Booker, D.J., Wells, N.C., Smith, I.P., 2008. Modelling the trajectories of migrating
Atlantic salmon (Salmo salar). Can. J. Fish. Aquat. Sci. 65, 352–361.
ranges while the rest of the processing circuit is largely conserved? Or
Brix, O., Forås, E., Strand, I., 1998. Genetic variation and functional properties of
are the receptor proteins responding in a similar manner to environ­ Atlantic cod hemoglobins: introducing a modified tonometric method for studying
mental temperature but the encoding of thermo-channel opening into fragile hemoglobins. Comp. Biochem. Physiol. Mol. Integr. Physiol. 119, 575–583.
sensory neuron activity is different? Or does the difference in thermal Budelli, G., Ni, L., Berciu, C., van Giesen, L., Knecht, Z.A., Chang, E.C., Kaminski, B.,
Silbering, A.F., Samuel, A., Klein, M., et al., 2019. Ionotropic receptors specify the
preference arise due to differences in central processing such as changes morphogenesis of phasic sensors controlling rapid thermal preference in Drosophila.
in how the activity of hindbrain temperature encoding neurons in­ Neuron 101, 738–747 e3.
fluences pre-motor neurons? Due to modern day cellular molecular Budick, S.A., O’Malley, D.M., 2000. Locomotor repertoire of the larval zebrafish:
swimming, turning and prey capture. J. Exp. Biol. 203, 2565–2579.
techniques including vastly reduced costs in genome sequencing, dif­ Butler, A.B., Hodos, W., 2005. Comparative Vertebrate Neuroanatomy: Evolution and
ferences in molecular composition or sensory physiology could likely be Adaptation. John Wiley & Sons.
answered in in vitro systems. Differences in central processing are Byron, C.J., Burke, B.J., 2014. Salmon ocean migration models suggest a variety of
population-specific strategies. Rev. Fish Biol. Fish. 24, 737–756.
considerably harder to address in non-model systems. However, some Carey, F.G., Teal, J.M., Kanwisher, J.W., Lawson, K.D., Beckett, J.S., 1971. Warm-bodied
comparison species like medaka or Danionella translucida are being fish. Integr. Comp. Biol. 11, 137–143.
established as neuroscience model systems and could be tested for their Casterlin, M.E., Reynolds, W.W., 1977. Behavioral fever in anuran amphibian larvae. Life
Sci. 20, 593–596.
thermal preference (Cook et al., 2018; Oda et al., 2017; Schulze et al., Caterina, M.J., Leffler, A., Malmberg, A.B., Martin, W.J., Trafton, J., Petersen-Zeitz, K.R.,
2018). Since medaka is often found in cooler habitats with temperatures Koltzenburg, M., Basbaum, A.I., Julius, D., 2000. Impaired nociception and pain
around 20 ◦ C (Fukuda, 2009) it could be a useful model for comparative sensation in mice lacking the capsaicin receptor. Science 288, 306–313.
Clark, D.A., Biron, D., Sengupta, P., Samuel, A.D.T., 2006. The AFD sensory neurons
neuroethology in thermoregulation. Furthermore, models developed for
encode multiple functions underlying thermotactic behavior in Caenorhabditis
larval zebrafish thermosensation as well as artificial neural network elegans. J. Neurosci. 26, 7444–7451.
comparison models that make clear predictions about the logic of

6
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

Clark, D.A., Gabel, C.V., Gabel, H., Samuel, A.D.T., 2007. Temporal activity patterns in Luo, L., Gershow, M., Rosenzweig, M., Kang, K., Fang-Yen, C., Garrity, P.A., Samuel, A.D.
thermosensory neurons of freely moving Caenorhabditis elegans encode spatial T., 2010. Navigational decision making in Drosophila thermotaxis. J. Neurosci. 30,
thermal gradients. J. Neurosci. 27, 6083–6090. 4261–4272.
Cook, E.S.B., Barrios, J.P., Bertram, E.P.L., Douglass, A.D., 2018. Establishment of the Madelaine, R., Lovett-Barron, M., Halluin, C., Andalman, A.S., Liang, J., Skariah, G.M.,
Miniature Fish Species Danionella Translucida as a Genetically and Optically Leung, L.C., Burns, V.M., Mourrain, P., 2017. The hypothalamic NPVF circuit
Tractable Neuroscience Model. modulates ventral raphe activity during nociception. Sci. Rep. 7, 41528.
Coutant, C.C., 1977. Compilation of temperature preference data. J. Fish. Res. Board Marques, J.C., Lackner, S., Félix, R., Orger, M.B., 2018. Structure of the zebrafish
Can. 34, 739–745. locomotor repertoire revealed with unsupervised behavioral clustering. Curr. Biol.
Crawshaw, L.I., Stitt, J.T., 1975. Behavioural and autonomic induction of prostaglandin 28, 181–195 e5.
E-1 fever in squirrel monkeys. J. Physiol. 244, 197–206. Mori, I., Ohshima, Y., 1995. Neural regulation of thermotaxis in Caenorhabditis elegans.
Cully, M., 2019. Zebrafish earn their drug discovery stripes. Nat. Rev. Drug Discov. 18, Nature 376, 344–348.
811–813. Nakamura, K., 2011. Central circuitries for body temperature regulation and fever. Am.
Dunn, T.W., Mu, Y., Narayan, S., Randlett, O., Naumann, E.A., Yang, C.-T., Schier, A.F., J. Physiol. Regul. Integr. Comp. Physiol. 301, R1207–R1228.
Freeman, J., Engert, F., Ahrens, M.B., 2016. Brain-wide mapping of neural activity Neill, W.H., 1979. Mechanisms of fish distribution in heterothermal environments.
controlling zebrafish exploratory locomotion. Elife 5, e12741. Integr. Comp. Biol. 19, 305–317.
Engeszer, R.E., Patterson, L.B., Rao, A.A., Parichy, D.M., 2007. Zebrafish in the wild: a Neill, W.H., Magnuson, J.J., 1974. Distributional ecology and behavioral
review of natural history and new notes from the field. Zebrafish 4, 21–40. thermoregulation of fishes in relation to heated effluent from a power plant at Lake
Fraenkel, G.S., Gunn, D.L., 1961. The Orientation of Animals: Kineses, Taxes and Monona, Wisconsin. Trans. Am. Fish. Soc. 103, 663–710.
Compass Reactions. Dover, New York, USA. Neill, W.H., Magnuson, J.J., Chipman, G.G., 1972. Behavioral thermoregulation by
Frank, D.D., Jouandet, G.C., Kearney, P.J., Macpherson, L.J., Gallio, M., 2015. fishes: a new experimental approach. Science 176, 1443–1445.
Temperature representation in the Drosophila brain. Nature 519, 358–361. Oda, M., Saito, K., Hatta, S., Kubo, Y., Saitoh, O., 2017. Chemical and thermal sensitivity
Fukuda, S., 2009. Consideration of fuzziness: is it necessary in modelling fish habitat of medaka TRPA1 analyzed in heterologous expression system. Biochem. Biophys.
preference of Japanese medaka (Oryzias latipes)? Ecol. Model. 220, 2877–2884. Res. Commun. 494, 194–201.
Gallio, M., Ofstad, T.A., Macpherson, L.J., Wang, J.W., Zuker, C.S., 2011. The coding of Ohlberger, J., Mehner, T., Staaks, G., Hölker, F., 2008. Temperature-related
temperature in the Drosophila brain. Cell 144, 614–624. physiological adaptations promote ecological divergence in a sympatric species pair
Gammon, J.R., 1998. The Wabash River Ecosystem. Indiana University Press. of temperate freshwater fish. Coregonus spp. Funct. Ecol. 22, 501–508.
Gau, P., Poon, J., Ufret-Vincenty, C., Snelson, C.D., Gordon, S.E., Raible, D.W., Dhaka, A., Paricio-Montesinos, R., Schwaller, F., Udhayachandran, A., Rau, F., Walcher, J.,
2013. The zebrafish ortholog of TRPV1 is required for heat-induced locomotion. Evangelista, R., Vriens, J., Voets, T., Poulet, J.F.A., Lewin, G.R., 2020. The sensory
J. Neurosci. 33, 5249–5260. coding of warm perception. Neuron 106 (105), 830–841.e3. https://doi.org/
Goyer, K., Bertolo, A., Pépino, M., Magnan, P., 2014. Effects of lake warming on 10.1016/j.neuron.2020.02.035.
behavioural thermoregulatory tactics in a cold-water stenothermic fish. PLoS One 9, Paster, E., Ryu, W.S., 2008. The thermal impulse response of Escherichia coli. Proc. Natl.
e92514. Acad. Sci. U.S.A. 105, 5373–5377.
Gracheva, E.O., Bagriantsev, S.N., 2015. Evolutionary adaptation to thermosensation. Petersen, M.F., Steffensen, J.F., 2003. Preferred temperature of juvenile Atlantic cod
Curr. Opin. Neurobiol. 34, 67–73. Gadus morhua with different haemoglobin genotypes at normoxia and moderate
Gracheva, E.O., Ingolia, N.T., Kelly, Y.M., Cordero-Morales, J.F., Hollopeter, G., hypoxia. J. Exp. Biol. 206, 359–364.
Chesler, A.T., Sánchez, E.E., Perez, J.C., Weissman, J.S., Julius, D., 2010. Molecular Pogorzala, L.A., Mishra, S.K., Hoon, M.A., 2013. The cellular code for mammalian
basis of infrared detection by snakes. Nature 464, 1006–1011. thermosensation. J. Neurosci. 33, 5533–5541.
Haesemeyer, M., Robson, D.N., Li, J.M., Schier, A.F., Engert, F., 2015. The structure and Prober, D.A., Zimmerman, S., Myers, B.R., McDermott Jr., B.M., Kim, S.-H., Caron, S.,
timescales of heat perception in larval zebrafish. Cell Syst 1, 338–348. Rihel, J., Solnica-Krezel, L., Julius, D., Hudspeth, A.J., et al., 2008. Zebrafish TRPA1
Haesemeyer, M., Robson, D.N., Li, J.M., Schier, A.F., Engert, F., 2018. A brain-wide channels are required for chemosensation but not for thermosensation or
circuit model of heat-evoked swimming behavior in larval zebrafish. Neuron 98, mechanosensory hair cell function. J. Neurosci. 28, 10102–10110.
817–831 e6. Rakus, K., Ronsmans, M., Vanderplasschen, A., 2017. Behavioral fever in ectothermic
Haesemeyer, M., Schier, A.F., Engert, F., 2019. Convergent temperature representations vertebrates. Dev. Comp. Immunol. 66, 84–91.
in artificial and biological neural networks. Neuron 103, 1123–1134 e6. Ran, C., Hoon, M.A., Chen, X., 2016. The coding of cutaneous temperature in the spinal
Hedgecock, E.M., Russell, R.L., 1975. Normal and mutant thermotaxis in the nematode cord. Nat. Neurosci. 19, 1201–1209.
Caenorhabditis elegans. Proc. Natl. Acad. Sci. U.S.A. 72, 4061–4065. Refinetti, R., 1999. Amplitude of the daily rhythm of body temperature in eleven
Ivlev, V.S., 1960. Analysis of the mechanism of distribution of fishes under the conditions mammalian species. J. Therm. Biol. 24, 477–481.
of a temperature gradient. Zool. Zhurnal 39, 494–499. Reynolds, W.W., Casterlin, M.E., 1979. Behavioral thermoregulation and the “final
Johnson, R.E., Linderman, S., Panier, T., Wee, C.L., Song, E., Herrera, K.J., Miller, A., preferendum” paradigm. Am. Zool. 19, 211–224.
Engert, F., 2020. Probabilistic models of larval zebrafish behavior reveal structure on Robson, D.N., 2013. Thermal Navigation in Larval Zebrafish. Doctoral dissertation.
many scales. Curr. Biol. 30, 70–82.e4. Harvard University. http://nrs.harvard.edu/urn-3:HUL.InstRepos:11158267.
Jordt, S.-E., McKemy, D.D., Julius, D., 2003. Lessons from peppers and peppermint: the Roger, P.A., 1996. Biology and Management of the Floodwater Ecosystem in Rice Fields.
molecular logic of thermosensation. Curr. Opin. Neurobiol. 13, 487–492. Int. Rice Res. Inst.
Karpov, A.K., Novikov, G.G., 1980. Hemoglobin alloforms in cod, Gadus morhua Rose, G.A., Leggett, W.C., 1988. Atmosphere–ocean coupling and atlantic cod
(Gadiformes, Gadidae), their functional characteristics and occurrence in migrations: effect of wind-forced variations in sea temperatures and currents or
populations. J. Ichthyol. 20, 45–50. nearshore distributions and catch rates of Gadus morhua. Can. J. Fish. Aquat. Sci. 45,
Kastenhuber, E., Gesemann, M., Mickoleit, M., Neuhauss, S.C.F., 2013. Phylogenetic 1234–1243.
analysis and expression of zebrafish transient receptor potential melastatin family Rosenzweig, M., Brennan, K.M., Tayler, T.D., Phelps, P.O., Patapoutian, A., Garrity, P.A.,
genes. Dev. Dynam. 242, 1236–1249. 2005. The Drosophila ortholog of vertebrate TRPA1 regulates thermotaxis. Genes
Kennedy, V.S., Mihursky, J., 1967. Bibliography on the Effects of Temperature in the Dev. 19, 419–424.
Aquatic Environment. Natural Resources Institute, University of Maryland. Rozin, P.N., Mayer, J., 1961. Thermal reinforcement and thermoregulatory behavior in
Kessler, A.R., 2002. Tourette syndrome associated with body temperature dysregulation: the goldfish, Carassius auratus. Science 134, 942–943.
possible involvement of an idiopathic hypothalamic disorder. J. Child Neurol. 17, Ryu, W.S., Samuel, A.D.T., 2002. Thermotaxis in Caenorhabditis elegans analyzed by
738–744. measuring responses to defined Thermal stimuli. J. Neurosci. 22, 5727–5733.
Klein, M., Afonso, B., Vonner, A.J., Hernandez-Nunez, L., Berck, M., Tabone, C.J., Sayeed, O., Benzer, S., 1996. Behavioral genetics of thermosensation and hygrosensation
Kane, E.A., Pieribone, V.A., Nitabach, M.N., Cardona, A., et al., 2015. Sensory in Drosophila. Proc. Natl. Acad. Sci. U.S.A. 93, 6079–6084.
determinants of behavioral dynamics in Drosophila thermotaxis. Proc. Natl. Acad. Schulze, L., Henninger, J., Kadobianskyi, M., Chaigne, T., Faustino, A.I., Hakiy, N.,
Sci. U.S.A. 112, E220–E229. Albadri, S., Schuelke, M., Maler, L., Del Bene, F., et al., 2018. Transparent Danionella
Knowlton, W.M., Bifolck-Fisher, A., Bautista, D.M., McKemy, D.D., 2010. TRPM8, but not translucida as a genetically tractable vertebrate brain model. Nat. Methods 15,
TRPA1, is required for neural and behavioral responses to acute noxious cold 977–983.
temperatures and cold-mimetics in vivo. Pain 150, 340–350. Schurmann, H., Steffensen, J., 1994. Spontaneous swimming activity OF atlantic cod
Liedtke, W.B., 2017. Deconstructing mammalian thermoregulation. Proc. Natl. Acad. Sci. GADUS morhua exposed to graded hypoxia at three temperatures. J. Exp. Biol. 197,
U.S.A. 114, 1765–1767. 129–142.
Lin, Q., Helmreich, M., Schlumm, F., Li, J.M., Robson, D.N., Engert, F., Schier, A., Schurmann, H., Steffensen, J.F., 1992. Lethal oxygen levels at different temperatures and
Nöbauer, T., Vaziri, A., 2019. Cerebellar Neurodynamics during Motor Planning the preferred temperature during hypoxia of the Atlantic cod, Gadus morhua L.
Predict Decision Timing and Outcome on Single-Trial Level. J. Fish. Biol. 41, 927–934.
Liu, W.W., Mazor, O., Wilson, R.I., 2015. Thermosensory processing in the Drosophila Schurmann, H., Steffensen, J.F., Lomholt, J.P., 1991. The influence of hypoxia on the
brain. Nature 519, 353–357. preferred temperature of rainbow trout Oncorhynchus mykiss. J. Exp. Biol. 157,
Löffler, S., Danos, P., Schillen, T.B., Klimke, A., 2008. Recurrent dysregulation of body 75–86.
temperature during antipsychotic pharmacotherapy. Psychiatr. Prax. 35, 91–93. Singer, H.S., 2010. Treatment of tics and tourette syndrome. Curr. Treat. Options Neurol.
Lovett-Barron, M., Chen, R., Bradbury, S., Andalman, A.S., Wagle, M., Guo, S., 12, 539–561.
Deisseroth, K., 2020. Multiple convergent hypothalamus–brainstem circuits drive Solinski, H.J., Hoon, M.A., 2019. Cells and circuits for thermosensation in mammals.
defensive behavior. Nat. Neurosci. 23 (8), 959–967. https://doi.org/10.1038/ Neurosci. Lett. 690, 167–170.
s41593-020-0655-1. Spencer, K.A., Belgacem, Y.H., Visina, O., Shim, S., Genus, H., Borodinsky, L.N., 2019.
Growth at cold temperature increases the number of motor neurons to optimize
locomotor function. Curr. Biol. 29, 1787–1799 e5.

7
M. Haesemeyer Molecular and Cellular Endocrinology 518 (2020) 110986

Story, G.M., Peier, A.M., Reeve, A.J., Eid, S.R., Mosbacher, J., Hricik, T.R., Earley, T.J., Wurtsbaugh, W.A., Neverman, D., 1988. Post-feeding thermotaxis and daily vertical
Hergarden, A.C., Andersson, D.A., Hwang, S.W., et al., 2003. ANKTM1, a TRP-like migration in a larval fish. Nature 333, 846–848.
channel expressed in nociceptive neurons, is activated by cold temperatures. Cell Xu, P., Zhang, X., Wang, X., Li, J., Liu, G., Kuang, Y., Xu, J., Zheng, X., Ren, L., Wang, G.,
112, 819–829. et al., 2014. Genome sequence and genetic diversity of the common carp, Cyprinus
Sullivan, C.M., 1954. Temperature reception and responses in fish. J. Fish. Res. Board carpio. Nat. Genet. 46, 1212–1219.
Can. 11, 153–170. Yahiro, T., Kataoka, N., Nakamura, Y., Nakamura, K., 2017. The lateral parabrachial
Tan, C.L., Knight, Z.A., 2018. Regulation of body temperature by the nervous system. nucleus, but not the thalamus, mediates thermosensory pathways for behavioural
Neuron 98, 31–48. thermoregulation. Sci. Rep. 7, 5031.
Van den Burg, E.H., Verhoye, M., Peeters, R.R., 2006. Activation of a sensorimotor Yarmolinsky, D.A., Peng, Y., Pogorzala, L.A., Rutlin, M., Hoon, M.A., Zuker, C.S., 2016.
pathway in response to a water temperature drop in a teleost fish. J. Health.com 209 Coding and plasticity in the mammalian thermosensory system. Neuron 92,
(Pt 11), 2015–2024. https://doi.org/10.1242/jeb.02240. 1079–1092.
Woods, I.G., Schoppik, D., Shi, V.J., Zimmerman, S., Coleman, H.A., Greenwood, J.,
Soucy, E.R., Schier, A.F., 2014. Neuropeptidergic signaling partitions arousal
behaviors in zebrafish. J. Neurosci. 34, 3142–3160.

You might also like