Skip to main content
F1000Research logoLink to F1000Research
. 2016 May 12;5:F1000 Faculty Rev-880. [Version 1] doi: 10.12688/f1000research.7958.1

Central control of body temperature

Shaun F Morrison 1,a
PMCID: PMC4870994  PMID: 27239289

Abstract

Central neural circuits orchestrate the behavioral and autonomic repertoire that maintains body temperature during environmental temperature challenges and alters body temperature during the inflammatory response and behavioral states and in response to declining energy homeostasis. This review summarizes the central nervous system circuit mechanisms controlling the principal thermoeffectors for body temperature regulation: cutaneous vasoconstriction regulating heat loss and shivering and brown adipose tissue for thermogenesis. The activation of these thermoeffectors is regulated by parallel but distinct efferent pathways within the central nervous system that share a common peripheral thermal sensory input. The model for the neural circuit mechanism underlying central thermoregulatory control provides a useful platform for further understanding of the functional organization of central thermoregulation, for elucidating the hypothalamic circuitry and neurotransmitters involved in body temperature regulation, and for the discovery of novel therapeutic approaches to modulating body temperature and energy homeostasis.

Keywords: Brown adipose tissue, shiver, cutaneous vasoconstriction, thermogenesis, fever, sympathetic nerve activity, preoptic hypothalamus, rostral raphe pallidus, dorsomedial hypothalamus, therapeutic hypothermia, obesity


Body temperature (T core) is a critical homeostatic parameter influencing cellular function and organismal survival. Life-threatening protein denaturation looms as T core increases, and reductions in membrane fluidity, ion fluxes, and enzyme performance accompany significant reductions in T core. The fundamental central neural circuits for thermoregulation orchestrate behavioral and autonomic repertoires that maintain T core during thermal challenges, sensed by thermal receptors, that arise from either the ambient or the internal (e.g., during exercise) environment. A variety of other neural circuits and neurochemical modulators impinge on the fundamental thermoregulatory pathways to produce the alterations in T core that occur with circadian and ultradian periodicities 1, 2 and those that accompany many behavioral states, such as fever during sickness, increased T core during psychological stress and at ovulation, and reductions in T core during sleep, sepsis, or exposure to metabolic distress (e.g., hypoxia and starvation). Recent research in the central nervous system (CNS) control of T core focuses principally on: (a) elaborating the CNS pathways and neurotransmitter systems involved in the fundamental central thermoregulatory network; (b) the modulation of activity within the fundamental central thermoregulatory network by non-thermal factors and behavioral states; and (c) pharmacological manipulation of the CNS thermoregulatory network for a variety of therapeutic goals (e.g., reducing T core to produce therapeutic hypothermia in stroke victims).

In its simplest form, the fundamental thermoregulatory network can be modeled as a reflex 3, 4 in which a central integrative circuit alters the activity of thermoeffector mechanisms in response to an input from the combination of peripheral (i.e., skin) and central (i.e., visceral and brain) thermoreceptors that provides a consolidated assessment of T core and, importantly, of imminent threats to T core. The primary thermoeffector mechanisms recruited for both cold defense and centrally driven hyperthermias (e.g., fever) include: (a) thermoregulatory behaviors to reduce the loss of heat produced during basal metabolism; (b) cutaneous vasoconstriction (CVC) to conserve heat in the body core and limit heat loss to the environment; and (c) heat production (thermogenesis). The principal sources of metabolic heat production, beyond those contributing to basal metabolic rate (e.g., pumping ions across membranes), are brown adipose tissue (BAT), whose sympathetic neural input fuels mitochondria that shunt proton fluxes into heat production, and shivering behavior in skeletal muscle, dependent on the inefficiency of ATP utilization to generate heat. Effector mechanisms for heat defense include: (a) thermoregulatory behavior to increase heat loss; (b) cutaneous vasodilation, which in humans includes a sympathetic vasodilator outflow 5, 6, combined with increased cardiac output 7 and visceral vasoconstriction 810 to facilitate core heat loss from the body surface; and (c) evaporative cooling (e.g., sweating).

Apart from important “first responders”, the CNS pathways controlling and mediating thermoregulatory behaviors remain incompletely defined 1115. Along these same lines, most research subjects are furry mammals that are strongly dependent on non-sweating mechanisms for evaporative cooling. Although this has limited the available detail on the CNS pathways regulating human sweating 6, 16, 17, considerable insight has been gained on the CNS regulation of other mechanisms of evaporative cooling, such as panting 1820. Additionally, since most of the basic neuroscience of CNS thermoregulatory pathways has been derived from experiments in rodents, including the exclusively in vitro studies of neurons with intrinsic thermosensitivity, the translation of the conclusions, including the circuit model in Figure 1, to humans must be cautiously undertaken (e.g., 21). Finally, although considerable progress has been achieved in the relatively young field of the neuroscience of thermoregulation, the synthesis ( Figure 1) of our understanding of this multifaceted neural network controlling multiple thermoeffectors represents a working model, with the expectation of revisions and added detail.

Figure 1. Functional neuroanatomical model for the fundamental pathways providing the thermoregulatory control and pyrogenic activation of cutaneous vasoconstriction (CVC) and brown adipose tissue (BAT) and shivering thermogenesis.

Figure 1.

Cool and warm cutaneous thermoreceptors transmit signals to respective primary sensory neurons in the dorsal root ganglia (DRG) which relay this information to second-order thermal sensory neurons in the dorsal horn (DH). Cool sensory DH neurons glutamatergically activate third-order sensory neurons in the external lateral subnucleus of the lateral parabrachial nucleus (LPB), while warm sensory DH neurons project to third-order sensory neurons in the dorsal subnucleus of the LPB. Thermosensory signals driving thermoregulatory responses are transmitted from the LPB to the preoptic area (POA), which contains the microcircuitry through which cutaneous and core thermal signals are integrated to regulate the balance of POA outputs that are excitatory (dashed green) and inhibitory (dashed red) to thermogenesis-promoting neurons in the dorsomedial hypothalamus (DMH) and to CVC sympathetic premotor neurons in the rostral raphe pallidus (rRPa). Within the POA, GABAergic interneurons (red) in the median preoptic (MnPO) subnucleus are postulated to receive a glutamatergic input from skin cooling-activated neurons in LPB and inhibit each of the distinct populations of warm-sensitive (W-S) neurons in the medial preoptic area (MPA) that control CVC, BAT, and shivering. In contrast, glutamatergic interneurons (dark green) in the MnPO are postulated to be excited by glutamatergic inputs from skin warming-activated neurons in LPB and, in turn, excite the populations of W-S neurons in MPA. Prostaglandin E 2 (PGE 2) binds to EP3 receptors, which are postulated to inhibit the activity of each of the classes of W-S neurons in the POA. Preoptic W-S neurons may provide inhibitory control of CVC by inhibiting CVC sympathetic premotor neurons in the rostral ventromedial medulla, including the rRPa, that project to CVC sympathetic preganglionic neurons (SPNs) in the intermediolateral nucleus (IML). Preoptic W-S neurons may provide inhibitory thermoregulatory control of BAT and shivering thermogenesis by inhibiting BAT sympathoexcitatory neurons and shivering-promoting neurons, respectively, in the DMH, which, when disinhibited during skin and core cooling, provide respective excitatory drives to BAT sympathetic premotor neurons and to skeletal muscle shivering premotor neurons in the rRPa. These, in turn, project, respectively, to BAT SPNs in the IML and to alpha (α) and gamma (γ) motoneurons in the ventral horn (VH) of the spinal cord.

The CNS thermoregulatory control of the sympathetic outflows mediating CVC and BAT thermogenesis and of the somatic motoneurons producing shivering is effected through parallel but distinct, effector-specific, integrative/efferent circuits ( Figure 1, and reviewed in 2225) that share common peripheral thermal sensory inputs. The hypothalamus contains the primary integrative and rostral efferent components of these circuits. Although many details of the preoptic area (POA) microcircuitry for thermoregulation remain to be elucidated, neurons in the POA are postulated to integrate ascending peripheral thermosensory signals with local thermosensitivity to regulate the output of BAT and shivering thermogenesis-promoting neurons in the dorsomedial hypothalamus (DMH) 26, 27 and of CVC-promoting neurons in the median preoptic nucleus (MnPO) 28, 29.

The POA regulation of DMH thermogenesis-promoting neurons represents the balance between a GABAergic inhibition 30, 31 and a glutamatergic excitation 32; the latter inputs, potentially arising from neurons in the MnPO that project to the DMH, are synaptically connected to BAT 33 and express the leptin receptor 34. These glutamatergic inputs to DMH 32 could provide the excitation required to drive the BAT sympathoexcitatory neurons and the shivering-promoting neurons in DMH when their POA inhibitory input is reduced during skin cooling or fever 35. Although intrinsically warm-sensitive (W-S) ( Figure 1), POA neurons 3638, generally located in the medial preoptic area (MPA) 39, are postulated to play a key role in central thermoregulation by providing a prominent core temperature-modulated, GABAergic 38 regulation of thermogenesis-promoting neurons in DMH ( Figure 1), the considerable direct functional evidence required to establish this attractive hypothesis has yet to be obtained. Different thermal sensitivities or neurochemical modulation among populations of temperature-sensitive POA neurons may underlie the differential responsiveness of different effectors to changes in cutaneous versus brain temperatures 40 as well as the significant alterations in thermoeffector activation during different sleep phases 41. Through their responses to immune signaling molecules, neurons in the POA are also the primary site for the organization and maintenance of the febrile response to inflammation and infection, which includes the stimulation of CVC, and shivering (“chills”) and BAT thermogenesis mediated by the action of prostaglandin E 2 (PGE 2) on its EP3 receptors 4245. Similarly, the fundamental thermoregulatory network mediates stress-induced hyperthermia 4648. Unraveling the complexity of the thermoregulatory circuitry in the hypothalamus 20, 28, 29, 4952, including the phenotypic characterization of the projection neurons 34 and their synaptic interactions that mediate the circadian 13 and many behavioral 53, 54 modulations in T core, continues to pose significant research challenges for understanding the “heart” of the CNS thermoregulatory network. The downstream targets of the hypothalamic projection neurons for thermoregulation are the sympathetic and somatic premotor neurons in the rostral ventromedial medulla, centered on the rostral raphe pallidus (rRPa). Midbrain (e.g., periaqueductal gray 5557 and retrorubral field 58) and medullary (e.g., ventrolateral medulla and nucleus of the solitary tract (NTS) 59, 60) pathways to these premotor neurons exert important modulatory influences on the thermoregulatory activation of thermoeffector organs (reviewed in 22, 24). These premotor neurons, in turn, excite CVC and BAT sympathetic preganglionic neurons and α-motoneurons and γ-motoneurons 61, 62 in the spinal cord.

Cold and warm thermoreceptors in the skin and viscera provide the extracranial thermal signals relating to skin temperature and T core. These are integrated with the brain temperature information potentially derived from the discharge of W-S, GABAergic preoptic neurons of the central thermoregulatory network to regulate thermoeffector activities. The membranes of thermoreceptor afferent neurons contain transient receptor potential (TRP) cation channels whose temperature-dependent conductances transduce skin temperature into primary thermoreceptor afferent neuronal activity. The TRPM8 channel, activated by menthol and cooling, is the strongest candidate for the cutaneous cold receptor (reviewed in 63). The identity of the peripheral warm receptor remains to be established, but the warm-sensing mechanism of preoptic neurons, though still debated 6466, is unlikely to involve a transient receptor potential (TRP) channel 67. Primary thermoreceptor dorsal root ganglion neurons synapse on thermoreceptive-specific, lamina I spinal (or trigeminal) dorsal horn cells 68, and these, in turn, collateralize to innervate the thalamus, providing the neural substrate for cutaneous thermosensory perception and localization 68, 69, and the pontine lateral parabrachial nucleus (LPB) 70, 71, which is responsible for triggering involuntary (e.g., autonomic and shivering) thermoregulatory responses. Spinal lamina I skin cooling-responsive neurons provide a glutamatergic excitation to neurons in the external lateral subdivision of the lateral parabrachial nucleus (LPBel), which, in turn, project principally to the MnPO of the POA 72, 73, while a parallel, spinoparabrachial glutamatergic pathway excites POA-projecting neurons in the dorsal subnucleus of the LPB (LPBd) 72, 74 in response to skin warming 74. Thus, activations of POA-projecting LPBd and LPBel neurons, driven respectively by cutaneous, and likely visceral, warm and cold thermoreceptor stimuli, initiate heat defense and cold defense inhibitions and excitations, respectively, in CVC sympathetic outflow and cutaneous blood flow, in BAT sympathetic outflow and BAT thermogenesis, and in shivering EMGs and shivering thermogenesis to maintain a homeostatic T core.

Synaptic integration sites throughout the core thermoregulatory network provide the substrate for a wide variety of non-thermal physiological parameters, disease processes, neuromodulators, and drugs to influence the central regulation of T core (reviewed in 25). For example, the high metabolic rate of BAT and shivering skeletal muscles during thermogenesis cannot be sustained without a dependable supply of metabolic fuels, particularly oxygen 75, lipolytic by-products 34, and glucose 76. Thus, the CNS networks driving cold-defensive and behavioral BAT activation or shivering are strongly inhibited by signals reflecting a reduction in the short- and long-term availability of these fuel molecules essential for BAT and skeletal muscle metabolism. Similarly, as hypovolemia progresses during dehydration, the ensuing hyperosmolarity reduces the thermoregulatory drive for sweating 20, 77, 78, which serves to prevent cardiovascular collapse 79. Some viscerosensory afferents with axons in the vagus nerve and synapsing on second-order neurons in the NTS can also influence BAT activity 25, 80 and shivering responses and thus are expected to influence the regulation of T core. For instance, vagal afferents convey the “metabolic” signals that produce the inhibition of BAT activity induced by upregulation of hepatic glucokinase 81 and the BAT activation following either intragastric delivery of the TRP agonist, capsiate 82, or the presence of lipids in the duodenum 83. Another influence on T core includes a prominent hypothermia during motion sickness and nausea 84, for which the pathways relating vestibular stimulation to inhibition of CVC and thermogenesis remain to be elaborated.

Providing an additional layer of complexity in the CNS regulation of T core is the presence of the thermally responsive TRPV1 channel in the membranes of a variety of classically non-thermal, unmyelinated afferents 8588 that may have access to central thermoregulatory circuits, including via NTS neurons that can inhibit thermogenesis 59. The finding that the TRPV1 responsiveness to local brain temperature alters spontaneous glutamate release from the terminals of unmyelinated vagal afferents 88, 89 provides a potential substrate for T core to modulate the activity of second-order sensory neurons in NTS that modulate thermoeffector activation. Importantly, TRPV1 channels are also activated by non-thermal factors, including low (or high) pH, inorganic cations, or endovanilloids, thereby providing a basis for such factors to influence thermoeffector activation and thus T core. For instance, TRPV1 channels, potentially on the terminals of afferents in the peritoneum, are stimulated by an endogenous ligand to tonically inhibit BAT thermogenesis, which results in a hyperthermic response to TRPV1 antagonist administration 85, 90. Although these afferents were not directly tested for their thermal responsiveness, the hyperthermic response to TRPV1 antagonism was not altered by changes in T core. The relative influences and the interactions of thermal and non-thermal stimuli on the conductance of the relevant TRPV1 channels could play a role in determining their effect on the level of thermoeffector activation to thermoreceptor stimuli.

Interest in pharmacological modulation of the central thermoregulatory network has focused: (a) on reducing CVC and thermogenesis to lower T core; and (b) on augmenting thermogenesis to elevate energy expenditure with the goal of weight loss through consumption of the high-energy lipid stores in white adipose tissue. Novel approaches to reducing T core would have immediate benefits in treating intractable fevers that are unresponsive to cyclooxygenase inhibitors. Therapeutic hypothermia can have beneficial effects on survival and on reducing brain and tissue damage in ischemic insults such as cardiac arrest, stroke, and neonatal encephalopathy 9194. Extended space travel (e.g., Mars One) also may require pharmacological induction of a hypothermic, hibernation-like state to reduce energy consumption and psychological stress.

Under basal metabolic and movement conditions, changes in T core must arise from changes in the level of activation of thermoeffector tissues. Although modulating neuronal discharge at any site within the central thermoregulatory network would be expected to alter T core, centrally generated hyperthermias, such as fever, that arise from altered activity in hypothalamic thermoregulatory neurons could be most effectively reduced by manipulating thermoeffector efferent pathways 95, 96. Indeed, directly inhibiting the discharge of neurons in the rRPa area, including the functionally significant premotor neurons that control the principal thermoeffectors, produced a fall in brain temperature of approximately 14°C in an ambient environment of 15°C 97, and stimulation of α2 adrenergic receptors in the rRPa could completely block or prevent a lipopolysaccharide-evoked fever 95. Similarly, central administration of an adenosine A1 receptor (A1AR) agonist reduced rat T core by 10°C in an ambient temperature of 15°C 60 and mouse T core by 5°C in an ambient temperature of 4°C 98. Of particular interest, in rats, this hypothermia, which was produced by a blockade of the cold-evoked activation of BAT and shivering thermogenesis, was long-lasting and paralleled by marked reductions in heart and respiratory rate, in EEG, and in behavior 60—all reminiscent of those occurring in hibernation and torpor, which require central stimulation of the A1AR 99101. The discovery of a role for TRP channels in thermal sensation and in thermoeffector activation has stimulated research into the pharmacological modulation of TRP channels in the central thermoregulatory network to abrogate the normal cold defense mechanisms and allow T core to fall in a cool ambient environment. By reducing the activation of cooling-responsive skin thermoreceptors, a TRPM8 antagonist reduced rat T core by approximately 1°C when the rats were in an ambient temperature of 19°C 63. Stimulating TRPV1 reduced mouse T core by approximately 12°C when mice were exposed to a 10°C ambient temperature 102, and this effect was potentiated by the addition of a TRPM8 antagonist 103. The location of the relevant, hypothermic TRPV1 channels remains unknown.

Not only is BAT a thermogenic thermoeffector, including in adult humans 104107, but through its consumption of lipid and glucose energy stores and oxygen, thermogenic metabolism in BAT is a neurally regulated contributor to energy homeostasis. Thus, particularly in the face of an elevated consumption of energy-rich food (e.g., a high-fat diet), a chronic reduction in cooling-evoked BAT thermogenesis would contribute to the augmented adipose energy stores that characterize obesity. Indeed, mice without BAT exhibit a propensity for obesity and diabetes 108, 109; conversely, overexpression of uncoupling protein-1 (UCP-1), principally responsible for thermogenesis in BAT, mitigates obesity induced by a high-fat diet 110. Several anti-obesity therapies currently being explored are based on increased activation of BAT thermogenesis, through either activation of the central thermoregulatory network to increase the sympathetic outflow to BAT 111 or an alteration in the cellular biochemical pathways in brown adipocytes or a hyperplasia of BAT to augment thermogenesis. The consistent findings that obese humans have significantly reduced cooling-activated BAT 104106, 112, and that the basal (i.e., principally cooling-evoked) sympathetic activation of BAT is reduced in rats fed a high-fat diet 113, 114, and that a vagal afferent input to the NTS mediates the reduced cooling-evoked BAT activity in rats fed a high-fat diet 115 not only support a role for reduced BAT activity in the excess adipose accumulation of obesity but also highlight the significance of non-thermal inputs to the central thermoregulatory network 25, 81, 83 in influencing even the most basic thermoregulatory responses.

Considerable progress has been achieved in revealing the functional organization of the dedicated thermoregulatory network within the CNS that provides the fundamental neural control of the thermoregulatory effectors: thermoregulatory behavior, CVC, and BAT and shivering thermogenesis, although many of the details of the neurophysiology and neuroanatomy of the central thermoregulatory network remain active areas of investigation. The changes in T core that accompany a wide range of behaviors and in response to many hormones and drugs arise through altered non-thermal inputs to, or neurochemical modulation of, the neural activity within the fundamental thermoregulatory network. The latter, as well as thermoreceptor-based strategies, are being researched as therapeutic approaches in which the central thermoregulatory networks are recruited to alter T core and metabolism.

Editorial Note on the Review Process

F1000 Faculty Reviews are commissioned from members of the prestigious F1000 Faculty and are edited as a service to readers. In order to make these reviews as comprehensive and accessible as possible, the referees provide input before publication and only the final, revised version is published. The referees who approved the final version are listed with their names and affiliations but without their reports on earlier versions (any comments will already have been addressed in the published version).

The referees who approved this article are:

  • Andrej A Romanovsky, Systemic Inflammation Laboratory (FeverLab), Trauma Research, St. Joseph’s Hospital and Medical Center, Phoenix, AZ, USA

  • Matteo Cerri, Department of Biomedical and NeuroMotor Sciences, University of Bologna, Bologna, Italy

  • Robin M McAllen, The Florey Institute of Neuroscience and Mental Health, University of Melbourne, Parkville, Victoria, Australia; Department of Anatomy and Neuroscience, University of Melbourne, Parkville, Victoria, Australia

  • Michael J McKinley, The Florey Institute of Neuroscience and Mental Health, University of Melbourne, Parkville, Victoria, Australia; Department of Physiology, University of Melbourne, Parkville, Victoria, Australia

Funding Statement

This work was supported by National Institutes of Health grants R01NS040987 and R01NS091066 to Shaun F. Morrison.

The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

[version 1; referees: 3 approved]

References

  • 1. Ootsuka Y, de Menezes RC, Zaretsky DV, et al. : Brown adipose tissue thermogenesis heats brain and body as part of the brain-coordinated ultradian basic rest-activity cycle. Neuroscience. 2009;164(2):849–61. 10.1016/j.neuroscience.2009.08.013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Vujovic N, Gooley JJ, Jhou TC, et al. : Projections from the subparaventricular zone define four channels of output from the circadian timing system. J Comp Neurol. 2015;523(18):2714–37. 10.1002/cne.23812 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Romanovsky AA: Do fever and anapyrexia exist? Analysis of set point-based definitions. Am J Physiol Regul Integr Comp Physiol. 2004;287(4):R992–5. 10.1152/ajpregu.00068.2004 [DOI] [PubMed] [Google Scholar]
  • 4. Werner J: System properties, feedback control and effector coordination of human temperature regulation. Eur J Appl Physiol. 2010;109(1):13–25. 10.1007/s00421-009-1216-1 [DOI] [PubMed] [Google Scholar]
  • 5. Wallin BG, Charkoudian N: Sympathetic neural control of integrated cardiovascular function: insights from measurement of human sympathetic nerve activity. Muscle Nerve. 2007;36(5):595–614. 10.1002/mus.20831 [DOI] [PubMed] [Google Scholar]
  • 6. Smith CJ, Johnson JM: Responses to hyperthermia. Optimizing heat dissipation by convection and evaporation: Neural control of skin blood flow and sweating in humans. Auton Neurosci. 2016;196:25–36. 10.1016/j.autneu.2016.01.002 [DOI] [PubMed] [Google Scholar]
  • 7. Crandall CG: Heat stress and baroreflex regulation of blood pressure. Med Sci Sports Exerc. 2008;40(12):2063–70. 10.1249/MSS.0b013e318180bc98 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Escourrou P, Freund PR, Rowell LB, et al. : Splanchnic vasoconstriction in heat-stressed men: role of renin-angiotensin system. J Appl Physiol Respir Environ Exerc Physiol. 1982;52(6):1438–43. [DOI] [PubMed] [Google Scholar]
  • 9. Kregel KC, Wall PT, Gisolfi CV: Peripheral vascular responses to hyperthermia in the rat. J Appl Physiol (1985). 1988;64(6):2582–8. [DOI] [PubMed] [Google Scholar]
  • 10. Minson CT, Wladkowski SL, Pawelczyk JA, et al. : Age, splanchnic vasoconstriction, and heat stress during tilting. Am J Physiol. 1999;276(1 Pt 2):R203–12. [DOI] [PubMed] [Google Scholar]
  • 11. Satinoff E, Rutstein J: Behavioral thermoregulation in rats with anterior hypothalamic lesions. J Comp Physiol Psychol. 1970;71(1):77–82. 10.1037/h0028959 [DOI] [PubMed] [Google Scholar]
  • 12. Nagashima K, Nakai S, Tanaka M, et al. : Neuronal circuitries involved in thermoregulation. Auton Neurosci. 2000;85(1–3):18–25. 10.1016/S1566-0702(00)00216-2 [DOI] [PubMed] [Google Scholar]
  • 13. Almeida MC, Steiner AA, Branco LG, et al. : Cold-seeking behavior as a thermoregulatory strategy in systemic inflammation. Eur J Neurosci. 2006;23(12):3359–67. 10.1111/j.1460-9568.2006.04854.x [DOI] [PubMed] [Google Scholar]
  • 14. Almeida MC, Steiner AA, Branco LG, et al. : Neural substrate of cold-seeking behavior in endotoxin shock. PLoS One. 2006;1(1):e1. 10.1371/journal.pone.0000001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Terrien J, Perret M, Aujard F: Behavioral thermoregulation in mammals: a review. Front Biosci (Landmark Ed). 2011;16:1428–44. 10.2741/3797 [DOI] [PubMed] [Google Scholar]
  • 16. Shibasaki M, Crandall CG: Mechanisms and controllers of eccrine sweating in humans. Front Biosci (Schol Ed). 2010;2:685–96. 10.2741/s94 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Farrell MJ, Trevaks D, Taylor NA, et al. : Regional brain responses associated with thermogenic and psychogenic sweating events in humans. J Neurophysiol. 2015;114(5):2578–87. 10.1152/jn.00601.2015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. McKinley MJ, McAllen RM, Whyte D, et al. : Central osmoregulatory influences on thermoregulation. Clin Exp Pharmacol Physiol. 2008;35(5–6):701–5. 10.1111/j.1440-1681.2007.04833.x [DOI] [PubMed] [Google Scholar]
  • 19. Poon CS: Optimal interaction of respiratory and thermal regulation at rest and during exercise: role of a serotonin-gated spinoparabrachial thermoafferent pathway. Respir Physiol Neurobiol. 2009;169(3):234–42. 10.1016/j.resp.2009.09.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. McKinley MJ, Yao ST, Uschakov A, et al. : The median preoptic nucleus: front and centre for the regulation of body fluid, sodium, temperature, sleep and cardiovascular homeostasis. Acta Physiol (Oxf). 2015;214(1):8–32. 10.1111/apha.12487 [DOI] [PubMed] [Google Scholar]
  • 21. McAllen RM, Farrell M, Johnson JM, et al. : Human medullary responses to cooling and rewarming the skin: a functional MRI study. Proc Natl Acad Sci U S A. 2006;103(3):809–13. 10.1073/pnas.0509862103 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Morrison SF, Nakamura K: Central neural pathways for thermoregulation. Front Biosci (Landmark Ed). 2011;16:74–104. 10.2741/3677 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Nakamura K: Central circuitries for body temperature regulation and fever. Am J Physiol Regul Integr Comp Physiol. 2011;301(5):R1207–28. 10.1152/ajpregu.00109.2011 [DOI] [PubMed] [Google Scholar]
  • 24. Morrison SF, Madden CJ: Central nervous system regulation of brown adipose tissue. Compr Physiol. 2014;4(4):1677–713. 10.1002/cphy.c140013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Morrison SF, Madden CJ, Tupone D: Central neural regulation of brown adipose tissue thermogenesis and energy expenditure. Cell Metab. 2014;19(5):741–56. 10.1016/j.cmet.2014.02.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Nakamura K, Morrison SF: Central efferent pathways mediating skin cooling-evoked sympathetic thermogenesis in brown adipose tissue. Am J Physiol Regul Integr Comp Physiol. 2007;292(1):R127–36. 10.1152/ajpregu.00427.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Nakamura K, Morrison SF: Central efferent pathways for cold-defensive and febrile shivering. J Physiol. 2011;589(Pt 14):3641–58. 10.1113/jphysiol.2011.210047 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Tanaka M, McKinley MJ, McAllen RM: Preoptic-raphé connections for thermoregulatory vasomotor control. J Neurosci. 2011;31(13):5078–88. 10.1523/JNEUROSCI.6433-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Tanaka M, McKinley MJ, McAllen RM: Role of an excitatory preoptic-raphé pathway in febrile vasoconstriction of the rat's tail. Am J Physiol Regul Integr Comp Physiol. 2013;305(12):R1479–89. 10.1152/ajpregu.00401.2013 [DOI] [PubMed] [Google Scholar]
  • 30. Cao WH, Fan W, Morrison SF: Medullary pathways mediating specific sympathetic responses to activation of dorsomedial hypothalamus. Neuroscience. 2004;126(1):229–40. 10.1016/j.neuroscience.2004.03.013 [DOI] [PubMed] [Google Scholar]
  • 31. Nakamura Y, Nakamura K, Matsumura K, et al. : Direct pyrogenic input from prostaglandin EP3 receptor-expressing preoptic neurons to the dorsomedial hypothalamus. Eur J Neurosci. 2005;22(12):3137–46. 10.1111/j.1460-9568.2005.04515.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Madden CJ, Morrison SF: Excitatory amino acid receptors in the dorsomedial hypothalamus mediate prostaglandin-evoked thermogenesis in brown adipose tissue. Am J Physiol Regul Integr Comp Physiol. 2004;286(2):R320–5. 10.1152/ajpregu.00515.2003 [DOI] [PubMed] [Google Scholar]
  • 33. Dimitrov EL, Kim YY, Usdin TB: Regulation of hypothalamic signaling by tuberoinfundibular peptide of 39 residues is critical for the response to cold: a novel peptidergic mechanism of thermoregulation. J Neurosci. 2011;31(49):18166–79. 10.1523/JNEUROSCI.2619-11.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Zhang Y, Kerman IA, Laque A, et al. : Leptin-receptor-expressing neurons in the dorsomedial hypothalamus and median preoptic area regulate sympathetic brown adipose tissue circuits. J Neurosci. 2011;31(5):1873–84. 10.1523/JNEUROSCI.3223-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Zhang YH, Yanase-Fujiwara M, Hosono T, et al. : Warm and cold signals from the preoptic area: which contribute more to the control of shivering in rats? J Physiol. 1995;485(Pt 1):195–202. 10.1113/jphysiol.1995.sp020723 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Boulant JA, Dean JB: Temperature receptors in the central nervous system. Annu Rev Physiol. 1986;48:639–54. 10.1146/annurev.ph.48.030186.003231 [DOI] [PubMed] [Google Scholar]
  • 37. Griffin JD, Kaple ML, Chow AR, et al. : Cellular mechanisms for neuronal thermosensitivity in the rat hypothalamus. J Physiol. 1996;492(Pt 1):231–42. 10.1113/jphysiol.1996.sp021304 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Lundius EG, Sanchez-Alavez M, Ghochani Y, et al. : Histamine influences body temperature by acting at H1 and H3 receptors on distinct populations of preoptic neurons. J Neurosci. 2010;30(12):4369–81. 10.1523/JNEUROSCI.0378-10.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Griffin JD, Saper CB, Boulant JA: Synaptic and morphological characteristics of temperature-sensitive and -insensitive rat hypothalamic neurones. J Physiol. 2001;537(Pt 2):521–35. 10.1111/j.1469-7793.2001.00521.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Ootsuka Y, McAllen RM: Comparison between two rat sympathetic pathways activated in cold defense. Am J Physiol Regul Integr Comp Physiol. 2006;291(3):R589–95. 10.1152/ajpregu.00850.2005 [DOI] [PubMed] [Google Scholar]
  • 41. Glotzbach SF, Heller HC: Central nervous regulation of body temperature during sleep. Science. 1976;194(4264):537–9. 10.1126/science.973138 [DOI] [PubMed] [Google Scholar]
  • 42. Scammell TE, Elmquist JK, Griffin JD, et al. : Ventromedial preoptic prostaglandin E2 activates fever-producing autonomic pathways. J Neurosci. 1996;16(19):6246–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Nakamura K, Kaneko T, Yamashita Y, et al. : Immunohistochemical localization of prostaglandin EP3 receptor in the rat nervous system. J Comp Neurol. 2000;421(4):543–69. [DOI] [PubMed] [Google Scholar]
  • 44. Nakamura K, Matsumura K, Kaneko T, et al. : The rostral raphe pallidus nucleus mediates pyrogenic transmission from the preoptic area. J Neurosci. 2002;22(11):4600–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Lazarus M, Yoshida K, Coppari R, et al. : EP3 prostaglandin receptors in the median preoptic nucleus are critical for fever responses. Nat Neurosci. 2007;10(9):1131–3. 10.1038/nn1949 [DOI] [PubMed] [Google Scholar]
  • 46. Lkhagvasuren B, Nakamura Y, Oka T, et al. : Social defeat stress induces hyperthermia through activation of thermoregulatory sympathetic premotor neurons in the medullary raphe region. Eur J Neurosci. 2011;34(9):1442–52. 10.1111/j.1460-9568.2011.07863.x [DOI] [PubMed] [Google Scholar]
  • 47. Kataoka N, Hioki H, Kaneko T, et al. : Psychological stress activates a dorsomedial hypothalamus-medullary raphe circuit driving brown adipose tissue thermogenesis and hyperthermia. Cell Metab. 2014;20(2):346–58. 10.1016/j.cmet.2014.05.018 [DOI] [PubMed] [Google Scholar]
  • 48. Mohammed M, Ootsuka Y, Blessing W: Brown adipose tissue thermogenesis contributes to emotional hyperthermia in a resident rat suddenly confronted with an intruder rat. Am J Physiol Regul Integr Comp Physiol. 2014;306(6):R394–400. 10.1152/ajpregu.00475.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Nakamura K, Morrison SF: Preoptic mechanism for cold-defensive responses to skin cooling. J Physiol. 2008;586(10):2611–20. 10.1113/jphysiol.2008.152686 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Tanaka M, McKinley MJ, McAllen RM: Roles of two preoptic cell groups in tonic and febrile control of rat tail sympathetic fibers. Am J Physiol Regul Integr Comp Physiol. 2009;296(4):R1248–57. 10.1152/ajpregu.91010.2008 [DOI] [PubMed] [Google Scholar]
  • 51. Yoshida K, Li X, Cano G, et al. : Parallel preoptic pathways for thermoregulation. J Neurosci. 2009;29(38):11954–64. 10.1523/JNEUROSCI.2643-09.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. Osaka T: Thermoregulatory responses elicited by microinjection of L-glutamate and its interaction with thermogenic effects of GABA and prostaglandin E 2 in the preoptic area. Neuroscience. 2012;226:156–64. 10.1016/j.neuroscience.2012.08.048 [DOI] [PubMed] [Google Scholar]
  • 53. Mohammed M, Kulasekara K, De Menezes RC, et al. : Inactivation of neuronal function in the amygdaloid region reduces tail artery blood flow alerting responses in conscious rats. Neuroscience. 2013;228:13–22. 10.1016/j.neuroscience.2012.10.008 [DOI] [PubMed] [Google Scholar]
  • 54. Ootsuka Y, Mohammed M: Activation of the habenula complex evokes autonomic physiological responses similar to those associated with emotional stress. Physiol Rep. 2015;3(2): pii: e12297. 10.14814/phy2.12297 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Zhang YH, Hosono T, Yanase-Fujiwara M, et al. : Effect of midbrain stimulations on thermoregulatory vasomotor responses in rats. J Physiol. 1997;503(Pt 1):177–86. 10.1111/j.1469-7793.1997.177bi.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Yoshida K, Maruyama M, Hosono T, et al. : Fos expression induced by warming the preoptic area in rats. Brain Res. 2002;933(2):109–17. 10.1016/S0006-8993(02)02287-4 [DOI] [PubMed] [Google Scholar]
  • 57. Yoshida K, Konishi M, Nagashima K, et al. : Fos activation in hypothalamic neurons during cold or warm exposure: projections to periaqueductal gray matter. Neuroscience. 2005;133(4):1039–46. 10.1016/j.neuroscience.2005.03.044 [DOI] [PubMed] [Google Scholar]
  • 58. Shibata M, Uno T, Hashimoto M: Disinhibition of lower midbrain neurons enhances non-shivering thermogenesis in anesthetized rats. Brain Res. 1999;833(2):242–50. 10.1016/S0006-8993(99)01532-2 [DOI] [PubMed] [Google Scholar]
  • 59. Cao WH, Madden CJ, Morrison SF: Inhibition of brown adipose tissue thermogenesis by neurons in the ventrolateral medulla and in the nucleus tractus solitarius. Am J Physiol Regul Integr Comp Physiol. 2010;299(1):R277–90. 10.1152/ajpregu.00039.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Tupone D, Madden CJ, Morrison SF: Central activation of the A1 adenosine receptor (A1AR) induces a hypothermic, torpor-like state in the rat. J Neurosci. 2013;33(36):14512–25. 10.1523/JNEUROSCI.1980-13.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Sato H, Hashitani T, Isobe Y, et al. : Descending influences from nucleus raphe magnus on fusimotor neurone activity in rats. J Therm Biol. 1990;15(3–4):259–65. 10.1016/0306-4565(90)90012-7 [DOI] [Google Scholar]
  • 62. Tanaka M, Owens NC, Nagashima K, et al. : Reflex activation of rat fusimotor neurons by body surface cooling, and its dependence on the medullary raphe. J Physiol. 2006;572(Pt 2):569–83. 10.1113/jphysiol.2005.102400 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63. Almeida MC, Hew-Butler T, Soriano RN, et al. : Pharmacological blockade of the cold receptor TRPM8 attenuates autonomic and behavioral cold defenses and decreases deep body temperature. J Neurosci. 2012;32(6):2086–99. 10.1523/JNEUROSCI.5606-11.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Tabarean IV, Behrens MM, Bartfai T, et al. : Prostaglandin E 2-increased thermosensitivity of anterior hypothalamic neurons is associated with depressed inhibition. Proc Natl Acad Sci U S A. 2004;101(8):2590–5. 10.1073/pnas.0308718101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Boulant JA: Counterpoint: Heat-induced membrane depolarization of hypothalamic neurons: an unlikely mechanism of central thermosensitivity. Am J Physiol Regul Integr Comp Physiol. 2006;290(5):R1481–4; discussion R1484. [DOI] [PubMed] [Google Scholar]
  • 66. Kobayashi S, Hori A, Matsumura K, et al. : Point: Heat-induced membrane depolarization of hypothalamic neurons: a putative mechanism of central thermosensitivity. Am J Physiol Regul Integr Comp Physiol. 2006;290(5):R1479–80; discussion R1484. 10.1152/ajpregu.00655.2005 [DOI] [PubMed] [Google Scholar]
  • 67. Wechselberger M, Wright CL, Bishop GA, et al. : Ionic channels and conductance-based models for hypothalamic neuronal thermosensitivity. Am J Physiol Regul Integr Comp Physiol. 2006;291(3):R518–29. 10.1152/ajpregu.00039.2006 [DOI] [PubMed] [Google Scholar]
  • 68. Craig AD: How do you feel? Interoception: the sense of the physiological condition of the body. Nat Rev Neurosci. 2002;3(8):655–66. 10.1038/nrn894 [DOI] [PubMed] [Google Scholar]
  • 69. Craig AD, Bushnell MC, Zhang ET, et al. : A thalamic nucleus specific for pain and temperature sensation. Nature. 1994;372(6508):770–3. 10.1038/372770a0 [DOI] [PubMed] [Google Scholar]
  • 70. Hylden JL, Anton F, Nahin RL: Spinal lamina I projection neurons in the rat: collateral innervation of parabrachial area and thalamus. Neuroscience. 1989;28(1):27–37. 10.1016/0306-4522(89)90229-7 [DOI] [PubMed] [Google Scholar]
  • 71. Li J, Xiong K, Pang Y, et al. : Medullary dorsal horn neurons providing axons to both the parabrachial nucleus and thalamus. J Comp Neurol. 2006;498(4):539–51. 10.1002/cne.21068 [DOI] [PubMed] [Google Scholar]
  • 72. Bratincsák A, Palkovits M: Activation of brain areas in rat following warm and cold ambient exposure. Neuroscience. 2004;127(2):385–97. 10.1016/j.neuroscience.2004.05.016 [DOI] [PubMed] [Google Scholar]
  • 73. Nakamura K, Morrison SF: A thermosensory pathway that controls body temperature. Nat Neurosci. 2008;11(1):62–71. 10.1038/nn2027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74. Nakamura K, Morrison SF: A thermosensory pathway mediating heat-defense responses. Proc Natl Acad Sci U S A. 2010;107(19):8848–53. 10.1073/pnas.0913358107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Madden CJ, Morrison SF: Hypoxic activation of arterial chemoreceptors inhibits sympathetic outflow to brown adipose tissue in rats. J Physiol. 2005;566(Pt 2):559–73. 10.1113/jphysiol.2005.086322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Madden CJ: Glucoprivation in the ventrolateral medulla decreases brown adipose tissue sympathetic nerve activity by decreasing the activity of neurons in raphe pallidus. Am J Physiol Regul Integr Comp Physiol. 2012;302(2):R224–32. 10.1152/ajpregu.00449.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Baker MA, Doris PA: Control of evaporative heat loss during changes in plasma osmolality in the cat. J Physiol. 1982;328(1):535–45. 10.1113/jphysiol.1982.sp014282 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Takamata A, Mack GW, Gillen CM, et al. : Osmoregulatory modulation of thermal sweating in humans: reflex effects of drinking. Am J Physiol. 1995;268(2 Pt 2):R414–22. [DOI] [PubMed] [Google Scholar]
  • 79. Kenney WL, Stanhewicz AE, Bruning RS, et al. : Blood pressure regulation III: what happens when one system must serve two masters: temperature and pressure regulation? Eur J Appl Physiol. 2014;114(3):467–79. 10.1007/s00421-013-2652-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. Székely M: The vagus nerve in thermoregulation and energy metabolism. Auton Neurosci. 2000;85(1–3):26–38. 10.1016/S1566-0702(00)00217-4 [DOI] [PubMed] [Google Scholar]
  • 81. Tsukita S, Yamada T, Uno K, et al. : Hepatic glucokinase modulates obesity predisposition by regulating BAT thermogenesis via neural signals. Cell Metab. 2012;16(6):825–32. 10.1016/j.cmet.2012.11.006 [DOI] [PubMed] [Google Scholar]
  • 82. Ono K, Tsukamoto-Yasui M, Hara-Kimura Y, et al. : Intragastric administration of capsiate, a transient receptor potential channel agonist, triggers thermogenic sympathetic responses. J Appl Physiol (1985). 2011;110(3):789–98. 10.1152/japplphysiol.00128.2010 [DOI] [PubMed] [Google Scholar]
  • 83. Blouet C, Schwartz GJ: Duodenal lipid sensing activates vagal afferents to regulate non-shivering brown fat thermogenesis in rats. PLoS One. 2012;7(12):e51898. 10.1371/journal.pone.0051898 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Ngampramuan S, Cerri M, Del Vecchio F, et al. : Thermoregulatory correlates of nausea in rats and musk shrews. Oncotarget. 2014;5(6):1565–75. 10.18632/oncotarget.1732 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Romanovsky AA, Almeida MC, Garami A, et al. : The transient receptor potential vanilloid-1 channel in thermoregulation: a thermosensor it is not. Pharmacol Rev. 2009;61(3):228–61. 10.1124/pr.109.001263 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Peters JH, McDougall SJ, Fawley JA, et al. : Primary afferent activation of thermosensitive TRPV1 triggers asynchronous glutamate release at central neurons. Neuron. 2010;65(5):657–69. 10.1016/j.neuron.2010.02.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87. Fawley JA, Hofmann ME, Andresen MC: Cannabinoid 1 and transient receptor potential vanilloid 1 receptors discretely modulate evoked glutamate separately from spontaneous glutamate transmission. J Neurosci. 2014;34(24):8324–32. 10.1523/JNEUROSCI.0315-14.2014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88. Fawley JA, Hofmann ME, Largent-Milnes TM, et al. : Temperature differentially facilitates spontaneous but not evoked glutamate release from cranial visceral primary afferents. PLoS One. 2015;10(5):e0127764. 10.1371/journal.pone.0127764 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Shoudai K, Peters JH, McDougall SJ, et al. : Thermally active TRPV1 tonically drives central spontaneous glutamate release. J Neurosci. 2010;30(43):14470–5. 10.1523/JNEUROSCI.2557-10.2010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90. Steiner AA, Turek VF, Almeida MC, et al. : Nonthermal activation of transient receptor potential vanilloid-1 channels in abdominal viscera tonically inhibits autonomic cold-defense effectors. J Neurosci. 2007;27(28):7459–68. 10.1523/JNEUROSCI.1483-07.2007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Hypothermia after Cardiac Arrest Study Group: Mild therapeutic hypothermia to improve the neurologic outcome after cardiac arrest. N Engl J Med. 2002;346(8):549–56. 10.1056/NEJMoa012689 [DOI] [PubMed] [Google Scholar]
  • 92. Shankaran S, Laptook AR, Ehrenkranz RA, et al. : Whole-body hypothermia for neonates with hypoxic-ischemic encephalopathy. N Engl J Med. 2005;353(15):1574–84. 10.1056/NEJMcps050929 [DOI] [PubMed] [Google Scholar]
  • 93. van der Worp HB, van Gijn J: Clinical practice. Acute ischemic stroke. N Engl J Med. 2007;357(6):572–9. 10.1056/NEJMcp072057 [DOI] [PubMed] [Google Scholar]
  • 94. Hemmen TM, Lyden PD: Hypothermia after acute ischemic stroke. J Neurotrauma. 2009;26(3):387–91. 10.1089/neu.2008.0574 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Madden CJ, Tupone D, Cano G, et al. : α2 Adrenergic receptor-mediated inhibition of thermogenesis. J Neurosci. 2013;33(5):2017–28. 10.1523/JNEUROSCI.4701-12.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Tupone D, Madden CJ, Morrison SF: Autonomic regulation of brown adipose tissue thermogenesis in health and disease: potential clinical applications for altering BAT thermogenesis. Front Neurosci. 2014;8:14. 10.3389/fnins.2014.00014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Cerri M, Mastrotto M, Tupone D, et al. : The inhibition of neurons in the central nervous pathways for thermoregulatory cold defense induces a suspended animation state in the rat. J Neurosci. 2013;33(7):2984–93. 10.1523/JNEUROSCI.3596-12.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Muzzi M, Blasi F, Masi A, et al. : Neurological basis of AMP-dependent thermoregulation and its relevance to central and peripheral hyperthermia. J Cereb Blood Flow Metab. 2013;33(2):183–90. 10.1038/jcbfm.2012.157 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Miyazawa S, Shimizu Y, Shiina T, et al. : Central A1-receptor activation associated with onset of torpor protects the heart against low temperature in the Syrian hamster. Am J Physiol Regul Integr Comp Physiol. 2008;295(3):R991–6. 10.1152/ajpregu.00142.2008 [DOI] [PubMed] [Google Scholar]
  • 100. Jinka TR, Tøien Ø, Drew KL: Season primes the brain in an arctic hibernator to facilitate entrance into torpor mediated by adenosine A 1 receptors. J Neurosci. 2011;31(30):10752–8. 10.1523/JNEUROSCI.1240-11.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Iliff BW, Swoap SJ: Central adenosine receptor signaling is necessary for daily torpor in mice. Am J Physiol Regul Integr Comp Physiol. 2012;303(5):R477–84. 10.1152/ajpregu.00081.2012 [DOI] [PubMed] [Google Scholar]
  • 102. Feketa VV, Balasubramanian A, Flores CM, et al. : Shivering and tachycardic responses to external cooling in mice are substantially suppressed by TRPV1 activation but not by TRPM8 inhibition. Am J Physiol Regul Integr Comp Physiol. 2013;305(9):R1040–50. 10.1152/ajpregu.00296.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Feketa VV, Zhang Y, Cao Z, et al. : Transient receptor potential melastatin 8 channel inhibition potentiates the hypothermic response to transient receptor potential vanilloid 1 activation in the conscious mouse. Crit Care Med. 2014;42(5):e355–63. 10.1097/CCM.0000000000000229 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104. Cypess AM, Lehman S, Williams G, et al. : Identification and importance of brown adipose tissue in adult humans. N Engl J Med. 2009;360(15):1509–17. 10.1056/NEJMoa0810780 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105. Saito M, Okamatsu-Ogura Y, Matsushita M, et al. : High incidence of metabolically active brown adipose tissue in healthy adult humans: effects of cold exposure and adiposity. Diabetes. 2009;58(7):1526–31. 10.2337/db09-0530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. van Marken Lichtenbelt WD, Vanhommerig JW, Smulders NM, et al. : Cold-activated brown adipose tissue in healthy men. N Engl J Med. 2009;360(15):1500–8. 10.1056/NEJMoa0808718 [DOI] [PubMed] [Google Scholar]
  • 107. Orava J, Nuutila P, Lidell ME, et al. : Different metabolic responses of human brown adipose tissue to activation by cold and insulin. Cell Metab. 2011;14(2):272–9. 10.1016/j.cmet.2011.06.012 [DOI] [PubMed] [Google Scholar]
  • 108. Lowell BB, S-Susulic V, Hamann A, et al. : Development of obesity in transgenic mice after genetic ablation of brown adipose tissue. Nature. 1993;366(6457):740–2. 10.1038/366740a0 [DOI] [PubMed] [Google Scholar]
  • 109. Hamann A, Flier JS, Lowell BB: Decreased brown fat markedly enhances susceptibility to diet-induced obesity, diabetes, and hyperlipidemia. Endocrinology. 1996;137(1):21–9. 10.1210/endo.137.1.8536614 [DOI] [PubMed] [Google Scholar]
  • 110. Kopecky J, Clarke G, Enerbäck S, et al. : Expression of the mitochondrial uncoupling protein gene from the aP2 gene promoter prevents genetic obesity. J Clin Invest. 1995;96(6):2914–23. 10.1172/JCI118363 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Hanssen MJ, van der Lans AA, Brans B, et al. : Short-term cold acclimation recruits brown adipose tissue in obese humans. Diabetes. 2015; pii: db151372. 10.2337/db15-1372 [DOI] [PubMed] [Google Scholar]
  • 112. Orava J, Nuutila P, Noponen T, et al. : Blunted metabolic responses to cold and insulin stimulation in brown adipose tissue of obese humans. Obesity (Silver Spring). 2013;21(11):2279–87. 10.1002/oby.20456 [DOI] [PubMed] [Google Scholar]
  • 113. Sakaguchi T, Arase K, Fisler JS, et al. : Effect of a high-fat diet on firing rate of sympathetic nerves innervating brown adipose tissue in anesthetized rats. Physiol Behav. 1989;45(6):1177–82. 10.1016/0031-9384(89)90106-6 [DOI] [PubMed] [Google Scholar]
  • 114. Levin BE: Reduced norepinephrine turnover in organs and brains of obesity-prone rats. Am J Physiol. 1995;268(2 Pt 2):R389–94. [DOI] [PubMed] [Google Scholar]
  • 115. Madden CJ: Consumption of a high fat diet inhibits sympathetic outflow to brown adipose tissue (BAT) via vagal afferent activation of neurons in the Nucleus Tractus Solitarius (NTS). Auton Neurosci. 2015;192:13 10.1016/j.autneu.2015.07.288 [DOI] [Google Scholar]

Articles from F1000Research are provided here courtesy of F1000 Research Ltd

RESOURCES