Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

journal article

Noise and Stress: A Comprehensive Approach

Environmental Health Perspectives

Vol. 41 (Oct., 1981)

, pp. 291-309 (19 pages)

Published By: The National Institute of Environmental Health Sciences

https://doi.org/10.2307/3429324

https://www.jstor.org/stable/3429324

Read and download

Log in through your school or library

Read Online (Free) relies on page scans, which are not currently available to screen readers. To access this article, please contact JSTOR User Support. We'll provide a PDF copy for your screen reader.

With a personal account, you can read up to 100 articles each month for free.

Get Started

Already have an account? Log in

Monthly Plan

  • Access everything in the JPASS collection
  • Read the full-text of every article
  • Download up to 10 article PDFs to save and keep
$19.50/month

Yearly Plan

  • Access everything in the JPASS collection
  • Read the full-text of every article
  • Download up to 120 article PDFs to save and keep
$199/year

Abstract

The fundamental purposes of hearing are to alert and to warn. As a result sound directly evokes emotions and actions. The processing of sound by the brain is outlined to provide a biological and psychological basis for understanding the way in which sound can become a human stressor. The auditory orienting response, startle reflex and defensive response translate sound stimuli into action and sometimes into stress induced bodily changes through "fight or flight" neural mechanisms. The literature on the health and mental health effects of noise then is reviewed in the context of an integrated model that offers a holistic approach to noise research and public policy formulation. The thesis of this paper is that research upon, and efforts to prevent or minimize the harmful effects of noise have suffered from the lack of a full appreciation of the ways in which humans process and react to sound.

Journal Information

\Environmental Health Perspectives (EHP) is a monthly peer-reviewed journal of research and news published with support from the National Institute of Environmental Health Sciences, National Institutes of Health, U.S. Department of Health and Human Services. The mission of EHP is to serve as a forum for the discussion of the interrelationships between the environment and human health by publishing high-quality research and news of the field. With an impact factor of 7.03, EHP is the third-ranked journal in Public, Environmental, and Occupational Health, the fourth-ranked journal in Toxicology, and the fifth-ranked journal in Environmental Sciences Current Issues of Environmental Health Perspectives are freely available to all users on the journal's website.

Publisher Information

The mission of the NIEHS is to reduce the burden of human illness and disability by understanding how the environment influences the development and progression of human disease. To have the greatest impact on preventing disease and improving human health, the NIEHS focuses on basic science, disease-oriented research, global environmental health, and multidisciplinary training for researchers.

Journal Article

Karel Pacák,

1Pediatric and Reproductive Endocrinology Branch (K.P.), National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, Maryland

2National Institute of Child Health and Human Development and Clinical Neurocardiology Section (K.P.), National Institute of Neurological Disorders and Stroke, National Institutes of Health, Bethesda, Maryland

*Address requests for reprints to: Karel Pacák, M.D., Ph.D., D.Sc., National Institute of Child Health and Human Development, National Institutes of Health, Building 10, Room 9D42, 10 Center Drive, Bethesda, Maryland 20892-1583.

Search for other works by this author on:

Miklós Palkovits

3Laboratory of Genetics (M.P.), National Institute of Mental Health, National Institutes of Health, Bethesda, Maryland 20892-1583

4Laboratory of Neuromorphology, Semmelweis University, 1094 Budapest, Hungary (M.P.)

Search for other works by this author on:

Published:

01 August 2001

  • PDF
  • Split View
    • Article contents
    • Figures & tables
    • Video
    • Audio
    • Supplementary Data
  • Cite

    Cite

    Karel Pacák, Miklós Palkovits, Stressor Specificity of Central Neuroendocrine Responses: Implications for Stress-Related Disorders, Endocrine Reviews, Volume 22, Issue 4, 1 August 2001, Pages 502–548, https://doi.org/10.1210/edrv.22.4.0436

    Close

    • Email
    • Twitter
    • Facebook
    • More

Close

Navbar Search Filter Microsite Search Term Search

Abstract

Despite the fact that many research articles have been written about stress and stress-related diseases, no scientifically accepted definition of stress exists. Selye introduced and popularized stress as a medical and scientific idea. He did not deny the existence of stressor-specific response patterns; however, he emphasized that such responses did not constitute stress, only the shared nonspecific component. In this review we focus mainly on the similarities and differences between the neuroendocrine responses (especially the sympathoadrenal and the sympathoneuronal systems and the hypothalamo-pituitary-adrenocortical axis) among various stressors and a strategy for testing Selye’s doctrine of nonspecificity. In our experiments, we used five different stressors: immobilization, hemorrhage, cold exposure, pain, or hypoglycemia. With the exception of immobilization stress, these stressors also differed in their intensities. Our results showed marked heterogeneity of neuroendocrine responses to various stressors and that each stressor has a neurochemical “signature.” By examining changes of Fos immunoreactivity in various brain regions upon exposure to different stressors, we also attempted to map central stressor-specific neuroendocrine pathways. We believe the existence of stressor-specific pathways and circuits is a clear step forward in the study of the pathogenesis of stress-related disorders and their proper treatment. Finally, we define stress as a state of threatened homeostasis (physical or perceived treat to homeostasis). During stress, an adaptive compensatory specific response of the organism is activated to sustain homeostasis. The adaptive response reflects the activation of specific central circuits and is genetically and constitutionally programmed and constantly modulated by environmental factors.

  • I.

    Introduction

  • II.

    Stress Concept

    • A.

      Definition of stress

    • B.

      Classification of stressful stimuli

    • C.

      Selye’s doctrine of nonspecificity revisited

  • III.

    Brain Regions Involved in Neuroendocrine Responses to Stress

    • A.

      Central autonomic system

    • B.

      Central aminergic systems

    • C.

      Noncatecholaminergic brainstem neurons

    • D.

      Thalamus

    • E.

      Neuroendocrine hypothalamus

    • F.

      Limbic system

  • IV.

    Methods used for Mapping Stressor-Specific Neuronal Circuits

    • A.

      Intracerebral microdialysis

    • B.

      Protooncogene-“immediate early genes”—immunohistochemistry

  • V.

    Stressor Specificity of Central Neuroendocrine Responses

    • A.

      Immobilization stress

    • B.

      Cold stress

    • C.

      Insulin-induced hypoglycemia

    • D.

      Hemorrhage

    • E.

      Pain stress

  • VI.

    Stressor-Specific Activation of Other Neuroendocrine Systems

  • VII.

    Clinical Relevance of Stressor Specificity and Future Perspectives

I. Introduction

HANS SELYE DESERVES much of the credit for introducing the term “stress,” and for popularizing the concept of stress in the scientific and medical literature of the 20th century. He wrote in his book, Stress of Life: “This lack of distinction between cause and effect was, I suppose, fostered by the fact that when I introduced the word stress into medicine in its present meaning, my English was not yet good enough for me to distinguish between the words stress and strain. It was not until several years later that the British Medical Journal called my attention to this fact, by the somewhat sarcastic remark that according to Selye stress is its own cause. Actually I should have called my phenomenon the strain reaction and that which causes it ‘stress,’ which would parallel the use of these terms in physics. However, by the time that this came to my attention, biological stress in my sense of the word was so generally accepted in various languages that I could not have redefined it. Hence, I was forced to create a neologism and introduce the word stressor, for the causative agent, into the English language, retaining stress for the resulting condition” (1).

Confusion still arises regarding what one believes defines and constitutes stress. Despite the fact that thousands of research articles have been written about stress and stress-related diseases, until now no scientifically accepted definition of stress exists (2). This results in the view that stress can be practically anything that contributes to virtually any disease in humans. Most scientists view stress as the situation when the hypothalamo-pituitary-adrenocortical (HPA) axis, represented mainly by elevated ACTH levels, is activated (3). Others suggest that activation of other systems with or without an elevation in ACTH may reflect stress-induced disturbed homeostasis (4, 5). Several review articles and book chapters have summarized data from hundreds of stress-related studies and drawn conclusions relating to different aspects of the stress response (2, 6–17). This review focuses on two major points: 1) evidence that specific stressors may elicit specific responses, and 2) different stressors may activate different brain systems by using specific pathways within the central nervous system. Particular attention has been paid to Selye’s doctrine of nonspecificity of stress responses, which has been tested in our laboratory. Based on our data using five different acute stressors (immobilization stress, hypoglycemia, nontraumatic hemorrhage, pain stress, and cold stress) and several previous reports by others (4, 18–29), we turned our attention to identifying stressor-specific neuronal circuits in brain and their involvement in stress-related diseases.

II. Stress Concept

A. Definition of stress

Cannon (30–32) was the first to introduce the term “homeostasis” to describe the “coordinated physiological processes which maintain most of the steady states in the organism.” He turned his attention to the sympathetic nervous system as an essential homeostatic system that serves to restore stress-induced disturbed homeostasis and to promote survival of the organism. Cannon was also the first to touch on the issue of specificity of stress responses since he showed, for example, that the specific stabilizing or homeostatic reaction to lack of oxygen is quite different from that with which the body responds to exposure to cold; this, in turn, is virtually the reverse of that required to resist heat (1). However, Cannon never used the term “stress.”

Selye introduced and popularized stress as a medical and scientific idea. The starting point for the elaboration of his stress theory was his report, published as a letter to Nature in 1936 (33), describing a pathological triad (adrenal enlargement, gastrointestinal ulceration, and thymicolymphatic involution) elicited by any of a variety of stressors. From this pathological triad he developed a theory of stress that attained wide popularity and aroused intense research interest but also incited controversy, which persists to the present. He defined stress as the nonspecific response (revealed after subtraction of the specific components from the total response) of the body to any demand, emphasizing that the same pathological triad— “stress syndrome”— would result from exposure to any stressor. According to Selye, these demands on the body included bacterial infection, toxins, x-irradiation, and various physical stimuli such as surgery and muscular exercise. Selye’s stress theory did not deny the existence of stressor-specific response patterns; however, he emphasized that such responses did not constitute stress, which was the shared nonspecific component.

Selye mainly focused on the HPA axis as the key effector of the stress response. He considered the adrenal cortex “the organ of integration which participates in the normal and pathological physiology of virtually all tissues in the body,” by virtue of its endocrine function (34). Indeed, administration of ACTH can elicit all three components of the pathological triad (34). However, Selye did not assert that HPA activation attending stress reflected the organ pathology in the pathological triad. If anything, Selye asserted the converse.

Selye also introduced the term “general adaptation syndrome” with its three successive phases: the alarm, resistance, and exhaustion stages. He wrote that during the stages of the “general adaptation syndrome” the intensity of the stress response might vary; however, the neural and endocrine patterns characterizing the stage of “alarm” would be essentially the same as those characterizing the other stages. He and others proposed an immense list of diseases of adaptation including hyperfunctional and dysfunctional conditions such as Cushing’s disease, hypertension, adrenal tumors, and others. Hypofunctional states included Addison’s disease and cancer (1, 2, 6, 7, 17, 34). Later, Selye proposed that most of the stressful stimuli induce two types of responses: 1) a general stress response, which is common to all stressors and involves the release of ACTH and adrenal corticosterone, and 2) individual stress responses mediated by “conditioning factors,” such as genetically determined predispositions (17).

In contrast to Selye, Cannon recognized the importance of psychological as opposed to physical responses during stress (30, 35). From an evolutionary perspective he questioned whether a stereotyped response pattern could be adaptive, recognizing that a nonspecific stress response would not have provided an advantage in natural selection and thus, would not have evolved. Others, like Mason, properly noted that in response to different stressors, activity of the HPA axis could increase, decrease, or remain unchanged, implying that the presence of a pathological triad may not indicate the occurrence of stress (2, 13–15). Mason proposed that elicitation of an emotion such as anxiety or fear constituted the basis for similar neuroendocrine responses to different stressors.

Many current views concerning what stress means and how to define and approach it exist, but none has been widely accepted. Many of these theories were discussed in detail and described by Goldstein (Ref. 2 and Table 1). Important contributions to these theories have been made by Weiner (36) and Chrousos (6) and Chrousos and Gold (7). Weiner correctly pointed to specificity of stressor responses by describing stressors as selective pressures from the physical and social environment that threaten or challenge an organism and elicit compensatory response patterns. Chrousos and Gold defined stress as a state of disharmony or of threatened homeostasis, evoking physiologically and behaviorally adaptive responses that can be specific to the stressor or generalized and nonspecific and that usually occur stereotypically, producing a “nonspecific” stress syndrome when the threat to homeostasis exceeds a threshold. They included genetic polymorphisms as well as alterations in the expression of genes and environmental factors as important determinants of individual stress responses.

TABLE 1

Summary of the effects of stressors on 1) Fos immunoreactivity, 2) PVN extracellular NE levels, 3) central nucleus of amygdala extracellular levels, and 4) plasma levels of various hormones

Stressors
ImmobilizationColdHypoglycemiaHemorrhagePain
1. Fos immunoreactivity in various brain regions            
           
Hypothalamus           
Paraventricular nucleus  +++  ++  +++ 
Supraoptic nucleus  ++  −  ±  ++  ++ 
Medial preoptic nucleus  +++  −  ++ 
Anterior hypothalamic nucleus  ++  −  − 
Arcuate nucleus  −  − 
Ventromedial nucleus  −  −  −  −  − 
Dorsomedial nucleus  ++ 
Supramamillary nucleus  ++  −  −  +++ 
Lateral hypothalamic area  ++  ++  ++ 
Thalamus           
Midline nuclei  ++  +++ 
Limbic system           
Central amygdala  ++  −  −  ++ 
Lateral septum  ++  −  −  ++ 
Medial habenula  −  −  −  −  ± 
Hippocampus  ±  −  −  −  ± 
Cingulate cortex  +++  −  +++ 
Piriform cortex  +++  −  +++ 
Midbrain           
Substantia nigra  −  −  −  −  − 
Central gray  ++  ±  ±  ++ 
Dorsal raphe  −  −  − 
Pons           
Parabrachial nuclei  ++  +++  −  ±  +++ 
Barrington nucleus  ++ 
Pontine nuclei  +++  ±  −  −  − 
Raphe nuclei  ++  −  − 
Cerebellum  −  −  −  −  − 
Medulla oblongata           
Spinal trigeminal nucleus  ++  −  −  ++ 
Peritrigemial nucleus  ++  +++  −  −  ± 
NTS  ++  ++ 
Area postrema  −  −  −  ++ 
Lateral reticular nucleus  −  ++ 
Catecholaminergic cell groups           
A1  +++  −  ++ 
A2  ++  ± 
A5  +++  −  −  −  ++ 
A6 (locus coeruleus)  +++  ±  −  +++ 
A7  +++  −  −  −  ++ 
A12 (dopamine)  −  −  −  −  − 
Spinal cord  ++  ++      +++ 
2. PVN microdialysis            
NE  +++  ±  ++ 
3. Central nucleus of amygdala            
NE  +++  ND  ND  ND  ND 
4. Plasma            
NE  +++  +++  ++  ±  +++ 
EPI  ++  ±  +++  ± 
ACTH  +++  ±  ++  ++ 
Corticosterone  +++  ++  ++  +++  +++ 

Stressors
ImmobilizationColdHypoglycemiaHemorrhagePain
1. Fos immunoreactivity in various brain regions            
           
Hypothalamus           
Paraventricular nucleus  +++  ++  +++ 
Supraoptic nucleus  ++  −  ±  ++  ++ 
Medial preoptic nucleus  +++  −  ++ 
Anterior hypothalamic nucleus  ++  −  − 
Arcuate nucleus  −  − 
Ventromedial nucleus  −  −  −  −  − 
Dorsomedial nucleus  ++ 
Supramamillary nucleus  ++  −  −  +++ 
Lateral hypothalamic area  ++  ++  ++ 
Thalamus           
Midline nuclei  ++  +++ 
Limbic system           
Central amygdala  ++  −  −  ++ 
Lateral septum  ++  −  −  ++ 
Medial habenula  −  −  −  −  ± 
Hippocampus  ±  −  −  −  ± 
Cingulate cortex  +++  −  +++ 
Piriform cortex  +++  −  +++ 
Midbrain           
Substantia nigra  −  −  −  −  − 
Central gray  ++  ±  ±  ++ 
Dorsal raphe  −  −  − 
Pons           
Parabrachial nuclei  ++  +++  −  ±  +++ 
Barrington nucleus  ++ 
Pontine nuclei  +++  ±  −  −  − 
Raphe nuclei  ++  −  − 
Cerebellum  −  −  −  −  − 
Medulla oblongata           
Spinal trigeminal nucleus  ++  −  −  ++ 
Peritrigemial nucleus  ++  +++  −  −  ± 
NTS  ++  ++ 
Area postrema  −  −  −  ++ 
Lateral reticular nucleus  −  ++ 
Catecholaminergic cell groups           
A1  +++  −  ++ 
A2  ++  ± 
A5  +++  −  −  −  ++ 
A6 (locus coeruleus)  +++  ±  −  +++ 
A7  +++  −  −  −  ++ 
A12 (dopamine)  −  −  −  −  − 
Spinal cord  ++  ++      +++ 
2. PVN microdialysis            
NE  +++  ±  ++ 
3. Central nucleus of amygdala            
NE  +++  ND  ND  ND  ND 
4. Plasma            
NE  +++  +++  ++  ±  +++ 
EPI  ++  ±  +++  ± 
ACTH  +++  ±  ++  ++ 
Corticosterone  +++  ++  ++  +++  +++ 

+++, High level activity; ++, moderate level activity; +, low level activity; ±, just detectable; −, undetectable. ND, own data not available.

TABLE 1

Summary of the effects of stressors on 1) Fos immunoreactivity, 2) PVN extracellular NE levels, 3) central nucleus of amygdala extracellular levels, and 4) plasma levels of various hormones

Stressors
ImmobilizationColdHypoglycemiaHemorrhagePain
1. Fos immunoreactivity in various brain regions            
           
Hypothalamus           
Paraventricular nucleus  +++  ++  +++ 
Supraoptic nucleus  ++  −  ±  ++  ++ 
Medial preoptic nucleus  +++  −  ++ 
Anterior hypothalamic nucleus  ++  −  − 
Arcuate nucleus  −  − 
Ventromedial nucleus  −  −  −  −  − 
Dorsomedial nucleus  ++ 
Supramamillary nucleus  ++  −  −  +++ 
Lateral hypothalamic area  ++  ++  ++ 
Thalamus           
Midline nuclei  ++  +++ 
Limbic system           
Central amygdala  ++  −  −  ++ 
Lateral septum  ++  −  −  ++ 
Medial habenula  −  −  −  −  ± 
Hippocampus  ±  −  −  −  ± 
Cingulate cortex  +++  −  +++ 
Piriform cortex  +++  −  +++ 
Midbrain           
Substantia nigra  −  −  −  −  − 
Central gray  ++  ±  ±  ++ 
Dorsal raphe  −  −  − 
Pons           
Parabrachial nuclei  ++  +++  −  ±  +++ 
Barrington nucleus  ++ 
Pontine nuclei  +++  ±  −  −  − 
Raphe nuclei  ++  −  − 
Cerebellum  −  −  −  −  − 
Medulla oblongata           
Spinal trigeminal nucleus  ++  −  −  ++ 
Peritrigemial nucleus  ++  +++  −  −  ± 
NTS  ++  ++ 
Area postrema  −  −  −  ++ 
Lateral reticular nucleus  −  ++ 
Catecholaminergic cell groups           
A1  +++  −  ++ 
A2  ++  ± 
A5  +++  −  −  −  ++ 
A6 (locus coeruleus)  +++  ±  −  +++ 
A7  +++  −  −  −  ++ 
A12 (dopamine)  −  −  −  −  − 
Spinal cord  ++  ++      +++ 
2. PVN microdialysis            
NE  +++  ±  ++ 
3. Central nucleus of amygdala            
NE  +++  ND  ND  ND  ND 
4. Plasma            
NE  +++  +++  ++  ±  +++ 
EPI  ++  ±  +++  ± 
ACTH  +++  ±  ++  ++ 
Corticosterone  +++  ++  ++  +++  +++ 

Stressors
ImmobilizationColdHypoglycemiaHemorrhagePain
1. Fos immunoreactivity in various brain regions            
           
Hypothalamus           
Paraventricular nucleus  +++  ++  +++ 
Supraoptic nucleus  ++  −  ±  ++  ++ 
Medial preoptic nucleus  +++  −  ++ 
Anterior hypothalamic nucleus  ++  −  − 
Arcuate nucleus  −  − 
Ventromedial nucleus  −  −  −  −  − 
Dorsomedial nucleus  ++ 
Supramamillary nucleus  ++  −  −  +++ 
Lateral hypothalamic area  ++  ++  ++ 
Thalamus           
Midline nuclei  ++  +++ 
Limbic system           
Central amygdala  ++  −  −  ++ 
Lateral septum  ++  −  −  ++ 
Medial habenula  −  −  −  −  ± 
Hippocampus  ±  −  −  −  ± 
Cingulate cortex  +++  −  +++ 
Piriform cortex  +++  −  +++ 
Midbrain           
Substantia nigra  −  −  −  −  − 
Central gray  ++  ±  ±  ++ 
Dorsal raphe  −  −  − 
Pons           
Parabrachial nuclei  ++  +++  −  ±  +++ 
Barrington nucleus  ++ 
Pontine nuclei  +++  ±  −  −  − 
Raphe nuclei  ++  −  − 
Cerebellum  −  −  −  −  − 
Medulla oblongata           
Spinal trigeminal nucleus  ++  −  −  ++ 
Peritrigemial nucleus  ++  +++  −  −  ± 
NTS  ++  ++ 
Area postrema  −  −  −  ++ 
Lateral reticular nucleus  −  ++ 
Catecholaminergic cell groups           
A1  +++  −  ++ 
A2  ++  ± 
A5  +++  −  −  −  ++ 
A6 (locus coeruleus)  +++  ±  −  +++ 
A7  +++  −  −  −  ++ 
A12 (dopamine)  −  −  −  −  − 
Spinal cord  ++  ++      +++ 
2. PVN microdialysis            
NE  +++  ±  ++ 
3. Central nucleus of amygdala            
NE  +++  ND  ND  ND  ND 
4. Plasma            
NE  +++  +++  ++  ±  +++ 
EPI  ++  ±  +++  ± 
ACTH  +++  ±  ++  ++ 
Corticosterone  +++  ++  ++  +++  +++ 

+++, High level activity; ++, moderate level activity; +, low level activity; ±, just detectable; −, undetectable. ND, own data not available.

AuthorDefinition of stressReference
Bernard  Introduced the term milieu interieur.  (496) 
Cannon  Introduced the term homeostasis.  (30) 
Selye  Introduced the terms stress, eustress, distress, and stressor. Introduced the general adaptation syndrome. Defined stress as the nonspecific response of the body to any demand upon it.  (1, 33, 34) 
Mason  Criticized Selye’s doctrine of nonspecificity. Anxiety and fear understood as the main factors contributing to nonspecific responses upon exposure to various stressors.  (14) 
Hennesy and Levine  Introduced a “psychoendocrine hypothesis” of stress and arousal.  (497) 
Krantz and Lazar  Defined psychological stress in terms of a “transaction” between an organism and its environment.  (498) 
Munck and Guyre  Incorporated inhibitory effects of glucocorticoids in the development of “diseases of adaptation”  (499) 
Levine and Ursin  Incorporated adaptive biological responses into the definition of stress.  (500) 
Weiner  Stress defined as an external experience or phenomena to the organism.  (36) 
Chrousos and Gold  Defined stress as a state of disharmony or of threatened homeostasis evoking both specific and nonspecific responses. Included genetic polymorphisms as important determinants of individual stress responses.  (7) 
Goldstein  Defined stress as a condition where expectations, whether genetically programmed, established by prior learning, or deduced from circumstances, do not match the current or anticipated perceptions of the internal or external environment. This discrepancy between what is observed or sensed and what is expected or programmed elicits patterned, compensatory responses.  (2) 
McEwen  Incorporated the term allostasis as the process of adaptation of the body upon the exposure to various stressors.  (11) 

AuthorDefinition of stressReference
Bernard  Introduced the term milieu interieur.  (496) 
Cannon  Introduced the term homeostasis.  (30) 
Selye  Introduced the terms stress, eustress, distress, and stressor. Introduced the general adaptation syndrome. Defined stress as the nonspecific response of the body to any demand upon it.  (1, 33, 34) 
Mason  Criticized Selye’s doctrine of nonspecificity. Anxiety and fear understood as the main factors contributing to nonspecific responses upon exposure to various stressors.  (14) 
Hennesy and Levine  Introduced a “psychoendocrine hypothesis” of stress and arousal.  (497) 
Krantz and Lazar  Defined psychological stress in terms of a “transaction” between an organism and its environment.  (498) 
Munck and Guyre  Incorporated inhibitory effects of glucocorticoids in the development of “diseases of adaptation”  (499) 
Levine and Ursin  Incorporated adaptive biological responses into the definition of stress.  (500) 
Weiner  Stress defined as an external experience or phenomena to the organism.  (36) 
Chrousos and Gold  Defined stress as a state of disharmony or of threatened homeostasis evoking both specific and nonspecific responses. Included genetic polymorphisms as important determinants of individual stress responses.  (7) 
Goldstein  Defined stress as a condition where expectations, whether genetically programmed, established by prior learning, or deduced from circumstances, do not match the current or anticipated perceptions of the internal or external environment. This discrepancy between what is observed or sensed and what is expected or programmed elicits patterned, compensatory responses.  (2) 
McEwen  Incorporated the term allostasis as the process of adaptation of the body upon the exposure to various stressors.  (11) 

AuthorDefinition of stressReference
Bernard  Introduced the term milieu interieur.  (496) 
Cannon  Introduced the term homeostasis.  (30) 
Selye  Introduced the terms stress, eustress, distress, and stressor. Introduced the general adaptation syndrome. Defined stress as the nonspecific response of the body to any demand upon it.  (1, 33, 34) 
Mason  Criticized Selye’s doctrine of nonspecificity. Anxiety and fear understood as the main factors contributing to nonspecific responses upon exposure to various stressors.  (14) 
Hennesy and Levine  Introduced a “psychoendocrine hypothesis” of stress and arousal.  (497) 
Krantz and Lazar  Defined psychological stress in terms of a “transaction” between an organism and its environment.  (498) 
Munck and Guyre  Incorporated inhibitory effects of glucocorticoids in the development of “diseases of adaptation”  (499) 
Levine and Ursin  Incorporated adaptive biological responses into the definition of stress.  (500) 
Weiner  Stress defined as an external experience or phenomena to the organism.  (36) 
Chrousos and Gold  Defined stress as a state of disharmony or of threatened homeostasis evoking both specific and nonspecific responses. Included genetic polymorphisms as important determinants of individual stress responses.  (7) 
Goldstein  Defined stress as a condition where expectations, whether genetically programmed, established by prior learning, or deduced from circumstances, do not match the current or anticipated perceptions of the internal or external environment. This discrepancy between what is observed or sensed and what is expected or programmed elicits patterned, compensatory responses.  (2) 
McEwen  Incorporated the term allostasis as the process of adaptation of the body upon the exposure to various stressors.  (11) 

AuthorDefinition of stressReference
Bernard  Introduced the term milieu interieur.  (496) 
Cannon  Introduced the term homeostasis.  (30) 
Selye  Introduced the terms stress, eustress, distress, and stressor. Introduced the general adaptation syndrome. Defined stress as the nonspecific response of the body to any demand upon it.  (1, 33, 34) 
Mason  Criticized Selye’s doctrine of nonspecificity. Anxiety and fear understood as the main factors contributing to nonspecific responses upon exposure to various stressors.  (14) 
Hennesy and Levine  Introduced a “psychoendocrine hypothesis” of stress and arousal.  (497) 
Krantz and Lazar  Defined psychological stress in terms of a “transaction” between an organism and its environment.  (498) 
Munck and Guyre  Incorporated inhibitory effects of glucocorticoids in the development of “diseases of adaptation”  (499) 
Levine and Ursin  Incorporated adaptive biological responses into the definition of stress.  (500) 
Weiner  Stress defined as an external experience or phenomena to the organism.  (36) 
Chrousos and Gold  Defined stress as a state of disharmony or of threatened homeostasis evoking both specific and nonspecific responses. Included genetic polymorphisms as important determinants of individual stress responses.  (7) 
Goldstein  Defined stress as a condition where expectations, whether genetically programmed, established by prior learning, or deduced from circumstances, do not match the current or anticipated perceptions of the internal or external environment. This discrepancy between what is observed or sensed and what is expected or programmed elicits patterned, compensatory responses.  (2) 
McEwen  Incorporated the term allostasis as the process of adaptation of the body upon the exposure to various stressors.  (11) 

Based on an intervening variable as a theoretical construct in psychology that links a stimulus to a behavioral response, Goldstein (2) recently introduced a new definition of stress. He defined stress as “a condition where expectations, whether genetically programmed, established by prior learning, or deduced from circumstances, do not match the current or anticipated perceptions of the internal or external environment, and this discrepancy between what is observed or sensed and what is expected or programmed elicits patterned, compensatory responses.”

Recently, McEwen (11) introduced the term “allostasis” into stress research. Allostasis, which may be defined as an ability to maintain stability of the internal milieu through change, was originally proposed by Sterling and Eyer (37). As discussed recently in detail by McEwen (11), allostasis refers to the active process of adaptation by productions of various mediators such as adrenal steroids, catecholamines, cytokines, tissue mediators, and immediate early genes. Upon exposure to a chronic stressful situation, physiological responses are initiated, leading to allostatic (adaptive) responses. These responses involve major systems similar to the stress effector systems that were described previously. If allostatic responses are efficient, adaptation occurs and the organism is protected from damage. In situations where allostatic responses are prolonged, inadequate, overstimulated by repeated “hits” from multiple stressors or if a lack of adaptation occurs, allostatic load results in maladaptation and damage to various organs (11, 16). In contrast to homeostatic mechanisms, allostatic regulations are broader and do not depend on set-point mechanisms, signals are not constant, and anticipation of need is an important element. Another aspect of this theory is that allostatic load also reflects aspects of lifestyle (e.g., eating a high-fat diet, lack of exercise, etc.) and disturbances of diurnal rhythms (e.g., sleep deprivation) that result from overexposure of various tissues to stress mediators. Allostatic theory also continues Selye’s notion of “conditioning factors” to explain individual differences in stress responses.

Based on our previous findings of the existence of stressor-specific neuroendocrine responses and mapping of stressor-specific central circuits that participate in these responses (see below), we attempted to define stress as a state of threatened homeostasis (physical or perceived treat to homeostasis). During stress, an adaptive compensatory specific response of the organism is activated to sustain homeostasis. The adaptive response reflects the activation of specific central circuits and is genetically and constitutionally programmed and constantly modulated by environmental factors.

Another “mainstream” theory of stress has been offered recently by molecular biologists regarding the role of heat shock proteins in cellular survival (38). Ironically, their theory posits essentially the same doctrine of nonspecificity that Selye espoused; regardless of the insult, cells respond in the same way.

Sapolsky and co-workers (39–41) and McEwen et al. (42) introduced and discussed in great detail new aspects of stress in terms of its adverse effects on various brain regions, especially the hippocampus. Upon exposure to stressors, glucocorticoids are released and act on target cells including brain cells. This central action of glucocorticoids is associated with behavioral, neurochemical, and neurodegenerative changes. Neurodegenerative changes are of great importance since they occur in the hippocampus, one of the brain regions involved in memory processes and other cognitive functions as well as in the regulation of the HPA axis (43, 44). Prolonged exposure to high glucocorticoid levels, as commonly seen upon exposure to chronic stress, causes premature age-related changes in hippocampal electrical activity (45) and dendritic and neuronal atrophy often accelerated upon exposure to neurological insults (e.g., hypoxia) (41). In contrast to these neurodegenerative changes, glucocorticoids also evoke responses that are neuroprotective during exposure to stress (46). For example, various stressors and glucocorticoids increase mRNA expression for oligodendrocyte markers such as glycerol-3-phosphate dehydrogenase and neuronal neurotrophin-3 (46, 47). According to Nichols et al. (46), the activation of these markers serves an important adaptive mechanism in promoting oligodendrocyte survival in response to high glucocorticoids levels. In general, glucocorticoids are viewed as key stress hormones that permit, stimulate, or suppress ongoing stress responses, or are preparative during exposure to a subsequent stressor (48).

Oxidative stress is another type of stressor that participates in neurodegeneration of brain cells (49, 50). Expression of mRNA for glial fibrillary acidic protein, an intermediate filament of astrocytes, is increased by oxidative stress, resulting in astrocyte hyperactivation and subsequent damage (49, 50). Recently, CRH and mifepristone, a potent antagonist of glucocorticoid and progesterone receptors, have been shown to protect against neuronal cell death upon exposure to oxidative stress (51, 52). CRH has a neuroprotective action in CRH receptor type 1-expressing neurons against oxidative cell death (52). This CRH protective function is accompanied by increased release of nonamyloidogenic soluble amyloid β-precursor protein and by suppression of nuclear factor-κB. The neuroprotective activity of these drugs may play an important role in new therapeutic interventions for neurodegenerative conditions such as stroke or Alzheimer’s disease (51).

B. Classification of stressful stimuli

A stressor may be viewed as a stimulus that disrupts homeostasis. In general, stressors can be divided into four main categories: 1) physical stressors that have either a negative or, in some situations, a positive psychological component; 2) psychological stressors that reflect a learned response to previously experienced adverse conditions; 3) social stressors reflecting disturbed interactions among individuals; and 4) stressors that challenge cardiovascular and metabolic homeostasis (4, 10, 53). Physical stressors include cold, heat, intense radiation, noise, vibration, and many others. Chemical stressors include all poisons. Pain stress may be elicited by many different chemical and physical agents. Psychological stressors profoundly affect emotional processes and may result in behavioral changes such as anxiety, fear, or frustration. Social stressors include an animal’s placement into the territory of a dominant animal, and in humans, unemployment and marital separation, among others, are considered social stressors. Stressors that disturb cardiovascular or metabolic homeostasis include exercise, orthostasis, upright tilt, heat exposure, hypoglycemia, and hemorrhage. Many of the stressors described above and used in animal research, however, are mixed and act in concert, such as handling, immobilization stress, anticipation of a painful stimulus, and hypotensive hemorrhage.

In terms of duration, stressors may be divided into two main categories: acute (single, intermittent, and time-limited exposure vs. continuous exposure) vs. chronic (intermittent and prolonged exposure vs. continuous exposure) stressors. It should be noted that many stressors differ in their intensity.

The adaptive responses that are elicited in response to an acute stressor include the physiological and behavioral processes that are essential to reestablish homeostatic balance. During an acute stress response, physiological processes are important to redirect energy utilization among various organs and selectively inhibit or stimulate various organ systems or their components to mobilize energy reserves and to be prepared for exposure to additional, unpredictable challenges. Thus, upon exposure to metabolic stressors, certain tissues tend to reduce their consumption of energy while others, especially those that are important for locomotor activity, receive sufficient nutrients to function properly. The central nervous system also has priority during metabolic stress responses and preferentially receives a sufficient amount of nutrients from the circulation. The increased supply of energy to “crucial” organs is achieved preferentially by release of catecholamines and glucocorticoids that, in general, increase gluconeogenesis and glycogenolysis, inhibit glucose uptake, and enhance proteolysis and lipolysis. The immune system is another essential component of these physiologically adaptive stress responses.

In terms of health consequences during exposure to various stressors, the mechanisms of coping with stress and relevant feedback mechanisms are essential for an organism to develop less severe stress-related health consequences and to survive (54–57). Coping responses during stress may be defined as cognitive and behavioral responses to manage stress (54–58). Cohen and Lazarus (59) defined five primary goals for successful coping with stress: 1) reduce harmful environmental conditions and enhance the prospect for recovery; 2) tolerate or adjust to negative events; 3) maintain a positive self-image; 4) maintain emotional equilibrium; and finally 5) preserve social relationships. There are a number of factors that determine whether an individual will cope effectively with a particular stressor. One of these factors, called the “relevant feedback,” is the appropriate feedback from coping responses (55–57). For example, if the relevant feedback to a stressor (unsignaled shock) is low, stress- related symptomatology, e.g., gastric ulceration, increases, while if the relevant feedback is high (signaled shock) less symptomatology is present. Other factors involve appropriate neuroendocrine responses.

The role of neuroendocrine responses in coping with stress is well recognized, since without these responses an organism would be less likely to survive many stressful situations (60). One important feature of successful coping with stress is that physiological systems are not only turned on efficiently by a particular stressor but are also turned off again after a stressor has ceased (60, 61). Thus, when these systems (e.g., neuroendocrine systems) are not rapidly mobilized and then appropriately reduced, elevated hormone levels become dangerous for an organism, resulting in various stress-related diseases (e.g., hypertension, stroke, diabetes, obesity, autoimmune and inflammatory disorders, etc.) (60). The extent to which an individual can cope with stressful situations varies, and these differences are a product of genetics, developmental influences, experience, training, social support, and current mental and physical health (58, 60, 61).

C. Selye’s doctrine of nonspecificity revisited

The parameters of stress observed by Selye were all derived from release of ACTH, which elicits hormonally mediated responses. Stressors, however, also elicit neuronally mediated sympatho-adrenomedullary responses, which although recognized by Selye from Cannon’s work, remained unmeasured and therefore were not considered in the syndrome described by Selye (1, 17, 34). Selye described that stressors do not differ in terms of the “patterns” of stress responses. Only after stressor-defined patterns were removed from consideration could one approach the stereotyped stress syndrome. This syndrome could be graded in intensity, but the pattern of response would not be defined by the stressor.

Testing Selye’s hypothesis in our laboratory was possible only by comparing the relative magnitudes of several independent neural and hormonal responses at different intensities of stressors using a simplifying assumption: that the magnitudes of both the specific and nonspecific components vary directly with the intensity of the stressor over the whole range of stressor intensities, i.e., that there is no ceiling for the specific component, and no threshold for the nonspecific component (4). Thus, if there is a single unitary response to all stressors, then at two different intensities of the same stressor, the ratios of the increments in the responses should be the same for all parameters, regardless of the stressor. By comparing ratios of differences in response to low- and high-intensity stressors, we examined the theory of nonspecific response patterns.

We measured arterial plasma NE, epinephrine (EPI), and ACTH concentrations in conscious Sprague Dawley rats after exposure to one of five different stressors: immobilization (2 h), hemorrhage (10% or 25% of estimated blood volume; the latter producing hemorrhagic hypotension), cold exposure (4 C or −3 C), pain (evoked by subcutaneous administration of 1% or 4% formalin), or hypoglycemia (evoked by intravenous injected insulin at one of three doses: 0.1, 1.0, or 3.0 IU/kg). For each plasma measure for each animal, an area under the curve (concentration × time) was calculated.

For all three plasma measures, net total responses varied by more than 50-fold across stressors (Fig. 1). At their highest intensity, all of the stressors resulted in significant increases in levels of ACTH, NE, and EPI compared with control values obtained after intravenous saline injection. Immobilization stress evoked large increases in plasma levels of ACTH, NE, and EPI, but other stressors induced disproportionately large NE or EPI responses compared with ACTH responses (Fig. 1). Thus, whereas for ACTH, immobilization stress evoked the largest responses and cold was relatively ineffective, for NE, cold evoked the largest responses, and for EPI, insulin evoked the largest responses. The largest increment in plasma EPI levels after administration of insulin is consistent with the homeostatic effect of EPI in antagonizing the actions of insulin and promoting release of glucose from the liver. Similarly, the largest NE responses after cold are consistent with sympathetic activation to conserve heat by piloerection, vasoconstriction, and energy expenditure. No clear specific responses were found for immobilization stress, hemorrhage, and pain stress; therefore, these stressors were suitable for testing of Selye’s doctrine of nonspecificity. Immobilization stress could not be used because the intensity of this stressor could not be varied. Thus, only data for ACTH, NE, and EPI responses to hemorrhage and formalin were appropriate for testing the doctrine of nonspecificity.

Fig. 1

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Responses of plasma ACTH, NE, and EPI during exposure of conscious rats to various stressors. Each bar represents the mean value for the net area under the curve (AUC) for that stressor, where net AUC for each animal was calculated from the baseline AUC subtracted from the total AUC. SAL, Saline; INS, insulin; HEM, hemorrhage; FORM, formalin; IMMO, immobilization.

The 4% formalin concentration elicited about a 2-fold larger plasma ACTH response and about a 4-fold larger plasma EPI response than did the 1% concentration, and the differences for both variables were statistically significant. Hemorrhage evoked small NE and EPI responses relative to ACTH responses. The 25% hemorrhage elicited about a 5-fold larger ACTH response than did the 10% hemorrhage. There were no differences in plasma NE and EPI responses when the 10% and 25% hemorrhages were compared.

As shown in Fig. 2, for plasma EPI, the ratio of the response for the 25% hemorrhage to the 10% hemorrhage was smaller than the ratio of the response for 4% formalin to 1% formalin. The doctrine of nonspecificity would predict that the difference between hemorrhage and formalin would also obtain for plasma ACTH; in fact, however, for plasma ACTH, the ratio of the response for the 25% hemorrhage to the 10% hemorrhage was much larger than the ratio of 4% formalin to 1% formalin. The increment in plasma EPI levels between the two intensities of formalin was larger than the increment in plasma ACTH levels; yet the increment in plasma EPI levels between the two intensities of hemorrhage was smaller than the increment in plasma ACTH levels. Clearly, the response patterns to these two stressors are not identical.

Fig. 2

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Ratios of plasma NE, EPI, and ACTH responses to greater or lesser intensities for hemorrhage and formalin. The application of Selye’s doctrine of nonspecificity would predict that arrows should be parallel to each other and of the same length.

Our results confirmed the marked heterogeneity of neuroendocrine responses. The present results are consistent with the alternative concept that each stressor has its own neurochemical “signature,” with quantitatively, if not qualitatively, distinct central mechanisms. Considering that our studies included assessments of activities of only three peripheral stress effector systems, one would expect that measurements of activities of other systems (e.g., vasopressin, oxytocin, renin-angiotensin, parasympathetic) would yield even more clearly distinct patterns.

The central nervous system plays a crucial role in elicitation and modulation of compensatory stress response patterns. Although a large number of neurotransmitters, neuropeptides, and neuromodulators are activated in various brain regions during exposure to stress, one can predict that specific neuronal circuits exist to optimize effective, rapid, and efficient responses to restore disturbed homeostasis and ensure minimal damage to the organism. This is supported by the elegant work of Gaillet et al. (62), who suggested a differential involvement of PVN noradrenergic pathways in the regulation of the HPA axis according to the nature of the stressor. Thus, identification of such “stressor-specific” anatomical and functional circuits would be extremely important in developing future therapies for stress-related disorders. In this review, based on our studies and the work of others, we have attempted to describe stressor-specific anatomical circuits. The mapping of these stressor-specific neuroanatomical circuits is the first step to move the field of stress research in a new direction.

III. Brain Regions Involved in Neuroendocrine Responses to Stress

Although the entire central nervous system is involved in the maintenance of internal homeostasis and participates in the organization of stress responses, some areas may have specific roles in these regulatory mechanisms. They are summarized briefly in this section.

Stressful stimuli may reach the central nervous system through somato- or viscerosensory pathways through spinal or brainstem sensory neurons (Fig. 3). Somatosensory signals are detected by noxious, mechanical, thermosensitive, etc., or specific (photic, acoustic, taste, equilibral) receptors and carried by spinal and cranial sensory nerves. Viscerosensory signals arise from the body and may reach spinal and supraspinal receptors by neural (from interoceptors) or humoral pathways. (Accordingly, stressful stimuli have been previously classified as neurogenic and systemic stressors.)

Fig. 3

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Neuronal circuits in the organization of stress responses. Horizontal thick and thin lines indicate “short circuit”: autonomic (sympathoadrenal and/or parasympathetic) and defense (withdrawal) spinal reflexes in response to stressful stimuli. Thin lines represent “long circuit”: ascending (afferent) and descending (efferent) neuronal loops between the spinal cord/medulla and “higher” brain centers. Dashed line indicates neurohumoral hypothalamo-pituitary outflow. CA, Brainstem catecholaminergic neurons.

Stress responses in general can be divided into short and long circuit categories (Fig. 3). Short circuit mechanisms are also called spinal stress responses based on spinal reflexes, while the “long circuits” are also called supraspinal stress responses. The maintenance of homeostasis requires precise coordination of autonomic, neuroendocrine, and behavioral responses to contend with constant perturbations of the internal and external environments. Therefore, the long circuits include higher centers such as the neuroendocrine hypothalamus, the limbic system, and the cerebral cortex. Each of them is neuronally connected with brainstem and spinal somato- and viscerosensory centers, and they are also interconnected with each other (Fig. 3). The output system (i.e., the realization of the stress response) involves two major routes: neuronal and neuroendocrine. The neuronal responses are carried by either somatomotor or visceromotor (autonomic) fibers in cranial or peripheral nerves. Thus, both motor and autonomic stress responses finally arise from brainstem or spinal neurons. The modulatory centers (hypothalamus, limbic system, neocortex) have no direct neuronal outputs to the periphery, but they may exert their effects through actions on brainstem or spinal motor or autonomic neurons (Fig. 3). The hypothalamus has a special neuroendocrine output route, the neurohumoral hypothalamo-pituitary system, which is involved in a prominent fashion in stress responses. Although the HPA axis is the most representative and probably the most effective neuroendocrine regulatory pathway in the stress response, hypothalamo-pituitary control of other endocrine organs, as well as control of body fluid and mineral homeostasis and food intake, also constitute important regulatory circuits that are involved in the organization of responses to stressful stimuli.

A. Central autonomic system

Preganglionic neuronal cell groups constitute the output in the effector loop of stress responses. These cholinergic neurons in the medulla and the spinal cord are activated during almost all types of stress responses that influence sympathetic or parasympathetic outflows.

The parasympathetic preganglionic neurons are located in the medulla oblongata and (a minor portion) in the sacral spinal cord. In the medulla, they form distinct cell groups (dorsal motor vagal nucleus, superior and inferior salivatory nuclei). In addition, cells are arranged diffusely in the caudal part of the medulla and form a cellular arc between the dorsal vagal and ambiguus nuclei. The sympathetic preganglionic neurons form a longitudinal cell line in the thoracic (and first lumbar) spinal cord at the lateral portion of the gray matter, and this is referred to as the intermediolateral cell column (IML).

Signals to both types of preganglionic neurons arise through two strictly organized projections: short circuit (reflex) and long circuit (modulatory) neurons (Fig. 3). The short circuit afferents to the parasympathetic preganglionic neurons in the dorsal motor vagal nucleus arise from spinal or cranial sensory (both somato- and viscerosensory) neurons.

Some of the sensory signals (like respiratory) reach the preganglionic cell directly (monosynaptic reflex), but the vast majority of the inputs are relayed by sensory neurons in the nucleus of the solitary tract (NTS). The efferent fibers of the short circuit reach the ganglionic cells (in the vegetative ganglia or intramural ganglionic cells) through the vagal nerve. In addition to the inputs to the dorsal motor vagal nucleus, sensory signals ascend from the NTS to brainstem (parabrachial), hypothalamic, and limbic areas constituting the ascending loop of the long circuit (Fig. 3). In addition, noradrenergic and adrenergic neurons within and around the NTS (A2 noradrenergic and C2 adrenergic cell groups, respectively) receive stressful stimuli through sensory pathways innervating the NTS (63). The long circuit afferents to the dorsal motor vagal neurons (descending loop of the circuit) (Fig. 3) arise from the limbic, hypothalamic, and brainstem nuclei, partly directly, partly relayed by neurons in brain regions such as the lateral hypothalamus, the bed nucleus of the stria terminalis, the parabrachial nuclei, and the periaqueductal central gray (64).

The short circuit afferent fibers to the spinal sympathetic preganglionic neurons in the IML arise in dorsal root ganglion cells with a relay by dorsal horn interneurons. Preganglionic efferents leave the spinal cord through the ventral roots and terminate on sympathetic ganglionic cells located either in the peripheral sympathetic ganglia or in the innervated organs (intramural ganglionic cells). Descending fibers to the IML (long loop efferents) (Fig. 3) arise in limbic, hypothalamic, and brainstem nuclei (64).

From a functional point of view, central biogenic amine-containing neurons can be considered as a part of the central autonomic system. While biogenic amines are present in the peripheral autonomic system, central aminergic neurons represent a very specific “one-way” regulatory system. Having neuronal inputs through both somato- and viscerosensory fibers and feedback signals from hypothalamic, limbic, and cortical areas, aminergic neurons are unique projecting neurons with hundreds of axon collaterals and ten of thousands of axon terminals. All of their nerve endings terminate within the central nervous system and none of them project to the periphery. Therefore, their characteristics are briefly summarized in a separate subsection.

B. Central aminergic systems

Brain adrenergic, noradrenergic, and serotonergic neurons are involved in the central processing of stress responses. The role of dopaminergic neurons is controversial in this respect.

Brainstem catecholaminergic neurons receive direct somatosensory input from spinal cord and trigeminal sensory neurons as well as viscerosensory input from the NTS. Their activation is stressor specific: certain stressors, such as immobilization or pain-related stimuli, activate them rapidly and substantially, while others may have only minor influences.

1. Norepinephrine synthesizing neurons. Neurons in the ventrolateral and the dorsomedial medulla oblongata are the major sources of noradrenergic nerve terminals in the hypothalamus and the limbic system (65–67). In addition to these, noradrenergic cells in the locus coeruleus also contribute to the central organization of the stress response (67, 68). Lesions of brainstem catecholaminergic cell groups or their ascending fibers block or reduce stress-induced changes in the HPA axis (21).

a. The A1 noradrenergic cell group consists of the most caudal noradrenergic cells in the ventrolateral medulla. They are topographically arranged from the level of the medulla -spinal cord junction up to the level of the area postrema. Axons of these cells, comprising the ventral noradrenergic bundle, ascend to the forebrain and innervate mainly hypothalamic and limbic structures. The highest density of noradrenergic terminals is found in the parvocellular subdivision of the PVN that contains the majority of CRH-synthesizing neurons (67, 69).

b. A2 noradrenergic cells are present in the dorsomedial medulla, partly in the nucleus of the solitary tract (NTS), but a number of NE cells are dispersed into the neighboring nuclei. Ascending noradrenergic fibers from this cell group join the ventral noradrenergic bundle and participate in the noradrenergic innervation of the neuroendocrine hypothalamus (65–67, 69).

c. Locus coeruleus neurons increase their activity dramatically in response to certain stressful stimuli (70). The cerebral cortex, the cerebellum, and the basal ganglia are the major targets of these neurons, but they also participate in the noradrenergic innervation of the hypothalamus and the spinal cord (67, 71, 72). The locus coeruleus is involved 1) in the conduction of stress signals to forebrain areas, and 2) in the organization of stress responses.

ad 1) Locus coeruleus neurons receive stress signals through somato- and viscerosensory pathways via the spinoreticulothalamic tract. Noradrenergic fibers from the locus coeruleus innervate almost the entire forebrain including cortical, limbic, and hypothalamic structures.

ad 2) Noradrenergic axons from neurons located in the locus coeruleus and in the subcoeruleus area descend in the pontomedullary reticular formation and the lateral spinal funiculus and innervate the spinal cord. The spinal projection of the locus coeruleus has been demonstrated by retrograde tract tracing (73–75). Large multipolar cells in the ventral part of the locus coeruleus and the subcoeruleus area project to the spinal cord (76). Using a transneuronal viral labeling technique, these neurons have been demonstrated 3 d after injection of pseudorabies virus directly into the lateral- intermediate zones of the thoracic-lumbar spinal cord (Fig. 4).

Fig. 4

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Retrogradely labeled, pseudorabies-infected cells in the locus coeruleus (ventral portion) and the subcoerulear area 4 d after microinjection of the virus into the thoracic spinal cord (into the intermediolateral cell column and the intermediate zones). LC, Locus coeruleus; MT, mesencephalic trigeminal nucleus; RF, pontine reticular formation; SC, subcoeruleus area; IV, fourth ventricle. Bar scale: 100 μm.

d. A5 and A7 noradrenergic cell groups are located in the ventrolateral and lateral pons, respectively. Their neurons project to the spinal cord with special high terminal density to the sympathetic preganglionic neurons in the intermediolateral cell column and to the sensory projecting neurons in the dorsal horn (77, 78). A5 neurons receive direct neuronal input from the PVN (79–81).

2. Epinephrine synthesizing neurons. Adrenergic neurons are present in the middle portion of the ventrolateral medulla (between the A1 and A5 cell groups, rostrocaudally). A separate population of C1 neurons gives rise to a long ascending projection to the endocrine hypothalamus, while others project to the spinal cord to innervate sympathetic preganglionic neurons in the intermediolateral cell column (82–85). The ascending fibers join the ventral noradrenergic bundle. In addition to C1 neurons, adrenergic neurons are also present in the dorsomedial medulla (C2 cell group) just rostral to the A2 noradrenergic cell group. Axons from these adrenergic neurons also join the ventral noradrenergic bundle and participate in the adrenergic innervation of the hypothalamus and the limbic system (84–86).

3. Serotonergic neurons. Serotonergic neurons are found in the lower brainstem (raphe nuclei) and in the hypothalamic dorsomedial nucleus. The rostral raphe nuclei (dorsal, midbrain and linear raphe nuclei) project to the hypothalamus and limbic regions (87, 88), while dorsomedial serotonergic neurons participate in the innervation of the pituitary gland (89, 90). Serotonergic neurons in the raphe magnus and raphe pallidus (rostral ventromedial medulla) project to the spinal cord (91, 92). TRH-, substance P-, and serotonin-synthesizing cells in the raphe obscurus (and probably in the raphe pallidus) innervate the dorsomedial medulla, including the dorsal motor vagal nucleus and the nucleus of the solitary tract (90).

Serotonergic neurons react sensitively to certain stressful stimuli (restraint, cold, pain) as has been demonstrated by increased c-fos activation. Especially, neurons in the raphe pallidus are very sensitive to immobilization stress and formalin-induced pain (refer to Section V). Despite a large number of studies, their contribution to the organization of stress responses is still not completely understood.

C. Noncatecholaminergic brainstem neurons

1. Medulla oblongata. The ventrolateral medulla contains stress-sensitive tyrosine hydroxylase-negative neurons. They are present in the lateral reticular and peritrigeminal nuclei. The latter neurons constitute the medullary thermosensitive area and respond to cold stress by rapid c-fos activation. In the dorsomedial medulla, NTS neurons are the principal recipients of first-order vagal and glossopharyngeal afferents, which carry viscerosensory signals (baroreceptor, respiratory, gastrointestinal, taste, etc.) to the central nervous system (93). In addition to catecholaminergic neurons (A2 and C2 cell groups), the NTS contains a variety of peptidergic neurons with hypothalamic and limbic projections (94, 95). In addition to these ascending fibers (they form the ascending loop of the “long circuit”), some of the NTS neurons serve as relay neurons that transfer viscerosensory signals directly to brainstem autonomic preganglionic (“short circuit”) and catecholaminergic neurons (Fig. 3).

Cells in the ventromedial medulla (serotonin-, TRH- and substance P-containing neurons in the raphe magnus, paragigantocellular, and magnocellular reticular nuclei) project to the spinal cord, both to the dorsal horn and the intermediolateral cell column (92). They may not receive direct nociceptive signals; painful stimuli are carried from the spinal cord by the spinomesencephalic tract initially to the periaqueductal central gray. From here, enkephalin- and dynorphin-synthesizing neurons project down to the ventromedial medulla and disinhibit γ-aminobutyric acid (GABA)ergic interneurons. The activated serotonin-, TRH-, substance P-containing neurons innervate dorsal horn-inhibitory (mainly enkephalin-containing) interneurons that can block or reduce acute pain modulation (96).

2. Pons. Neurons in the parabrachial nuclei (medial, lateral, and Kölliker-Fuse nuclei) may serve important roles as intermediate stations that are modulated by both ascending and descending pathways. The lateral parabrachial nucleus is the main site for the relay of viscerosensory information from the NTS to the forebrain (97–102). The parabrachial nuclei also receive direct neuronal information from the spinal cord and the spinal trigeminal nucleus (98).

3. Midbrain. Cell columns of the periaqueductal gray matter are involved in behavioral, autonomic, and antinociceptive changes. These neurons respond to several stressful stimuli with c-fos activation. Considerable data have been accumulated regarding central gray inhibition of pain through the activation of neurons in the rostral ventromedial medulla. Neurons in the lateral and ventrolateral cell columns of the periaqueductal gray matter project to the medullary parasympathetic preganglionic neurons (ambiguus and dorsal motor vagal nuclei) as well as to viscerosensory NTS neurons (103, 104). Other midbrain structures like the colliculi and the geniculate bodies may participate in the organization of responses to specific (optic, audiogenic) stressful stimuli.

D. Thalamus

The midline and intralaminar thalamic nuclei are strongly involved, especially in mammals, in nociceptive mechanisms (refer to the section on Pain for details). Fibers of the spinoreticulothalamic tract terminate here, and nociceptive signals transfer to limbic cortical areas (cingulate, piriform, entorhinal cortex). These neurons influence behavioral responses to certain stressful stimuli. The other sensory thalamic nucleus (ventral posterior thalamic nucleus), which receives nociceptive signals through spinothalamic, trigeminothalamic, and medial lemniscus fibers, represents the subcortical relay center for the discriminative and topographic recognition of sensory signals that terminate in the somatosensory cortex.

Neurons of the midline thalamic nuclei respond to stress with rapid c-fos activation (63, 105–108). In various experimental conditions, even the mildest interventions, such as handling and control saline injections, may elicit c-fos activation in the midline thalamic nuclei. Thus, Fos positivity in these nuclei after exposure to various stressors should be assessed with caution.

E. Neuroendocrine hypothalamus

Almost all of the medial hypothalamic nuclei participate in the organization of responses to some stressors. The paraventricular, arcuate, and medial preoptic nuclei project to both the median eminence (neurohumoral output) and brainstem/spinal cord autonomic centers (neuronal output). Descending fibers may terminate on autonomic preganglionic neurons directly (64, 109–115) or they may exert their effect through brainstem (A5) catecholaminergic neurons (Fig. 5). Other nuclei, like the ventromedial, dorsomedial, perifornical, and supramamillary nuclei, contain stress- responsive neurons with mainly intrahypothalamic projections. The magnocellular neurosecretory neurons in the paraventricular, supraoptic, and accessory magnocellular nuclei are sensitive to stressors influencing body water and electrolyte homeostasis.

Fig. 5

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Stress-related hypothalamo-autonomic neuronal circuits. 1, Stressful stimuli; 2, spinal and medullary viscero- and somatosensory inputs to the hypothalamus; 3, spinal and medullary signals to the hypothalamus through brainstem catecholaminergic pathways; 4, direct hypothalamic projections to spinal and medullary autonomic centers; 5, hypothalamic projections to spinal and medullary autonomic centers through brainstem catecholaminergic neurons; 6, hypothalamic feedback to sensory neurons; 7, stress response to the periphery through the central autonomic system. 1. 8. 7., Short neuronal (periphery-spinal/brain stem-periphery) circuit in response to stressful stimuli.

Various inputs mediating stress converge upon neurons of the parvocellular subdivisions of the PVN. These neurons, which synthesize and release CRH and vasopressin (116–118), represent the origin of a final common pathway for neurohormonal regulation of ACTH-corticosterone release.

The lateral hypothalamus may be viewed as comprising a combination of several ascending and descending fibers between the medial hypothalamus, the limbic system, and the autonomic nervous system with thousands of interneurons. Almost all of the stress-conducting fibers enter the hypothalamus in this lateral area. These fibers and a high percentage of medial hypothalamic afferent and efferent fibers are relayed here.

F. Limbic system

Both cortical and subcortical limbic structures are involved in the organization of stress responses. The subcortical areas include the amygdala, septum, habenula, and related structures, while the limbic cortex consists of the hippocampal formation (hippocampus, dentate gyrus, subiculum) and entorhinal, piriform, prelimbic, intralimbic, and cingulate cortices (119, 120).

A great variety of behavioral responses to stress are organized by the limbic system. Accordingly, limbic areas receive neuronal input from brainstem and spinal viscero- and somatosensory neurons (ascending loop of the “long circuit”), and they project to brainstem and spinal autonomic preganglionic neurons (descending loop in the “long circuit”; Fig. 3). With neuronal projections to the hypothalamus, limbic areas may influence the activity of the neuroendocrine hypothalamo-pituitary system (69, 120, 121).

The central nucleus of the amygdala occupies a special position in the organization of stress responses. This nucleus, with four subdivisions, contains various types of peptidergic (CRH, somatostatin, neurotensin, enkephalin, galanin) neurons and receptors (122, 123). They receive brainstem and hypothalamic inputs and project back to these regions directly or through the bed nucleus of the stria terminalis (69, 124). The hypothalamic targets are the PVN and cells within and between the arcuate and ventromedial nuclei (Fig. 6). A significant portion of the brainstem/spinal cord projections of the central amygdala and the bed nucleus of the stria terminalis are relayed by the parabrachial nuclei (Fig. 6).

Fig. 6

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Projections from the central nucleus of the amygdala to the hypothalamus and to the central autonomic system. AN, Arcuate nucleus; HPA, hypothalamo-pituitary-adrenocortical axis; NIST, bed nucleus of the stria terminalis; VM, ventromedial nucleus.

The septum constitutes an interface between the hippocampus and the hypothalamus. Among other hippocampal neuronal connections, the supramamillary-medial septal- hippocampal afferent projections and the hippocampus-lateral septum-PVN efferent projections are frequently activated by different stressors (69, 120).

Limbic cortical regions are sensitive to stress, especially if the stressor exceeds a noxious threshold. These regions are neuronally connected (directly or through entorhinal neurons) with the hippocampus and are responsible for stress-related motivational and behavioral responses (57).

IV. Methods Used for Mapping Stressor-Specific Neuronal Circuits

Several approaches have been introduced to localize and characterize brain structures and neuronal pathways that are involved in the organization of stress responses. Among them, intracerebral microdialysis, immunohistochemistry and in situ hybridization for neurotransmitters, neuropeptides and protooncogenes, and tract-tracing techniques are very useful tools, especially when combined with experimental brain surgery (lesioning of brain nuclei, transection of neuronal pathways). Here, we discuss two powerful techniques: intracerebral microdialysis and c-fos immunohistochemistry, in detail.

A. Intracerebral microdialysis

Initially, the participation of various brain regions in stress-induced neuroendocrine responses were studied by measuring tissue concentrations of neurotransmitters or related substances. When lower tissue concentrations of a neurocompound of interest were found in a particular brain area, it was assumed that activation of that area had occurred, and the area was considered to be part of a stressor-specific anatomical and functional circuit. However, a comprehensive understanding of neuronal regulation required the development of approaches for assessing simultaneously the rate of delivery of various neurotransmitters into the synaptic cleft and the magnitude of receptor-mediated postsynaptic responses. It was assumed that a positive correlation would exist, at least in acute stress responses, between neurotransmitter release, its synaptic cleft concentrations, and the activation of an effector system. Much attention was also paid to developing in vivo methods that would be applicable in awake animals. Therefore, microdialysis as a new in vivo method for manipulating and monitoring neurotransmitter release and inactivation, as well as for evaluating receptor-mediated biochemical effects, was introduced (125–129).

The microdialysis technique makes use of a simple principle. The dialysis membrane is permeable to water and solutes below a certain molecular mass. Perfusion of the microdialysis probe with artificial cerebrospinal fluid or a solution containing a drug of interest creates a concentration gradient across the membrane, causing diffusion of substances across the membrane. In particular, catecholamines, in the extracellular fluid space, diffuse across the dialysis membrane and enter the perfusate, assays of which can then reflect extracellular fluid concentrations of the endogenous compounds. Liquid chromatographic-electrochemical assays of substances in the microdialysate require only small amounts of material, and the samples, under some circumstances, can be assayed directly because the dialysate lacks protein. Microdialysis may also be combined with other techniques such as lesions, local chemical stimulation, pharmacological interventions, and anatomical evaluations, or localized delivery of drugs to specific brain regions, enhancing the value of microdialysis as a tool in experimental neuroendocrinology, neurology, and pharmacology. Our group has applied microdialysis in several studies relating to catecholaminergic innervation of the PVN and stressor-specific NE release in the PVN and its relationship to activation of the HPA axis (24, 25, 130–132). We introduced and described the usefulness of microdialysis in small brain areas such as the central nucleus of the amygdala and the bed nucleus of the stria terminalis (133). We also introduced simultaneous measurements of extracellular fluid levels of NE and its metabolites in various brain regions to provide a comprehensive assessment of synthesis, turnover, release, metabolism, and uptake of monoaminergic neurotransmitters under basal and pathophysiological conditions (24). In all of our studies, we performed our microdialysis experiments no earlier than 20 h after a probe was inserted to minimize factors that could affect neurotransmitter release into extracellular fluid (e.g., surgical stress, acute tissue damage after probe insertion, etc.).

B. Protooncogene-“immediate early genes”- immunohistochemistry

With an ever-expanding knowledge of brain function, attempts have been made to study the activity of individual neuroendocrine cells. Immediate early genes such as c-fos, c-jun, jun D, or zif268 represent one promising avenue of research. These genes are expressed immediately in response to appropriate extracellular stimuli and then may play important roles in signal transduction and transcriptional regulation in normal cells. Thus, the expression of various immediate early genes in particular neurons is proposed to correlate with their functional activation, and their activation is followed by the production of cell-specific neuroactive substances (28, 28A, 107, 134, 135).

c-fos Has been the most frequently used immediate early gene. The pattern of brain activation in response to acute stressful stimuli is examined by using Fos immunohistochemistry (immunostaining the Fos protein product) or in situ hybridization (expression of c-fos) as markers for neuronal activity. After an appropriate stimulus, c-fos expression occurs rapidly, usually within a few minutes, with a peak response within 30 min from the time of the initiation of stress. Fos protein is detectable by immunohistochemistry somewhat later, with maximal levels at 60–90 min after the stressor. The synthesized Fos protein has a half-life of about 2 h.

Initially, it was proposed that responses of immediate early genes to different stressors were rather similar and stereotyped but an increasing number of later studies have clearly demonstrated stressor specificity of c-fos responses in different brain regions (for review, see Refs. 19, 28, 107 , and 135).

As with other methods, detection of Fos immunoreactivity has some limitations. The identification of Fos in individual neurons can be used for functional anatomical mapping with one caveat, that an absence of c-fos induction does not necessarily indicate a lack of neuronal activity (134). The absence of Fos may indicate 1) that a population of neurons does not express c-fos, 2) some other immediate early genes and their products are responsible for neuronal stimulation, 3) the signal at the cell body may be insufficient to induce c-fos expression, 4) the thresholds for c-fos induction may differ in different neurons, or 5) the activating transmitter or the second messenger needed to induce c-fos expression was not present, functioned abnormally, or was bypassed (134). In contrast, the appearance of c-fos mRNA is not necessarily followed by the production of Fos protein. Another situation in which c-fos activation is dissociated from neuronal firing is the presence of a persistent or recurring stimulus. Sustained expression of c-fos in response to long-lasting stimuli has been observed in a variety of stressful conditions. Chronic stress may cause persistent stimulation of c-fos gene expression, or neurons of certain brain nuclei or regions are activated alternatively. Whether this is due to activation of other intracellular mechanisms or to the intensity of the stimulus remains to be determined.

Recently, NGFI-B (nerve growth factor, type 1-B), also known as nur77, N10, and TISI in mice, has also been used to map neuronal activation at multiple levels of stress-related neuroendocrine circuitry (for review, see Ref. 135). Stressor-specific induction of NGFI-B and c-fos expression established NGFI-B as a useful alternative or adjunct marker to c-fos for revealing neuronal activation in the neuroendocrine hypothalamus.

V. Stressor Specificity of Central Neuroendocrine Responses

Various stressors elicit different neural and endocrine responses. We analyzed five different stressors under acute conditions by recording such parameters as plasma ACTH, corticosterone, NE, and EPI levels and extracellular NE levels in the PVN (and in some experiments in limbic nuclei) and investigated c-fos activation by Fos immunostaining, and CRH mRNA expression by in situ hybridization. Thus, we were able to collect enough information to propose the possible routes and targets of these stressors in the central nervous system. We found that stressors differ not only in their evoked responses but also in their neuronal circuits. It should be pointed out, however, that each stressor may activate several brain structures, and the proposed pathways are the most probable but not the exclusive ones.

A. Immobilization stress

Hans Selye was the first researcher who used immobilization stress, which led in rats to the manifestation of his stress syndrome, i.e., adrenal hypertrophy, gastric ulceration, and thymicolymphatic involution (136). Selye’s original restraint procedure involved tying a rat’s legs together and wrapping the rat tightly in a towel.

All types of restraint stress should be viewed as a mixture of physical and psychological stressors, including decreased body temperature and pain stress as important components of some restraint procedures. Thus, immobilization stress-induced patterns of activation of various stress effector systems result from restraint, pain stress, and changes in body temperature.

The maximal responses of the stress effector systems are usually seen within the first 30 min after the beginning of immobilization stress. The magnitude of central stress responses usually diminishes upon exposure to chronic intermittent immobilization stress, most likely reflecting habituation as well as exhaustion of stress effector systems. This is well documented by continuous basal or stress-induced decreases in NE release into extracellular fluid in the PVN, most likely due to maximal NE release that is not matched by its ongoing stimulated synthesis in rats exposed to chronic immobilization stress (132, 137).

In our studies, immobilization stress consisted of taping each rat’s limbs to a metal frame with each rat kept in a prone position. Our data are in substantial agreement with a number of previous reports of c-fos activation in brain nuclei after various types of immobilization or restraint stress.

1. c-fos Expression in the central nervous system after immobilization stress. Strong c-fos activation are observed in several brain regions 30–120 min after immobilization stress, indicating that numerous systems were influenced by this stressor (Table 1).

a. Central catecholaminergic system. Induction of c-fos expression in response to immobilization stress is evident within brainstem catecholaminergic cell groups previously shown to play a role in stress-induced activation of the HPA axis (Table 1). Two hours of immobilization stress induces a strong Fos-like immunostaining in the A1, A2, A5, A6 (locus coeruleus), and A7 noradrenergic cell groups (26, 29, 63). Almost all of the tyrosine hydroxylase-positive cells in the ventrolateral medulla and the locus coeruleus are also Fos positive. In contrast to these, only 40% of tyrosine hydroxylase-positive cells (A2 and C2 cell groups) showed Fos immunopositivity in the dorsomedial medulla (Fig. 7). The double-labeled neurons are mainly in the commissural part of the NTS. Tyrosine hydroxylase-positive dopaminergic neurons in the higher and lower brainstem all remain Fos negative after immobilization stress. These included cells in the substantia nigra and the ventral tegmental area (A8, A9, and A10 cell groups), the A11 cell group in the posterior hypothalamus, the A12 cell group in the arcuate nucleus, the A13 cell group in the zona incerta, and the A14 cell group in the preoptic and hypothalamic periventricular nuclei. Acute immobilization stress-induced Fos expression in several brainstem catecholaminergic areas and various brain nuclei has been reported by others as well (19, 28, 106, 107, 138, 139).

Fig. 7

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Fos immunoreactive cell nuclei in tyrosine hydroxylase-positive A2 and tyrosine-hydroxylase-negative neurons in the nucleus of the solitary tract after 3 h of immobilization stress in rats. Note: a number of tyrosine hydroxylase-positive neurons contain Fos-immunonegative nuclei. ST, Solitary tract. Bar scale, 50 μm.

b. Brainstem noncatecholaminergic neurons. Several tyrosine hydroxylase-negative neurons in the ventrolateral medulla, some of them in close proximity to the ventral surface of the medulla or the ventral edge of the spinal trigeminal nucleus, are also Fos positive (26, 29, 140). Immobilization stress-induced increases in Fos immunoreactivity in neurons of the spinal trigeminal, peritrigeminal, and raphe nuclei may be related to immobilization stress-induced reductions in body temperature (133, 133A). Almost all neurons in the paratrigeminal nucleus, a cell group that projects to the pontine tegmentum (141, 142), showed Fos immunoreactivity after immobilization stress. Cells in the lateral parabrachial and Kölliker-Fuse nuclei that receive direct neuronal input from the spinal cord and trigeminal nociceptive neurons, but not the NTS (139), show some Fos immunoreactivity after immobilization stress.

Immobilization stress also activates noncatecholaminergic neurons in the dorsomedial medulla (26, 28, 63, 107, 139). A fairly high percentage of tyrosine hydroxylase-negative, neuropetidergic neurons are activated in the NTS after immobilization stress (Fig. 7). These neurons project directly to the hypothalamus and the limbic system (94, 95). Whether Fos immunoreactivity in the paratrigeminal and parabrachial nuclei, including the Kölliker-Fuse nucleus, after immobilization stress is dependent on somatosensory or unconditioned aversive stimuli is unknown.

c. Thalamus. A widespread pattern of Fos immunoreactivity and c-fos mRNA expression (26, 138) are detected in the midline thalamic nuclei 60–120 min after immobilization stress. Like after noxious stimuli (26, 28, 105, 107), the central medial, paraventricular, rhomboid, anterodorsal, reuniens, and intermediodorsal thalamic nuclei, as well as the medial subdivision of the lateral habenula, exhibit marked c-fos activation in response to immobilization stress.

d. Hypothalamus. Strong, bilateral Fos immunoreactivity was found in noncatecholaminergic neurons in the hypothalamus after immobilization stress (Table 1). Thirty minutes after immobilization stress, cells in the parvocellular PVN (Fig. 8A) and in the dorsal part of the supraoptic nucleus, most likely oxytocin-containing cells, show c-fos activation. However, 2 to 3 h after immobilization stress, pronounced Fos immunoreactivity can be observed in the entire supraoptic nucleus. Thus, it is possible that alterations in salt and water balance as well as changes in blood volume could contribute to these time-related changes in vasopressin- synthesizing supraoptic neurons. Similar to our results, Miyata et al. (143) also described increases in Fos immunoreactivity in oxytocin-containing neurons in the PVN and supraoptic nucleus. Cells in the magnocellular subdivision of the PVN and vasopressin-containing cells in the supraoptic nucleus remained Fos negative.

Fig. 8

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Fos-immunoreactive cell nuclei of hypothalamic paraventricular neurons. A, After 3 h of immobilization stress; B, 60 min after a 4% formalin injection (subcutaneous) into the hind paw; C, 60 min after withdrawing 25% of the estimated blood volume (hypotensive hemorrhage), and D, 90 min after administration of 3.0 IU/kg insulin (acute hypoglycemia). m, Magnocellular paraventricular subdivision; pm, medial parvocellular paraventricular subdivision; V, third ventricle. Bar scales, 100 μm.

In the parvocellular PVN, Fos-immunopositive cells were distributed through the medial and dorsal subdivisions where the majority of CRH-immunoreactive cells are located (Refs. 19, 28, 107, 138, 144, 145 and Fig. 8A). Two hours of immobilization stress activated c-fos in CRH and glucocorticoid receptor-containing PVN neurons (146). The expression was markedly reduced by 120 min after immobilization stress (63). Dopaminergic cells in the PVN appeared to be Fos negative after immobilization stress.

In the preoptic area, dense populations of Fos-positive cells can be located in the ventral and commissural subdivisions of the medial preoptic nucleus but not in its central part or in the lateral and periventricular preoptic areas. In the anterior hypothalamus, Fos-positive cells occupied the ventral subdivision of the anterior hypothalamic nucleus and the lateral hypothalamic area. Fos-immunopositive but not tyrosine hydroxylase-positive cells were seen in the dorsomedial and arcuate nuclei. The ventromedial nucleus and nuclei in the premamillary region were devoid of Fos immunoreactivity (63).

A prominent increase in c-fos mRNA expression was detected in the medial parvocellular subdivision and to a much lesser extent in the magnocellular subdivision of the PVN, as well as in the dorsomedial, arcuate, supramamillary, and posterior hypothalamic nuclei, and in the medial preoptic area (138). c-fos mRNA levels generally peaked at 30 min after stress, and by 120 min were markedly reduced or had returned to basal levels (138).

e. Limbic system. After three hours of immobilization stress, Fos immunoreactivity was found in the lateral subdivision of the central amygdaloid nucleus and in the medial amygdaloid nucleus. CRH-immunopositive cells in the intermediate subdivision of the central nucleus of the amygdala failed to show c-fos activation in response to immobilization stress. Arnold et al. (144) and Chen and Herbert (139) did not find c-fos activation in the central nucleus of the amygdala after 15–60 min of restraint stress. Similarly, Senba and Ueyama (28) and Arnold et al. (144) did not report any changes in c-fos activation in the central nucleus of the amygdala after 15–120 min of immobilization stress. Moderate numbers of immobilization stress-induced Fos-immunopositive neurons were demonstrated in the bed nucleus of the stria terminalis, mainly in its dorsolateral subdivision, which contains CRH cell bodies. Strong c-fos activation was observed in limbic cortical areas (piriform, cingulate, and entorhinal cortices) after immobilization stress (26, 63). Marked c-fos mRNA expression was also detected in the bed nucleus of the stria terminalis, medial and central amygdaloid nuclei, the lateral septal nucleus, medial preoptic area, and the cingulate cortex (138). A marked increase in c-fos mRNA expression was detected in the bed nucleus of the stria terminalis, medial and central amygdaloid nuclei, the lateral septal nucleus, and the cingulate cortex 30 min after immobilization stress (138).

f. Cerebral cortex. Neurons in the parietal somatosensory cortex, mainly in layers 3–5, showed c-fos activation 30–120 min after immobilization stress (26, 138).

g. Activation of other immediate early genes by immobilization stress. In an extensive study by Cullinan et al. (138) c-jun and zif/268 mRNA expression was studied after exposure of rats to restraint stress (30 min in plastic cylinders).

Induction of c-jun mRNA after immobilization stress was detected in the NTS, the medial preoptic area, the dorsal premamillary nucleus, the ventral subdivision of the lateral septal nucleus, superficial layers of the parietal cortex, and the infralimbic cortex (138). Prominent expression of transcripts encoding zif/268 was noted in the parvocellular subdivision of the PVN, the anterior and lateral hypothalamic areas, and the lateral septal nucleus. Moderate increases in zif/268 mRNA expression were found in the dorsomedial nucleus, the premamillary region, the supramamillary nucleus, and the zona incerta. Relatively minor changes were detected in the brainstem. Thus, as is evident from this study, immobilization stress-induced activation of various brain regions included those regions that participate in the activation of the HPA axis as well as in its inhibition to maintain optimal glucocorticoid levels to “repair” disturbed homeostasis. In terms of HPA axis inhibition, GABA-ergic projections from the anterior hypothalamic area, the medial portions of the bed nucleus of the stria terminalis, and the retrochiasmatic area were proposed to participate in the inhibition of CRH neurons in the PVN (138, 147).

2. Expression of CRH mRNA after immobilization stress.

a. Paraventricular nucleus. We have demonstrated that 3 h of immobilization stress significantly increased CRH mRNA expression in the PVN (148). Bartanusz et al. (149) reported that acute immobilization stress increased the levels of both CRH and vasopressin mRNAs in the PVN. Similarly, Kalin and co-workers (150) found that CRH mRNA levels in the PVN were significantly elevated 1 h after cessation of 1 h of restraint. Harbuz and associates (151–153) also demonstrated significant increases in CRH mRNA expression in the PVN 4 and 8 h after exposure to 1 h of immobilization stress. Ceccatelli and Orazzo (18) also examined changes in expression of rat PVN CRH as well as TRH, enkephalin, and neurotensin mRNAs after exposure of animals to 1 h of immobilization stress. There were few significant changes in expression of these mRNAs; enkephalin and neurotensin mRNA expression increased only slightly after immobilization stress. Exposure to acute or chronic immobilization stress was also associated with significant increases in CRH receptor mRNA expression in the parvocellular subdivision of the PVN (154).

It is well established that PVN vasopressin potentiates the ability of CRH to promote ACTH release upon exposure of animals to various stressors. Immobilization stress has been shown to increase vasopressin mRNA expression and vasopressin synthesis in the parvocellular subdivision of the PVN, vasopressin accumulation in the median eminence, and its release into the hypophyseal portal system (149, 155, 156). In contrast, no changes in vasopressin mRNA expression were seen in either the posterior magnocellular subdivision of the PVN or the supraoptic nucleus in response to acute or chronic immobilization stress (149, 155, 156).

b. Amygdala. We also examined the effects of 3 h of immobilization stress on CRH mRNA expression in the central nucleus of the amygdala. Despite marked immobilization stress-induced NE release in the central nucleus of the amygdala, immobilization stress did not alter CRH mRNA expression in this nucleus (148). In contrast to our findings, Kalin et al. (150) reported that 1 h of restraint stress increased CRH mRNA expression in the whole amygdala measured 1 h after cessation of restraint stress. Similarly, Mamalaki et al. (157) demonstrated increased CRH mRNA expression in the central nucleus of the amygdala immediately after a 2-h period of immobilization stress. The discrepancy between our results and previous reports might be related to the experimental design, the severity of the stressor, or time-dependent changes in immobilization stress-induced CRH mRNA expression.

3. Expression of TRH mRNA after immobilization stress.Acute and chronic immobilization stress exerts inhibitory effects on the hypothalamo-pituitary-thyroid axis, as indicated by decreases in circulating levels of TSH and thyroid hormones in rats (158, 159). Our previous observations that hypothalamic TRH mRNA expression is decreased concomitantly during 2 h of immobilization stress in F344 male rats suggests that this inhibitory effect is in part centrally mediated (160). As investigated further by Cizza et al. (161), immobilization stress-induced decreases in TRH mRNA expression could result from stress-induced activation of the NE/CRH system. CRH either directly or via stimulation of somatostatin secretion could inhibit TRH production. In contrast to TRH, after exposure to acute immobilization stress, pituitary TSH mRNA is unchanged in adult rats (161).

Immobilization stress-induced changes in the levels of brain TRH and its receptors were demonstrated by Takayama and co-workers (162). He showed that exposure to 30–180 min of immobilization stress resulted in decreased TRH concentrations in the pons, medulla oblongata, frontal cortex, septum, and amygdala but increased levels in the spinal cord. TRH receptor binding was decreased in the hypothalamus, amygdala, and septum.

4. Concentrations of brain catecholamines after immobilization stress. Acute immobilization stress significantly reduced concentrations of both NE and dopamine (DA) in the arcuate nucleus but not in the supraoptic and ventromedial nuclei and the median eminence. Kvetnansky et al. (163) demonstrated the effects of acute (20 min) and chronic immobilization stress (150 min daily for 40 d) on NE levels in additional brain nuclei and regions in rats. Acute immobilization stress was found to decrease tissue NE levels in the PVN, supraoptic, ventromedial, dorsomedial, and arcuate nuclei. Chronic immobilization stress resulted in significant increases in tissue NE levels in the PVN, supraoptic, and dorsomedial nuclei. DA concentrations were significantly increased only in the PVN and dorsomedial nuclei. In a study conducted by Tanaka et al. (164), exposure to 15 or 30 min of immobilization stress did not result in any significant changes in rat tissue NE levels in the whole hypothalamus, amygdala, thalamus, hippocampus, cerebral cortex, pons and medulla oblongata, midbrain, and basal ganglia. In contrast, exposure to immobilization stress lasting from 60 to 120 min resulted in significant decreases in NE levels in the hypothalamus, amygdala, thalamus, hippocampus, and cerebral cortex. In the basal ganglia, tissue NE concentrations were significantly increased while the midbrain, pons, and medulla oblongata failed to show any significant changes.

Hellriegel and D’Mello (165) and Tanaka et al. (164, 166) reported time-related changes in immobilization stress- induced enhancement of noradrenergic activity in individual brain regions and their correlation with behavior of rats during immobilization stress. Soon after rats are immobilized, they exhibit struggling, vocalization, urination, and defecation, which could be attributed to activation of the hypothalamus, amygdala, and thalamus, brain regions in which NE responses to immobilization stress are first observed and which could participate in the initial coping responses to the stressor. After vigorous struggling behavior, a period of relative quiescence typically ensues, which could reflect activation of brain centers involved in memory processes (8). Using an immobilization technique that involved wrapping rats in an adhesive band, Lachuer et al. (23) demonstrated immobilization stress-induced increases in tissue levels of dihydroxyphenylacetic acid, a DA metabolite, in A1/C1, A2/C2, and A6 (locus coeruleus) catecholaminergic cell groups. Only 5 min of immobilization stress led to immediate increases in dihydroxyphenylacetic acid in all three catecholaminergic cell groups, and maximal increases of dihydroxyphenylacetic acid and plasma corticosterone levels were attained 15 min after the onset of immobilization stress. Return to baseline was achieved within 2 h. Continuous immobilization stress for up to 60 min only slightly increased dihydroxyphenylacetic acid in the A1/C1 and A6 areas while a progressive increase in dihydroxyphenylacetic acid and plasma corticosterone levels were seen over 60 min of immobilization stress. Thus, findings of region-specific immobilization stress-induced noradrenergic activation could be related to different functions mediated by noradrenergic systems in these areas during the course of immobilization stress.

Several research groups investigated the effects of immobilization stress on brain EPI levels. Tissue EPI levels were significantly decreased only after 240 min of immobilization stress in A1/C1, A2/C2, and A6 catecholaminergic cell groups (167).

5. In vivo microdialysis studies. Pacak et al. (132) were the first to demonstrate immobilization stress-induced NE increases in the PVN. In vivo microdialysis was combined with simultaneous measurement of NE and compounds related to NE turnover and synthesis. Two hours of immobilization stress produced rapid, marked, proportionately similar increases in extracellular fluid NE and dihydroxyphenylglycol levels. The responses of NE were larger, both absolutely and proportionately, in the repeatedly (120 min of immobilization stress for 7 consecutive days) stressed group compared with the singly stressed group. This was because of significantly decreased baseline levels of NE in the repeatedly stressed group. Although absolute and relative responses of dihydroxyphenylacetic acid levels did not differ significantly between the groups, baseline dihydroxyphenylacetic acid levels tended to increase with chronic repetition of immobilization stress, suggesting that tyrosine hydroxylase activity was increased. Extracellular dihydroxyphenylglycol concentration reflects two intraneuronal processes: reuptake of released NE and net leakage of NE from vesicular stores (for review, see Refs. 24 and 132). The relative contribution of these two processes to total dihydroxyphenylglycol release depends on neuronal activity. The similarity of the proportionate increases of NE and dihydroxyphenylglycol, as described above, suggests that the increment in microdialysate dihydroxyphenylglycol was from reuptake of released NE. In brain regions, extracellular dihydroxyphenylacetic acid levels reflect oxidative deamination of axoplasmic DA (for review, see Refs. 24 and 132). Increases in extracellular levels of dihydroxyphenylacetic acid therefore reflect noradrenergic as well as dopaminergic activation. Since stress-induced increases in dihydroxyphenylacetic acid are delayed compared with increases in dihydroxyphenylglycol, stress-induced changes in dihydroxyphenylacetic acid tend to support the view that the increases may reflect increased NE synthesis (132). Thus, our findings that immobilization stress increases release of NE, dihydroxyphenylglycol, and dihydroxyphenylacetic acid suggest that acute and chronic immobilization stress increased NE release, reuptake, and metabolism in the PVN. Moreover, after 6 d of repeated immobilization stress, baseline microdialysate NE concentrations were significantly lower, whereas levels of dihydroxyphenylglycol, the deaminated neuronal metabolite of NE, were unchanged. These findings indicate either that chronic intermittent stress enhances reuptake or diminishes release of NE in the PVN. Since the NE/dihydroxyphenylglycol ratio was lower in chronically immobilized rats, we favor chronic stress-induced increases in NE reuptake as an explanation of the findings.

Tanaka et al. (168) and Pacak et al. (133) found immobilization stress-induced profound increases in extracellular NE levels in the amygdala of rats. In contrast to the PVN, in the central nucleus of the amygdala, repetitive responses to immobilization stress did not differ from acute responses. However, basal NE turnover was decreased in the central nucleus of the amygdala but tended to increase in the PVN in repeatedly stressed animals, suggesting that chronic stress may differentially affect noradrenergic activity in different brain areas.

Using an in vivo microdialysis technique, we also examined the effects of 2 h of immobilization stress on NE release in the bed nucleus of the stria terminalis, a relay station for conveying information from the amygdala to the PVN, the lower brainstem and spinal cord autonomic preganglionic neurons (169–174). Microdialysate levels of NE and its intraneuronal metabolite, dihydroxyphenylglycol, were markedly increased above basal levels during the first hour of immobilization stress and decreased gradually thereafter. Levels of dihydroxyphenylacetic acid increased gradually over the entire period of immobilization stress. These findings indicate that in rats, a single bout of immobilization stress results in increased synthesis, release, and reuptake of NE within the bed nucleus of the stria terminalis.

A previously described highly positive correlation between NE release in the PVN and plasma ACTH and corticosterone levels and the existence of noradrenergic synapses on CRH cells in the PVN supported the hypothesis that NE released in the PVN could be stimulatory to CRH neurons (175, 176). A strong positive correlation between NE release in the PVN and plasma ACTH responses during immobilization stress in intact rats demonstrated in our studies (24, 25, 130) further supported a stimulatory role of NE for CRH neurons. To clarify further a role for NE in the control of the HPA axis, we assessed the effects of brainstem hemisections on immobilization stress-induced activation of the HPA axis. Lower brainstem hemisections, which disrupt ascending noradrenergic pathways from A1 and A2 areas to the PVN, decreased microdialysate baseline levels of NE and its metabolites in the ipsilateral PVN and markedly attenuated increments in these levels during immobilization stress (148). We also demonstrated that rats with similar hemisections had lower PVN CRH mRNA expression ipsilateral to the lesion and markedly blunted responses after immobilization stress, compared with values in sham-operated rats (148). Surgical or chemical lesions of the ventral noradrenergic bundle markedly decreased CRH immunoreactivity in the PVN and decreased CRH concentrations in hypophysial portal blood (177, 178).

In conclusion, our findings support the view that CRH mRNA expression in the PVN depends at least partially on ipsilaterally ascending noradrenergic pathways arising from the medulla and that increments in CRH mRNA expression in the PVN during immobilization stress depend on an intact ventral ascending noradrenergic bundle as suggested previously (153, 179–181). However, brainstem hemisections failed to alter plasma ACTH and corticosterone levels either at baseline or in response to immobilization stress. Bilateral and redundant noradrenergic innervation of the PVN as well as disruption of many other stimulatory and inhibitory pathways by “nonselective” surgical hemisection could explain the lack of attenuation of plasma ACTH and corticosterone responses during immobilization stress in lesioned animals.

Using unilateral surgical transections of the stria terminalis, the medial forebrain bundle at the rostral hypothalamic level, or the lower brainstem between the obex and the locus coeruleus, Kiss et al. (182) demonstrated the participation of different neuronal pathways in the regulation of basal and immobilization stress-induced expression of CRH and CRH receptor mRNAs in the PVN in rats. Only brainstem hemisections significantly decreased basal and 3 h of immobilization stress-induced CRH mRNA expression in the ipsilateral PVN, suggesting that CRH mRNA levels in the PVN are under tonic stimulatory influences of the lower brainstem. The ability to increase CRH mRNA expression in rats with brainstem hemisections even to a lesser extent than in sham-operated rats further suggests involvement of factors other than NE in the regulation of PVN CRH neurons. Transection of the stria terminalis and the medial forebrain bundle had no effect on basal CRH receptor mRNA levels, but lower brainstem transections decreased significantly basal CRH receptor mRNA levels ipsilateral to the knife cut. None of these lesions affected immobilization stress-induced increases in CRH receptor mRNA expression, suggesting that stress-induced changes in PVN CRH receptor mRNA are mediated either by neural inputs from brain areas other than those investigated or by humoral factors. None of these transections affected vasopressin mRNA expression in the magnocellular PVN. Oxytocin mRNA expression in the caudal magnocellular PVN was significantly decreased only after brainstem hemisections.

6. Proposed existence of neuronal circuits/pathways activated by immobilization stress. Immobilization stress evokes extremely variable endocrine, physiological, and behavioral responses by activating motor, autonomic, and HPA systems. Information from Fos immunohistochemistry after immobilization stress indicates that several neuronal cell groups are involved in this stress response.

Several somatosensory signals arise from mechano- and nociceptive receptors carried by spinal and trigeminal sensory fibers. These first-order neurons (mainly through dorsal horn or NTS interneurons) activate spinal and medullary defense (withdrawal) and autonomic reflexes (“short circuit” mechanism in stress response, Figs. 3 and 5). As a consequence of this “alarm” reaction, the viscerosensory (cardiovascular, respiratory, etc.) system as well as the sympathoadrenal system become activated.

The ascending loop of the “long neuronal circuit” involves several neuronal pathways (Fig. 9). Viscerosensory signals carried by vagal and glossopharyngeal fibers enter the brainstem to synapse in the NTS. Here, these fibers and their collaterals terminate on both the viscerosensory and the catecholaminergic (A2 cells) neurons of the NTS. The viscerosensory NTS neurons relay information bilaterally to the pontine parabrachial nuclei or directly to forebrain (hypothalamic, limbic) structures. The parabrachial nuclei, in turn, relay visceral information to the thalamus, amygdala, and the insular cortex. This pathway appears to contribute to the autonomic organization of visceral responses to stress involving cardiovascular, respiratory, and gastrointestinal activities. Direct NTS-forebrain projections may relay in the lateral hypothalamus to terminate on neuroendocrine (mainly PVN) and limbic (amygdala) neurons (Fig. 9).

Fig. 9

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Neuronal efferent (ascending) pathways to the hypothalamic paraventricular nucleus that are activated by immobilization stress. 1, First-order somatosensory neurons; 2, spinoreticular tract; 3, spinosolitary tract; 4, ventral noradrenergic bundle from A1, A2, and locus coeruleus catecholaminergic neurons; 5, ascending noncatecholaminergic pathway from the nucleus of the solitary tract; 6, spinohypothalamic tract. A1 and A2 represent A1 and A2 noradrenergic cell groups; LC, locus coeruleus.

Somatosensory signals through first-order sensory neurons enter the dorsal root in the spinal cord (terminating mainly in layers I-V and X of the dorsal horn), while signals from the head and neck regions are carried by the trigeminal, facial, glossopharyngeal, and vagal somatosensory fibers to the spinal trigeminal nucleus. From these primary sensory regions, stress-induced signals transfer through at least four major pathways (Fig. 9).

1. Nociceptive signals to the sensory ventral posterolateral and posteromedial thalamic nuclei are carried by the spinothalamic, trigeminothalamic, and medial lemniscus pathways. From here, thalamo-cortical fibers arise and terminate in the somatosensory cortex.

2. Some of the dorsal horn neurons constitute the spinoreticulothalamic tract. This multisynaptic pathway relays in various cell groups of the brainstem reticular formation to organize brainstem stress responses, and finally reach the midline and intralaminar thalamic nuclei. From here, signals run to limbic cortical areas that are involved in the organization of behavioral responses to stressful stimuli.

3. A direct spinohypothalamic pathway has also been described (183–185). Fibers arise in dorsal horn neurons and ascend without synaptic contacts in any brainstem neurons up to the lateral hypothalamus. From here, lateral hypothalamic relay neurons transfer the signals to the medial hypothalamus, mainly to the paraventricular, arcuate and ventromedial nuclei (M. Palkovits, K. Gallatz, A. Bratincsak, and Zs. Toth, unpublished observations).

4. Direct projections have been localized from spinal cord sensory neurons to brainstem catecholaminergic (A1 and A2) cells (spinoreticular pathway). Ascending fibers with bilateral terminations reach the ventrolateral medulla (186–191). Some of the fibers in the medulla project to the A2 cells in the NTS (spinosolitary fibers), which also reach viscerosensory inputs from the vagus and glossopharyngeal nerves (93, 192, 193). It should be noted that these viscerosensory inputs are relayed from the NTS to catecholaminergic and noncatecholaminergic neurons in the ventrolateral medulla (194–197). From A1 and A2 neurons, the ventral noradrenergic bundle arises, and with additional adrenergic fibers from brainstem C1-C2 neurons and some noradrenergic fibers from the subcoeruleus area, ascend to the forebrain and innervate the hypothalamus and the major parts of the limbic system (Fig. 9).

The descending loop of the immobilization stress-evoked stress pathway arises from cortical, limbic, hypothalamic, and some lower brainstem neurons (neuronal efferents), which activate motor and autonomic output systems (Fig. 3). The components of the descending loops are interconnected with each other (cortico-hypothalamic, limbic-hypothalamic, limbic-brainstem, hypothalamo-brainstem fibers) to coordinate responses to stressful stimuli.

Pathways from the hypothalamus project to the median eminence (hypothalamo-hypophysial tract) and the autonomic centers (Fig. 10). The autonomic centers contain mainly fibers that directly innervate brainstem parasympathetic preganglionic and spinal cord sympathetic preganglionic neurons (64, 109–115), and some others that innervate A5 noradrenergic neurons (113).

Fig. 10

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Neurohormonal and neuronal hypothalamic paraventricular efferent pathways activated by immobilization stress. 1, Hypothalamo-hypophysial tract; 2, descending paraventricular fibers to brainstem and spinal cord autonomic preganglionic, and A5 catecholaminergic neurons; 3, noradrenergic fibers from A5 cells to autonomic neurons; 4, parasympathetic preganglionic fibers from the dorsal vagal and ambiguus nuclei; 5, sympathetic preganglionic fibers from the intermediolateral cell column. A5 represents the A5 noradrenergic cell group; Am, ambiguus nucleus; DVN, dorsal motor nucleus of the vagus; IML, intermediolateral cell column; ME, median eminence; PVN, paraventricular nucleus; SG, suprarenal autonomic ganglia; VG, vagal ganglionic cells.

Immobilization stress also activates the somatomotor system as indicated by strong c-fos activation of brainstem and spinal cord motoneurons and in the pontine nuclei (cortico-ponto-cerebellar system). Signal transfer from the somatosensory to the somatomotor system is organized at a neocortical level, while the ventral pallidum occupies a position as an interface between the limbic and the somatomotor systems. It has been hypothesized that the ventral pallidum mediates the flow of motivationally relevant information to motor systems (198, 199).

B. Cold stress

The existence of temperature-sensitive central elements and structures that control body temperature was first reported by Bergmann (200) and Tscheschichin (201). Liebermeister (202) suggested homeostatic mechanisms in temperature control by introducing the concept of a “set-point” for body temperature. An increasing number of rather controversial experiments finally indicated that normal temperature regulation required the integrity of the hypothalamus (for review, see Refs. 203–207). Some other brain areas, such as subthalamic, midbrain, and brainstem regions, were considered to be important structures in thermoregulation as well (for review, see Ref. 208).

During the past two decades, it has been shown that the preoptic region of the hypothalamus is the major organizing center for thermoregulation. A variety of thermoregulatory responses can be elicited by activating a group of medial preoptic neurons (209–225). These neurons receive synaptic input from the periphery through spinal and medullary thermoreceptive pathways. Lipton et al. (215) attempted a more detailed mapping of cold-sensitive pathways. Rats in which the preoptic/anterior hypothalamic region was destroyed or disconnected from brainstem structures lost the ability to regulate body temperature properly in response to heat. They demonstrated that 1) communication between the preoptic and anterior hypothalamic (most probably paraventricular) nuclei is required for normal regulation against cold or heat; and 2) a major portion of pathways for regulation against cold passes through the medial forebrain bundle in the lateral hypothalamic area.

1. Cold stress-evoked c-fos expression in brain areas. In our experiments (26, 29, 108, 133A), we exposed rats to cold stress in a temperature-controlled chamber, with temperature maintained at −3 C. Animals were kept in the chamber for 3 h and killed immediately after termination of cold stress. Increased Fos immunoreactivity was found in several brain regions (Table 1). Fos-like immunoreactivity was found in the marginal zone (lamina I) of the dorsal horn, and in the ventrolateral medulla. This group of cells at the ventromedial edge of the spinal trigeminal tract (called the peritrigeminal nucleus) exhibited very strong Fos-like immunoreactivity after cold exposure. Cold stress also activated cells in the paratrigeminal nucleus (a group of cells just dorsal to the spinal trigeminal nucleus). This cell group projects to the pontine tegmentum and especially to the parabrachial nuclei (141, 142), where hundreds of cells showed Fos-like immunoreactivity in response to cold exposure. In noncatecholaminergic cells of the NTS, cold exposure induced Fos-like immunoreactivity in a fairly high number of cells but almost exclusively in the lateral part of the nucleus. Cold exposure elicited mild influences on cells in the raphe nuclei. A small group of cells in the dorsomedial portion of the pontine reticular formation, known as the “pontine thermoregulatory area” in guinea pig (226), also showed Fos-like immunostaining after cold stress. These Fos-immunoreactive cells were located just medial to the subcoeruleus area and ventral to the dorsal tegmental nucleus. Only a very few catecholaminergic neurons in the A1 but not in the A2, A5, A6, and A7 cell groups showed Fos immunoreactivity after cold exposure (up to 10–15% of the tyrosine hydroxylase-positive neurons at maximum). In the hypothalamus, strong c-fos expression was found in the preoptic area, especially in the lateral subdivision of the medial preoptic nucleus and in the ventral part of the anterior hypothalamic nucleus (Table 1). Cold exposure had a relatively minor effect on c-fos activation in the parvocellular PVN (29).

Miyata et al. (227) investigated the effects of acute and chronic cold stress (10 C) on c-fos expression in various brain regions. Fos-positive regions were classified into three groups on the basis of the expression period of Fos protein. In the first group, which included the locus coeruleus, the NTS, the dorsal tegmental nucleus, the parvocellular PVN, the posterior hypothalamic area, the supramamillary nucleus, and the lateral septal nucleus, Fos-positive cells were seen 3 h and 24 h after cold stress, but were not observed 14 d after cold stress. The second group consisted of brain areas in which Fos-like immunoreactivity was found 3 h and 1 and 14 d after cold exposure. This group included the medial preoptic nucleus, lateral preoptic area, zona incerta, paraventricular thalamic nucleus, and the lateral parabrachial nucleus. The ventromedial hypothalamic nucleus and the spinal cord were classified into the third group, where Fos immunoreactivity was higher 14 d after cold stress than after 3 h or 24 h of cold stress. In the spinal cord, Fos-positive cells were present in dorsal horn laminae II and III neurons, which participate in thermoregulatory responses such as shivering and vasoconstriction. The time-dependent appearance and disappearance of Fos protein could thus indicate the activation of additional neural pathways participating in homeostasis after exposure to chronic cold stress.

Acute heat exposure induces c-fos expression in brain areas in a fashion similar to acute cold exposure (191, 228–232). Heat activates C polymodal nociceptors and heat-sensitive A∂ high threshold nociceptors in the periphery and stimulates laminae I and II neurons in the dorsal horn of the spinal cord (191, 228–230). Among brain regions where cold evoked strong c-fos activation, the medial preoptic neurons were very sensitive to heat (231). Thermal stimuli in a noxious range evoked c-fos expression in pain-sensitive brain areas such as the midline thalamic, paraventricular, and supramamillary nuclei (229).

2. Effects of cold stress on the hypothalamo-pituitary-adrenal and other axes. Harbuz and Lightman (152) measured time- dependent responses of CRH mRNA in the PVN and POMC mRNA in the anterior pituitary after exposure to cold stress. Rats were placed in a cold room (4 C) for 1 h, after which they were returned to their vivarium room and killed 1, 2, 4, or 8 h later. Although there were small increases in levels of CRH and POMC mRNAs, especially 4 and 8 h after termination of cold stress, these increments were not significantly different from control animals. In contrast, Zoeller et al. (233) found significantly increased expression of CRH mRNA in rat PVN after exposure of rats to acute (6 h) and extended (30 h) periods of cold stress (5 C).

Hauger and co-workers (234, 235) investigated changes in CRH receptors in various brain regions of rats upon exposure to chronic cold stress (60 h at 4 C). Although plasma ACTH levels increased by 3-fold, CRH receptor concentrations in the anterior pituitary were unchanged, whereas in the neurointermediate lobe CRH receptor levels were increased. No changes in CRH receptor levels were demonstrated in the amygdala and the frontal cortex. Immunoreactive CRH content in the intermediate lobe was decreased slightly and increased in the median eminence. Thus, these data support the view that cold stress-induced ACTH release from the anterior pituitary may be at least partly mediated by other mechanisms/factors and pathways than CRH neurons.

Angulo et al. (236) studied the effect of cold stress or combined cold and isolation stress (animals housed individually after exposure to cold) on expression of vasopressin mRNA in the paraventricular and supraoptic nuclei. Cold stress (4 C) for 3 h daily for 4 consecutive days increased vasopressin mRNA levels in the posterior magnocellular PVN and had no significant effect on vasopressin mRNA levels in the supraoptic nucleus. In cold-isolated animals, vasopressin mRNA levels were further increased in the posterior magnocellular PVN but not in the supraoptic nucleus. These findings suggest that cold-induced synthesis and subsequent release of vasopressin could have further stimulatory effects on ACTH release at the anterior pituitary level during exposure to cold stress. Furthermore, isolation as a form of stress has been shown to activate the posterior pituitary which, in some clinical settings, could also contribute further to enhanced activity of the HPA axis with some behavioral consequences due to adverse effects of high cortisol levels on brain (11, 237–239).

Acute or chronic cold stress also increases TRH mRNA expression in the PVN, TRH levels in the hypothalamus, and plasma TSH levels, supporting the concept that TRH mRNA levels are reflective of TRH secretion in these neurons (233, 240–243). The effects of cold stress appear to be specific for TRH expression in the PVN, since cold stress does not affect cellular levels of TRH mRNA in other brain regions (233). In support of these findings, Arancibia et al. (244–246) demonstrated rapid cold stress-induced TRH release from the median eminence using a push-pull perfusion technique.

Previously, many investigators have reported changes in plasma and brain catecholamine levels upon exposure of experimental animals as well as humans to cold stress (247–256). In our experiments, rats with both an indwelling femoral arterial catheter and with a PVN microdialysis probe were exposed to cold stress at one of two temperatures (4 C and −3 C) (4, 130). Cold stress evoked only very small increments in extracellular PVN NE levels. In plasma, cold stress produced much larger proportionate increments in NE levels than in EPI, ACTH, or corticosterone levels (Fig. 2). These results are consistent with numerous previous reports showing cold stress-induced depletion of hypothalamic catecholamines and selective activation of the peripheral sympathoneuronal system (63). Furthermore, small increments in PVN NE levels and plasma ACTH and corticosterone levels and no correlation between cold stress-induced release of NE in the PVN and activation of the HPA axis suggest that other stress effector systems, such as the hypothalamo- pituitary-thyroid axis, may play a more important role in the maintenance of homeostasis during cold stress. Extracellular serotonin in the PVN was significantly increased in cold-exposed rats (257), suggesting that serotonergic fibers terminating in the PVN may influence TRH synthesis. Serotonin fibers terminating in the PVN arise from the raphe nuclei. Their possible role in body temperature regulation is supported by experimental studies and observations (258–260).

Other stressors, such as acute immobilization, decreased body temperature by 2–4 C, and it remained low for up to 24 h (261). In these animals, brain regions where cold stress was effective (peritrigeminal, paratrigeminal, lateral parabrachial, and medial preoptic nuclei), showed strong Fos immunoreactivity after 2 h of immobilization stress (Table 1) (26).

3. Proposed existence of neuronal circuits/pathways activated by cold stress. Cold stress produces a coordinated response through metabolic, endocrine, autonomic, and behavioral systems. Accordingly, several brain areas and pathways are involved in response to cold-evoked stress. As a working model, we distinguish between extero- and interoceptive thermal stimuli. Temperature changes in the environment activate thermoreceptors. Signals are carried by first-order spinal and cranial nerve fibers, which activate dorsal horn (191, 228) and brainstem (peri- and paratrigeminal) neurons. Thermal stimulation from lamina I and II dorsal horn neurons may reach the ventromedial medulla through the spinoreticular tract, while noxious heat signals are carried by the spinoreticulothalamic tract up to the supramamillary midline and intralaminar thalamic nuclei (229).

Signals from temperature-sensitive interoceptors may be carried by viscerosensory fibers to the NTS (c-fos activation in the lateral part of the NTS after cold stress), or influence other thermosensitive neurons through a humoral pathway (by acute changes in blood temperature, which has been observed during certain types of anesthesia or during exposure to subacute stressful stimuli for several hours). It is not yet known whether neurons in brainstem regions where c-fos activation was found in cold-stressed animals, i.e., the parabrachial nuclei and the subcoeruleus area, receive first- or second-order thermosensitive signals.

The lateral subdivision of the medial preoptic nucleus and the median preoptic nucleus are considered to be the center of thermoregulation in the forebrain (227, 230–232, 262–265). These neurons react to thermal stimuli with strong c-fos activation. Here, acute cold exposure induces c-fos activation in similar fashion as acute heat exposure. Using a combination of tract-tracing and Fos immunohistochemical techniques, projections from lower brainstem areas to the preoptic thermoregulatory cell group were demonstrated in cold-exposed rats (A. Bratincsàk and M. Palkovits, unpublished observations). This pathway (Fig. 11) may represent one of the possible ascending routes for cold-stress signals to the forebrain.

Fig. 11

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Proposed thermoregulatory neuronal pathways activated by cold stress. 1, Ascending pathway from medullary and pontine thermosensitive areas to the medial preoptic nucleus (bilateral pathway is shown only on the right side) (green); 2–6, efferent projections (bilateral pathways are shown only on the left side); 2, medial preoptic projections to the PVN (yellow); 3, hypothalamo-hypophysial tract (blue); 4, medial preoptic projections to brainstem and spinal autonomic centers (yellow); 5, Sympathetic (blue) and 6, parasympathetic preganglionic fibers (blue). Am, Ambiguus nucleus; DVN, dorsal motor nucleus of the vagus; GG, Gasserian ganglion; IML, intermediolateral cell column; LHA, lateral hypothalamic area; LP, lateral preoptic area; ME, median eminence; MP, medial preoptic nucleus; Mpl, lateral subdivision of the medial preoptic nucleus; PG, parasympathetic ganglia; PL, lateral parabrachial nucleus; PR, pontine raphe nucleus; PT, peritrigeminal nucleus; PTC, pontine thermoregulatory center; Sal, salivary glands; SCG, superior cervical ganglion; SN, superior salivatory nucleus; T, thyroid gland; V, roots of the trigeminal nerve; VII, roots of the facial nerve; X, roots of the vagal nerve.

The descending (effector) loop of the cold stress response has not yet been localized. Very little is known about the precise topography of axonal projections of preoptic neurons contributing to thermoregulation. Two proposed pathways should be considered (Fig. 11):

1. The median and medial preoptic nuclei project to the hypothalamic paraventricular nucleus (266). Paraventricular neurons may regulate heat-loss functions through neurohumoral and neuronal mechanisms (Fig. 11): a) Cold stress appears to have no effect on paraventricular CRH (152) but increases TRH mRNA expression (233, 267). Thus, the hypothalamo-pituitary-thyroid axis is one of the candidates for the effector loop of cold-evoked stress responses. b) Preoptic axons terminate on dorsal parvicellular neurons of the paraventricular nucleus, which provide the major descending projections to the sympathetic preganglionic neurons in the intermediolateral cell column of the thoracic spinal cord (266).

2. Neurons in the preoptic area have descending projections to several brain areas. Excitatory signals are conveyed to vasodilatatory neurons in the lateral hypothalamus (268). Preoptic axons terminate in the ventral tegmental area, the periaqueductal central gray, midbrain, pontine, and medullary reticular formation (265, 269). Three further medial preoptic projections have special importance in cold-evoked stress responses (Fig. 11): a) Preoptic axons innervate the rostral medullary reticular formation (269) where parasympathetic preganglionic neurons constitute the superior salivatory nucleus (270). Electrical stimulation of the lateral part of the medial preoptic nucleus produces a marked increase in secretion from the ipsilateral salivary glands. Warming of the preoptic area induces salivary secretion, which is suppressed by cooling of the skin (263). b) Medial preoptic neurons establish bilateral neuronal connections with brainstem serotoninergic raphe nuclei (269, 271), which innervate the spinal cord, including the intermediolateral sympathetic preganglionic cells. Serotonin may play a role in thermoregulation: serotonin synthesis inhibitors completely abolish stress-induced hypothermia (261). c) Preoptic projections terminate in the dorsomedial medulla (266). In conclusion, medial preoptic neurons influence brainstem and spinal autonomic neurons, which regulate the activity of the salivary, thyroid, and adrenal glands, as well as cutaneous blood vessels (Fig. 11), which are significant structures for integrated temperature regulation.

C. Insulin-induced hypoglycemia

Brain function depends mainly on glucose as the major substrate for energy production since, under normal conditions, brain tissue consumes about 50% of the total glucose produced by the liver. Upon exposure to hypoglycemia, neurophysiological and neuropsychological changes occur and develop quite rapidly. They include confusion, abnormal behavior, loss of consciousness, and seizures. Other symptoms such as hunger, sweating, tremor, restlessness, lightheadedness, chills, or sensations of warmth reflect activation of the adrenomedullary system. Together with increased activity of the HPA axis, these stress effector systems increase peripheral plasma glucose levels and delivery of glucose to brain to maintain fundamental brain processes important for an organism’s survival (272).

A central neural “glucostat” determines the neuroendocrine responses to glucoprivation, since an intracarotid injection of glucose attenuates, and ganglionic blockade or spinal transection abolishes, the adrenomedullary response to hypoglycemia (273, 274). Although lower brainstem centers can initiate adrenomedullary responses to hypoglycemia, the hypothalamus plays a critical role in hypoglycemia-induced responses of adrenomedullary, sympathoneuronal, and other systems. Cellular glucoprivation induced by local administration of 2-deoxyglucose into the ventromedial hypothalamus evokes rapid and marked increases in plasma levels of glucose, glucagon, NE, and especially EPI (275). Ventromedial hypothalamic lesions attenuate these responses substantially (276), indicating that glucoprivation of cells in the ventromedial hypothalamus triggers glucose counterregulatory responses. As was demonstrated in one of the first “classic” neuroendocrine experiments, hypothalamic paraventricular lesions result in hypoglycemia (277). Nevertheless, identification of neuroanatomical sites and pathways that participate in hypoglycemia-induced responses of various brain regions has not been available until recently, when Fos staining was introduced.

1. Hypoglycemia-induced c-fos expression in brain areas. Hypoglycemia was repeatedly found to induce Fos immunoreactivity in the lateral hypothalamus, reflecting activation of glucose-sensitive neurons located in this area (278, 279), but no Fos immunoreactivity was detected in the ventromedial nucleus, a region containing glucoreceptor neurons (278, 280, 281). Three peptidergic neuronal populations, melanin- concentrating hormone, PRL-like immunoreactive peptide, and orexin synthesizing neurons, have been identified recently in the lateral hypothalamic area and are considered to play an important role in feeding behavior and glucose homeostasis (278, 279, 282–284). A few melanin-concentrating hormone neurons were found to exhibit Fos-like immunoreactivity in contrast to about 80% of PRL-like immunoreactive peptide neurons that expressed c-fos after exposure to hypoglycemia. Orexin-containing neurons in the lateral hypothalamus are activated during hypoglycemia. Fos-like immunoreactivity was detected in approximately 33% of orexin-containing neurons in the lateral hypothalamus after a single insulin (20 IU/kg) injection (284). Since insulin receptors are not expressed in the lateral hypothalamus, the activation of these lateral hypothalamic neurons most likely resulted from either local hypoglycemia or activation of some other brain regions that project to the lateral hypothalamus (281). Indeed, it has been established that the lateral hypothalamus is connected with glucose receptor-positive neurons in the medial hypothalamus (ventromedial and dorsomedial nuclei) and other brain regions such as the septum, dentate gyrus, and other areas (285).

Niimi et al. (281) evaluated c-fos responses to insulin- induced hypoglycemia in other hypothalamic and brain regions. After administration of an insulin dose of 3 IU/kg, blood glucose fell to values less than 40 mg/dl in 3 h when Fos immunoreactivity was most prominent. Acute hypoglycemia induced c-fos expression in various hypothalamic nuclei such as the parvocellular division of the PVN, the periventricular, dorsomedial, and arcuate nuclei, and the lateral hypothalamic areas. Only a few Fos-positive cells were detected in the preoptic and anterior hypothalamic areas. Although this study did not examine the colocalization of Fos and CRH, increased Fos immunoreactivity appeared in most of the hypothalamic CRH-producing neurons.

A few cells of the magnocellular division of the PVN and supraoptic nucleus, which contain arginine vasopressin and oxytocin neurons, expressed Fos during hypoglycemia. Similarly, Griffond et al. (286) reported that a small number of oxytocin neurons (12–18%) showed Fos immunoreactivity, and these neurons were mainly found in both the parvocellular and magnocellular divisions of the PVN. In the hypothalamic periventricular nucleus, hypoglycemia-induced Fos immunoreactivity corresponded to somatostatin- producing neurons. This is consistent with previous reports in which hypoglycemia was shown to increase hypothalamic somatostatin mRNA levels and to inhibit GH secretion in rats (287). Hypoglycemia also induced increases in Fos immunoreactivity in the bed nucleus of the stria terminalis, septum, amygdala, cerebral cortex, and the thalamic paraventricular nucleus (281). It should be noted, however, that Fos immunohistochemistry is an exceptionally sensitive tool to visualize neuronal activity, and experimental conditions may influence c-fos expression nonspecifically. Transfer of animal cages or the animals in a short interval before injection or cannulation and handling of the animals without at least 5–7 d adaptation can result in activated c-fos expression in the hypothalamic, midline thalamic, and limbic nuclei.

In our experiments, we exposed rats to acute hypoglycemia using an insulin dose of 3 IU/kg to investigate changes in Fos immunoreactivity in brainstem neurons and other brain regions (26, 29). Rats were killed 90 min after a single insulin injection. Although blood glucose decreased substantially with a nadir of less than 40 mg/dl 30 min after an insulin injection, hypoglycemia had only minor effects on Fos immunoreactivity in tyrosine hydroxylase-positive cells in the A2 area (0–2 cells per section), but not in the A1 cell group or the locus coeruleus. Fos-like immunopositivity was also detected in some of the tyrosine hydroxylase-negative neurons in the dorsomedial medulla (26). In the diencephalon, Fos-like immunoreactivity was found in the zona incerta and the dorsomedial hypothalamic nucleus 90 min after administration of 3.0 IU/kg insulin. In the PVN, Fos-like immunoreactivity was found only in a few dorsal and caudal parvocellular neurons (Fig. 8D), mainly in those that project to brainstem and spinal cord (29). In our conditions, no c-fos expression was found in the limbic system in acute hypoglycemia except a well circumscribed area in the ventrolateral part of the central nucleus of the amygdala. None of the known neuropeptides, neurotransmitters, or receptors exhibit such a pattern of distribution. Similar to our results, Porter and Bokil (288) did not find increased c-fos expression in brain nuclei 60 min after intracerebroventricular, or 90 min after intravenous insulin infusions under euglycemic conditions. When blood glucose decreased, they found c-fos activation in the arcuate nucleus, the NTS, and the ventrolateral medulla, similar to our experiment (26). Marked c-fos activation was found in the hypothalamic paraventricular, arcuate, and ventromedial nuclei after intraperitoneal injections of a bolus of insulin plus glucose (288A).

Insulin-induced hypoglycemia did not activate catecholaminergic neurons in the locus coeruleus; Rusnak et al. (289) reported no changes in tyrosine hydroxylase mRNA in the locus coeruleus measured 5 h after single or repeated insulin injections (5 IU/kg), which caused a greater than 50% reduction in plasma glucose levels. As discussed later, these findings are also supported by our previous studies that did not demonstrate any hypoglycemia-induced extracellular NE increments in the PVN (130). Chemical lesions of the ascending noradrenergic bundle did not block insulin- induced activation of ACTH and corticosterone secretions (290). In contrast to insulin, single or repeated 2-deoxyglucose administration increased tyrosine hydroxylase mRNA in the locus coeruleus (289). In contrast to insulin, 2-deoxyglucose causes cellular glucodeprivation, its mechanism of action is different, and blood glucose levels are elevated after its administration. Thus, it cannot be excluded that 2-deoxyglucose acts directly to affect tyrosine hydroxylase gene expression.

2. Effects of hypoglycemia on the hypothalamo-pituitary-adrenocortical axis. In clinical practice, insulin-induced hypoglycemia is used to assess the functional integrity of the HPA axis. Delayed responses of the HPA axis to hypoglycemia suggest that these responses are mediated by neural structures that are sensitive to hypoglycemia and carry information either to CRH neurons or ACTH- or cortisol-producing cells (273, 291). Among central regulators that receive such information, CRH and vasopressin play an important role in hypoglycemia-induced activation of the HPA axis since administration of CRH antibodies reduced ACTH responses to insulin-induced hypoglycemia (27, 292).

Hypoglycemia increases CRH mRNA and c-fos mRNA expression in the PVN (293–295), CRH turnover in the median eminence (296), CRH and POMC mRNA levels in the anterior pituitary (292, 297, 298), CRH levels in the hypophysial portal and peripheral blood, and depletion of ACTH content in the anterior pituitary (293, 299, 300). Some of these findings are in contrast to the study by Plotsky (27), who found unchanged CRH concentrations but elevated vasopressin levels in portal plasma during hypoglycemia. Administration of a vasopressin V1 receptor antagonist attenuated hypoglycemia-induced plasma ACTH increments, and intracerebroventricular administration of vasopressin significantly decreased hypophyseal portal plasma concentrations of CRH (27). These results suggest that hypoglycemia-induced ACTH secretion may be mediated by dynamic changes in portal vasopressin concentrations and that CRH has a permissive role in hypoglycemia-induced plasma ACTH responses. Furthermore, these results provide support for the hypothesis of stressor-specific hypophysiotropic coding.

Hypoglycemia-induced ACTH increments seem to follow the early induction of c-fos in PVN CRH neurons (295). c-fos mRNA levels in the PVN are increased significantly 30 min after a single insulin injection of 5 IU/kg, reach a peak at 60 min, and return to baseline by 120 min (294). c-jun mRNA concentrations reach a peak at 90 min and remain elevated 120 min after an insulin injection. CRH mRNA accumulation is unchanged until 30 min after an insulin injection; however, a significant increase is observed at 60 min, reaching a plateau at 90 min (294). Chronic hypoglycemia (blood glucose below 80 mg/dl for 5 d) also provokes marked increases in c-fos and CRH mRNA expression in the PVN (295).

To detect neural pathways participating in hypoglycemia-induced activation of the HPA axis, several surgical approaches have been used. Using hypothalamic deafferentation and/or lesions of the medial basal hypothalamus, or hypophysectomy, Kárteszi et al. (291) and Ježová et al. (301) and Weidenfeld et al. (302) demonstrated that hypoglycemia mainly stimulated corticosteroid production through neural pathways to the hypothalamus from an anterolateral direction. These workers (302) investigated insulin-induced activation of the HPA axis in rats with various types of hypothalamic deafferentation and measured responses of plasma ACTH and corticosterone after injection of low (0.04 IU/100 g of body weight) or high (0.2 UI/100 g of body weight) insulin doses. The higher insulin dose reduced serum glucose by approximately 50% and elicited a marked increase in serum ACTH and corticosterone levels. The lower dose of insulin, which reduced serum glucose by approximately 30%, elicited a significant adrenocortical response only in intact rats and rats with posterior hypothalamic deafferentation. They concluded that during mild hypoglycemia, activation of neural pathways impinging upon CRH neurons from a rostral direction is crucial for responses of the HPA axis. In contrast, during severe hypoglycemia, adrenocortical responses are mediated by systemic mechanisms that act directly on the medial basal hypothalamus. In addition to these major pathways, hypoglycemia was found to stimulate corticosterone responses through pathways independent of the medial basal hypothalamus but requiring the pituitary (291, 303). This is well supported by studies of Mezey et al. (303), who reported that insulin-induced hypoglycemia stimulated ACTH secretion from stalk-sectioned rats. Thus, it appears that both peripheral (humoral) and central (neuronal) pathways are necessary for activation of the HPA axis, and the severity of hypoglycemia determines which pathways are activated (304).

Insulin-induced hypoglycemia is also a primary stimulus for the secretion of both posterior pituitary hormones, vasopressin, and oxytocin (305–307). Vasopressin synergistically enhances CRH-induced ACTH secretion (308–310). Hypoglycemia does not affect vasopressin mRNA in the whole PVN (298) but selectively activates a vasopressin-containing subset of CRH neurons in the parvocellular subdivision of the nucleus (307). Double-labeling experiments demonstrated that the number of CRH mRNA-containing cell bodies that also contained vasopressin mRNA was doubled during insulin-induced hypoglycemia (311). These data, together with increases in the turnover of vasopressin in the median eminence (296) and vasopressin levels in hypophysial portal blood (293), suggest that increases in plasma vasopressin levels during hypoglycemia are derived from activation of PVN CRH neurons.

Insulin-induced hypoglycemia is also known to inhibit the activity of the hypothalamic GnRH pulse generator (312). There is no direct inhibition of GnRH neurons since these neurons remain capable of releasing their hormone during hypoglycemia when stimulated by N-methyl-d,l-aspartate (313). Recent studies in which the suppression of GnRH neurons caused by insulin-induced hypoglycemia was blocked by a CRH antagonist or a CRH inhibitor, together with observations that CRH synapses on these neurons, suggest that CRH is an important intermediate involved in this inhibition (314–316).

Orexins are a family of recently identified neuropeptides that have been implicated in feeding behavior: food deprivation for 48 h was shown to up-regulate prepro-orexin mRNA levels in rat hypothalamus (282). They stimulate food intake after injection into the paraventricular, dorsomedial, and perifornical nuclei, or into the lateral hypothalamus (283). Orexin-synthesizing neurons in the lateral hypothalamus project to the arcuate nucleus to innervate neuropeptide Y (NPY)-containing cells, which contain leptin receptors receptors (317, 318). These findings suggest that orexins may play an important role in stress-related changes in feeding behavior.

Among several other factors, the adipocyte-derived hormone leptin (319) is strongly involved in both pathways (320–326). Insulin is thought to stimulate leptin secretion. Increased plasma leptin concentrations are found in hyperinsulinemia, and insulin-induced hypoglycemia is associated with a decrease in leptin concentrations (321, 323–325). The cocaine- and amphetamine-regulated transcript (CART) peptide could be an endogenous factor in brain mediating the effects of stress on appetite (327). Food intake increases CART mRNA expression and, in turn, CART reduces food intake significantly (323, 328). At the hypothalamic level (especially in the arcuate nucleus), this activity of CART neurons is closely associated with the action of leptin and NPY. NPY is a potent stimulator of food intake and regulates feeding behavior (318). Food deprivation increases the activity of arcuate NPY neurons that project to the PVN: Leptin decreases NPY synthesis in arcuate neurons and reduces NPY release in the PVN (326). Taken together, these findings suggest that insulin-induced hypoglycemia may influence the insulin-leptin-CART-NPY circuit. In addition to this mechanism, the existence of a lateral hypothalamic- arcuate-PVN-orexin-NPY circuit has also been proposed in the central control of food intake (329).

Selective adrenomedullary activation during hypoglycemia constitutes key evidence for the differential regulation of sympathoneuronal and adrenomedullary outflow during exposure to different stressors (4, 330). Despite profound adrenomedullary stimulation during systemic glucoprivation, microdialysate concentrations of NE in the PVN increase relatively little (130). During hypoglycemia, increments in plasma ACTH levels relative to those of NE in the PVN exceed increases during responses to other stressors, e.g., pain stress (4, 130). These and other findings led to the concept that in response to stressors requiring metabolic demands without involving marked discomfort or distress, ACTH release does not require a mechanism involving major activation of catecholaminergic pathways terminating in the PVN. In contrast, glucoprivation-induced suppression of LH secretion does appear to depend on NE release in the PVN (331).

3. Proposed existence of neuronal circuits/pathways activated by insulin-induced hypoglycemia. Insulin-evoked hypoglycemia may induce stress responses in two major pathways: 1) neuronal and humoral activation of the HPA axis, and 2) activation of the central autonomic system.

a. Activation of the HPA axis. The activation of the HPA axis in insulin-induced acute hypoglycemia has been discussed in detail in Section V.C.2. Both CRH and vasopressin are involved in rapid ACTH release in response to insulin-induced acute hypoglycemia (27, 296, 298, 300, 307, 332). Inputs to hypothalamic CRH and vasopressin neurons that are influenced by blood glucose and insulin levels may arise from the dorsomedial medulla. From the periphery, glucose-sensitive fibers run in the vagus nerve and terminate in the NTS. Insulin exerts direct effects on neurons in the area postrema and the NTS (333, 334). Indeed, hypothalamic lesion-induced hyperinsulinemia can be blocked by vagotomy (335). Although several neuronal pathways from the dorsomedial medulla to the PVN have been reported, the route of information regarding blood glucose or insulin levels from the medulla to the hypothalamus has not yet been localized. It is more likely that neither catecholamines nor serotonin are involved in this neuronal circuit: chemical lesions of ascending NE or serotonin bundles did not block insulin-induced activation of ACTH and corticosterone (290, 336). Lesions of the lateral hypothalamus did not inhibit a full response to insulin-induced hypoglycemia (302).

b. Activation of the central autonomic system. Descending hypothalamic and limbic axons terminate on medullary parasympathetic and spinal sympathetic neurons. These neurons provide preganglionic fibers to peripheral ganglionic cells that innervate the adrenal gland, as well as the pancreas. Cells in the dorsal motor vagal nucleus are the site of origin for parasympathetic innervation of insulin-producing cells in the endocrine pancreas (Fig. 12).

Fig. 12

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Proposed neuronal pathways involved in the response to experimental hypoglycemia by a single insulin injection. Hypothalamic projections to autonomic centers, and autonomic innervation of the pancreas. AN, Arcuate nucleus; CG, celiac ganglion; D, duodenum; DM, dorsomedial nucleus; DVN, dorsal motor nucleus of the vagus; GN, sensory gastric fibers; IML, intermediolateral cell column; VG, vagal ganglionic cells; X, vagus nerve.

1. From the hypothalamus, the PVN provides a major descending pathway to the intermediolateral cell column and the dorsal motor vagal nucleus of the vagus (69, 109, 110, 115, 116, 288A, 337–340) innervating, among others, neurons with pancreatic projections (103, 341). This pathway consists of a variety of peptidergic axons. At least eight possible neuropeptides (oxytocin, vasopressin, met-enkephalin, CRH, somatostatin, angiotensin II, neurotensin, TRH) have been identified in PVN-spinal/vagal neurons (67, 69, 112, 342).

2. ter Horst and Luiten (343) suggested that hypothalamic dorsomedial neurons provide a substantial direct innervation of vagal-pancreatic neurons. In our studies, immunohistochemistry revealed that insulin evokes c-fos activation in the dorsomedial nucleus (Table 1).

3. POMC neurons in the arcuate neurons project to the dorsal motor vagal and ambiguus nucleus (344) and thoracic sympathetic preganglionic neurons (320). Some of the arcuate POMC neurons contain CART peptide and are targets of leptin (320). These POMC/CART neurons project to the spinal cord innervating sympathetic preganglionic neurons. This pathway may contribute to the influence of leptin on the autonomic nervous system (320). A fall in leptin concentrations after insulin-induced hypoglycemia correlates with an increase in plasma catecholamine concentrations (321, 324, 325). Insulin itself may not have a direct action on arcuate POMC/CART neurons since an acute injection of insulin did not result in c-fos activation in the arcuate nucleus (26, 29, 288) and did not induce POMC mRNA expression (297). A direct action of insulin on arcuate neurons was unlikely since insulin receptors have not been visualized in this nucleus (318).

4. Autonomic preganglionic neurons receive direct input from two limbic forebrain regions, the central nucleus of the amygdala and the bed nucleus of the stria terminalis (119, 122, 123, 170, 341). Insulin-induced c-fos activation in a specific group of central amygdaloid neurons suggests that these cells may contribute to acute hypoglycemic stress.

D. Hemorrhage

Depending upon whether hypotensive or nonhypotensive hemorrhage occurs, various physiological responses are elicited upon activation of peripheral mechanoreceptors that are sensitive to volume and pressure alterations. Hemorrhage-induced responses include activation of four main systems, sympathoadrenal, HPA, renin-angiotensin, and hypothalamic atrial natriuretic hormone-vasopressin systems. These responses are important counter-regulatory mechanisms to maintain homeostasis during hypovolemia and may differ due to a variety of factors, including the magnitude and rate of blood loss, starvation, prehemorrhagic hypoxia, or prior exposure to another stressor (345–350).

In experimental and clinical studies, acute hypotensive hemorrhage (usually more than 20% of blood volume is lost) is an extremely dangerous situation since it may lead to shock, a situation in which practically all stress effector systems are activated. If severe hypovolemia is present for a longer time, death is inevitable. Clinical symptoms of nonhypotensive hemorrhage include a low energy level, fatigue, a feeling of being cold, dizziness, and sleepiness; clinical signs of hypotensive hemorrhage include cool or mottled extremities, increased heart rate, low blood pressure, pallor, and altered mental status, ranging from restlessness and agitation to coma. Restoration of blood volume is a high priority, although there are some reports that fluid replacement that is too rapid may be harmful since it may shut down the level of activity of the sympathoadrenal system in the critical period when survival depends on the activity of this stress effector system.

Karplus and Kreidl (351) were the first to implicate the hypothalamus in central autonomic and endocrine regulation of blood pressure. Later, the anterior hypothalamus was found to play an important role in depressor responses, whereas the posterior hypothalamus was found to be involved in pressor responses (352, 353).

Neurons situated in brainstem and pontine catecholaminergic cell groups also appear to be involved in the control of cardiovascular and neuroendocrine function upon exposure to hemorrhage (347, 354–358). Brainstem neurons relay baroreceptor information to other brain regions such as the hypothalamus, brainstem, and spinal cord. Using several experimental approaches including trans-synaptic tracing, antidromic mapping, and c-fos expression, findings point to the NTS as the principal site of termination of cardiovascular mechano- and chemoreceptor afferents (for review, see 354). Most NTS neurons activated upon exposure to hemorrhage are nonaminergic and project to the hypothalamus, the bed nucleus of the stria terminalis, the amygdala, and the ventrolateral medulla (94, 354, 359), where volume- and pressure-related information is interconnected and where regulation of neuroendocrine brain regions occurs directly or indirectly.

Pontine regions, like the locus coeruleus and the parabrachial nuclei, receive cardiovascular information most likely from the NTS via the ventrolateral medulla (for a review, see Ref. 354) and relays information to higher brain regions. Bilateral electrolytic lesions in the anterior portion of the locus coeruleus do not affect blood pressure during nonhypotensive and hypotensive (20% of total blood volume) hemorrhage (355). However, animals with lesions in the posterior part of the locus coeruleus show a greater decrease in blood pressure during the hypotensive hemorrhage. The decrease in blood pressure is similar in sham-operated and lesioned animals exposed to severe (30 or 40% of total blood volume) hemorrhage, suggesting that other brain regions and pathways are involved to compensate for the lack of pressor actions of the locus coeruleus (355).

Brainstem and pontine aminergic neurons control the activity of spinal sympathetic preganglionic neurons and are involved in the regulation of neuroendocrine systems, especially the HPA axis (85, 360–362). Posterior hypothalamic knife cuts that disrupt fibers of the medial forebrain bundle at the level of the mamillary body or 6-hydroxydopamine lesions of the dorsal, but not ventral, noradrenergic bundle significantly decrease hypotensive hemorrhage-induced responses of the sympatho-adrenomedullary system, ACTH, renin, and vasopressin (363–365).

1. Hemorrhage-induced c-fos expression in brain areas. In our laboratory (26, 29), we exposed rats to hypotensive hemorrhage by withdrawing 25% of their estimated blood volume through a femoral vein cannula. Rats were killed 60 min after hemorrhage. Hemorrhage activated Fos in a small number of catecholaminergic and noncatecholaminergic neurons in the A1 and A2 cell groups and the locus coeruleus (Table 1). In the PVN, hypotensive hemorrhage induced marked c-fos expression in the magnocellular subdivision and had only minor effects on c-fos activation in the parvocellular subdivision (Ref. 29 and Fig. 8C). Hemorrhage also had minor effects on c-fos activity in the supraoptic nucleus. Most of the neurons in the reticular formation, pontine nuclei, and raphe nuclei did not show any Fos immunoreactivity after exposure to hemorrhage (Table 1).

Several other laboratories have used immunocytochemical labeling of Fos protein to determine the effect of hemorrhage on activation of different brain regions, but especially those previously considered to be involved in cardiovascular responses to this stressor. Hypotension caused by hemorrhage or nitroprusside administration results in an increase in Fos immunoreactivity in the locus coeruleus (354, 366). More detailed studies of anatomical distribution showed that after hypotensive (20% of total blood volume) hemorrhage, increased Fos immunoreactivity was detected mainly in the posterior region of the locus coeruleus, suggesting a possible participation of this brain area in blood pressure control (355). Badoer et al. (367) examined whether neurons in the rostral ventrolateral medulla that project to the upper thoracic spinal cord and the NTS express c-fos after either nonhypotensive or hypotensive hemorrhage. In rats exposed to hypotensive hemorrhage, large numbers of Fos immunoreactive cells were detected in the ventrolateral medulla, the NTS, and the area postrema. In the ventrolateral medulla, Fos immunoreactive cells were present across all rostrocaudal levels and coincided with the region occupied by spinally projecting neurons with presumed vasomotor function (82, 83, 368, 369). In rats with nonhypotensive hemorrhage, the number of Fos-positive cell nuclei was clearly reduced and did not differ consistently from that observed in control rats. In the rostral ventrolateral medulla of rats exposed to hypotensive hemorrhage, the proportion of spinally projecting neurons expressing Fos was significantly greater compared with rats exposed to nonhypotensive hemorrhage or to controls. The proportion of retrogradely labeled cells in the rostral ventrolateral medulla expressing Fos after nonhypotensive hemorrhage did not differ significantly from controls. Chan and Sawchenko (354) calculated that after hypotensive hemorrhage, the proportion of Fos immunoreactive catecholaminergic neurons in A2 and C2 areas that projected to the PVN was 4.5% and 19.3%, respectively. In contrast, more substantial proportions of Fos immunoreactive catecholaminergic neurons in A1 (29.6%) and C1 (32.0%) areas projected to the PVN. About 14.5% and 6.9% of Fos immunoreactive catecholaminergic neurons from the locus coeruleus and A5 cell group, respectively, projected to the PVN.

Both the parvocellular and magnocellular subdivisions of the PVN also exhibit increases in Fos immunoreactivity after hypotensive or nonhypotensive hemorrhage (135, 367, 370, 371). Hypotensive hemorrhage was attended by a significant increase in Fos immunoreactivity in magnocellular neurosecretory neurons in the PVN (including the anterior, medial, and posterior subnuclei). Among the parvocellular neurons of the PVN, dense concentrations of Fos immunoreactive cells were observed in the medial subnucleus, a region known to project to the brainstem and to thoracolumbar spinal cord sympathetic preganglionic vasomotor neurons. Fos-positive cell nuclei were also found in all other parvocellular subnuclei. During nonhypotensive hemorrhage, Fos immunoreactivity increased in the magnocellular and parvocellular subdivisions of the PVN, including the ventromedial subdivision known to contain neuronal projections to brainstem autonomic centers and spinal cord sympathetic preganglionic neurons. Both the parvocellular and magnocellular subdivisions of the PVN also exhibited increased Fos immunoreactivity during isovolemic hypotension (366).

Combined immuno- and hybridization histochemical approaches were used to determine the effect of hypotensive hemorrhage on expression of NGFI-B and c-fos mRNAs and Fos immunoreactivity in the endocrine hypothalamus, especially the PVN (135). Hypotensive hemorrhage resulted in expression of NGFI-B mRNA in the dorsomedial part of the parvocellular subdivision of the PVN, a region in which CRH neurons are located. In the magnocellular subdivision of the PVN, hemorrhage-induced expression of NGFI-B mRNA was uniformly distributed across areas containing vasopressin and oxytocin neurons. Hemorrhage also increased NGFI-B mRNA expression in the supraoptic nucleus. Hemorrhage-induced changes in Fos-immunoreactivity and c-fos mRNA expression in individual paraventricular subdivisions and the supraoptic nucleus followed closely that of NGFI-B mRNA. Quantitative assessments indicated that 92–98% of all Fos immunoreactive neurons also displayed a positive hybridization signal for NGFI-B mRNA.

By combining tract-tracing techniques with Fos immunohistochemistry, Badoer et al. (367, 370) determined whether PVN neurons that project to the rostral ventrolateral medulla, the nucleus of the solitary tract, and the upper thoracic spinal cord and spinally projecting neurons of the rostral ventrolateral medulla express Fos after nonhypotensive or hypotensive hemorrhage. Retrogradely labeled neurons in the PVN that expressed Fos after hemorrhage were present in all parvocellular subnuclei. The proportion of retrogradely labeled neurons expressing Fos was highest in rats with hypotensive hemorrhage and lower, but still significantly different, from controls in rats with nonhypotensive hemorrhage. Similar to the rostral ventrolateral medulla where only 22% of spinally projecting neurons expressed Fos after hypotensive hemorrhage, 14% of spinally projecting neurons in the PVN expressed Fos after hypotensive hemorrhage. Only a small (5%) subpopulation of PVN neurons projecting to the rostral ventrolateral medulla was activated by hypotensive hemorrhage. Furthermore, Chan and Sawchenko (354) demonstrated that with the exception of the C1 cell group, only a small portion of dopamine-β-hydroxylase-positive neurons in C2, A5, and A6 and projecting to the spinal cord showed increased Fos immunoreactivity after hypotensive hemorrhage.

In a study by Krukoff et al. (372), approximately 33% and 16% of neurons projecting from the PVN to the NTS and the caudal ventrolateral medulla, respectively, showed Fos immunoreactivity after exposure to hypotensive hemorrhage. The targets of activated PVN neurons in the NTS and the caudal ventrolateral medulla were found to represent the same brainstem neurons that are activated by hemorrhage and that project to the PVN, suggesting that neurons in the brainstem and the PVN may reciprocally conduct cardiovascular-related information (373). About one-third of PVN neurons projecting to the brainstem were identified as nitric oxide-producing neurons, and nitric oxide is an important neurotransmitter in blood pressure regulation. In the study by Dun et al. (374, 375), about 60%–80% of tyrosine hydroxylase-positive and 60% of phenylethanolamine-N-methyltransferase-positive neurons projecting from the NTS and from the caudal and rostral ventrolateral medulla, and the rostral ventrolateral medulla to the PVN showed Fos immunoreactivity in response to hypotensive hemorrhage. In contrast, only a very few tyrosine hydroxylase- and phenylethanolamine-N-methyltransferase-positive neurons in the NTS displayed Fos immunoreactivity after hypotensive hemorrhage. Chan and Sawchenko (376) measured levels of mRNA for tyrosine hydroxylase in medullary catecholaminergic neurons after hypotensive hemorrhage. They reported significantly increased levels of tyrosine hydroxylase mRNA that preceded c-fos expression that were detected in Fos immunoreactive A1, A2, C1, and C2 cell groups. Maximal increases were detected within 0.5 to 1 and 2 h post hemorrhage in A1 and A2 noradrenergic cell groups and in C1 and C2 adrenergic cell groups, respectively. Tyrosine hydroxylase mRNA in non-Fos immunoreactive cells showed a rapid but transient increase (A1, A2, C1 cell groups) or did not change significantly over the course of the experiment (C2 cell group). As discussed by Chan and Sawchenko (354), hemorrhage-induced higher levels of tyrosine hydroxylase mRNA in noradrenergic compared with adrenergic cell groups could reflect the distinctive neuronal inputs of these brain regions. For example, more A2 than C2 neurons receive direct vagal sensory innervation, and this may explain prompt tyrosine hydroxylase mRNA responses to hemorrhage.

To separate blood pressure effects from those associated with changes in blood volume, Krukoff et al. (373) investigated the effects of isovolemic hypotension induced by administration of nitroprusside and hypovolemic hypotension on c-fos expression in the ventrolateral medulla and the NTS. No significant differences were found in patterns or numbers of activated Fos immunoreactive neurons in the NTS and the ventrolateral medulla after hypotensive hemorrhage or nitroprusside treatment. More than 90% and 75% of Fos immunoreactive neurons in the NTS and the ventrolateral medulla, respectively, did not project to the PVN. In the recent study by Chan and Sawchenko (354), withdrawal of 15% of total blood volume that caused hypotension provoked significant and progressive increases in the number and distribution of Fos immunoreactive nuclei and c-fos mRNA in the NTS, A1, A5, and C1 cell groups, the locus coeruleus, and the lateral parabrachial nucleus. Both Fos immunoreactivity and c-fos mRNA first appeared 0.5 h after the beginning of hemorrhage. Fos immunoreactivity peaked at 2.0–2.5 h and c-fos expression peaked at 1.0–1.5 h in the medulla and the pons. The characterization of Fos immunoreactive neurons showed that approximately 20% and 86% of Fos immunoreactive cells in the NTS and the ventrolateral medulla, respectively, displayed dopamine-β-hydroxylase immunoreactivity. In pontine regions, approximately 22% and 35% of Fos immunoreactive cells in the locus coeruleus and the A5 cell groups, respectively, were also found to be immunoreactive for dopamine-β-hydroxylase. Cells displaying both Fos and dopamine-β-hydroxylase immunoreactivity were not identified in the parabrachial nucleus. Nitroprusside-induced hypotension yielded a pattern of c-fos expression similar to that seen after hemorrhage, except in the A1 cell group, area postrema, and the subpostrema regions of the NTS, where the responses were muted or lacking. These results are consistent with the fact that these structures are involved in volume-induced neuroendocrine responses; A1 afferents regulate vasopressin secretion, and the area postrema is involved in regulation of the renin-angiotensin system (354). Moreover, several previous studies showed no effect of nitroprusside on plasma vasopressin levels (377–379). Different intensities and duration of hemorrhage could explain different results across these studies.

The finding that less than 50% of Fos positive PVN neurons project to the ventrolateral medulla, the NTS, or the thoracic spinal cord suggests that PVN neurons projecting to other brain regions may also be involved in activation of neuronal circuitry upon exposure to hemorrhage. Nevertheless, these results suggest that brainstem and spinally projecting neurons in the PVN or the rostral ventrolateral medulla are an important component of brain circuits involved in neuroendocrine and autonomic adjustments to hemorrhage.

In studies by Chan and Sawchenko (354) and Badoer et al. (380), hypotensive hemorrhage, but not nitroprusside administration, gave rise to increased Fos immunoreactivity in the area postrema and the subpostrema region. Systemic infusion of sarile, an angiotensin receptor antagonist, disrupted hemorrhage-induced c-fos expression in both regions, suggesting the presence of angiotensin binding sites in these regions. The anatomical connections between the area postrema and the NTS (381) and the finding that hemorrhage-induced c-fos expression in the NTS is attenuated by administration of sarile suggest that the area postrema participates in hemorrhage-induced neuroendocrine regulation.

Hemorrhage was also shown to induce Fos expression in several other brain regions (354, 367, 371, 375, 380). After hypotensive hemorrhage, a large number of Fos immunoreactive cells were observed in the supraoptic and dorsomedial hypothalamic nuclei, the bed nucleus of the stria terminalis, the central amygdaloid nucleus, the preoptic area, the bed nucleus of the lamina terminalis, the piriform cortex, several auditory and precerebellar nuclei, the spinal trigeminal nucleus, and the spinal intermediolateral cell column (Table 1). A few Fos-positive neurons were occasionally noted in various thalamic nuclei, tegmental nuclei, and the raphe nuclei. Double-labeling experiments showed that most of the Fos immunoreactive neurons in the supraoptic nucleus but not in the suprachiasmatic nucleus were immunopositive for vasopressin, supporting the idea that expression of Fos is stimulus-specific (375).

2. Effects of hemorrhage on the hypothalamo-pituitary-adrenocortical axis. The HPA axis is activated upon exposure to acute and chronic hemorrhage as is evident from several previous reports. CRH mRNA expression, portal plasma levels of CRH, and plasma ACTH and corticosteroids are all increased after exposure to hemorrhage (27, 304, 345–347, 356, 357, 382–385). Reduced hemorrhage-induced ACTH responses in rats with surgical or chemical lesions of the PVN, the locus coeruleus, or the NTS demonstrate that these two brain regions are essential components of the neural pathways for ACTH release after hemorrhage (358, 386, 387). Experiments with dogs demonstrate that hypotensive hemorrhage- induced increases in plasma corticosteroid levels are not always preceded by measurable increases in plasma ACTH concentrations above control, suggesting that initial (first 10–15 min) corticosteroid responses to hemorrhage may be stimulated by a factor(s) other than ACTH (356). Similarly, using in vivo microdialysis in the PVN, we found only small changes in extracellular concentrations of NE, a potent CRH stimulator, after exposure to hypotensive hemorrhage. Thus, we have proposed that other neurotransmitters, factors, or neuronal pathways conveying information into the PVN or bypassing it completely participate in hemorrhage-induced activation of the HPA axis. In response to osmotic stimuli, increased expression of CRH and CRH type 1 receptor have been observed in the magnocellular paraventricular and supraoptic nuclei (387A).

Experimentally induced hypotensive hemorrhage has been shown to increase vasopressin and oxytocin release in the PVN (388–390), portal plasma levels of vasopressin and oxytocin (382, 391, 392), and plasma vasopressin, oxytocin, PRL, renin, and interleukin-6 levels (356, 382, 388, 393–397). Hemorrhage-induced increases in c-fos expression in PVN vasopressin and oxytocin neurons have been repeatedly demonstrated (398–400). Plasma levels of vasopressin in rats with lesions of the hypothalamic supraoptic nuclei do not increase to levels attained in sham-operated rats after hypotensive hemorrhage, suggesting that the supraoptic nucleus is the primary regulatory site for vasopressin release in response to this stressor (401).

Deafferentation of the posterior hypothalamus reduces PRL responses to hemorrhage but not to immobilization stress (394). An anterolateral knife cut around the mediobasal hypothalamus eliminates both hemorrhage- and immobilization stress-induced PRL responses. In contrast, denervation of the posterior pituitary only slightly reduces PRL responses after immobilization stress but completely abolishes PRL responses to hemorrhage. These experiments support the existence of stressor-specific pathways and neuroendocrine responses (394).

Using in vivo microdialysis in freely moving rats, we have recently reported effects of nonhypotensive and hypotensive hemorrhage on NE release in the PVN (130). In response to both nonhypotensive and hypotensive hemorrhage, there were only modest increments in microdialysate NE levels. In contrast, several previous studies reported increased extracellular NE levels in various hypothalamic nuclei or regions after hemorrhage. In a study by Morris et al. (402), hypotensive hemorrhage resulted in a nonsignificant increase in extracellular NE in the rat PVN, but in the presence of desimipramine, there was a significant NE release when compared with baseline levels. Van Huysse and Bealer (403) demonstrated increased extracellular NE levels in the PVN/anterior hypothalamic area and in the dorsomedial medulla in rats after hypotensive hemorrhage and isovolemic hypotension. Thrivikraman et al. (404) have shown increased catecholaminergic activity in the cat PVN during hemorrhage but they did not determine which catecholamine was released. Using the push-pull cannula technique in conscious rats, Qualy and Westfall (405) demonstrated increased NE release in the PVN after administration of nitroprusside. Different experimental conditions, especially the fact that dialysate NE levels in several studies described above represented NE released from a number of brain sites/nuclei rather than from the PVN alone, most likely contributed to different results across these studies.

In the magnocellular subdivision of the PVN, hemorrhage is known to induce NE release, whose role is predominantly stimulatory to vasopressin neurons (406, 407) since treatment with 6-hydroxydopamine eliminates activation of vasopressin neurons upon stimulation of the A1 area (408). Increased NE turnover has also been described in the locus coeruleus (409); however, the functional significance of activation of this brain region is unclear since its hypothalamic and spinal projections do not appear to modulate the adaptive neuroendocrine and autonomic responses to hemorrhage (65). This may be partially explained by our recent finding that only a small portion of locus coeruleus noradrenergic neurons project to the PVN (131). Furthermore, elimination of baroreceptor afferent input to the brain produced by chronic lesions of the NTS does not alter vasopressin release during hypotensive hemorrhage in conscious rats (410).

As originally suggested by Blessing and Willoughby (411) and discussed in detail by Chan and Sawchenko (354), one possible pathway for regulation of paraventricular vasopressin upon exposure to hypovolemia involves stimulation of cardiac volume receptors. Their stimulation results in stimulation of vagal afferents that in turn stimulate A1 noradrenergic neurons, which project to the paraventricular magnocellular subdivision and to the supraoptic nucleus to facilitate release of vasopressin. This is further supported by studies of Makino et al. (364) and Lightman et al. (363, 365), who demonstrated decreased plasma vasopressin concentrations upon exposure of rats to hemorrhage after surgical lesions of either the medial forebrain or the dorsal noradrenergic bundle. Furthermore, Head et al. (412) and Gieroba et al. (413) found marked reductions in plasma vasopressin release after electrolytic lesions or chemical blockade of the A1 area in rabbits or rats and Day and Renaud (408) reported that stimulation of the A1 area enhanced the activity of vasopressin-secreting supraoptic neurons. The activation could be prevented by local injection of 6-hydroxydopamine, although it seems that during severe hemorrhage, vasopressin cells are activated by additional vasopressin stimulatory pathways that bypass the A1 region (414, 415).

3. Proposed existence of neuronal circuits/pathways activated by hemorrhage. Hemorrhage rapidly activates the HPA axis. Both parvicellular CRH and magnocellular vasopressin neurons in the PVN are involved in this activation. In response to acute hemorrhage, increased CRH mRNA synthesis has been demonstrated in the PVN (384) and increases in CRH concentrations in the hypophyseal portal blood (416).

Signals from peripheral baro-, osmo-, and angiotensin II receptors converge to a neuronal circuit in the hypothalamus and the preoptic area. To activate the HPA axis, at least three neuropeptides, angiotensin II, atrial natriuretic polypeptide, and vasopressin, are involved. Inputs to the preoptic-hypothalamic neuronal circuit could be 1) humoral, or 2) neuronal (Fig. 13).

Fig. 13

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Hypothalamic neuropeptidergic circuit in the organization of body fluid and mineralocorticoid homeostasis. Humoral and neuronal inputs to the hypothalamic regulatory circuit. 1, Humoral inputs to the subfornical organ; 2, humoral inputs through the area postrema; 3, viscerosensory neuronal inputs to the nucleus of the solitary tract; 4, ascending projections from the nucleus of the solitary tract to the perifornical and the paraventricular nuclei, as well as to the anteroventral third ventricle area; 5, angiotensin II-containing neurons in the subfornical organ with their projections to the medial preoptic area; 6, atrial natriuretic polypeptide-containing neurons in the organum vasculosum of the laminae terminalis and the medial preoptic area with supraoptic and paraventricular projections; 7, supraoptic and paraventricular vasopressinergic projections to the posterior pituitary; 8, angiotensin II-containing projections from the perifornical nucleus to the subfornical organ; 9, ascending noradrenergic fibers from the ventrolateral medulla to preoptic and magnocellular hypothalamic nuclei. AC, Anterior commissure; AP, area postrema; F, fornix; ME, median eminence; NTS, nucleus of the solitary tract; OC, optic chasm; OVLT, organum vasculosum of the laminae terminalis; PF, perifornical nucleus; PP, posterior pituitary lobe; SFO, subfornical organ; SO, supraoptic nucleus.

a. Humoral inputs. Extracellular dehydration caused by acute hemorrhage (for review, see Ref. 417) increases plasma angiotensin II concentrations and alters the activity of angiotensin II receptors in the subfornical organ and the area postrema. These two circumventricular organs lie outside the blood-brain barrier, and they express angiotensin II receptors at the highest level in the central nervous system (for review see Ref. 418). Both circumventricular organs are sensitive to circulating angiotensin II (354, 419), and angiotensin II as well as hemorrhage elicit Fos induction in their neurons (354, 420). From here hemorrhage-elicited humoral signals are transported further by neuronal routes: subfornical cells project to the preoptic-hypothalamic circuit while the area postrema neurons reach the NTS (Fig. 13).

b. Neuronal pathways. 1) Cells in the subfornical organ project to magnocellular supraoptic and paraventricular nuclei either directly or their signals are relayed by atrial natriuretic polypeptide-containing neurons in the organum vasculosum laminae terminalis and the medial (periventricular) preoptic nucleus (421). Transection of subfornical organ neuronal connections diminishes the pressor response to intravenous infused angiotensin II (422, 423). Hemorrhage activates vasopressin- and oxytocin-synthesizing hypothalamic neurons and leads to the release of these substances from the posterior pituitary. Unilateral lesions of the ventrolateral catecholaminergic system significantly reduced the hypotensive hemorrhage-induced activation of hypothalamic vasopressin, oxytocin, and medial parvocellular paraventricular (probably CRH) cells (363–365, 413, 424). 2) In addition to circulating angiotensin II, which may reach the NTS through the area postrema, neuronal afferents from cardiac and arterial baroreceptors terminate on NTS neurons. These receptors are activated by reduced blood pressure and blood volume caused by acute hemorrhage. 3) Information from the NTS may be transmitted to the PVN through either catecholaminergic (65–67, 85) or peptidergic neurons (94, 95). 4) Aminergic neurons in the caudal ventrolateral medulla (A1 noradrenergic and C1 adrenergic cell groups) are preferentially activated by hypovolemic input (354). These neurons receive noncatecholaminergic neuronal input from the NTS (67, 197). A1 and C1 neurons contribute to the noradrenergic and adrenergic innervation of the neuroendocrine hypothalamus. The A1 cell groups constitute the major source of ascending noradrenergic projections to vasopressinergic neurons in the hypothalamus (65, 67, 408, 425). The activity of A1 neurons is negatively related to baroreceptor input (411, 426). 5) NTS neurons innervate hypothalamic perifornical cells that project to the subfornical organ (423). This route may provide neuronal feedback to angiotensin II-containing cells in the subfornical organ, or it may synchronize baroreceptor-related humoral signals to brain angiotensin II receptors (Fig. 13).

Paraventricular, perifornical, and dorsal posterior hypothalamic neurons project to sympathetic preganglionic neurons in the thoracic spinal cord (see Section V.C.3). Through these descending systems the sympathoadrenal system can be activated and involved in hemorrhage-induced stress responses.

E. Pain stress

Subcutaneous injection of formalin is an excellent tool to investigate pain-evoked stress responses (427). A small volume (0.015–0.2 ml/100 g body weight in rats) of 1–5% formalin injected subcutaneously into the paw elevates plasma ACTH, corticosterone, and catecholamine levels immediately and induces marked CRH expression in the PVN (4, 130, 428–431). Formalin administration induces an acute phase of nociception (up to 1–5 min) followed by a second stage, termed “tonic,” which lasts for about 1 h (432). This effect is mediated by C-fiber afferents that induce excitatory responses in relevant dorsal horn neurons. Here, a non-N-methyl-d-aspartate ionotropic glutamate receptor (kainate GluR5 receptor subtype) mediates the nociceptive response to formalin (433).

Other painful stimuli, like mustard oil (228) or sc injection of capsaicin (19), are used less frequently to investigate pain-induced c-fos activation in the central nervous system.

1. c-fos Expression in the central nervous system after pain stress. Strong c-fos activation was observed in several brain regions 30–120 min after administration of formalin, indicating that numerous systems were influenced by this stressor (Table 1). The distribution patterns of Fos immunoreactive neurons in the brain and the spinal cord are quite similar in animals subjected to pain or immobilization stress:

a. Spinal cord. Pain induces neurons to express c-fos in a site-specific manner in areas of the spinal cord (Fig. 14): in the marginal layer (lamina I), the outer part of the substantia gelatinosa (lamina IIo), deep layers (laminae V and VI), lamina X, and the intermediolateral cell column (29, 63, 107, 191, 229, 430, 434–437). These regions are predominantly implicated in the reception, processing, and transmission of nociceptive information. Labeled cells extend from lumbar (L3) to sacral (S1) segments. Fos-immunopositive cell nuclei were detected only ipsilateral to the formalin injection (430).

Fig. 14

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Fos-immunoreactive cell nuclei in dorsal horn neurons (segment lumbar 4) 60 min after a 4% formalin injection (subcutaneous) into the hind paw. Fos-immunoreactive cells are concentrated mainly in the marginal layer (lamina I and Iio) and nociceptive neurons in laminae IV, V, VI, and X, ipsilateral to the site of the injection. DC, Dorsal columns; LF, lateral funiculus. The central canal is indicated by an arrow. Bar scale: 200μ m.

b. Central catecholaminergic and other systems. In the medulla oblongata 60 min after a unilateral formalin injection, strong bilateral labeling was seen in the ventromedial medulla (Table 1). Here, both epinephrine-containing C1 and norepinephrine-containing A1 neurons express c-fos in response to formalin, similar to the pattern after immobilization (19, 26, 63, 107). A fairly high number of neurons were Fos labeled in the NTS, almost all of them in the locus coeruleus and the lateral parabrachial nucleus bilaterally (26, 63, 107, 229). Formalin exerted effects on Fos activity in other pontine noradrenergic neurons (A5 and A7 cell groups), also bilaterally (26). The bilateral c-fos responses in brainstem and forebrain nuclei after a unilateral injection of formalin into the fore- or hindlegs clearly indicates a cross-over of pain-conducting fibers below or at the level of the medulla oblongata. This condition is supported by other observations: unilateral surgical transection at the medullary-spinal cord junction did not influence formalin-induced plasma corticosterone and ACTH levels or release of NE in both sides of the PVN (430). In hemisected rats, a formalin injection elicited bilateral c-fos expression in the PVN and thalamic nuclei in a pattern that was similar to sham-operated rats (430). Similar to a formalin injection, in response to footshock, Fos immunoreactivity was strongly induced in noradrenergic neurons of all brainstem catecholaminergic cell groups, as well as in central gray neurons and serotoninergic neurons in the dorsal raphe nucleus (438).

Formalin also activated neurons in the periaqueductal central gray (107). Endogenous opioid neurons of this region are implicated in pain modulation (96, 439, 440). These neurons receive nociceptive inputs from the spinal cord through the spinomesencephalic tract (441, 442) and exert their antinociceptive action through rostral ventromedial medullary neurons from where facilitatory signals descend to the spinal serotoninergic system (443). These medial ventromedial medullary neurons are activated (through GABAergic disinhibition) by enkephalin and dynorphin neurons of the periaqueductal central gray. Some of the neurons in the reticular formation and contribute to persistent pain by enhanced release of substance P from primary afferent nociceptive fibers in the spinal cord (444).

After a formalin injection in the midline and intralaminar thalamic nuclei, marked expression of c-fos is induced (Fig. 15) where the spinoreticulothalamic fibers terminate (26, 28, 63, 229). This area includes the central medial, thalamic paraventricular, rhomboid, anterodorsal, mediodorsal, intermediodorsal, and reuniens thalamic nuclei and the medial subdivision of the lateral habenular nucleus. These sensory-discriminative pain fibers carried by the spinothalamic and trigeminothalamic tracts terminate in the ventral posterolateral and posteromedial thalamic nuclei. These nuclei showed c-fos activation on the contralateral side after a formalin injection (105).

Fig. 15

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Fos-immunoreactive cell nuclei of midline and intralaminar thalamic nuclei 60 min after a 4% formalin injection (sc) into the hind paw. A - coronal section through the anterior (1.8 mm caudal to the level of the bregma) part of the dorsal thalamus. AM, Anteromedial thalamic nucleus; CM, centromedian thalamic nucleus; IA, interanteromedian thalamic nucleus; MD, mediodorsal thalamic nucleus; PC, paracentral thalamic nucleus; PV, paraventricular thalamic nucleus. Bar scale, 200 μm.

Painful stimuli resulted in marked c-fos expression in the hypothalamus, particularly in the medial parvicellular subdivision of the PVN (Fig. 8B), overlapping an area where CRH-immunoreactive neurons are highly concentrated (26, 63, 107, 229). Neurons in the lateral hypothalamus, mainly in the middle part of the hypothalamus and in the medial preoptic and the supramamillary nuclei, were also activated by formalin injections similar to previous reports (107).

The amygdala has been considered to play a role in stress-related changes in pain sensitivity. It may exert an antinociceptive action through the central gray-medial ventromedullary system (445). Among the amygdaloid nuclei, formalin elicited c-fos expression in the medial and central amygdaloid nuclei. Like formalin-evoked pain, foot shock also induced c-fos expression in the amygdala (106).

Noxious stimulation of the cornea in rats produced a characteristic distribution of Fos in the spinal trigeminal nucleus (446).

2. In vivo microdialysis studies during painful stimuli. Using an in vivo microdialysis technique, we also examined the effects of 4% formalin administration into the hind paw on NE release in the PVN. Subcutaneous injection of 4% formalin caused marked increases in extracellular PVN NE, dihydroxyphenylglycol, and dihydroxyphenylacetic acid concentrations (4, 130). Formalin-induced increases in PVN NE concentrations reached a peak 30 min after injection. Spino-medullary hemisection did not affect baseline and formalin-induced increases in PVN NE concentrations (430). Ponto-medullary hemisections did not affect formalin-induced immediate and marked NE release in the PVN contralateral to the knife cut (430). Similar to intact rats, 5- to 8-fold increases in NE levels were measured independently of whether formalin was injected into the right or the left leg. The hemisection attenuated formalin-induced elevations of NE in the PVN ipsilateral to the knife cut, regardless of whether formalin was injected into the left or the right leg. In these animals, PVN NE concentrations were twice as low as those measured in the contralateral PVN, but twice as high as in rats with subcutaneous saline injection. Our data further indicate that medullary NE afferents that project to the PVN ascend mainly on the ipsilateral side. As indicated by Fos expression studies, the main portion of formalin-induced NE release from the PVN is derived from the A1 catecholaminergic cell group and a smaller portion is derived from catecholaminergic inputs from the locus coeruleus. By simultaneous measurements of plasma ACTH concentrations after formalin injection, we demonstrated a positive correlation between NE release in the PVN and plasma ACTH responses (130), suggesting that PVN NE is stimulatory to PVN CRH neurons during formalin-induced pain.

3. Proposed existence of neuronal circuits/pathways activated by pain. There appear to be numerous neuronal pathways leading to the hypothalamus, mainly to the PVN and the median eminence, which convey pain-related signals. The nociceptive signals may reach the hypothalamo-pituitary system directly through the spinohypothalamic tract, or through activation of brainstem catecholaminergic and noncatecholaminergic neurons (Fig. 16).

Fig. 16

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Spinoreticulothalamic tract. Ascending nociceptive pathways to the hypothalamus and the limbic system. 1, Projections to brainstem catecholaminergic and reticular formation neurons; 2, nerve terminals in the intralaminar and midline thalamic nuclei; 3, thalamic projections to the cingulate and piriform cortex. A1, A2, A1 and A2 noradrenergic cell groups; CC, cingulate cortex; CN, central amygdaloid nucleus; CP, piriform cortex; GR, gigantocellular reticular nucleus; LC, locus coeruleus, LHA, lateral hypothalamic area; PL, lateral parabrachial nucleus.

First-order sensory neurons (neuronal perikarya in dorsal root and Gasserian ganglia) terminate in various cell types (directly on projection neurons or excitatory interneurons) in the dorsal horn of the spinal cord and the sensory trigeminal nucleus. Second (and/or third) order nociceptive fibers run in the spinothalamic, spinoreticulothalamic, spinoreticular, and spinohypothalamic tracts. The spinothalamic tract, which is a phylogenetically younger structure and well developed in humans, projects from the spinal cord directly to the ventral posterolateral thalamic nucleus, from where it projects to sensory-discriminative cortical areas. The spinoreticulothalamic tract is an older structure (it is also called the paleospinothalamic tract or medial spinothalamic tract), a multisynaptic system, which terminates in several neurons in the reticular formation and finally reaches the midline and intralaminar thalamic nuclei. The final destination of these projections is the motivational effective (limbic) cortical areas (cingulate, piriform, and entorhinal cortex). The spinohypothalamic pathway (183, 185) conducts direct signals up to the lateral hypothalamus from where they are relayed to the PVN.

Visceral pain signals are mainly carried by vagal and glossopharyngeal nerves to the NTS. From here, this information is relayed bilaterally to the parabrachial nucleus and further up to the thalamus. The pathway terminates in the viscerosensory cortical areas (prelimbic, intralimbic, and anterior cingulate cortex).

Directly or indirectly, pain-related ascending signals reach the PVN, and they are crucial to the organization of stress responses. Complete unilateral ponto-medullary brainstem surgical cuts reduced CRH immunoreactivity (64, 131, 177, 178, 447–449) and the expression of CRH mRNA in the PVN (148, 182).

Some nociceptive signals are relayed via brainstem catecholaminergic neurons and are conveyed by the ascending ventral noradrenergic bundle to the hypothalamus. Lesions of this bundle have a blocking effect on responses to stressful, including noxious, stimuli (62). The majority of noradrenergic fibers ascend on the same side and innervate the ipsilateral PVN (64). Catecholaminergic input to the PVN is known to have a significant effect on stress-induced CRH expression and release (24).

Spinal projections from nociceptive neurons to the ventrolateral medulla, including the A1 noradrenergic and C1 adrenergic cell groups, have been demonstrated with tract-tracing techniques (187, 188, 190, 191). All segments of the spinal cord contribute to the projection to the ventrolateral medulla. The ventrolateral medullary neurons (including A1 and C1 cells) are involved in central processing of interoceptive information. They receive visceral sensory input gated initially through the NTS (195).

Retrograde transport studies have also shown that a considerable number of spinal cord neurons, including nociceptive neurons in the superficial layers of the dorsal horn, project to the NTS (93). In addition to these, the NTS, which includes the A2 noradrenergic and C2 adrenergic cell groups, also receives vagal and glossopharyngeal viscerosensory inputs (450). NTS neurons project to the ventrolateral medulla (65, 67, 77, 93, 197), to the locus coeruleus (67, 451), and to diverse forebrain regions including the PVN (65, 67, 77, 95, 197). Signals may reach the hypothalamus and the amygdala directly or through the parabrachial nuclei (97). Several noncatecholaminergic neurons in the NTS receive nociceptive signals from spinal cord neurons, as has been demonstrated by formalin-induced c-fos expression in tyrosine hydroxylase-negative NTS cells (29, 452). Axons primarily from peptidergic neurons terminate in the forebrain, including the PVN (90, 94, 95).

Nociceptive fibers (or their axon collaterals) may reach the locus coeruleus through the spinoreticulothalamic tract. Locus coeruleus neurons increase their activity in response to painful stimuli (70). As a sign of the activation, marked c-fos expression has been demonstrated in locus coeruleus neurons after formalin injection (26, 107, 453), or after other painful stimuli such as footshock (438). Tyrosine hydroxylase-positive cells in the locus coeruleus project to the PVN (67, 454, 455). In addition, CRH may act as a neurotransmitter to activate noradrenergic neurons in the locus coeruleus (456–458). CRH content in the locus coeruleus is increased after acute or chronic pain stress (456).

The parabrachial nuclei appear to function as a secondary integration site for autonomic regulation (cardiovascular, respiratory, gustatory) and also for nociception. Fibers from the superficial laminae of the lumbar dorsal horn and from the spinal trigeminal nucleus ascend to the parabrachial nuclei along with the ventral spinocerebellar tract (98). The parabrachial nuclei are also the main sites for the relay of visceral information from the NTS to the forebrain.

Primary afferent fibers arising through the trigeminal, facial, glossopharyngeal, and vagal fibers are collected in the spinal trigeminal tract and terminate along the rostro-caudal axis of the spinal trigeminal nucleus. From here, secondary sensory neurons join the spinothalamic tract (lateral trigeminothalamic tract) and pass through the lateral hypothalamus on their way to the thalamic nuclei. Reportedly, nociceptive signals carried by the trigeminal nerve and the tract may reach the lateral hypothalamus (459).

The physiological response to painful stimuli is conducted by two major systems (Fig. 16): 1) activation of the HPA axis, and 2) activation of the central autonomic system. The role of PVN neurons is significant in both pathways. Here, CRH is one of the most potent neuropeptides in the stress response. It acts as a neurohormone in the HPA axis and mediates transmission through neuronal pathways that regulate autonomic outflow and visceral activity. The topography of the fibers in the formalin-evoked stress response is similar to the pattern that has been described for immobilization stress.

VI. Stressor-Specific Activation of Other Neuroendocrine Systems

Many other neuroendocrine axes are also activated upon exposure to acute or chronic stressors (7, 53, 460, 461). In terms of the hypothalamic-gonadal axis, exposure of animals and humans to acute stress is associated with a small and often short-lived increase in plasma LH and androgens (462, 463). Acute stress-induced increases in androgen appear to be caused by an alteration in plasma volume and decrease in androgen metabolic clearance, rather than as a result of a true alteration of androgen release (462). The mechanisms by which LH is increased under various stress conditions are not clear, but one possibility is that high levels of ACTH may stimulate GnRH neurons (462). Stress-induced responses are also dependent on estrogen levels; thus, animals in proestrus or gonadectomized animals do not show LH rise upon exposure to acute stress, but instead a quick decrease can be found (460). Biogenic amines such as NE and serotonin and IL-1α also participate in the regulation of the hypothalamic-pituitary-gonadal axis during acute stress reaction and most likely exert inhibitory effect on GnRH neurons or LH (462). FSH may follow the same stress response pattern as LH, but often the changes are very small or do not even exist. In contrast, upon exposure to chronic stress, reproductive functions decrease. Such an inhibition occurs at all levels including the hypothalamus (inhibition of GnRH secretion), the anterior pituitary (inhibition of LH release), and gonads (sex hormones secretion). Stress-induced activation of the hypothalamic-pituitary-adrenocortical axis is one of the main factors contributing to this inhibition since CRH and glucocorticoids are known to inhibit the GnRH, and glucocorticoids themselves exert inhibitory effects at the level of anterior pituitary and gonads (462).

It is also well documented that acute exposure to various stressors (e.g., hypoglycemia, exercise, pain) increases secretion of GH from the pituitary gland. In contrast, other stressors such as cold, handling, hypertonic glucose, and electric shock produce marked decrease in plasma GH levels (464). Acute glucocorticoid administration also increases GH concentrations in plasma (6). Thus, it appears that acute stress responses that are associated with acute plasma glucocorticoid elevation may increase plasma GH levels through stimulation of the GH gene through glucocorticoid-responsive elements in its promoter region (464). In contrast, chronic stress inhibits GH release from the anterior pituitary gland and increased resistance of peripheral tissues to insulin-like growth factors (e.g., IGF-I). Such inhibition may occur via CRH-induced increase in somatostatin secretion. Thus, dwarfism that occurs in the setting of severe chronic psychosocial deprivation may reflect chronically inhibited growth axis (for review, see Ref. 6). How catecholamines and other neurotransmitters contribute to stress-induced activation or inhibition of GH secretion is unclear.

In terms of the hypothalamic-pituitary-thyroid axis, acute (most often) stressors with high intensity induce inhibition of the thyroid axis. This is reflected by the decrease in release/production of hypothalamic TRH, decreased production/secretion of thyroid-stimulating hormone, and by inhibition of conversion of T4 to more biological active T3 in peripheral tissues (for review, see Ref. 6). CRH, somatostatin, and cytokines (e.g., IL-1 and IL-6) most likely contribute to acute stress-induced inhibition of the thyroid axis. The exception is the exposure to acute cold stress, forced swimming, and noise stress when profound thyroid axis activation occurs (12). Mechanisms by which cold-induced activation of TRH occurs are not fully understood, although catecholamines are thought to participate in this interaction.

PRL can be viewed as another anterior pituitary hormone that is increased during exposure to various stressors such as exercise, hypoglycemia, hemorrhage, immobilization, and pain stress (6, 7, 460, 461). The mechanisms by which stress increased PRL secretion remain unknown. The physiological consequences of PRL hypersecretion during stress are not clear, although it is considered that PRL may enhance immune function.

VII. Clinical Relevance of Stressor Specificity and Future Perspectives

At the present time, it is well recognized that stress plays an important role in the pathogenesis of chronic diseases. Recent studies by Chrousos and others (6) suggest that chronic diseases such as affective disorders, panic disorders, anxiety, anorexia, chronic inflammatory disorders, and gastrointestinal disturbances, among others, reflect dysregulation of stress responses and are expressed either as hyperfunctional or as hypofunctional states. In melancholic depression, hyperarousal, anxiety, hypervigilance, and insomnia reflect maladaptive generalized stress responses, including tachyphylaxis of the mesocorticolimbic system and hypersecretion of CRH in brain (6). Females exposed to chronic stress may present with delayed puberty, anovulatory cycles, and spontaneous abortions, and their infants have increased mortality (465). In males, chronic stress induces inhibition of testosterone secretion associated with abnormal spermatogenesis and decreased libido (466). Psychosocial dwarfism is another example of the effects of chronic emotional stress (467). Chronic fatigue syndrome, increased vulnerability to infection, and tumor development are examples of chronic stress-induced immunosuppression (6, 7, 468). Finally, osteoporosis and syndrome X may result from chronic hyperactivity of stress effector systems as seen, for example, in melancholic depression (6, 469). Disturbances of central noradrenergic neurons repeatedly described in patients with typical depression may play an important role in the pathogenesis of this mood disorder (469). Based on previous reports showing stimulatory role of NE on CRH neurons, it was hypothesized that stress-induced NE release in the brain increases CRH neuron activity, which in turn feeds back to increase locus coeruleus firing (470). CRH acts either directly via its CRH receptors or by stimulating the HPA axis (for review, see Ref. 470). Recent studies show that transgenic mice overexpressing CRH exhibit an increase in anxiogenic behavior (471). In contrast, administration of a CRH receptor antagonist reduces anxiety behavior, and these drugs are now potential targets for a new generation of antidepressants (472). In terms of direct glucocorticoid action on brain function, transgenic mice in which glucocorticoid receptor gene expression is decreased serve as an excellent model to study the role of glucocorticoids in behavioral changes that resemble signs and symptoms found in human depression (473, 474). These behavioral deficits can be reversed by administration of antidepressants that inhibit monoamine oxidase (473). This inhibition results in increased synaptic cleft NE levels and glucocorticoid receptor signaling efficiency that may not necessarily depend on higher NE concentrations (for review, see Ref. 470).

To understand fully how each stressor operates to elicit particular organ dysfunction, one must determine whether stressor-specific pathways/circuits exist and whether stress-induced activation or inhibition of such pathways is associated with abnormal organ function, known as stress disorders. Our present studies are a starting point in developing a foundation of research in this field and for introducing for the future specific therapies for stress-related disorders.

Differences in neuroendocrine responses of an organism upon exposure to various stressors are not consistent with the existence of a unitary stress syndrome. Instead, we have introduced a new proposal that each stressor has its own unique “signature.” Thus, stress may be viewed as a response that reflects the existence of specific central anatomical and functional circuits. This view of stressor-specific circuits is, however, not new and has been discussed recently by several researchers working in the field of stress research (21, 53, 469, 475–482). For example, Herman and Cullinan (21) outlined interactions between stress-sensitive brain circuitry and neuroendocrine neurons of the PVN, especially as they relate to activation of the HPA axis. They divided stressors into two categories: those involving an immediate physiological threat, called systemic stressors, and those requiring interpretation by higher brain structures, called processive stressors. According to their theory, systemic stressors are relayed directly, probably via brainstem catecholaminergic projections, to the PVN. In contrast, processive stressors are channeled through limbic forebrain sites that upon their activation or inhibition affect the PVN via GABA-containing neurons. Recently, Li et al. (479) suggested the existence of different central circuits upon exposure to systemic vs. emotional stressors. In the biochemical switching hypothesis, Swanson (478) proposed that upon exposure to a stressor, stressor-specific changes in synthesis, transport, and storage of individual neurotransmitters exist, which determine a stressor-specific ratio of neuroactive substances at the axon terminal and serve as the final signal given to an effector system.

In our studies, we also demonstrated stressor-specific responses of the HPA axis and the sympathoadrenal system. In contrast to other studies, our experiments combined in vivo and ex vivo measurements of central and peripheral endocrine responses to various stressors. Figure 17 presents our proposed mechanisms by which noradrenergic neurons in the PVN may be involved in the regulation of the HPA axis after exposure to different stressors. When animals are exposed to metabolic stressors (e.g., hypoglycemia), activation of the HPA axis does not require noradrenergic neurons terminating in the PVN. Upon exposure to painful stressors (e.g., subcutaneous formalin), NE released in the PVN is the main stimulus of CRH neurons in the PVN. Activation of the HPA axis during immobilization stress, combining physical and psychological stressors, depends on both PVN noradrenergic and PVN or extra-PVN nonnoradrenergic mechanisms. Finally, maintaining homeostasis in an organism during cold stress does not require significant activation of the HPA axis and central noradrenergic neurons but does involve activation of other systems such as the hypothalamo-pituitary-thyroid axis. All of these data suggest the existence of stressor-specific central pathways that participate differentially in the regulation of sympathoneuronal and adrenomedullary outflow as well as activity of the HPA axis. Our conclusions further support our view and suggestions by others (483) that the neural representations of stress and other disease-related symptoms and signs cannot be described as a matter of altered function of one brain region or neurotransmitter. Rather, it reflects the activation of several circuits to orchestrate an optimal pattern of neuroendocrine and other responses. It is very likely that one of these circuits is the primary circuit, others are subordinate, and all combine to form the functional circuit to guarantee maximal plasticity of stress responses during acute as well as chronic stress conditions. Stressor-specific formation of new connections and disconnections of subordinate circuits with the primary one, anatomical and biochemical switches within the functional circuit, previous experience of an organism to an individual stressor, and genetically programmed neuronal and cellular functions precisely tune responses to stress and maximally protect from its deleterious effects. For example, in the response to an acute stressor, the failure of the primary circuit to function properly may have disastrous effects for an organism as is evident from neuroendocrine responses in patients with hypothalamic lesions; failure of subordinate circuits may contribute to milder symptoms and signs.

Fig. 17

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor?

Proposed schematic diagram of the role of PVN, NE, and other mechanisms on the regulation of the HPA axis during stress. Upon exposure to metabolic stressors (hypoglycemia, hemorrhage) PVN NE is only slightly elevated, and it plays a minor role in the activation of the HPA axis. Upon exposure to pain, PVN NE is increased significantly and is considered to be a potent stimulator of PVN CRH neurons. Although immobilization stress leads to a significant increase in PVN NE, a correlation between PVN NE and plasma ACTH levels suggests that immobilization stress-induced activation of the HPA axis depends on both PVN noradrenergic and PVN or extra-PVN nonnoradrenergic mechanisms. Finally, maintaining homeostasis in an organism exposed to cold stress does not require significant activation of PVN NE and the HPA axis, but the activation of other systems such as the hypothalamo-pituitary-thyroid axis is essential.

We are now left with the question of how the existence of stressor-specific circuits and the involvement of individual brain regions coordinating neuroendocrine responses of stress effector systems may help us in developing novel approaches to the pathogenesis of stress-related disorders and their possible treatment. For example, as proposed by Maier and Watkins (475), upon exposure of an organism to infection, IL-1 is released from macrophages and binds to receptors on paraganglia that surround vagal nerve fibers. Specific vagal nerve fibers carry a neural signal to the NTS and from there to specific brain areas such as the hypothalamus and hippocampus. Increased local production and release of PGs in the preoptic-anterior hypothalamic area alters the activity of temperature-sensitive neurons and fever occurs. If vagotomy is performed, PG release in the preoptic region is prevented and no fever occurs (484).

Another promising area of research in which the mapping of stressor-specific pathways may markedly contribute to the future development of therapies for stress-related disorders is the development of drugs that will specifically target neurons and their products, other neurocompounds, or neuroreceptors in individual brain regions. This is well documented by recent progress in the development of novel analgesic drugs (485). After peripheral nerve injury, a variety of pain syndromes, such as causalgia, thermal hyperalgesia, or reflex sympathetic dystrophy, occurs. Nociceptive signals are carried to the dorsal horn of the spinal cord where the processing of nociceptive information, including the participation of different neurotransmitters and other substances, has been recognized and described in great detail (485). Thus, one rational strategy for altering pain processing at the spinal level would be to affect specific substances (glutamate, substance P, galanin, vasoactive intestinal polypeptide, calcitonin gene-related peptide) released from primary afferents to prevent postsynaptic spinal excitation. Recent attention has been paid to glutamate and neurokinin receptor antagonists that have been shown to have potent antinociceptive effects (485–487). Pain-induced excitation of spinal postsynaptic neurons is coupled with an increase in intracellular Ca2+ followed by activation of nitric oxide synthase and cyclooxygenase. Not surprisingly, spinal delivery of a cyclooxygenase-2 inhibitor or L-type Ca2+ channel blockers was shown to attenuate the hyperalgesic state induced by tissue injury (488, 489). Targeted drugs and toxins are another promising area of pharmacological research that may be used in the regulation of stress responses by selective inactivation of transmitter release mechanisms of the targeted cell (485). Spinal delivery of substance P-saporin, which kills local neurons expressing the NK1 receptor and, thus, produces a potent antinociception, is one example (490).

Within the context of stressor specificity, individual stressors in turn down-regulate most “housekeeping” genes while, at the same time, inducing a small set of stressor-specific genes (9). Such stress genes can produce specific proteins (e.g., heat-shock proteins) that may play a critical role in preventing cells from sustaining damage and possibly undergoing apoptosis. For example, as discussed by Macario and Macario (9), cells can be stressed after exposure to heat shock. If they survive, they can tolerate a second heat shock at a much higher temperature that would otherwise be lethal for normal cells. The acquired resistance to a lethal temperature is caused by stress proteins, in this particular case by heat-shock proteins. These cells are also resistant to other stressors as well. This process is referred to as “cross-tolerance.” The potential clinical applications of this phenomenon can be tremendous. One can design a specific compound that could induce a mild, innocuous stress response in various cells that would then become preconditioned cells resistant to harmful stressors. A similar analogy is seen in the process of immunization (491, 492). Thus, we predict that in the near future we may identify a hereditary or acquired deficiency in the stress response caused by “defective” stress proteins.

Introducing knockout technology has tremendous potential in medical research, since knockout animal models help describe the pathogenesis of stress-related diseases and serve as important tools to develop new therapeutic approaches. At the present time there is no knockout mouse model that is characterized by increased susceptibility to stress-related disorders. Nevertheless, CRH, or CRH receptor knockout, knockdown, or overexpression mice represent a first attempt at addressing this hypothesis (491–494). We also envision that tissue-specific knockout technology will be applied to brain research and allow us to study the involvement of specific neurocompounds in particular brain regions in stress responses. The use of oligonucleotide antisense for specific mRNAs to prevent the expression of specific products (proteins) as a result of stressor-induced activation of individual neurons represents another approach for controlling stress responses (485).

Finally, new trends in brain imaging may help to identify stressor-specific neuroanatomical and functional circuitry. The field of nuclear medicine has evolved rapidly in recent years, especially given that new radiocompounds have been introduced to visualize a wide variety of disease processes at the cellular as well as molecular levels (495). Using positron emission tomography and functional magnetic resonance imaging in human stress research will open new directions and perspectives regarding ways to correlate stress-induced dysfunctions of peripheral organs with changes in activity of specific neurons in individual brain regions. The possibility of imaging stress-induced activity in various brain areas simultaneously and precisely will contribute further to the identification of stressor-specific circuits.

Despite the many difficulties in studying and defining stress, stress is an important area of medicine. We do not think that one magic stress hormone exists and we do not think that stress disorders can be treated simply. Nevertheless, the basic studies described in the present review illustrate the importance of continuing neuroscience research and applying new information to clinical practice. The proposed existence of stressor-specific pathways and circuits is definitely a step forward in the study of the pathogenesis of stress-related disorders. The identification of stressor- specific circuits helps us find “crucial” neurotransmitters or other neurocompounds and determine disordered sites (e.g., nuclei) located within individual stressor-specific brain regions. Specific targeting of these nuclei, their genetic equipment using tissue/region-specific knockout technology or their neurotransmitters and neurocompounds will bring novel information of their involvement in the pathogenesis of stress-related diseases and may promote the development of new therapeutic approaches. Furthermore, introducing new imaging techniques such as positron emission tomography or functional magnetic resonance imaging into stress research may bring unique information about the function of the “living” human brain. New imaging techniques will also allow for simultaneous observations of stressor-specific changes in central processes such as blood flow, neurotransmitter release, and receptor binding characteristics in individual regions with responses of stress effector systems. However, this may take time, since current basic and clinical research has just begun to unravel pathophysiological mechanisms for some major stress-related diseases. Meanwhile, we hope that researchers will elaborate on this theory of stressor specificity into their research and in further explorations of the pathogenesis of stress-related diseases.

Acknowledgments

We thank Drs. D. S. Goldstein and I. J. Kopin, who contributed in important ways to this work and without whom many of our previous studies could not have been started. We also thank Dr. R. McCarty for his valuable comments and suggestions during the preparation of this manuscript and R. Patel and T. Hoff for technical assistance.

Abbreviations:

     
  • CART,

    cocaine- and amphetamine-regulated transcript;

  •  
  • DA,

  •  
  • EPI,

  •  
  • GABA,

  •  
  • HPA,

    hypothalamo-pituitary-adrenocortical axis;

  •  
  • IML,

    intermediolateral cell column;

  •  
  • NE,

  •  
  • NGFI-B,

    nerve growth factor, type I-B;

  •  
  • NPY,

  •  
  • NTS,

    nucleus of the solitary tract;

  •  
  • PVN,

    hypothalamic paraventricular nucleus.

1

Selye

H

1974

Stress without distress.

New York

:

New York American Library

2

Goldstein

DS

1995

Stress, catecholamines, and cardiovascular disease.

New York

:

Oxford University Press

3

Ganong

WF

1995

Review of medical physiology. Norwalk, CT, Appleton & Lange;

327

351

4

Pacak

K

,

Palkovits

M

,

Yadid

G

,

Kvetnansky

R

,

Kopin

IJ

,

Goldstein

DS

1998

Heterogeneous neurochemical responses to different stressors: a test of Selye’s doctrine of nonspecificity. Am J Physiol 275:R1247–R1255.

5

Vigas

M

1985

Neuroendokrinna reakci v strese u cloveka. In: Veda (ed.), Bratislava.

6

Chrousos

GP

1998

Stressors, stress, and neuroendocrine integration of the adaptive response.

Ann NY Sci

851

:

311

335

7

Chrousos

GP

,

Gold

PW

1992

The concepts of stress and stress system disorders.

JAMA

267

:

1244

1252

8

Glavin

GB

1985

Stress and brain noradrenaline: a review.

Neurosci Biobehav Rev

9

:

233

243

9

Macario

AJL

,

de Macario

EC

1997

Stress genes: an introductory overview.

Stress

1

:

123

134

10

McCarty

R

1989

Stress research: principles, problems and prospects. In: Van

Loon

GR

,

Kvenansky

R

,

McCarty

R

,

Axelrod

J

, eds. Stress: neurochemical and humoral mechanisms.

New York

:

Gordon and Breach Science Publishers

;

3

13

11

McEwen

BS

1998

Protective and damaging effects of stress mediators.

N Engl J Med

338

:

171

179

12

Munck

A

,

Naray-Fejes-Toth

A

1994

Glucocorticoids and stress: permissive and suppressive actions.

Ann NY Acad Sci

746

:

115

133

13

Munck

A

,

Guyre

PM

,

Holbrook

NJ

1984

Physiological functions of glucocorticoids in stress and their relation to pharmacological actions.

Endocr Rev

5

:

25

44

14

Mason

JW

1971

A re-evaluation of the concept of “non-specificity” in stress theory.

J Psych Res

8

:

323

333

15

Mason

JW

1975

A historical view of the stress field. II.

J Hum Stress

1

:

6

–12,

22

36

16

Schulkin

J

,

McEwen

BS

,

Gold

PW

1994

Allostasis, amygdala, and anticipatory angst.

Neurosci Biobehav Rev

18

:

385

396

.

17

Selye

H

1976

Stress in health and disease.

Boston

:

Butterworth

18

Ceccatelli

S

,

Orazzo

C

1993

Effect of different types of stressors on peptide messanger ribonucleic acids in the hypothalamic paraventricular nucleus.

Acta Endocrinol (Copenh)

128

:

485

492

.

19

Ceccatelli

S

,

Villar

MJ

,

Goldstein

M

,

Hökfelt

T

1989

Expression of c-fos immunoreactivity in transmitter-characterized neurons after stress.

Proc Natl Acad Sci USA

86

:

9569

9573

20

Harbuz

MS

,

Jessop

DS

,

Lightman

SL

,

Chowdrey

HS

1994

The effects of restraint or hypertonic saline stress on corticotropin-releasing factor, arginine vasopressin, and proenkephalin A mRNAs in the CFY, Sprague-Dawley and Wistar strains of rat.

Brain Res

667

:

6

12

21

Herman

JP

,

Cullinan

WE

1997

Neurocircuitry of stress: central control of the hypothalamo-pituitary-adrenocortical axis.

Trends Neurosci

20

:

78

84

22

Komesaroff

PA

,

Funder

JW

1994

Differential glucocorticoid effects on catecholamine responses to stress

.

Am J Physiol

266

:

E118

E128

23

Lachuer

J

,

Gaillet

S

,

Barbagli

B

,

Buda

M

,

Tappaz

M

1991

Differential early time course activation of the brainstem catecholaminergic groups in response to various stresses.

Neuroendocrinology

53

:

589

596

24

Pacak

K

,

Palkovits

M

,

Kopin

IJ

,

Goldstein

DS

1995

Stress-induced norepinephrine release in the hypothalamic paraventricular nucleus and pituitary-adrenocortical and sympathoadrenal activity: in vivo microdialysis studies.

Front Neuroendocrinol

16

:

89

150

25

Pacak

K

,

Baffi

JS

,

Kvetnansky

R

,

Goldstein

DS

,

Palkovits

M

1998

Stressor-specific activation of catecholaminergic systems: implications for stress-related hypothalamic-pituitary-adrenocortical responses.

Adv Pharmacol

42

:

561

564

26

Palkovits

M

,

Baffi

JS

,

Pacak

K

1997

Stress-induced Fos-like immunoreactivity in the pons and the medulla oblongata of rats.

Stress

1

:

155

168

27

Plotsky

PM

,

Bruhn

TO

,

Vale

W

1985

Hypophysiotrophic regulation of adrenocorticotropin secretion in response to insulin- induced hypoglycemia.

Endocrinology

117

:

323

329

28

Senba

E

,

Ueyama

T

1997

Stress-induced expression of immediate early genes in the brain and peripheral organs of the rat.

Neurosci Res

29

:

183

207

28A

Herdegen

T

,

Leah

JD

1998

Inducible and constitutive transcription factors in the mammalian nervous system: control of gene expression by Jun, Fos, and Krox, and CREB/ATF protein.

Brain Res Rev

28

:

370

490

29

Baffi

J

,

Pacak

K

,

Szabó

S

,

Palkovits

M

1996

Different stressors have different effects on c-fos activity of catecholaminergic neurons. In:

McCarty

R

,

Aguilera

G

,

Sabban

E

,

Kvetnansky

R

, eds. Stress: molecular genetic and neurological advances.

New York

:

Gordon and Breach

;

143

155

30

Cannon

WB

1929

Organization for physiological homeostasis.

Physiol Rev

9

:

399

431

31

Cannon

WB

1929

Bodily changes in pain, hunger, fear, and rage

.

New

York

:

Appleton

32

Cannon

WB

1939

The wisdom of the body.

New York

:

WW Norton

33

Selye

H

1936

A syndrome produced by diverse nocious agents.

Nature

138

:

32

34

Selye

H

1950

The physiology and pathology of exposure to stress.

A treatise based on the concepts of the general adaptation syndrome and the diseases of adaptation. Montreal

:

Acta Inc.

35

Cannon

WB

1914

The interrelations of emotions as suggested by recent physiological researches.

Am J Physiol

25

:

256

282

36

Weiner

H

1991

Behavioral biology of stress and psychosomatic medicine. In:

Brown

MR

,

Koob

GF

,

Rivier

C

, eds. Stress.

Neurobiology and neuroendocrinology. New York

:

Marcel Dekker

;

23

51

37

Sterling

P

,

Eyer

J

1981

Allostasis: a new paradigm to explain arousal pathology. In: Fisher S, Reason HS, eds. Handbook of life stress, cognition, and health.

New York

:

John Wiley and Sons

38

Minowada

G

,

Welch

WJ

1995

Clinical implications of the stress response.

J Clin Invest

95

:

3

12

39

Sapolsky

RM

,

Pulsinelli

WA

1985

Glucocorticoids potentiate ischemic injury to neurons: therapeutic implications.

Science

229

:

1397

1400

40

Sapolsky

RM

1985

A mechanism for glucocorticoid toxicity in the hippocampus: increased neuronal vulnerability to metabolic insults.

J Neurosci

5

:

1228

1232

41

Sapolsky

RM

,

Uno

H

,

Rebert

CS

,

Finch

CE

1990

Hippocampal damage associated with prolonged glucocorticoid exposure in primates.

J Neurosci

10

:

2897

2902

42

McEwen

BS

,

De Kloet

ER

,

Rostene

W

1986

Adrenal steroid receptors and actions in the nervous system.

Physiol Rev

66

:

1121

1188

43

Herman

JP

,

Schäfer

MK-H

,

Young

EA

,

Thompson

R

,

Douglas

J

,

Akil

H

,

Watson

SJ

1989

Evidence for hippocampal regulation of neuroendocrine neurons of the hypothalamo-pituitary-adrenocortical axis.

J Neurosci

9

:

3072

3082

44

Herman

JP

,

Prewitt

CM

,

Cullinan

WE

1996

Neuronal circuit regulation of the hypothalamo-pituitary-adrenocortical stress axis.

Crit Rev Neurobiol

10

:

371

394

45

Kerr

DS

,

Campbell

LW

,

Applegate

MD

,

Brodish

A

,

Landfield

PW

1991

Chronic stress-induced acceleration of electrophysiologic and morphometric biomarkers of hippocampal aging.

J Neurosci

11

:

1316

1324

46

Nichols

NR

,

Dokas

L

,

Ting

SM

,

Kumar

S

,

de Vellis

J

,

Shors

TJ

,

Uenishi

N

,

Thompson

RF

,

Finch

CE

1996

Hippocampal responses to corticosterone and stress, one of which is the 35,000 M(r) protein, glycerol phosphate dehydrogenase.

J Neuroendocrinol

8

:

867

876

47

Nichols

NR

,

Masters

JN

,

Finch

CE

1990

Changes in gene expression in hippocampus in response to glucocorticoids and stress.

Brain Res Bull

24

:

659

662

48

Sapolsky

RM

,

Romero

LM

,

Munck

AU

2000

How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions.

Endocr Rev

21

:

55

89

49

Morgan

TE

,

Xie

Z

,

Goldsmith

S

,

Yoshida

T

,

Lanzrein

AS

,

Stone

D

,

Rozovsky

I

,

Perry

G

,

Smith

MA

,

Finch

CE

1999

The mosaic of brain glial hyperactivity during normal ageing and its attenuation by food restriction.

Neuroscience

89

:

687

699

50

Morgan

TE

,

Rozovsky

I

,

Goldsmith

SK

,

Stone

DJ

,

Yoshida

T

,

Finch

CE

1997

Increased transcription of the astrocyte gene GFAP during middle-age is attenuated by food restriction: implications for the role of oxidative stress.

Free Radic Biol Med

23

:

524

528

51

Behl

C

,

Trapp

T

,

Skutella

T

,

Holsboer

F

1997

Protection against oxidative stress-induced neuronal cell death—a novel role for RU486.

Eur J Neurosci

9

:

912

920

52

Lezoualc’h

F

,

Engert

S

,

Berning

B

,

Behl

C

2000

Corticotropin-releasing hormone-mediated neuroprotection against oxidative stress is associated with the increased release of nonamyloidogenic amyloid β precursor protein and with the suppression of nuclear factor-κB.

Mol Endocrinol

14

:

147

159

53

van

de Kar

LD

,

Blair

ML

1999

Forebrain pathways mediating stress-induced hormone secretion.

Front Neuroendocrinol

20

:

1

48

54

Weiss

JM

1968

Effects of coping responses on stress.

J Comp Physiol Psychol

65

:

251

260

55

Weiss

JM

1971

Effects of coping behavior with and without a feedback signal on stress pathology in rats.

J Comp Physiol Psychol

77

:

22

30

56

Weiss

JM

1971

Effects of coping behavior in different warning signal conditions on stress pathology in rats.

J Comp Physiol Psychol

77

:

1

13

57

Weiss

JM

1971

Effects of punishing the coping response (conflict) on stress pathology in rats.

J Comp Physiol Psychol

77

:

14

21

58

DeLongis

A

,

Preece

M

2000

Coping skills. In:

Fink

G

, ed. Encyclopedia of stress, vol.

1. New York

:

Academic Press

;

532

541

59

Cohen

F

,

Lazarus

R

1979

Coping with the stresses of illness. In:

Stone

GC

,

Cohen

F

,

Adler

NE

, eds. Health physiology: a handbook.

San Francisco

:

Jossey-Bass

;

217

254

60

McEwen

BS

1997

Stress, brain, and behavior: life-long effects upon health and disease. In:

Kinney

JM

,

Tucker

HN

, eds. Physiology, stress, and malnutrition.

Functional correlates of nutritional intervention. Philadelphia

:

Lippincott-Raven

;

113

130

61

Sapolsky

RM

1994

Why zebras don’t get ulcers.

New York

:

W.H. Freeman and Company

62

Gaillet

S

,

Lachuer

J

,

Malaval

F

,

Assenmacher

I

,

Szafarczyk

A

1991

The involvement of noradrenergic ascending pathways in the stress-induced activation of ACTH and corticosterone secretions is dependent on the nature of stressors.

Exp Brain Res

87

:

173

180

63

Palkovits

M

,

Baffi

JS

,

Dvori

S

1995

Neuronal organization of stress response. Pain-induced c-fos expression in brain stem catecholaminergic cell groups.

Ann NY Acad Sci

771

:

313

326

64

Palkovits

M

1999

Interconnections between the neuroendocrine hypothalamus and the autonomic system.

Front Neuroendocrinol

20

:

1

26

65

Cunningham

ET

,

Sawchenko

PE

1988

Anatomical specificity of noradrenergic inputs to the paraventricular and supraoptic nuclei of the rat hypothalamus.

J Comp Neurol

274

:

60

76

66

Palkovits

M

,

Záborszky

L

,

Feminger

A

,

MezeyÉ , Fekete

MIK

,

Herman

JP

,

Kanyicska

B

,

Szabó

D

1980

Noradrenergic innervation of the rat hypothalamus: experimental biochemical and electron microscopic studies.

Brain Res

191

:

161

171

67

Sawchenko

PE

,

Swanson

LW

1982

The organization of noradrenergic pathways from the brainstem to the paraventricular and supraoptic nuclei in the rat.

Brain Res Rev

4

:

275

325

68

Sawchenko

PE

,

Swanson

LW

1983

The organization of forebrain afferents to the paraventricular and supraoptic nuclei of the rat.

J Comp Neurol

218

:

121

144

69

Swanson

LW

1987

The hypothalamus. In:

Björklund

A

,

Hökfelt

T

,

Swanson

LW

, eds. Handbook of chemical neuroanatomy, vol. 5: integrated systems of the CNS, pt. 1, chapt.

1. Amsterdam

:

Elsevier

;

1

124

70

Foote

SL

,

Bloom

FE

,

Aston-Jones

G

1983

The nucleus locus coeruleus: new evidence of anatomical and physiological specificity.

Physiol Rev

63

:

844

914

71

Nygren

L-G

,

Olson

L

1977

A new major projection from locus coeruleus: the main source of noradrenergic nerve terminals in the ventral and dorsal columns of the spinal cord.

Brain Res

132

:

85

93

72

Clark

FM

,

Proudfit

HK

1991

The projection of locus coeruleus to the spinal cord in the rat determined by anterograde tracing combined with immunohistochemistry.

Brain Res

538

:

231

245

73

Loughlin

SE

,

Foote

SL

,

Grzanna

R

1986

Efferent projections of nucleus locus coeruleus: morphologic subpopulations have different efferent targets.

Neuroscience

18

:

307

319

74

Satoh

K

,

Tohyama

M

,

Yamamoto

K

,

Sakumoto

T

,

Shimizu

N

1980

Noradrenaline innervation of the spinal cord studied by the horseradish method combined with monoamine oxidase staining.

Exp Brain Res

30

:

175

186

75

Westlund

KN

,

Bowker

RM

,

Ziegler

MC

,

Coulter

JD

1983

Noradrenergic projections to the spinal cord of the rat.

Brain Res

263

:

15

31

76

Loughlin

SE

,

Foote

SL

,

Bloom

FE

1986

Efferent projections of nucleus locus coeruleus: topographic organization of cells of origin demonstrated by three-dimensional reconstruction.

Neuroscience

18

:

291

306

77

Loewy

AD

,

McKellar

S

,

Saper

CB

1979

Direct projections from the A5 catecholamine cell group to the intermediolateral cell column.

Brain Res

174

:

309

314

78

Byrum

CE

,

Guyenet

PG

1987

Afferent and efferent connections of the A5 noradrenergic cell group in the rat.

J Comp Neurol

261

:

529

542

79

Hosoya

Y

,

Sugiura

Y

,

Ito

R

,

Kohno

K

1990

Descending projections from the hypothalamic paraventricular nucleus to the A5 area, including the superior salivatory nucleus, in the rat.

Exp Brain Res

82

:

513

518

80

Clark

FM

,

Proudfit

HK

1991

The projection of noradrenergic neurons in the A7 catecholamine cell groups to the spinal cord in the rat demonstrated by anterograde tracing combined with immunocytochemistry.

Brain Res

547

:

279

288

81

Shafton

AD

,

Ryan

A

,

Badoer

E

1998

Neurons in the hypothalamic paraventricular nucleus send collaterals to the spinal cord and to the rostral ventrolateral medulla in the rat.

Brain Res

801

:

239

243

82

Ross

CA

,

Armstrong

DM

,

Ruggiero

DA

,

Pickel

VM

,

Joh

TH

,

Reis

DJ

1981

Adrenaline neurons in the rostral ventrolateral medulla innervate thoracic spinal cord: a combined immunocytochemical and retrograde transport demonstration.

Neurosci Lett

25

:

257

262

83

Ross

CA

,

Ruggiero

DA

,

Park

DH

,

Joh

TH

,

Sved

AF

,

Fernandez-Pardal

J

,

Saavedra

JM

,

Reis

DJ

1984

Tonic vasomotor control by the rostral ventrolateral medulla: effect of electrical or chemical stimulation of the area containing C1 adrenaline neurons on arterial pressure, heart rate, and plasma catecholamines and vasopressin.

J Neurosci

4

:

474

494

84

Tucker

DC

,

Saper

CB

,

Ruggiero

DA

,

Reis

DJ

1987

Organization of central adrenergic pathways. I. Relationships of ventrolateral medullary projections to the hypothalamus and spinal cord.

J Comp Neurol

259

:

591

603

85

Cunningham

ETJ

,

Bohn

MC

,

Sawchenko

PE

1990

Organization of adrenergic inputs to the paraventricular and supraoptic nuclei of the hypothalamus in the rat.

J Comp Neurol

292

:

651

667

86

Palkovits

M

,

Mezey É, Skirboll

LR

,

Hökfelt

T

1992

Adrenergic projections from the lower brainstem to the hypothalamic paraventricular nucleus, the lateral hypothalamic area and the central nucleus of the amygdala in rats.

J Chem Neuroanat

5

:

407

415

87

Brownstein

MJ

,

Palkovits

M

1984

Catecholamines, serotonin, acetylcholine, and γ-aminobutyric acid in the rat brain: biochemical studies. In:

Björklund

A

,

Hökfelt

T

, eds. Handbook of chemical neuroanatomy, vol. 2: classical transmitters in the CNS, pt. 1, chapt.

2. Amsterdam

:

Elsevier

;

23

54

88

Steinbusch

HWM

1984

Serotonin-immunoreactive neurons and their projections in the CNS. In:

Björklund

A

,

Hökfelt

T

,

Kuhar

MJ

, eds. Handbook of chemical neuroanatomy, vol. 3: classical transmitter receptors in the CNS, pt. 2, chapt.

3. Amsterdam

:

Elsevier

;

68

125

89

Friedman

E

,

Krieger

DT

,

Mezey É, Léránth

C

,

Brownstein

MJ

,

Palkovits

M

1983

Serotonergic innervation of the rat pituitary intermediate lobe: decrease after stalk section.

Endocrinology

112

:

1943

1947

90

Palkovits

M

,

Mezey É, Chiueh

CG

,

Krieger

DT

,

Gallatz

K

,

Brownstein

MJ

1986

Serotonin-containing elements of the rat pituitary intermediate lobe.

Neuroendocrinology

42

:

522

525

91

Bowker

RM

,

Westlund

KN

,

Sullivan

MC

,

Wilber

JF

,

Coulter

JD

1986

Innervation of the nucleus of the solitary tract and the dorsal vagal nucleus by thyrotropin-releasing hormone-containing raphe neurons.

Brain Res

373

:

246

251

92

Johansson

O

,

Hökfelt

T

,

Pernow

B

,

Jeffcoate

SL

,

White

N

,

Steinbusch

HWM

,

Verhofstad

AAJ

,

Emson

PC

,

Spindel

E

1981

Immunohistochemical support for three putative transmitters in one neuron: coexistence of 5-hydroxytryptamine, substance P and thyrotropin-releasing hormone-like immunoreactivity in medullary neurons projecting to the spinal cord.

Neuroscience

6

:

1857

1881

93

Menétrey

D

,

Basbaum

AI

1987

Spinal and trigeminal projections to the nucleus of the solitary tract. A possible substrate for somatovisceral and viscerovisceral reflex activation.

J Comp Neurol

255

:

439

450

94

Riche

D

,

Pommery

JD

,

Menétrey

D

1990

Neuropeptides and catecholamines in efferent projections of the nuclei of the solitary tract in the rat.

Neuroscience

293

:

399

424

95

Sawchenko

PE

,

Arias

C

,

Bittencourt

JC

1990

Inhibin β, somatostatin, and enkephalin immunoreactivities coexist in caudal medullary neurons that project to the paraventricular nucleus of the hypothalamus.

J Comp Neurol

291

:

269

280

96

Basbaum

AI

,

Fields

HL

1984

Endogenous pain control system: brainstem spinal pathways and endorphin circuitry.

Annu Rev Neurosci

7

:

309

338

97

Saper

CB

,

Loewy

AD

1980

Efferent connections of the parabrachial nucleus in the rat.

Brain Res

197

:

291

317

98

Cechetto

DF

,

Standaert

DG

,

Saper

CB

1985

Spinal and trigeminal dorsal horn projections to the parabrachial nucleus in the rat.

J Comp Neurol

240

:

153

160

99

Bernard

J-F

,

Alden

M

,

Besson

J-M

1993

The organization of the efferent projections from the pontine parabrachial area to the amygdaloid complex: a Phaseolus vulgaris leucoagglutinin (PHA-L) study in the rat.

J Comp Neurol

329

:

201

229

100

Jia

H-G

,

Rao

Z-R

,

Shi

J-W

1994

An indirect projection from the nucleus of the solitary tract to the central nucleus of the amygdala via the parabrachila nucleus in the rat: a light and electron microscopic study.

Brain Res

663

:

181

190

101

Alden

M

,

Besson

JM

,

Bernard

JF

1994

Organization of the efferent projections from the pontine parabrachial area to the bed nucleus of the stria terminalis and neighbouring regions: a PHA-L study in the rat.

J Comp Neurol

341

:

289

314

102

Saleh

TM

1997

Visceral afferent stimulation-evoked changes in the release of peptides into the parabrachial nucleus in vivo.

Brain Res

778

:

56

63

103

Hopkins

DA

,

Bieger

D

,

De Vente

J

,

Steinbusch

HWM

1996

Vagal efferent projections: viscerotopy, neurochemistry, and effects of vagotomy.

Prog Brain Res

107

:

79

96

104

Farkas

E

,

Jansen

ASP

,

Loewy

AD

1997

Periaqueductal gray matter projection to vagal preganglionic neurons and the nucleus tractus solitarius.

Brain Res

764

:

257

261

105

Bullitt

E

1989

Induction of c-fos-like protein within the lumbar spinal cord and thalamus of the rat following peripheral stimulation.

Brain Res

493

:

391

397

106

Pezzone

MA

,

Lee

W-S

,

Hoffman

GE

,

Rabin

BS

1992

Induction of c-Fos immunoreactivity in the rat forebrain by conditioned and unconditioned aversive stimuli.

Brain Res

597

:

41

50

107

Senba

E

,

Matsunaga

K

,

Tohyama

M

,

Noguchi

K

1993

Stress-induced c-fos expression in the rat brain: activation mechanism of sympathetic pathway.

Brain Res Bull

31

:

329

341

108

Palkovits

M

,

Dvori

S

,

Pacak

K

,

Baffi

J

1996

Stress-induced functional activity in neuronal cells of catecholaminergic hypothalamic and amygdaloid nuclei. In:

McCarty

R

,

Aguilera

G

,

Sabban

E

,

Kvetnansky

R

, eds. Stress: molecular genetic and neurobiological advances.

New York

:

Gordon and Breach

;

23

47

109

ter Horst

GJ

,

Luiten

PG

,

Kuipers

F

1984

Descending pathways from hypothalamus to dorsal motor vagus and ambiguus nuclei in the rat.

J Auton Nerv Syst

11

:

59

75

110

Luiten

PG

,

ter Horst

GJ

,

Karst

H

,

Steffens

AB

1985

The course of paraventricular hypothalamic efferents to autonomic structures in medulla and spinal cord.

Brain Res

329

:

374

378

111

Tucker

DC

,

Saper

CB

1985

Specificity of spinal projections from hypothalamic and brainstem areas which innervate sympathetic preganglionic neurons.

Brain Res

360

:

159

164

112

Cechetto

DF

,

Saper

CB

1988

Neurochemical organization of the hypothalamic projection to the spinal cord in the rat.

J Comp Neurol

272

:

579

604

113

Hosoya

Y

,

Sugiura

Y

,

Okado

N

,

Loewy

AD

,

Kohno

K

1991

Descending input from the hypothalamic paraventricular nucleus to sympathetic preganglionic neurons in the rat.

Exp Brain Res

85

:

10

20

114

Portillo

F

,

Carrasco

M

,

Vallo

JJ

1998

Separate populations of neurons within the paraventricular hypothalamic nucleus of the rat project to vagal and thoracic autonomic preganglionic levels and express c-Fos protein induced by lithium chloride.

J Chem Neuroanat

14

:

95

102

115

Tóth

ZE

,

Gallatz

K

,

Fodor

M

,

Palkovits

M

1999

Decussations of the descending paraventricular pathways to the brainstem and spinal cord autonomic centers.

J Comp Neurol

414

:

255

266

116

Sawchenko

PE

,

Swanson

LW

1982

Immunohistochemical identification of neurons in the paraventricular nucleus of the hypothalamus that project to the medulla or to the spinal cord in the rat.

J Comp Neurol

205

:

260

272

117

Palkovits

M

1989

Neuroanatomical overview of brain neurotransmitters in stress. In: Van

Loon

GR

,

Kvetnansky

R

,

McCarty

R

,

Axelrod

J

, eds. Stress: neurochemical and humoral mechanisms, vol.

1. New York

:

Gordon & Breach Scientific Publishers

;

31

42

118

Palkovits

M

1992

Peptidergic neurotransmitters in the endocrine hypothalamus. In: Functional anatomy of the neuroendocrine hypothalamus.

Ciba Foundation Symposium 168; A Wiley-Interscience Publication; Chichester-New York

:

John Wiley & Sons; 3–15

119

Gray

TS

,

Carney

ME

,

Magnuson

DJ

1989

Direct projections from the central amygdaloid nucleus to the hypothalamic paraventricular nucleus: possible role in stress-induced adrenocorticotropin release.

Neuroendocrinology

50

:

433

446

120

Swanson

LW

,

Köhler

C

,

Björklund

A

1987

The limbic region. I. The septohippocampal system. In:

Björklund

A

,

Hökfelt

T

,

Swanson

LW

, eds. Handbook of chemical neuroanatomy, vol. 5: integrated systems of the CNS, pt. 1, chapt.

2. Amsterdam

:

Elsevier

;

125

277

121

Price

JL

,

Russchen

FT

,

Amaral

DG

1987

The limbic region. II. The amygdaloid complex. In:

Björklund

A

,

Hökfelt

T

,

Swanson

LW

, eds. Handbook of chemical neuroanatomy, vol. 5: integrated systems of the CNS, pt.1, chapt.

3. Amsterdam

:

Elsevier

;

279

388

122

Veening

JG

,

Swanson

LW

,

Sawchenko

PE

1984

The organization of projections from the central nucleus of the amygdala to brainstem sites involved in central autonomic regulation: a combined retrograde transport-immunohistochemical study.

Brain Res

303

:

337

357

123

Gray

TS

,

Magnuson

DJ

1987

Neuropeptide neuronal efferents from the bed nucleus to the stria terminalis and central amygdaloid nucleus to the dorsal vagal complex in the rat.

J Comp Neurol

262

:

365

374

124

Wallace

DM

,

Magnuson

DJ

,

Gray

TS

1989

The amygdalobrainstem pathway: selective innervation of dopaminergic, noradrenergic and adrenergic cells in the rat.

Neurosci Lett

97

:

252

258

125

Benveniste

H

,

Huttemeier

CH

1990

Microdialysis: theory and application.

Prog Neurobiol

35

:

195

215

126

Connelly

CA

1999

Microdialysis update: optimizing the advantages.

J Physiol (Lond)

514

:

303

127

Delgado

JMR

,

Feudis

FV

,

Roth

RH

,

Ryugo

DK

,

Mitruka

BM

1972

Dialytrode for long-term intravertebral perfusion in awake monkeys.

Arch Int Pharmacodyn Ther

198

:

9

21

128

Ungerstedt

U

,

Backstrom

T

,

Hallström

A

,

Grande

PO

,

Mellergard

P

,

Nordström

C-H

1997

Microdialysis in normal and injured human brain. In:

Kinney

JM

,

Tucker

HN

, eds. Physiology, stress, and malnutrition.

Functional correlates and nutritional intervention. Philadelphia

:

Lippincott-Raven Publishers

;

361

374

129

Zetterstrom

T

,

Vernet

L

,

Ungerstedt

U

,

Tossman

U

,

Jonzon

B

,

Fredholm

BB

1982

Purine levels in the intact rat brain. Studies with an implanted perfused hollow fibre.

Neurosci Lett

29

:

111

115

130

Pacak

K

,

Palkovits

M

,

Kvetnansky

R

,

Yadid

G

,

Kopin

IJ

,

Goldstein

DS

1995

Effects of various stressors on in vivo norepinephrine release in the hypothalamic paraventricular nucleus and on the pituitary-adrenocortical axis.

Ann NY Acad Sci

771

:

115

130

131

Pacak

K

,

Palkovits

M

,

Kvetnansky

R

,

Kopin

IJ

,

Goldstein

DS

1993

Stress-induced norepinephrine release in the paraventricular nucleus of rats with brainstem hemisections: a microdialysis study.

Neuroendocrinology

58

:

196

201

132

Pacak

K

,

Armando

I

,

Fukuhara

K

,

Kvetnansky

R

,

Palkovits

M

,

Kopin

IJ

,

Goldstein

DS

1992

Noradrenergic activation in the paraventricular nucleus during acute and chronic immobilization stress in rats: an in vivo microdialysis study.

Brain Res

589

:

91

96

133

Pacak

K

,

Palkovits

M

,

Kvetnansky

R

,

Fukuhara

K

,

Armando

I

,

Kopin

IJ

,

Goldstein

DS

1993

Effects of single or repeated immobilization on release of norepinephrine and its metabolites in the central nucleus of the amygdala in conscious rats.

Neuroendocrinology

57

:

626

633

133A

Baffi

J

,

Palkovits

M

2000

Fine topography of brain areas activated by cold stress.

Neuroendocrinology

72

:

102

113

134

Hoffman

GE

,

Smith

MS

,

Verbalis

JG

1993

C-Fos and related immediate early gene products as markers of activity in neuroendocrine systems.

Front Neuroendocrinol

14

:

173

213

135

Chan

RKW

,

Brown

ER

,

Ericsson

A

,

Kovacs

KJ

,

Sawchenko

PE

1993

A comparison of two immediate-early genes, c-fos and NGFI-B, as markers for functional activation in stress-related neuroendocrine circuitry.

J Neurosci

13

:

5126

5138

136

Selye

H

1936

Thymus and adrenals in the response of the organism to injuries and intoxications.

Br J Exp Pathol

17

:

234

248

137

Shibasaki

T

,

Tsumori

C

,

Hotta

M

,

Imaki

T

,

Yamada

K

,

Demura

H

1995

The response pattern of noradrenaline release to repeated stress in the hypothalamic paraventricular nucleus differs according to the form of stress in rats.

Brain Res

670

:

169

172

138

Cullinan

WE

,

Herman

JP

,

Battaglia

DF

,

Akil

H

,

Watson

SJ

1995

Pattern and time course of immediate early gene expression in rat brain following acute stress.

Neuroscience

64

:

477

505

139

Chen

X

,

Herbert

J

1995

Regional changes in c-fos expression in the basal forebrain and brainstem during adaptation to repeated stress: correlations with cardiovascular, hypothermic and endocrine responses.

Neuroscience

64

:

675

685

140

Herbert

J

,

Howes

SR

1993

Interactions between corticotropin-releasing factor and endogenous opiates on the cardioaccelerator, hypothermic, and corticoid responses to restraint in the rat.

Peptides

14

:

145

152

141

Fodor

M

,

Palkovits

M

1991

Neuropeptide Y-containing neuronal pathway from the spinal trigeminal nucleus to the pontine peribrachial region in the rat.

Neurosci Lett

133

:

195

198

142

Slugg

RM

,

Light

AR

1994

Spinal cord and trigeminal projections to the pontine parabrachial region in the rat as demonstrated with Phaseolus vulgaris leucoagglutinin.

J Comp Neurol

339

:

49

61

143

Miyata

S

,

Lin

SH

,

Kawarabayashi

T

,

Nakashima

T

,

Kiyohara

T

1995

Maintenance of ultrastructural plasticity of the hypothalamic supraoptic nucleus in the ovariectomized rat.

Brain Res Bull

37

:

405

409

144

Arnold

FJL

,

de Lucas Bueno

M

,

Shiers

H

,

Hancock

DC

,

Evans

GI

,

Herbert

J

1992

Expression of c-fos in regions of the basal limbic forebrain following intracerebroventricular corticotropin-releasing factor in unstressed or stressed male rats.

Neuroscience

51

:

377

390

145

Kononen

J

,

Honkaniemi

J

,

Alho

H

,

Koistinaho

J

,

Iadarola

M

,

Pelto-Huikko

M

1992

Fos-like immunoreactivity in the rat hypothalamic-pituitary axis after immobilization stress.

Endocrinology

130

:

3041

3047

146

Covenas

R

,

de León

M

,

Cintra

A

,

Bjelke

B

,

Gustafsson

J-A

,

Fuxe

K

1993

Coexistence of c-fos and glucocorticoid receptor immunoreactivities in the CRF immunoreactive neurons of the paraventricular hypothalamic nucleus of the rat after acute immobilization stress.

Neurosci Lett

149

:

149

152

147

Roland

BL

,

Sawchenko

PE

1993

Local origins of some GABAergic projections to the paraventricular and supraoptic nuclei of the hypothalamus in the rat.

J Comp Neurol

332

:

123

143

148

Pacak

K

,

Palkovits

M

,

Makino

S

,

Kopin

IJ

,

Goldstein

DS

1996

Brainstem hemisection decreases corticotropin-releasing hormone mRNA in the paraventricular nucleus but not in the central amygdaloid nucleus.

J Neuroendocrinol

8

:

543

551

149

Bartanusz

V

,

Jezova

D

,

Bertini

LT

,

Tilders

FJ

,

Aubry

JM

,

Kiss

JZ

1993

Stress-induced increase in vasopressin and corticotropin- releasing factor expression in hypophysiotrophic paraventricular neurons.

Endocrinology

132

:

895

902

150

Kalin

NH

,

Takahashi

LK

,

Chen

F-L

1994

Restraint stress increases corticotropin-releasing hormone mRNA content in the amygdala and paraventricular nucleus.

Brain Res

656

:

182

186

151

Harbuz

MS

,

Chalmers

J

,

De Souza

L

,

Lightman

SL

1993

Stress-induced activation of CRF and c-fos mRNAs in the paraventricular nucleus are not affected by serotonin depletion

.

Brain Res

609

:

167

173

152

Harbuz

MS

,

Lightman

SL

1989

Responses of hypothalamic and pituitary mRNA to physical and psychological stress in the rat.

J Endocrinol

122

:

705

711

153

Harbuz

MS

,

Chowdrey

HS

,

Jessop

DS

,

Biswas

S

,

Lightman

SL

1991

Role of catecholamines in mediating messenger RNA and hormonal responses to stress.

Brain Res

551

:

52

57

154

Luo

X

,

Kiss

A

,

Makara

GB

,

Lolait

S

,

Aguilera

G

1994

Stress-specific regulation of corticotropin releasing hormone receptor expression in the paraventricular and supraoptic nuclei of the hypothalamus in the rat.

J Neuroendocrinol

6

:

689

696

155

Bartanusz

V

,

Aubry

JM

,

Jezova

D

,

Baffi

J

,

Kiss

JZ

1993

Up-regulation of vasopressin mRNA in paraventricular hypophysiotrophic neurons after acute immobilization stress.

Neuroendocrinology

58

:

625

629

156

Herman

JP

1995

In situ hybridization analysis of vasopressin gene transcription in the paraventricular and supraoptic nuclei of the rat: regulation by stress and glucocorticoids.

J Comp Neurol

363

:

15

27

157

Mamalaki

E

,

Kvetnansky

R

,

Brady

LS

,

Gold

PW

,

Herkenham

M

1992

Repeated immobilization stress alters tyrosine hydroxylase, corticotropin-releasing hormone and corticosteroid receptor messenger ribonucleic acid levels in rat brain

.

J Neuroendocrinol

4

:

689

699

158

Simpkins

JW

,

Hodson

CA

,

Meites

J

1978

Differential effects of stress on release of thyroid-stimulating hormone in young and old male rats.

Proc Soc Exp Biol Med

157

:

144

147

159

Armario

A

,

Marti

O

,

Gavalda

A

,

Giralt

M

,

Jolin

T

1993

Effects of chronic immobilization stress on GH and TSH secretion in the rat: response to hypothalamic regulatory factors.

Psychoneuroendocrinology

18

:

405

413

160

Cizza

G

,

Brady

LS

,

Pacak

K

,

Blackman

MR

,

Gold

PW

,

Chrousos

GP

1995

Stress-induced inhibition of the hypothalamic-pituitary-thyroid axis is attenuated in the aged Fischer 344/N male rat.

Neuroendocrinology

62

:

506

513

161

Cizza

G

,

Brady

LS

,

Esclapes

M

,

Blackman

MR

,

Gold

PW

,

Chrousos

GP

1996

Age and gender influence basal and stress-modulated hypothalamic-pituitary-thyroidal function in Fischer 344/N rats.

Neuroendocrinology

64

:

440

448

162

Takayama

H

,

Ota

Z

,

Ogawa

N

1986

Effect of immobilization stress on neurupeptides and their receptors in rat central nervous system.

Regul Pept

15

:

239

248

163

Kvetnansky

R

,

Palkovits

M

,

Mitro

A

,

Torda

T

,

Mikulaj

L

1977

Catecholamines in individual hypothalamic nuclei of acutely and repeatedly stressed rats.

Neuroendocrinology

23

:

257

267

164

Tanaka

MY

,

Kohno

Y

,

Nakagawa

R

,

Ida

Y

,

Takeda

S

,

Nagasaki

N

1982

Time-related differences in noradrenaline turnover in rat brain regions by stress.

Pharmacol Biochem Behav

16

:

315

319

165

Hellriegel

ET

,

D’Mello

PD

1997

The effect of acute, chronic and chronic intermittent stress on the central noradrenergic system.

Pharmacol Biochem Behav

67

:

207

214

166

Tanaka

MY

,

Kohno

Y

,

Nakagawa

R

,

Ida

Y

,

Takeda

S

,

Nagasaki

N

,

Noda

Y

1983

Regional characteristics of stress-induced increases in brain noradrenaline release in rats.

Pharmacol Biochem Behav

19

:

543

547

167

Saavedra

JM

,

Kvetnansky

R

,

Kopin

IJ

1979

Adrenaline, noradrenaline and dopamine levels in specific brain stem areas of acutely immobilized rats.

Brain Res

160

:

271

280

168

Tanaka

T

,

Yokoo

H

,

Mizoguchi

K

,

Yoshida

M

,

Tsuda

A

,

Tanaka

M

1991

Noradrenaline release in the rat amygdala is increased by stress: studies with intracerebral microdialysis

.

Brain Res

544

:

174

176

169

Cullinan

WE

,

Herman

JP

,

Watson

SJ

1993

Ventral subicular interaction with the hypothalamic paraventricular nucleus: evidence for a relay in the bed nucleus of the stria terminalis.

J Comp Neurol

332

:

1

20

170

Herman

JP

,

Cullinan

WE

,

Watson

SJ

1994

Involvement of the bed nucleus of the stria terminalis in tonic regulation of paraventricular hypothalamic CRH and AVP mRNA expression.

J Neuroendocrinol

6

:

433

442

171

Pacak

K

,

McCarty

R

,

Palkovits

M

,

Kopin

IJ

,

Goldstein

DS

1995

Effects of immobilization on in vivo release of norepinephrine in the bed nucleus of the stria terminalis in conscious rats.

Brain Res

688

:

242

246

172

Dunn

JD

1987

Plasma corticosterone responses to electrical stimulation of the bed nucleus of the stria terminalis.

Brain Res

407

:

327

331

173

Feldman

S

,

Conforti

N

,

Saphier

D

1990

The preoptic area and bed nucleus of the stria terminalis are involved in the effects of the amygdala on adrenocortical secretion.

Neuroscience

37

:

775

779

174

Sun

N

,

Roberts

L

,

Cassell

MD

1991

Rat central amygdaloid nucleus projections to the bed nucleus of the stria terminalis.

Brain Res Bull

27

:

651

662

175

Olschowka

JA

,

Molliver

ME

,

Grzanna

R

,

Rice

FL

,

Coyle

JT

1981

Ultrastructural demonstrations of noradrenergic synapses in the rat central nervous system by dopamine-β-hydroxylase immunocytochemistry.

J Histochem Cytochem

29

:

271

280

176

Liposits

Z

,

Phelix

C

,

Paull

WK

1986

Electron microscopic analysis of tyrosine hydroxylase, dopamine-β-hydroxylase and phenylethanolamine-N-methyltransferase immunoreactive innervation of the hypothalamic paraventricular nucleus in the rat.

Histochemistry

84

:

105

120

177

Sawchenko

PE

1988

Effects of catecholamine-depleting medullary knife cuts on corticotropin-releasing factor and vasopressin immunoreactivity in the hypothalamus of normal and steroid manipulated rats.

Neuroendocrinology

48

:

459

470

178

Alonso

G

,

Szafarczyk

A

,

Balmefrezol

M

,

Assenmacher

I

1986

Immunocytochemical evidence for stimulatory control by the ventral noradrenergic bundle of parvocellular neurons of the paraventricular nucleus secreting corticotropin-releasing hormone and vasopressin in rats.

Brain Res

397

:

297

307

179

Sawchenko

PE

,

Swanson

LW

1981

Central noradrenergic pathways for the integration of hypothalamic neuroendocrine and autonomic responses.

Science

214

:

685

687

180

Weidenfeld

J

,

Feldman

S

1991

Effect of hypothalamic norepinephrine depletion on median eminence CRF-41 content and serum ACTH in control and adrenalectomized rats.

Brain Res

542

:

204

210

181

Szafarczyk

A

,

Alonso

G

,

Ixart

G

,

Malaval

F

,

Assenmacher

I

1985

Diurnal-stimulated and stress-induced ACTH release inrats is mediated by ventral noradrenergic bundle

.

Am J Physiol

249

:

E219

E226

182

Kiss

A

,

Palkovits

M

,

Aguilera

G

1996

Neural regulation of corticotropin releasing hormone (CRH) and CRH receptor mRNA in the hypothalamic paraventricular nucleus in the rat.

J Neuroendocrinol

8

:

103

112

183

Burstein

R

,

Cliffer

KD

,

Giesler

GJJ

1987

Direct somatosensory projections from the spinal cord to the hypothalamus and telencephalon.

J Neurosci

7

:

4159

4164

184

Burstein

R

,

Cliffer

KD

,

Giesler

GJJ

1990

Cells of origin of the spinothalamic tract in the rat.

J Comp Neurol

291

:

329

344

185

Cliffer

K

,

Burstein

R

,

Giesler

GJJ

1991

Distributions of spinothalamic, spinohypothalamic, and spinotelencephalic fibers revealed by anterograde transport of PHA-L in rats.

J Neurosci

11

:

852

868

186

Kevetter

GA

,

Willis

WD

1982

Spinothalamic cells in the rat lumbar cord with collaterals to the medullary reticular formation.

Brain Res

238

:

181

185

187

Menétrey

D

,

Roudier

F

,

Besson

JM

1983

Spinal neurons reaching the lateral reticular nucleus as studied in the rat by retrograde transport of horseradish peroxidase.

J Comp Neurol

220

:

439

452

188

Shokunbi

MT

,

Hrycyshin

AW

,

Flumerfelt

BA

1985

Spinal projections to the lateral reticular nucleus in the rat: a retrograde labeling study using horseradish peroxidase.

J Comp Neurol

239

:

216

226

189

Lima

D

,

Coimbra

A

1989

Morphological types of spinomesencephalic neurons in the marginal zone (lamina I) of the rat spinal cord as shown after retrograde labelling with cholera toxin subunit B.

J Comp Neurol

279

:

327

339

190

Lima

D

,

Mendes-Ribeiro

JA

,

Coimbra

A

1991

The spino-latero-reticular system of the rat: projections from the superficial dorsal horn and structural characterization of marginal neurons involved.

Neuroscience

45

:

137

152

191

Tavares

I

,

Lima

D

,

Coimbra

A

1993

Neurons in the superficial dorsal horn of the rat spinal cord projecting to the medullary ventrolateral reticular formation express c-fos after noxious stimulation of the skin.

Brain Res

623

:

278

286

192

Esteves

F

,

Lima

D

,

Coimbra

A

1993

Structural types of spinal cord marginal (lamina-1) neurons projecting to the nucleus of the tractus solitarius in the rat.

Somatosens Mot Res

10

:

203

216

193

Otake

K

,

Reis

DJ

,

Ruggiero

DA

1994

Afferents to the midline thalamus issue collaterals to the nucleus tractus solitarii: an anatomical basis for thalamic and visceral reflex integration.

J Neurosci

14

:

5694

5707

194

Aicher

SA

,

Kurucz

OS

,

Reis

DJ

,

Milner

TA

1995

Nucleus tractus solitarius efferent terminals synapse on neurons in the caudal ventrolateral medulla that project to the rostral ventrolateral medulla.

Brain Res

693

:

51

63

195

Chan

RKW

,

Peto

CA

,

Sawchenko

PE

1995

A1 catecholamine cell group: fine structure and synaptic input from the nucleus of the solitary tract.

J Comp Neurol

351

:

62

80

196

Hancock

MB

1988

Evidence for direct projections from the nucleus of the solitary tract onto medullary adrenaline cells.

J Comp Neurol

276

:

460

467

197

Ross

CA

,

Ruggiero

DA

,

Reis

DJ

1985

Projections from the nucleus tractus solitarii to the rostral ventrolateral medulla.

J Comp Neurol

242

:

511

534

198

Kalivas

PW

1993

Neurotransmitter regulation of dopamine neurons in the ventral tegmental area.

Brain Res Rev

18

:

75

113

199

Mogenson

GJ

,

Jones

DL

,

Yim

CY

1980

From motivation to action: functional interface between the limbic system and the motor system.

Prog Neurobiol

14

:

69

97

200

Bergmann

C

1845

[Nichtchemischer Beitrag zur Kritik der Lehre von Calor animalis]. Arch Anat Physiol

300

319

201

Tscheschichin

J

1866

Zur Lehre von der thierischen Wärme.

Arch Anat Physiol

11

:

151

179

202

Liebermeister

Cv

1860

Physiologische Untersuchungen über die quantitativen Veränderungen der Warmeproduction. Arch Anat Physiol 520–541,

589

623

203

Teague

RS

,

Ranson

SW

1936

The role of the anterior hypothalamus in temperature regulation.

Am J Physiol

117

:

562

570

204

Isenschmid

R

,

Krehl

L

1912

Uber den Einfluss des Gehirns auf die Warmeregulation.

Arch exp Path Pharmacol

70

:

149

182

205

Isenschmid

R

,

Schnitzler

W

1914

Beitrage zur Lokalisation des der Warmeregulation vorstehenden Zentralapparates im Zwischenhirn.

Arch exp Path Pharmacol

76

:

202

233

206

Keller

AD

,

Hare

WK

1932

The hypothalamus and heat regulation.

Proc Soc Exp Biol Med

29

:

1069

1070

207

Clark

G

,

Magoun

HW

,

Ranson

SW

1939

Hypothalamic regulation of body temperature.

J Neurophysiol

2

:

61

80

208

Bligh

J

1966

The thermosensitivity of the hypothalamus and thermoregulation in mammals.

Biol Rev

41

:

317

367

209

Pachomov

N

1962

The effects of posterior and anterior hypothalamic lesions on the maintenance of body temperature in the rat.

J Neuropathol Exp Neurol

21

:

450

460

210

Satinoff

E

,

Rutstein

J

1970

Behavioral thermoregulation in rats with anterior hypothalamic lesions.

J Comp Physiol Psychol

71

:

77

82

211

Satinoff

E

,

Shan

SY

1971

Loss of behavioral thermoregulation after lateral hypothalamic lesions in rats.

J Comp Physiol Psychol

77

:

302

312

212

Satinoff

E

1978

Neural organization and evolution of thermal regulation in mammals.

Science

201

:

16

22

213

Lipton

JM

1971

Thermal stimulation of the medulla alters behavioral temperature regulation.

Brain Res

26

:

439

442

214

Lipton

JM

1973

Thermosensitivity of medulla oblongata in control of body temperature.

Am J Physiol

224

:

890

897

215

Lipton

JM

,

Dwyer

P

,

Fossler

DE

1974

Effects of brainstem lesions on temperature regulation in hot and cold environments.

Am J Physiol

226

:

1356

1365

216

Lipton

JM

,

Clark

WG

1986

Neurotransmitters in temperature control.

Annu Rev Physiol

48

:

613

623

217

Mosso

JA

,

Kruger

L

1972

Spinal trigeminal neurons excited by noxious and thermal stimuli.

Brain Res

38

:

206

210

218

Boulant

JA

,

Bignall

KE

1973

Hypothalamic neuronal responses to peripheral and deep-body temperatures.

Am J Physiol

225

:

1371

1374

219

Boulant

JA

,

Hardy

JD

1974

The effect of spinal and skin temperatures on the firing rate and thermosensitivity of preoptic neurones.

J Physiol (Lond)

240

:

639

660

220

Boulant

JA

,

Demieville

HN

1977

Responses of thermosensitive preoptic and septal neurons to hippocampal and brain stem stimulation.

J Neurophysiol

40

:

1356

1368

221

Boulant

JA

1980

Hypothalamic control of thermoregulation. In:

Morgane

PJ

,

Panksepp

J

, eds. Handbook of the hypothalamus, vol.

2. New York

:

Marcell Decker

;

1

82

222

Boulant

JA

,

Dean

JB

1986

Temperature receptors in the central nervous system.

Annu Rev Physiol

48

:

639

654

223

Boulant

JA

1992

Properties of hypothalamic temperature sensitive and insensitive neurones.

Physiol Res

41

:

83

84

224

Boulant

JA

1998

Cellular mechanisms of temperature sensitivity in hypothalamic neurons.

Prog Brain Res

115

:

3

8

225

Trzcinka

GP

,

Lipton

JM

,

Hawkins

M

,

Clark

WG

1977

Effects on temperature of morphine injected into the preoptic/anterior hypothalamus, medulla oblongata, and peripherally in unrestrained and restrained rats.

Proc Soc Exp Biol Med

156

:

523

526

226

Hinckel

P

,

Schroder-Rosenstock

K

1981

Responses of pontine units to skin-temperature changes in the guinea-pig.

J Physiol (Lond)

314

:

189

194

227

Miyata

S

,

Ishiyama

M

,

Shido

O

,

Nakashima

T

,

Shibata

M

,

Kiyohara

T

1995

Central mechanism of neural activation with cold acclimation of rats using fos immunohistochemistry.

Neurosci Res

22

:

209

218

228

Hunt

SP

,

Pini

A

,

Evan

G

1987

Induction of c-fos-like protein in spinal cord neurons following sensory stimulation.

Nature

328

:

632

634

229

Bullitt

E

1990

Expression of c-fos-like protein as a marker for neuronal activity following noxious stimulation in the rat.

J Comp Neurol

296

:

517

530

230

Scammell

TE

,

Price

KJ

,

Sagar

SM

1993

Hyperthermia induce c-fos expression in the preoptic area.

Brain Res

618

:

303

307

231

Patronas

P

,

Horowitz

M

,

Simon

E

,

Gerstberger

R

1998

Differential stimulation of c-fos expression in hypothalamic nuclei of the rat brain during short-term heat acclimation and mild dehydration.

Brain Res

798

:

127

139

232

Kiyohara

T

,

Miyata

S

,

Nakamura

T

,

Shido

O

,

Nakashima

T

,

Shibata

M

1995

Differences in Fos expression in the rat brains between cold and warm ambient exposures

.

Brain Res Bull

38

:

193

201

233

Zoeller

RT

,

Kabeer

N

,

Albers

HE

1990

Cold exposure elevates cellular levels of messenger ribonucleic acid encoding thyrotropin-releasing hormone in paraventricular nucleus despite elevated levels of thyroid hormones.

Endocrinology

127

:

2955

2962

234

Hauger

RL

,

Milan

MA

,

Lorang

M

,

Harwood

JP

,

Aguilera

G

1988

Corticotropin-releasing factor receptors and pituitary adrenal responses during immobilization stress.

Endocrinology

123

:

396

405

235

Hauger

RL

,

Lorang

M

,

Irwin

M

,

Aguilera

G

1990

CRF receptor regulation and sensitization of ACTH responses to acute ether stress during chronic immobilization stress.

Brain Res

532

:

34

40

236

Angulo

JA

,

Ledoux

M

,

McEwen

BS

1991

Genomic effects of cold and isolation stress on magnocellular vasopressin mRNA-containing cells in the hypothalamus of the rat.

J Neurochem

56

:

2033

2038

237

McEwen

BS

,

Angulo

J

,

Cameron

H

1992

Paradoxical effects of adrenal steroids on the brain: protection vs. degeneration.

Biol Psychiatry

31

:

177

199

238

McEwen

BS

,

Conrad

CD

,

Kuroda

Y

,

Frankfurt

M

,

Magarinos

AM

,

McKittrick

C

1997

Prevention of stress-induced morphological and cognitive consequences. Eur Neuropsychopharmacol 7 (Suppl 3):

323

328

239

Magarinos

AM

,

McEwen

BS

1995

Stress-induced atrophy of apical dendrites of hippocampal CA3c neurons: involvement of glucocorticoid secretion and excitatory amino acid receptors.

Neuroscience

69

:

89

98

240

Ishikawa

K

,

Kakegawa

T

,

Suzuki

M

1984

Role of hypothalamic paraventricular nucleus in the secretion of thyrotropin under adrenergic and cold-stimulated conditions in the rat.

Endocrinology

114

:

352

358

241

Lin

MT

,

Wang

PS

,

Chuang

J

,

Fan

LJ

,

Won

SJ

1989

Cold stress or a pyrogenic substance elevates thyrotropin-releasing hormone levels in the rat hypothalamus and induces thermogenic reactions.

Neuroendocrinology

50

:

177

181

242

Rondeel

JMM

,

de Greef

WJ

,

Hop

WCJ

,

Rowland

DL

,

Visser

TJ

1991

Effect of cold exposure on the hypothalamic release of thyrotropin-releasing hormone and catecholamines.

Neuroendocrinology

54

:

477

481

243

Fukuhara

K

,

Kvetnansky

R

,

Cizza

G

,

Pacak

K

,

Ohara

H

,

Goldstein

DS

,

Kopin

IJ

1996

Interrelations between sympathoadrenal system and hypothalamo-pituitary-adrenocortical/thyroid systems in rats exposed to cold stress.

J Neuroendocrinol

8

:

533

541

244

Arancibia

S

,

Tapia-Arancibia

L

,

Roussel

JP

,

Assenmacher

I

,

Astier

H

1986

Effects of morphine on cold-induced TRH release from the median eminence of unanesthetized rats.

Life Sci

38

:

59

66

245

Arancibia

S

,

Tapia-Arancibia

L

,

Assenmacher

I

,

Astier

H

1983

Direct evidence of short-term cold-induced TRH release in the median eminence of unanesthetized rats.

Neuroendocrinology

37

:

225

228

246

Arancibia

S

,

Rage

F

,

Astier

H

,

Tapia-Arancibia

L

1996

Neuroendocrine and autonomous mechanisms underlying thermoregulation in cold environment.

Neuroendocrinology

64

:

257

267

247

Thoenen

H

1970

Induction of tyrosine hydroxylase in peripheral and central adrenergic neurones by cold-exposure of rats.

Nature

228

:

861

862

248

Richard

F

,

Faucon-Biguet

N

,

Labutu

R

,

Rollet

D

,

Mallet

J

,

Buda

M

1988

Modulation of tyrosine hydroxylase gene expression in rat brain and adrenals by exposure to cold.

J Neurosci Res

20

:

32

37

249

Rostrup

M

,

Kjeldsen

SE

,

Eide

IK

1990

Awareness of hypertension increases blood pressure and sympathetic responses to cold pressure test.

Am J Hypertens

3

:

912

917

250

Konarska

M

,

Stewart

RE

,

McCarty

R

1990

Predictability of chronic intermittent stress: effects on sympathetic-adrenal medullary responses of laboratory rats.

Behav Neural Biol

53

:

231

243

251

Konarska

M

,

Stewart

RE

,

McCarty

R

1989

Sensitization of sympathetic-adrenal medullary responses to a novel stressor in chronically stressed laboratory rats.

Physiol Behav

46

:

129

135

252

Fukuhara

K

,

Kvetnansky

R

,

Weise

VK

,

Ohara

H

,

Yoneda

R

,

Goldstein

DS

,

Kopin

IJ

1996

Effects of continuous and intermittent cold (SART) stress on sympathoadrenal system activity in rats.

J Neuroendocrinol

8

:

65

72

253

Suemaru

S

,

Hashimoto

K

,

Ota

Z

1985

Brain corticotropin-releasing factor (CRF) and catecholamine responses in acutely stressed rats.

Endocrinol Jpn

32

:

709

718

254

Bhatnagar

S

,

Mitchell

JB

,

Betito

K

,

Boksa

P

,

Meaney

MJ

1995

Effects of chronic intermittent cold stress on pituitary adrenocortical and sympathetic adrenomedullary functioning.

Physiol Behav

57

:

633

639

255

Frank

SM

,

Raja

SN

,

Bulcao

CF

,

Goldstein

DS

1999

Relative contribution of core and cutaneous temperatures to thermal comfort and autonomic responses in humans.

J Appl Physiol

86

:

1588

1593

256

Marino

F

,

Sockler

JM

,

Fry

JM

1998

Thermoregulatory, metabolic and sympathoadrenal responses to repeated brief exposure to cold.

Scand J Clin Lab Invest

58

:

537

545

257

Ohtani

N

,

Sugano

T

,

Ohta

M

1999

Alterations in monoamines and GABA in the ventromedial and paraventricular nuclei of the hypothalamus following cold exposure: a reduction in noradrenaline induces hyperphagia.

Brain Res

842

:

6

14

258

Bruck

K

,

Hinckel

P

1980

Thermoregulatory noradrenergic and serotonergic pathways to hypothalamic units.

J Physiol (Lond)

304

:

193

202

259

Szelenyi

Z

,

Hinckel

P

1987

Changes in cold- and heat-defence following electrolytic lesions of raphe nuclei in the guinea-pig.

Pflüger’s Arch Pathol

409

:

175

181

260

Yang

H

,

Wu

SV

,

Ishikawa

T

,

Tache

Y

1994

Cold exposure elevates thyrotropin-releasing hormone gene expression in medullary raphe nuclei: relationship with vagally mediated gastric erosions.

Neuroscience

61

:

655

663

261

Tanaka

M

,

Nishikawa

T

,

Kohno

Y

,

Nagasaki

N

,

Noda

Y

,

Inanaga

K

1981

Hypothermia and gastric lesions in rats exposed to immobilization stress.

Kurume Med J

28

:

247

253

262

Lipton

JM

,

Avery

DD

,

Marotto

DR

1970

Determinants of behavioral thermoregulation against heat: thermal intensity and skin temperature levels.

Physiol Behav

5

:

1083

1088

263

Kanosue

K

,

Nakayama

T

,

Tanaka

H

,

Yanase

M

,

Yasuda

H

1990

Modes of action of local hypothalamic and skin thermal stimulation on salivary secretion in rats.

J Physiol (Lond)

424

:

459

471

264

Kanosue

K

,

Hosono

T

,

Zhang

Y-H

,

Chen

X-M

1998

Neuronal networks controlling thermoregulatory effectors. In:

Sharma

HS

,

Westman

J

, eds. Brain function in hot environment, vol.

115. Amsterdam

:

Elsevier

;

49

62

265

Zhang

YH

,

Yamada

K

,

Hosono

T

,

Chen

XM

,

Shiosaka

S

,

Kanosue

K

1997

Efferent neuronal organization of thermoregulatory vasomotor control.

Ann NY Acad Sci

813

:

117

122

266

Simerly

RB

,

Swanson

LW

1986

The organization of neuronal inputs to the medial preoptic nucleus of the rat.

J Comp Neurol

246

:

312

342

267

Lin

LS

,

Chiu

WT

,

Shih

CJ

,

Lin

MT

1989

Influence of thermal stress and various agents on the brain edema formation in rats following a cryogenic brain lesion.

Chin J Physiol

32

:

41

47

268

Kanosue

K

,

Yanase-Fujiwara

M

,

Hosono

T

1994

Hypothalamic network for thermoregulatory vasomotor control

.

Am J Physiol

267

:

R283

R288

269

Murphy

AZ

,

Rizvi

TA

,

Ennis

M

,

Shipley

MT

1999

The organization preoptic-medullary circuits in the male rat: evidence for interconnectivity of neural structures involved in reproductive behavior, antinociception and cardiovascular regulation.

Neuroscience

91

:

1103

1116

270

Tóth

IE

,

Boldogköi

Z

,

Medveczky

I

,

Palkovits

M

1999

Lacrimal preganglionic neurons form a subdivision of the superior salivatory nucleus of rat: transneuronal labelling by pseudorabies virus.

J Auton Nerv Syst

77

:

45

54

271

Smith

JE

,

Jansen

AS

,

Gilbey

MP

,

Loewy

AD

1998

CNS cell groups projecting to sympathetic outflow of tail artery: neural circuits involved in heat loss in the rat.

Brain Res

786

:

153

164

272

Fish

HR

,

Chernow

B

,

O’Brian

J

1986

Endocrine and neurophysiologic responses of the pituitary to insulin-induced hypoglycemia: a review.

Metabolism

35

:

763

780

273

Keller-Wood

ME

,

Dallman

MF

1984

Corticosteroid inhibition of ACTH secretion.

Endocr Rev

5

:

1

24

274

Gagner

JP

,

Gauthier

S

,

Sourkes

TL

1985

Descending spinal pathways mediating the responses of adrenal tyrosine hydroxylase and catecholamines to insulin and 2-deoxyglucose

.

Brain Res

325

:

187

197

275

Borg

WP

,

During

MJ

,

Sherwin

RS

,

Borg

MA

,

Brines

ML

,

Shulman

GI

1994

Ventromedial hypothalamic lesions in rats suppress counterregulatory responses to hypoglycemia.

J Clin Invest

93

:

1677

1682

276

Borg

WP

,

Sherwin

RS

,

During

MJ

,

Borg

MA

,

Shulman

GI

1995

Local ventromedial hypoyhalamus glucopenia triggers counterregulatory hormone release.

Diabetes

44

:

180

184

277

Barris

RW

,

Ingram

WR

1936

The effects of experimental hypothalamic lesions upon blood sugar.

Am J Physiol

114

:

555

561

278

Oomura

Y

,

Ono

T

,

Ooyama

H

,

Wayner

MJ

1969

Glucose and osmosensitive neurones of the rat hypothalamus

.

Nature

222

:

282

284

279

Bahjaoui-Bouhaddi

M

,

Fellmann

D

,

Bugnon

C

1994

Induction of fos-immunoreactivity in prolactin-like containing neurons of the rat lateral hypothalamus after insulin injection.

Neurosci Lett

168

:

11

15

280

Oomura

Y

,

Ooyama

H

,

Sugimori

M

,

Nakamura

T

,

Yamada

Y

1974

Glucose inhibition of the glucose-sensitive neurone in the rat lateral hypothalamus.

Nature

247

:

284

286

281

Niimi

M

,

Sato

M

,

Tamaki

Y

,

Wada

Y

,

Takahara

J

,

Kawanishi

K

1995

Induction of Fos protein in the rat hypothalamus elicited by insulin-induced hypoglycemia.

Neurosci Res

23

:

361

364

282

Sakurai

T

,

Amemiya

A

,

Ishii

M

,

Matsuzaki

I

,

Chemelli

RM

,

Tanaka

H

,

Williams

SC

,

Richardson

JA

,

Kozlowski

GP

,

Wilson

S

,

Arch

JRS

,

Buckingham

RE

,

Haynes

AC

,

Carr

SA

,

Annan

RS

,

McNulty

DE

,

Liu

W-S

,

Terrett

JA

,

Elshourbagy

NA

,

Bergsma

DJ

,

Yanagisawa

M

1998

Orexins and orexin receptors: a family of hypothalamic neuropeptides and G protein-coupled receptors that regulate feeding behavior.

Cell

92

:

573

585

283

Dube

MG

,

Kalra

SP

,

Kalra

PS

1999

Food intake elicited by central administration of orexin/hypocrtins: identification of hypothalamic sites of action.

Brain Res

842

:

473

477

284

Moriguchi

T

,

Sakurai

T

,

Nambu

T

,

Yanagisawa

M

,

Goto

K

1999

Neurons containing orexin in the lateral hypothalamic area of the adult rat brain are activated by insulin-induced acute hypoglycemia.

Neurosci Lett

264

:

101

104

285

Marks

JL

,

Porte

DJR

,

Stahl

WL

,

Baskin

DG

1990

Localization of insulin receptor mRNA in rat brain by in situ hybridization.

Endocrinology

127

:

3234

3236

286

Griffond

B

,

Deray

A

,

Bahjaoui-Bouhaddi

M

,

Colard

C

,

Bugnon

C

,

Fellmann

D

1994

Induction of Fos-like immunoreactivity in rat oxytocin neurons following insulin injections.

Neurosci Lett

178

:

119

123

287

Murao

K

,

Sato

M

,

Mizobuchi

M

,

Niimi

M

,

Ishida

T

,

Takahara

J

1994

Acute effects of hypoglycemia on hypothalamic growth hormone-releasing hormone and somatostatin gene expression in the rat.

Endocrinology

134

:

418

423

288

Porter

JP

,

Bokil

HS

1997

Effect of intracerebroventricular and intravenous insulin on Fos-immunoreactivity in the rat brain.

Neurosci Lett

224

:

161

164

288A

Carrasco

M

,

Portillo

F

,

Larsen

PJ

,

Vallo

JJ

2001

Insulin and glucose administration stimulates Fos expression in neurons of the paraventricular nucleus that project to autonomic preganglionic structures.

J Neuroendocrinol

13

:

339

346

289

Rusnak

M

,

Jelokova

J

,

Vietor

I

,

Sabban

EL

,

Kvetnansky

R

1998

Different effects of insulin and 2-deoxy-d-glucose administration on tyrosine hydroxylase gene expression in the locus coeruleus and the adrenal medulla in rats.

Brain Res Bull

46

:

447

452

290

Gaillet

S

,

Malaval

F

,

Barbanel

G

,

Pelletier

G

,

Assenmacher

I

,

Szafarczyk

A

1991

Inhibitory interactions betweenα 2-adrenergic and opioid but not NPY mechanisms controlling the CRF-ACTH axis in the rat.

Regul Pept

36

:

249

261

291

Kárteszi

M

,

Dallman

MF

,

Makara

GB

,

Stark

E

1982

Regulation of the adrenocortical response to insulin-induced hypoglycemia.

Endocrinology

111

:

535

541

292

Suda

T

,

Nakano

Y

,

Towaza

F

,

Sumitomo

T

,

Sato

Y

,

Yamada

M

,

Demura

H

1992

The role of corticotropin-releasing factor and vasopressin in hypoglycemia-induced proopiomelanocortin gene expression in the rat anterior pituitary gland.

Brain Res

579

:

303

308

293

Suda

T

,

Tomori

N

,

Sumimoto

T

,

Nakagami

Y

,

Tozawa

F

,

Ushiyama

T

,

Yamada

M

,

Sato

A

,

Demura

H

,

Shinzumi

K

1988

Feedback- and hypoglycemia-induced regulation of secretion and synthesis of ACTH and CRF. In:

Imura

H

, ed. Neuroendocrine control of the hypothalamo-pituitary system.

Tokyo

:

Japanese Scientific Society Press

;

23

30

294

Itoi

K

,

Horiba

N

,

Tozawa

F

,

Sakai

Y

,

Sakai

K

,

Abe

K

,

Demura

H

,

Suda

T

1996

Major role of 3′,5′-cyclic adenosine monophosphate-dependent protein kinase A pathway in corticotropin-releasing factor gene expression in the rat hypothalamus in vivo.

Endocrinology

137

:

2389

2396

295

Brown

ER

,

Sawchenko

PE

1997

Hypophysiotropic CRF neurons display a sustained immediate-early gene response to chronic stress but not to adrenalectomy.

J Neuroendocrinol

9

:

307

316

296

Berkenbosch

F

,

De Goeji

DCE

,

Tilders

FJH

1989

Hypoglycemia enhances turnover of corticotropin-releasing factor and vasopressin in the zona externa of the rat median eminence.

Endocrinology

125

:

28

34

297

Tozawa

F

,

Suda

T

,

Yamada

M

,

Ushiyama

T

,

Tomori

N

,

Sumitomo

T

,

Nakagami

Y

,

Demura

H

,

Shizume

K

1988

Insulin-induced hypoglycemia increases proopiomelanocortin messenger ribonucleic acid levels in rat anterior pituitary gland.

Endocrinology

122

:

1231

1235

298

Robinson

BG

,

Mealy

K

,

Wilmore

DW

,

Majzoub

JA

1992

The effect of insulin-induced hypoglycemia on gene expression in the hypothalamic-pituitary-adrenal axis of the rat

.

Endocrinology

130

:

920

925

299

Sumomito

T

,

Suda

T

,

Tomori

N

,

Yajima

F

,

Nakagami

Y

,

Ushiyama

T

,

Demura

H

,

Shizume

K

1987

Immunoreactive corticotropin-releasing factor in rat plasma.

Endocrinology

120

:

1391

1396

300

Suda

T

,

Tozawa

F

,

Yamada

M

,

Ushiyama

T

,

Tomori

N

,

Sumimoto

T

,

Nakagami

Y

,

Demura

H

,

Shizume

K

1988

Insulin-induced hypglycemia increases corticotropin-releasing factor messenger ribonucleic acid levels in rat hypothalamus.

Endocrinology

123

:

1371

1375

301

Jezova

D

,

Kvetnansky

R

,

Kovács

K

,

Oprsalova

Z

,

Vigas

M

,

Makara

GB

1987

Insulin-induced hypoglycemia activates the release of adrenocorticotrophin predominantly through via central and propranolol insensitive systems.

Endocrinology

120

:

409

415

302

Weidenfeld

J

,

Siegel

RA

,

Feldman

S

,

Conforti

N

,

Chowers

I

1982

ACTH and corticosterone secretion following insulin in intact and in variously hypothalamic deafferentated male rats.

Exp Brain Res

48

:

306

308

303

Mezey

E

,

Reisine

TD

,

Brownstein

MJ

,

Palkovits

M

,

Axelrod

J

1984

β-Adrenergic mechanism of insulin-induced adrenocorticotropin release from the anterior pituitary.

Science

226

:

1085

1087

304

Plotsky

PM

1985

Hypophysiotrophic regulation of adenohypophyseal adrenocorticotropin secretion.

Fed Proc

44

:

207

213

305

Fisher

BM

,

Bayliss

PH

,

Frier

BM

1987

Plasma oxytocin, arginine vasopressin and atrial natriuretic peptide responses to insulin-induced hypoglycemia in man.

Clin Endocrinol (Oxf)

26

:

179

185

306

Chiodera

P

,

Volpi

R

,

Capretti

L

,

Speroni

G

,

Marcato

A

,

Rossi

G

,

Coiro

V

1992

Hypoglycemia-induced arginine vasopressin and oxytocin release is mediated by glucoreceptors located inside the blood-brain barrier.

Neuroendocrinology

55

:

655

659

307

Whitnall

M

1989

Stress selectively activates the vasopressin- containing subset of corticotropin-releasing hormone neurons.

Neuroendocrinology

50

:

702

707

308

Gillies

GE

,

Linton

EA

,

Lowry

PJ

1982

Corticotropin-releasing activity of the new CRF is potentiated several times by vasopressin.

Nature

289

:

676

679

309

Vale

W

,

Vaughan

J

,

Smith

M

,

Yamomoto

G

,

Rivier

J

,

Rivier

C

1983

Effects of synthetic ovine corticotropin-releasing factor, glucocorticoids, catecholamines, neurohypophysial peptides, and other substances on cultured corticotropic cells.

Endocrinology

113

:

1121

1131

310

Rivier

C

,

Vale

W

1983

Interaction of corticotropin releasing factor and arginie vasopressin on adrenocorticotropin secretion in vivo.

Endocrinology

113

:

939

942

311

Paulmyer-Lacroix

O

,

Anglade

G

,

Grino

M

1994

Insulin-induced hypoglycemia increases colocalization of corticotrophin-releasing factor and arginie vasopresin mRNAs in the rat hypothalamic paraventricular nucleus.

J Mol Endocrinol

13

:

313

320

312

Chen

MD

,

O’Byrne

KT

,

Chiappini

SE

,

Hotchkiss

J

,

Knobil

E

1992

Hypoglycemic ’stress’ and gonadotropin-releasing hormone pulse generator activity in the rhesus monkey: role of the ovary.

Neuroendocrinology

56

:

666

673

313

Bucholtz

DC

,

Vidwans

NM

,

Herbosa

CG

,

Schillo

KK

,

Foster

DL

1995

Metabolic interfaces between growth and reproduction. V. Pulsatile lutenizing hormone secretion is dependant on glucose availability.

Endocrinology

137

:

601

607

314

Chen

M

,

Ordog

T

,

O’Byrne

K

,

Goldsmith

J

,

Connaughton

M

,

Knobil

E

1996

The insulin hypoglycemia induced inhibition of gonadotrophin-releasing hormone pulse generator activity in the rhesus monkey: roles of vasopressin and corticotropin-releasing factor.

Endocrinology

137

:

2012

2021

315

MacLusky

N

,

Naftolin

F

,

Léránth

C

1988

Immunocytochemical evidence for direct synaptic connections between corticotrophin-releasing factor and gonadotropin-releasing hormone-containing neurons in the preoptic area of the rat.

Brain Res

439

:

391

395

316

van Vugt

DA

,

Washburn

DLS

,

Farley

AE

,

Reid

RL

1997

Hypoglycemia-induced inhibition of LH and stimulation of ACTH secretion in the rhesus monkey is blocked by alprazolam.

Neuroendocrinology

65

:

344

352

317

Horvath

TL

,

Diano

S

,

van den Pol

AN

1999

Synaptic interaction between hypocretin (orexin) and neuropeptide Y cells in the rodent and primate hypothalamus: a novel circuit implicated in the metabolic and endocrine regulation.

J Neurosci

19

:

1072

1078

318

Kalra

SP

,

Dube

MG

,

Pu

S

,

Xu

B

,

Horvath

TL

,

Kalra

PS

1999

Interacting appetite-regulating pathways in the hypothalamic regulation of body weight.

Endocr Rev

20

:

68

100

319

Zhang

Y

,

Proenca

R

,

Maffei

M

,

Barone

M

,

Leopold

L

,

Friedman

JM

1994

Positional cloning of the mouse obese gene and its human homologue.

Nature

372

:

425

432

320

Elias

CF

,

Lee

C

,

Kelly

J

,

Aschkenasi

C

,

Ahima

RS

,

Couceyro

PR

,

Kuhar

MJ

,

Saper

CB

,

Elmquist

JK

1998

Leptin activates hypothalamic CART neurons projecting to the spinal cord.

Neuron

21

:

1375

1385

321

Fritsche

A

,

Wahl

HG

,

Metzinger

E

,

Renn

W

,

Kellerer

M

,

Haring

H

,

Stumwoll

M

1998

Evidence for inhibition of leptin secretion by catecholamines in man.

Exp Clin Endocrinol Diabetes

106

:

415

418

322

Heiman

ML

,

Ahima

RS

,

Craft

LS

,

Schoner

B

,

Stephens

TW

,

Flier

JS

1997

Leptin inhibition of the hypothalamic-pituitary adrenal axis response to stress.

Endocrinology

138

:

3859

3863

323

Kristensen

P

,

Judge

ME

,

Thim

L

,

Ribel

U

,

Christjansen

KN

,

Wulff

BS

,

Clausen

JT

,

Jensen

PB

,

Madesen

OD

,

Wrang

N

,

Larsen

PJ

,

Hastrup

S

1998

Hypothalamic CART is a new anorectic peptide regulated by leptin.

Nature

393

:

72

76

324

Nye

EJ

,

Bornstein

SR

,

Grice

JE

,

Tauchnitz

R

,

Hockings

GI

,

Strakosch

CR

,

Jackson

RV

,

Torpy

DJ

2000

Interactions between the stimulated hypothalamic-pituitary-adrenal axis and leptin in humans.

J Neuroendocrinol

12

:

141

145

325

Schmitz

O

,

Fisker

S

,

Orskov

L

,

Hove

KY

,

Nyholm

B

,

Moller

N

1997

Effects of hyperinsulinaemia and hypoglycaemia on circulating leptin concentrations in healthy lean males.

Diabetes Metab

23

:

80

83

326

Wang

Q

,

Bing

C

,

Al-Barazanji

K

,

Mossakowaska

DE

,

Wang

X-M

,

McBay

DL

,

Neville

WA

,

Taddayon

M

,

Pickavance

L

,

Dryden

S

,

Thomas

MEA

,

McHale

MT

,

Gloyer

IS

,

Wilson

S

,

Buckingham

R

,

Arch

JRS

,

Trayhurn

P

,

Williams

G

1997

Interactions between leptin and hypothalamic neuropeptide Y neurons in the control of food intake and energy homeostasis in the rat.

Diabetes

46

:

335

341

327

Kask

A

,

Schiöth

HB

,

Mutulis

F

,

Wikberg

JES

,

Rägo

L

2000

Anorexigenic cocaine- and amphetamine-regulated transcript peptide intensifies fear reactions in rats.

Brain Res

857

:

283

285

328

Lambert

PD

,

Couceyro

PR

,

McGirr

KM

,

Dall Vechia

SE

,

Smith

Y

,

Kuhar

MJ

1998

CART peptides in the central control of feeding and interactions with neuropeptide Y.

Synapse

29

:

293

298

329

Yamanaka

A

,

Kunii

K

,

Nambu

T

,

Tsujino

N

,

Sakai

A

,

Matsuzaki

I

,

Miwa

Y

,

Goto

K

,

Sakurai

T

2000

Orexin-induced food intake involves neuropeptide Y pathway.

Brain Res

859

:

404

409

330

Watabe

T

,

Tanaka

K

,

Kumagae

M

,

Itoh

S

,

Takeda

F

,

Morio

K

,

Hasegawa

M

,

Horiuchi

T

,

Miyabe

S

,

Shimizu

N

1987

Hormonal responses to insulin-induced hypoglycemia in man.

J Clin Endocrinol Metab

65

:

1187

1191

331

Nagatani

S

,

Tsukamura

H

,

Murahashi

K

,

Bucholtz

DC

,

Foster

DL

,

Maeda

K

1996

Paraventricular norepinephrine release mediates glucoprivic suppression of pulsatile luteinizing hormone secretion.

Endocrinology

137

:

3183

3186

332

Goeij

DCE

,

Binnekade

R

,

Tilders

FJH

1992

Chronic stress enhances vasopressin but not corticotropin-releasing factor secretion during hypoglycemia

.

Am J Physiol

263

:

E394

E399

333

Carpenter

DO

,

Briggs

DB

1986

Insulin excites neurons of the area postrema and causes emesis.

Neurosci Lett

68

:

85

89

334

Mizuno

Y

,

Oomura

Y

1984

Glucose responding neurons in the nucleus tractus solitarius of the rat: in vitro study.

Brain Res

307

:

109

116

335

Berthoud

H-R

,

Jeanrenaud

B

1979

Acute hyperinsulinemia and its reversal by vagotomy after lesions of the ventromedial hypothalamus in anesthetized rats.

Endocrinology

105

:

146

151

336

Fuller

RW

1992

The involvement of serotonin in regulation of pituitary-adrenocortical function

.

Front Neuroendocrinol

13

:

250

270

337

Saper

CB

,

Loewy

AD

,

Swanson

LW

,

Cowan

WM

1976

Direct hypothalamo-autonomic connections.

Brain Res

117

:

305

312

338

Swanson

LW

,

Kuypers

HGJM

1980

The paraventricular nucleus of the hypothalamus: cytoarchitectonic subdivisions and organization of projections to the pituitary, dorsal vagal complex and spinal cord as demonstrated by retrograde fluorescence double labeling methods.

J Comp Neurol

194

:

555

570

339

Lawrence

D

,

Pittman

QJ

1985

Interaction between descending paraventricular neurons and vagal motor neurons.

Brain Res

332

:

158

160

340

Portillo

F

,

Carrasco

M

,

Vallo

JJ

1996

Hypothalamic neuron projection to autonomic preganglionic levels related with glucose metabolism: a fluorescent labelling study in the rat.

Neurosci Lett

210

:

197

200

341

Loewy

AD

,

Haxhiu

MA

1993

CNS cell groups projecting to pancreatic parasympathetic preganglionic neurons

.

Brain Res

620

:

323

330

342

Sofroniew

MV

,

Schrell

U

1981

Evidence for a direct projection from oxytocin and vasopressin neurons in the hypothalamic paraventricular nucleus to the medulla oblongata: immunohistochemical visualization of both horseradish peroxidase transported and the peptide produced by the same neurons

.

Neurosci Lett

22

:

211

217

343

ter Horst

GJ

,

Luiten

PG

1986

The projections of the dorsomedial hypothalamic nucleus in the rat.

Brain Res Bull

16

:

231

248

344

Palkovits

M

1987

Anatomy of neural pathways affecting CRH secretion.

Ann NY Acad Sci

512

:

139

148

345

Graessler

J

,

Kvetnansky

R

,

Jezova

D

,

Dobrakovova

M

,

Van Loon

GR

1989

Prior immobilization stress alters adrenal hormone responses to hemorrhage in rats

.

Am J Physiol

257

:

R661

R667

346

Darlington

DN

,

Keil

LC

,

Dallman

MF

1989

Potentiation of hormonal responses to hemorrhage and fasting, but not hypoglycemia in conscious adrenalectomized rats.

Endocrinology

125

:

1398

1406

347

Darlington

DN

,

Shinsako

J

,

Dallman

MF

1986

Responses of ACTH, epinephrine, norepinephrine, and cardiovascular system to hemorrhage

.

Am J Physiol

251

:

H612

H618

348

Bereiter

DA

,

Gann

DS

1986

Potentiation of hemorrhage-evoked catecholamine release by prior blood loss in cats

.

Am J Physiol

250

:

E18

E23

349

Bereiter

DA

,

Zaid

AM

,

Gann

DS

1986

Effect of rate of hemorrhage on sympathoadrenal catecholamine release in cats

.

Am J Physiol

250

:

E69

E75

350

Ljunqvist

O

,

Effendic

S

,

Eneroth

P

,

Hamberger

B

,

Nylander

G

,

Ware

J

1986

Nutritional status and endocrine response to hemorrhage.

Can J Physiol Pharmacol

64

:

1185

1188

351

Karplus

JP

,

Kreidl

A

1927

Gehirn und Sympathicus. VII. Mitteilung. Ueber Beziehungen der Hypothalamuszentren zu Blutdruck und innerer Sekretion.

Pflugers Arch Physiol

215

:

667

670

352

Feuerstein

G

,

Johnso

AK

,

Erbe

RL

,

Davis-Kramer

R

,

Faden

AI

1984

Anteroventral hypothalamus and hemorrhagic shock: cardiovascular and neuroendocrine responses

.

Am J Physiol

246

:

R551

R557

353

Allen

GV

,

Cechetto

DF

1992

Functional and anatomical organization of cardiovascular pressor and depressor sites in the lateral hypothalamic area. I. Descending projections.

J Comp Neurol

315

:

313

332

354

Chan

RKW

,

Sawchenko

PE

1994

Spatially and temporally differentiated patterns of c-fos expression in brainstem catecholaminergic cell groups induced by cardiovascular challenges in the rat.

J Comp Neurol

348

:

433

460

355

Anselmo-Franci

JA

,

Peres-Polon

VL

,

da Rocha-Barros

VM

,

Moreira

ER

,

Franci

CR

,

Rocha

MJA

1998

C-fos expression and electrolytic lesions studies reveal activation of the posterior region of locus coeruleus during hemorrhage induced by hypotension.

Brain Res

799

:

278

284

356

Wood

CE

,

Shinsako

J

,

Keil

LC

,

Ramsay

DJ

,

Dallman

MF

1982

Apparent dissociation of adrenocorticotropin and corticosteroid responses to 15 ml/kg hemorrhage in conscious dogs.

Endocrinology

110

:

1416

1421

357

Wood

CE

,

Shinsako

J

,

Keil

L

,

Dallman

MF

1982

Adrenal sensitivity to adrenocorticotropin in normovolemic and hypovolemic conscious dogs.

Endocrinology

110

:

1422

1429

358

Darlington

DN

,

Shinsako

J

,

Dallman

MF

1986

Medullary lesions eliminate ACTH responses to hypotensive hemorrhage

.

Am J Physiol

251

:

R106

R115

359

van der Kooy

D

,

Koda

LY

,

McGinty

JF

,

Gerfen

CR

,

Bloom

FE

1984

The organization of projections from the cortex, amygdala, and hypothalamus to the nucleus if the solitary tract in the rat.

J Comp Neurol

224

:

1

24

360

Sawchenko

PE

,

Bohn

MC

1989

Glucocorticoid receptor immunoreactivity in C1, C2 and C3 adrenergic neurons that project to the hypothalamus or to the spinal cord in the rat.

J Comp Neurol

205

:

107

116

361

Ruggiero

DA

,

Cravo

SL

,

Aragano

V

,

Reis

DJ

1989

Central control of the circulation by the rostral ventrolateral reticular nucleus, anatomical substrates. In:

Ciriello

J

,

Caverson

MM

,

Polosa

C

, eds. The central neural organization of cardiovascular control, vol.

81. Amsterdam

:

Elsevier

;

49

79

362

Morrison

SF

,

Milner

TA

,

Pickel

VM

,

Reis

DJ

1988

Reticulospinal vasomotor neurons of the rat rostral ventrolateral medulla (RVL), relationship to sympathetic nerve activity and the C1 adrenergic cell group.

J Neurosci

8

:

1286

1301

363

Lightman

SL

,

Todd

K

,

Everitt

BJ

,

Brown

MJ

,

Causon

RC

1984

Ascending brain-stem noradrenergic pathways modulate the renin response to haemorhage.

Clin Sci

67

:

269

272

364

Makino

S

,

Hashimoto

K

,

Ota

Z

1990

Effect of posterior hypothalamic knife cuts on the baroflex and hemorrhage-induced hormonal responses.

Acta Med Okayama

44

:

93

102

365

Lightman

SL

,

Todd

K

,

Everitt

BJ

1984

Ascending noradrenergic projections from the brainstem: evidence for a major role in the regulation of blood pressure and vasopressin secretion.

Exp Brain Res

55

:

145

151

366

Li

YW

,

Dampney

RAL

1994

Expression of fos-like protein in brain following sustained hypertension and hypotension in conscious rabbits.

Neuroscience

61

:

613

634

367

Badoer

E

,

McKinley

MJ

,

Oldfield

BJ

,

McAllen

RM

1993

A comparison of hypotensive and non-hypotensive hemorrhage on Fos expression in spinally projecting neurons of the paraventricular nucleus and rostral ventrolateral medulla.

Brain Res

610

:

216

223

368

Ross

CA

,

Ruggiero

DA

,

Joh

TH

,

Park

DH

,

Reis

DJ

1984

Rostral ventrolateral medulla: Selective projections to the thoracic autonomic cell column from the region containing C1 adrenaline neurons.

J Comp Neurol

228

:

168

185

369

Ross

CA

,

Ruggiero

DA

,

Joh

TH

,

Park

DH

,

Reis

DJ

1983

Adrenaline synthesizing neurons in the rostral ventrolateral medulla: a possible role in vasomotor control.

Brain Res

273

:

356

361

370

Badoer

E

,

Merolli

J

1998

Neurons in the hypothalamic paraventricular nucleus that project to the rostral ventrolateral medulla are activated by haemorrhage.

Brain Res

791

:

317

320

371

Blair

M

,

Piekut

D

,

Want

A

,

Olschowka

JA

1996

Role of the hpothalamic paraventricular nucleus in cardiovascular regulation.

Clin Exp Pharmacol Physiol

23

:

161

165

372

Krukoff

TL

,

MacTavish

D

,

Jhamandas

JH

1997

Activation by hypotension of neurons in the hypothalamic paraventricular nucleus that project to the brainstem.

J Comp Neurol

385

:

285

296

373

Krukoff

TL

,

MacTavish

D

,

Harris

K

,

Jhamandas

JH

1995

Changes in blood volume and pressure induce c-fos expression in brainstem neurons that project to the paraventricular nucleus of the hypothalamus.

Mol Brain Res

34

:

99

108

374

Dun

NJ

,

Dun

SL

,

Chiaia

NL

1993

Hemorrhage induces Fos immunoreactivity in rat medullary catecholaminergic neurons.

Brain Res

608

:

223

232

375

Dun

NJ

,

Dun

SL

,

Shen

E

,

Tang

H

,

Huang

R

,

Chiu

TH

1995

C-fos expression as a marker of central cardiovascular neurons.

Biol Signals

4

:

117

123

376

Chan

RKW

,

Sawchenko

PE

1995

Hemodynamic regulation of tyrosine hydroxylase messenger RNA in medullary catecholamine neurons: a c-fos-guided hybridization histochemical study.

Neuroscience

66

:

377

390

377

Brizzee

BL

,

Russ

RD

,

Walker

BR

1991

Role of vasopressin release in acutely altered baroreflex sensitivity during hemorrhage in rats

.

Am J Physiol

261

:

R667

R685

378

Matsukawa

S

,

Keil

LC

,

Reid

IA

1991

Role of endogenous angiotensin II in the control of vasopressin secretion during hypovolemia and hypotension in conscious rabbits.

Endocrinology

128

:

204

210

379

Wehberg

KE

,

Gala

GL

,

Brunner

MJ

1991

Comparison of carotid baroreflex control of plasma AVP concentration in concscious and anesthetized dogs

.

Am J Physiol

261

:

R950

R956

380

Badoer

E

,

McKinley

MJ

,

Oldfield

BJ

,

McAllen

RM

1992

Distribution of hypothalamic, medullary and lamina terminalis neurons expressing Fos after hemorrhage in conscious rats.

Brain Res

582

:

323

328

381

Cunningham

ETJ

,

Miselis

RR

,

Sawchenko

PE

1994

The relationship of efferent projections from the area postrema to vagal motor and brain stem catecholamine-containing cell groups: an axonal transport and immunohistochemical study in the rat.

Neuroscience

58

:

635

648

382

Plotsky

PM

,

Bruhn

TO

,

Vale

W

1985

Evidence for multifactor regulation of the adrenocorticotropin secretory response to hemodynamic stimuli.

Endocrinology

116

:

633

639

383

Darlington

DN

,

Neves

RB

,

Ha

T

,

Chew

G

,

Dallman

MF

1990

Fed, but not fasted, adrenalectomized rats survive the stress of hemorrhage and hypovolemia.

Endocrinology

127

:

759

765

384

Darlington

DN

,

Barraclough

CA

,

Gann

DS

1992

Hypotensive hemorrhage elevates corticotropin-releasing hormone messenger ribonucleic acid (mRNA) but not vasopressin mRNA in the rat hypothalamus.

Endocrinology

130

:

1281

1288

385

Thrivikraman

KV

,

Plotsky

PM

1993

Absence of glucocorticoid negative feedback to moderate hemorrhage in conscious rats

.

Am J Physiol

264

:

E497

E503

386

Thrivikraman

KV

,

Plotsky

PM

1992

Locus coeruleus noradrenergic modulation of the pituitary-adrenocortical responses to hemorrhage and novel environment in awake rats. Program of the 74th Annual Meeting of The Endocrine Society, San Antonio, TX, 1992, p 328

387

Darlington

DN

,

Shinsako

J

,

Dallman

MF

1988

Paraventricular lesions: hormonal and cardiovascular responses to hemorrhage.

Brain Res

439

:

289

301

387A

Imaki

T

,

Katsumata

H

,

Miyata

M

,

Naruse

M

,

Imaki

J

,

Minami

S

2001

Expression of corticotropin releasing factor (CRF), urocortin and CRF type 1 receptors in hypothalamic-hypophyseal systems under osmotic stimulation.

J Neuroendocrinol

13

:

328

338

388

Ota

M

,

Crofton

J

,

Share

L

1994

Hemorrhage-induced vasopressin release in paraventricular nucleus measured by in vivo microdialysis.

Brain Res

658

:

49

54

389

Hattori

T

,

Morris

M

,

Alexander

N

,

Sundberg

DK

1990

Exracellular oxytocin in the paraventricular nucleus: hyperosmotic stimulation by in vivo microdialysis.

Brain Res

506

:

169

171

390

Landgraf

R

,

Ludwig

M

1991

Vasopressin release within the supraoptic and paraventricular nuclei of the rat brain: osmotic stimulation via microdialysis.

Brain Res

558

:

191

196

391

Plotsky

PM

,

Vale

W

1984

Hemorrhage-induced secretion of corticotropin-releasing factor-like immunoreactivity into the rat hypophysial portal circulation and its inhibition by glucocorticoids.

Endocrinology

114

:

164

169

392

Stricker

MJ

,

Verbalis

JG

1986

Interaction of osmotic and volume stimuli in regulation of neurohypophyseal secretion in rats

.

Am J Physiol

250

:

R267

R275

393

Jurcovicova

J

,

Dobrakovova

M

,

Jezova

D

,

Oprsalova

Z

,

Kvetnansky

R

,

Vigas

M

1988

Different pattern of prolactin release under various acute stress stimuli in rats.

Endocrinol Exp

22

:

235

242

394

Jurcovicova

J

,

Kvetnansky

R

,

Dobrakovova

M

,

Jezova

D

,

Kiss

A

1990

Prolactin response to immobilization stress and hemorrhage: the effect of hypothalamic deafferentations and posterior pituitary denervation.

Endocrinology

126

:

2527

2533

395

Zerbe

RL

,

Bayorh

MA

,

Feuerstein

G

1982

Vasopressin: an essential pressor factor for blood pressure recovery following hemorrhage.

Peptides

3

:

509

514

396

Lilly

MP

,

Engeland

MC

,

Gann

DS

1986

Pituitary-adrenal responses to repeated small hemorrhage in conscious dogs.

Am J Physiol

251

:

1200

1207

397

Lilly

MP

,

Engeland

MC

,

Gann

DS

1986

Adrenal medullary responses to repeated hemorrhage in conscious dogs.

Am J Physiol

251

:

1193

1199

398

Schiltz

JC

,

Hoffman

GE

,

Stricker

EM

,

Sved

AF

1997

Decreases in arterial pressure activate oxytocin neurons in conscious rats. Am J Physiol 273:R1474–R1483.

399

Shoji

M

,

Kimura

T

,

Kawarabayasi

Y

,

Ota

K

,

Inoue

M

,

Yamamoto

T

,

Sato

K

,

Ohta

M

,

Funyu

T

,

Yamamoto

T

,

Abe

K

1993

Effects of acute hypotensive hemorrhage on arginine vasopressin gene transcription in the rat brain.

Neuroendocrinology

58

:

630

636

400

Roberts

MM

,

Robonson

AG

,

Fitzsimmons

MD

,

Grant

F

,

Lee

WS

,

Hoffman

GE

1993

c-Fos expression in vasopressin and oxytocin neurons reveals functional heterogeneity within magnocellular neurons.

Neuroendocrinology

57

:

388

400

401

Feuerstein

G

,

Zerbe

RL

,

Siren

AL

1991

The supraoptic nuclei in vasopressin and hemodynamic responses to hemorrhage in rats.

Neuroreport

2

:

612

614

402

Morris

MJ

,

Hastings

JA

,

Pavia

JM

1994

Catecholamine release in the rat hypothalamic paraventricular nucleus in response to haemorrhage, desipramine and potassium.

Brain Res

665

:

5

12

403

Van Huysse

JW

,

Bealer

SL

1991

Central nervous system norepinephrine release during hypotension and hyperosmolality in conscious rats

.

Am J Physiol

260

:

R1071

R1076

404

Thrivikraman

KV

,

Plotsky

PM

,

Gann

DS

1993

Alterations of locus coeruleus noradrenergic activity in relation to pituitary secretion after hemorrhage in cats.

Neurosci Lett

161

:

85

88

405

Qualy

JM

,

Westfall

TC

1989

Release of norepinephrine from the paraventricular hypothalamic nucleus of hypertensive rats

.

Am J Physiol

254

:

H993

H1003

406

Kendrick

KM

,

Leng

G

1988

Hemorrhage-induced release of noradrenaline, 5-hydroxytyptamine and uric acid in the supraoptic nucleus of the rat, measure by microdialysis.

Brain Res

440

:

402

406

407

Randle

JCR

,

Mazurek

M

,

Kneifel

D

,

Dufresne

J

,

Renaud

LP

1986

A1-adrenergic receptor activation releases vasopressin and oxytocin from perfused hypothalamic explants

.

Neurosci Lett

65

:

219

223

408

Day

TA

,

Renaud

LP

1984

Electrophysiological evidence that noradrenergic afferents selectively facilitate the activity of supraoptic vasopressin neurons.

Brain Res

303

:

233

240

409

Thrivikraman

KV

,

Carlson

DE

,

Gann

DS

1988

Noradrenergic turnover increases in locus coeruleus after hemorrhage in cats

.

Am J Physiol

254

:

R296

R301

410

Schreihofer

AM

,

Hoffman

GE

,

Sved

AF

1997

The kidneys stimulate vasopressin release during hemorrhage in rats with chronic NTS lesions

.

Am J Physiol

272

:

R1540

R1551

411

Blessing

WW

,

Willoughby

JO

1985

Inhibiting the rabbit caudal ventrolateral medulla prevents the baroreceptor-initiated secretion of vasopressin.

J Physiol (Lond)

363

:

253

265

412

Head

GA

,

Quail

AW

,

Woods

RL

1987

Lesions of A1 noradrenergic cells affect AVP release and heart rate during hemorrhage

.

Am J Physiol

253

:

H1012

H1017

413

Gieroba

ZJ

,

Shapoval

LN

,

Blessing

WW

1994

Inhibition of the A1 area prevents hemorrhage-induced secretion of vasopressin in rats.

Brain Res

657

:

330

332

414

Smith

DW

,

Sibbald

JR

,

Khanna

S

,

Day

TA

1995

Rat vasopressin cell responses to stimulated hemorrhage: stimulus-dependent role for A1 noradrenergic neurons

.

Am J Physiol

268

:

R1336

R1342

415

Khanna

S

,

Sibbald

JR

,

Smith

DW

,

Day

TA

1994

Initiation of rat vasopressin cell responses to stimulated hypotensive hemorrhage

.

Am J Physiol

267

:

R1142

R1149

416

Plotsky

PM

1989

Regulation of the adrenocortical axis: hypophysiotropic coding, catecholamines and glucocorticoids. In:

Rose

FC

, ed. The control of the hypothalamo-pituitary-adrenocortical axis.

Madison WI

:

International University Press

;

131

146

417

Grossman

SP

1990

Thirst and sodium appetite.

San Diego, CA

:

Academic Press

418

Lenkei

Z

,

Palkovits

M

,

Corvol

P

,

Llorens-Cortes

C

1997

Expression of angiotensin type-1 (AT1) and type-2 (AT2) receptor mRNAs in the adult rat brain: a functional neuroanatomical review.

Front Neuroendocrinol

18

:

383

439

419

Phillips

MI

1987

Functions of angiotensin II in the central nervous system.

Annu Rev Physiol

49

:

413

435

420

McKinley

MJ

,

Badoer

E

,

Oldfield

BJ

1992

Intravenous angiotensin II induces Fos-immunoreactivity in circumventricular organs of the lamina terminalis.

Brain Res

594

:

295

300

421

Palkovits

M

,

Bahner

U

,

Geiger

H

1995

Preoptic neuronal circuit: atrial natriuretic peptide-containing neurons are sensitive to acute and chronic alterations in body fluid homeostasis.

Miner Electrolyte Metab

21

:

423

427

422

Lind

RW

,

Ohman

LE

,

Lansing

MB

,

Johnson

AK

1983

Transection of subfornical organ connections diminishes the pressor response to intravenously infused angiotensin II.

Brain Res

275

:

361

364

423

Lind

RW

,

Swanson

LW

,

Ganten

D

1994

Angiotensin II immunoreactivity in the neuronal afferents and efferents of the subfornical organ of the rat.

Brain Res

321

:

209

215

424

Buller

MK

,

Smith

DW

,

Day

TA

1999

Differential recruitment of hypothalamic neuroendocrine and ventrolateral medulla catecholamine cells by non-hypotensive and hypotensive hemorrhages.

Brain Res

834

:

42

54

425

Cowley

AW

1992

Long-term control of arterial bood pressure.

Physiol Rev

72

:

231

300

426

McAllen

RM

,

Blessing

WW

1987

Neurons (presumably A1 cells) projecting from the caudal ventrolateral medulla to the region of the supraoptic nucleus respond to baroreceptor inputs in the rabbit.

Neurosci Lett

73

:

247

252

427

Abbott

FV

,

Hong

Y

,

Blier

P

1997

Persisting sensitization of the behavioural response to formalin- induced injury in the rat through activation of serotonin2A receptors.

Neuroscience

77

:

575

584

428

Makara

GB

,

Stark

E

,

Mihály

K

1969

Corticotrophin release induced by injection of formalin in rats with hemisection of the spinal cord.

Acta Physiol Acad Sci Hung

35

:

331

333

429

Stark

EG

,

Makara

B

,

Palkovits

M

,

Mihaly

K

1970

Afferent pathways of stressful stimuli:their dependence on strength and the time elapsed after the onset of stimulation.

Acta Physiol Acad Sci Hung

38

:

43

49

430

Palkovits

M

,

Baffi

JS

,

Pacak

K

1999

The role of ascending neuronal pathways in stress-induced release of noradrenaline in the hypothalamic paraventricular nucleus of rats.

J Neuroendocrinol

11

:

529

539

431

Taylor

BK

,

Akana

SF

,

Peterson

MA

,

Dallman

MF

,

Basbaum

AI

1998

Pituitary-adrenocortical responses to persistent noxious stimuli in the awake rat: endogenous corticosterone does not reduce nociception in the formalin test.

Endocrinology

139

:

2407

2413

432

Dickenson

AH

,

Sullivan

AF

1987

Subcutaneous formalin-induced activity of dorsal horn neurones in the rat: differential response to an intrathecal opiate administered pre or post formalin.

Pain

30

:

349

360

433

Simmons

RMA

,

Li

DL

,

Hoo

KH

,

Deverill

M

,

Ornstein

PL

,

Iyengar

S

1998

Kainate GluR5 receptor subtype mediates the nociceptive response to formalin in the rat.

Neuropharmacology

37

:

25

36

434

Gogas

KR

,

Presley

RW

,

Levine

JD

,

Basbaum

AI

1991

The antinociceptive action of supraspinal opioids results from an increase in descending inhibitory control: correlation of nociceptive behavior and c-fos expression.

Neuroscience

42

:

617

628

435

Presley

RM

,

Menétrey

D

,

Levine

JD

,

Basbaum

AI

1990

Systemic morphine suppresses noxious stimulus-evoked fos protein-like immunoreactivity in the rat spinal cord.

J Neurosci

10

:

323

335

436

Zhang

RX

,

Wang

R

,

Chen

JY

,

Qiao

JT

1994

Effects of descending inhibitory systems on the c-Fos expression in the rat spinal cord during formalin-induced noxious stimulation.

Neuroscience

58

:

299

304

437

Abbadie

C

,

Lombard

M-C

,

Morain

F

,

Besson

JM

1992

Fos-like immunoreactivity in the rat superficial dorsal horn induced by formalin injection in the forepaw: effects of dorsal rhizotomies.

Brain Res

578

:

17

25

438

Pezzone

MA

,

Lee

W-S

,

Hoffman

GE

,

Pezzone

KM

,

Rabin

BS

1993

Activation of brainstem catecholaminergic neurons by conditioned and unconditioned aversive stimuli as revealed by c-Fos immunoreactivity.

Brain Res

68

:

310

318

439

Besson

JM

,

Chaouch

A

1987

Peripheral and spinal mechanisms of nociception.

Physiol Rev

67

:

67

181

440

Fields

HI

,

Heinricher

MM

,

Mason

P

1991

Neurotransmitters in nociceptive modulatory circuits.

Annu Rev Neurosci

14

:

219

245

441

Menétrey

D

,

Chaouch

A

,

Binder

D

,

Besson

JM

1982

The origin of spinomesencephalic tract in the rat: an anatomical study using the retrograde transport of horseradish peroxidase.

J Comp Neurol

206

:

193

207

442

Liu

RP

1986

Spinal neuronal collaterals to the intralaminar thalamic nuclei and periaqueductal gray.

Brain Res

365

:

145

150

443

Zhou

M

,

Gebhart

GF

1991

Spinal serotonin receptors mediate descending facilitation of a nociceptive reflex from the nuclei reticularis gigantocellularis and gigantocellularis pars α in the rat.

Brain Res

550

:

35

48

444

Inoue

A

,

Hashimoto

T

,

Hilde

I

,

Nishio

H

,

Nakata

Y

1997

5- Hydroxytryptamine-facilitated release of substance P from rat spinal cord slices is mediated by nitric oxide and cyclic GMP.

J Neurochem

68

:

128

133

445

Helmstetter

FJ

,

Tershner

SA

,

Poore

LH

,

Bellgowan

PSF

1998

Antinociception following opioid stimulation of the basolateral amygdala is expressed through the periaqueductal gray and rostral ventromedial medulla.

Brain Res

779

:

104

118

446

Meng

ID

,

Bereiter

DA

1996

Differential distribution of Fos-like immunoreactivity in the spinal trigeminal nucleus after noxious and innocuous thermal and chemical stimulation of rat cornea.

Neuroscience

72

:

243

254

447

Swanson

LW

,

Simmons

DM

1989

Differential steroid hormone and neural influences on peptide mRNA levels in CRH cells of the paraventricular nucleus: a hybridization histochemical study in the rat.

J Comp Neurol

285

:

413

435

448

Szafarczyk

A

,

Guillame

V

,

Conte-Devolx

B

,

Alonso

G

,

Malaval

F

,

Pares-Herbute

N

,

Oliver

C

,

Assenmacher

I

1988

Central catecholaminergic system stimulates secretion of CRH at different sites

.

Am J Physiol

255

:

E463

E468

449

Mezey

E

,

Palkovits

M

1991

CRF-containing neurons in the hypothalamic paraventricular nucleus: Regulation, especially by catecholamines.

Front Neuroendocrinol

12

:

23

37

450

Sumal

KK

,

Blessing

WW

,

Joh

TH

,

Reis

DJ

,

Pickel

VM

1983

Synaptic interactions of vagal afferents and catecholaminergic neurons in the rat nucleus solitarius.

Brain Res

277

:

31

40

451

Cederbaum

JM

,

Aghajanian

GK

1978

Afferent projections to the rat locus coeruleus as determined by a retrograde tracing method.

J Comp Neurol

178

:

1

16

452

Menétrey

D

,

Gannon

A

,

Levine

JD

,

Basbaum

AI

1989

Expression of c-fos protein in interneurons and projection neurons of the rat spinal cord in response to noxious somatic, articular and visceral stimulation.

J Comp Neurol

285

:

177

195

453

Palkovits

M

,

Mezey

E

,

Fodor

M

,

Ganten

D

,

Bahner

U

,

Geiger

H

,

Heidland

A

1995

Neurotransmitters and neuropeptides in the baroreceptor reflex arc: connections between the nucleus of the solitary tract and the ventrolateral medulla oblongata in the rat.

Clin Exp Hypertens

17

:

101

113

454

Kobayashi

RM

,

Palkovits

M

,

Kopin

IJ

,

Jacobowitz

DM

1974

Biochemical mapping of noradrenergic nerves arising from the rat locus coeruleus.

Brain Res

77

:

269

279

455

Záborszky

L

,

Feminger

A

,

Palkovits

M

1979

Afferent brain stem connections of the central amygdaloid nucleus.

Verh Anat Ges

73

:

1117

1120

456

Chappell

PB

,

Smith

MA

,

Kilts

CD

,

Bissette

G

,

Ritchie

J

,

Anderson

C

,

Nemeroff

CB

1986

Alterations in corticotropin-releasing factor-like immunoreactivity in discrete rat brain regions after acute and chronic stress.

J Neurosci

6

:

2908

2914

457

Valentino

RJ

,

Foote

SL

1988

Corticotropin-releasing hormone increases tonic but not sensory-evoked activity of noradrenergic locus coeruleus neurons in unanesthetized rats.

J Neurosci

8

:

1016

1025

458

Lechner

SM

,

Valentino

RJ

1999

Glucocorticoid receptor-immunoreactivity in corticotrophin-releasing factor afferents to the locus coeruleus.

Brain Res

816

:

17

28

459

Hamba

M

,

Hisamitsu

H

,

Muro

M

1990

Nociceptive projection from tooth-pulp to the lateral hypothalamus in rats.

Brain Res Bull

25

:

355

364

460

Marti

O

,

Armario

A

1998

Anterior pituitary response to stress: time-related changes and adaptation.

Int J Dev Neurosci

16

:

241

260

461

Makara

GB

1985

Mechanisms by which stressful stimuli activate the pituitary-adrenal system.

Fed Proc

44

:

149

153

462

Rivier

C

,

Rivest

S

1991

Effect of stress on the activity of the hypothalamic-pituitary-gonadal axis: peripheral and central mechanisms.

Biol Reprod

45

:

523

532

463

Patchev

VK

,

Almeida

OF

1998

Gender specificity in the neural regulation of the response to stress: new leads from classical paradigms.

Mol Neurobiol

16

:

63

77

464

Buckingham

JC

,

Cowell

A-M

,

Gillies

GE

,

Herbison

AE

,

Stell

JH

1997

The neuroendocrine system: anatomy, physiology and responses to stress. In:

Buckingham

JC

,

Gillies

GE

,

Cowell

A-M

, eds. Stress, stress hormones and the immune system, vol.

1. New York

:

John Wiley & Sons, Ltd.

;

9

47

465

Johnson

EO

,

Kamilaris

TC

,

Chrousos

GP

,

Gold

PW

1992

Mechanisms of stress: a dynamic overview of hormonal and behavioral homeostasis.

Neurosci Biobehav Rev

16

:

115

130

466

McGrady

AV

1984

Effects of psychological stress on male reproduction: a review.

Arch Androl

13

:

1

7

467

Chrousos

GP

1997

The neuroendocrinology of stress: its relation to the hormonal milieu, growth, and development.

Growth Genet Horm

13

:

1

8

468

Garssen

B

,

Goodkin

K

1999

On the role of immunological factors as mediators between psychosocial factors and cancer progression.

Psychiatry Res

85

:

51

61

469

Gold

PW

,

Chrousos

GP

1999

The endocrinology of melancholic and atypical depression: relation to neurocircuitry and somatic consequences.

Proc Assoc Am Physicians

111

:

22

34

470

Holsboer

F

,

Barden

N

1996

Antidepressants and hypothalamic-pituitary-adrenocortical regulation.

Endocr Rev

17

:

187

205

471

Stenzel-Poore

MP

,

Heinrichs

SC

,

Rivest

S

,

Koob

GF

,

Vale

WW

1994

Overproduction of corticotropin-releasing factor in transgenic mice: a genetic model of anxiogenic behavior.

J Neurosci

14

:

2579

2584

472

Holsboer

F

1999

The rationale for corticotropin-releasing hormone receptor (CRH-R) antagonists to treat depression and anxiety.

J Psychiatr Res

33

:

181

214

473

Montkowski

A

,

Barden

N

,

Wotjak

C

,

Stec

I

,

Ganster

J

,

Meaney

M

,

Engelmann

M

,

Reul

JM

,

Landgraf

R

,

Holsboer

F

1995

Long-term antidepressant treatment reduces behavioural deficits in transgenic mice with impaired glucocorticoid receptor function.

J Neuroendocrinol

7

:

841

845

474

Dijkstra

I

,

Tilders

FJ

,

Aguilera

G

,

Kiss

A

,

Rabadan-Diehl

C

,

Barden

N

,

Karanth

S

,

Holsboer

F

,

Reul

JM

1998

Reduced activity of hypothalamic corticotropin-releasing hormone neurons in transgenic mice with impaired glucocorticoid receptor function.

J Neurosci

18

:

3909

3918

475

Maier

SF

,

Watkins

LR

1998

Cytokines for psychologists: Implications of bidirectional immune-to-brain communication for understanding behavior, mood, and cognition.

Psychol Rev

105

:

83

107

476

Valentino

RJ

,

Curtis

AL

,

Page

ME

,

Pavcovich

LA

,

Florin-Lechner

SM

1998

Activation of the locus ceruleus brain noradrenergic system during stress: circuitry, consequences, and regulation.

Adv Pharmacol

42

:

781

784

477

Chrousos

GP

,

Torpy

DJ

,

Gold

PW

1998

Interactions between the hypothalamic-pituitary-adrenal axis and the female reproductive system: clinical implications.

Ann Intern Med

129

:

229

240

478

Swanson

LW

1991

Biochemical switching in hypothalamic circuits mediating responses to stress.

Prog Brain Res

87

:

181

200

479

Li

H-Y

,

Ericsson

A

,

Sawchenko

PE

1996

Distinct mechanisms underline activation of hypothalamic neurosecretory neurons and their medullary catecholaminergic afferents in categorically different stress paradigms.

Proc Natl Acad Sci USA

93

:

2359

2364

480

Dallman

MF

1993

Stress update: adaptation of the hypothalamic-pituitary-adrenal axis to chronic stress.

Trends Endocrinol Metab

4

:

62

69

481

Orchinik

M

1998

Glucocorticoids, stress, and behavior: shifting the timeframe.

Horm Behav

34

:

320

327

482

Mulders

WH

,

Meek

J

,

Hafmans

TG

,

Cools

AR

1997

Plasticity in the stress-regulating circuit: decreased input from the bed nucleus of the stria terminalis to the hypothalamic paraventricular nucleus in Wistar rats following adrenalectomy.

Eur J Neurosci

9

:

2462

2471

483

Hyman

SE

1998

Brain neurocircuitry of anxiety and fear: implications for clinical research and practice.

Biol Psychiat

44

:

1201

1203

484

Sehic

T

,

Blatteis

ME

1996

Blockade of lipopolysaccharide-induced fever by subdiaphragmatic vagotomy in guinea pigs.

Brain Res

726

:

160

166

485

Yaksh

TL

1999

Spinal systems and pain processing: development of novel analgesic drugs with mechanistically defined models.

Trends Pharmacol Sci

20

:

329

337

486

Yamamoto

T

,

Yaksh

TL

1991

Stereospecific effects of a nonpeptidic NK1 selective antagonist, CP-96,345: antinociception in the absence of motor dysfunction.

Life Sci

49

:

1955

1963

487

Nishiyama

T

,

Yaksh

TL

,

Weber

E

1998

Effects of intrathecal NMDA and non-NMDA antagonists on acute thermal nociception and their interaction with morphine.

Anesthesiology

89

:

715

722

488

Dirig

DM

,

Isakson

PC

,

Yaksh

TL

1998

Effect of COX-1 and COX-2 inhibition on induction and maintenance of carrageenan-evoked thermal hyperalgesia in rats.

J Pharmacol Exp Ther

285

:

1031

1038

489

Yaksh

TL

,

Chaplan

SR

,

Malmberg

AB

1995

Medications development for the treatment of pregnant addicts and their infants. In:

Chiang

CN

,

Finnegan

FP

, eds. NIDA Research Monograph Series, vol.

149. Rockville, MD

:

US Department of Health and Human Services

;

84

101

490

Mantyh

PW

,

Rogers

SD

,

Honore

P

,

Allen

BJ

,

Ghilardi

JR

,

Li

J

,

Daughters

RS

,

Lappi

DA

,

Wiley

RG

,

Simone

DA

1997

Inhibition of hyperalgesia by ablation of lamina I spinal neurons expressing the substance P receptor.

Science

278

:

275

279

491

Marber

MS

1994

Stress proteins and myocardial protection.

Clin Sci

86

:

375

381

492

Benjamin

IJ

,

Williams

RS

1994

Expression and function of stress proteins in the ischemic heart. In:

Morimoto

RI

,

Tissieres

A

,

Georgopolous

C

, eds. The biology of heat shock proteins and molecular chaperones.

Plainview, NY

:

Cold Spring Harbor Laboratory Press

;

533

552

493

Heinrichs

SC

1998

Stress-axis, coping and dementia: gene-manipulation studies.

J Neurosci

18

:

5938

5947

494

Muglia

L

,

Jacobson

L

,

Dikkes

P

,

Majzoub

JA

1995

Corticotropin-releasing hormone deficiency reveals major fetal but not adult glucocoticoid need.

Nature

373

:

427

432

495

Tewson

TJ

,

Krohn

KA

1998

PET radiopharmaceuticals: state-of-the-art and future prospects.

Semin Nucl Med

28

:

221

234

496

Bernard

C

1878

Lecons sur les phenomenones de la vie communs aux animaux et aux vegetaux

.

In: Bernard C, ed.

Paris

:

Balliere

497

Hennesy

JW

,

Levine

S

1979

Stress, arousal, and the pituitary-adrenal system: a psychoendocrine hypothesis.

Prog Psychobiol Physiol Psychol

8

:

133

178

498

Krantz

DS

,

Lazar

JD

1987

Behavioral factors in hypertension. In:

Julius

S

,

Bassett

DR

, eds. The stress concept: issues and measurements.

New York

:

Elsevier Science

;

43

58

499

Munck

A

,

Guyre

PM

1986

Glucocorticoid physiology, pharmacology, and stress. In:

Chrousos

GP

,

Loriaux

DL

,

Lipsett

MB

, eds. Steroid hormone resistance.

New York

:

Plenum Press

;

81

96

500

Levine

S

,

Ursin

H

1991

What is stress? In:

Brown

MR

,

Koob

GF

,

Rivier

C

, eds. Stress: neurobiology and neuroendocrinology.

New York

:

Dekker

;

3

21

Copyright © 2001 by The Endocrine Society

Copyright © 2001 by The Endocrine Society

Citations

Views

Altmetric

Email alerts

More on this topic

Citing articles via

  • Latest

  • Most Read

  • Most Cited

More from Oxford Academic

Which branch of the autonomic nervous system is activated when a person is agitated by a stressor quizlet?

What is the parasympathetic branch of the nervous system responsible for? The branch of the autonomic nervous system that is activated when a person is exposed to a stressor. The hormone released by the pituitary gland during the stress response.

What branch of the autonomic nervous system is activated by stress?

Sympathetic Nervous System The most noticeable function of the sympathetic branch is during the fight-or-flight response. During conditions that are considered threatening or stressful, the sympathetic system activates, providing an automatic response.

Which part of the autonomic nervous system is activated during trauma?

A branch of the autonomic nervous system responsible for rapid activation of the fight-flight components of the body's emergency survival response in the face of actual or perceived threat, emotional or physiological overwhelm, or traumatic memories, reminders and conditioned reactions.

What triggers the sympathetic nervous system?

Danger or stress activates your sympathetic nervous system, which can cause several things to happen in your body. In response to danger or stress, your sympathetic nervous system may affect your: Eyes: Enlarge your pupils to let more light in and improve your vision.