www.fgks.org   »   [go: up one dir, main page]

Warning: The NCBI web site requires JavaScript to function. more...

U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Institute of Medicine (US) Committee on Health and Behavior: Research, Practice, and Policy. Health and Behavior: The Interplay of Biological, Behavioral, and Societal Influences. Washington (DC): National Academies Press (US); 2001.

Cover of Health and Behavior

Health and Behavior: The Interplay of Biological, Behavioral, and Societal Influences.

Show details

2Biobehavioral Factors in Health and Disease

Research into the bidirectional and multilevel relationships between behavior and health has been aided by technology and by conceptual advances in the behavioral, biological, and medical sciences. Our understanding of the interactions between brain function and behavior has been enriched by advances in behavioral neurobiology, neuroscience, and neuroendocrinology from molecular mechanisms to psychological systems. Real-time imaging of the living human brain during different behavioral states has promoted our understanding of the links between human behavior and basic neurochemical processes or specific neuroanatomic pathways. Common availability of monoclonal antibodies, routine production of genetically altered animals, and new understanding of the genetic code have contributed to exploration of how genetics interacts with development and early experiences to influence both vulnerability to disease and resistance to age-related decline. Yet much of the research knowledge is highly compartmentalized, and there is a need to integrate isolated pockets of information.

This chapter addresses the interplay among biological, behavioral, and social factors in health and disease, with an emphasis on biologic factors. Subsequent chapters address behavioral and social factors in greater detail.

STRESS, HEALTH, AND DISEASE

The Stress Response

Over the past 60 years or so, the study of stress has provided a major link in explaining the behavioral variables and the biological factors that influence physical health. Stress both causes and modulates a diversity of physiological effects that can enhance resistance to disease or cause damage and thereby promote disease. For example, stress-related hormones, such as cortisol and epinephrine, have protective and adaptive functions as well as damaging effects. This idea, first introduced by Hans Selye (1956), is reemerging in contemporary biobehavioral research (McEwen, 1998). A characteristic set of physiological effects—the “stress response” —has been identified and investigated in humans and animals (Chrousos, 1998). The primary and secondary effects of the stress response constitute the biologic pathways along which a person's experiences, living and working conditions, interpersonal relations, lifestyle, diet, personality traits, and general socioeconomic status can affect the body. Individual behavior is important because it increases or decreases the pathophysiological cost of stress through diet, exercise, and other activities.

The stress response is an important component of the body's regulatory systems. The maintenance of constant and appropriate internal conditions and functioning in the face of changing environmental demands is called homeostasis, an idea first developed by Walter Cannon (1936). The stress response, however, primarily involves reaction in an emergency. This function evolved over millions of years and is critical to the survival of most animals, including humans, when external threats and dangers, such as predation, are encountered. The stress response consists of many coadapted and simultaneous shifts in the physiological functioning of the cardiovascular, respiratory, muscular, metabolic, immune, and central nervous systems. Physiological changes can be accompanied by altered emotional responses, enhanced vigilance, heightened appraisal of risk, enhanced memory storage and retrieval, and changes in motivation. The stress response is a rapid and pervasive adjustment of internal states to prepare an organism to adapt to a threat—to respond to the rigors of “fight or flight” (Chrousos, 1998).

Many aspects of the stress response, however, are inappropriate or maladaptive in the context of modern postindustrial societies. The threats posed here are different from those our evolutionary ancestors faced. We do not commonly confront acute, life-threatening assault. Instead, contemporary humans face ill-defined, diffuse, often chronic threats that cannot be resolved by fight or flight. Nevertheless, the ancient physiologic stress response is triggered when one experiences, for example, a threat to social position, damage to important interpersonal relationships, loss of possessions, or barriers to the achievement of goals. Because many difficulties of contemporary life and their accompanying stress cannot be rapidly resolved—as could many physical stressors—the stress response persists, homeostasis is not restored, and the response becomes dysfunctional rather than adaptive. An increasing body of evidence indicates that stress is a potent contributor to illness (Cohen and Herbert, 1996; Cohen et al., 1991; Hermann et al., 1995; Kiecolt-Glaser et al., 1996; McEwen, 1998). The continued and unproductive activation of the stress response, including the failure to shut off this response when it is not needed, called allostatic load, is discussed below.

The stress response is one aspect of an array of biologic and behavioral processes that either protect or cause damage. For example, secretion of stress-related hormones, such as cortisol and the catecholamines (epinephrine and norepinephrine), typically varies in a daily rhythm that is entrained by the light/dark cycle and by sleep/waking patterns that are part of normal daily life. But chronic increase in cortisol throughout the diurnal cycle is associated with negative consequences, such as accelerated bone mineral loss and hyperglycemia. Because the subjective experience of stress does not always correlate with physiological response (Kirschbaum et al., 1999), long-term measurement of hormone concentrations and of the processes that they regulate (for example, blood cholesterol concentration, fat accumulation, immune function, atrophy of brain structures, blood pressure), constitute an important way to connect life experience and the risk of disease.

Allostasis and Allostatic Load

An important new attempt to understand the relationships between environmental and behavioral challenges and stressors, the physiological responses to these events, and disease uses the terms allostasis and allostatic load. Allostasis is the maintenance of overall stability (homeostasis) through the constant adjustment and balancing of various components in the process of adapting to challenge. Sterling and Eyer (1988) first used the term to describe cardiovascular system adjustments in response to rest and activity states. Later, the idea was generalized to other physiologic mediators, such as adrenal cortisol and the catecholamines. Allostatic load is the wear and tear the body experiences as a result of repeated allostatic response (McEwen, 1998; McEwen and Stellar, 1993).

Allostasis and allostatic load operate in all systems of the body and focus attention on the protective, as well as the damaging, property of the primary mediators of the stress response: cortisol and the catecholamines. The major aspects are summarized in Figure 2-1. First, the brain integrates and coordinates behavioral and physiologic responses (hormonal and autonomic) to challenge. Some challenges can be perceived as stressful; others are related to circadian rhythms and to coordination of the functions of sleep and waking with the environment. Second, individual differences in the capacity to cope with challenges are based on multilevel relationships between genetic, developmental, and experiential influences. Third, intrinsic to the autonomic, neuroendocrine, and behavioral responses to challenge is the capacity to adapt (allostasis); indeed, neuroendocrine responses, such as the release of cortisol, are by nature protective and acute. Problems arise only when they persist, so efficient initiation and cessation of these responses is vital. Negative effects result when allostatic responses to challenge or stress occur inappropriately or are terminated inefficiently. Fourth, allostasis has a price that is related to the degree of inefficiency in the response and to the number of challenges and stressors a person experiences. Allostatic load is more than chronic stress. It can also reflect a genetically or developmentally induced failure to cope efficiently with the daily challenges related to the sleep/waking cycle and other experiences. And it also includes contributions of lifestyle factors, such as diet, alcohol and tobacco use, physical activity, and sleep, through their influences on the production of stress hormones.

FIGURE 2-1. The Stress Response and Development of Allostatic Load.

FIGURE 2-1

The Stress Response and Development of Allostatic Load. Individuals experience objective psychological and environmental conditions that are conducive to stress, referred to as stressors. The perception of stress is influenced by social, psychological, (more...)

Protective and Damaging Effects of Stress Mediators

A behavioral response to challenge or stress can be protective or damaging. The risk of harm or disease can be increased by such patterns of behavior as hostility or aggression, and it can be reduced by cooperation and conciliation. Cigarette-smoking, excessive alcohol consumption, high fat consumption, and exposure to physical hazards increase the risk, as does insufficient physical activity. The link of allostasis and allostatic load can be applied to various behavioral responses: Such behaviors as smoking, high alcohol consumption, and consumption of high-fat foods all have some perceived adaptive effects in the short-term but damaging effects if they persist. Behavior can attenuate some of the damaging effects of physiologic responses. For example, even a brief period of exercise can enhance glucose uptake by reducing the insulin resistance of muscle tissue (Perseghin et al.,1996).

The mediators of protective and damaging effects of allostatic responses are mainly adrenal steroids and catecholamines. Other hormones—such as dehydroepiandrosterone, prolactin, growth hormones, and the cytokines—also mediate adaptive or maladaptive effects, but their consequences are often specific to an organ or a system. Once the mediators are released, they produce their effects by acting on cellular receptors. The effects can be classified as primary effects; secondary outcomes, which are risk factors for disease; and tertiary outcomes, which are diseases themselves (McEwen and Seeman, 1999). The actions of the mediators adrenal glucocorticoids and catecholamines are shown in Figure 2-2. These substances act via receptors that trigger changes throughout the target cell (including changes in gene expression) that have long-lasting consequences for cell function. It is important to consider the short- and long-term consequences of hormone release for cell function. There are many examples of beneficial and adverse effects of the mediators of allostatic responses. These factors are introduced here and discussed in more detail later.

FIGURE 2-2. Allostasis in the Autonomic Nervous System and HPA Axis.

FIGURE 2-2

Allostasis in the Autonomic Nervous System and HPA Axis. Allostatic systems respond to stress (upper panel) by initiating the adaptive response, sustaining it until the stress ceases, and then shutting it off (recovery). Allostatic responses are initiated (more...)

In the central nervous system, catecholamines and adrenal steroids promote the storage and retrieval of memories of events, pleasant and unpleasant, associated with arousal. However, adrenal steroids acting with excitatory amino acid neurotransmitters are associated with cognitive dysfunction involving various mechanisms that promote atrophy and, in some extreme cases, the death of neurons, particularly in the hippocampal region.

In the cardiovascular system, autonomic responses, in part because of catecholamines, promote allostasis (adaptation) by adjusting heart rate and blood pressure according to the changing demands of sleeping, waking, and physical exertion. Damaging allostatic load occurs as a result of a failure to terminate blood pressure surges efficiently. This accelerates atherosclerosis and synergizes with metabolic hormones to accelerate non-insulin-dependent diabetes.

The immune system is particularly responsive to the mediators of allostatic response. Adrenal steroids and catecholamines promote the movement of immune cells to organs or tissues where they are needed to resist infection or other challenge, thereby enhancing the effectiveness of immune responses. But adrenal steroids also can increase allostatic load and suppress immune system response when they are secreted chronically or when their release from the adrenal cortex is not terminated properly.

Allostatic load is associated with at least four patterns of long-term harm to the body. The first is a perception of excessive stress. This can take the form of repeated events of various types that cause recurring increases in the release of stress mediators. For example, the amount and frequency of economic hardship are good predictors of decline in physical and mental functioning and even death (Lynch et al., 1997b). The second pattern involves a failure to adapt to recurrence of the same stressor. This leads to overexposure to stress mediators because of the failure to dampen response to a repeated event. Most people, for example, adapt to repeated public-speaking challenges, but some continue to show elevated cortisol concentrations, which indicate a failure to adapt (Kirschbaum et al., 1995). The third pattern entails the failure to terminate the hormonal stress response or the lack of appearance of the normal trough in the daily cortisol release pattern. Examples are increased blood pressure caused by work-related stress (Gerin and Pickering, 1995), increased evening cortisol and hyperglycemia caused by sleep deprivation (Van Cauter et al., 1997), and the chronically elevated cortisol that often accompanies depressive illness (Michelson et al., 1996). The fourth pattern involves inadequate release of hormones, thus allowing other systems, such as inflammatory cytokines, to become overactive. In the Lewis rats, for example, inadequate release of cortisol is associated with increased susceptibility to inflammatory and autoimmune disturbances (Sternberg, 1997; Sternberg et al., 1996).

Early Development Influences Long-Term Effects of Stress

Developmental influences are implicated in susceptibility to stress-related disorders. Classic work by Levine et al. (1967), Denenberg and Haltmeyer (1967), and Ader (1968), shows that the handling of neonatal rats by experimenters leads to reduced emotionality and stress hormone reactivity throughout life. In contrast, prenatal stress increases emotionality and stress hormone reactivity throughout the life of the animal. Post-natal handling reverses the effects of prenatal stress (Fride et al., 1986; Wakschlak and Weinstock, 1990). Handling is believed to increase maternal licking and grooming of pups, which are associated with reduced reactivity of the hypothalamic/pituitary/adrenal (HPA) axis (Liu et al., 1997). Use of animal models has revealed that age-related brain deterioration is increased by high-stress reactivity and reduced by low-stress reactivity (Dellu et al., 1996; Liu et al., 1997; Meaney et al., 1988). Animals that show high-stress reactivity also show a high propensity for substance abuse (Dellu et al., 1994). Studies in nonhuman primates show that early maternal deprivation alters brain serotonin functioning, increases alcohol preference, increases aggressive behavior, and decreases affiliate behaviors (Higley et al., 1996a, b; Kraemer et al., 1997). In marmosets, altered HPA axis function was found in offspring that experienced negative parenting (Johnson et al., 1996a).

Individual differences in human brain aging are correlated with plasma cortisol concentration (see Lupien et al., 1994, 1998; Seeman et al., 1997), although it is not known whether there are connections to early life events. Some data suggest that extremely low birth weight (Barker, 1997; Wadhwa, 1998) and trauma in early life are risk factors that influence later health in humans (Bremner et al., 1997; Felitti et al., 1998). Studies that link life experience, especially early experience, with stress, allostatic load, and later health risks (Felitti et al., 1998; Bremner et al., 1997; De Bellis et al., 1999a, 1999b) signal the importance of research on the effects of early life experiences on later stress reactivity and health.

Some evidence suggests that even prenatal experiences can have long-term health consequences. In laboratory animals, prenatal stress has been linked to alterations in adrenocortical and central serotonergic and dopaminergic circuits (Nelson and Bloom, 1997). These observations led to the hypothesis (Barker and Sultan, 1995) that disease vulnerabilities in childhood and adult life result from “fetal programming” of homeostatic response set points. This is supported by Eriksson et al, (1999), who report that death from coronary heart disease (CHD) among Finnish men was associated with low birth weight and low ponderal index at birth. In addition, prematurity and low birth weight resulting from maternal behaviors has been seen to increase the life-long risk of CHD (Wadhwa, 1998) and diabetes mellitus (Rich-Edwards et al., 1999).

THE BRAIN AS INTERPRETER, REGULATOR, AND TARGET

The brain is influenced by experiences, including stress, and it is a target of allostatic load, or long-term wear and tear. Since the IOM (1982) report, advances in basic neuroscience and the development of imaging technology have combined to enhance our understanding of how different regions of the brain control behavior. They reveal the plasticity and vulnerability of the brain to effects of life experiences. Advances in the neurobiology of learning and memory have provided important insights into the dynamic nature of brain function throughout life and from the level of the gene to the level of nerve-cell structure and function. Studies of the interplay between brain development and early experience have emphasized their importance in establishing patterns of brain function that persist through life. This section discusses advances in neurobiology that are particularly relevant to the brain as an interpreter of life events, as a regulator of the stress response and daily biological rhythms, and as a target of the protective and the damaging functions of stress mediators. Developmental issues are discussed later in the chapter.

Neurotransmitters, Experience, and Behavior

Changes in balance among neurotransmitters in the brain can influence behavioral responses to potentially stressful situations, can alter the interpretation of stimuli, and might be associated with anxiety and depression. Research that has accelerated rapidly over the past few decades reveals that the human brain has multiple neurotransmitters and neuromodulators. The release of a neurotransmitter sends a message from one neuron to another. Neuromodulators share properties of neurotransmitters and hormones to regulate the general tone of neural systems. Many substances act as neuromodulators, including acetylcholine, histamine, serotonin, the catecholamines, excitatory and inhibitory amino acids, and a host of neuropeptides. In this discussion we emphasize two, serotonin and corticotropin-releasing hormone (CRH)—each with links to the stress response—while recognizing that brain function is a result of the simultaneous interactions of many hormones, neuromodulators, and neurotransmitters.

Serotonin is a neurotransmitter with widespread influences throughout the brain. Most neurons that release serotonin are found in the raphe nuclei in the midbrain. Axons of serotonergic cells have many branches, and they project widely throughout the forebrain, cerebellum, and spinal cord (Cooper et al., 1996). The serotonin system exerts widespread influence over mood and mood disorders, such emotional responses as hostility and aggression, arousal, sensory perception, and higher cognitive functions. For example, low concentrations of brain serotonin are associated with increased incidence of suicide (Brown et al., 1982; Mann, 1998), impulsive aggression (Brown et al., 1982; Higley et al., 1996a, 1996b), and the abuse of alcohol and other substances (Higley et al., 1991).

Experience can alter brain chemistry. One well-studied, common pathway through which the environment effects changes in the brain is the HPA axis. Briefly, to maintain appropriate internal conditions (allostasis) in response to life's stresses, the hypothalamus releases CRH, which travels to the anterior pituitary where it causes the release of adrenocorticotropic hormone. The adrenocorticotropic hormone, in turn, controls the release of other hormones, such as cortisol, from the adrenal cortex (located on top of the kidneys). The HPA axis has been shown to be exquisitely responsive to social environment in rats (Albeck et al, 1997) and in primates (Johnson et al., 1996b; Shively et al., 1997). Social status affects the response of the HPA axis, such that hypothalamic CRH release is deficient in subordinate rats, which also show reduced testosterone concentrations under these conditions (Blanchard et al., 1993; Albeck et al., 1997). CRH involvement in the stress response is complex, with at least one other brain system, involving the amygdala, showing an increase in CRH expression linked to anxiety disorders and depression (Schulkin et al, 1998). Furthermore, behavioral and chemical profiles often depend on the context of the stressful situation (Johnson et al., 1996b).

Arousal and Memory Modulation

Many people report vivid memories of events they associate with intense arousal or strong emotion. Sometimes called “flashbulb memories,” these illustrate that memory storage processes are modulated by the degree of arousal associated with an event. The amygdala is a complex collection of nuclei that has several functions, including control of some autonomic responses, elucidation of innate emotional responses, modulation of memory, and control of some aspects of male sexual behavior. The pathway for encoding memories involves the interaction of neural systems in the amygdala and other brain areas, such as the hippocampus, and hormones released by the adrenal cortex (cortisol) and the adrenal medulla (the catecholamines epinephrine and norepinephrine). The amygdala is associated with the modulation of memories of important and arousing life events—events that have strong positive or negative emotional impact (affect; Cahill and McGaugh, 1998; LeDoux, 1996). It mediates the effects of the degree of arousal on the storage of memories of all kinds, including those that pertain to spatial information, events, and learning habits.

In relating autonomic and neuroendocrine function to memory processes, it is noteworthy that the encoding of memories is strengthened by glucocorticoids from the adrenal and by norepinephrine released from nerve terminals in the amygdala, hippocampus, and other brain regions. Substances associated with arousal, such as adrenal epinephrine and various peptide hormones, circulating outside the blood-brain barrier act to stimulate receptors in the periphery that then send neural messages to the brain via the vagus nerve (Clark et al., 1995; Williams and Jensen 1991, 1993). Modulated neural messages are then passed via the nucleus of the solitary tract to brain regions where memories are actually encoded (deQuervain et al., 1998; McGaugh et al., 1996; Roozendaal et al., 1996). In support of this mechanism, antagonism of peripheral b-receptors that respond to epinephrine released from the adrenals attenuates the memory-modulated effects of arousal (Cahill et al., 1994; Nielsen and Jensen, 1995). Those findings could be relevant to the biology of post-traumatic stress disorder and depression, which seem to involve overactive functioning of the amygdala (Cahill and McGaugh, 1998; Drevets et al., 1997; Sheline et al.,1998).

Some important advances in animal model studies related to emotional experiences have recently been carried over to human brain function. Human brain-imaging studies have shown that emotionally arousing information (pleasant or unpleasant) activates the amygdala and induces the formation of strong, long-term, episodic memories of that information (Cahill et al., 1996; Hamann et al., 1999). Electric stimulation of the vagus nerve in laboratory rats enhances memory (Clark et al., 1995); this finding was recently expanded to human memory for word recognition (Clark et al., 1999).

Studies of learning and memory have revealed neural plasticity involving structural changes in brain cells and changes in gene expression. It can be seen in the remodeling of neuron structure brought about by training (Greenough and Bailey, 1988). Transcription factors involved in regulating expressions of groups of genes in brain cells also appear essential to the formation of long-term memories in species as varied as fruit flies and mice (Guzowski and McGaugh, 1997; Martin and Kandel, 1996).

Short-term responses of the brain to novel, arousing, or potentially threatening situations are adaptive and result in enhanced learning and in the acquisition of new behavioral strategies for coping. However, repeated stress can increase allostatic load and cause cognitive deficits. The hippocampus is important in declarative, spatial, and contextual memory processes. It also works in processing the contextual associations of strong emotions (Eichenbaum, 1997; Gray, 1982). Atrophy of the hippocampus with age has been reported in animals (Meaney et al., 1988) and humans (Lupien et al., 1998; Lupien et al., 1994) and is accompanied by cognitive impairment. The cumulative effects of life stress, as expressed by the concept of allostatic load, can cause impairment by at least four mechanisms (McEwen, 1997, 1999b): impairing neural excitability, causing atrophy of neurons in the Ammon's horn region of the hippocampus, inhibiting neurogenesis in the dentate gyrus of the hippocampus, and causing permanent loss of neurons in the hippocampus. Each process can occur somewhat independently of the others, and each contributes to some degree to different pathophysiologic conditions associated with traumatic stress, depression, or aging.

Those laboratory findings have been carried over to the human brain by magnetic resonance imaging. Hippocampal atrophy and cognitive impairment have been reported in conditions as diverse as Cushing's syndrome, post-traumatic stress disorder, and recurrent major depression (for reviews, see McEwen et al., 1997; Sapolsky, 1996). The hippocampus is susceptible to those effects and is likely not the only brain region so affected. Atrophy of the amygdala and the prefrontal cortex also has been reported in depressive illness (Drevets et al., 1997; Sheline et al., 1998, 1999). The reversibility or prevention of such atrophy clearly is an important topic for research, as are its implications for cognitive function.

IMMUNE SYSTEM FUNCTION IN HEALTH AND DISEASE

The rebirth of integrative physiology in the context of the growing pool of information about genes and gene regulation has led to the development of new interdisciplinary fields of study, such as neuroendocrine immunology and psychoneuroimmunology—the latter includes psychology and an eclectic mix of other disciplines. At the same time, the more traditional field of immunology has advanced with growing knowledge about the messengers of the immune system, the cytokines and chemokines. Those advances underline the importance of explaining how the immune system functions in the living body, as opposed to examining only its theoretical workings. The immune system is highly integrated with other physiologic systems. It is sensitive to virtually every hormone in the body, and sympathetic, parasympathetic, and sensory nerves innervate the organs of the immune system. These organs also produce hormones that affect cells in the rest of the body. The remarkable capacity to distinguish “self” from “nonself” to protect the body from infectious and malignant challenges is a hallmark of the immune system that has been studied in great detail at the cellular and molecular levels over the past 50 years.

Much progress has been made recently in explaining how the body responds to environmental challenges through immunologic pathways. Research has led to development of the notion of a “danger” hypothesis in which tissue damage (and perhaps a threat to the very survival of the organism) is a signal that activates inflammatory and immune responses (Fuchs and Matzinger, 1996). The response of the immune system to environmental challenge represents a coordinated pattern of gene expression that results in rapid and sustained production of activated cells, secretory products, and effector mechanisms that persist until the challenge is eliminated.

Like many other bodily and behavioral responses to challenge, immune system responses are initially beneficial, but when they are sustained, they can damage healthy host tissue or augment damage produced by pathogens. That potential threat is reduced by the high level of regulation of the immune system and its close integration with the autonomic and endocrine systems. Bidirectional interaction of common chemical messengers and cellular receptors connects the immune system with the nervous system and the endocrine system. In this interactive communication between systems, sensory stimuli to the central nervous system that activates the HPA axis result in the peripheral release of adrenal steroids and catecholamines, both of which can have immunoregulatory effects. Similarly, a challenge that induces an inflammatory response causes the release of cytokines that stimulate the peripheral and central nervous systems (Figure 2-3). This provides an important link through which the neuroendocrine response modulates the development of an inflammatory response at the site of challenge. That is especially important in the lungs, heart, brain, and kidneys, where there are limits to the magnitude an inflammatory response can reach without causing loss of critical function (Hermann et al., 1995).

FIGURE 2-3. Neuroendocrine Regulation of Immunity.

FIGURE 2-3

Neuroendocrine Regulation of Immunity. Tissue damage, infection, and malignancy are hypothesized to generate a danger signal. This signal is transduced at the cellular and molecular levels leading to activation, trafficking and expression of effector (more...)

The movement of immune cells to organs and tissues where they are needed to fight infection or other challenge is highly regulated at multiple levels, including interactions among neural, endocrine, and immune systems (Fauci, 1975; Hermann et al., 1995). Steroid hormones and catecholamines released from the adrenals promote the movement of immune cells. These steroid hormones and catecholamines are immuno-suppressive when they are secreted chronically or when their release is not terminated properly—in which case they can actually increase susceptibility to infection and malignancy (Dobbs et al., 1993; Hermann et al., 1993). A deficiency of circulating glucocorticoids and catecholamines, however, allows other immune mediators, such as cytokines, to overreact and thereby increase the risk of autoimmune and inflammatory disorders (Sternberg et al., 1992, 1989; Wilder, 1995).

Neural and Endocrine Effects on the Immune System

The immune system is integrated with the nervous and endocrine systems, which modulate aspects of its function. Many chemical messengers of the nervous and endocrine systems are immunomodulatory, and these substances are important in regulating inflammatory and immune responses (Figure 2.3) (Felten et al., 1987). The presence of receptors— for hormones, neuropeptides, and neurotransmitters—on cells of the immune system suggests some ways by which modulation occurs. The receptors provide a mechanism for other physiologic systems to modulate immune responses. Activation of the HPA axis, which leads to systemic release of potent anti-inflammatory substances, such as cortisol and other glucocorticoids, regulates inflammatory and immune responses (Berczi, 1998). The autonomic system's part in immunoregulation also became clearer once it was demonstrated that primary and secondary lymphoid tissues are highly innervated (Bulloch and Pomerantz, 1984; Felten et al., 1987) and that immune cells in these tissues have receptors for neuro-modulators and neuro transmitters.

Effects of Inflammatory and Immune Responses on the Nervous System

Gaining a clear understanding of the bidirectional flow of neuroendocrine and immune interactions had to wait for a revolution in cellular and molecular immunology. Advances in cytokine research provide a context for the study of soluble products produced by activated monocytes and lymphoid cells (circulating and stationary immune cells, respectively).

These low-molecular-weight products of mononuclear cells are secreted during inflammatory and immune responses. They were shown to be produced by several types of cells; to have several kinds of biologic activity; and to be pleiotropic, affecting cells in many physiologic systems (Arai et al., 1990). The cytokines initiate action by binding to specific receptors on target cell surfaces, and many of them have multiple signaling functions including autocrine (secretion of a substance that stimulates the secretory cell itself), paracrine (the target cell is close to the secretory cell, such as neurotransmitters in the brain), and endocrine (the chemical signal can travel long distances from the secretory cells to the target tissue, via the blood or lymph systems). The pleiotropy of cytokines led to investigations of inflammatory and immune interactions with the nervous system and to the development of the idea that the immune system communicates with the brain through the release of proinflammatory cytokines (Besedovsky et al., 1986). Proinflammatory cytokines released in peripheral tissues function as hormones, and biologically are associated with the development and expression of behaviors associated with illness (Dantzer et al., 1998; Maier et al., 1998) and can induce chronic stress responses (Shanks et al., 1998).

The recognition that cytokines, particularly those that are proinflammatory, communicate with the brain and influence activation of the HPA axis led to investigations of neuroendocrine responses in the etiology of inflammatory (Chrousos, 1995) and autoimmune diseases. Associations between stress responses and autoimmunity have been studied in conditions as diverse as rheumatoid arthritis (Heijnen et al., 1996; Sternberg et al., 1989, 1992), inflammatory bowel disease (Anton and Shanahan, 1998), systemic lupus erythematosus (Utz et al., 1997), and multiple sclerosis (Griffin and Whitacre, 1991). Other studies have examined neuroendocrine responses and immune system reactivity in asthma (Barnes, 1986; Busse et al., 1995; Kang et al., 1998), atopic dermatitis (Buske-Kirschbaum et al., 1998), and allergy (Anderzen et al., 1997).

Stress and Immune System Function

The recognition of the importance of bidirectional communication between neural, endocrine, and immune systems through shared ligands and receptors led to a major research emphasis on immunoregulation by hormones, peptide neuromodulators, and neurotransmitters. The primary function of the immune system is to protect the host from infectious and malignant challenges. Acute stress enhances immune function, and it does so in part by promoting immune cell translocation to sites of immune challenge (Dhabhar et al., 1995, 1996), whereas chronic stress has the opposite effect: it impairs immune function (Dhabhar and McEwen, 1999; Hermann et al., 1995). Various aspects of immune function in states of stress-induced neuroendocrine activation, with a primary emphasis on negative, immunosuppressive outcomes, have been reported (Dobbs et al., 1993; Kiecolt-Glaser et al., 1996).

One important factor in whether a person will develop respiratory infection after challenge with an infectious virus is lack of social support (Cohen, 1995; Cohen et al., 1991). In studies of two-party relationships, marital discord was found to affect general health and immunity significantly (Kiecolt-Glaser et al., 1997).

Because immune system functioning is studied best when the system has been provoked, examinations of stress and immunity have considered the effectiveness and durability of the immune response after the administration of vaccines in humans. The stress of taking a university examination (Glaser et al., 1992) and the chronic stress of being a caregiver (Kiecolt-Glaser et al., 1996; Glaser et al., 1998) were used to characterize responses to vaccination during a stressful period. In each case, the anti-body responses to vaccines were poorer in the stressed than in the nonstressed groups.

Similar studies of the mechanisms of neuroendocrine/immune interactions have been performed in animals. Studies of experimental viral infections in mice demonstrate that both the HPA axis and the sympathetic system alter virus-induced pathophysiology under conditions of imposed experimental stress (Hermann et al., 1993). The stress response also can suppress specific components of natural resistance and adaptive immune responses to viral infection, both acute (Dobbs et al., 1993; Sheridan et al., 1991) and latent (Bonneau et al., 1991; Kusnecov et al., 1992). Environmental stress suppresses immunity and enhances the pathogenicity of bacteria, particularly that caused by facultative intracellular parasites, such as mycobacteria (Brown et al., 1993).

Wound healing and other physiologic processes that require substantial proinflammatory responses (including cytokine, chemokine, and growth factor gene expression) are affected by environmental and behavioral stress. The stress of long-term caregiving to dementia patients delays the healing of full-thickness, cutaneous-punch biopsy wounds (Kiecolt-Glaser et al., 1995). Acute stress induced by taking academic examinations delayed the healing of mucosal wounds in the oral cavity; the delay was associated with diminished proinflammatory cytokine responses in the peripheral blood of those who experienced the stress (Marucha et al., 1998). In animal models, restraint stress caused activation of the HPA axis, which was shown to suppress movement of immune cells to wound sites (Padgett et al., 1998).

The effects of disaster-related stress responses on the immune system have been studied (Ironson et al., 1997; Solomon et al., 1997). Major effects of distress of natural disasters include alterations in natural and adaptive immunity, as indicated by lower natural killer-cell cytotoxicity (NKCC) and lower numbers of circulating T lymphocytes. People who were tested after surviving Hurricane Andrew had lower NKCC and fewer suppressor T cells (CD8+) and helper T cells (CD4+) than did comparison subjects (Ironson et al., 1997). Alterations in NKCC were related to psychologic and behavioral factors: Survivors reported greater loss of resources, greater post traumatic disorder symptomology, and more negative intrusive thoughts than did control subjects. Those observations are consistent with conclusions drawn from a growing literature on psychologic stressors and immunity, which has shown NKCC to be diminished by bereavement (Irwin et al., 1987), marital discord (Kiecolt-Glaser et al., 1987), and exposure to earthquakes (Solomon et al., 1997).

The observed reductions in measures of natural and adaptive immunity were statistically significant in a stressed population but did not suggest increased risk for infection or disease in any individual. However, these studies demonstrate that natural disasters, industrial accidents, and psychosocial events are stressors that can affect human immunophysiology and thereby affect both mental and physical well-being.

ADDITIONAL FACTORS INFLUENCE LONG-TERM EFFECTS OF STRESS

Resilience

People differ widely in resilience to and recovery from illness, injury, or surgery and in overcoming adversity. However, relatively little is known about the physiology of resilience and good health. Resilience undoubtedly consists of more than just the absence of allostatic load. It is thought to be the product of cellular processes that protect and build cells and tissues—processes that involve some reserve capacity and resistance to the damaging effects of stressors. Promising research includes how anabolic hormones, such as growth hormone and insulin, and neurotrophic factors work in the brain as they are related to voluntary exercise and to recovery from injury and illness. For example, voluntary exercise in rats (running in an activity wheel) increases expression of messenger RNA for a neurotrophin that protects neurons from death and that promotes neuroplasticity and synaptic transmission (Oliff et al., 1998). It is not known, however, what advantages increased neurotrophin concentrations confer on the brains of exercising animals. The animals might be more resilient in the face of severe stress, or their brains might deteriorate more slowly with age. Although the role of neurotrophin regulation in exercise is not known, it has been reported that voluntary exercise increases production of new neurons in the dentate gyrus of the hippocampus, the brain region that is important in spatial and declarative memory (van Praag et al., 1999).

The study of factors that promote resilience, still poorly defined, is important as a complement to the more traditional approach of studying the damaging effects of stress mediators. Therefore, it will be important for research to relate human life histories, stress, and allostatic load to the production of such substances as the neurotrophins, which are related to tissue growth and repair. It also will be important to identify the influence of social support mechanisms and of individual attitudes that promote beneficial physiologic states associated with the capacity to repair damaged tissues and to protect against pathogens and toxic agents, such as free radicals (Epel et al., 1998; Ryff and Singer, 1998; Seeman and McEwen, 1996; Singer et al., 1998; Taylor et al., 1997).

Coping

Coping efforts are important moderators of the impact of stress on health (Baum and Posluszny, 1999). Coping is defined as volitional management of stressful events or conditions and regulation of cognitive, behavioral, emotional, and physiological responses to stress (Compas et al., 1999; Lazarus and Folkman, 1984). Various classifications of coping responses have been proposed, including coping to solve a problem versus coping to manage emotions, cognitive versus behavioral coping, approach versus avoidance coping, and coping intended to achieve (primary) control over the stressor (the source of stress) versus (secondary) control over response to the stress (emotions).

Coping efforts are important in the process of adaptation to illness. Several consistent findings have emerged from prospective longitudinal studies of breast cancer patients from diagnosis through treatment and recovery (Carver et al., 1993; Epping-Jordan et al., 1999; Stanton and Snider, 1993). Successful coping is facilitated by optimism—the tendency to anticipate positive outcomes. Through the use of strategies including acceptance, positive thinking, and problem solving, optimism is associated with lower psychological distress (reduced symptoms of anxiety and depression). Conversely, pessimistic thinking is associated with coping that involves avoidance and social withdrawal, which are related to higher symptoms of anxiety and depression (Carver et al., 1993; Epping-Jordan et al., 1999). Patients who are more prone to poor coping have histories of social isolation, recent losses, or multiple obligations (Rowland, 1990).

Breast cancer patients who learn to use more direct and confrontational coping strategies are less distressed than are those who use avoidance and denial (Holland and Rowland, 1990). Furthermore, a “fighting spirit” about the illness leads to a probability of longer survival (Green and Berlin, 1987; Greer et al., 1979; Watson et al., 1990). Research suggests that the belief that one has control over the cause of the disease leads to poor outcome, whereas belief in control over the course of the disease leads to better outcome (Watson et al., 1990). Psychosocial stress has been reported to lead to higher relapse rates in metastatic breast cancer (Ramirez et al., 1989). However, several studies report no significant effect of psychosocial variables on the course of carcinoma (Angell, 1985; Cassileth et al., 1985; Jamison et al., 1987).

Although stress can affect immune function and health, most of the observed effects are relatively small and within the range of normal immune function (Glaser et al., 1999). Therefore, stress-induced immune system changes that are related to disease are likely to be the result of multiple small simultaneous changes in the immune system. Measurement of multiple aspects of the immune system and their interactions is thus necessary to reveal the subtle and complex relationships among stress, immune function, and disease. Study of the effects of coping efforts on stress and immunologic responses is also important because coping might be a crucial mediator of the stress/immune relationship that can be modified through behavioral interventions.

In addition to possible effects on disease onset and etiology, stress also can disrupt the behaviors that normally enhance healthy functioning and protect a person from illness (discussed further in Chapter 3). This can be seen in the association between stress and health risk behaviors, such as smoking (Shiffman et al., 1996), poor diet, and lack of exercise (Greeno and Wing, 1994), and in health-compromising responses to stress that include increased autonomic arousal and elevated blood pressure (Baum and Posluszny, 1999).

Extensive research documents that expression of emotions has beneficial effects on both emotional and physical well-being (Esterling et al., 1999; Pennebaker, 1997). But emotional regulation does not involve the unmodulated ventilation of emotions or the containment or suppression of feelings; rather, successful regulation of emotion appears to involve the controlled and modulated expression and release of feelings in ways that contribute to an increased understanding of those emotions and their meaning. Research by Pennebaker and colleagues (1997) shows that writing about deep feelings is a powerful way to regulate emotional expression. Writing about emotions is associated with improved mood, fewer health problems, and enhanced immune function (Petrie et al., 1995, 1998). The specific mechanisms through which regulation of emotional expression affects health are not fully understood and are the subject of continuing research. However, the regulated expression of emotions through writing is a potentially important component of interventions to change health behavior.

Gender

In animal models and possibly also in humans, there are gender differences in vulnerability to brain damage or brain remodeling as a result of stress (see Galea et al., 1997; Uno et al., 1989). Although the gonadal hormones are important influences in the development of gender differences in early life, hormones and experience also can change brain structure and function in adult life (Greenough and Bailey, 1988; McEwen and Alves, 1999; McEwen, 1999b)—an indication that there is considerable life-long plasticity in the nervous system. And although the role of hormones is a hallmark of sexual differentiation, experience and social factors also are critical, especially in humans and nonhuman primates (Goy, 1970; McEwen, 1999a; Reinisch et al., 1987).

Social Influences

People live in social groups and as members of societies. It is well known that social class or socioeconomic position has a profound effect on health through multiple pathways (Adler et al., 1994; Antonovsky, 1967; Marmot et al., 1991; Syme and Berkman, 1976). Since the 1982 predecessor to this report (IOM, 1982), it has become evident that the degree of a given country's social inequality is related to health in that society (Kaplan et al., 1996; Wilkinson, 1992). And the degree of social integration or connection and the social networks in which people are embedded are related to morbidity and mortality (Berkman, 1995; House et al., 1988). Like economic inequality, social cohesion and social capital are associated with health (Kawachi et al., 1997). Moreover, there are characteristics of the work environment that can produce job stress and significantly influence workers' health (Karasek and Theorell, 1990).

These topics are developed more fully in Chapter 4. Physiologic systems that could mediate the effects of stressful social circumstances on health are discussed here.

CARDIOVASCULAR HEALTH AND DISEASE

Cardiovascular health and disease provide an example of the interactions of behavioral, psychologic, and social factors. This discussion of CHD will be used to point out the biological effects of stress and the psychosocial influences that exist. Despite progress in elucidating the role of genetics in human disease, it is clear that no single cause of CHD can be identified and that these conditions develop as a result of complex interactions among multiple factors.

One example of the effects of disparate factors on the incidence of cardiovascular disease is provided by a recent analysis of changing mortality patterns in Russia (Notzon et al., 1998). Over a 4-year period after the breakup of the former Soviet Union, mean life expectancy declined by 5 years. Most of the decline could be attributed to increased mortality in men aged 25 to 64 because of accidents and cardiovascular disease. Factors implicated in the dramatic change in death rates included economic instability, chronic stress, depression, and the increased use of alcohol and tobacco.

Stress and Cardiovascular Function

Stress clearly is important in cardiovascular health and disease. There is general agreement that acute stress can trigger acute cardiovascular events (Muller and Tofler, 1990), but the more subtle influences of chronic stress and allostatic load are not well understood. The effects of psychosocial stressors are mediated through the central nervous system, so it is relevant to review several pathways through which the brain affects bodily processes related to cardiovascular function. Much new information about function and measurement has contributed to our explanation of the relationships.

The autonomic nervous system regulates internal bodily functions, including all aspects of cardiovascular function. The autonomic system maintains appropriate internal states (homeostasis) and enables the body to respond to external threats perceived as stressors. It has two primary divisions: the sympathetic and parasympathetic. The sympathetic nervous system permits response to extreme conditions: fight or flight. The parasympathetic nervous system modulates functions under resting conditions. Both blood pressure and heart rate are modulated through the autonomic nervous system.

There is strong evidence that increased sympathetic activity is a feature of many cases of hypertension in young adults. Cardiac output increases in the early stages of hypertension and decreases with advancing age. With age, peripheral vascular resistance increases, largely because of remodeling (rerouting) and hypertrophy (overgrowth) of blood vessel walls. Sympathetic activity also can affect the development of atherosclerosis. Mechanisms include increasing insulin resistance, a known risk factor for cardiovascular disease; hemodynamic effects on the arterial wall; and direct metabolic effects, such as increased plasma triglycerides and alteration in the metabolism of low-density lipoproteins (Julius, 1993). Furthermore, increased sympathetic activity can increase the risk of cardiovascular disease through the effects of adrenal epinephrine on platelet aggregation and the development of left ventricular hypertrophy. There is experimental evidence that increased heart rate (Beere et al., 1984) and increased blood pressure variability are both risk factors for atherosclerosis (Sasaki et al., 1994). Decreased heart rate variability itself is associated with the presence of CHD and is a risk factor for cardiovascular morbidity. But it is not known whether this association is causal. For example, reduced heart rate variability might be a consequence of artherosclerotic damage to the carotid sinuses, which could cause impaired baroreceptor reflexes.

Laboratory studies demonstrate that cardiovascular disease can be produced by chronic social stress. Hypertension can be elicited in some strains of mice, but not in others (Henry et al., 1986), and the hypertensive consequences of behavioral stress can be potentiated in genetically normotensive animals by a high-sodium diet (Anderson, 1994). In animals, the combination of emotional stress and high sodium intake has been associated with a greater increase in blood pressure than results from either factor alone (Staesson et al., 1994).

Other studies show that subordinate female cynomolgus monkeys have more atherosclerosis than do dominant females, and the difference appears to be related to suppression of the release of cardioprotective ovarian hormones (Shively and Clarkson, 1994). Atherosclerosis develops faster in dominant male monkeys when they are defending their social position or re-establishing it in an unstable social hierarchy (Manuck et al., 1995). The combination of a high-fat diet with psychosocial stress accelerates the disease process (Brindley and Holland, 1989).

Studies of chronic stress among people have yielded inconsistent findings: some show activation of the HPA axis and others show its suppression (Ockenfels et al., 1995). Although anticipation or experience of acute stress activates the HPA axis (Smyth et al., 1998), the degree of activation with repeated exposure to stress is greatly variable (Kirschbaum et al., 1995).

The importance of personality, emotion, and social environment in the development of cardiovascular disease is a subject of controversy, but there is evidence that anger, whether expressed openly or repressed, is associated with an increased risk of hypertension (Everson et al., 1998). Job-related stress is also important. The combination of high job demands and low control is associated with hypertension (Schnall et al., 1992). Blood pressure tends to be highest in the workplace, but the increase in blood pressure in people with high-strain jobs is seen at work, at home, and during sleep (Schnall et al., 1992). An imbalance between income and expenditure is associated with high blood pressure (Chin-Hong and McGarvey, 1996; Dressler, 1991).

The prevalence of hypertension in humans varies greatly from one society to another, and it appears to be strongly influenced by society and culture factors. For example, epidemiologic studies indicate that the transition from life in traditional tribal community to urbanized Western society is associated with an increase in blood pressure (Cruz-Coke, 1987; Poulter et al., 1988, 1990), although it is unclear whether this effect is the result of changes in diet or of psychosocial stress.

Behavioral and Psychosocial Factors

Psychosocial factors can influence the course of chronic human disease along several pathways. Behavior that has perceived short-term benefits, such as mood-enhancement induced by cigarette-smoking or excessive alcohol consumption, but that causes long-term injury constitutes one (Chapter 3). Another involves the influence of social and environmental factors, such as socioeconomic status or stress on disease processes (Chapter 4). A third consists of individual psychological factors, such as hostility and depression, that interact with the other two pathways to increase susceptibility to illness. The evidence for a role for these psychological factors in cardiovascular disease is described below.

Hostility

Hostility is the psychosocial variable most often associated with the incidence of CHD (Booth-Kewley and Friedman, 1987). In the context of physical health, hostility is defined usually as a stable attribute characterized by mistrusting cynicism that leads to antagonistic or aggressive behavior and feelings of anger (Miller et al., 1996). The extent to which hostility is a personality trait or a behavioral coping response to environmental stimuli, however, is not known. Most of the research on hostility has been done in men.

Interest in hostility and CHD evolved from earlier research on the type A behavior pattern, an idea originally formulated by Friedman and Rosenman (1974). Type A behavior was characterized by a sense of time urgency, loud and explosive speech, hostility, and competitiveness. Early studies supported an association between type A behavior and the development of CHD (Review Panel, 1981), but later research failed to confirm the association (Case et al., 1985; Shekelle et al., 1985a, 1985b). The original type A behavior data-set was reanalyzed by two teams of investigators to examine inconsistencies and identify variables within the multifaceted type A behavior patterns that were most predictive of CHD (Chesney et al., 1988; Matthews et al., 1977). These analyses revealed that hostility was the best variable for distinguishing men who developed heart disease from men who did not (Hecker et al., 1988; Matthews et al., 1977). Many prospective studies confirmed the relationship between hostility, as assessed by interviews and questionnaires, and CHD incidence (Barefoot et al., 1983; 1989; Dembroski et al., 1989; Houston and Kelly, 1987; Shekelle et al., 1983). Significant associations also have been found between hostility and cardiac mortality (Koskenvuo et al., 1988; Shekelle et al., 1991). Considered together, the cumulative findings constitute substantial evidence of the link between hostility and various aspects of CHD. Although some studies have not found an association (Hearn et al., 1989; Leon et al., 1988; McCranie et al., 1986), the positive reports outnumber negative ones (Scheier and Bridges, 1995). One reason for this inconsistency is that the assessment of hostility often relies on self reports, and people might tend to underreport this socially undesirable trait (Helmers et al., 1995).

There is a hypothesis that people who are hostile have exaggerated cardiovascular reactivity to stress and that this either contributes to the development of atherosclerosis (Matthews et al., 1998) or triggers acute events (Rozanski et al., 1999). However, hostility also is correlated with increased likelihood of smoking, with decreased likelihood of quitting smoking (Lipkus et al., 1994), and with lower socioeconomic status (Barefoot et al., 1991; Carroll et al., 1997). Each of these will increase allostatic load.

Anger

Anger is a psychological state thought to be related to hostility. Expression of anger has been shown to trigger myocardial infarction. In a study of patients undergoing coronary angiography, recall of anger was a potent stimulus that induced vasoconstriction in diseased coronary arteries, but not in healthy arteries (Boltwood et al., 1993). The recall of anger can also produce an acute impairment in ventricular function in patients with CHD (Ironson et al., 1992).

Vital Exhaustion

One common premonitory symptom of myocardial infarction is vital exhaustion, a state of excessive fatigue, increased irritability, and demoralization (Appels et al., 1987). A prospective study of 3877 city employees in Rotterdam, The Netherlands, compared the risk of coronary heart disease among those scoring in the highest third on a measurement scale of exhaustion to those with lower scores. Vital exhaustion predicted myocardial infarction with a relative risk of 2.28—a relatively robust effect for a behavioral predictor (Appels and Mulder, 1989). There appears to be no correlation between the severity of CHD and vital exhaustion score, so it is unlikely that subclinical coronary disease causes the observed fatigue (Kop et al., 1996). Vital exhaustion also has been reported to predict recurrence of arterial blockage after coronary angioplasty (Kop et al., 1994). Job stress is associated with vital exhaustion and is a risk factor for cardiovascular disease (Everson et al., 1997; Keltikangas-Jarvinen et al., 1996a, 1996b; Kop et al., 1998; Lynch et al., 1997a; Raikkonen et al., 1996).

Depression

Depression affects about half of patients who experience myocardial infarction. Depression predicts significantly poorer outcome with heart disease (Denollet et al., 1996; Denollet and Brutsaert, 1998; King, 1997) and roughly doubles the risk of recurrent cardiovascular events (Barefoot et al., 1996; Barefoot and Schroll, 1996; Frasure-Smith et al., 1995). About half of postinfarction patients with depression have a history of depression before the onset of CHD, and there is some evidence suggesting depression as a risk factor for a first infarction (Sesso et al., 1998). The association between depression and mortality seems to be the same in men and women (Frasure-Smith et al., 1999). However, the prevalence of postinfarction depression is about twice as high in women as in men (Carney et al., 1990). It is unlikely that depression is a consequence of CHD, inasmuch as the occurrence of depression often precedes any disease symptoms and there is no relationship between severity of depression and severity of coronary arterial disease (Carney et al., 1995).

Depression is associated with increased sympathetic and decreased parasympathetic tone, as manifested by increased plasma catecholamine concentrations, increased heart rate, and decreased heart rate variability. Myocardial infarctions tend to happen most commonly between 6 A.M. and noon, the time of day that parallels the normal circadian rhythm of sympathetic activity. But the cycles of catecholamines and cortisol are disturbed in people who have depression, peaking earlier in the day than in nondepressed people. Depressed people are more likely than nondepressed people to have myocardial infarction during the night or very early in the morning (Carney et al., 1995).

Twenty years ago it was suggested that the presence of depression predicted a higher subsequent incidence of cancer (Shekelle et al., 1981). Although a large cohort study of employees at Western Electric reported an elevated rate of subsequent cancers among those diagnosed with depression, this finding was not confirmed in a more recent large-scale cohort trial (Zonderman et al., 1989) found no relationship between two measures of depressive symptoms and cancer morbidity or mortality in a large population. The researchers used continuous and not categorical measures of depression, leaving open the possibility that severe clinical depression could be associated with elevated cancer risk. However, this and earlier studies lend little support to the idea that depression increases cancer risk (Fox, 1989). Fox's reanalysis of the original observation suggests that a combination of depression and exposure to toxins could have accounted for the apparent association (Fox, 1989). However, a study by Penninx et al. (1998) did find in a sample of 5000 elderly people that consistent symptoms of depression were predictive of an almost 2-fold elevation in risk of cancer incidence. Thus, depression does not seem to predict cancer incidence, but it is elevated among those who have cancer.

Anxiety, Worry, and Hope

Anxiety and worry have recently received renewed attention as risk factors for cardiovascular disease. Two prospective studies have shown that anxiety predicts the development of CHD (Sloan et al., 1999), and worrying is an important component of anxiety. Men who worry a lot were found to be at increased risk for CHD (Kubzansky et al., 1997).

Hope and optimism, in contrast, have been suggested as important components of psychological well-being and as factors that can contribute to good physical health (Scheier and Carver, 1985; Snyder et al., 1991). A lack of hope is commonly thought to adversely affect health (Scheier and Carver, 1992). However, only recently has there been empirical support for this. One major challenge for researchers and health care providers was to develop ways to measure hope and hopelessness. Hopelessness, as assessed by one question on a four-item questionnaire designed to measure depressed affect, reliably predicted fatal and nonfatal CHD events in a cohort of more than 2800 initially healthy men and women (Anda et al., 1993). Similarly, a two-item hopelessness scale significantly predicted all-cause mortality, the incidence of myocardial infarction and cancer, and death from violence and injury in a sample of 2428 men in the Kuopio ischemic heart disease study in Finland (Everson et al., 1996). Those and similar findings support the general idea that psychosocial factors are important determinants of physical health and disease.

DEVELOPMENTAL TRAJECTORIES

Development is important in the biological and behavioral processes that preserve health or lead to human disease throughout life. Cumulative experience, adaptive plasticity, physical and social exchange with surrounding environments, and genetic predisposition interact to influence development. The unique physiology of each person, partly encoded in the genome and partly determined by prior physical exposures and social experiences, generates the individual behaviors that influence morbidity and mortality.

Although developmental status is a continuing factor in health outcomes over life, it has heightened salience for immediate and long-term health responses during infancy, early childhood, and adolescence. These periods are characterized by extremely rapid biological and psychosocial change. The resolution of the developmental challenges faced at these times determine set points for homeostatic systems, as well as for adopting crucial health-related attitudes and behaviors. These outcomes, in turn, determine trajectories for subsequent biobehavioral functioning that can have long-term effects on health. Therefore, the periods of infancy, early childhood, and adolescence are highlighted here.

Early childhood, infancy, and even prenatal experiences appear to have long-term consequences for health because they influence the biological mechanisms that underlie stress reactivity. There are sociocultural consequences of these experiences as well (NRC, 2000). A secure attachment with a parenting person provides a protective modulator of the environmental influences on an infant. The mother/child attachment is affected by the infant's temperament, which is characterized by reactions to stimuli, the tone of the emotional expression (positive or negative), activity level, and sociability. There is increasing evidence that temperament has a biological base (Boyce et al., 1992), including a genetic component that is heavily influenced by experience (van der Boom, 1994). The minority of infants who have difficult temperaments can experience attachment problems and high levels of stress, with consequences for their stress responses as adults. Unresponsive, insensitive, or abusive parenting also can lead to atypical emotional development. Research with infants of depressed mothers, for example, has shown that diminished parental responsiveness is associated with changes in infant emotional regulation and the balance of left vs. right frontal cortical activation (Dawson et al., 1997). A disproportionate number of severely deprived infants raised in Romanian orphanages exhibited abnormalities in cognitive, emotional, and linguistic development (Fisher et al., 1997). They also showed retarded growth and higher rates of infectious diseases (Albers et al., 1997; Hostetter et al, 1991).

Stress in young children can influence future health. Although currently limited to retrospective analyses, studies of physical and sexual abuse in childhood suggest variable but elevated risk for later depression, somatization, excessive rates of health care use, homelessness, and other indicators of behavioral maladaptation (Cheatsy et al., 1998; Herman et al., 1997; Salmon and Calderbank, 1996; Styron and Janoff-Bulman, 1997). A recent prospective controlled study (Heim et al., 2000) found that, during a relatively mild stress task involving public speaking and mental arithmetic, those with a history of sexual abuse showed significantly elevated concentrations of adrenocorticotropic hormone. This was most pronounced in participants who exhibited symptoms of major depression at the time of testing. The results provide support for increased stress vulnerability and altered HPA axis function in adults who were abused as children (Heim et al., 2000). Elevated prolactin response to serotonin challenge has been found among abused children (Kaufman et al., 1998; Pine et al., 1997), suggesting that the experience could be associated with dysregulation of serotonergic neurotransmission. Activation of specific neuroendocrine systems, such as the HPA axis, also has been found in conditions of normative, acute stress, such as that which accompanies the transition to primary school (Boyce et al., 1995), and prolonged extreme neglect and sensory deprivation, such as adverse rearing in a Romanian orphanage (Gunnar, 1998). Furthermore, chronic stressors in early childhood could impair the emergence of higher cognitive processes such as memory (Nelson and Carver, 1998). The potential for “recovery” from prolonged, severe early adverse experiences through later enrichment is still unknown. All are important areas for research.

Behavioral factors, such as physical activity and diet, are significant in setting health trajectories in children. Nearly half of American children are not regularly physically active, and physical activity declines dramatically among older children (U.S. Department of Health and Human Services, 1996). In the United States, children currently are 20–30% less active than is recommended by the World Health Organization (Salbe et al., 1997). Childhood activity is associated with relative weight, parental obesity, and the proportion of time spent outdoors (Klesges et al., 1990). These statistics are important because the prevalence of overweight children between the ages of 6 and 11 in the National Health and Nutrition Examination Study increased from 5% to 22% between 1976–1980 and 1988–1991 (Troiano et al., 1995). In the Bogalusa Heart Study, the prevalence of overweight 5- to 24-year-olds doubled between 1973 and 1994 (Freedman et al., 1997). Evidence is growing that obesity in childhood has psychosocial consequences and portends greater risk of disease in adulthood (for reviews, see Dietz, 1998; Must and Strauss, 1999).

Adolescence provides another important period for promoting healthy behaviors. Many of the behaviors associated with adult morbidities and even mortality, such as cigarette smoking, alcohol and drug abuse, unsafe sexual practices, and violent or aggressive responses to stress often begin in adolescence. But because adolescents, in general, are curious about and interested in their bodies, that time of life also provides opportunities to promote good health and to involve young people in decision making about themselves. Effective early intervention can prevent the onset of health-compromising behaviors and can work to prevent their becoming less firmly established as life-long patterns (Millstein et al., 1993).

REFERENCES

  1. Ader R. Effects of early experiences on emotional and physiological reactivity in the rat. Journal of Comparative Physiology and Psychology. 1968;66:264–268. [PubMed: 5722035]
  2. Adler NE, Boyce T, Chesney MA, Cohen S, Folkman S, Kahn RL, Syme LS. Socioeconomic status and health: The challenge of the gradient. American Psychologist. 1994;49:15–24. [PubMed: 8122813]
  3. Albeck DS, McKittrick CR, Blanchard DC, Blanchard RJ, Nikulina J, McEwen BS, Sakai RR. Chronic social stress alters levels of corticotropin-releasing factor and arginine vasopressin mRNA in rat brain. Journal of Neuroscience. 1997;17:4895–4903. [PMC free article: PMC6573358] [PubMed: 9169547]
  4. Albers L, Johnson DE, Hostetter M, Iverson S, Georgieff M, Miller L. Health of children adopted from the former Soviet Union and Eastern Europe: Comparison with pre-adoptive medical records. Journal of the American Medical Association. 1997;278:922–924. [PubMed: 9302245]
  5. Anda R, Williamson D, Jones D, Macera C, Eaker E, Glassman A. Depressed affect, hopelessness, and the risk of ischemic heart disease in a cohort of U.S. adults. Epidemiology. 1993;4:285–294. [PubMed: 8347738]
  6. Anderson DE. Behavior analysis and the search for the origins of hypertension. Journal of the Experimental Analysis of Behavior. 1994;61:255–261. [PMC free article: PMC1334413] [PubMed: 8169574]
  7. Anderzen I, Arnetz BB, Soderstrom T, Soderman E. Stress and sensitization in children: a controlled prospective psychophysiological study of children exposed to international relocation. Journal of Psychosomatic Research. 1997;43:259–269. [PubMed: 9304552]
  8. Angell M. Disease as a reflection of the psyche. New England Journal of Medicine. 1985;312:1570–1572. [PubMed: 4000189]
  9. Anton PA, Shanahan F. Neuroimmunomodulation in inflammatory bowel disease. How far from “bench” to “bedside”? Annals of the New York Academy of Sciences. 1998;840:723–734. [PubMed: 9629299]
  10. Antonovsky A. Social class, life expectancy, and overall mortality. Milbank Memorial Fund Quarterly. 1967;45:31–73. [PubMed: 6034566]
  11. Appels A, Mulder P. Fatigue and heart disease. The association between ‘vital exhaustion' and past, present and future coronary heart disease. Journal of Psychosomatic Research. 1989;33:727–738. [PubMed: 2621676]
  12. Appels A, Hoppener P, Mulder P. A questionnaire to assess premonitory symptoms of myocardial infarction. International Journal of Cardiology. 1987;17:15–24. [PubMed: 3666994]
  13. Arai K, Lee F, Miyajima A, Miyatake S, Arai N, Yokota T. Cytokines: Coordinators of immune and inflammatory responses. Annual Review of Biochemistry. 1990;59:783–836. [PubMed: 1695833]
  14. Barefoot JC, Schroll M. Symptoms of depression, acute myocardial infarction, and total mortality in a community sample. Circulation. 1996;93:1976–1980. [PubMed: 8640971]
  15. Barefoot JC, Dahlstrom WG, Williams RBJ. Hostility, CHD incidence, and total mortality: a 25-year follow-up study of 255 physicians. PsychosomaticMedicine. 1983;45:59–63. [PubMed: 6844529]
  16. Barefoot JC, Helms MJ, Mark DB, Blumenthal JA, Califf RM, Haney TL, O'Connor CM, Siegler IC, Williams RB. Depression and long-term mortality risk in patients with coronary artery disease. American Journal of Cardiology. 1996;78:613–617. [PubMed: 8831391]
  17. Barefoot JC, Peterson BL, Dahlstrom WG, Siegler IC, Anderson NB, Williams RB. Hostility patterns and health implications: Correlates of Cook-Medley Hostility Scale scores in a national survey. Health Psychology. 1991;10:18–24. [PubMed: 2026126]
  18. Barefoot JC, Peterson BL, Harrell FEJ, Hlatky MA, Pryor DB, Haney TL, Blumenthal JA, Siegler IC, Williams RBJ. Type A behavior and survival: A follow-up study of 1,467 patients with coronary artery disease. American Journal of Cardiology. 1989;64:427–432. [PubMed: 2773785]
  19. Barker DJP. The fetal origins of coronary heart disease. Acta Paediatrica Supplement. 1997;422:78–82. [PubMed: 9298799]
  20. Barker DJP, Sultan HY. Fetal programming of human disease. In: Hanson MA, Spencer JAD, Rodeck JH, editors. Growth. Cambridge, UK: Cambridge University Press; 1995.
  21. Barnes PJ. Asthma as an axon reflex. Lancet. 1986;1(8475):242–245. [PubMed: 2418322]
  22. Baum A, Posluszny DM. Health psychology: Mapping biobehavioral contributions to health and illness. Annual Review of Psychology. 1999;50:137–164. [PubMed: 10074676]
  23. Beere PA, Glagov S, Zarins CK. Retarding effect of lowered heart rate on coronary atherosclerosis. Science. 1984;226:180–182. [PubMed: 6484569]
  24. Berczi I. The stress concept and neuroimmunoregulation in modern biology. Annals of the New York Academy of Sciences. 1998;851:3–12. [PubMed: 9668600]
  25. Berkman LF. The role of social relations in health promotion. Psychosomatic Medicine. 1995;57:245–254. [PubMed: 7652125]
  26. Besedovsky H, del Rey A, Sorkin E, Dinarello CA. Immunoregulatory feedback between interleukin-1 and glucocorticoid hormones. Science. 1986;233:652–654. [PubMed: 3014662]
  27. Blanchard DC, Sakai RR, McEwen BS, Weiss SM, Blanchard RJ. Subordination stress: Behavioral, brain and neuroendocrine correlates. Behavioral Brain Research. 1993;58:113–121. [PubMed: 8136039]
  28. Boltwood MD, Taylor CB, Burke MB, Grogin H, Giacomini J. Anger report predicts coronary artery vasomotor response to mental stress in atherosclerotic segments. American Journal of Cardiology. 1993;72:1361–1365. [PubMed: 8256727]
  29. Bonneau RH, Sheridan JF, Feng N, Glaser R. Stress-induced effects on cell-mediated innate and adaptive memory components of the murine immune response to herpes simplex virus infection. Brain Behavior and Immunity. 1991;5:274–295. [PubMed: 1659472]
  30. Booth-Kewley S, Friedman HS. Psychological predictors of heart disease: A quantitative review. Psychological Bulletin. 1987;101:343–362. [PubMed: 3602244]
  31. Boyce WT, Adams S, Tschann JM, Cohen F, Wara D, Gunnar MR. Adrenocortical and behavioral predictors of immune responses to starting school. Pediatric Research. 1995;38:1009–1017. [PubMed: 8618776]
  32. Boyce WT, Barr RG, Zeltzer LK. Temperament and the psychobiology of childhood stress. Pediatrics. 1992;90:483–486. [PubMed: 1513612]
  33. Bremner JD, Randall P, Vermetten E, Staib L, Bronen RA, Mazure C, Capelli S, McCarthy G, Innis RB, Charney DS. Magnetic resonance imaging-based measurement of hippocampal volume in posttraumatic stress disorder related to childhood physical and sexual abuse—a preliminary report. Biological Psychiatry. 1997;41:23–32. [PMC free article: PMC3229101] [PubMed: 8988792]
  34. Brindley D, Rolland Y. Possible connections between stress, diabetes, obesity, hypertension and altered lipoprotein metabolism that may result in atherosclerosis. Clinical Science. 1989;77:453–461. [PubMed: 2684477]
  35. Brown DH, Sheridan JF, Pearl D, Zwilling BS. Regulation of Mycobacterial growth by the hypothalamus-pituitary-adrenal axis: Differential responses of Mycobacterium bovis BCG-resistant and susceptible mice. Infection and Immunity. 1993;61:4793–4800. [PMC free article: PMC281236] [PubMed: 8406880]
  36. Brown GL, Ebert MH, Goyer DC. Aggression, suicide and serotonin: relationships to CSF amine metabolites. American Journal of Psychiatry. 1982;139:741–746. [PubMed: 6177256]
  37. Bulloch K, Pomerantz W. Autonomic nervous system innervation of thymic related lymphoid tissue in wild-type and nude mice. Journal of Comparative Neurology. 1984;228:57–68. [PubMed: 6480908]
  38. Buske-Kirschbaum A, Jobst S, Hellhammer DH. Altered reactivity of the hypothalamus-pituitary-adrenal axis in patients with atopic dermatitis: Pathologic factor or symptom? Annals of the New York Academy of Sciences. 1998;840:747–754. [PubMed: 9629301]
  39. Busse WW, Kiecolt-Glaser JK, Coe C, Martin RJ, Weiss ST, Parker SR. NHLBI workshop summary: Stress and asthma. American Journal of Respiratory and Critical Care Medicine. 1995;151:249–252. [PubMed: 7812562]
  40. Cahill L, McGaugh JL. Mechanisms of emotional arousal and lasting declarative memory. Trends in Neurosciences. 1998;21:294–299. [PubMed: 9683321]
  41. Cahill L, Haier RJ, Fallon J, Alkire MT, Tang C, Keator D, Wu J, McGaugh JL. Amygdala activity at encoding correlated with long-term, free recall of emotional information. Proceedings of the National Academy of Science. 1996;93:8016–8021. [PMC free article: PMC38867] [PubMed: 8755595]
  42. Cahill L, Prins B, Weber M, McGaugh JL. Beta-adrenergic activation and memory for emotional events. Nature. 1994;371:702–704. [PubMed: 7935815]
  43. Cannon WB. The role of emotions in disease. Annals of Internal Medicine. 1936;11:1453–1465.
  44. Carney RM, Freedland KE, Jaffe AS. Insomnia and depression prior to myocardial infarction. Psychosomatic Medicine. 1990;52:603–609. [PubMed: 2287700]
  45. Carney RM, Freedland KE, Riggs B, Jaffe AS. Depression as a risk factor for cardiac events in established coronary heart disease: a review of possible mechanisms. Annals of Behavioral Medicine. 1995;17:142–149. [PubMed: 18425665]
  46. Carroll D, Smith GD, Sheffield D, Shipley MJ, Marmot MG. The relationship between socioeconomic status, hostility, and blood pressure reactions to mental stress in men: Data from the Whitehall II study. Health-Psychology. 1997;16:131–136. [PubMed: 9269883]
  47. Carver CS, Pozo C, Harris SD, Noriega V, Scheier MF, Robinson DS, Ketcham AS, Moffat FL Jr, Clark KC. How coping mediates the effect of optimism on distress: A study of women with early stage breast cancer. Journal of Personality and Social Psychology. 1993;65:375–390. [PubMed: 8366426]
  48. Multicenter Post-Infarction Research Group. Case RB, Heller SS, Case NB, Moss AJ. Type A behavior and survival after acute myocardial infarction. The New England Journal of Medicine. 1985;312:737–741. [PubMed: 3974650]
  49. Cassileth BR, Lusk EJ, Miller DS, Brown LL, Miller C. Psychosocial correlates of survival in advanced malignant disease? New England Journal of Medicine. 1985;312:1551–1555. [PubMed: 4000186]
  50. Cheasty M, Clare AW, Collins C. Relation between sexual abuse in childhood and adult depression: Case-control study. British Medical Journal. 1998;6:198–201. [PMC free article: PMC2665415] [PubMed: 9468687]
  51. Chesney MA, Hecker MHL, Black GW. Coronary-prone components of Type A behavior in the WCGS: A new methodology. In: Houston BK, Snyder CR, editors. Type A Behavior Pattern: Research, Theory, and Intervention. New York: Wiley; 1988. pp. 168–188.
  52. Chin-Hong PV, McGarvey ST. Lifestyle incongruity and adult blood pressure in Western Samoa. Psychosomatic Medicine. 1996;58:131–137. [PubMed: 8849629]
  53. Chrousos GP. The hypothalamic-pituitary-adrenal axis and immune-mediated inflammation. New England Journal of Medicine. 1995;332:1351–1362. [PubMed: 7715646]
  54. Chrousos GP. Stressors, stress and neuroendocrine integration of the adaptive response. Annals of the New York Academy of Science. 1998;851:311–335. [PubMed: 9668623]
  55. Clark KB, Krahl SE, Smith DC, Jensen RA. Post-training unilateral vagal stimulation enhances retention performance in the rat. Neurobiology of Learning and Memory. 1995;63:213–216. [PubMed: 7670833]
  56. Clark KB, Naritoku DK, Smith DC, Browning RA, Jensen RA. Enhanced recognition memory following vagus nerve stimulation in humans. Nature Neuroscience. 1999;2:94–98. [PubMed: 10195186]
  57. Cohen S. Psychological stress and susceptibility to upper respiratory infections. American Journal of Respiratory and Critical Care Medicine. 1995;152:S53–S58. [PubMed: 7551414]
  58. Cohen S, Herbert TB. Health psychology: Psychological factors and physical disease from the perspective of human psychoneuroimmunology. Annual Review of Psychology. 1996;47:113–142. [PubMed: 8624135]
  59. Cohen S, Tyrrell D, Smith A. Psychological stress and susceptibility to the common cold. New England Journal of Medicine. 1991;325:606–612. [PubMed: 1713648]
  60. Compas BE, Connor JK, Saltzman H, Thomsen AH, Wadsworth M. Getting specific about coping: Effortful and involuntary responses to stress in development. In: Lewis M, Ramsay D, editors. Soothing and Stress. Mahwah, NJ: Lawrence Erlbaum Associates; 1999. pp. 229–256.
  61. Cooper JR, Bloom FE, Roth RH. The Biochemical Basis of Neuropharmacology. New York: Oxford University Press; 1996.
  62. Cruz-Coke R. Correlation between prevalence of hypertension and degree of acculturation. Journal of Hypertension. 1987;5:47–50. [PubMed: 3584962]
  63. Dantzer R, Bluthe RM, Laye S, Bret-Dibat JL, Parnet P, Kelley KW. Cytokines and sickness behavior. Annals of the New York Academy of Sciences. 1998;840:586–590. [PubMed: 9629285]
  64. Dawson G, Frey K, Panagiotides H, Osterling J, Hessl D. Infants of depressed mothers exhibit atypical frontal brain activity: A replication and extension of previous findings. Journal of Child Psychology and Psychiatry. 1997;38:179–186. [PubMed: 9232464]
  65. De Bellis MD, Baum AS, Birmaher B, Keshavan MS, Eccard CH, Boring AM, Jenkins FJ, Ryan ND. A. E. Bennett research award. Developmental traumatology. Part I: Biological stress systems. Biological Psychiatry. 1999a;45:1259–1270. [PubMed: 10349032]
  66. De Bellis MD, Keshavan MS, Clark DB, Casey BJ, Giedd JN, Boring AM, Frustaci K, Ryan ND. A. E. Bennett research award. Developmental traumatology. Part II: Brain development. Biological Psychiatry. 1999b;45:1271–1284. [PubMed: 10349033]
  67. de Quervain DJ, Roozendaal B, McGaugh JL. Stress and glucocorticoids impair retrieval of long-term spatial memory. Nature. 1998;394:787–790. [PubMed: 9723618]
  68. Dellu F, Mayo W, Vallee M, LeMoal M, Simon H. Reactivity to novelty during youth as a predictive factor of cognitive impairment in the elderly: a longitudinal study in rats. Brain Research. 1994;653:51–56. [PubMed: 7982075]
  69. Dellu F, Mayo W, Vallee M, Maccari S, Piazza PV, Le Moal M, Simon H. Behavioral reactivity to novelty during youth as a predictive factor of stress-induced corticosterone secretion in the elderly—a life-span study in rats. Psychoneuroendocrinology. 1996;21:441–453. [PubMed: 8888367]
  70. Dembroski TM, MacDougall JM, Costa JM Jr, Grandits GA. Components of hostility as predictors of sudden death and myocardial infarction in the Multiple Risk Factor Intervention Trial. Psychosomatic Medicine. 1989;51:514–522. [PubMed: 2678209]
  71. Denenberg VH, Haltmeyer GC. Test of the monotonicity hypothesis concerning infantile stimulation and emotional reactivity. Journal of Comparative and Physiological Psychology. 1967;63:394–396. [PubMed: 6064382]
  72. Denollet J, Brutsaert DL. Personality, disease severity, and the risk of long-term cardiac events in patients with a decreased ejection fraction after myocardial infarction. Circulation. 1998;97:167–173. [PubMed: 9445169]
  73. Denollet J, Sys SU, Stroobant N, Rombouts H, Gillebert TC, Brutsaert DL. Personality as independent predictor of long-term mortality in patients with coronary heart disease. Lancet. 1996;347:417–421. [PubMed: 8618481]
  74. Dhabhar F, McEwen B. Enhancing versus suppressive effects of stresshormones on skin immune function. Proceedings of the National Academy of Sciences. 1999;96:1059–1064. [PMC free article: PMC15350] [PubMed: 9927693]
  75. Dhabhar FS, Miller AH, McEwen BS, Spencer RL. Effects of stress on immune cell distribution. Dynamics and hormonal mechanisms. Journal of Immunology. 1995;154:5511–5527. [PubMed: 7730652]
  76. Dhabhar FS, Miller AH, McEwen BS, Spencer RL. Stress-induced changes in blood leukocyte distribution: Role of adrenal steroid hormones. Journal of Immunology. 1996;157:1638–1644. [PubMed: 8759750]
  77. Dietz WH. Childhood weight affects adult morbidity and mortality. Journal of Nutrition. 1998;128(2 Suppl):411S–414S. [PubMed: 9478038]
  78. Dobbs CM, Vasquez M, Glaser R, Sheridan JF. Mechanisms of stress-induced modulation of viral pathogenesis and immunity. Journal of Neuroimmunology. 1993;48:151–160. [PubMed: 8227313]
  79. Dressler WW. Social support, lifestyle incongruity, and arterial blood pressure in a southern black community. Psychosomatic Medicine. 1991;53:608–620. [PubMed: 1758946]
  80. Drevets WC, Price JL, Simpson JR Jr, Todd RD, Reich T, Vannier M, Ralchle ME. Subgenual prefrontal cortex abnormalities in mood disorders. Nature. 1997;386:824–827. [PubMed: 9126739]
  81. Eichenbaum H. How does the brain organize memories? Science. 1997;277:330–332. [PubMed: 9518364]
  82. Epel N, McEwen B, Ickovics J. Embodying psychological thriving: Physical thriving in response to stress. Journal of Social Issues. 1998;54:301–322.
  83. Epping-Jordan JE, Compas BE, Osowiecki DM, Oppedisano G, Gerhardt C, Primo K, Krag DN. Psychological adjustment to breast cancer: Processes of emotional distress. Health Psychology. 1999;18:1–12. [PubMed: 10431932]
  84. Eriksson JG, Forsén T, Tuomilehto J, Winter PD, Osmond C, Barker DJP. Catch-up growth in childhood and death from coronary heart disease: Longitudinal study. British Medical Journal. 1999;318:427–431. [PMC free article: PMC27731] [PubMed: 9974455]
  85. Esterling BA, L'Abate L, Murray EJ, Pennebaker JW. Empirical foundations for writing in prevention and psychotherapy: Mental and physical health outcomes. Clinical Psychology Review. 1999;19:79–96. [PubMed: 9987585]
  86. Everson SA, Goldberg DE, Kaplan GA, Cohen RD, Pukkala E, Tuomilehto J, Salonen JT. Hopelessness and risk of mortality and incidence of myocardial infarction and cancer. Psychosomatic Medicine. 1996;58:113–121. [PubMed: 8849626]
  87. Everson SA, Goldberg DE, Kaplan GA, Julkunen J, Salonen JT. Anger expression and incident hypertension. Psychosomatic Medicine. 1998;60:730–735. [PubMed: 9847033]
  88. Everson SA, Lynch JW, Chesney MA, Kaplan GA, Goldberg DE, Shade SB, Cohen RD, Salonen R, Salonen JT. Interaction of workplace demands and cardiovascular reactivity in progression of carotid atherosclerosis: Population based study. British Medical Journal. 1997;314:553–558. [PMC free article: PMC2126071] [PubMed: 9055713]
  89. Fauci AS. Mechanisms of corticosteroid action on lymphocyte subpopulations. I. Redistribution of circulating T and B lymphocytes to the bone marrow. Immunology. 1975;28:669–680. [PMC free article: PMC1445841] [PubMed: 1080130]
  90. Felitti VJ, Anda RF, Nordenberg D, Williamson DF, Spitz AM, Edwards V, Koss MP, Marks JS. Relationship of childhood abuse and household dysfunction to many of the leading causes of death in adults. The adverse childhood experiences (ACE) study. American Journal of Preventive Medicine. 1998;14:245–258. [PubMed: 9635069]
  91. Felten DL, Felten S, Bellinger DL, Carlson SL, Ackerman KD, Madden KS, Olschowki JA, Livnat S. Noradrenergic sympathetic neural interactions with the immune system: Structure and function. Immunological Reviews. 1987;100:225–260. [PubMed: 3326822]
  92. Fisher L, Ames EW, Chisholm K, Savoie L. Problems reported by parents of Romanian orphans adopted to British Columbia. International Journal Behavioral Development. 1997;20:67–82.
  93. Fox BH. Depressive symptoms and risk of cancer. Journal of the American Medical Association. 1989;262:1231. [PubMed: 2761063]
  94. Frasure-Smith N, Lesperance F, Talajic M. Depression and 18-month prognosis after myocardial infarction. Circulation. 1995;91:999–1005. [PubMed: 7531624]
  95. Frasure-Smith N, Lesperance F, Juneau M, Talajic M, Bourassa MG. Gender, depression, and one-year prognosis after myocardial infarction. Psychosomatic Medicine. 1999;61:26–37. [PubMed: 10024065]
  96. Freedman DS, Srinivasan SR, Valdez RA, Williamson DF, Berenson GS. Secular increases in relative weight and adiposity among children over two decades: The Bogalusa Heart Study. Pediatrics. 1997;99:420–426. [PubMed: 9041299]
  97. Fride E, Dan Y, Feldon J, Halevy G, Weinstock M. Effects of prenatal stress on vulnerability to stress in prepubertal and adult rats. Physiology and Behavior. 1986;37:681–687. [PubMed: 3774900]
  98. Friedman M, Rosenman RH. Type A Behavior and Your Heart. New York: Knopf; 1974.
  99. Fuchs EJ, Matzinger P. Is cancer dangerous to the immune system? Seminarsin Immunology. 1996;8:271–280. [PubMed: 8956455]
  100. Galea LAM, McEwen BS, Tanapat P, Deak T, Spencer RL, Dhabhar FS. Sex differences in dendritic atrophy of CA3 pyramidal neurons in response to chronic restraint stress. Neuroscience. 1997;81:689–697. [PubMed: 9316021]
  101. Gerin W, Pickering TG. Association between delayed recovery of blood pressure after acute mental stress and parental history of hypertension. Journal of Hypertension. 1995;13:603–610. [PubMed: 7594416]
  102. Glaser R, Kiecolt-Glaser JK, Bonneau RH, Malarkey W, Kennedy S, Hughes J. Stress-induced modulation of the immune response to recombinant hepatitis B vaccine. Psychosomatic Medicine. 1992;54:22–29. [PubMed: 1553399]
  103. Glaser R, Kiecolt-Glaser JK, Malarkey W, Sheridan JF. The influence of psychological stress on the immune response to vaccines. Annals of the New York Academy of Sciences. 1998;840:649–655. [PubMed: 9629291]
  104. Glaser R, Rabin B, Chesney M, Cohen S, Natelson B. Stress-induced immunomodulation: Implications for infectious diseases? Journal of the American Medical Association. 1999;281:2268–2270. [PubMed: 10386538]
  105. Goy RW. Early hormonal influences on the development of sexual and sex-related behavior. In: Schmitt FO, editor. The Neurosciences: Second Study Program. New York: Rockefeller University Press; 1970. pp. 196–206.
  106. Gray JA. Precis of the neuropsychology of anxiety: An enquiry into the functions of the septo-hippocamal system. The Behavioral and Brain Sciences. 1982;5:469–534.
  107. Green MA, Berlin MA. Five psychosocial variables related to the existence of post-traumatic stress disorder symptoms. Journal of Clinical Psychology. 1987;43:643–649. [PubMed: 3693553]
  108. Greeno CG, Wing RR. Stress-induced eating. Psychological Bulletin. 1994;115:444–464. [PubMed: 8016287]
  109. Greenough WT, Bailey CH. The anatomy of a memory: Convergence of results across a diversity of tests. Trends in Neurosciences. 1988;11:142–147.
  110. Greer S, Morris T, Pettingale KW. Psychological response to breast cancer: Effect on outcome. Lancet. 1979;2(8146):785–787. [PubMed: 90871]
  111. Griffin AC, Whitacre CC. Sex and strain differences in the circadian rhythm fluctuation of endocrine and immune function in the rat: Implications for rodent models of autoimmune disease. Journal of Neuroimmunology. 1991;35:53–64. [PubMed: 1955572]
  112. Gunnar MR. Minneapolis: Institute of Child Development. University of Minnesota; 1998. Neuroendocrine activity in children adopted from Romania: Relations with cognitive and social functioning.
  113. Guzowski JF, McGaugh JL. Antisense oligodeoxynucleotide-mediated disruption of hippocampal cAMP response element binding protein levels impairs consolidation of memory for water maze training. Proceedings of the National Academy of Sciences. 1997;94:2693–2698. [PMC free article: PMC20151] [PubMed: 9122258]
  114. Hamann SB, Ely TD, Grafton ST, Kilts CD. Amygdala activity related to enhanced memory for pleasant and aversive stimuli. Nature Neuroscience. 1999;3:289–293. [PubMed: 10195224]
  115. Hearn MD, Murray DM, Luepker RV. Hostility, coronary heart disease, and total mortality: A 33-year follow-up study of university students. Journal of Behavioral Medicine. 1989;12:105–121. [PubMed: 2788220]
  116. Hecker MH, Chesney MA, Black GW, Frautschi N. Coronary-prone behaviors in the Western Collaborative Group Study. Psychosomatic Medicine. 1988;50:153–164. [PubMed: 3375405]
  117. Heijnen CJ, Rouppe van der Voort C, Wulffraat N, van der Net J, Kuis W, Kavelaars A. Functional alpha-1 adrenergic receptors on leukocytes of patients with polyarticular juvenile rheumatoid arthritis. Journal of Neuroimmunology. 1996;71:223–226. [PubMed: 8982123]
  118. Heim C, Newport DJ, Heit S, Graham YP, Wilcox M, Bonsall R, Miller AH, Nemeroff CB. Pituitary-adrenal and autonomic responses to stress in women after sexual and physical abuse in childhood. Journal of the Americian Medical Association. 2000;284:592–597. [PubMed: 10918705]
  119. Helmers KF, Krantz DS, Merz CN, Klein J, Kop WJ, Gottdiener JS, Rozanski A. Defensive hostility: Relationship to multiple markers of cardiac ischemia in patients with coronary disease. Health Psychology. 1995;14:202–209. [PubMed: 7641660]
  120. Henry JP, Stephens PM, Ely DL. Psychosocial hypertension and the defence and defeat reactions. Journal of Hypertension. 1986;4:687–697. [PubMed: 3546493]
  121. Herman DB, Susser ES, Struening EL, Link BL. Adverse childhood experiences: Are they risk factors for adult homelessness? American Journal of Public Health. 1997;87:249–255. [PMC free article: PMC1380802] [PubMed: 9103105]
  122. Hermann G, Beck FM, Sheridan JF. Stress-induced glucocorticoid response modulates mononuclear cell trafficking during an experimental influenza viral infection. Journal of Neuroimmunology. 1995;56:179–186. [PubMed: 7860713]
  123. Hermann G, Tovar CA, Beck FM, Allen C, Sheridan JF. Restraint stress differentially affects the pathogenesis of an experimental influenza viral infection in three inbred strains of mice. Journal of Neuroimmunology. 1993;47:83–94. [PubMed: 8397217]
  124. Higley JD, Hasert MF, Suomi SJ, Linnoila M. Nonhuman primate model of alcohol abuse: Effects of early experience, personality, and stress on alcohol consumption. Proceedings of the National Academy of Science. 1991;88:7261–7265. [PMC free article: PMC52274] [PubMed: 1871131]
  125. Higley JD, Mehlman PT, Higley SB, Fernald B, Vickers J, Lindell SG, Taub DM, Suomi SJ, Linnoila M. Excessive mortality in young free-ranging male nonhuman primates with low cerebrospinal fluid 5-hydroxyindoleacetic acid concentrations. Archives of General Psychiatry. 1996a;53:537–543. [PubMed: 8639037]
  126. Higley JD, Mehlman PT, Poland RE, Taub DM, Vickers J, Suomi SJ, Linnoila M. CSF testosterone and 5-HIAA correlate with different types of aggressive behaviors. Biological Psychiatry. 1996b;40:1067–1082. [PubMed: 8931909]
  127. Holland JC, Rowland JH, editors. Handbook of Psychooncology: PsychologicalCare of the Patient with Cancer. New York: Oxford University Press; 1990.
  128. Hostetter MK, Iverson S, Thomas W, McKenzie D, Dole K, Johnson DE. Prospective medical evaluation of internationally adopted children. New England Journal of Medicine. 1991;325:479–485. [PubMed: 1649404]
  129. House JS, Landis KR, Umberson R. Social relationships and health. Science. 1988;241:540–545. [PubMed: 3399889]
  130. Houston BK, Kelly KE. Type A behavior in housewives: Relation to work, marital adjustment, stress, tension, health, fear-of-failure and self esteem. Journal of Psychosomatic Research. 1987;31:55–61. [PubMed: 3820146]
  131. Ironson G, Taylor CB, Boltwood M, Bartzokis T, Dennis C, Chesney M, Spitzer S, Segall GM. Effects of anger on left ventricular ejection fraction in coronary artery disease. American Journal of Cardiology. 1992;70:281–285. [PubMed: 1632389]
  132. Ironson G, Wynings C, Schneiderman N, Baum A, Rodriguez M, Greenwood D, Benight C, Antoni M, LaPerriere A, Huang HS, Klimas N, Fletcher MA. Posttraumatic stress symptoms, intrusive thoughts, loss, and immune function after Hurricane Andrew. Psychosomatic Medicine. 1997;59:128–141. [PubMed: 9088048]
  133. Irwin M, Daniels M, Smith TL, Bloom E, Weiner H. Impaired natural killer cell activity during bereavement. Brain, Behavior and Immunity. 1987;1:98–104. [PubMed: 3451784]
  134. Israel BA, Schurman SJ. Social support, control and the stress process. In: Glanz K, Lewis FM, Rimer BK, editors. Health, Behavior and Health Education: Theory, Research, and Practice. San Francisco: Jossey-Bass; 1990. pp. 179–205.
  135. Jamison RN, Burish TG, Wallston KA. Psychogenic factors in predicting survival of breast cancer patients. Journal of Clinical Oncology. 1987;5:768–772. [PubMed: 3572466]
  136. Johnson EO, Kamilaris TC, Calogero AE, Gold PW, Chrousos GP. Effects of early parenting on growth and development in a small primate. Pediatric Research. 1996a;39:999–1005. [PubMed: 8725261]
  137. Johnson EO, Kamilaris TC, Carter CS, Calogero AE, Gold PW, Chrousos GP. The biobehavioral consequences of psychogenic stress in a small, social primate (Callithrix jacchus jacchus). Biological Psychiatry. 1996b;40:317–337. [PubMed: 8874833]
  138. Julius S. Corcoran Lecture. Sympathetic hyperactivity and coronary risk in hypertension. Hypertension. 1993;21:886–893. [PubMed: 8505097]
  139. Kang DH, Coe CL, Karaszewski J, McCarthy DO. Relationship of social support to stress responses and immune function in healthy and asthmatic adolescents. Research in Nursing and Health. 1998;21:117–128. [PubMed: 9535404]
  140. Kaplan GA, Pamuk E, Lynch JW, Cohen RD, Balfour JL. Income inequality and mortality in the United States: Analysis of mortality and potential pathways. British Medical Journal. 1996;312:999–1003. [PMC free article: PMC2350835] [PubMed: 8616393]
  141. Karasek R, Theorell T. Healthy Work. New York: Basic Books; 1990.
  142. Kaufman J, Birmaher B, Perel J, Dahl RE, Stull S, Brent D, Trubnick L, al-Shabbout M, Ryan ND. Serotonergic functioning in depressed abused children: Clinical and familial correlates. Biological Psychiatry. 1998;44:973–981. [PubMed: 9821561]
  143. Kawachi I, Kennedy B, Lochner K, Prothrow-Stith D. Social capital, income inequality, and mortality. American Journal of Public Health. 1997;87:1491–1498. [PMC free article: PMC1380975] [PubMed: 9314802]
  144. Keltikangas-Jarvinen L, Raikkonen K, Hautanen A. Type A behavior and vital exhaustion as related to the metabolic hormonal variables of the hypothalamic-pituitary-adrenal axis. Behavioral Medicine. 1996a;22:15–22. [PubMed: 8805957]
  145. Keltikangas-Jarvinen L, Raikkonen K, Hautanen A, Adlercreutz H. Vital exhaustion, anger expression, and pituitary and adrenocortical hormones: Implications for the insulin resistance syndrome. Arteriosclerosis, Thrombosis, and Vascular Biology. 1996b;16:275–280. [PubMed: 8620343]
  146. Kiecolt-Glaser JK, Fisher LD, Ogrocki P, Stout JC, Speicher CE, Glaser R. Marital quality, marital disruption, and immune function. Psychosomatic Medicine. 1987;49:13–34. [PubMed: 3029796]
  147. Kiecolt-Glaser JK, Glaser R, Cacioppo JT. Marital conflict in older adults: endocrinological and immunological correlates. Psychosomatic Medicine. 1997;59:339–349. [PubMed: 9251151]
  148. Kiecolt-Glaser JK, Glaser R, Gravenstein S, Malarkey WB, Sheridan JF. Chronic stress alters the immune response to influenza virus vaccine in older adults. Proceedings of the National Academy of Sciences. 1996;93:3043–3047. [PMC free article: PMC39758] [PubMed: 8610165]
  149. Kiecolt-Glaser JK, Marucha PT, Malarkey WB, Mercado AM, Glaser R. Slowing of wound healing by psychological stress. Lancet. 1995;346:1194–1196. [PubMed: 7475659]
  150. King KB. Psychologic and social aspects of cardiovascular disease. Annals of Behavioral Medicine. 1997;19:264–270. [PubMed: 9603700]
  151. Kirschbaum C, Kudielka BM, Gaab J, Schommer NC, Hellhammer DH. Impact of gender, menstrual cycle phase and oral contraceptive use on the activity of the hypothalamo-pituitary-adrenal axis. Psychosomatic Medicine. 1999;61:154–162. [PubMed: 10204967]
  152. Kirschbaum C, Prussner JC, Stone AA, Federenko I, Gaab J, Lintz D, Schommer N, Hellhammer DH. Persistent high cortisol responses to repeated psychological stress in a subpopulation of healthy men. Psychosomatic Medicine. 1995;57:468–474. [PubMed: 8552738]
  153. Klesges RC, Eck LH, Hanson CL, Haddock CK, Klesges LM. Effectsof obesity, social interactions, and physical environment on physical activity in preschoolers. Health Psychology. 1990;9:435–449. [PubMed: 2373068]
  154. Kop WJ, Appels AP, Mendes DLC, Bar FW. The relationship between severity of coronary artery disease and vital exhaustion. Journal of Psychosomatic Research. 1996;40:397–405. [PubMed: 8736420]
  155. Kop WJ, Appels AP, Mendes dLC, de Swart HB, Bar FW. Vital exhaustion predicts new cardiac events after successful coronary angioplasty. Psychosomatic Medicine. 1994;56:281–287. [PubMed: 7972608]
  156. Kop WJ, Hamulyak K, Pernot C, Appels A. Relationship of blood coagulation and fibrinolysis to vital exhaustation. Psychosomatic Medicine. 1998;60:352–358. [PubMed: 9625224]
  157. Koskenvuo M, Kaprio J, Rose RJ, Kesaniemi A, Sarna S, Heikkila K, Langinvainio H. Hostility as a risk factor for mortality and ischemic heart disease in men. Psychosomatic Medicine. 1988;50:330–340. [PubMed: 3413267]
  158. Kraemer GW, Schmidt DE, Ebert MH. The behavioral neurobiology of self-injurious behavior in rhesus monkeys. In: Stoff DM, Mann JJ, editors. The Neurobiology of Suicide: From the Bench to the Clinic. Vol. 836. Annals of the New York Academyof Sciences; 1997. pp. 12–38. [PubMed: 9616792]
  159. Kubzansky LD, Kawachi I, Spiro A, Weiss ST, Vokonas PS, Sparrow D. Is worrying bad for your heart? A prospective study of worry and coronary heart disease in the Normative Aging Study. Circulation. 1997;95:818–824. [PubMed: 9054737]
  160. Kusnecov AV, Grota LJ, Schmidt SG, Bonneau RH, Sheridan JF, Glaser R, Moynihan JA. Decreased herpes simplex viral immunity and enhanced pathogenesis following stressor administration in mice. Journal of Neuroimmunology. 1992;38:129–137. [PubMed: 1315793]
  161. Lazarus RS, Folkman S. Stress, Appraisal, and Coping. New York: Springer-Verlag; 1984.
  162. LeDoux JE. The Emotional Brain. New York: Simon and Schuster; 1996.
  163. Leon GR, Finn SE, Murray D, Bailey JM. Inability to predict cardiovascular disease from hostility scores or MMPI items related to type A behavior. Journal of Consulting and Clinical Psychology. 1988;56:597–600. [PubMed: 3198819]
  164. Levine S, Click D, Nakane PK. Adrenal and plasma corticosterone and vitamin A in rat adrenal glands during postnatal development. Endocrinology. 1967;80:910–914. [PubMed: 4290303]
  165. Lipkus IM, Barefoot JC, Williams RB, Siegler IC. Personality measures as predictors of smoking initiation and cessation in the UNC Alumni Heart Study. Health Psychology. 1994;13:149–155. [PubMed: 8020458]
  166. Liu D, Diorio J, Tannenbaum B, Caldji C, Francis D, Freedman A, Sharma S, Pearson D, Plotsky PM, Meaney MJ. Maternal care, hippocampal glucocorticoid receptors, and hypothalamic-pituitary-adrenal responses to stress. Science. 1997;277:1659–1662. [PubMed: 9287218]
  167. Lupien S, Lecours AR, Lussier I, Schwartz G, Nair NPV, Meaney MJ. Basal cortisol levels and cognitive deficits in human aging. Journal of Neuroscience. 1994;14:2893–2903. [PMC free article: PMC6577490] [PubMed: 8182446]
  168. Lupien SJ, DeLeon MJ, De Santi S, Convit A, Tarshish C, Nair NPV, Thakur M, McEwen BS, Hauger RL, Meaney MJ. Cortisol levels during human aging predict hippocampal atrophy and memory deficits. Nature Neuroscience. 1998;1:69–73. [PubMed: 10195112]
  169. Lynch J, Krause N, Kaplan GA, Tuomilehto J, Salonen JT. Workplace conditions, socioeconomic status, and the risk of mortality and acute myocardial infarction: The kuopio ischemic heart disease risk factor study. American Journal of Public Health. 1997a;87:617–622. [PMC free article: PMC1380842] [PubMed: 9146441]
  170. Lynch JW, Kaplan GA, Shema SJ. Cumulative impact of sustained economic hardship on physical, cognitive, psychological, and social functioning. New England Journal of Medicine. 1997b;337:1889–1895. [PubMed: 9407157]
  171. Maier S, Goehler LE, Fleshner M, Watkins LR. The role of the vagus in cytokine to brain communication. Annals of the New York Academy of Sciences. 1998;840:289–300. [PubMed: 9629257]
  172. Mann JJ. The neurobiology of suicide. Nature Medicine. 1998;4:25–30. [PubMed: 9427602]
  173. Manuck SB, Kaplan JR, Adams MR, Clarkson TB. Studies of psychosocial influences on coronary artery atherosclerosis in cynomolgus monkeys. Health Psychology. 1995;7:113–124. [PubMed: 3371304]
  174. Marmot MG, Davey Smith G, Stansfeld S, Patel C, North F, Head J, White I, Brunner E, Feeney A. Health inequalities among British civil servants: The Whitehall II study. Lancet. 1991;337:1387–1393. [PubMed: 1674771]
  175. Martin KC, Kandel ER. Cell adhesion molecules, CREB, and the formation of new synaptic connections. Neuron. 1996;17:567–570. [PubMed: 8893013]
  176. Marucha PT, Kiecolt-Glaser JK, Favagehi M. Mucosal wound healing is impaired by examination stress. Psychosomatic Medicine. 1998;60:362–365. [PubMed: 9625226]
  177. Matthews KA, Glass DC, Rosenman RH, Bortner RW. Competitivedrive, pattern A, and coronary heart disease: A further analysis of some data from the Western Collaborative Group Study. Journal of Chronic Diseases. 1977;30:489–498. [PubMed: 893653]
  178. Matthews KA, Owens JF, Kuller LH, Sutton-Tyrrell K, Jansen-McWilliams L. Are hostility and anxiety associated with carotid atherosclerosis in healthy postmenopausal women? Psychosomatic. Medicine. 1998;60:633–638. [PubMed: 9773770]
  179. McCranie EW, Watkins LO, Brandsma JM, Sisson BD. Hostility, coronary heart disease (CHD) incidence, and total mortality: Lack of association in a 25-year follow-up study of 478 physicians. Journal of Behavioral Medicine. 1986;9:119–125. [PubMed: 3712425]
  180. McEwen B, Biron C, Brunson K, Bulloch K, Chambers WH, Dhabhar FS, Goldfarb RH, Kitson RP, Miller AH, Spencer RL, Weiss JM. The role of adrenocorticoids as modulators of immune function in health and disease: Neural, endocrine, and immune interactions. Brain Research Reviews. 1997;23:79–133. [PubMed: 9063588]
  181. McEwen BS. Possible mechanisms for atrophy of the human hippocampus. Molecular Psychiatry. 1997;2:255–262. [PubMed: 9152991]
  182. McEwen BS. Protective and damaging effects of stress mediators. New England Journal of Medicine. 1998;338:171–179. [PubMed: 9428819]
  183. McEwen BS. Permanence of brain sex differences and structural plasticity of the adult brain. Proceedings of the National Academy of Sciences. 1999a;96:7128–7130. [PMC free article: PMC33584] [PubMed: 10377379]
  184. McEwen BS. Stress and hippocampal plasticity. Annual Review of Neuroscience. 1999b;22:105–122. [PubMed: 10202533]
  185. McEwen BS, Alves SH. Estrogen actions in the central nervous system. Endocrine Reviews. 1999;20:279–307. [PubMed: 10368772]
  186. McEwen BS, Seeman T. Protective and damaging effects of mediators of stress: Elaborating and testing the concepts of allostasis and allostatic load. Annals of the New York Academy of Sciences. 1999;896:30–47. [PubMed: 10681886]
  187. McEwen BS, Stellar E. Stress and the individual: Mechanisms leading to disease. Archives of Internal Medicine. 1993;153:2093–2101. [PubMed: 8379800]
  188. McGaugh JL, Cahill L, Roozendaal B. Involvement of the amygdala in memory storage: Interaction with other brain systems. Proceedings of the National Academy of Sciences. 1996;93:13508–13514. [PMC free article: PMC33638] [PubMed: 8942964]
  189. Meaney M, Aitken D, Berkel H, Bhatnager S, Sapolsky R. Effect of neonatal handling on age-related impairments associated with the hippocampus. Science. 1988;239:766–768. [PubMed: 3340858]
  190. Michelson D, Stratakis C, Hill L, Reynolds J, Galliven E, Chrousos G, Gold P. Bone mineral density in women with depression. New England Journal of Medicine. 1996;335:1176–1181. [PubMed: 8815939]
  191. Miller TQ, Smith TW, Turner CW, Guijarro ML, Hallet AJ. A meta-analytic review of research on hostility and physical health. Psychological Bulletin. 1996;119:322–348. [PubMed: 8851276]
  192. Millstein SG, Petersen AC, Nightingale EO, editors. Promoting the Health of Adolescents: New Directions for the Twenty-first Century. New York: Oxford University Press; 1993.
  193. Muller JE, Tofler GH. A symposium: triggering and circadian variation of onset of acute cardiovascular disease. American Journal of Cardiology. 1990;66:1–70. [PubMed: 2239705]
  194. Must A, Strauss RS. Risks and consequences of childhood and adolescent obesity. International Journal of Obesity and Related Metabolic Disorders. 1999;23(suppl. 2):S2–S11. [PubMed: 10340798]
  195. National Research Council (NRC). From Neurons to Neighborhoods: The Science of Early Childhood Development. Washington, DC: National Academy Press; 2000. [PubMed: 25077268]
  196. Nelson C, Bloom F. Child development and neuroscience. Child Development. 1997;68:970–987. [PubMed: 29106726]
  197. Nelson CA, Carver LJ. The effects of stress and trauma on brain and memory: A view from developmental cognitive neuroscience. Development and Psychopathology. 1998;10:793–809. [PubMed: 9886227]
  198. Nielsen KA, Jensen RA. Beta-adrenergic receptor antagonist anti-hypertensive medications impair arousal-induced modulation of working memory in elderly humans. Behavioral and Neural Biology. 1995;62:190–200. [PubMed: 7857241]
  199. Notzon FC, Komarov YM, Ermakov SP, Sempos CT, Marks JS, Sempos EV. Causes of declining life expectancy in Russia. Journal of the American Medical Association. 1998;279:793–800. [PubMed: 9508159]
  200. Ockenfels MC, Porter L, Smyth J, Kirschbaum C, Hellhammer DH, Stone AA. Effect of chronic stress associated with unemployment on salivary cortisol: Overall cortisol levels, diurnal rhythm, and acute stress reactivity. Psychosomatic Medicine. 1995;57:460–467. [PubMed: 8552737]
  201. Oliff HS, Berchtold NC, Isackson P, Cotman CW. Exercise-induced regulation of brain-derived neurotrophic factor (BDNF) transcripts in the rat hippocampus. Brain Research Molecular Brain Research. 1998;61:147–153. [PubMed: 9795193]
  202. Padgett DA, Marucha PT, Sheridan JF. Restraint stress slow cutaneous wound healing in mice. Brain, Behavior and Immunity. 1998;12:64–73. [PubMed: 9570862]
  203. Pennebaker JW. Writing about emotional experiences as a therapeutic process. Psychological Science. 1997;8:162–166.
  204. Penninx BW, Guralnik JM, Pahor M, Ferrucci L, Cerhan JR, Wallace RB, Havlik RJ. Chronically depressed mood and cancer risk in older persons. Journal of the National Cancer Institute. 1998;90:1888–1893. [PubMed: 9862626]
  205. Perseghin G, Price TB, Petersen KF, Roden M, Cline GW, Gerow K, Rothman DL, Shulman GI. Increased glucose transport-phosphorylation and muscle glycogen synthesis after exercise training in insulin-resistant subjects. New England Journal of Medicine. 1996;335:1357–1362. [PubMed: 8857019]
  206. Petrie KJ, Booth RJ, Pennebaker JW. The immunological effects of thought suppression. Journal of Personality and Social Psychology. 1998;75:1264–1272. [PubMed: 9866186]
  207. Petrie KJ, Booth RJ, Pennebaker JW, Davison KP, Thomas MG. Disclosure of trauma and immune response to a hepatitis B vaccination program. Journal of Consulting and Clinical Psychology. 1995;63:787–792. [PubMed: 7593871]
  208. Pine DS, Coplan JD, Wasserman GA, Miller LS, Fried JE, Davies M, Cooper TB, Greenhill L, Shaffer D, Parsons B. Neuroendocrine response to fenfluramine challenge in boys. Associations with aggressive behavior and adverse rearing. Archives of General Psychiatry. 1997;54:839–846. [PubMed: 9294375]
  209. Poulter NR, Khaw KT, Sever PS. Higher blood pressures of urban migrants from an African low-blood pressure population are not due to selective migration. American Journal of Hypertension. 1988;1:143S–145S. [PubMed: 3415788]
  210. Poulter NR, Khaw KT, Hopwood BE, Mugambi M, Peart WS, Rose G, Sever PS. The Kenyan Luo migration study: Observations on the initiation of a rise in blood pressure. British Medical Journal. 1990;300:967–972. [PMC free article: PMC1662695] [PubMed: 2344502]
  211. Raikkonen K, Hautanen A, Keltikangas-Jarvinen L. Feelings of exhaustion, emotional distress, and pituitary and adrenocortical hormones in borderline hypertension. Journal of Hypertension. 1996;14:713–718. [PubMed: 8793693]
  212. Ramirez AJ, Craig TK, Watson JP, Fentiman IS, North WR, Rubens RD. Stress and relapse of breast cancer. Biomedical Journal. 1989;298:291–293. [PMC free article: PMC1835577] [PubMed: 2493899]
  213. Reinisch JM, Rosenblum LA, Sanders SA, editors. Masculinity/Femininity. New York: Oxford University Press; 1987.
  214. Review Panel. Coronary-prone behavior and coronary heart disease: A critical review. Circulation. 1981;63:1199–1215. [PubMed: 7014026]
  215. Rich-Edwards JW, Colditz GA, Stampfer MJ, Willett WC, Gillman MW, Hennekens CH, Speizer FE, Manson JE. Birthweight and the risk of type 2 diabetes mellitus in adult women. Annals of Internal Medicine. 1999;130:278–284. [PubMed: 10068385]
  216. Roozendaal B, Portillo-Marquez G, McGaugh JL. Basolateral amygdala lesions block glucocorticoid-induced modulation of memory for spatial learning. Behavioral Neuroscience. 1996;110:1074–1083. [PubMed: 8919010]
  217. Rowland J. Interpersonal resources: Coping. In: Holland JC, Holland JR, editors. Psychooncology: Psychological Care of the Patient with Cancer. New York: Oxford University Press; 1990. pp. 44–57.
  218. Rozanski A, Blumenthal JA, Kaplan J. Impact of psychological factors on the pathogenesis of cardiovascular disease and implications for therapy. Circulation. 1999;99:2192–2217. [PubMed: 10217662]
  219. Ryff CD, Singer B. The contours of positive health. Psychological Inquiry. 1998;9:1–28.
  220. Salbe AD, Fontvieille AM, Harper IT, Ravussin E. Low levels of physical activity in 5-year-old children. Journal of Pediatrics. 1997;131:423–429. [PubMed: 9329420]
  221. Salmon P, Calderbank S. The relationship of childhood physical and sexual abuse to adult illness behavior. Journal of Psychosomotic Research. 1996;40:329–336. [PubMed: 8861129]
  222. Sapolsky RM. Why stress is bad for your brain. Science. 1996;273:749–750. [PubMed: 8701325]
  223. Sasaki S, Yoneda Y, Fujita H, Uchida A, Takenaka K, Takesako T, Itoh H, Nakata T, Takeda K, Nakagawa M. Association of blood pressure variability with induction of atherosclerosis in cholesterol-fed rats. American Journal of Hypertension. 1994;7:453–459. [PubMed: 8060580]
  224. Scheier MF, Bridges MW. Person variables and health: Personality predispositions and acute psychological states as shared determinants for disease. Psychosomatic Medicine. 1995;57:255–268. [PubMed: 7652126]
  225. Scheier MF, Carver CS. Optimism, coping, and health: Assessment andimplications of generalized outcome expectancies. Health Psychology. 1985;4:219–247. [PubMed: 4029106]
  226. Scheier MF, Carver CS. Effects of optimism on psychological and physical well-being: Theoretical overview and empirical update. Cognitive Therapy and Research. 1992;16:201–228.
  227. Schnall PL, Schwartz JE, Landsbergis PA, Warren K, Pickering TG. Relation between job strain, alcohol, and ambulatory blood pressure. Hypertension. 1992;19:488–494. [PubMed: 1568768]
  228. Schulkin J, Gold PW, McEwen BS. Induction of corticotropin-releasing hormone gene expression by glucocorticoids: Implication for understanding the states of fear and anxiety and allostatic load. Psychoneuroendocrinology. 1998;23:219–243. [PubMed: 9695128]
  229. Seeman TE, McEwen BS. The impact of social environment characteristics on neuroendocrine regulation. Psychosomatic Medicine. 1996;58:459–471. [PubMed: 8902897]
  230. Seeman TE, McEwen BS, Singer BH, Albert MS, Rowe JW. Increase in urinary cortisol excretion and memory declines: MacArthur studies of successful aging. Journal of Clinical Endocrinology and Metabolism. 1997;82:2458–2465. [PubMed: 9253318]
  231. Selye H. The Stress of Life. New York: McGraw-Hill; 1956.
  232. Sesso HD, Kawachi I, Vokonas PS, Sparrow D. Depression and the risk of coronary heart disease in the Normative Aging Study. American Journal of Cardiology. 1998;82:851–856. [PubMed: 9781966]
  233. Shanks N, Harbuz MS, Jessop DS, Perks P, Moore PM, Lightman SL. Inflammatory disease as chronic stress. Annals of the New York Academy of Sciences. 1998;840:599–607. [PubMed: 9629287]
  234. Shekelle RB, Raynor WJ Jr, Ostfeld AM, Garron DC, Bieliauskas LA, Liu SC, Maliza C, Paul O. Psychological depression and 17-year risk of death from cancer. Psychosomatic Medicine. 1981;43:117–125. [PubMed: 7267935]
  235. Shekelle RB, Gale M, Norusis M. Type A score (Jenkins Activity Survey) and risk of recurrent coronary heart disease in the Aspirin Myocardial Infarction Study. American Journal of Cardiology. 1985a;56:221–225. [PubMed: 3895879]
  236. Shekelle RB, Gale M, Ostfeld AM, Paul O. Hostility, risk of coronary heart disease, and mortality. Psychosomatic Medicine. 1983;45:109–114. [PubMed: 6867229]
  237. Shekelle RB, Hulley SB, Neaton JD, Billings JH, Borhani NO, Gerace TA, Jacobs DR, Lasser NL, Mittlemark MB, Stamler J. for the Multiple Risk Factor Intervention Trial Research Group. The MRFIT behavior pattern study, Type A behavior and the incidence of coronary heart disease. American Journalof Epidemiology. 1985b;122:559–570. [PubMed: 4025299]
  238. Shekelle RB, Vernon SW, Ostfeld AM. Personality and coronary heart disease. Psychosomatic Medicine. 1991;53:176–184. [PubMed: 2031071]
  239. Sheline YI, Gado MH, Price JL. Amygdala core nuclei volumes are decreased in recurrent major depression. Neuro Report. 1998;9:2023–2028. [PubMed: 9674587]
  240. Sheline YI, Sanghavi M, Mintun MA, Gado MH. Depression duration but not age predicts hippocampal volume loss in medically healthy women with recurrent major depression. Journal of Neuroscience. 1999;19:5034–5043. [PMC free article: PMC6782668] [PubMed: 10366636]
  241. Sheridan JF, Feng NG, Bonneau RH, Allen CM, Huneycutt BS, Glaser R. Restraint stress differentially affects anti-viral cellular and humoral immune responses in mice. Journal of Neuroimmunology. 1991;31:245–255. [PubMed: 1847396]
  242. Shiffman S, Hickcox M, Paty JA, Gnys M, Kassel JD, Richards TJ. Progression from a smoking lapse to relapse: Prediction from abstinence violation effects, nicotine dependence, and lapse characteristics. Journal of Consulting andClinical Psychology. 1996;64:993–1002. [PubMed: 8916628]
  243. Shively CA, Clarkson TB. Social status and coronary artery atherosclerosis in female monkeys. Arteriosclerosis and Thrombosis. 1994;14:721–726. [PubMed: 8172850]
  244. Shively CA, Laber-Laird K, Anton RF. Behavior and physiology of social stress and depression in female cynomolgus monkeys. Biological Psychiatry. 1997;41:871–882. [PubMed: 9099414]
  245. Singer B, Ryff CD, Carr D, Magee WJ. Linking life histories and mental health: A person-centered strategy. Sociological Methodology. 1998;28:1–51.
  246. Sloan RP, Shapiro PA, Bagiella E, Myers M, German JM. Cardiac autonomic control buffers blood pressure variability responses to challenge: A psychophysiologic model of coronary artery disease. Psychosomatic. Medicine. 1999;61:58–68. [PubMed: 10024068]
  247. Smyth J, Ockenfels MC, Porter L, Kirschbaum C, Hellhammer DH, Stone AA. Stressors and mood measured on a momentary basis are associated with salivary cortisol secretion. Psychoneuroendocrinology. 1998;23:353–370. [PubMed: 9695136]
  248. Snyder CR, Harris C, Anderson JR, Holleran SA, Irving LM, Sigmon ST, Yoshinobu L, Gibb J, Langelle C, Harney P. The will and the ways: Development and validation of an individual-differences measure of hope. Journal of Personality and Social Psychology. 1991;60:570–585. [PubMed: 2037968]
  249. Solomon GF, Segerstrom SC, Grohr P, Kemeny M, Fahey J. Shaking up immunity: Psychological and immunologic changes after a natural disaster. Psychosomatic Medicine. 1997;59:114–127. [PubMed: 9088047]
  250. Staessen JA, Poulter NR, Fletcher AE, Markowe HL, Marmot MG, Shipley MJ, Bulpitt CJ. Psycho-emotional stress and salt intake may interact to raise blood pressure. Journal of Cardiovascular Risk. 1994;1:45–51. [PubMed: 7614417]
  251. Stanton AL, Snider PR. Coping with a breast cancer diagnosis: A prospective study. Health Psychology. 1993;12:16–23. [PubMed: 8462494]
  252. Sterling P, Eyer J. Allostasis: A new paradigm to explain arousal pathology. In: Fisher S, Reason J, editors. Handbook of Life Stress, Cognition and Health. New York: John Wiley and Sons; 1988. pp. 629–649.
  253. Sternberg EM. Neural-immune interactions in health and disease. Journal of Clinical Investigation. 1997;100:2641–2647. [PMC free article: PMC508465] [PubMed: 9389725]
  254. Sternberg EM, Chrousos GP, Wilder RI, Gold PW. The stress response and the regulation of inflammatory disease. Annals of Internal Medicine. 1992;117:854–866. [PubMed: 1416562]
  255. Sternberg EM, Hill JM, Chrousos GP. Inflammatory mediator-induced hypothalamic-pituitary-adrenal axis activation is defective in streptococcal cell wall arthritis susceptible Lewis rats. Proceedings of the National Academy of Sciences. 1996;86:2374–2378. [PMC free article: PMC286915] [PubMed: 2538840]
  256. Sternberg EM, Hill JM, Chrousos GP, Kamilaris T, Listwak SJ, Gold PW, Wilder RI. Inflammatory mediator-induced hypothalamic-pituitary-adrenal axis activation is defective in streptococcal cell wall arthritis susceptible Lewis rats. Proceedings of the National Academy of Sciences. 1989;86:2374–2378. [PMC free article: PMC286915] [PubMed: 2538840]
  257. Styron T, Janoff-Bulman R. Childhood attachment and abuse: Long-term effects on adult attachment, depression, and conflict resolution. Child Abuse and Neglect. 1997;21:1015–1023. [PubMed: 9330802]
  258. Syme S, Berkman L. Social class, susceptibility and sickness. American Journal of Epidemiology. 1976;104:1–8. [PubMed: 779462]
  259. Taylor SE, Repetti RL, Seeman T. Health Psychology: What is an unhealthy environment and how does it get under the skin? Annual Review of Psychology. 1997;48:411–447. [PubMed: 9046565]
  260. Troiano RP, Flegal KM, Kuczmarski RJ, Campbell SM, Johnson CL. Overweight prevalence and trends for children and adolescents. Archives of Pediatricsand Adolescent Medicine. 1995;149:1085–1091. [PubMed: 7550810]
  261. U.S. Department of Health and Human Services. Physical Activity and Health: AReport of the Surgeon General. Washington, DC: Centers for Disease Control and Prevention National Center for Chronic Disease Prevention and Health Promotion. The President's Council on Physical Fitness and Sports; 1996.
  262. Uno H, Ross T, Else J, Suleman M, Sapolsky R. Hippocampal damage associated with prolonged and fatal stress in primates. Journal of Neuroscience. 1989;9:1705–1711. [PMC free article: PMC6569823] [PubMed: 2723746]
  263. Utz PJ, Hottelet M, Schur PH, Anderson P. Proteins phosphorylated during stress-induced apoptosis are common targets for autoantibody production in patients with systemic lupus erythematous. Journal of Experimental Medicine. 1997;185:843–854. [PMC free article: PMC2196161] [PubMed: 9120390]
  264. van Cauter E, Polonsky KS, Scheen AJ. Roles of circadian rhythmicity and sleep in human glucose regulation. Endocrine Reviews. 1997;18:716–738. [PubMed: 9331550]
  265. van den Boom DC. The influence of temperament and mothering on attachment and exploration: An experimental manipulation of sensitive responsiveness among lower-class mothers with irritable infants. Child Development. 1994;65:1457–1477. [PubMed: 7982362]
  266. van Praag H, Kempermann G, Gage FH. Running increases cell proliferation and neurogenesis in the adult mouse dentate gyrus. Nature Neuroscience. 1999;2:266–270. [PubMed: 10195220]
  267. Wadhwa PD. Encyclopedia of Mental Health. San Diego: Academic Press; 1998. Prenatal stress and life-span development.
  268. Wakschlak A, Weinstock M. Neonatal handling reverses behavioral abnormalities induced in rats by prenatal stress. Physiology and Behavior. 1990;48:289–92. [PubMed: 2255733]
  269. Watson M, Pruyn J, Greer S, van den Borne B. Locus of control and adjustment to cancer. Psychology Representative. 1990;66:39–48. [PubMed: 2326428]
  270. Wilder RL. Neuroendocrine-immune system interactions and autoimmunity. Annual Review of Immunology. 1995;13:307–338. [PubMed: 7612226]
  271. Wilkinson RG. Income distribution and life expectancy. British Medical Journal. 1992;304:165–168. [PMC free article: PMC1881178] [PubMed: 1637372]
  272. Williams CL, Jensen RA. Effects of vagotomy on Leu-enkephalin-induced changes in memory storage processes. Physiology and Behavior. 1993;54:659–663. [PubMed: 8248342]
  273. Williams CL, Jensen RA. Vagal afferents: A possible mechanism for the modulation of memory by peripherally acting agents. In: Frederickson RCA, McGaugh JL, Felten DL, editors. Peripheral Signaling of the Brain: Role in Neural-Immune Interactions, Learning and Memory. Lewiston, NY: Hogrefe and Huber; 1991. pp. 467–472.
  274. Zonderman AB, Costa PT Jr, McCrae RR. Depression as a risk for cancer morbidity and mortality in a nationally representative sample. Journal of the American Medical Association. 1989;262:1191–1195. [PubMed: 2761060]
Copyright © 2001, National Academy of Sciences.
Bookshelf ID: NBK43737

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this title (5.9M)

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...
statistics