A mechanistic explanation for the tolerance limits of animals at high temperatures is still missing, but one potential target for thermal failure is the electrical signaling off cells and tissues. With this in mind, here I review the effects of high temperature on the electrical excitability of heart, muscle and nerves, and refine a hypothesis regarding high temperature-induced failure of electrical excitation and signal transfer [the temperature-dependent deterioration of electrical excitability (TDEE) hypothesis]. A central tenet of the hypothesis is temperature-dependent mismatch between the depolarizing ion current (i.e. source) of the signaling cell and the repolarizing ion current (i.e. sink) of the receiving cell, which prevents the generation of action potentials (APs) in the latter. A source–sink mismatch can develop in heart, muscles and nerves at high temperatures owing to opposite effects of temperature on source and sink currents. AP propagation is more likely to fail at the sites of structural discontinuities, including electrically coupled cells, synapses and branching points of nerves and muscle, which impose an increased demand of inward current. At these sites, temperature-induced source–sink mismatch can reduce AP frequency, resulting in low-pass filtering or a complete block of signal transmission. In principle, this hypothesis can explain a number of heat-induced effects, including reduced heart rate, reduced synaptic transmission between neurons and reduced impulse transfer from neurons to muscles. The hypothesis is equally valid for ectothermic and endothermic animals, and for both aquatic and terrestrial species. Importantly, the hypothesis is strictly mechanistic and lends itself to experimental falsification.

Most animal life is limited to a temperature range between −2°C and +46°C (Clarke, 2014). This is the thermal window in which cellular macromolecules (e.g. nucleic acids, proteins and phospholipids) and body fluids allow the correct functioning of cellular life processes (Precht et al., 1973; Cossins and Bowler, 1987; Fields, 2001). Despite numerous studies on the temperature responses of animal bodies, organs, tissues, cells and macromolecules, the mechanistic basis of animal temperature tolerance is still unresolved. Although several ideas have been proposed as possible mechanisms for thermal failure, there are few formulated hypotheses that can guide research on the problem. The concept of oxygen- and capacity-limited thermal tolerance (OCLTT) has been the main formulated hypothesis on physiological thermal tolerance of ectotherms for approximately 20 years (Pörtner, 2001) (Box 1). This hypothesis proposes that the respiratory and cardiovascular systems are impaired at temperature extremes, thus limiting the availability of oxygen to tissues and cells. This results in hypoxemic deterioration of performance and fitness of the animal. The OCLTT hypothesis has had significant impact on research: its predictions have been tested experimentally in hundreds of studies. However, the OCLTT hypothesis is heavily disputed and it might be valid only for a limited number of (mainly) aquatic ectotherms (Clark et al., 2013; Lefevre, 2016). Clearly, alternative hypotheses are needed to guide and stimulate research on this fundamental and timely topic.

List of symbols and abbreviations
     
  • AP

    action potential

  •  
  • AV

    atrioventricular

  •  
  • CD

    critical depolarization

  •  
  • Cm

    membrane capacitance

  •  
  • ECG

    electrocardiogram

  •  
  • EE

    electrical excitability

  •  
  • fH

    heart rate

  •  
  • ICa

    inward Ca2+ current

  •  
  • ICl

    outward Cl current

  •  
  • INa

    inward Na+ current

  •  
  • IK1

    outward K+ current

  •  
  • [K+]o

    extracellular K+ concentration

  •  
  • OCLTT

    oxygen- and capacity-limited thermal tolerance

  •  
  • Ri

    resistance to axial current flow

  •  
  • Rin

    input resistance of a cell

  •  
  • Rm

    resistance of the resting plasma membrane

  •  
  • TBP

    break point temperature

  •  
  • TDEE

    temperature-dependent deterioration of electrical excitability

  •  
  • Vrest

    resting membrane potential

  •  
  • Vth

    threshold potential

Glossary

Axial current flow

The flow of current within the cell in the direction of action potential propagation.

Bradycardia

Slowing of heart rate.

Breakpoint temperature

The temperature at which the rate of a function starts to decrease.

Cardiac output

The volume of blood pumped by the heart in a unit of time.

Charge transfer

The number of charges transferred by an ion current, i.e. integral of current density.

Conduction block

Failure of impulse transmission along a nerve fiber or between electrically excitable cells.

Critical depolarization

The size of depolarization needed to trigger an action potential.

Current density

The size of ion current per unit membrane area.

His bundle

A component of the mammalian cardiac conduction system that extends from the atrioventricular node to the walls of the ventricles.

Hyperthermia

Abnormally high body temperature owing to the failure of thermoregulation.

Input resistance

Resistance of a cell membrane of the whole cell to current flow.

Membrane capacitance

The storage of electric charge by a lipid membrane.

P wave

A wave on an ECG that is generated by atrial depolarization.

Resistance of the resting plasma membrane

Resistance of cell membrane to current flow per unit membrane area.

Resting membrane potential

The negative membrane potential of a quiescent cell relative to external medium.

QRS complex

A waveform on an ECG that is generated by ventricular depolarization.

Safety factor

The excess charge delivered to a cell, beyond the amount required to trigger an action potential.

T wave

A wave of an ECG that is generated by ventricular repolarization.

Box 1. Oxygen- and capacity-limited thermal tolerance (OCLTT) and temperature-dependent deterioration of electrical excitability (TDEE)

The OCLTT hypothesis attempts to explain ecologically relevant temperature tolerance limits of animals as an outcome of thermally induced failure of aerobic physical and metabolic performance (aerobic scope, defined as the difference between standard and maximum metabolic rates). According to the OCLTT hypothesis, depression of aerobic scope at low and high temperatures is due to insufficient oxygen availability resulting from a temperature-induced decrease in respiratory and circulatory capacities. Thus, vital functions, such as movement, growth and reproduction, and the fitness of the animal are impaired (Pörtner, 2001, 2002). The OCLTT hypothesis is built on the early concepts of temperature dependence of aerobic swimming performance in fishes (Fry and Hart, 1948; Brett, 1971). Revival and reformulation of these concepts have had a strong stimulating effect on research in the field of thermal physiology. Some of the studies support the OCLTT hypothesis, whereas others contradict it. Currently, the hypothesis is highly disputed, mainly because of the ambiguity about the central tenets of the most recent version of the hypothesis (Pörtner et al., 2017), and the absence of clear-cut mechanistic cause–effect explanations. In its current form, the hypothesis is not considered to provide clear and testable predictions (Clark et al., 2013; Jutfelt et al., 2018). The source–sink mismatch concept, which is a central feature of TDEE, does not contradict the OCLTT hypothesis. In fact, concerning the heat tolerance of fishes, it provides a well-defined mechanistic explanation for high temperature-induced failure of cardiovascular function (Haverinen and Vornanen, 2020 preprint). In the OCLTT hypothesis, thermal tolerance is examined at the level of an intact organism and explanations are given in relation to lower-level functions (e.g. mitochondrial respiration) that are determined experimentally. In contrast to the top-down approach of the OCLTT hypothesis, TDEE explains the higher-level functions of the animal from the fundamental electrical properties of cellular signaling and information transfer, i.e. in a bottom-up manner. TDEE is not limited in application to the physical and metabolic performance of the animal but is also relevant to mental and cognitive functions such as sensation, learning and memory.

A scientific hypothesis should be simple and lead to clear-cut predictions, which can be tested experimentally. The results of such experiments should either conform to or refute the hypothesis, i.e. the hypothesis should, at least in principle, be falsifiable. A physiological hypothesis should also be truly mechanistic, explaining findings at all levels of biological organization, from molecules to cells, tissues, organs, individuals and the operation of animals in their natural habitats (Platt, 1964; Beard et al., 2005). Recently, I proposed a hypothesis of temperature-dependent deterioration of electrical excitability (TDEE) to explain the high temperature-induced depression of cardiac function in fish (Vornanen, 2016). Using a reductionist approach, TDEE starts from the molecular functions of a single cardiac myocyte in vitro and uses them to explain cardiac function in intact fish as it appears in the electrocardiogram (ECG) in vivo. In brief, temperature dependencies of the inward Na+ current (INa) and the outward K+ current (IK1) – two antagonistic ion currents critical for the initiation of cardiac action potentials (APs) (Varghese, 2016) – are widely different. INa is depressed at relatively low temperatures, whereas IK1 is heat resistant and increases almost linearly with rising temperature (Vornanen et al., 2014; Badr et al., 2017b). The temperature-induced mismatch between INa and IK1 is considered to result in temperature-dependent depression of electrical excitability (EE) (Vornanen, 2016).

The objective of this Review is twofold: to further elaborate the TDEE hypothesis and to expand its scope of application from the heart to neuronal and muscular tissues. To this end, I first define EE, then describe how differential effects of temperature on inward and outward ion currents of fish cardiac myocytes could impair EE at high temperatures. I go on to collect data from the literature on the temperature responses of muscle and nerves and integrate them into the TDEE hypothesis. Finally, I propose that a source–sink mismatch between the inward current flow from the signaling (upstream) cell and the outward current flow (ion leak) of the resting receiving (downstream) cell is common for all electrically excitable cells and tissues at high temperatures. At critically high temperatures, source–sink mismatch compromises sensory, motor and cardiac processes, with consequent impairment of higher-level functions such as neuronal integration, behavior, locomotion, circulation and body homeostasis.

There is a rich literature on different aspects of EE, but the term itself is rarely defined. According to Boyett and Jewell (1980), ‘Excitability of a cell can be defined loosely as the ease with which a response may be triggered. It is often expressed as the minimum current required to depolarize the membrane to the threshold potential’ (Vth, see Glossary). Vth is the value of membrane potential where the density (see Glossary) of voltage-activated inward currents exceeds the density of outward currents (membrane leak) and elicits an active membrane response, an all-or-none, self-propagating action potential (Fozzard, 1977; Koester and Siegelbaum, 2000) (Figs 1 and 2).

Fig. 1.

Effect of warming on electrical excitability (EE) of fish ventricular myocytes. (Ai) EE of a cell is dependent on passive [resistance of the resting plasma membrane (Rm), membrane capacitance (Cm), resistance to axial current flow (Ri)] and active [inward Na+ current (INa), (ICa), outward K+ current (IK1), (ICl)] properties of the cell membrane. (Aii) An action potential (AP) is initiated when the charge transfer by INa depolarizes the membrane from the resting membrane potential (Vrest) to the threshold potential (Vth). At Vth, charge transfer of INa exceeds that of IK1. A subthreshold stimulus causes a local non-propagating voltage change, which exponentially decays to Vrest. (Bi) At high temperature, the outward leak of K+ via IK1 hyperpolarizes Vrest: the critical depolarization (CD=VthVrest) becomes bigger and more inward current/charge is needed to depolarize the membrane to Vrest. (Bii,iii) When temperature rises, the charge transfer by INa becomes smaller owing to the fast inactivation of INa and because the increased K+ current density makes the AP shorter. If the reduction is large enough to depress the safety factor below 1.0 excitation will fail.

Fig. 1.

Effect of warming on electrical excitability (EE) of fish ventricular myocytes. (Ai) EE of a cell is dependent on passive [resistance of the resting plasma membrane (Rm), membrane capacitance (Cm), resistance to axial current flow (Ri)] and active [inward Na+ current (INa), (ICa), outward K+ current (IK1), (ICl)] properties of the cell membrane. (Aii) An action potential (AP) is initiated when the charge transfer by INa depolarizes the membrane from the resting membrane potential (Vrest) to the threshold potential (Vth). At Vth, charge transfer of INa exceeds that of IK1. A subthreshold stimulus causes a local non-propagating voltage change, which exponentially decays to Vrest. (Bi) At high temperature, the outward leak of K+ via IK1 hyperpolarizes Vrest: the critical depolarization (CD=VthVrest) becomes bigger and more inward current/charge is needed to depolarize the membrane to Vrest. (Bii,iii) When temperature rises, the charge transfer by INa becomes smaller owing to the fast inactivation of INa and because the increased K+ current density makes the AP shorter. If the reduction is large enough to depress the safety factor below 1.0 excitation will fail.

Fig. 2.

Effects of warming and increased capacitive load on safety factor (SF) and impulse transmission between ventricular myocytes. (Ai) At physiological body temperature, the upstream cell (source; upper bucket) provides more depolarizing charge into the downstream cell (sink; lower bucket) than is needed for electrical excitation, because ion leak of the resting cell is low. (Aii) At high temperature, ion leak becomes larger; thus, Vrest is more negative and critical depolarization (CD=VthVrest) becomes larger. Simultaneously, charge transfer by the upstream cell via INa reduces (shown as less liquid in the upper bucket), and therefore Vrest of the downstream cell cannot be depolarized to Vth. (B) When a small cell is connected to a larger cell, there is an increased capacitive load (a larger area of leaky membrane): the small cell cannot provide enough depolarizing charge to bring Vrest of the downstream cell to Vth.

Fig. 2.

Effects of warming and increased capacitive load on safety factor (SF) and impulse transmission between ventricular myocytes. (Ai) At physiological body temperature, the upstream cell (source; upper bucket) provides more depolarizing charge into the downstream cell (sink; lower bucket) than is needed for electrical excitation, because ion leak of the resting cell is low. (Aii) At high temperature, ion leak becomes larger; thus, Vrest is more negative and critical depolarization (CD=VthVrest) becomes larger. Simultaneously, charge transfer by the upstream cell via INa reduces (shown as less liquid in the upper bucket), and therefore Vrest of the downstream cell cannot be depolarized to Vth. (B) When a small cell is connected to a larger cell, there is an increased capacitive load (a larger area of leaky membrane): the small cell cannot provide enough depolarizing charge to bring Vrest of the downstream cell to Vth.

The electrical excitation of a cell requires that an inward current flows into the cell and depolarizes the resting membrane potential (Vrest; see Glossary) to Vth (Spector, 2013; Ma and Huguenard, 2013; Varghese, 2016). In multicellular tissue, the current flow occurs between electrically coupled cells. The excited signaling cell functions as a ‘source’ of depolarizing current, which flows into the quiescent receiving cell, the latter forming the current ‘sink’ owing to the ion leak of the membrane. Each cell in a multicellular tissue functions both as a source and sink for the current flow. Robust propagation of impulses through a tissue requires that the source current is larger than the sink current at each location of the cell/tissue. Voltage- and time-dependent inward Na+ (INa) and Ca2+ (ICa) currents usually form the current source, which promotes EE. The passive electrical properties of the cell constitute the current sink. Thus, EE is dependent on both passive and active properties of the cell membrane, which are discussed in more detail below (Koester and Siegelbaum, 2000; Khodorov, 2012).

Passive membrane responses

The ‘ease’ with which an electrical response is triggered in a cell is dependent on the passive electric properties of the excitable cell membrane at rest. It can be presented as a function of (i) resistance and (ii) capacitance in parallel at the plane of the plasma membrane, and (iii) the conductive/resistive intracellular solution, i.e. the cytoplasm of the cell (Fig. 1Ai) (Fozzard, 1977; Koester and Siegelbaum, 2000). Of these three passive membrane properties, the resistance of the resting plasma membrane (Rm; see Glossary) is most important for excitability, because it strongly influences the response of the cell to current flow from the neighboring cells. According to Ohm's law, Rm determines how much current needs to be injected into the cell to increase Vrest to the Vth of an AP (ΔV=I·Rm): the higher the Rm, the less inward current is needed to change the membrane potential from Vrest to VthV) and to trigger an AP.

The magnitude of Rm is dependent on ion flux through Na+-, Ca2+-, K+- and Cl-specific ion channels in the cell membrane (Fig. 1Ai). If there are many open channels in the membrane of the quiescent cell, the ion leakage is large, and Rm is low. The leakage current can be the flow of positive ions out of the cell (heart/nerve) or the flow of negative ions into the cell (muscle), both of which hyperpolarize the cell membrane. In resting cardiac myocytes, the ion leakage occurs mainly through Kir2 K+ channels, which maintain the negative Vrest (Hibino et al., 2010). If the leakage is large, a large inward current flow is needed to bring the membrane potential to Vth, because a greater part of the inward current is short-circuited by the leak (Ma and Huguenard, 2013). The second passive membrane property introduced above, membrane capacitance (Cm; see Glossary), is directly related to the surface area of the cell membrane, and it indicates how much charge the membrane can store upon voltage change. Cell size is important for excitability, because a large cell has more cell surface area and therefore a larger number of leak channels than a smaller cell (Koester and Siegelbaum, 2000). Finally, resistivity of the cytoplasm determines the resistance to axial current flow (Ri; ­see Glossary) and affects the rate of AP conduction. Experimentally, the composite of the passive electrical properties of a cell is measured by injecting a small current into the cell and recording the corresponding change in the membrane potential: input resistance (Rin; see Glossary) is obtained from the Ohm's equation (RinV/I). Rin takes into account all three passive membrane properties: it is proportional to Rm and Ri and inversely proportional to cell size (Cm). Because Rm is much (typically about two orders of magnitude) larger than Ri, Rin is a good measure for the membrane leak.

The passive electrical properties of a cell membrane are temperature dependent, and this has important implications for the function of electrically excitable cells. In this respect, Rm is the most important passive electrical property of the cell. When temperature increases, Rm decreases (because ion leakage increases) and Vrest becomes more negative. Therefore, more depolarizing current is needed to excite the cell to the Vth (Figs 1B, 2). This is a common property of the surface membrane of all excitable cells (Table 1). Ri is suggested to slightly decrease with temperature and, in some preparations, Cm may slightly increase with temperature (Adams, 1987, 1989b). Both changes will have a slight negative effect on excitability at high temperatures.

Table 1.

Temperature-induced changes in passive membrane properties and action potentials (APs) when temperature is acutely increased

Temperature-induced changes in passive membrane properties and action potentials (APs) when temperature is acutely increased
Temperature-induced changes in passive membrane properties and action potentials (APs) when temperature is acutely increased

Active membrane response

The ‘response’ of an electrically excitable cell to a suprathreshold stimulus current is an AP, which – once triggered – propagates through the cell with constant amplitude and velocity. A weaker subthreshold current elicits a smaller voltage change, which remains local and diminishes in amplitude with distance from the site of current injection (Fig. 1Aii).

EE is based on the asymmetric distribution of Na+, K+, Ca2+ and Cl ions across the cell membrane and passive flow of ions down their electrochemical gradients (Fozzard, 1977; Koester and Siegelbaum, 2000). The former is the result of energy-requiring ion transfer by ATP-driven pumps, and the latter is caused by the opening and closing of ion channels in the plasma membrane. A quiescent cell has a Vrest of −60 to −90 mV, with the inside of the cell being negative relative to the external solution. In cardiac myocytes and nerve cells, Vrest is largely due to diffusion of K+ out of the cell through K+-specific ion channels, whereas in skeletal muscle fibers, Cl channels can make a significant contribution to Vrest (Hutter and Noble, 1960; Adams, 1989b). K+ efflux and/or Cl influx maintain a value of Vrest that is close to the Nernst equilibrium potential of K+ and/or Cl ions. The initiation and conduction of an AP is a function of voltage- and time-dependent opening and closing of Na+-, Ca2+-, K+- and Cl-specific ion channels (Fozzard, 1977; Koester and Siegelbaum, 2000).

Safety factor

The robustness of AP propagation under varying physiological conditions and stresses is based on the excess supply of the inward charge of the active signaling cell relative to the charge demand of the receiving cell (Hodgkin, 1948; Spector, 2013). When this ratio is ≥1.0, impulse propagation will occur; if the ratio is <1.0, impulse propagation will fail (Fig. 2). The reserve capacity of the inward charge of the active cell is called the ‘safety factor’ (see Glossary; Rasminsky, 1973; Targ and Kocsis, 1985; Wood and Slater, 2001). Most estimates of the safety factor for the vertebrate heart vary between 1.5 and 3.5, whereas the value is placed at between 2 and 5 for the vertebrate neuromuscular junction (Harris and Ribchester, 1979; Wang and Rudy, 2000; Wood and Slater, 2001; Ruff, 2011). Therefore, in healthy tissues, small decreases in inward current or increases in outward leak current do not endanger the propagation of APs. However, the safety factor is not uniform throughout an excitable cell/tissue, and it may be lost under heat stress, as indicated by conduction block (see Glossary) in neurons, muscles and heart (Hodgkin and Katz, 1949; Rasminsky, 1973; Westerfield et al., 1978; Vornanen et al., 2014; Badr et al., 2016). The sites where impulse conduction is most vulnerable to failure are found at functional and structural discontinuities of cells and tissues, such as synapses, branching points of neurons and muscle fibers and coupled cardiac cells (Fig. 3). Such discontinuities can create a local source–sink mismatch – an imbalance between the current provided by a smaller mass of tissue (source) and the current necessary to bring to threshold an adjacent larger mass of tissue (sink). Source–sink mismatch can induce a local slowing of conduction, therefore resulting in reduced frequency of response, i.e. low-pass filtering or complete conduction block (Rasminsky, 1973; Aslanidi et al., 2009) (see ‘Functional consequences of impaired excitability’, below).

Fig. 3.

Effect of warming on impulse transmission. Transmission between cardiac myocytes (A), at electrical and chemical synapses (Bi,ii) and in a branching neuron (C). Line tracings show membrane potential during an AP at high (red) and low (blue) temperatures. AP transmission is impaired by a mismatch between inward and outward current/charge flow across the plasma membrane. Warming increases passive ion leak of the resting cell. Simultaneously, the inward charge transfer is decreased, owing to the reductions in duration and amplitude of AP. At high temperatures, the safety factor falls below 1.0, and excitation of the downstream cell fails. Branching points of axons represent sites of reduced safety factor due to a higher total capacitive load of the branches. Outward leak current is shown by blue and red arrows for low and high temperature, respectively. The inward charge during the upstroke of AP is reduced by warming owing to reduction in the duration and amplitude of AP. Functionally, the source–sink mismatch of ion current/charge flow may appear as a low-pass filter, resulting in reduced frequency of heartbeat and neuronal APs. d1, d2 and d3 are diameters of axon branches. Arrows show the direction of AP propagation; dashed arrows in C indicate the blocked propagation of AP.

Fig. 3.

Effect of warming on impulse transmission. Transmission between cardiac myocytes (A), at electrical and chemical synapses (Bi,ii) and in a branching neuron (C). Line tracings show membrane potential during an AP at high (red) and low (blue) temperatures. AP transmission is impaired by a mismatch between inward and outward current/charge flow across the plasma membrane. Warming increases passive ion leak of the resting cell. Simultaneously, the inward charge transfer is decreased, owing to the reductions in duration and amplitude of AP. At high temperatures, the safety factor falls below 1.0, and excitation of the downstream cell fails. Branching points of axons represent sites of reduced safety factor due to a higher total capacitive load of the branches. Outward leak current is shown by blue and red arrows for low and high temperature, respectively. The inward charge during the upstroke of AP is reduced by warming owing to reduction in the duration and amplitude of AP. Functionally, the source–sink mismatch of ion current/charge flow may appear as a low-pass filter, resulting in reduced frequency of heartbeat and neuronal APs. d1, d2 and d3 are diameters of axon branches. Arrows show the direction of AP propagation; dashed arrows in C indicate the blocked propagation of AP.

When temperature rises, the metabolic rate and oxygen demand of the tissues of ectothermic animals increases. It has been suggested that, in fish, oxygen convection of the circulatory system cannot keep pace with the increased oxygen demand, and thus the aerobic performance of the animal decreases (Farrell, 2009). Below, I describe how high temperature may cause heart failure as a result of the collapse of EE.

Cardiac output in fish is frequency-modulated under heat stress

Cardiac output (see Glossary) is the product of heart rate (fH) and stroke volume. When temperature is raised acutely, fH increases in fish up to a point called the breakpoint temperature (TBP; see Glossary), above which fH declines and finally ceases. In contrast to fH, stroke volume is largely independent of temperature (Stevens et al., 1972; Steinhausen et al., 2008; Mendonca and Gamperl, 2010). This means that cardiac output under acute temperature changes is closely correlated with fH (i.e. it is frequency modulated).

The rate and rhythm of the fish heart are determined by the primary pacemaker, the site of origin of cardiac APs at the border between the sinus venosus and the atrium (Yamauchi and Burnstock, 1968; Haverinen and Vornanen, 2007). APs propagate from the primary pacemaker to the atrium and, from there, via the atrioventricular (AV) canal to the ventricle, thus triggering sequential contractions of the atrium and ventricle. To understand the mechanism of temperature limitation of heart function, it is necessary to know the location(s) in the heart where temperature impairs AP generation/conduction. This can be achieved using electrocardiography, which produces electrocardiograms (ECGs) that show the familiar patterns of P waves, QRS complexes and T waves (see Glossary). Fig. 4A shows the beating rate of brown trout (Salmo trutta) cardiomyocytes as a function of temperature at the level of single pacemaker cells in vitro and in the intact animal in vivo (Haverinen et al., 2017). In single pacemaker cells, the frequency of APs increases almost exponentially in response to increases in temperature, reaching a rate of 191.4 beats min−1 at 24°C. By contrast, in the intact fish, fHV (calculated from QRS complexes) reaches its peak value of 76.5 beats min−1 at the TBP of 15.7°C. This strongly suggests that the warming-induced depression of fH is not caused by the failure of the primary pacemaker mechanism. ECGs taken from roach (Rutilus rutilus), rainbow trout (Oncorhynchus mykiss) and brown trout at temperatures above TBP show that atrial P waves occur at regular intervals, whereas several of the ventricular QRS complexes are missing (Vornanen et al., 2014; Badr et al., 2016). Indeed, there is often a regular 2:1, 3:1 or 4:1 P wave:QRS complex pattern, indicating various extents of AV block (Fig. 4C,D). This indicates that the atrium can follow the pacemaker rate, but that excitation of the ventricle fails, resulting in dissociation of atrial and ventricular beating rates: the temperature-induced bradycardia (see Glossary) is specific for the ventricle.

Fig. 4.

Warming-induced dissociation of sinoatrial and ventricular beating in fish hearts. (A) Action potential rate of sinoatrial pacemaker (PM) cells and ventricular rate (fHV) in vivo of 4°C-acclimated brown trout (Haverinen et al., 2017). (B) In vivo atrial and ventricular heart rates (fHA and fHV, respectively) of rainbow trout (acclimated at 12°C). Atrial and ventricular heart rates dissociate at approximately 25°C (Haverinen and Vornanen, 2020). (C,D) Original ECG tracings of brown trout acclimated at 4°C (C) and rainbow trout acclimated at 12°C (D) showing 2:1, 3:1 and 4:1 AV block. P, atrial depolarization.

Fig. 4.

Warming-induced dissociation of sinoatrial and ventricular beating in fish hearts. (A) Action potential rate of sinoatrial pacemaker (PM) cells and ventricular rate (fHV) in vivo of 4°C-acclimated brown trout (Haverinen et al., 2017). (B) In vivo atrial and ventricular heart rates (fHA and fHV, respectively) of rainbow trout (acclimated at 12°C). Atrial and ventricular heart rates dissociate at approximately 25°C (Haverinen and Vornanen, 2020). (C,D) Original ECG tracings of brown trout acclimated at 4°C (C) and rainbow trout acclimated at 12°C (D) showing 2:1, 3:1 and 4:1 AV block. P, atrial depolarization.

What causes the AV block? It could be caused either by the failure of impulse conduction in the AV canal or by an inability of the ventricular tissue to respond to APs arriving from the AV canal. In mammalian hearts, there are three main categories of AV block: first-, second- and third-degree blocks (Merideth and Pruitt, 1973). In the first-degree AV block, AP conduction in the AV node is delayed but there are no dropped ventricular beats (thus, the P wave is always associated with a QRS complex). The two types (type I and type II) of the second-degree AV block are both associated with the normal rhythm of P waves but with intermittent absence of QRS complexes. The type-I block is characterized by a delay in AV node conduction, which manifests in the ECG as a prolonged PQ interval. The high temperature-induced AV block in roach and trout is not associated with prolongation of the PQ interval, suggesting that the block is not analogous to the first-degree block or type I of the second-degree block of the mammalian heart. In mammals, the type-II second-degree block is also characterized by regular P waves with the intermittent absence of QRS complexes but does not involve prolongation of the PQ interval. This kind of ECG pattern is explained by the failure of impulse propagation in the conduction pathways of the ventricles. If the block is in the branch of the His bundle (see Glossary), the type-II AV block is characterized by widening of the QRS complex. In the third-degree AV block, impulse transmission to the ventricles is completely prevented. Interestingly, the temperature-induced AV block of the trout and roach hearts resembles the second-degree type-II AV block of the mammalian heart with skipped ventricular beats and widening of the QRS complex (Vornanen et al., 2014; Badr et al., 2016). This suggests that the temperature-induced AV block is not due to the failure of impulse conduction in the AV canal, but is somehow related to the depression of ventricular excitation. If temperature is high enough, it will cause a complete standstill of the ventricle (third-degree block). I discuss how this might occur in more detail below.

Heat-induced ventricular failure by source–sink mismatch

The excitability of atrial and ventricular myocytes is governed by two antagonistic ion currents: the inwardly rectifying K+ current (IK1) and the fast inward INa (Varghese, 2016). INa is the source current, which provides the necessary inward charge movement for depolarization of Vrest to Vth and determines the rate of AP propagation in the atrium and ventricle. IK1 forms the current sink by allowing the outward leak of K+ in quiescent myocytes. (Within the voltage range of an AP, IK1 is always an outward current, but the current decreases when membrane voltage becomes more positive than approximately −50 mV, i.e. it rectifies in inward direction.) At physiological temperatures, INa is large enough to enable normal excitation of the ventricle (Fozzard, 1977; Shaw and Rudy, 1997). However, at critically high temperatures, the safety factor reduces and may be totally lost owing to opposing effects of high temperature on INa and IK1 (Box 2). Initially, moderate increases in temperature increase the density of both INa and IK1, and the excitation of ventricular myocytes is ensured (Vornanen et al., 2014; Badr et al., 2017b; Badr et al., 2018). However, the charge transfer (see Glossary) by INa starts to decline immediately when temperature rises, because the fast inactivation of the channels allows less time for Na+ influx (Fig. 1Bii,iii). This means that the surplus of INa that was present at the acclimation temperature is lost when temperature approaches the upper thermal tolerance limit of the fish. The charge transferred by INa declines so much that it cannot exceed the leak by IK1; consequently, a mismatch between source and sink develops and excitation fails.

Box 2. Temperature dependence of ion channel function

Ion channels form water-filled pores in the plasma membrane that allow ion flux into and out of the cell. Opening and closing of the pore result from conformational changes of the channel: the pore becomes conducting only when both the activation and inactivation gate move to the open position. Ion flux stops when the inactivation gate closes. Notably, the inactivation gate may sometimes close before the activation gate opens, rendering the channel non-conducting (Armstrong, 2006). This is termed ‘closed-state inactivation’.

Activation energy for the voltage-gating of ion channels comes from changes in the membrane potential. However, the kinetic energy of temperature also contributes to activation energy: temperature affects the function of ion channels in the same way that it affects the function of enzyme proteins (Liang et al., 2009; Fields et al., 2015). Warming increases the rate of opening (activation) and closing (inactivation) of the gates and the size of the ion current (Collins and Rojas, 1982; Cavalié et al., 1985; Liang et al., 2009). The temperature dependence of ion channels consists of two components: (i) the effect of temperature on the movement of the gates of the channel and (ii) the effect of temperature on the passage of ions through the pore. The latter is relatively weakly dependent on temperature, with a 1.2- to 1.6-fold increase in rate for each 10°C increase in temperature (Q10 value). In contrast, conformational changes of the channel are strongly dependent on temperature, with typical Q10 values between 2 and 6.

Temperature sensitivity of ion channel function is also affected by the general physical properties and specific components of the lipid membrane and the cytoskeleton of the cell to which the channel may be attached (Maltsev and Undrovinas, 1997; Tillman and Cascio, 2003; Duncan et al., 2020). If the attached cytoskeletal protein and/or the surrounding lipid membrane are rigid, the gating is slow. If the temperature exceeds the optimum of the ion channel, then either the components of the channel assembly start to dissociate, or the conformational changes may become disorganized, with a consequent decrease in ion current (Charalambous et al., 2009).

The significance of the temperature-dependent kinetics of INa in the reduction of safety factor has been studied in rainbow trout ventricular myocytes. In these cells, the open Na+ channels quickly inactivate, and this happens faster at high temperatures, thereby reducing the integral of INa (Fig. 1Bii,iii). In rainbow trout ventricular myocytes, the peak density of INa at 25°C is approximately 10% larger than at 12°C, whereas the charge transfer by INa is 56% smaller at 25°C than 12°C (Haverinen and Vornanen, 2020). Collectively, experimental findings suggest that ventricular bradycardia is caused by high temperature-induced mismatch between source (INa) and sink (IK1). It is notable, however, that in enzymatically isolated ventricular myocytes, APs can be elicited much above the TBP of the in vivo fH (Vornanen et al., 2014). This indicates that when the stimulus – applied from an amplifier – is sufficiently large, ventricular myocytes can generate APs at temperatures considerably above the upper critical temperature of the fish. However, in the intact heart, the source current is finite and becomes a limiting factor.

Why are the nodal tissues and atrium resistant to thermal deterioration?

The ionic mechanisms underlying AP generation by the fish nodal tissues (the sinoatrial pacemaker and AV canal) are poorly understood, and therefore we must largely rely on data from other vertebrates in our attempts to understand why increases in temperature do not affect the nodal tissues and atrium (Efimov et al., 2004; Monfredi et al., 2010; George et al., 2017). In vertebrate hearts, the cells of the primary pacemaker and AV node do not have INa and IK1, or if they are expressed the currents are small. In the absence of IK1, the nodal myocytes do not have a stable Vrest but are self-excitatory, i.e. spontaneously active. The AP upstroke of nodal cells is based on the L-type Ca2+ current (ICaL), which is resistant to high temperatures in fish (Vornanen et al., 2014; Badr et al., 2017b). Thus, nodal cells do not experience the antagonism between INa and IK1, at least not to the same extent as ventricular myocytes.

In contrast to the nodal cells, fish atrial myocytes have a stable negative Vrest and they express IK1. However, the density of the atrial IK1 is small and therefore the Rin of atrial myocytes is about an order of magnitude higher than in ventricular myocytes (Vornanen et al., 2002; Haverinen and Vornanen, 2009; Abramochkin and Vornanen, 2015; Badr et al., 2017a). In contrast, the density of atrial INa is similar to or higher than that of ventricular myocytes (Haverinen and Vornanen, 2006; Badr et al., 2017b). The favorable INa:IK1 ratio (large safety factor) makes the atrial myocytes easily excitable and protects them against heat-dependent deterioration of excitability. It is noteworthy that EE of atrial myocytes can be depressed by inducing an outward K+ current by acetylcholine (IKAch) (Abramochkin and Vornanen, 2017). However, the temperature-induced increase in IKAch is not large enough to prevent atrial excitability in vivo, as shown by the persistence of P waves in the ECG (Fig. 4C,D).

Although the pacemaker rate is not suppressed by high temperatures, there are marked changes in the shape of the pacemaker AP as temperatures increase. Both the duration and amplitude of the AP decrease when temperature is acutely increased (Haverinen et al., 2017). Although not yet documented, similar effects are expected to occur in the AV nodal cells under warming. Because of their smaller and shorter AP, the AV nodal cells may be less able to provide depolarizing INa to trigger ventricular APs at higher temperatures.

Of course, AP propagation is not only important for cardiac function; for example, the transmission of different forms of sensory information and the activation and coordination of motor functions are mediated by trains of APs. Changes in the strength and quality of sensory information and in the force of muscle contraction are achieved by differences in the frequency of APs (i.e. sensory and motor information is largely frequency coded; Gerstner et al., 1997). Therefore, the same principles and mechanisms that apply to the thermal responses of the fish heart could also operate in nervous and muscular tissues under heat stress (Fig. 5).

Fig. 5.

Basic principles of the source–sink mismatch hypothesis. Warming causes mismatch between the depolarizing current of the active cell (source) and repolarizing currents of the resting (sink) cell, which increases the voltage and current threshold for excitation. When mismatch is so large that the safety factor drops below 1.0, excitation of the downstream cell will fail. In the heart, this causes ventricular bradycardia, whereas in nervous and muscle tissues it causes reduced frequency of conducted APs (i.e. low-pass filtering of information coding).

Fig. 5.

Basic principles of the source–sink mismatch hypothesis. Warming causes mismatch between the depolarizing current of the active cell (source) and repolarizing currents of the resting (sink) cell, which increases the voltage and current threshold for excitation. When mismatch is so large that the safety factor drops below 1.0, excitation of the downstream cell will fail. In the heart, this causes ventricular bradycardia, whereas in nervous and muscle tissues it causes reduced frequency of conducted APs (i.e. low-pass filtering of information coding).

Effects of high temperature on the current sink of neurons and muscle cells

High temperatures compromise the EE of neurons and muscle cells in both ectotherms and endotherms (White, 1983; Rutkove et al., 1997; Racinais and Oksa, 2010). Considering the potential importance of sarcolemmal ion leak in the source–sink mismatch of fish ventricular myocytes, the temperature dependence of the passive membrane properties of neurons and muscle cells is of interest. Fortunately, passive membrane properties have been extensively studied in nervous and muscular tissues, both in invertebrates and in ectothermic and endothermic vertebrates. A coherent picture is emerging from these studies. In general, high temperatures reduce Rin via increases in K+ or Cl leak across the cell membrane. Because the diffusion of K+ and/or Cl maintain Vrest in nerve and muscle cells, Vrest is regularly hyperpolarized at high temperatures (Table 1). Although the ion channels that mediate the resting leak may differ at least partly in heart, muscles and neurons, all tissues show increased outward current leak in response to high temperatures. Consequently, the current sink of neurons and muscle cells is larger at higher temperatures.

Effect of temperature on current source in neurons and muscle cells

INa is the main depolarizing current in most neurons, skeletal muscle fibers and atrial and ventricular myocytes, and it is also present in smooth muscle cells (Catterall, 2012; Ulyanova and Shirokov, 2018). Temperature-induced increases in the outward leak reduce the safety factor for AP generation/conduction unless the inward current is simultaneously increased (Adams, 1987). Several indirect findings indicate that rising temperatures have a greater (Q10) effect on outward K+ than inward Na+ currents, with a consequent reduction in the source-to-sink ratio (Hodgkin and Katz, 1949; Westerfield et al., 1978; Volgushev et al., 2000). Extracellular recordings of APs indicate that the amplitude and duration of APs are reduced in the myelinated axons of the frog (Rana pipiens) when temperature is raised from 8°C to 26°C (Schoepfle and Erlanger, 1941). Notably, the effect of temperature is much stronger on the descent (K+ current) than the ascent (Na+ current) of the AP. Similarly, Hodgkin and Katz (1949) found that higher temperatures (2–40°C) have a much greater effect on the duration of the falling than the rising phase of the AP in the giant axons of the squid (Loligo forbesi). This results in complete block of conduction at temperatures between 35°C and 40°C. Later, similar results were reported for the giant nerve of Loligo pealeii and several other preparations from different species (Westerfield et al., 1978) (Table 1). Humans are no exceptions in this respect; electromyograph recordings of APs have shown that an increase in muscle temperature of a few degrees diminishes the Na+ current of muscle fibers (Rutkove et al., 1997; Racinais et al., 2008).

With the advent of the voltage-clamp method it was possible to directly demonstrate that the thermal responses of APs are caused by differential effects of temperature on Na+ and K+ currents (Huxley, 1959). Still, relatively few studies have examined the effects of acute temperature changes on the size of neuron and muscle INa over a wide temperature range (most of this work has been done in fish cardiac myocytes). In mammals, the peak amplitude of the neuronal INa increases with rising temperature up to the normal body temperature (<36.5–37.5°C). However, at 37–43°C, the peak INa stops increasing or decreases (Volgushev et al., 2000; Touska et al., 2018). Relatively little is known about temperature effects on INa density; however, the temperature dependence of INa kinetics is better understood (Schwarz, 1979; Collins and Rojas, 1982; Schmidtmayer, 1989). In fact, the inactivation kinetics of Na+ channels are more important than the peak INa amplitude for excitability, because the amount of transferred charge (integral of INa) is decisive for excitation (Hodgkin, 1951). Na+ channels open very rapidly and spontaneously inactivate from the open state a few milliseconds later (Yue et al., 1989; Catterall, 2012). Because the opening of Na+ channels is almost instantaneous, changes in the rate of activation do not affect charge transfer as much as the rate of inactivation (Fig. 1Bii,iii). Owing to the accelerated inactivation of INa at higher temperatures, the charge transfer through Na+ channels decreases throughout the rising temperature range (Volgushev et al., 2000). Furthermore, at high temperatures, a larger fraction of Na+ channels could inactivate from the closed state (Box 2), which reduces the availability of Na+ channels for opening and therefore decreases the peak density of INa (Irvine et al., 1999). Collectively, a significant body of data indicates that charge transfer by INa decreases at rising temperatures and thereby reduces the safety factor in various excitable cells and tissues (Volgushev et al., 2000) (Table 1).

Functionally competent neuronal and neuromuscular connections are necessary for sensory, motor and behavioral responses of animals (Friedlander et al., 1976; Janssen, 1992). As long as neuronal, sensory and motor functions remain intact, normal body functions are maintained under acute and chronic temperature changes. Conversely, impairment of the nervous system is suggested to play a critical role in determining the upper and lower thermal limits of behavior and survival (Hamby, 1975; Prosser and Nelson, 1981; Cossins and Bowler, 1987). Complex polysynaptic behaviors are considered to be more susceptible to high temperature than simple reflexes (Prosser and Nelson, 1981; Montgomery and Macdonald, 1990). Thermal failure of excitability often appears as conduction block, in particular at branching points of cells and at synaptic connections between cells.

Conduction block

The normal functioning of the nervous system involves several processes that may be susceptible to thermal failure (i.e. initiation and conduction of APs, synaptic transmission) (Fig. 3B). Owing to the fast inactivation of INa and increase in the amplitude of K+ currents at high temperatures, acute temperature increases result in the reduction of AP duration and amplitude (Table 1). Small, short APs carry less inward current and are therefore weak triggers for AP initiation/conduction, eventually resulting in conduction block. If high temperature increases membrane leak, the small and short APs cannot provide enough depolarizing source current to trigger an AP in the receiving cell. Demyelinated axons are especially vulnerable to high temperature, because removal of the resistive myelin layer increases ion leakage of the axolemma. In fact, in the demyelinated axons of the rat spinal ventral root preparation, conduction block can appear at normal body temperature (Bostock et al., 1978). Notably, AP conduction can be restored by reducing temperature to below the body temperature or by exposing the axons to venom from the scorpion Leiurus quinquestriatus; both of these measures delay the inactivation of INa and prolong AP duration, and therefore increase source current (Bostock et al., 1978).

Another example of temperature-sensitive cardiac conduction failure appears in mice that are deficient in an intracellular Na+ channel modulator, the fibroblast growth factor homologous factor (FHF2) (Park et al., 2016). In mice missing this protein, INa inactivates faster and is therefore unable to provide enough depolarizing charge for excitation at high temperature (+40°C). In addition, temperature-dependent conduction block can also be observed in invertebrates: in the fruit fly (Drosophila), functional defects in neuronal Na+ channels result in temperature-sensitive phenotypes. In flies carrying a single gene mutation, parats, heating from 22–25°C to 29–38°C causes an almost instant but reversible paralysis of the skeletal muscles. This is due to temperature-induced failure of motoneuron function (Suzuki et al., 1971; Siddiqi and Benzer, 1976; Wu et al., 1978; Benshalom and Dagan, 1981). The number of functional Na+ channels is too low to raise Vrest to Vth (Loughney et al., 1989) because K+ currents overwhelm the small INa (O'Dowd et al., 1989), a clear source–sink mismatch. It is noteworthy that the flight and leg muscles of the fly remain fully functional, because depolarization of the muscle fiber is based on ICa (Ganetzky and Wu, 1986; Peron et al., 2009). Although the above examples are from animals with defective Na+ channels, a conduction block can be induced by high temperature in healthy animals and intact preparations (Hodgkin and Katz, 1949).

Branching

Once initiated, AP conduction is not particularly sensitive to high temperature when the AP is conducted along a structurally homogeneous plasma membrane (see ‘Conduction block’, above) (Li and Gouras, 1958; Jensen, 1972; Macdonald and Montgomery, 1982). However, nerves and muscle are geometrically complex tissues, with extensive branching and variations in cell diameter. In particular, the branching points of axons and muscle fibers are sensitive to conduction block when temperature rises (Spira et al., 1976; Westerfield et al., 1978; Grossman et al., 1979; Smith, 1980) (Figs 2B, 3). After the branching point, the total membrane area and therefore the summed membrane leak are bigger than those of the main axon (Fig. 3). Owing to this geometry, the branches have a low total Rin, and the safety factor is smaller at the branching point than before the branching point. This means that more inward charge is needed to drive Vrest of the branches to Vth (Swadlow et al., 1980). When the temperature rises, increased ion leak may depress the safety factor below 1.0 and prevent AP conduction to the branches (Fig. 3) (Westerfield et al., 1978; Smith, 1980). In the branching axon of the squid L.pealeii, a 2°C increase in temperature from 5°C to 7°C causes a conduction block. Mathematical modeling suggests that the failure occurs because the shortened AP is unable to provide enough inward charge to exceed the capacitive load (sink) of the branches (Westerfield et al., 1978). Similarly, in the crayfish Orconectes viris, motor axon conduction failure at the branching point is preceded by a reduction in INa, i.e. not only is the leak increased but the source current is also depressed (Smith, 1980). Clearly, the low safety factor of branching points makes these sites prone to conduction failure. Functionally, this appears as a reduction in the firing frequency of APs (Fig. 3C). This low-pass filtering of signals may set constraints on the motor and behavioral performance of animals under rising temperatures.

Synaptic transmission

Besides the branching points, synapses (both chemical and electrical) are thermally the weakest sites of nervous and motor systems (Krnjević and Miledi, 1959; Swadlow et al., 1980; White, 1983; Robertson and Money, 2012) (Fig. 3B). High temperatures have been shown to depress or cause complete failure of synaptic transmission in several invertebrates and vertebrates, including locusts (Locusta migratoria, Schistocerca gregaria), fruit fly (Drosophila melanogaster), crayfish (Procambarus clarkii), goldfish (Carassius auratus), frog (Xenopus laevis) and humans (Eusebi and Miledi, 1983; Burrows, 1989; Adams, 1989a; Dawson-Scully and Robertson, 1998; Heitler and Edwards, 1998; Barclay and Robertson, 2000). Depression of synaptic transmission is known to impair neuromuscular function in several invertebrate species, including locust (L. migratoria), fruit fly (D. melanogaster) and crayfish (P. clarkii); this is also observed in vertebrate ectotherms (e.g. C. auratus) and in humans (Friedlander et al., 1976; White, 1983; Karunanithi et al., 1999; Barclay and Robertson, 2000). Similar to all excitable membranes, ion leak of the postsynaptic membrane increases with warming (Heitler and Edwards, 1998). This may be one of the reasons why the amplitude of postsynaptic potentials decreases at high temperatures (Jensen, 1972; Macdonald and Montgomery, 1982). Even if the inward currents of the postsynaptic membrane were increased, they could be offset by the increased leak.

In the electrical synapses of invertebrates, the situation is analogous to the signal transmission between ventricular myocytes of the vertebrate heart. Charge is transferred directly, without the involvement of neurotransmitters, from the presynaptic neuron to the postsynaptic neuron or muscle. Experiments on the crayfish (P.clarkii and Pacifastacus leniusculus) neuromotor system have shown that synaptic transmission between the command nerve and the giant motoneuron fails at high temperature. In this system, the shortened presynaptic AP of the command nerve cannot provide enough depolarizing charge to the postsynaptic giant motoneuron, for which Rin is lowered by high temperature (Heitler and Edwards, 1998). Consequently, the excitatory postsynaptic potential remains too small to depolarize the membrane to the Vth.

Results on human subjects indicate that a 2°C increase in core temperature or local heating of the leg muscles reduces voluntary muscle activity (Dewhurst et al., 2005; Racinais et al., 2008). The decrement of activity is explained by failures in synaptic transmission at the neuromuscular junction owing to shortening of the depolarization time and therefore reduced Na+ entry through the muscle membrane, i.e. shortage of the source current (Rutkove, 2001; Racinais et al., 2008). Similar failures of synaptic transmission have been reported for Drosophila neuromuscular junction and mouse brain stem slices (Karunanithi et al., 1999; Kelty et al., 2002). Interestingly, the 2°C increase in body temperature in humans is associated with decreased memory capacity, which can be prevented by brain cooling (Racinais et al., 2008). These findings indicate that relatively mild hyperthermia (see Glossary) can weaken locomotor and learning skills at peripheral and central levels (i.e. thermal effects can appear as relatively subtle decrements in human/animal performance).

Collectively, these findings suggest that high temperature-induced increase in the sink current and/or decrease in the source current of either presynaptic or postsynaptic membrane can cause the failure of synaptic transmission.

Warming may also reduce EE via increases in extracellular K+ concentration ([K+]o, hyperkalemia), which can be substantial (2- to 3-fold) (Bowler, 1963; O'Sullivan et al., 2017). Hyperkalemia reduces source current and increases sink current. Increased [K+]o depolarizes Vrest with a consequent reduction in the number of available Na+ channels for opening (by steady-state inactivation); this depresses INa, reducing the rate of AP upstroke and conduction velocity (Nielsen and Gesser, 2001; Badr et al., 2018; Abramochkin et al., 2019). Increased [K+]o may occur when APs are generated at high frequency, because of the K+ efflux from the cell via the repolarizing K+ currents (Frankenhaeuser and Hodgkin, 1956; Spira et al., 1976). Tight diffusion-restricted extracellular space between cardiac myocytes and between the axolemma and myelin sheath around axons promotes increases in [K+]o (Frankenhaeuser and Hodgkin, 1956; Castel et al., 1976; Kline and Morad, 1978). Temperature- and frequency-dependent hyperkalemia stimulates outward K+ currents, further aggravating the situation (Badr et al., 2018). Reductions in INa and increases in outward K+ current reduce the amplitude and duration of APs, which lowers the safety factor and may result in frequency limitation (low-pass filtering) or a complete conduction block of APs (Parnas et al., 1976). Indeed, depression of the rate of AP upstroke and shortening of AP duration by high [K+]o has been documented for fish ventricular myocytes (Badr et al., 2018; Abramochkin et al., 2019).

The TDEE hypothesis is based on experiments on fish hearts (Vornanen, 2016). Here, the hypothesis is generalized to all excitable cells and tissues, including neurons, skeletal muscle fibers, cardiac myocytes and smooth muscle cells. The basic idea remains the same: high temperature affects inward currents (INa and ICa) and outward K+ currents (IK and ICl) disproportionally, and therefore Vrest cannot be depolarized to Vth at high temperatures, and excitation/conduction will fail (Vornanen, 2016). Because inward and outward currents of excitation may differ between cells and tissues, the central tenet of the hypothesis – the relationship between inward and outward currents – is now more generally expressed as a temperature-dependent mismatch between the depolarizing source current and the repolarizing sink current. Therefore, the hypothesis could be called the source–sink mismatch hypothesis (Fig. 5). Furthermore, in its current form, the hypothesis includes the concept of safety factor and indicates the thermally vulnerable tissue morphometries where the safety factor is low and may fall below the critical value of 1.0. According to the source–sink mismatch hypothesis, the function of electrically excitable cell membranes can be impaired at critically high temperatures. This agrees with Ken Bowler's notion that thermal limitation of animal life is a cellular membrane event (Bowler, 1981, 2018). However, the membrane malfunction may not require any cellular or molecular damage in order to compromise excitability: denaturation of ion channel proteins or damage to the lipid membrane need not occur. Instead, the impairment of EE occurs as an outcome of its molecular properties, which evolved for fast impulse transmission. High-frequency impulse transmission requires a large and fast inward charge transfer and rapid recovery of ion conductance after excitation. To this end, the voltage-gated Na+ channels evolved from the Ca2+ channels (Liebeskind et al., 2011). They provide a large, non-toxic and transient surge of inward charge without interfering with intracellular Ca2+ signaling (Nishino and Okamura, 2017). Na+ channels respond to membrane depolarization by almost instantaneous opening, followed by fast inactivation that terminates the Na+ influx (Mangold et al., 2017). This on–off behavior of voltage-gated Na+ channels underlies the coding of neuronal information and muscle function. Although the fast gating process is necessary for high-frequency neural coding, this ultimately results in temperature-dependent failure of excitation, as charge influx decreases with increasing temperature. Indeed, experiments on mouse pain receptors have shown that Na+ channels can operate at high temperatures (>46°C), if their kinetics are sufficiently slow. In the pain receptors this is possible because the relevant Na+ channel isoforms, NaV1.8 and NaV1.9, have very slow activation and inactivation kinetics (Touska et al., 2018).

The function of neuronal networks and smooth/skeletal/cardiac muscles are based on electrically excitable cells. If excitability of these tissues is threatened by high temperatures, is one of the tissues more likely to fail earlier than the other, or will they concurrently lose their function? Experiments on fish show that the functionally specialized parts of the heart are not equally sensitive to temperature (see ‘Heat-induced ventricular failure by source–sink mismatch’). A similar principle also applies to different parts of the neuromuscular system (Hodgkin, 1951; Orr, 1955b; Grainger, 1973). According to the source–sink mismatch hypothesis, the sensitivity of tissues to heat depends on the ion channel composition (source/sink channels), the structural complexity of tissues (branching) and the magnitude of the safety factor (surplus of the inward current). In addition, animals can defend tissues against high temperature by triggering various defenses (e.g. neuromodulation, heat shock response, differential gene expression) at different time scales (Ramirez et al., 1999; Haverinen and Vornanen, 2004; Money et al., 2005; Robertson and Money, 2012; Städele et al., 2015). Finally, in the light of this hypothesis, it is easy to see why the effects of temperature on vital functions and animal performance are gradual. At lower temperatures, dysfunction may be mild, and onset may be slow and difficult to detect (e.g. impairment of memory, learning and motor functions). At higher temperatures, the effects are faster and more intense, and they can quickly lead to death (e.g. cessation of heartbeat, loss of consciousness/equilibrium).

The source–sink mismatch hypothesis does not exclude any alternative hypotheses. Indeed, it is unlikely that there would be a single body function or process that limits the thermal tolerance of an animal. The limiting mechanism may differ depending on stage of ontogenetic development, phase of the life cycle, time scale of the temperature change or plasticity of the species' genomics (Orr, 1955a,b; Hollingsworth and Bowler, 1966; Davison, 1969; Irvin, 1974; Gracey et al., 2004; Vornanen et al., 2005). Practically all organs, tissues, cells and macromolecules are adapted to some extent to function properly in the thermal habitat of the animal: energy production by mitochondria, the structure of membrane lipids, the activity of enzymes and the oxidative stress response of cells may all be compromised under thermal stress. However, if an integrative model of thermal tolerance/failure, consisting of several body processes, is to be constructed (MacMillan, 2019), temperature dependence of EE should be an integral part of it.

The principles of the source–sink mismatch hypothesis should be valid for all multicellular animals, only excluding sponges (phylum Porifera) and placozoans (phylum Placozoa), which do not have excitable cells similar to the neurons and muscle fibers of other metazoans (Liebeskind, 2011). The hypothesis does not make any distinction between ectotherms and endotherms or between aquatic and terrestrial animals. Excitable cells are vital constituents of most tissues and organs and are needed for practically all basic functions of life, including sensation, locomotion, circulation, digestion and reproduction. Body homeostasis, learning and behavior are critically dependent on the function of excitable cells. In principle, each of these functions could be impaired by high temperature owing to the source–sink mismatch. The objective of outlining the source–sink hypothesis here is to stimulate research on this fundamental and timely topic of thermal physiology; in the future I hope to see more studies of thermal physiology from the perspective of EE. This would likely deepen our understanding of the effects of climate warming on animal life and motivate measures to mitigate its negative effects.

I would like to thank my post docs Jaakko Haverinen and Minna Hassinen and three anonymous reviewers for providing encouraging and constructive feedback on an earlier version of the Review.

Funding

This study was supported by a grant from the Academy of Finland to the author (grant 15051).

Abramochkin
,
D. V.
and
Vornanen
,
M.
(
2015
).
Seasonal acclimatization of the cardiac potassium currents (IK1 and IKr) in an arctic marine teleost, the navaga cod (Eleginus navaga)
.
J. Comp. Physiol. B.
185
,
883
-
890
.
Abramochkin
,
D. V.
and
Vornanen
,
M.
(
2017
).
Seasonal changes of cholinergic response in the atrium of Arctic navaga cod (Eleginus navaga)
.
J. Comp. Physiol. B
187
,
329
-
338
.
Abramochkin
,
D. V.
,
Haverinen
,
J.
,
Mitenkov
,
Y. A.
and
Vornanen
,
M.
(
2019
).
Temperature- and external K+-dependence of electrical excitation in ventricular myocytes of cod-like fishes
.
J. Exp. Biol.
222
,
jeb193607
.
Adams
,
B. A.
(
1987
).
Thermal dependence of passive electrical properties of lizard muscle fibers
.
J. Exp. Biol.
133
,
169
-
182
.
Adams
,
B. A.
(
1989a
).
Temperature and synaptic efficacy in frog skeletal muscle
.
J. Physiol.
408
,
443
-
455
.
Adams
,
B. A.
(
1989b
).
Temperature effects on membrane chloride conductance and electrical excitability of lizard skeletal muscle fibres
.
J. Exp. Biol.
144
,
551
-
563
.
Armstrong
,
C. M.
(
2006
).
Na channel inactivation from open and closed states
.
Proc. Natl. Acad. Sci. USA
103
,
17991
-
17996
.
Aslanidi
,
O. V.
,
Stewart
,
P.
,
Boyett
,
M. R.
and
Zhang
,
H.
(
2009
).
Optimal velocity and safety of discontinuous conduction through the heterogeneous Purkinje-ventricular junction
.
Biophys. J.
97
,
20
-
39
.
Badr
,
A.
,
El-Sayed
,
M. F.
and
Vornanen
,
M.
(
2016
).
Effects of seasonal acclimatization on temperature-dependence of cardiac excitability in the roach, Rutilus rutilus
.
J. Exp. Biol.
219
,
1495
-
1504
.
Badr
,
A.
,
Hassinen
,
M.
,
El-Sayed
,
M. F.
and
Vornanen
,
M.
(
2017a
).
Effects of seasonal acclimatization on action potentials and sarcolemmal K+ currents in roach (Rutilus rutilus) cardiac myocytes
.
Comp. Biochem. Physiol. A
205
,
15
-
27
.
Badr
,
A.
,
Korajoki
,
H.
,
Abu-Amra
,
E. S.
,
El-Sayed
,
M. F.
and
Vornanen
,
M.
(
2017b
).
Effects of seasonal acclimatization on thermal tolerance of inward currents in roach (Rutilus rutilus) cardiac myocytes
.
J. Comp. Physiol. B
188.2
,
255
-
269
.
Badr
,
A.
,
Abu-Amra
,
E. S.
,
El-Sayed
,
M. F.
and
Vornanen
,
M.
(
2018
).
Electrical excitability of roach (Rutilus rutilus) ventricular myocytes: effects of extracellular K+, temperature and pacing frequency
.
Am. J. Physiol.
315
,
R303
-
R311
.
Barclay
,
J. W.
and
Robertson
,
R. M.
(
2000
).
Heat-shock-induced thermoprotection of hindleg motor control in the locust
.
J. Exp. Biol.
203
,
941
-
950
.
Beard
,
D. A.
,
Bassingthwaighte
,
J. B.
and
Greene
,
A. S.
(
2005
).
Computational modeling of physiological systems
.
Physiol. Genomics
23
,
1
-
3
.
Benshalom
,
G.
and
Dagan
,
D.
(
1981
).
Electrophysiological analysis of the temperature-sensitive paralytic Drosophila mutant, parats
.
J. Comp. Physiol.
144
,
409
-
417
.
Bostock
,
H.
,
Sherratt
,
R. M.
and
Sears
,
T. A.
(
1978
).
Overcoming conduction failure in demyelinated nerve fibres by prolonging action potentials
.
Nature
274
,
385
.
Bowler
,
K.
(
1963
).
A study of the factors involved in acclimatization to temperature and death at high temperatures in Astacus pallipes. II. Experiments at the tissue level
.
J. Cell. Comp. Physiol.
62
,
133
-
146
.
Bowler
,
K.
(
1981
).
Heat death and cellular heat injury
.
J. Therm. Biol.
6
,
171
-
178
.
Bowler
,
K.
(
2018
).
Heat death in poikilotherms: is there a common cause?
J. Therm. Biol.
76
,
77
-
79
.
Boyett
,
M. R.
and
Jewell
,
B. R.
(
1980
).
Analysis of the effects of changes in rate and rhythm upon electrical activity in the heart
.
Prog. Biophys. Mol. Biol.
36
,
1
-
52
.
Brett
,
J. R.
(
1971
).
Energetic responses of salmon to temperature. A study of some thermal relations in the physiology and freshwater ecology of sockeye salmon
.
Am. Zool.
11
,
99
-
113
.
Burrows
,
M.
(
1989
).
Effects of temperature on a central synapse between identified motor neurons in the locust
.
J. Comp. Physiol. B
165
,
687
-
695
.
Castel
,
M.
,
Spira
,
M. E.
,
Parnas
,
I.
and
Yarom
,
Y.
(
1976
).
Ultrastructure of region of a low safety factor in inhomogeneous giant axon of the cockroach
.
J. Neurophysiol.
39
,
900
-
908
.
Catterall
,
W. A.
(
2012
).
Voltage-gated sodium channels at 60: structure, function and pathophysiology
.
J. Physiol.
590
,
2577
-
2589
.
Cavalié
,
A.
,
McDonald
,
T. F.
,
Pelzer
,
D.
and
Trautwein
,
W.
(
1985
).
Temperature-induced transitory and steady-state changes in the calcium current of guinea pig ventricular myocytes
.
Pflugers Arch.
405
,
294
-
296
.
Charalambous
,
K.
,
O'Reilly
,
A. O.
,
Bullough
,
P. A.
and
Wallace
,
B. A.
(
2009
).
Thermal and chemical unfolding and refolding of a eukaryotic sodium channel
.
BBA Biomembranes
1788
,
1279
-
1286
.
Clark
,
T. D.
,
Sandblom
,
E.
and
Jutfelt
,
F.
(
2013
).
Aerobic scope measurements of fishes in an era of climate change: respirometry, relevance and recommendations
.
J. Exp. Biol.
216
,
2771
-
2782
.
Clarke
,
A.
(
2014
).
The thermal limits to life on Earth
.
J. Astrobiol.
13
,
141
-
154
.
Collins
,
C. A.
and
Rojas
,
E.
(
1982
).
Temperature dependence of the sodium channel gating kinetics in the node of Ranvier
.
Quart. J. Exp. Physiol.
67
,
41
-
55
.
Colton
,
C. K.
and
Freeman
,
A. R.
(
1975
).
Dual response of lobster muscle fibers to L-glutamate
.
Comp. Biochem. Physiol. C
51
,
275
-
284
.
Cossins
,
A. R.
and
Bowler
,
K.
(
1987
).
Temperature Biology of Animals
.
Canbridge
:
Cambridge University Press
.
Dalton
,
J. C.
and
Hendrix
,
D. E.
(
1962
).
Effects of temperature on membrane potentials of lobster giant axon
.
Am. J. Physiol.
202
,
491
-
494
.
Davison
,
T. F.
(
1969
).
Changes in temperature tolerance during the life cycle of Calliphora erythrocephala
.
J. Insect Physiol.
15
,
977
-
988
.
Dawson-Scully
,
K.
and
Robertson
,
R. M.
(
1998
).
Heat shock protects synaptic transmission in flight motor circuitry of locusts
.
Neuroreport
9
,
2589
-
2593
.
Dewhurst
,
S.
,
Riches
,
P. E.
,
Nimmo
,
M. A.
and
De Vito
,
G.
(
2005
).
Temperature dependence of soleus H-reflex and M wave in young and older women
.
Eur. J. Appl. Physiol.
94
,
491
-
499
.
Dierolf
,
B. M.
and
McDonald
,
H. S.
(
1969
).
Effects of temperature acclimation on electrical properties of earthworm giant axons
.
J. Comp. Physiol. A
62
,
284
-
290
.
Duncan
,
A. L.
,
Song
,
W.
and
Sansom
,
M. S. P.
(
2020
).
Lipid-dependent regulation of ion channels and G protein–coupled receptors: insights from structures and simulations
.
Annu. Rev. Pharmacol. Toxicol.
60
,
31
-
50
.
Efimov
,
I. R.
,
Nikolski
,
V. P.
,
Rothenberg
,
F.
,
Greener
,
I. D.
,
Li
,
J.
,
Dobrzynski
,
H.
and
Boyett
,
M.
(
2004
).
Structure-function relationship in the AV junction
.
Anat. Rec.
280
,
952
-
965
.
Eusebi
,
F.
and
Miledi
,
R.
(
1983
).
Divalent cations and temperature dependent block of impulse propagation at the frog neuromuscular junction
.
Muscle Nerve
6
,
602
-
605
.
Farrell
,
A. P.
(
2009
).
Environment, antecedents and climate change: lessons from the study of temperature physiology and river migration of salmonids
.
J. Exp. Biol.
212
,
3771
-
3780
.
Fatt
,
P.
and
Katz
,
B.
(
1953
).
The electrical properties of crustacean muscle fibres
.
J. Physiol.
120
,
171
-
204
.
Fields
,
P. A.
(
2001
).
Review: protein function at thermal extremes: balancing stability and flexibility
.
Comp. Biochem. Physiol. A
129A
,
417
-
431
.
Fields
,
P. A.
,
Dong
,
Y.
,
Meng
,
X.
and
Somero
,
G. N.
(
2015
).
Adaptations of protein structure and function to temperature: there is more than one way to ‘skin a cat
’.
J. Exp. Biol.
218
,
1801
-
1811
.
Fozzard
,
H. A.
(
1977
).
Cardiac muscle: excitability and passive electrical properties
.
Prog. Cardiovasc. Dis.
19
,
343
-
359
.
Frankenhaeuser
,
B.
and
Hodgkin
,
A. L.
(
1956
).
The after-effects of impulses in the giant nerve fibres of Loligo
.
J. Physiol.
131
,
341
-
376
.
Friedlander
,
M. J.
,
Kotchabhakdi
,
N.
and
Prosser
,
C. L.
(
1976
).
Effects of cold and heat on behavior and cerebellar function in goldfish
.
J. Comp. Physiol.
112
,
19
-
45
.
Fry
,
F. E. J.
and
Hart
,
J. S.
(
1948
).
The relation of temperature to oxygen consumption in the goldfish
.
Biol. Bull.
94
,
66
-
77
.
Ganetzky
,
B.
and
Wu
,
C. F.
(
1986
).
Neurogenetics of membrane excitability in Drosophila
.
Annu. Rev. Genet.
20
,
13
-
44
.
George
,
S. A.
,
Faye
,
N. R.
,
Murillo-Berlioz
,
A.
,
Lee
,
K. B.
,
Trachiotis
,
G. D.
and
Efimov
,
I. R.
(
2017
).
At the atrioventricular crossroads: dual pathway electrophysiology in the atrioventricular node and its underlying heterogeneities
.
Arrhythm. Electrophysiol. Rev.
6
,
179
-
185
.
Gerstner
,
W.
,
Kreiter
,
A. K.
,
Markram
,
H.
and
Herz
,
A. V. M.
(
1997
).
Neural codes: firing rates and beyond
.
Proc. Natl. Acad. Sci. USA
94
,
12740
-
12741
.
Gladwell
,
R. T.
,
Bowler
,
K.
and
Duncan
,
C. J.
(
1976
).
Heat death in the crayfish Austropotamobius pallipes—ion movements and their effects on excitable tissues during heat death
.
J. Therm. Biol.
1
,
79
-
94
.
Gracey
,
A. Y.
,
Fraser
,
E. J.
,
Li
,
W.
,
Fang
,
Y.
,
Taylor
,
R. R.
,
Rogers
,
J.
,
Brass
,
A.
and
Cossins
,
A. R.
(
2004
).
Coping with cold: An integrative, multitissue analysis of the transcriptome of a poikilothermic vertebrate
.
Proc. Natl. Acad. Sci. USA
101
,
16970
-
16975
.
Grainger
,
J. N. R.
(
1973
).
A study of heat death in nerve-muscle preparations of Rana temporaria L
.
P. R. Irish Acad. B
73
,
283
-
290
.
Grossman
,
Y.
,
Parnas
,
I.
and
Spira
,
M. E.
(
1979
).
Differential conduction block in branches of a bifurcating axon
.
J. Physiol.
295
,
283
-
305
.
Hamby
,
R. J.
(
1975
).
Heat effects on a marine snail
.
Biol. Bull.
149
,
331
-
347
.
Harri
,
M. N. E.
and
Florey
,
E.
(
1977
).
The effects of temperature on a neuromuscular system of the crayfish, Astacus leptodactylus
.
J. comp. Physiol.
117
,
47
-
61
.
Harris
,
J. B.
and
Ribchester
,
R. R.
(
1979
).
The relationship between end-plate size and transmitter release in normal and dystrophic muscles of the mouse
.
J. Physiol.
296
,
245
-
265
.
Haverinen
,
J.
and
Vornanen
,
M.
(
2004
).
Temperature acclimation modifies Na+ current in fish cardiac myocytes
.
J. Exp. Biol.
207
,
2823
-
2833
.
Haverinen
,
J.
and
Vornanen
,
M.
(
2006
).
Significance of Na+ current in the excitability of atrial and ventricular myocardium of the fish heart
.
J. Exp. Biol.
209
,
549
-
557
.
Haverinen
,
J.
and
Vornanen
,
M.
(
2007
).
Temperature acclimation modifies sinoatrial pacemaker mechanism of the rainbow trout heart
.
Am. J. Physiol.
292
,
R1023
-
R1032
.
Haverinen
,
J.
and
Vornanen
,
M.
(
2009
).
Responses of action potential and K+ currents to temperature acclimation in fish hearts: phylogeny or thermal preferences?
Physiol. Biochem. Zool.
82
,
468
-
482
.
Haverinen
,
J.
and
Vornanen
,
M.
(
2020
)
Reduced ventricular excitability causes atrioventricular block and depression of heart rate in fish at critically high temperatures
.
J. Exp. Biol. 223, jeb225227.
Haverinen
,
J.
,
Abramochkin
,
D. V.
,
Kamkin
,
A.
and
Vornanen
,
M.
(
2017
).
The maximum heart rate in brown trout (Salmo trutta fario) is not limited by firing rate of pacemaker cells
.
Am. J. Physiol.
312
,
R165
-
R171
.
Heitler
,
W. J.
and
Edwards
,
D. H.
(
1998
).
Effect of temperature on a voltage-sensitive electrical synapse in crayfish
.
J. Exp. Biol.
201
,
503
-
513
.
Hibino
,
H.
,
Inanobe
,
A.
,
Furutani
,
K.
,
Murakami
,
S.
,
Findlay
,
I.
and
Kurachi
,
Y.
(
2010
).
Inwardly rectifying potassium channels: their structure, function and physiological role
.
Physiol. Rev.
90
,
291
-
366
.
Hodgkin
,
A. L.
(
1948
).
The local electric changes associated with repetitive action in a non-medullated axon
.
J. Physiol.
107
,
165
-
181
.
Hodgkin
,
A. L.
(
1951
).
The ionic basis of electrical activity in nerve and muscle
.
Biol. Rev.
26
,
339
-
409
.
Hodgkin
,
A. L.
and
Katz
,
B.
(
1949
).
The effect of temperature on the electrical activity of the giant axon of the squid
.
J. Physiol.
109
,
240
-
249
.
Hodgkin
,
A. L.
and
Nakajima
,
S.
(
1972
).
The effect of diameter on the electrical constants of frog skeletal muscle fibres
.
J. Physiol.
221
,
105
-
120
.
Hollingsworth
,
M. J.
and
Bowler
,
K.
(
1966
).
The decline in ability to withstand high temperature with increase in age in Drosophila subobscura
.
Exp. Gerontol.
1
,
251
-
257
.
Hutter
,
O. F.
and
Noble
,
D.
(
1960
).
The chloride conductance of frog skeletal muscle
.
J. Physiol.
151
,
89
-
102
.
Huxley
,
A. F.
(
1959
).
Ion movement during nerve activity. Part I. Physical and chemical aspects of nerve impulse conduction
.
Ann. N. Y. Acad. Sci.
81
,
221
-
246
.
Irvin
,
D. N.
(
1974
).
Temperature tolerance of early developmental stages of Dover sole, Solea solea (L.)
. In
The Early Life History of Fish
(ed.
J. H. S.
Blaxter
), pp.
449
-
463
.
Heidelberg
:
Springer
.
Irvine
,
L. A.
,
Jafri
,
M. S.
and
Winslow
,
R. L.
(
1999
).
Cardiac sodium channel Markov model with temperature dependence and recovery from inactivation
.
Biophys. J.
76
,
1868
-
1885
.
Janssen
,
R.
(
1992
).
Thermal influences on nervous system function
.
Neurosci. Behav. Rev.
16
,
399
-
413
.
Jensen
,
D. W.
(
1972
).
The effect of temperature on transmission at the neuromuscular junction of the sartorius muscle of Rana pipiens
.
Comp. Biochem. Physiol.
41
,
685
-
695
.
Johnson
,
B. R.
,
Peck
,
J. H.
and
Harris-Warrick
,
R. M.
(
1991
).
Temperature sensitivity of graded synaptic transmission in the lobster stomatogastric ganglion
.
J. Exp. Biol.
156
,
267
-
285
.
Jutfelt
,
F.
,
Norin
,
T.
,
Ern
,
R.
,
Overgaard
,
J.
,
Wang
,
T.
,
McKenzie
,
D. J.
,
Lefevre
,
S.
,
Nilsson
,
G. E.
,
Metcalfe
,
N. B.
,
Hickey
,
A. J. R.
, et al. 
(
2018
).
Oxygen- and capacity-limited thermal tolerance: blurring ecology and physiology
.
J. Exp. Biol.
221
,
jeb169615
.
Karunanithi
,
S.
,
Barclay
,
J. W.
,
Robertson
,
R. M.
,
Brown
,
I. R.
,
Atwood
,
H. L.
(
1999
).
Neuroprotection at Drosophila synapses conferred by prior heat shock
.
J. Neurosci.
19
,
4360
-
4369
.
Kelty
,
J. D.
,
Noseworthy
,
P. A.
,
Feder
,
M. E.
,
Robertson
,
R. M.
and
Ramirez
,
J.-M.
(
2002
).
Thermal preconditioning and heat-shock protein 72 preserve synaptic transmission during thermal stress
.
J. Neurosci.
22
,
RC193
.
Khodorov
,
B. I.
(
2012
).
The Problem of Excitability: Electrical Excitability and Ionic Permeability of the Nerve Membrane
.
New York and London
:
Plenum Press
.
Kivivuori
,
L. A.
and
Lagerspetz
,
K. Y. H.
(
1982
).
Temperature acclimation of axonal functions in the crayfish Astacus astacus L
.
J. Therm. Biol.
7
,
221
-
225
.
Kivivuori
,
L. A.
,
Lehti
,
S.
and
Lagerspetz
,
K. Y. H.
(
1990
).
Effect of temperature acclimation on thermal dependence and hysteresis of the resting membrane potential of the stretch receptor neuron in crayfish [Astacus astacus (L.)]
.
J. Therm. Biol.
15
,
9
-
14
.
Klee
,
M. R.
,
Pierau
,
F. K.
and
Faber
,
D. S.
(
1974
).
Temperature effects on resting potential and spike parameters of cat motoneurons
.
Exp. Brain. Res.
19
,
478
-
492
.
Klein
,
M. G.
and
Prosser
,
C. L.
(
1985
).
The effects of temperature acclimation on the resting membrane of skeletal muscle fibres from green sunfish
.
J. Exp. Biol.
114
,
563
-
579
.
Kline
,
R. P.
and
Morad
,
M.
(
1978
).
Potassium efflux in heart muscle during activity: extracellular accumulation and its implications
.
J. Physiol.
280
,
537
-
558
.
Koester
,
J.
and
Siegelbaum
,
S. A.
(
2000
).
Chapter 8: Local signaling: passive electrical properties of the neuron
. In
Principle of Neural Science
(ed.
E. R.
Kandel
,
J. H.
Schwartz
and
T. M.
Jessell
), pp.
4
-
13
.
New York
:
Elsevier
.
Kornhuber
,
M. E.
and
Walther
,
C.
(
1987
).
The electrical constants of the fibres from two leg muscles of the locust Schistocerca gregaria
.
J. Exp. Biol.
127
,
173
-
189
.
Krnjević
,
K.
and
Miledi
,
R.
(
1959
).
Presynaptic failure of neuromuscular propagation in rats
.
J. Physiol.
149
,
1
-
22
.
Lefevre
,
S.
(
2016
).
Are global warming and ocean acidification conspiring against marine ectotherms? A meta-analysis of the respiratory effects of elevated temperature, high CO2 and their interaction
.
Conserv. Physiol.
4
,
cow009
.
Li
,
C.-L.
and
Gouras
,
P.
(
1958
).
Effect of cooling on neuromuscular transmission in the frog
.
Am. J. Physiol.
192
,
464
-
470
.
Liang
,
S.
,
Yang
,
F.
,
Zhou
,
C.
,
Wang
,
Y.
,
Li
,
S.
,
Sun
,
C. K.
,
Puglisi
,
J. L.
,
Bers
,
D.
,
Sun
,
C.
and
Zheng
,
J.
(
2009
).
Temperature-dependent activation of neurons by continuous near-infrared laser
.
Cell Biochem. Biophys.
53
,
33
.
Liebeskind
,
B. J.
(
2011
).
Evolution of sodium channels and the new view of early nervous system evolution
.
Commun. Integr. Bio.
4
,
679
-
683
.
Liebeskind
,
B. J.
,
Hillis
,
D. M.
and
Zakon
,
H. H.
(
2011
).
Evolution of sodium channels predates the origin of nervous systems in animals
.
Proc. Natl. Acad. Sci. USA
108
,
9154
-
9159
.
Loughney
,
K.
,
Kreber
,
R.
and
Ganetzky
,
B.
(
1989
).
Molecular analysis of the para locus, a sodium channel gene in Drosophila
.
Cell
58
,
1143
-
1154
.
Ma
,
Y.
and
Huguenard
,
J. R.
(
2013
).
Reemerging role of cable properties in action potential initiation
.
Proc. Natl. Acad. Sci. USA
110
,
3715
-
3716
.
Macdonald
,
J. A.
and
Montgomery
,
J. C.
(
1982
).
Thermal limits of neuromuscular function in an Antarctic fish
.
J. Comp. Physiol. A
147
,
237
-
250
.
MacMillan
,
H. A.
(
2019
).
Dissecting cause from consequence: a systematic approach to thermal limits
.
J. Exp. Biol.
222
,
jeb191593
.
Maltsev
,
V. A.
and
Undrovinas
,
A. I.
(
1997
).
Cytoskeleton modulates coupling between availability and activation of cardiac sodium channel
.
Am. J. Physiol.
273
,
H1831
-
H1840
.
Mangold
,
K. E.
,
Brumback
,
B. D.
,
Angsutararux
,
P.
,
Voelker
,
T. L.
,
Zhu
,
W.
,
Kang
,
P. W.
,
Moreno
,
J. D.
and
Silva
,
J. R.
(
2017
).
Mechanisms and models of cardiac sodium channel inactivation
.
Channels
11
,
517
-
533
.
Mendonca
,
P. C.
and
Gamperl
,
A. K.
(
2010
).
The effects of acute changes in temperature and oxygen availability on cardiac performance in winter flounder (Pseudopleuronectes americanus)
.
Comp. Biochem. Physiol. A
155
,
245
-
252
.
Merideth
,
J.
and
Pruitt
,
R. D.
(
1973
).
Disturbances in cardiac conduction and their management
.
Circulation
47
,
1098
-
1107
.
Money
,
T. G. A.
,
Anstey
,
M. L.
and
Robertson
,
R. M.
(
2005
).
Heat stress-mediated plasticity in a locust looming-sensitive visual interneuron
.
J. Neurophysiol.
93
,
1908
-
1919
.
Monfredi
,
O.
,
Dobrzynski
,
H.
,
Mondal
,
T.
,
Boyett
,
M. R.
and
Morris
,
G. M.
(
2010
).
The anatomy and physiology of the sinoatrial node—a contemporary review
.
PACE Pacing. Clin. Electrophysiol.
33
,
1392
-
1406
.
Montgomery
,
J. C.
and
Macdonald
,
J. A.
(
1990
).
Effects of temperature on nervous system: implications for behavioral performance
.
Am. J. Physiol.
259
,
R191
-
R196
.
Nielsen
,
J. S.
and
Gesser
,
H.
(
2001
).
Effects of high extracellular [K+] and adrenaline on force development, relaxation and membrane potential in cardiac muscle from freshwater turtle and rainbow trout
.
J. Exp. Biol.
204
,
261
-
268
.
Nishino
,
A.
and
Okamura
,
Y.
(
2017
).
Evolutionary history of voltage-gated sodium channels
. In
Voltage-gated Sodium Channels: Structure, Function and Channelopathies
(ed.
M.
Chahine
), pp.
3
-
32
.
Springer
.
O'Dowd
,
D. K.
,
Germeraad
,
S. E.
and
Aldrich
,
R. W.
(
1989
).
Alterations in the expression and gating of Drosophila sodium channels by mutations in the para gene
.
Neuron
2
,
1301
-
1311
.
Orr
,
P. R.
(
1955a
).
Heat death. I. Time-temperature relationships in marine animals
.
Physiol. Zool.
28
,
290
-
294
.
Orr
,
P. R.
(
1955b
).
Heat death. II. Differential response of entire animal (Rana pipiens) and several organ systems
.
Physiol. Zool.
28
,
294
-
302
.
O'Sullivan
,
J. D. B.
,
MacMillan
,
H. A.
and
Overgaard
,
J.
(
2017
).
Heat stress is associated with disruption of ion balance in the migratory locust, Locusta migratoria
.
J. Therm. Biol.
68
,
177
-
185
.
Park
,
D. S.
,
Shekhar
,
A.
,
Marra
,
C.
,
Lin
,
X.
,
Vasquez
,
C.
,
Solinas
,
S.
,
Kelley
,
K.
,
Morley
,
G.
,
Goldfarb
,
M.
and
Fishman
,
G. I.
(
2016
).
Fhf2 gene deletion causes temperature-sensitive cardiac conduction failure
.
Nat. Commun.
7
,
12966
.
Parnas
,
I.
,
Hochstein
,
S.
and
Parnas
,
H.
(
1976
).
Theoretical analysis of parameters leading to frequency modulation along an inhomogeneous axon
.
J. Neurophysiol.
39
,
909
-
923
.
Peron
,
S.
,
Zordan
,
M. A.
,
Magnabosco
,
A.
,
Reggiani
,
C.
and
Megighian
,
A.
(
2009
).
From action potential to contraction: neural control and excitation–contraction coupling in larval muscles of Drosophila
.
Comp. Biochem. Physiol. A
154
,
173
-
183
.
Platt
,
J. R.
(
1964
).
Strong inference: certain systematic methods of scientific thinking may produce much more rapid progress than others
.
Science
146
,
347
-
353
.
Pörtner
,
H.-O.
(
2001
).
Climate change and temperature-dependent biogeography: oxygen limitation of thermal tolerance in animals
.
Naturwissenschaften
88
,
137
-
146
.
Pörtner
,
H.-O.
(
2002
).
Physiological basis of temperature-dependent biogeography: trade-offs in muscle design and performance in polar ectotherms
.
J. Exp. Biol.
205
,
2217
-
2230
.
Pörtner
,
H.-O.
,
Bock
,
C.
and
Mark
,
F. C.
(
2017
).
Oxygen- and capacity-limited thermal tolerance: bridging ecology and physiology
.
J. Exp. Biol.
220
,
2685
-
2696
.
Precht
,
H.
,
Christophersen
,
J.
,
Hensel
,
H.
and
Larcher
,
W.
(
1973
).
Temperature and Life
.
Berlin, Heidelberg, New York
:
Springer-Verlag
.
Prosser
,
C. L.
and
Nelson
,
D. O.
(
1981
).
The role of nervous systems in temperature adaptation of poikilotherms
.
Annu. Rev. Physiol.
43
,
281
-
300
.
Racinais
,
S.
and
Oksa
,
J.
(
2010
).
Temperature and neuromuscular function
.
Scand. J. Med. Sci. Sports
20
,
1
-
18
.
Racinais
,
S.
,
Gaoua
,
N.
and
Grantham
,
J.
(
2008
).
Hyperthermia impairs short-term memory and peripheral motor drive transmission
.
J. Physiol.
586
,
4751
-
4762
.
Ramirez
,
J. M.
,
Elsen
,
F. P.
and
Robertson
,
R. M.
(
1999
).
Long-term effects of prior heat shock on neuronal potassium currents recorded in a novel insect ganglion slice preparation
.
J. Neurophysiol.
81
,
795
-
802
.
Rasminsky
,
M.
(
1973
).
The effects of temperature on conduction in demyelinated single nerve fibers
.
Arch. Neurol.
28
,
287
-
292
.
Robertson
,
R. M.
and
Money
,
T. G. A.
(
2012
).
Temperature and neuronal circuit function: compensation, tuning and tolerance
.
Curr. Opin. Neurobiol.
22
,
724
-
734
.
Ruff
,
R. L.
(
2011
).
Endplate contributions to the safety factor for neuromuscular transmission
.
Muscle Nerve
44
,
854
-
861
.
Rutkove
,
S. B.
(
2001
).
Effects of temperature on neuromuscular electrophysiology
.
Muscle Nerve
24
,
867
-
882
.
Rutkove
,
S. B.
,
Kothari
,
M. J.
and
Shefner
,
J. M.
(
1997
).
Nerve, muscle, and neuromuscular junction electrophysiology at high temperature
.
Muscle Nerve
20
,
431
-
436
.
Schmidtmayer
,
J.
(
1989
).
Voltage and temperature dependence of normal and chemically modified inactivation of sodium channels
.
Pflüg. Arch. Eur. J. Physiol.
414
,
273
-
281
.
Schoepfle
,
G. M.
and
Erlanger
,
J.
(
1941
).
The action of temperature on the excitability, spike height and configuration, and the refractory period observed in the responses of single medullated nerve fibers
.
Am. J. Physiol.
134
,
694
-
704
.
Schwarz
,
W.
(
1979
).
Temperature experiments on nerve and muscle membranes of frogs
.
Pflüg. Arch. Eur. J. Physiol.
382
,
27
-
34
.
Shaw
,
R. M.
and
Rudy
,
Y.
(
1997
).
Ionic mechanisms of propagation in cardiac tissue. Roles of the sodium and L-type calcium currents during reduced excitability and decreased gap junction coupling
.
Circ. Res.
81
,
727
-
741
.
Siddiqi
,
O.
and
Benzer
,
S.
(
1976
).
Neurophysiological defects in temperature-sensitive paralytic mutants of Drosophila melanogaster
.
Proc. Natl. Acad. Sci. USA
73
,
3253
-
3257
.
Smith
,
D. O.
(
1980
).
Mechanisms of action potential propagation failure at sites of axon branching in the crayfish
.
J. Physiol.
301
,
243
-
259
.
Spector
,
P.
(
2013
).
Principles of cardiac electric propagation and their implications for re-entrant arrhythmias
.
Circ. Arrhythm Electrophysiol.
6
,
655
-
661
.
Spira
,
M. E.
,
Yarom
,
Y.
and
Parnas
,
I.
(
1976
).
Modulation of spike frequency by regions of special axonal geometry and by synaptic inputs
.
J. Neurophysiol.
39
,
882
-
899
.
Städele
,
C.
,
Heigele
,
S.
and
Stein
,
W.
(
2015
).
Neuromodulation to the rescue: compensation of temperature-induced breakdown of rhythmic motor patterns via extrinsic neuromodulatory input
.
PLoS Biol.
13
,
e1002265
.
Steinhausen
,
M. F.
,
Sandblom
,
E.
,
Eliason
,
E. J.
,
Verhille
,
C.
and
Farrell
,
A. P.
(
2008
).
The effect of acute temperature increases on the cardiorespiratory performance of resting and swimming sockeye salmon (Oncorhynchus nerka)
.
J. Exp. Biol.
211
,
3915
-
3926
.
Stephens
,
P. J.
(
1988
).
The effects of temperature and ethanol on the properties of the fast excitatory axon to the crab limb bender muscle
.
Comp. Biochem. Physiol. A
90
,
341
-
347
.
Stephens
,
P. J.
,
Frascella
,
P. A.
and
Mindrebo
,
N.
(
1983
).
Effects of ethanol and temperature on a crab motor axon action potential: a possible mechanism for peripheral spike generation
.
J. Exp. Biol.
103
,
289
-
301
.
Stevens
,
E. D.
,
Bennion
,
G. R.
,
Randall
,
D. J.
and
Shelton
,
G.
(
1972
).
Factors affecting arterial pressures and blood flow from the heart in intact, unrestrained lingcod, Ophiodon elongatus
.
J. Exp. Biol.
43
,
681
-
695
.
Suzuki
,
D. T.
,
Grigliatti
,
T.
and
Williamson
,
R.
(
1971
).
Temperature-sensitive mutations in Drosophila melanogaster. VII. A mutation (parats) causing reversible adult paralysis
.
Proc. Natl. Acad. Sci. USA
68
,
890
-
893
.
Swadlow
,
H. A.
,
Kocsis
,
J. D.
and
Waxman
,
S. G.
(
1980
).
Modulation of impulse conduction along the axonal tree
.
Annu. Rev. Biophys. Bioeng.
9
,
143
-
179
.
Targ
,
E. F.
and
Kocsis
,
J. D.
(
1985
).
4-Aminopyridine leads to restoration of conduction in demyelinated rat sciatic nerve
.
Brain Res.
328
,
358
-
361
.
Tillman
,
T. S.
and
Cascio
,
M.
(
2003
).
Effects of membrane lipids on ion channel structure and function
.
Cell Biochem. Biophys.
38
,
161
-
190
.
Touska
,
F.
,
Turnquist
,
B.
,
Vlachova
,
V.
,
Reeh
,
P. W.
,
Leffler
,
A.
and
Zimmermann
,
K.
(
2018
).
Heat-resistant action potentials require TTX-resistant sodium channels NaV1. 8 and NaV1. 9
.
J. Gen. Physiol.
150
,
1125
-
1144
.
Ulyanova
,
A. V.
and
Shirokov
,
R. E.
(
2018
).
Voltage-dependent inward currents in smooth muscle cells of skeletal muscle arterioles
.
PLoS ONE
13
,
e0194980
.
Varghese
,
A.
(
2016
).
Reciprocal modulation of IK1 and INa extends excitability in cardiac ventricular cells
.
Front. Physiol.
7
,
542
.
Volgushev
,
M.
,
Vidyasagar
,
T. R.
,
Chistiakova
,
M.
,
Yousef
,
T.
and
Eysel
,
U. T.
(
2000
).
Membrane properties and spike generation in rat visual cortical cells during reversible cooling
.
J. Physiol.
522
,
59
-
76
.
Vornanen
,
M.
(
2016
).
The temperature-dependence of electrical excitability of fish heart
.
J. Exp. Biol.
219
,
1941
-
1952
.
Vornanen
,
M.
,
Ryökkynen
,
A.
and
Nurmi
,
A.
(
2002
).
Temperature-dependent expression of sarcolemmal K+ currents in rainbow trout atrial and ventricular myocytes
.
Am. J. Physiol.
282
,
R1191
-
R1199
.
Vornanen
,
M.
,
Hassinen
,
M.
,
Koskinen
,
H.
and
Krasnov
,
A.
(
2005
).
Steady-state effects of temperature acclimation on the transcriptome of the rainbow trout heart
.
Am. J. Physiol.
289
,
R1177
-
R1184
.
Vornanen
,
M.
,
Haverinen
,
J.
and
Egginton
,
S.
(
2014
).
Acute heat tolerance of cardiac excitation in the brown trout (Salmo trutta fario)
.
J. Exp. Biol.
217
,
299
-
309
.
Wang
,
Y.
and
Rudy
,
Y.
(
2000
).
Action potential propagation in inhomogeneous cardiac tissue: safety factor considerations and ionic mechanism
.
Am. J. Physiol.
278
,
H1019
-
H1029
.
Westerfield
,
M.
,
Joyner
,
R. W.
and
Moore
,
J. W.
(
1978
).
Temperature-sensitive conduction failure at axon branch points
.
J. Neurophysiol.
41
,
1
-
8
.
White
,
R. L.
(
1983
).
Effects of acute temperature change and acclimation temperature on neuromuscular function and lethality in crayfish
.
Physiol. Zool.
56
,
174
-
194
.
Wood
,
S. J.
and
Slater
,
C. R.
(
2001
).
Safety factor at the neuromuscular junction
.
Prog. Neurobiol.
64
,
393
-
429
.
Wu
,
C. F.
,
Ganetzky
,
B.
,
Jan
,
L. Y.
,
Jan
,
Y. N.
and
Benzer
,
S.
(
1978
).
A Drosophila mutant with a temperature-sensitive block in nerve conduction
.
Proc. Natl. Acad. Sci. USA
75
,
4047
-
4051
.
Yamauchi
,
A.
and
Burnstock
,
G.
(
1968
).
An electronmicroscopic study on the innervation of the trout heart
.
J. Comp. Neurol.
132
,
567
-
588
.
Yue
,
D. T.
,
Lawrence
,
J. H.
and
Marban
,
E.
(
1989
).
Two molecular transitions influence cardiac sodium channel gating
.
Science
244
,
349
-
352
.

Competing interests

The authors declare no competing or financial interests.