04.01.2015 Views

ABSTRACT OF DISSERTATION Deanna Lynn McCullers The ...

ABSTRACT OF DISSERTATION Deanna Lynn McCullers The ...

ABSTRACT OF DISSERTATION Deanna Lynn McCullers The ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>ABSTRACT</strong> <strong>OF</strong> <strong>DISSERTATION</strong><br />

<strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong><br />

<strong>The</strong> Graduate School<br />

University of Kentucky<br />

2001


ADRENOCORTICOSTEROID RECEPTOR EFFECTS<br />

ON HIPPOCAMPAL NEURON VIABILITY<br />

____________________________________<br />

<strong>ABSTRACT</strong> <strong>OF</strong> <strong>DISSERTATION</strong><br />

____________________________________<br />

A dissertation submitted in partial fulfillment of the<br />

requirements for the degree of Doctor of Philosophy in the<br />

College of Medicine at the University of Kentucky<br />

By<br />

<strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong><br />

Lexington, Kentucky<br />

Director: Dr. James P. Herman, Associate Professor of Anatomy and Neurobiology<br />

Lexington, Kentucky<br />

2001<br />

Copyright © <strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong> 2001


<strong>ABSTRACT</strong> <strong>OF</strong> <strong>DISSERTATION</strong><br />

ADRENOCORTICOSTEROID RECEPTOR EFFECTS<br />

ON HIPPOCAMPAL NEURON VIABILITY<br />

Glucocorticoid activation of two types of adrenocorticosteroid receptors (ACRs), the<br />

mineralocorticoid receptor (MR) and the glucocorticoid receptor (GR), influences hippocampal<br />

neuron vulnerability to injury. Excessive activation of GR may compromise hippocampal<br />

neuron survival after several types of challenge including ischemic, metabolic, and excitotoxic<br />

insults. In contrast, MR prevents adrenalectomy-induced loss of granule neurons in the dentate<br />

gyrus. <strong>The</strong> present thesis addresses the respective roles of MR and GR in modulating neuronal<br />

survival following two forms of neuronal injury, excitotoxicity and traumatic brain injury. Male<br />

Sprague-Dawley rats were pretreated with MR antagonist spironolactone or GR antagonist<br />

mifepristone (RU486) and subsequently injected with kainic acid, an excitotoxic glutamate<br />

analog, or injured with a controlled cortical impact. Twenty-four hours following injury,<br />

hippocampal neuron survival was measured to test the hypotheses that MR blockade would<br />

endanger and GR blockade would protect hippocampal neurons following injury. Messenger<br />

RNA levels of viability-related genes including bcl-2, bax, p53, BDNF, and NT-3 were also<br />

measured to test the hypothesis that ACR regulation of these genes would


correlate with neuronal survival. In addition, ACR mRNA levels were measured following<br />

receptor blockade and injury to test the hypothesis that glucocorticoid signaling is altered<br />

following neuronal injury via regulation of ACR expression.<br />

Mineralocorticoid receptor blockade with spironolactone increased neuronal vulnerability to<br />

excitotoxic insult in hippocampal field CA3, and GR blockade with RU486 prevented neuronal<br />

loss after traumatic brain injury in field CA1. <strong>The</strong>se results are consistent with the hypotheses<br />

that MR protects and GR endangers hippocampal neurons. Adrenocorticosteroid receptor<br />

blockade decreased mRNA levels of the anti-apoptotic gene bcl-2 in select regions of uninjured<br />

hippocampus, yet ACR regulation of bcl-2 did not consistently correspond with measures of<br />

neuronal survival after injury. Kainic acid decreased MR mRNA levels in CA1 and CA3, while<br />

both kainic acid and controlled cortical impact dramatically decreased GR mRNA levels in<br />

dentate gyrus. <strong>The</strong>se data suggest that injury modulation of glucocorticoid signaling through<br />

regulation of ACR expression may influence hippocampal neuron viability following injury.<br />

KEYWORDS: Mineralocorticoid Receptor, Glucocorticoid Receptor, Messenger RNA, Kainic<br />

Acid, Traumatic Brain Injury<br />

<strong>Deanna</strong> L. <strong>McCullers</strong><br />

December 6, 2001


ADRENOCORTICOSTEROID RECEPTOR EFFECTS<br />

ON HIPPOCAMPAL NEURON VIABILITY<br />

By<br />

<strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong><br />

Dr. James P. Herman<br />

Director of Dissertation<br />

Dr. Brian MacPherson<br />

Director of Graduate Studies<br />

December 6, 2001


RULES FOR THE USE <strong>OF</strong> <strong>DISSERTATION</strong>S<br />

Unpublished dissertations submitted for the Doctor’s degree and deposited in the University of<br />

Kentucky Library are as a rule open for inspection, but are to be used only with due regard to the<br />

rights of the authors. Bibliographical references may be noted, but quotations or summaries of<br />

parts may be published only with the permission of the author, and with the usual scholarly<br />

acknowledgments.<br />

Extensive copying or publication of the dissertation in whole or in part also requires the consent<br />

of the Dean of the Graduate School of the University of Kentucky.


<strong>DISSERTATION</strong><br />

<strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong><br />

<strong>The</strong> Graduate School<br />

University of Kentucky<br />

2001


ADRENOCORTICOSTEROID RECEPTOR EFFECTS<br />

ON HIPPOCAMPAL NEURON VIABILITY<br />

____________________________________<br />

<strong>DISSERTATION</strong><br />

____________________________________<br />

A dissertation submitted in partial fulfillment of the<br />

requirements for the degree of Doctor of Philosophy in the<br />

College of Medicine at the University of Kentucky<br />

By<br />

<strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong><br />

Lexington, Kentucky<br />

Director: Dr. James P. Herman, Associate Professor of Anatomy and Neurobiology<br />

Lexington, Kentucky<br />

2001<br />

Copyright © <strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong> 2001


ACKNOWLEDGMENTS<br />

I would like to acknowledge my advisor and Dissertation Chair, Dr. James Herman, for his<br />

instrumental role in my training and in the completion of this work. For many years Dr. Herman<br />

has been an excellent role model as a scientist, a teacher, and a person.<br />

I also thank the<br />

members of my committee, Dr. Brian Davis, Dr. Steve Estus, Dr. Kurt Hauser, Dr. Jim Geddes,<br />

and Dr. Joe Springer, and the outside examiner, Dr. Thomas Foster, for their comments, insight,<br />

and commitment to my academic development. <strong>The</strong> former and current Directors of Graduate<br />

Studies, Dr. Harold Traurig and Dr. Brian MacPherson, have also been of great assistance<br />

throughout the dissertation process. I would further like to acknowledge Dr. Stephen Scheff for<br />

providing assistance and resources necessary for performing the studies on traumatic brain<br />

injury.<br />

Expert technical assistance was provided by several members of the Herman lab including:<br />

Garrett Bowers, Mark Dolgas, Dr. Helmer Figueiredo, Erin Murphy, Megan Paskitti, Dr. Kim<br />

Sipe, and Dr. Dana Ziegler. Dr. Patrick Sullivan, Katie Kraft, and other members of the Scheff<br />

lab provided technical assistance essential for the completion of the studies on traumatic brain<br />

injury, and Cynthia Long assisted with several histological procedures. I would also like to<br />

express my appreciation to Kelly Bolte Svec, Susan Herman, Liz Jones, Betty Newsom, Kim<br />

Stroth, and Kim Wilkerson for their valuable administrative assistance. Finally, I am indebted to<br />

my parents, Lynda J. <strong>McCullers</strong> and L. Arnold <strong>McCullers</strong>, for their continuing support.<br />

<strong>The</strong> contents of Chapter Two have previously been published as an article in the journal<br />

Neuroreport titled, “Mineralocorticoid receptors regulate bcl-2 and p53 mRNA expression in<br />

hippocampus”, by <strong>Deanna</strong> L. <strong>McCullers</strong> and James P. Herman, Copyright © 1998 Lippincott<br />

iii


Williams & Wilkins. This work is reprinted with permission of Lippincott Williams & Wilkins.<br />

<strong>The</strong> contents of Chapter Three have been previously published as an article in the Journal of<br />

Neuroscience Research titled, “Adrenocorticosteriod receptor blockade and excitotoxic<br />

challenge regulate adrenocorticosteroid receptor mRNA levels in hippocampus”, by <strong>Deanna</strong> L.<br />

<strong>McCullers</strong> and James P. Herman, Copyright © 2001 Wiley-Liss Inc. This work is reprinted with<br />

permission of Wiley-Liss, Inc., a subsidiary of John Wiley & Sons, Inc. <strong>The</strong> contents of Chapter<br />

Four are currently in press to be published as an article in the journal Neuroscience titled,<br />

“Mifepristone protects CA1 hippocampal neurons following traumatic brain injury in rat”, by<br />

D.L. <strong>McCullers</strong>, P.G. Sullivan, S.W. Scheff, and J.P. Herman, Copyright © 2001 IBRO. Work<br />

contained in the following dissertation was supported by AG12962 (J.P.H.), AG10836 (J.P.H),<br />

AG00242 (D.L.M), NS39828 (S.W.S.), and KSCHIRT (S.W.S.).<br />

iv


TABLE <strong>OF</strong> CONTENTS<br />

Acknowledgments.............................................................................................................. iii<br />

List of Tables .................................................................................................................... vii<br />

List of Figures.................................................................................................................. viii<br />

List of Files ..........................................................................................................................x<br />

Chapter One: Introduction ..................................................................................................1<br />

Adrenocorticosteroid receptors in brain.........................................................................1<br />

<strong>The</strong> hypothalamic-pituitary-adrenal axis .................................................................3<br />

Properties of adrenocorticosteroid receptors ...........................................................5<br />

Central nervous system distribution of adrenocorticosteroid receptors...................6<br />

Regulation of adrenocorticosteroid receptor expression..........................................7<br />

Adrenocorticosteroid receptor activation ................................................................8<br />

Adrenocorticosteroid receptor regulation of gene transcription ..............................9<br />

Adrenocorticosteriod receptor roles in neuronal viability ...........................................12<br />

Excessive glucocorticoid receptor activation compromises hippocampal<br />

neuron viability ...................................................................................................12<br />

Mechanisms for glucocorticoid receptor endangerment of hippocampal<br />

neurons................................................................................................................14<br />

Neurotrophic effects of mineralocorticoid receptor activation..............................16<br />

Kainic acid as a model of excitotoxicity......................................................................17<br />

Glutamate receptors ...............................................................................................17<br />

Kainic acid excitotoxicity ......................................................................................18<br />

Glucocorticoid modulation of kainic acid excitotoxicity.......................................21<br />

Regulators of cell survival and death.....................................................................22<br />

Controlled cortical impact as a model of traumatic brain injury .................................25<br />

Controlled cortical impact pathology.....................................................................26<br />

Glucocorticoid modulation of traumatic brain injury ............................................27<br />

Regulators of cell survival and death.....................................................................27<br />

Chapter Two: Mineralocorticoid receptors regulate bcl-2 and p53 mRNA expression<br />

in hippocampus ...............................................................................................................29<br />

Introduction..................................................................................................................29<br />

Materials and Methods.................................................................................................30<br />

Results..........................................................................................................................31<br />

Discussion....................................................................................................................38<br />

Chapter Three: Adrenocorticosteriod receptor blockade and excitotoxic challenge<br />

regulate adrenocorticosteroid receptor mRNA levels in hippocampus ..........................41<br />

Introduction..................................................................................................................41<br />

Materials and Methods.................................................................................................42<br />

v


Results..........................................................................................................................45<br />

Discussion....................................................................................................................50<br />

Chapter Four: Mifepristone protects CA1 hippocampal neurons following traumatic<br />

brain injury in rat ............................................................................................................55<br />

Introduction..................................................................................................................55<br />

Materials and Methods.................................................................................................56<br />

Results..........................................................................................................................61<br />

Discussion....................................................................................................................68<br />

Chapter Five: Traumatic brain injury regulates adrenocorticosteroid receptor mRNA<br />

levels in rat hippocampus................................................................................................75<br />

Introduction..................................................................................................................75<br />

Materials and Methods.................................................................................................76<br />

Results..........................................................................................................................79<br />

Discussion....................................................................................................................86<br />

Chapter Six: Discussion.....................................................................................................91<br />

Adrenocorticosteriod receptor regulation of hippocampal neuron viability................91<br />

Adrenocorticosteriod receptor regulation of hippocampal neuron survival ..........91<br />

Adrenocorticosteriod receptor regulation of viability-related gene expression.....92<br />

Adrenocorticosteriod receptor regulation of hippocampal neuron survival –<br />

potential mechanisms..........................................................................................93<br />

Adrenocorticosteriod receptor modulation of calcium homeostasis................93<br />

Adrenocorticosteriod receptor suppression of inflammation...........................94<br />

Adrenocorticosteriod receptor regulation of hippocampal neurotrophic<br />

factor expression ...........................................................................................96<br />

Summary..........................................................................................................98<br />

Injury regulation of adrenocorticosteroid receptor mRNA levels .............................100<br />

Injury regulation of adrenocorticosteroid receptor mRNA levels – potential<br />

mechanisms.......................................................................................................100<br />

Glutamate regulation of adrenocorticosteriod receptor mRNA levels ..........100<br />

Calcium regulation of adrenocorticosteriod receptor mRNA levels..............101<br />

Injury regulation of adrenocorticosteroid receptor mRNA levels – potential<br />

influence on hippocampal neuron viability.......................................................101<br />

Potential effects of GR downregulation on hippocampal circuitry ...............101<br />

Potential effects of GR downregulation on neurogenesis..............................102<br />

Summary....................................................................................................................104<br />

Appendix..........................................................................................................................105<br />

References........................................................................................................................106<br />

Vita...................................................................................................................................139<br />

vi


LIST <strong>OF</strong> TABLES<br />

Table 3.1. Effects of adrenocorticosteroid receptor blockade and kainic acid on<br />

plasma corticosterone levels .........................................................................45<br />

Table 3.2. Effects of adenocorticosteroid receptor blockade and kainic acid on<br />

hippocampal neuron viability .......................................................................49<br />

Table 4.1. Mean coefficient of error .................................................................................59<br />

Table 4.2. Mean coefficient of variation...........................................................................59<br />

Table 5.1. Effects of adrenocorticosteroid receptor blockade and controlled cortical<br />

impact on plasma corticosterone levels ........................................................79<br />

vii


LIST <strong>OF</strong> FIGURES<br />

Figure 1.1. <strong>The</strong> hypothalamic-pituitary-adrenal axis..........................................................4<br />

Figure 1.2. Functional domains of the rat glucocorticoid receptor and<br />

mineralocorticoid receptor ..........................................................................5<br />

Figure 1.3. Microdistribution of mineralocorticoid receptor and glucocorticoid<br />

receptor mRNAs in rat hippocampus...........................................................7<br />

Figure 1.4. Models for transcriptional regulation by adrenocorticosteriod receptors.......11<br />

Figure 1.5. Schematic diagram of the intrinsic circuitry of the hippocampal formation..19<br />

Figure 1.6. Photomicrograph illustrating the effects of saline versus kainic acid on<br />

CA1 hippocampal pyramidal neurons .......................................................20<br />

Figure 2.1. Photomicrographs illustrating bcl-2, bax, and p53 mRNA distribution in<br />

hippocampus of saline- and kainic acid-treated rats..................................32<br />

Figure 2.2. Effects of mineralocorticoid receptor blockade and kainic acid on bcl-2<br />

mRNA expression in hippocampus ...........................................................33<br />

Figure 2.3. Effects of mineralocorticoid receptor blockade and kainic acid on bax<br />

mRNA expression in hippocampus ...........................................................34<br />

Figure 2.4. Effects of mineralocorticoid receptor blockade and kainic acid on p53<br />

mRNA expression in hippocampus ...........................................................35<br />

Figure 2.5. Effects of mineralocorticoid receptor blockade and kainic acid on<br />

brain-derived neurotrophic factor expression in hippocampus..................36<br />

Figure 2.6. Effects of mineralocorticoid receptor blockade and kainic acid on<br />

neurotrophin-3 mRNA expression in hippocampus ..................................37<br />

Figure 3.1. Photomicrographs illustrating the effects of saline and kainic acid on<br />

mineralocorticoid receptor and glucocorticoid receptor mRNA levels<br />

in hippocampus of vehicle-pretreated rats .................................................46<br />

Figure 3.2. Effects of adrenocorticosteroid receptor blockade and kainic acid<br />

treatments on mineralocorticoid receptor mRNA levels in rat<br />

hippocampus ..............................................................................................47<br />

Figure 3.3. Effects of adrenocorticosteroid receptor blockade and kainic acid<br />

treatments on glucocorticoid receptor mRNA levels in rat<br />

hippocampus ..............................................................................................48<br />

Figure 4.1. Effects of mineralocorticoid receptor blockade, glucocorticoid receptor<br />

blockade, and controlled cortical impact injury on hippocampal neuron<br />

viability as measured by the optical fractionator method ..........................62<br />

Figure 4.2. Photomicrographs illustrating bcl-2, bax, and p53 mRNA levels in<br />

hippocampus of vehicle-pretreated, injured rats at twenty-four hours ......63<br />

Figure 4.3. Effects of mineralocorticoid receptor blockade, glucocorticoid receptor<br />

blockade, and controlled cortical impact injury on bcl-2 mRNA levels<br />

in hippocampus .........................................................................................65<br />

Figure 4.4. Effects of mineralocorticoid receptor blockade, glucocorticoid receptor<br />

blockade, and controlled cortical impact injury on bax mRNA levels<br />

in hippocampus ..........................................................................................66<br />

Figure 4.5. Effects of mineralocorticoid receptor blockade, glucocorticoid receptor<br />

blockade, and controlled cortical impact injury on p53 mRNA levels<br />

in hippocampus ..........................................................................................67<br />

viii


Figure 5.1. Twenty-four hour time course analysis of plasma adrenocorticotropin<br />

releasing hormone levels following controlled cortical impact.................81<br />

Figure 5.2. Twenty-four hour time course analysis of plasma corticosterone levels<br />

following controlled cortical impact..........................................................81<br />

Figure 5.3. Photomicrographs illustrating the effects of sham operation and<br />

controlled cortical impact on mineralocorticoid receptor and<br />

glucocorticoid receptor mRNA levels in hippocampus of vehicle<br />

pretreated rats.............................................................................................82<br />

Figure 5.4. Effects of mineralocorticoid receptor blockade, glucocorticoid receptor<br />

blockade, and controlled cortical impact injury on mineralocorticoid<br />

receptor mRNA levels in hippocampus .....................................................84<br />

Figure 5.5. Effects of mineralocorticoid receptor blockade, glucocorticoid receptor<br />

blockade, and controlled cortical impact injury on glucocorticoid<br />

receptor mRNA levels in hippocampus .....................................................85<br />

Figure 6. Potential mechanisms for glucocorticoid modulation of hippocampal<br />

neuron survival after injury........................................................................99<br />

ix


LIST <strong>OF</strong> FILES<br />

mcculler.pdf ............................................................................................................ 2.49 MB<br />

x


Chapter One: Introduction<br />

Adrenocorticosteroid receptors in brain<br />

Glucocorticoids are steroid hormones that play an essential role in adaptation to stress.<br />

Named for their profound effects on glucose metabolism, glucocorticoids stimulate<br />

gluconeogenesis, mobilize stored energy, and inhibit subsequent energy storage during the stress<br />

response. In addition to regulating metabolism, glucocorticoids also increase cardiovascular<br />

tone, inhibit growth and reproductive capabilities, restrain immune and inflammatory responses,<br />

and influence learning and memory (Munck et al., 1984; Sapolsky et al., 2000). Glucocorticoids<br />

are essential for surviving stress, but excessive exposure can have detrimental effects such as<br />

hypertension, steroid diabetes, immunosuppression, inhibition of growth, and infertility (Munck<br />

et al., 1984; Sapolsky et al., 1986; McEwen and Stellar, 1993). Elevated levels of the steroid<br />

may also contribute to the development of affective disorders such as depression (De Kloet et al.,<br />

1998) and neuropathological processes including Alzheimer’s disease (Landfield et al., 1992)<br />

and age-related neurodegeneration (Sapolsky et al., 1986; Landfield and Eldridge, 1991;<br />

McEwen, 1992). Furthermore, many studies have demonstrated that excessive exposure to<br />

glucocorticoids increases neuronal vulnerability to many types of challenge including ischemia,<br />

excitotoxicity, and oxidative stress (see Sapolsky, 1994).<br />

At the cellular level, glucocorticoid effects are mediated by two types of ligand-activated<br />

receptors, the mineralocorticoid receptor (MR) and the glucocorticoid receptor (GR). <strong>The</strong>se<br />

adrenocorticosteroid receptors (ACRs) are differentially activated in different glucocorticoid<br />

contexts; MR is extensively activated at basal hormone levels, while GR is activated by high<br />

glucocorticoid concentrations such as those experienced during stress or circadian peak (Reul<br />

and de Kloet, 1985; Reul et al., 1987). This arrangement suggests that the MR mediates tonic<br />

influences of glucocorticoids, while GR drives glucocorticoid actions aimed at restoring<br />

homeostasis (De Kloet and Reul, 1987). <strong>The</strong> dichotomous activation of the receptors also<br />

implies that MR maintains normal neuronal function while GR mediates the deleterious effects<br />

of excessive glucocorticoids. Studies demonstrating rescue of dentate gyrus granule neurons<br />

from adrenalectomy-induced death provide some support for the notion that MR has<br />

neurotrophic properties (Woolley et al., 1991). In addition, numerous studies have demonstrated<br />

1


that chronic GR activation compromises neuronal survival in hippocampus (see Sapolsky, 1994),<br />

a structure particularly rich in ACR expression (Reul and de Kloet, 1985) and particularly<br />

vulnerable to neuronal injury and degeneration (see Sapolsky, 1994).<br />

Improved understanding of the mechanisms underlying pathogenic effects of glucocorticoids<br />

is of crucial importance to facilitate treatment of neurodegenerative and psychiatric illness. <strong>The</strong><br />

negative effects of chronic exposure to supraphysiological glucocorticoid levels have been well<br />

established in animal models (Sapolsky, 1985; Sapolsky and Pulsinelli, 1985; Sapolsky, 1986a,<br />

b; Elliott et al., 1993; Stein-Behrens et al., 1994a), and in vitro studies provide convincing<br />

evidence that deleterious glucocorticoid effects are mediated by GR (Packan and Sapolsky,<br />

1990; Behl et al., 1997b). Some studies demonstrate that GR blockade with antagonist drug<br />

mifepristone (RU486) can protect cultured cells against glucocorticoid toxicity (Packan and<br />

Sapolsky, 1990; Behl et al., 1997b; Behl et al., 1997a). However, the potential for attenuating<br />

physiological glucocorticoid neurotoxicity with GR blockade has not been thoroughly examined<br />

in vivo. One previous study has demonstrated transient RU486 protection against ischemiainduced<br />

hippocampal cell loss (Antonawich et al., 1999), but RU486 protection against other<br />

forms of neuronal injury has not been studied in animal models. Furthermore, the possible<br />

neuroprotective role of normal MR activation has not been previously addressed.<br />

<strong>The</strong> current thesis aims to evaluate the respective roles of MR and GR on hippocampal<br />

neuron survival in vivo under conditions of neuronal challenge. Adrenocorticosteroid receptor<br />

regulation of neuronal viability was assessed after administration of ACR antagonist drugs and<br />

injury with kainic acid, an excitotoxic glutamate analog, or controlled cortical impact, a model of<br />

experimental traumatic brain injury. Following antagonist pretreatments and injury, survival of<br />

challenged neurons, neuronal expression of viability-related genes, and neuronal expression of<br />

ACR mRNAs were measured to test the hypotheses that (1) MR blockade would decrease<br />

hippocampal neuron survival of injury and expression of genes associated with neuronal<br />

viability, (2) GR blockade would increase hippocampal neuron survival of injury and expression<br />

of genes associated with neuronal viability, and (3) glucocorticoid signaling is altered through<br />

regulation of ACR expression following injury.<br />

2


<strong>The</strong> hypothalamic-pituitary-adrenal axis<br />

Glucocorticoid secretion is regulated by the hypothalamic-pituitary-adrenal (HPA) axis<br />

(Figure 1.1). Perception of stress or circadian drive stimulates the paraventricular nucleus of the<br />

hypothalamus to release corticotropin-releasing hormone (CRH) and other adrenocorticotropinreleasing<br />

hormone (ACTH) secretogogues into the hypophysial portal system in the median<br />

eminence. <strong>The</strong> neuropeptides stimulate anterior pituitary release of ACTH, which travels though<br />

systemic circulation and promotes synthesis and release of glucocorticoids from the adrenal<br />

cortex [see (Antoni, 1986; Jones and Gillham, 1988)]. Circulating glucocorticoids subsequently<br />

inhibit their own release by acting at multiple levels of the HPA system including brain and<br />

pituitary (Keller-Wood and Dallman, 1984; Jones and Gillham, 1988). This negative feedback<br />

system participates in termination of the stress response and prevents overexposure to catabolic<br />

glucocorticoid effects that might threaten homeostasis (Munck et al., 1984). Hypothalamicpituitary-adrenal<br />

dysregulation and elevated basal cortisol levels have been associated with<br />

affective disease and cognitive decline in humans (Sapolsky et al., 1986; Martignoni et al., 1992;<br />

Axelson et al., 1993).<br />

3


Figure 1.1. <strong>The</strong> hypothalamic-pituitary-adrenal axis. Stress or circadian drive stimulate the<br />

paraventricular nucleus of the hypothalamus (PVN) to release corticotropin releasing hormone<br />

(CRH), which stimulates release of adrenocorticotropin-releasing hormone (ACTH) from the<br />

anterior pituitary. ACTH promotes glucocorticoid release from the adrenal cortex.<br />

Glucocorticoids act at multiple levels to inhibit further HPA activation.<br />

4


Properties of adrenocorticosteroid receptors<br />

As members of the nuclear hormone receptor superfamily, MR and GR share a characteristic<br />

three-domain structure consisting of an N-terminal transactivation domain, a central DNAbinding<br />

domain, and a carboxy-terminal ligand-binding domain (Evans, 1988) (Figure 1.2). <strong>The</strong><br />

non-conserved N-terminal domain provides unique transactivation properties to each receptor<br />

(Evans, 1988). <strong>The</strong> highly conserved DNA-binding domain suggests that both receptor types<br />

bind to identical response elements on target genes (Arriza et al., 1987). In addition to<br />

interaction with hormone ligand, the carboxy-terminal domain also mediates binding to<br />

chaperone proteins, nuclear translocation, dimerization, and transactivation (Bertorelli et al.,<br />

1998). Rat and human ACR sequences are highly homologous; comparisons indicate complete<br />

identity in MR and GR DNA-binding domains and high coincidence in steroid-binding domains<br />

(Hollenberg et al., 1985; Miesfeld et al., 1986; Patel et al., 1989).<br />

Figure 1.2. Functional domains of the rat glucocorticoid receptor (GR) and mineralocorticoid<br />

receptor (MR). Each protein consists of three main functional domains: an N-terminal<br />

transactivation domain, a central DNA binding domain, and a carboxy-terminal hormone-binding<br />

domain (Miesfeld et al., 1986; Picard and Yamamoto, 1987; Evans, 1988; Patel et al., 1989).<br />

Approximate percentages of amino acid identity within each domain are based on human MR<br />

and GR homology (Evans, 1988).<br />

5


Differences in ligand binding affinity underlie the differential activation of MR and GR with<br />

increasing plasma glucocorticoid levels. <strong>The</strong> MR, also termed the “type I” receptor, binds<br />

corticosterone with high affinity (Kd ∼ 0.5-1 nM), resulting in extensive occupation even at low<br />

circulating glucocorticoid levels (Reul and de Kloet, 1985). <strong>The</strong> GR, or “type II” receptor, has a<br />

lower affinity for corticosterone (Kd ∼ 2.5-5 nM) and is activated (in addition to MR) only at<br />

high CORT levels as seen at the circadian peak or during stress (Reul and de Kloet, 1985). In<br />

contrast to a simple dose-response relationship, this two-receptor system is capable of generating<br />

unique cellular responses based on the strength of glucocorticoid signal.<br />

Central nervous system distribution of adrenocorticosteroid receptors<br />

<strong>The</strong> GR is widely expressed in the brain, in both neurons and glial cells (Fuxe et al., 1985;<br />

van Eekelen et al., 1987; Ahima and Harlan, 1990). Particularly high levels of GR expression<br />

are found in the hippocampal formation; other regions expressing moderate to high levels of GR<br />

include the paraventricular nucleus of the hypothalamus, the lateral septum, brainstem<br />

monoaminergic neurons, thalamic nuclei, the central amygdala, striatum, and layers of cortex<br />

(Aronsson et al., 1988; Van Eekelen et al., 1988; Chao et al., 1989). Expression of MR is limited<br />

mainly to limbic areas, with highest expression in the hippocampal formation (Reul and de<br />

Kloet, 1985; Ahima and Harlan, 1990). Colocalization of both MR and GR has been reported in<br />

dorsal subiculum, septum, amygdala, and particularly in hippocampus (Reul and de Kloet, 1985;<br />

Van Eekelen et al., 1988). <strong>The</strong> distribution of ACRs varies by subregion in hippocampus<br />

(Herman et al., 1989a; Herman, 1993) (Figure 1.3). <strong>The</strong> rich expression of ACRs in<br />

hippocampal neurons may contribute to their unusual vulnerability to challenge and their<br />

extreme sensitivity to the neurotoxic effects of glucocorticoids.<br />

6


Figure 1.3. Microdistribution of mineralocorticoid receptor (MR) (panel A) and glucocorticoid<br />

receptor (GR) (panel B) mRNAs in rat hippocampus. MR and GR mRNAs are heterogeneously<br />

distributed within the hippocampal formation (Herman et al., 1989a; Herman, 1993). MR<br />

mRNA levels are highest in CA2 and comparable in CA1, CA3, and dentate gyrus (DG) (panel<br />

A). GR signal is highly expressed in CA1 and DG and low in CA3 (panel B).<br />

Regulation of adrenocorticosteroid receptor expression<br />

Adrenocorticosteroid receptor expression is subject to short-term and long-term regulation by<br />

many factors, including glucocorticoids. Autoregulation of hippocampal ACR expression has<br />

been demonstrated in adrenalectomized animals and in subjects treated with ACR antagonists<br />

(i.e. receptor expression is increased in the absence of receptor activation). Results from several<br />

studies are consistent with regulation of MR expression by MR (Chao et al., 1998a; <strong>McCullers</strong><br />

and Herman, 2001) and with regulation of GR expression by both MR (O'Donnell and Meaney,<br />

1994; Chao et al., 1998a; <strong>McCullers</strong> and Herman, 2001) and GR (Herman et al., 1989a; Reul et<br />

al., 1989; Chao et al., 1998a; Herman and Spencer, 1998; <strong>McCullers</strong> and Herman, 2001)<br />

(Chapter Three). Furthermore, ACR expression is subject to regulation by glucocorticoid<br />

secretion during acute stress (Herman and Watson, 1995; Paskitti et al., 2000), chronic stress<br />

(Sapolsky et al., 1984; Gomez et al., 1996; Paskitti et al., 2000), social conflict (Johren et al.,<br />

1994), and the circadian rising phase (Herman et al., 1993). It is important to note that ACR<br />

expression is differentially regulated by brain region, particularly across subfields of the<br />

hippocampus (Herman, 1993).<br />

Adrenocorticosteroid receptor expression is also regulated by numerous factors other than<br />

glucocorticoids, emphasizing the importance of ACR regulation as a means for modulating<br />

glucocorticoid signaling. Neurotransmitters appear to play a prominent role in regulating ACR<br />

7


levels; there is evidence for serotonergic (Seckl et al., 1990; Mitchell et al., 1992), noradrenergic<br />

(Yau and Seckl, 1992), and dopaminergic (Antakly et al., 1987) regulation of neuronal ACR<br />

mRNAs and/or receptor binding. Antidepressant drugs also affect ACR density, but whether<br />

antidepressants act directly to alter ACR expression or indirectly by increasing monoamine<br />

levels is unclear (Seckl and Fink, 1992). Disruption of hippocampal circuitry also influences<br />

ACR mRNA expression; lesions of the medial septum (Yau and Seckl, 1992), dentate gyrus<br />

(Brady et al., 1992), entorhinal cortex (O'Donnell et al., 1993), and central 5-HT lesions (Yau et<br />

al., 1994) can regulate hippocampal ACR mRNA levels. Excitatory amino acids (Lowy, 1992)<br />

and growth factors (Sarrieau et al., 1996) are among numerous other factors shown to regulate<br />

ACR expression.<br />

Expression of ACR mRNAs is regulated at multiple levels including transcription,<br />

polyadenylation, and/or alternative splicing (see De Kloet et al., 1998). Splicing at the 3’-end of<br />

human GR heteronuclear RNA results in at least two GR mRNA variants, GRα and GRβ<br />

(Hollenberg et al., 1985). Notably, GRβ is not expressed in rat (Otto et al., 1997). Rat GR and<br />

MR mRNAs exhibit heterogeneity in 5’-untranslated exon 1 (McCormick et al., 2000) (Kwak et<br />

al., 1993; Zennaro et al., 1995). Interestingly, expression of rat GR mRNA variants can be<br />

regulated by environmental manipulations (i.e. postnatal handling) (McCormick et al., 2000),<br />

and select MR transcripts are regulated by glucocorticoids (Kwak et al., 1993). Splice variants<br />

may exhibit different stabilities and/or translation efficiencies, influencing overall levels of ACR<br />

expression based on the differential expression of transcripts (Kwak et al., 1993; Zennaro et al.,<br />

1995).<br />

Adrenocorticosteroid receptor activation<br />

Adrenocorticosteroid receptors are activated upon the binding of glucocorticoid ligand,<br />

specifically cortisol in human or corticosterone (CORT) in the rat. Unliganded ACRs reside in<br />

the cytoplasm as part of a multi-protein complex containing two molecules of heat shock protein<br />

(HSP) 90, a molecule each of hsp70, hsp56, p23, and an immunophilin (Pratt, 1993; Bruner et<br />

al., 1997; Bertorelli et al., 1998). <strong>The</strong> chaperone complex facilitates hormone binding to the<br />

receptor and participates in receptor trafficking (Pratt, 1993). Glucocorticoid binding induces a<br />

conformational change in the ACR molecule that favors dissociation from the chaperone<br />

complex (Pratt, 1993). Following dissociation, the liganded receptors become<br />

8


hyperphosphorylated (Orti et al., 1992) and translocate into the nucleus (Picard and Yamamoto,<br />

1987) where they act as transcription factors (Yamamoto, 1985).<br />

Adrenocorticosteroid receptor regulation of gene transcription<br />

Models proposed to explain ACR regulation of gene transcription identify chromatin<br />

remodeling and stabilization of the transcriptional preinitiation complex as essential mechanisms<br />

(Jenster et al., 1997). Adrenocorticosteroid receptor regulation of these processes is<br />

accomplished via binding to DNA response elements or through interactions with other<br />

transcription factors (Figure 1.4). In the case of the former, the activated receptors can bind as<br />

homo- or heterodimers (Luisi et al., 1991; Trapp et al., 1994) to specific DNA sequences called<br />

glucocorticoid response elements (GREs). <strong>The</strong>se palindromic sequences are located in or near<br />

promoter regions of glucocorticoid-responsive genes and serve as inducible enhancer elements<br />

(see Yamamoto, 1985). Response element-bound ACRs recruit coactivator proteins capable of<br />

remodeling chromatin structure via histone acetyltransferase activity. By acetylating histone N-<br />

termini, coactivators loosen nucleosomal structure and increase accessibility for other<br />

transcription factors. <strong>The</strong> steroid receptors and coactivators may also stabilize assembly of the<br />

preinitiation complex, thereby increasing the rate of transcription initiation (Jenster et al., 1997).<br />

<strong>The</strong> MR and GR can repress gene transcription by recruiting corepressor proteins or by acting at<br />

negative GREs, which differ in sequence from positive GREs (Drouin et al., 1993). It is<br />

important to note that interactions with coactivator proteins will vary depending on dimer<br />

configuration, with distinct transactivation properties conferred by each type of homo- or<br />

heterodimer (Trapp et al., 1994).<br />

Ligand-activated ACRs may also influence transcription of genes via interactions with other<br />

transcription factors (Karin, 1998) (Figure 1.4). For instance, GR modulates activity of<br />

transcription factors including activator protein-1 (AP-1), cyclic AMP response element binding<br />

protein (CREB), and nuclear factor-κB (NF-κB), in the absence of DNA binding (Herrlich,<br />

2001). Steroid receptor cross-talk with transcription factors most frequently results in negative<br />

interference, but synergism is also possible under specific conditions (i.e. when AP-1 is a<br />

homodimer of c-Jun) (Diamond et al., 1990). <strong>The</strong> exact molecular mechanism of steroid<br />

receptor cross-talk has yet to be elucidated. Models for direct protein-protein interactions<br />

between GR and other transcription factors have been proposed (McEwan et al., 1997).<br />

9


Adrenocorticosteroid receptors may also repress gene transcription by inducing specific<br />

inhibitors of other transcription factors. For example, GR may repress NF-κB activity by<br />

inducing IκB, an inhibitor of NF-κB (Auphan et al., 1995; Scheinman et al., 1995). <strong>The</strong><br />

evidence indicates that multiple mechanisms are likely to account for glucocorticoid-mediated<br />

gene repression (see De Bosscher et al., 2000).<br />

10


Figure 1.4. Models for transcriptional regulation by adrenocorticosteriod receptors. Activated<br />

mineralocorticoid receptors (MRs) or glucocorticoid receptors (GRs) bind DNA hormone<br />

response elements as dimers or interact with other transcription factors to influence transcription<br />

(McEwan et al., 1997; Karin, 1998; De Bosscher et al., 2000). Possible mechanisms for<br />

modulating transcription include: recruitment of transcription coactivators or repressors (A, B);<br />

binding to negative hormone response elements (nGREs)(C); interference with activity of other<br />

transcription factors (D); competive DNA binding (E, F); and induction of inhibitors of other<br />

transcription factors (G).<br />

11


Adrenocorticosteriod receptor roles in neuronal viability<br />

Differential ACR activation at varying glucocorticoid levels can elicit unique cellular<br />

responses due to the distinct regulatory effects of MR and GR. Collaborative binding of MR/GR<br />

heterodimers to GREs demonstrates the potential for ACR cooperation in regulating gene<br />

expression (Trapp et al., 1994). In some cases, however, MR and GR effects appear to be quite<br />

opposite. For instance, predominant MR activation at low glucocorticoid concentrations results<br />

in very small depolarization-induced calcium (Ca 2+ ) currents in CA1 pyramidal neurons (Karst et<br />

al., 1994), while the additional occupation of GR at high glucocorticoid levels results in<br />

relatively large Ca 2+ currents (Kerr et al., 1992). Further, MR activation by low glucocorticoid<br />

levels or aldosterone decreases accommodation and afterhyperpolarization amplitude in CA1,<br />

enhancing excitatory transmission (Joels and de Kloet, 1990); in contrast, high glucocorticoid<br />

levels and GR agonist RU28362 increase accommodation and the duration of<br />

afterhyperpolarization, suppressing CA1 excitability (Kerr et al., 1989). Differential activation<br />

of GR can also elicit structural effects on hippocampal neurons that do not occur at lower, MRactivating<br />

levels. For instance, chronic exposure to GR-saturating glucocorticoid levels achieved<br />

by 21-day corticosterone (CORT) administration (Woolley et al., 1990; Magarinos and McEwen,<br />

1995) or repeated stress (Magarinos and McEwen, 1995) reversibly decreases both numbers and<br />

length of apical dendritic branches in CA3 pyramidal neurons.<br />

<strong>The</strong> preferential activation of MR or GR can also influence the survival of hippocampal<br />

neurons. <strong>The</strong> vast majority of literature to date examines negative GR effects on hippocampal<br />

neuron viability, as GR activation is associated with the exacerbation of several forms of<br />

neuronal insult (Sapolsky, 1985; Sapolsky and Pulsinelli, 1985; Sapolsky, 1986a, b; Sapolsky et<br />

al., 1988; Packan and Sapolsky, 1990; Elliott et al., 1993; Stein-Behrens et al., 1994a; Goodman<br />

et al., 1996; McIntosh and Sapolsky, 1996a; Behl et al., 1997b). In contrast, MR activation is not<br />

associated with reduced neuronal survival (Packan and Sapolsky, 1990), and there is some<br />

evidence to indicate that MR may have trophic or neuroprotective effects (Woolley et al., 1991).<br />

Excessive glucocorticoid receptor activation compromises hippocampal neuron viability<br />

Chronic exposure to high concentrations of glucocorticoids can be toxic to neurons, with<br />

preferential damage to the hippocampus (see Sapolsky, 1994). Studies describing hippocampal<br />

12


degeneration in aging male rats with elevated basal glucocorticoid levels provided the first<br />

evidence of a physiological role for glucocorticoid neurotoxicity. Landfield and colleagues<br />

demonstrated a correlation between the degree of glucocorticoid hypersecretion and the amount<br />

of hippocampal degeneration in mid-aged rats (Landfield et al., 1978). Removal of<br />

glucocorticoids via adrenalectomy was subsequently shown to prevent the hippocampal cell loss<br />

and glial reactivity (Landfield et al., 1981). <strong>The</strong>se studies led to the formation of the<br />

“glucocorticoid hypothesis of brain aging and neurodegeneration”, which proposes that<br />

glucocorticoids contribute to age-related degeneration of the hippocampus (Landfield and<br />

Eldridge, 1991). Sapolsky and colleagues later demonstrated that chronic exposure to high<br />

levels of glucocorticoids, achieved by repeated injection of corticosterone, resulted in the death<br />

of hippocampal pyramidal neurons in rat (Sapolsky et al., 1985). It remains unclear whether<br />

glucocorticoids alone are sufficient to induce neuronal death in hippocampus.<br />

While glucocorticoids may not kill neurons directly, there is ample support for the notion that<br />

chronic exposure to high levels of glucocorticoids increases the vulnerability of hippocampal<br />

neurons to subsequent challenge (Sapolsky, 1985). In rat, administration of high physiologic<br />

concentrations of CORT potentiates ischemic injury to CA3 neurons produced by the four-vessel<br />

occlusion method of forebrain ischemia (Sapolsky and Pulsinelli, 1985). High CORT levels also<br />

enhance damage induced in CA3 by the excitotoxin kainic acid (Sapolsky, 1985, 1986a, b;<br />

Elliott et al., 1993; Stein-Behrens et al., 1994a) and in DG by the antimetabolite toxin 3-<br />

acetylpyridine (Sapolsky, 1985, 1986b). Glucocorticoid exacerbation of excitotoxic (Sapolsky et<br />

al., 1988; Packan and Sapolsky, 1990; Goodman et al., 1996) and metabolic (Sapolsky et al.,<br />

1988) challenge is also seen in primary hippocampal cultures. Furthermore, in vitro experiments<br />

have shown that glucocorticoids increase the amount of hippocampal cell death induced by<br />

oxidative stress (Sapolsky et al., 1988; Goodman et al., 1996; McIntosh and Sapolsky, 1996a;<br />

Behl et al., 1997b) and amyloid β-peptide toxicity (Goodman et al., 1996).<br />

Despite the graded activation of GR with increasing CORT levels, it remains difficult to<br />

differentiate between specific MR and GR effects at high CORT concentrations, as both receptor<br />

types are activated. Likewise, both MR and GR are inactivated in adrenalectomized animals due<br />

to the removal of endogenous CORT. In order to dissect the individual effects of each receptor<br />

type, selective MR/GR agonists may be administered in adrenalectomized animals or in culture<br />

to study the selective activation of a receptor type. Conversely, MR/GR antagonists may be<br />

13


administered to intact animals or in culture to observe the effects of loss of function of a specific<br />

ACR type. A number of studies using ACR agonists and antagonists in vitro have provided<br />

convincing evidence for the hypothesis that the endangering effects of excessive glucocorticoid<br />

exposure are mediated by the GR. In primary hippocampal cultures, a specific GR agonist,<br />

dexamethasone, exacerbates kainic acid damage, where the MR agonist aldosterone has no effect<br />

(Packan and Sapolsky, 1990). In the same culture system, GR antagonist RU486 protects against<br />

glucocorticoid/kainic acid synergy, but MR antagonist RU28318 does not prevent the damaging<br />

glucocorticoid effects (Packan and Sapolsky, 1990). Other in vitro studies provide strong<br />

support for GR enhancement of amyloid β protein, hydrogen peroxide, and glutamate toxicity, as<br />

dexamethasone exacerbates (Behl et al., 1997b) and RU486 prevents (Behl et al., 1997b; Behl et<br />

al., 1997a) negative glucocorticoid effects with these types of insult. In vivo, RU486 transiently<br />

protects gerbil CA1 neurons from ischemia-induced cell loss, indicating that blockade of GR can<br />

prevent the detrimental effects of post-ischemia CORT elevations (Antonawich et al., 1999).<br />

Mechanisms for glucocorticoid receptor endangerment of hippocampal neurons<br />

Glucocorticoids can increase neuronal susceptibility to injury by inducing a state of<br />

metabolic vulnerability (Sapolsky, 1985). In the absence of additional challenge, temporary<br />

energy crisis induced by glucocorticoids is not likely to threaten neuronal survival. However, by<br />

compromising cellular energy supplies, glucocorticoids can greatly intensify the toxicity of a<br />

wide variety of insults that impair generation of energy (i.e. ischemia) or place pathogenic<br />

demands on the neuron for energy (i.e. excitotoxicity) (Sapolsky, 1990). Glucocorticoids impair<br />

glucose metabolism in regions including hippocampus, locus coeruleus, hypothalamic<br />

paraventricular nucleus, median eminence, and pituitary anterior lobe (Kadekaro et al., 1988).<br />

Activation of GR may compromise neuronal energy stores by impairing glucose transport, as<br />

demonstrated in cultured hippocampal neurons and glia in vitro (Horner et al., 1990). Chronic<br />

glucocorticoid exposure also markedly induces expression of ATP-requiring enzymes such as<br />

glutamine synthetase and glycerol-phosphate dehydrogenase, which may result in the further<br />

depletion of cellular ATP stores (Nichols and Finch, 1994). Energy supplementation with<br />

glucose or mannose has been shown to attenuate glucocorticoid neurotoxicity in vivo (Sapolsky,<br />

1986b) and in vitro (Sapolsky et al., 1988).<br />

14


Neuronal energy depletion following excessive glucocorticoid exposure has been proposed to<br />

exacerbate hippocampal insult by contributing to excitotoxic cascades (Sapolsky, 1990). Under<br />

normal conditions, glutamate interaction with postsynaptic receptors mobilizes free cytosolic<br />

Ca 2+ in the postsynaptic neuron in a highly regulated fashion. Glutamate is removed from the<br />

synapse by energy-dependent glutamate transporters (Kanai et al., 1993). Calcium in the<br />

postsynaptic neuron is effectively buffered by Ca 2+ binding proteins including calbindin and<br />

parvalbumin, sequestered by cellular organelles such as smooth endoplasmic reticulum and<br />

mitochondria, or extruded via ATP-dependent transporters including a Na + /Ca 2+ exchanger and<br />

the plasma membrane calcium pump (PMCA) (Blaustein, 1988). However, energy depletion and<br />

necrotic insult can lead to the pathologic accumulation of synaptic excitatory amino acids,<br />

leading to excessive influx of Ca 2+ into the postsynaptic neuron. Energy crisis compromises<br />

glutamate removal from the synapse and may even result in reversal of glutamate transporters<br />

(Takahashi et al., 1997), resulting in the further depolarization-induced opening of voltage-gated<br />

Ca 2+ channels and NMDA-gated Ca 2+ channels on the postsynaptic neuron. Energy deficits can<br />

also lead to the failure of ATP-driven Ca+ sequestration and extrusion systems, exacerbating<br />

increases in intracellular Ca 2+ levels (see Tymianski and Tator, 1996). Excessive Ca 2+ influx can<br />

set in motion many destructive processes including mitochondrial failure, the degradation of<br />

structural proteins by Ca 2+ -activated proteases (i.e. calpain)(Siman et al., 1989), and the<br />

generation of free radicals [see (Choi, 1992; Tymianski and Tator, 1996)].<br />

Many experimental studies support the theory that glucocorticoids contribute to excitotoxic<br />

glutamate/Ca 2+ cascades. Chronic exposure to high levels of glucocorticoids has been shown to<br />

increase extracellular levels of excitory amino acids following kainic acid administration (Stein-<br />

Behrens et al., 1992; Stein-Behrens et al., 1994b). Glucocorticoids also prolong intracellular<br />

Ca 2+<br />

elevation after kainic acid treatment (Elliott and Sapolsky, 1992) and increase basal<br />

intracellular Ca 2+ levels (Elliott and Sapolsky, 1993) in culture. Furthermore, glucocorticoids<br />

exacerbate Ca 2+ -mediated degenerative events in the hippocampus following intrahippocampal<br />

kainic acid infusion, as measured by spectrin proteolysis and altered tau immunoreactivity<br />

(Elliott et al., 1993; Stein-Behrens et al., 1994a). Glucocorticoid exacerbation of 3-<br />

acetylpyridine toxicity is eliminated with administration of an NMDA receptor antagonist,<br />

aminophosphonovaleric acid, providing evidence that negative glucocorticoid effects are<br />

dependent on activation of the NMDA receptor (Armanini et al., 1990).<br />

15


Glucocorticoids can also influence neuronal survival via genomic mechanisms. High<br />

glucocorticoid concentrations may modulate Ca 2+ homeostasis by increasing P/Q- and L- type<br />

Ca 2+ channel subunit mRNA levels, as demonstrated in isolated CA1 neurons (Nair et al., 1998).<br />

Glucocorticoids also decrease hippocampal mRNA levels of plasma membrane calcium pump<br />

isoform 1 (PMCA1) (Bhargava et al., 2000), which regulates intracellular Ca 2+ levels by<br />

coupling ATP hydrolysis with Ca 2+ extrusion. <strong>The</strong>se changes in gene expression, if indicative of<br />

changes in protein expression, are consistent with increased intracellular Ca 2+ accumulation with<br />

GR activation. Furthermore, glucocorticoid regulation of neurotrophin expression may lead to<br />

decreased neurotrophic support and the exacerbation of neuronal insults. Immobilization stress<br />

(Smith et al., 1995a) and CORT administration (Chao and McEwen, 1994; Smith et al., 1995a)<br />

decrease hippocampal brain-derived neurotrophic factor (BDNF) mRNA levels and increase<br />

hippocampal neurotrophin-3 (NT-3) levels in hippocampus.<br />

Neurotrophic effects of mineralocorticoid receptor activation<br />

<strong>The</strong> potentially damaging effects of GR activation at high glucocorticoid concentrations are<br />

well documented. In contrast, the MR, which is persistently activated even at basal<br />

glucocorticoid levels (Reul and de Kloet, 1985), is not associated with neuronal energy depletion<br />

or exacerbation of neuronal insult (Packan and Sapolsky, 1990). On the contrary, MR may<br />

function in a neurotrophic manner, as activation of MR is required for the survival of dentate<br />

gyrus granule neurons. Removal of endogenous glucocorticoids by adrenalectomy causes<br />

granule cell loss within as little as three days (Gould et al., 1990), and nearly all granule cells are<br />

lost within three to four months after surgery (Sloviter et al., 1989). Administration of the<br />

selective MR agonist aldosterone prevents adrenalectomy-induced granule neuron loss, but<br />

treatment with GR agonist RU28362 provides only partial protection for DG granule neurons<br />

(Woolley et al., 1991). Thus, MR occupation is necessary for maintenance of DG neuron<br />

survival, but the role for GR is less clear (Woolley et al., 1991). MR activation also regulates<br />

granule cell neurogenesis in DG, maintaining a balance between granule cell birth and cell death<br />

(Gould et al., 1991). MR may regulate neurogenesis by enhancing excitatory input to the region<br />

(Gould, 1994) or indirectly through regulation of neurotrophin expression (Lindholm et al.,<br />

1994; Qiao et al., 1996).<br />

16


Kainic acid as a model of excitotoxicity<br />

Kainic acid (KA) is a glutamate analog notable for its excitotoxic and epileptogenic effects.<br />

Systemic or intracerebral injection of KA induces epileptiform seizures and cell loss in<br />

hippocampus in a pattern similar to human temporal lobe epilepsy (Nadler et al., 1978; Ben-Ari<br />

and Cossart, 2000). <strong>The</strong> extensive literature demonstrating glucocorticoid modulation of KA<br />

excitotoxicity, the selective toxicity of KA in hippocampus, and the substantial but incomplete<br />

neuronal loss following seizure indicate KA as an ideal model for use in studying ACR effects<br />

on hippocampal neuron viability.<br />

Glutamate receptors<br />

Glutamate receptors are divided into two major categories - metabotropic and ionotropic<br />

receptors. Metabotropic glutamate receptors are coupled to G-protein-mediated second<br />

messenger pathways; ionotropic glutamate receptors allow the direct influx of cations upon<br />

activation. Ionotropic glutamate receptors are responsible for the majority of fast excitatory<br />

synaptic transmission in the central nervous system. Three types of ionotropic glutamate<br />

receptors have been identified based on their preferential activation by each compound – the<br />

AMPA, kainate, and NMDA receptors (Dingledine and McBain, 1999).<br />

Non-NMDA glutamate receptor subunits are approximately 900 amino acid residues in<br />

length, have a calculated molecular weight of approximately 100 kDa, and are predicted to have<br />

four transmembrane domains (Keinanen et al., 1990). Non-NMDA receptors are formed by<br />

homomeric or heteromeric associations of AMPA (GluR-1/-2/-3/-4) or kainate (GluR-5/-6/-7 and<br />

KA-1/-2) receptor subunits. Subunit composition may confer distinct functional and<br />

pharmacological properties (Hollmann et al., 1991; Bettler and Mulle, 1995). Kainate receptors<br />

exhibit significant structural homology to AMPA receptors and can be expressed in the same<br />

cells as AMPA receptors, yet KA receptor subunits and AMPA receptor subunits do not seem to<br />

coassemble (Frerking and Nicoll, 2000). Kainate receptors generate excitatory postsynaptic<br />

potentials (EPSPs) in response to glutamate release (Frerking and Nicoll, 2000), and kainate<br />

responses undergo rapid and total desensitization from which recovery is slow (Lerma et al.,<br />

1997). As compared to AMPA receptor-mediated EPSPs, kainate-mediated EPSPs have smaller<br />

17


peak amplitude and slower decay kinetics. A controversial presynaptic role for kainate receptors<br />

has also been proposed (see Malva et al., 1998).<br />

NMDA receptors are believed to play a key role in glutamate neurotoxicity due to their high<br />

Ca 2+ permeability as compared to non-NMDA ionotropic glutamate receptors (Rothman, 1987).<br />

NMDA receptors are unique in that they require bound ligand and the voltage-dependent<br />

removal of a Mg2+-dependent block to initiate activation, and the presence of glycine is required<br />

for full activation (Malva et al., 1998). NMDA receptors are insensitive to AMPA and kainate<br />

(Hollmann and Heinemann, 1994) but may be activated by kainate-induced seizures or<br />

excitotoxic cascades (Ben-Ari and Gho, 1988).<br />

Kainic acid excitotoxicity<br />

<strong>The</strong> distribution of KA binding sites roughly corresponds to regional susceptibility to KA<br />

excitotoxicity. Regions displaying KA binding and sensitivity to KA toxicity include the<br />

hippocampus, amygdala, entorhinal cortex, piriform cortex and insular cortex (Kohler and<br />

Schwarcz, 1983). High-affinity kainate receptors are expressed at very high density in rat brain<br />

in the mossy fiber synaptic region, or stratum lucidum of CA3 (Foster et al., 1981; Represa et al.,<br />

1987). <strong>The</strong> epileptogenic effects of KA are attributed to excitation of CA3 neurons by postsynaptic<br />

kainate receptors at mossy fiber synapses (Ben-Ari and Gho, 1988). It is important to<br />

note that degeneration of CA3 neurons is probably due to the sustained release of glutamate<br />

subsequent to recurrent seizures, not to the activation of KA receptors per se (Coyle et al., 1981;<br />

Ben-Ari, 1985; Ben-Ari and Cossart, 2000). Seizures generated in CA3 propagate to CA1 via a<br />

network of recurrent collateral glutamatergic axons that interconnect pyramidal neurons.<br />

Secondary activation of CA1, the major source of hippocampal output, leads to the spread of<br />

seizures to other limbic structures including entorhinal cortex (Figure 1.5) (Ben-Ari and Cossart,<br />

2000).<br />

18


Figure 1.5. Schematic diagram of the intrinsic circuitry of the hippocampal formation (adapted<br />

from Amaral and Witter, 1989). <strong>The</strong> hippocampal formation is composed of four basic regions:<br />

the entorhinal cortex (EC), the dentate gyrus (DG), the hippocampus proper (which may be<br />

divided into subfields CA1, CA2, and CA3), and the subicular complex (S). <strong>The</strong> major<br />

connections linking hippocampal subfields are glutamatergic and largely unidirectional. Input<br />

from EC reaches DG via the perforant path. Granule cells of the DG project to the CA3 field via<br />

mossy fibers. CA3 pyramidal neurons give rise to Schaffer collaterals, which terminate on CA1<br />

pyramidal neurons. CA1 pyramidal neurons project to S, which projects to regions including<br />

EC.<br />

<strong>The</strong> behavioral effects of systemic KA administration have been described in detail (Sperk et<br />

al., 1983). Approximately five minutes after injection, most animals assume a catatonic posture<br />

with staring behavior. Within approximately one hour, subjects develop generalized clonic-tonic<br />

convulsions accompanied by frequent rearing, falling over to one side, “wet dog shakes”, and<br />

hemorrhagic foam from the mouth. Convulsions often subside two hours after injection and are<br />

replaced by a phase of motor hyperactivity. In more severely affected animals, the convulsive<br />

phase may last up to 4-5 hours after KA injection. After twenty-four hours the animals appear<br />

physically exhausted. Rare convulsions may still occur spontaneously or in response to noise or<br />

handling (Sperk et al., 1983).<br />

Kainic acid lesion preferentially destroys hippocampal pyramidal cells, particularly those of<br />

region CA3 (Nadler et al., 1978). Neurons of CA1 are vulnerable to a lesser degree, and DG<br />

19


granule cells are relatively insensitive to KA excitotoxicity (Nadler et al., 1978; Kohler and<br />

Schwarcz, 1983). As early as one to three hours after exposure, damaged neurons appear<br />

shrunken, with darkly stained cytoplasm and pyknotic nuclei (Sperk et al., 1983) (Figure 1.6).<br />

Some neurons appear swollen, containing small vacuoles in mitochondria and in cytoplasm due<br />

to swelling of endoplasmic reticulum and Golgi apparatus (Sperk et al., 1983). Cell death<br />

induced by KA administration has been classified as necrotic (Sperk et al., 1983; Fujikawa et al.,<br />

2000) yet exhibits some characteristics consistent with apoptosis such as TUNEL positivity and<br />

DNA laddering (Fujikawa et al., 2000). Damaged neurons appear to be partially removed by<br />

three days after injection but may be seen for at least two weeks (Nadler et al., 1978). Neuronal<br />

degeneration is accompanied by edema, perivenous hemorrhage, gliosis, and vacuolization of the<br />

neuropil (Nadler et al., 1978; Sperk et al., 1983). <strong>The</strong> entorhinal cortex, piriform cortex, frontal<br />

cortex, and medial thalamic nuclei are also subject to degrees of KA-induced degeneration<br />

(Nadler et al., 1978; Kohler and Schwarcz, 1983; Sperk et al., 1983).<br />

Figure 1.6. Photomicrograph illustrating the effects of (A) saline versus (B) kainic acid on CA1<br />

hippocampal pyramidal neurons (40X magnification). Twenty-four hours following systemic<br />

administration, saline-treated neurons are healthy and densely packed with rounded borders and<br />

prominent nucleoli. Neurons damaged by kainic acid appear pyknotic and shriveled with darkstaining<br />

somata and shrunken processes.<br />

20


Kainate-induced death of hippocampal neurons is likely due to the sustained release of<br />

glutamate during recurrent seizures (Coyle et al., 1981; Ben-Ari, 1985; Ben-Ari and Cossart,<br />

2000) and subsequent increases in intracellular Ca 2+ levels (Hori et al., 1985; Sun et al., 1992).<br />

Excessive accumulation of glutamate in the synapse will activate post-synaptic glutamate<br />

receptors, resulting in Ca 2+ influx. Calcium accumulation is mediated predominantly by the<br />

NMDA receptor, but AMPA or kainate receptors may also contribute to rises in intracellular<br />

Ca 2+ (Choi, 1992). <strong>The</strong> pathologic energy demands of repeated seizures lead to impairment of<br />

Ca 2+ sequestration and extrusion systems, contributing to Ca 2+ elevations (see Tymianski and<br />

Tator, 1996). Calcium homeostasis may be further disrupted by the release of Ca 2+ from internal<br />

stores, activation of certain enzymes (such as calpains and phospholipases), and changes in<br />

immediate early gene expression. Glutamate efflux from injured neurons may further exacerbate<br />

excitotoxic cascades (Choi, 1992).<br />

Glucocorticoid modulation of kainic acid excitotoxicity<br />

Glucocorticoids exacerbate KA toxicity (Sapolsky, 1985, 1986a, b; Sapolsky et al., 1988;<br />

Packan and Sapolsky, 1990; Stein-Behrens et al., 1994a), making KA administration an excellent<br />

model for studying ACR effects on neuronal viability. Glucocorticoids intensify KA-induced<br />

accumulation of extracellular glutamate (Stein-Behrens et al., 1992; Stein-Behrens et al., 1994b)<br />

and KA-induced calcium elevations (Elliott and Sapolsky, 1992, 1993). Chronic exposure to<br />

glucocorticoids also impairs antioxidant defenses in response to KA (McIntosh et al., 1998a) and<br />

increases the electophysiological sensitivity to KA in hippocampal neurons (Talmi et al., 1995).<br />

In addition, glucocorticoids influence immediate early gene responses to KA (Unlap and Jope,<br />

1995a) and exacerbate KA-induced spectrin proteolysis (Elliott et al., 1993; Lowy et al., 1994).<br />

Glucocorticoids may also influence neuronal viability to KA via genomic mechanisms.<br />

Kainate administration results in the altered expression of a multitude of genes, including: cell<br />

survival/death genes (bcl-2, bax, p53, KROX-20, COX-2, cyclin D1, and GADD45) (Sakhi et<br />

al., 1994; Gillardon et al., 1995; Vreugdenhil et al., 1996; Zhu et al., 1997; Hashimoto et al.,<br />

1998; Timsit et al., 1999), transcription factors (AP-1 and NFκB) (Kaminska et al., 1994;<br />

Pennypacker and Hong, 1995; Won et al., 1999), neurotrophic factors [BDNF and nerve growth<br />

factor (NGF)] (Ballarin et al., 1991; Gall et al., 1991; Hashimoto et al., 1998), neuropeptides<br />

(Pennypacker and Hong, 1995; Friedman, 1997), heat shock proteins (Zhang et al., 1997;<br />

21


Hashimoto et al., 1998), immediate early genes (Honkaniemi and Sharp, 1999), GABA-A<br />

receptor subunits (Tsunashima et al., 1997), and countless others. Altered gene expression is<br />

likely to influence cellular response to KA toxicity; accordingly, glucocorticoid modulation of<br />

gene expression may modulate cell survival. Experiments in the current thesis focus on<br />

glucocorticoid regulation of members of the Bcl-2 family, p53, and neurotrophins BDNF and<br />

NT-3, all of which have been implicated in KA-induced neuronal death (see below).<br />

Regulators of cell survival and death<br />

Members of the bcl-2 family of genes are critical regulators of cell viability. Interactions<br />

between pro-survival genes, including bcl-2, bcl-x L , and bcl-w, and pro-death genes, such as bax<br />

and bad, are thought to influence cellular susceptibility to necrotic and apoptotic death (Oltvai et<br />

al., 1993; Korsmeyer, 1995; Merry and Korsmeyer, 1997). Bcl-2 family members influence cell<br />

survival via multiple mechanisms including stabilization of mitochondria and interference with<br />

cell death pathways (see below).<br />

A key event associated with cell death is the opening of a large conductance channel called<br />

the mitochondrial permeability transition pore. Permeability transition can be induced by a<br />

number of stimuli, including sustained increases in intracellular Ca 2+ levels, oxidative stress,<br />

caspase activation (see below), uncoupling of the respiratory chain, and modification of the Bcl-<br />

2 complex (Green and Reed, 1998). Opening of the nonselective permeability transition pore<br />

allows molecules up to approximately 1.5 kD to pass through the mitochondrial inner membrane<br />

into the matrix. <strong>The</strong> loss of inner membrane selectivity causes collapse of the mitochondrial<br />

transmembrane potential and dissipation of the H + gradient, resulting in the uncoupling of the<br />

electron transport chain, a decrease in ATP production, and an increase in production of reactive<br />

oxygen species. In addition, permeability transition pore opening causes release of caspaseactivating<br />

proteins such as cytochrome c and apoptosis-inducing factor (AIF) into the cytoplasm,<br />

promoting apoptosis [see (Kroemer et al., 1997; Green and Reed, 1998)].<br />

Mitochondrial release of cytochrome c promotes apoptosis by initiating protease cascades<br />

programmed to dismantle the cell (Eldadah and Faden, 2000). Cytochrome c facilitates<br />

interaction between apoptotic protease activating factor-1 (Apaf-1) and procaspase-9 (Li et al.,<br />

1997; Zou et al., 1997). Subsequent processing yields caspase-9, which activates “executioner”<br />

caspases such as caspase-3 (Li et al., 1997; Pan et al., 1998b). Caspase-3 cleaves ICAD, the<br />

22


inhibitor of CAD, a caspase-activated deoxyribonuclease (Enari et al., 1998; Sakahira et al.,<br />

1998). Once released from ICAD inhibition, CAD is responsible for internucleosomal DNA<br />

degradation characteristic of apoptosis (Enari et al., 1998; Sakahira et al., 1998).<br />

Conditions favoring apoptosis trigger bax translocation from the cytoplasm to mitochondrial<br />

membranes (Green and Reed, 1998), promoting the release of cytochrome c from the<br />

intramembranous compartment (Jurgensmeier et al., 1998). Anti-apoptotic proteins such as Bcl-<br />

2 and Bcl-xL may inhibit apoptosis by preventing molecules such as Apaf-1 from activating<br />

caspases (Pan et al., 1998a). Bcl-2 can also prevent mitochondrial permeability transition (Green<br />

and Reed, 1998), the release of cytochrome c (Kluck et al., 1997; Yang et al., 1997), and the<br />

release of AIF from mitochondria (Green and Reed, 1998), each of which could prevent caspase<br />

activation. Although the functional significance is not fully understood, Bcl-2 family members<br />

can also form cation-selective pores, which could speculatively influence ion transport, regulate<br />

volume of the matrix space, control pH in the intermembrane space, or induce permeability<br />

transition in mitochondria (Green and Reed, 1998). In addition, Bcl-2 acts at cell nuclei (Marin<br />

et al., 1996), endoplasmic reticula (He et al., 1997), and mitochondria (Green and Reed, 1998) to<br />

influence subcellular Ca 2+ distribution. Whether the many effects of Bcl-2 are separate functions<br />

or consequences of Bcl-2 regulation of a single cell death process remains uncertain.<br />

Accumulating evidence supports a role for bcl-2 family members in modulating neuronal<br />

viability after KA-induced seizure. Systemic KA administration decreases bcl-2 mRNA levels in<br />

rat hippocampus (<strong>McCullers</strong> and Herman, 1998) (Chapter Two) and Bcl-2 immunoreactivity in<br />

mouse hippocampus and cortex (Gillardon et al., 1995). In contrast, systemic KA increases bax<br />

mRNA levels in rat hippocampus (<strong>McCullers</strong> and Herman, 1998; Lopez et al., 1999) (Chapter<br />

Two) and mouse brain (Gillardon et al., 1995). Increased Bcl-w expression has also been<br />

observed in rat hippocampus and piriform cortex following KA microinjection into the amygdala<br />

(Henshall et al., 2001). Bcl-2 promotes neuronal survival of KA insult, as Bcl-2 overexpression<br />

with viral vectors protects hippocampal neurons against KA toxicity in culture and in vivo<br />

(Phillips et al., 2000). Furthermore, KA-induced degeneration is attenuated in Bax-deficient<br />

cultured motor neurons (Bar-Peled et al., 1999), demonstrating Bax participation in KA-induced<br />

cell death.<br />

<strong>The</strong> p53 tumor-suppressor gene has been described as a guardian of the genome. In response<br />

to DNA damage, the transcription factor induces either growth arrest or apoptosis depending<br />

23


upon the cell type, environment, and expression of other regulatory factors. In either scenario,<br />

the end result is the genetic death of the damaged cell, preventing potential neoplastic risk (Evan<br />

and Littlewood, 1998). <strong>The</strong> mechanisms by which p53 promote apoptosis are not fully<br />

delineated, but many studies indicate the involvement of repression or induction of specific<br />

target genes such as bcl-2 and bax (Miyashita et al., 1994; Miyashita and Reed, 1995; Evan and<br />

Littlewood, 1998). Transrepression of anti-apoptotic genes and non-transcriptional mechanisms<br />

may also be involved (Evan and Littlewood, 1998). <strong>The</strong> signaling pathways by which insults<br />

activate p53 are mostly unknown (Evan and Littlewood, 1998).<br />

Experimental evidence suggests that p53 influences hippocampal vulnerability to KAinduced<br />

neuronal degeneration. P53 mRNA is induced in KA-sensitive brain regions including<br />

hippocampus, amygdala, piriform cortex, and thalamus following systemic KA injection (Sakhi<br />

et al., 1994) (Chapter Two), and nuclear accumulation of p53 protein has been observed in<br />

degenerating neurons after KA administration (Sakhi et al., 1996). Furthermore, p53-deficient<br />

mice (“knock-out” mice) exhibit little or no hippocampal cell loss after KA treatment (Morrison<br />

et al., 1996; Xiang et al., 1996).<br />

Neurotrophic factors (i.e. nerve growth factor, brain-derived neurotrophic factor,<br />

neurotrophins-3 and 4/5, basic fibroblast growth factor, ciliary neurotrophic factor, and insulinlike<br />

growth factors) support neuronal differential and survival during development (Barde, 1989)<br />

and may protect against select forms of neuronal injury (Lindsay et al., 1994; Mattson and<br />

Scheff, 1994; Mocchetti and Wrathall, 1995). A variety of neurotrophic factors have been<br />

reported to protect against injury resulting from physical, ischemic, excitotoxic, metabolic, and<br />

oxidative insults (Mattson and Scheff, 1994; Mocchetti and Wrathall, 1995). <strong>The</strong><br />

neuroprotective function of these factors is proposed to involve enhanced maintenance of Ca 2+<br />

homeostasis and free radical metabolism (Mattson and Scheff, 1994).<br />

Several studies have examined the involvement of two members of the neurotrophin family,<br />

brain-derived neurotrophic factor (BDNF) and neurotrophin-3 (NT-3), in relation to KA<br />

excitotoxicity in hippocampus. Intrahippocampal (Ballarin et al., 1991) or systemic (Dugich-<br />

Djordjevic et al., 1992; Lee et al., 1997) KA administration dramatically increases hippocampal<br />

BDNF mRNA levels. Levels of BDNF mRNA remain elevated for as long as two weeks<br />

following intracranial ventricular injection of KA (Garcia et al., 1997a). Increased BDNF<br />

protein levels have also been demonstrated following systemic KA administration (Rudge et al.,<br />

24


1998; Katoh-Semba et al., 1999). In contrast, NT-3 mRNA levels are decreased following<br />

systemic (Lee et al., 1997; Katoh-Semba et al., 1999) or intracranial ventricular (Garcia et al.,<br />

1997a) KA administration, with decreases lasting up to two weeks post-injection in the latter<br />

model (Garcia et al., 1997a). NT-3 protein levels correspond with altered mRNA levels,<br />

decreasing following systemic KA injection (Katoh-Semba et al., 1999). Experiments indicate<br />

that BDNF and NT-3 protect cultured rat hippocampal, septal, and cortical neurons against<br />

damage induced by glucose deprivation or glutamate neurotoxicity by attenuating harmful rises<br />

in intracellular Ca 2+ levels (Cheng and Mattson, 1994). BDNF also protects cultured rat<br />

cerebellar granule neurons against glutamate neurotoxicity (Lindholm et al., 1993). In contrast, a<br />

study examining the effects of in vivo intrahippocampal neurotrophin administration reports no<br />

neuronal protection against KA toxicity with NT-3 infusion and exacerbation of KA-induced<br />

injury in CA3 with BDNF administration (Rudge et al., 1998).<br />

Controlled cortical impact as a model of traumatic brain injury<br />

Hypothalamic-pituitary-adrenal abnormalities are prevalent in human cases of head injury<br />

(King et al., 1970; Pentelenyi and Kammerer, 1984; Chiolero and Berger, 1994; Lieberman et al.,<br />

2001), indicating the urgent need for research examining the physiological effects of<br />

glucocorticoid secretion after traumatic brain injury (TBI). Synthetic glucocorticoid steroids<br />

such as methylprednisolone have been extensively used in the clinical treatment of central<br />

nervous system trauma. In 1990, the American National Acute Spinal Cord Injury Study<br />

demonstrated significant benefits of high-dose methylprednisolone in the treatment of human<br />

spinal cord injury (Bracken et al., 1990). However, the utility of these compounds with respect<br />

to brain injury requires further evaluation (see McIntosh et al., 1998b). Not surprisingly,<br />

neuroprotection conferred by methylprednisolone is thought to be independent of hormonal<br />

(receptor-mediated) glucocorticoid activities; rather, the primary neuroprotective mechanism of<br />

methylprednisolone is hypothesized to be the inhibition of oxygen free-radical induced lipid<br />

peroxidation via direct membrane effects (Hall et al., 1992). While the nonhormonal effects of<br />

exogenous synthetic steroids have been researched extensively, the physiological role for<br />

endogenous glucocorticoids after TBI remains largely unexamined. Increased cortisol levels<br />

reported with human TBI (King et al., 1970; Pentelenyi and Kammerer, 1984) emphasize the<br />

25


potential importance of receptor-mediated glucocorticoid effects on neuronal viability after<br />

trauma.<br />

Many animal models of mechanical brain injury have been developed in an attempt to<br />

reproduce aspects of human closed-head injury. Of these, several commonly used techniques<br />

are: fluid percussion, which produces brain injury by the rapid injection of fluid into the closed<br />

cranial cavity (Dixon et al., 1987); closed head injury, in which a weight-drop device delivers a<br />

standard blow to the skull (Shapira et al., 1988); and the cortical impact model, which uses a<br />

pneumatic impactor to deliver a controlled impact to the exposed dura, with measurable velocity<br />

and cortical deformation (Dixon et al., 1991). Experiments in the present thesis employ the<br />

controlled cortical impact (CCI) model of TBI in conjunction with ACR antagonist treatments in<br />

order to study ACR effects on hippocampal neuron viability following TBI.<br />

Controlled cortical impact pathology<br />

Traumatic injury of the human brain causes neuronal loss in regions including cortex,<br />

hippocampus, thalamus, and cerebellum (Adams et al., 1985; Kotapka et al., 1992; Ross et al.,<br />

1993). Neuronal degeneration caused by TBI is the product of direct mechanical insult (primary<br />

injury) and delayed pathological events including ion disequilibration, free radical formation,<br />

membrane damage, protease activation, and cytoskeletal disruption (secondary injury) (Faden,<br />

1993). <strong>The</strong> delayed component of TBI pathology is of great significance as it provides<br />

opportunity for therapeutic intervention after injury. <strong>The</strong> pattern of neuronal loss in<br />

experimentally injured rat brain mimics that of human TBI, with neuronal degeneration observed<br />

in cortical areas overlying the evolving contusion, in hippocampus, and in lateral thalamus<br />

(Sutton et al., 1993; Dietrich et al., 1994; Hicks et al., 1996). Dark, shrunken neurons and<br />

swollen astrocytes can be detected in the acute phase, as early as ten minutes following injury<br />

(Hicks et al., 1996). With the controlled cortical impact model at moderate severity, neuronal<br />

loss in CA3 is evident by one hour after injury (Baldwin et al., 1997). Furthermore, there is no<br />

significant difference in the percentage of surviving CA3 neurons estimated 24 hours or 30 days<br />

post-injury, indicating that neuronal loss is rapid and complete by 24 hours (Baldwin et al.,<br />

1997). Moderate levels of experimental TBI result in neuronal death by both necrosis and<br />

apoptosis, with the greatest loss caused by necrosis (Raghupathi et al., 2000; Zipfel et al., 2000).<br />

26


Secondary neuronal degeneration following traumatic insult is driven by an array of<br />

physiological and biochemical changes. Excessive glutamate release (Faden et al., 1989;<br />

Bullock and Fujisawa, 1992), increased intracellular Ca 2+ levels (Young, 1992; Nilsson et al.,<br />

1993; McIntosh et al., 1998c), mitochondrial dysfunction (Sullivan et al., 1998; Sullivan et al.,<br />

1999), metabolic changes (Hovda et al., 1992), tissue edema (Wahl et al., 1988), lipid<br />

peroxidation (Braughler and Hall, 1992), free radical formation (Lewen et al., 2000), vascular<br />

disruption (Cortez et al., 1989), neurochemical alterations (Hayes et al., 1992), inflammation<br />

(Feuerstein et al., 1997), astroglial responses (Cortez et al., 1989), axonal injury (Povlishock and<br />

Christman, 1995), proteolysis (Kampfl et al., 1997), activation of cell death pathways (Beer et<br />

al., 2000; Raghupathi et al., 2000), and altered gene expression (Mattson and Scheff, 1994;<br />

Hayes et al., 1995; Raghupathi et al., 2000) are among the many factors identified as contributors<br />

to trauma-induced degeneration (see Faden, 1993).<br />

Glucocorticoid modulation of traumatic brain injury<br />

Many factors contributing to secondary neuronal degeneration following TBI are likely<br />

influenced by glucocorticoids. Glucocorticoids reduce edema (Xu et al., 1992), influence<br />

neurotransmitter systems (Meijer and de Kloet, 1998), reduce inflammation (De Bosscher et al.,<br />

2000), and interact with cell death pathways (Reagan and McEwen, 1997). Steroid hormones<br />

also prevent lipid hydrolysis and the subsequent formation of oxygen free radicals, as noted<br />

above (Hall et al., 1992). Furthermore, glucocorticoids exacerbate cognitive deficits associated<br />

with experimental TBI (Hicks et al., 1993; White-Gbadebo and Hamm, 1993; Smith et al., 1994;<br />

Scheff et al., 1997). Experiments in the current thesis focus on glucocorticoid and traumainduced<br />

regulation of viability-related genes, specifically members of the bcl-2 family and<br />

tumor-suppressor p53.<br />

Regulators of cell survival and death<br />

Bcl-2 family members are critical regulators of cell viability, and an increasing body of<br />

evidence suggests that bcl-2 related proteins control neuronal survival following TBI. Bcl-2<br />

expression increases in injured human brain (Clark et al., 1999) and in rat hippocampal and<br />

cortical neurons surviving cortical impact with secondary hypoxia (Clark et al., 1997). In<br />

contrast, bcl-2 expression is reported to decrease in rat cortical neurons following closed head<br />

27


injury (Lu et al., 2000). A neuroprotective role for Bcl-2 after TBI has been supported by studies<br />

demonstrating reduced neuronal loss in injured cortex and hippocampus of Bcl-2 overexpressing<br />

mice (Raghupathi et al., 1998; Nakamura et al., 1999). Bax may also participate in the<br />

regulation of TBI-induced neuronal death, as bax translocation to nuclei of injured cortical and<br />

hippocampal neurons has been reported forty-eight hours after cortical contusion (Kaya et al.,<br />

1999).<br />

P53, a regulator of bcl-2 and bax expression (Miyashita et al., 1994), is also associated with<br />

experimental TBI in the rat. Induction of p53 mRNA has been observed in cortex, hippocampus,<br />

and thalamus following lateral fluid-percussion injury (Napieralski et al., 1999), and<br />

translocation of p53 proteins to nuclei of injured neurons has been reported using the controlled<br />

cortical impact model (Kaya et al., 1999). Neuronal apoptosis in thalamus is attenuated in p53 -<br />

/- mice (Martin et al., 2001), but a study measuring cortical lesion volume after controlled<br />

cortical impact in p53 -/- mice reported no significant differences compared to wild-type<br />

(Tomasevic et al., 1999).<br />

Several studies suggest that neurotrophic factors may also influence hippocampal neuron<br />

survival following TBI. Nerve growth factor mRNA levels are increased in dentate gyrus after<br />

injury by lateral fluid percussion (Truettner et al., 1999), while NT-3 mRNA levels are decreased<br />

(Hicks et al., 1997). In addition, increased hippocampal BDNF mRNA levels are observed<br />

following injury with CCI (Yang et al., 1996) and lateral fluid percussion (Hicks et al., 1997;<br />

Truettner et al., 1999) models of TBI, suggesting the potential for enhanced neuronal survival<br />

due to altered neurotrophin expression. In support of a neuroprotective role for BDNF<br />

expression after fluid percussion injury, BDNF mRNA levels are decreased in cortical regions<br />

containing degenerating neurons, generally unchanged in adjacent areas, and increased in more<br />

remote regions such as piriform cortex (Hicks et al., 1999). However, in a recent study<br />

intrahippocampal infusion of BDNF did not improve measures of neuronal loss, neuromotor<br />

function, learning, or memory following lateral fluid percussion injury (Blaha et al., 2000).<br />

28


Chapter Two: Mineralocorticoid receptors regulate bcl-2 and p53 mRNA<br />

expression in hippocampus<br />

Introduction<br />

Glucocorticoids regulate gene expression in hippocampal neurons through mineralocorticoid<br />

receptor (MR) and glucocorticoid receptor (GR) proteins. MR binds glucocorticoids with high<br />

affinity but low capacity, whereas GR binds with low affinity and high capacity (Reul and de<br />

Kloet, 1985). <strong>The</strong>se differences in binding efficacy allow MR and GR to regulate gene<br />

expression in a concentration-dependent manner; MR is predominantly active at low<br />

concentrations of glucocorticoids, while GR becomes increasingly bound and dominant at higher<br />

glucocorticoid concentrations (Reul and de Kloet, 1985). MR activity at low glucocorticoid<br />

levels is associated with increased cell excitability (Joels and de Kloet, 1990), long-term<br />

potentiation (Pavlides et al., 1995), and dentate gyrus (DG) granule cell survival (Woolley et al.,<br />

1991). In contrast, high levels of glucocorticoids acting mainly through GR have been shown to<br />

reduce long-term potentiation (Pavlides et al., 1993), shrink hippocampal pyramidal cell<br />

dendrites (Woolley et al., 1990), and exacerbate damaging effects of toxic insults (Sapolsky et<br />

al., 1988; Packan and Sapolsky, 1990). Thus, MR and GR seem to promote differential cellular<br />

effects; MR appears to enhance and GR appears to compromise neuronal viability.<br />

<strong>The</strong> influence of MR and GR on survival of hippocampal neurons is likely exerted at the<br />

level of gene regulation. To address the role of the MR in this regulation, we tested the effect of<br />

spironolactone (SPIRO), a specific, competitive MR antagonist (Sutanto and de Kloet, 1991), on<br />

basal and kainic acid (KA) stimulated expression of mRNAs associated with apoptosis or<br />

neuronal survival. Messenger RNAs include those that encode: 1) bcl-2, an inhibitor of<br />

apoptosis which is down-regulated by KA (Gillardon et al., 1995); 2) bax, a death-inducing<br />

member of the bcl-2 gene family which is up-regulated by KA (Gillardon et al., 1995); 3) p53, a<br />

tumor-suppressor which is induced by KA-related DNA damage (Sakhi et al., 1994); 4) brainderived<br />

neurotrophic factor (BDNF), a neurotrophin which is induced by KA excitotoxicity<br />

(Ballarin et al., 1991); and 5) neurotrophin-3 (NT-3). To investigate the hypothesis that MR<br />

promotes neuroprotective gene activity, we evaluated the effects of MR blockade and KA<br />

29


treatment by in situ hybridization analysis of mRNA expression in CA1, CA3, and DG of the rat<br />

hippocampal formation.<br />

Materials and Methods<br />

Animal treatment and tissue preparation<br />

All animal handling and protocols were approved by the University of Kentucky IACUC.<br />

Young adult male Sprague-Dawley rats were subcutaneously injected twice daily, in morning<br />

and afternoon, with MR antagonist spironolactone (50 mg/kg) or sesame oil vehicle. Treatment<br />

continued for three days until a total of five injections were administered. On day three, two<br />

hours after the final morning injection, half of each group received intraperitoneal injection of<br />

kainic acid (12 mg/kg), while the remaining half received 0.9% saline vehicle. Rats were<br />

monitored for 8 hours following KA injections; only animals exhibiting behavioral changes<br />

indicative of KA seizures as described previously (Sperk et al., 1983) were included in the study.<br />

Animals were decapitated on day four, twenty-four hours after onset of kainic acid-induced<br />

seizure activity. Brains were rapidly removed and frozen in isopentane (Sigma) cooled to -30 o<br />

to -40 o C on dry ice. Tissue was sectioned (16 um) with a Bright-Hacker cryostat, thaw-mounted<br />

onto Ultrastick Slides (Becton Dickinson), and stored at -20 o C.<br />

In situ hybridization<br />

Assessment of mRNA expression was accomplished using probes specific for 1) bcl-2, 2)<br />

bax, 3) p53, 4) BDNF, and 5) NT-3. Probes were labeled by standard in vitro transcription,<br />

using 33 P-UTP, according to previously published methods (Seroogy and Herman, 1996).<br />

Frozen tissue sections removed from a -20 o C freezer were placed in 4% paraformaldehyde for<br />

10 min. Slides were rinsed with 5mM KPBS, 5mM KPBS with 0.2% glycine, and again with<br />

5mM KPBS. Following 10 min in 0.1 M TEA with 0.25% acetic anhydride and rinses in 0.2X<br />

SSC, sections were dehydrated. Labeled probes were added to a hybridization buffer containing<br />

50% formamide. 50 uL (1 x 10 6 cpm) of diluted probe were applied to each slide. Slides were<br />

coverslipped, placed in moistened chambers, and incubated overnight at 55 o C. Following<br />

hybridization, coverslips were removed in 2X SSC, rinsed once in fresh 2X SSC, rinsed 3X in<br />

30


0.2X SSC for 10 min. per wash, followed by a 1 hour wash in 0.2X SSC at 60 o C. After<br />

dehydration, slides were exposed for 14 - 21 days to Kodak BioMAX film.<br />

Data analysis<br />

Densitometric measurements were taken from subfields CA1, CA3, and dentate gyrus using<br />

NIH Image 1.55 software for Macintosh. Data were analyzed by two-way ANOVA, with<br />

planned comparisons used to distinguish significance of KA and SPIRO effects. <strong>The</strong><br />

relationship between bax and p53 mRNAs was assessed by regression analysis.<br />

Results<br />

Figure 2.1 illustrates expression of bcl-2, bax, and p53 in hippocampus of saline- and KAtreated<br />

rats. Bcl-2 mRNA levels were significantly affected by SPIRO (F(1, 25) = 4.305; p <<br />

0.05) and KA (F(1, 25) = 5.845; p < 0.05) in CA1 and by KA in CA3 (F(1,25) = 6.468; p <<br />

0.05). Planned comparisons indicate that SPIRO decreased basal bcl-2 mRNA levels in CA1 and<br />

CA3 of saline-treated subjects (p < 0.05). Kainic acid significantly decreased bcl-2 mRNA<br />

levels in CA1 (p < 0.05) and CA3 (p < 0.01) of oil-treated subjects (Figure 2.2).<br />

Kainic acid treatment significantly influenced bax mRNA expression in CA1 (F(1, 12) =<br />

5.555; p < 0.05) and CA3 (F(1, 24) = 8.049; p < 0.01) (Figure 2.3). Planned comparisons<br />

revealed that following KA injection, bax mRNA expression significantly increased in CA1 (p =<br />

0.05) of SPIRO-treated animals and in CA3 (p < 0.05) of oil-treated animals. MR blockade had<br />

no significant effect on bax mRNA expression.<br />

Spironolactone (F(1,20) = 4.481; p < 0.05) and KA treatment (F(1, 20) = 15.819; p < 0.001)<br />

elicited significant main effects on p53 mRNA expression in CA3, the only hippocampal region<br />

in which p53 was observed in consistently quantifiable amounts (Figure 2.4). Kainic acid<br />

significantly increased p53 mRNA levels in CA3 of oil-treated animals (p < 0.001), while<br />

SPIRO pre-treatment significantly attenuated this increase (p < 0.05).<br />

31


Figure 2.1. Photomicrographs illustrating bcl-2 (A, B), bax (C, D), and p53 (E, F) mRNA<br />

distribution in hippocampus of saline- (A, C, E) and KA-treated (B, D, F) rats. KA decreased<br />

bcl-2 mRNA expression in CA1 and CA3, increased bax mRNA expression in CA3, and<br />

increased p53 mRNA expression in CA3.<br />

32


Figure 2.2. Effects of MR blockade and kainic acid on bcl-2 mRNA expression in hippocampus.<br />

MR blockade with SPIRO significantly decreased basal bcl-2 mRNA expression in CA1 and<br />

CA3 of saline-treated animals (*p < 0.05). KA decreased bcl-2 mRNA expression in CA1 and<br />

CA3 of oil-treated animals ( ‡ p < 0.05).<br />

33


Figure 2.3. Effects of MR blockade and kainic acid on bax mRNA expression in hippocampus.<br />

Kainic acid (KA) significantly increased bax mRNA expression in CA1 of SPIRO-treated<br />

animals and in CA3 of oil-treated animals ( ‡ p < 0.05).<br />

34


Figure 2.4. Effects of MR blockade and kainic acid on p53 mRNA expression in hippocampus.<br />

Kainic acid significantly increased p53 mRNA levels in CA3 ( ‡ p < 0.05). MR blockade with<br />

SPIRO significantly attenuated the KA-induced increase (*p < 0.05).<br />

Kainic acid treatment significantly altered BDNF mRNA levels in CA1 (F(1, 23) = 26.952; p<br />

< 0.0001) and CA3 (F(1, 23) = 13.511; p < 0.01); expression increased significantly (p < 0.05) in<br />

CA1 and CA3 of both oil- and SPIRO-treated animals (Figure 2.5). NT-3 mRNA levels were<br />

affected by KA in DG (F(1, 24) = 107.00; p < 0.0001), showing significant decreases in<br />

expression in both oil- and SPIRO-treated animals (p < 0.0001) (Figure 2.6). MR blockade had<br />

no significant effect on BDNF or NT-3 mRNAs.<br />

35


Figure 2.5. Effects of MR blockade and kainic acid on brain derived neurotrophic factor<br />

(BDNF) expression in hippocampus. Kainic acid significantly increased BDNF mRNA<br />

expression in CA1 and CA3 ( ‡ p < 0.05).<br />

36


Figure 2.6. Effects of MR blockade and kainic acid on neurotrophin-3 (NT-3) expression in<br />

hippocampus. KA decreased NT-3 mRNA expression in DG ( ‡ p < 0.05).<br />

37


Discussion<br />

Within the bcl-2 gene family, dimerization among death-repressing members, such as Bcl-2,<br />

and death-inducing members, such as Bax, is thought to determine cellular response to apoptotic<br />

signals. <strong>The</strong> ratio of life-promoting to death-promoting proteins determines if dimerization<br />

favoring cell survival will occur (Oltvai et al., 1993). We report that MR blockade by SPIRO<br />

decreases basal bcl-2 mRNA expression in CA1 and CA3, demonstrating that MR activity is<br />

essential in maintaining normal levels of bcl-2 mRNA in these regions. This finding supports the<br />

hypothesis that MR functions as a neuroprotective factor, as decreases in bcl-2 mRNA could lead<br />

to the alteration of cellular Bcl-2/Bax ratios and predispose cells to apoptotic cell death. We also<br />

report a decrease in bcl-2 mRNA expression in CA1 and CA3 of oil-treated animals with KA<br />

administration, an effect previously demonstrated in mouse hippocampus (Gillardon et al.,<br />

1995). Bcl-2 mRNA levels were similarly decreased in SPIRO and oil-treated animals after KAtreatment.<br />

Thus, whereas SPIRO pretreatment decreased bcl-2 mRNA levels in control rats, it<br />

did not exacerbate decreases in bcl-2 mRNA expression induced by KA treatment. Overall, the<br />

data suggest that MR modulates bcl-2 expression in hippocampus and may therefore play a role<br />

in pyramidal cell survival.<br />

<strong>The</strong> observed induction of bax mRNA expression in CA1 and CA3 following KA treatment<br />

is consistent with previous studies (Gillardon et al., 1995). Notably, levels of bax mRNA did not<br />

differ significantly between SPIRO or oil-treated groups prior to or following KA administration.<br />

In combination with decreased bcl-2 mRNA expression, these results suggest a possible reduced<br />

Bcl-2/Bax ratio, a scenario believed to favor the neuroendangering actions of Bax (Oltvai et al.,<br />

1993).<br />

In agreement with previous studies (Sakhi et al., 1994), we demonstrated that KA treatment<br />

increases p53 mRNA levels in rat hippocampus. We additionally report that SPIRO attenuated<br />

this KA-induced increase in CA3, indicating MR involvement in p53 mRNA expression under<br />

these conditions. As accumulation of p53 is associated with induction of apoptosis (Yonish-<br />

Rouach et al., 1991; Sakhi et al., 1996), these data suggest that, contrary in part to our principal<br />

hypothesis, the MR may play a role in regulating both protective and endangering gene products.<br />

Bax mRNA responses parallel those of p53 mRNA in this study. SPIRO appeared to<br />

attenuate the KA-induced increase of bax mRNA in CA3 and also significantly reduced the KA<br />

38


induction of p53 mRNA in the same region. <strong>The</strong>se changes in bax and p53 mRNA levels after<br />

KA treatment are significantly correlated (r = 0.74; p < 0.05), indicating that the expression of<br />

bax and p53 mRNAs are regulated in parallel. P53 is capable of inducing the human bax gene<br />

(Miyashita and Reed, 1995) and is thought to participate in a signaling pathway involving bax<br />

and bcl-2 that predisposes cells to apoptosis (Miyashita et al., 1994). For these reasons, the<br />

similar responses observed in this study invite speculation that increased bax mRNA levels may<br />

result from upregulation by p53. However, p53 overexposure has been shown to cause apoptosis<br />

in hippocampal pyramidal neurons without bax induction (Jordan et al., 1997), so the possibility<br />

remains that the increases in bax and p53 mRNA are not causally linked.<br />

Consistent with previous reports (Ballarin et al., 1991), this study demonstrated that BDNF<br />

mRNA levels increase in rat hippocampus following KA treatment. As elevated glucocorticoid<br />

levels have been shown to decrease the expression of BDNF mRNA (Schaaf et al., 1997), this<br />

induction is most likely a consequence of another aspect of KA excitotoxicity, although possibly<br />

influenced by glucocorticoid input. In addition, NT-3 mRNA expression was decreased in DG<br />

of animals receiving KA. While neurotrophins are clearly susceptible to glucocorticoid<br />

regulation, as demonstrated by previous studies (Chao and McEwen, 1994; Smith et al., 1995a),<br />

there does not appear to be any differential regulation of BDNF or NT-3 mRNA with MR<br />

blockade prior to or following excitotoxic insult with KA.<br />

It should be noted that blockade of the MR may alter mRNA expression via activity of an<br />

unopposed GR. Without the MR available to form MR-GR heterodimers and otherwise compete<br />

with GR homodimers at the GC response element, an increase in GR-mediated transactivation<br />

may underlie changes in transcriptional regulation. Overall, it is clear that MR blockade affects<br />

bcl-2 and p53 mRNA levels, but whether this change is due to a loss of MR or increase in GR<br />

activity remains to be determined. Another important consideration is the substantial<br />

hippocampal pyramidal cell loss that accompanies KA treatment after 24 hours. It is possible<br />

that changes observed in bcl-2, bax, and p53 mRNAs could represent non-specific actions of<br />

dying cells, rather than specific regulatory events, or possibly reflect changes in surviving cell<br />

numbers.<br />

Our current understanding of the function of GC receptors and cell viability is far from<br />

complete. <strong>The</strong> MR likely participates in apoptosis regulation, but the present evidence does not<br />

convincingly support an exclusively neuroprotective role. One problematic issue is the<br />

39


complexity of receptor interaction involved. It is unclear at this time whether changes in mRNA<br />

expression following MR blockade are attributable to an absence of MR activity or an<br />

exaggerated presence of GR. Based on the available data, it appears that MR promotes<br />

protective gene induction under normal conditions, which may have important implications for<br />

survival of pyramidal cells following injury.<br />

40


Chapter Three: Adrenocorticosteriod receptor blockade and excitotoxic<br />

challenge regulate adrenocorticosteroid receptor mRNA levels in<br />

hippocampus<br />

Introduction<br />

Glucocorticoids are steroid hormones that convey hypothalamic-pituitary-adrenal (HPA) axis<br />

information throughout the body. Two types of adrenocorticosteroid receptors (ACRs) mediate<br />

glucocorticoid action in brain: the mineralocorticoid receptor (MR) and the glucocorticoid<br />

receptor (GR). Once the receptors bind glucocorticoid ligand, cortisol in humans and<br />

corticosterone (CORT) in rodents, MR and GR act as transcription factors (Yamamoto, 1985) to<br />

modulate genes controlling many physiological processes including development, metabolism,<br />

memory, inflammation, immune function, and cell death (Munck et al., 1984; McEwen et al.,<br />

1986).<br />

Because the MR has greater affinity for binding CORT and the GR greater capacity, the MR<br />

is mainly active at basal glucocorticoid concentrations while the GR is extensively bound during<br />

periods of elevated glucocorticoid levels (Reul and de Kloet, 1985; Reul et al., 1987).<br />

Predominant activation of the MR or GR at different CORT concentrations appears to have<br />

opposing effects on neuronal viability in hippocampus, an area of MR and GR colocalization<br />

(Reul and de Kloet, 1985; Herman et al., 1989a) that is particularly susceptible to glucocorticoidmediated<br />

effects. For example, activation of the MR protects hippocampal dentate gyrus (DG)<br />

granule cells from adrenalectomy (ADX) induced death (Woolley et al., 1991). In contrast,<br />

prolonged elevation of glucocorticoid levels, acting mainly through the GR, shrinks pyramidal<br />

cell dendrites (Woolley et al., 1990), impairs calcium regulation (Elliott and Sapolsky, 1993),<br />

and exacerbates neuronal insults (Sapolsky, 1985; Sapolsky and Pulsinelli, 1985; Sapolsky,<br />

1986a; Sapolsky et al., 1988; Goodman et al., 1996; McIntosh and Sapolsky, 1996a). Thus,<br />

evidence exists that the MR protects and the GR endangers hippocampal neurons.<br />

Considering the differential effects of MR and GR on hippocampal cell survival, ACR<br />

activation may be of critical importance in the cellular response to injury. To address this issue,<br />

the present study is designed to investigate the roles of the MR and the GR in regulating ACR<br />

expression and neuron viability under basal conditions and following excitotoxic insult. Subjects<br />

41


were pretreated with vehicle, the MR antagonist spironolactone (SPIRO), or the GR antagonist<br />

mifepristone (RU486) and subsequently challenged with systemic injection of kainic acid (KA).<br />

In situ hybridization was performed to assess the effects of ACR blockade and subsequent<br />

challenge with KA on relative MR and GR mRNA levels. In addition, hippocampal neuronal<br />

loss was evaluated following MR or GR blockade and KA insult to determine the role of ACRs<br />

in hippocampal vulnerability to injury in vivo. <strong>The</strong> results of this study provide further evidence<br />

for homo- and heterologous regulation of ACR expression and indicate that ACR mRNA<br />

regulation is altered by glutamate receptor stimulation.<br />

Materials and Methods<br />

Subjects<br />

Young adult male Sprague Dawley rats (Harlan Sprague Dawley, Inc., Indianapolis, IN)<br />

weighing between 200-250 gm were maintained on a 12 hr light/dark cycle in an environment<br />

with constant temperature, humidity, and ad libitum access to food and water. All animal<br />

procedures were performed in accordance with the National Institutes of Health guidelines and<br />

approved by the University of Kentucky Institutional Animal Care and Use Committee.<br />

In vivo protocols<br />

All rats were injected s.c. twice daily, in morning and afternoon, with propylene glycol<br />

vehicle, MR antagonist spironolactone (SPIRO) (50 mg/kg) (Herman and Spencer, 1998)<br />

(Sigma, St. Louis, MO), or GR antagonist RU486 (25 mg/kg) (Ratka et al., 1989) (Sigma).<br />

Subjects were administered a final pretreatment injection at 900h on day three, followed by an<br />

i.p. injection of either saline vehicle or kainic acid (KA) (12 mg/kg)(Sigma) at 1100h. Subjects<br />

were monitored for seizure activity for 6h following KA injections, and only those displaying<br />

behavioral changes indicative of KA intoxication (Sperk et al., 1983) were included in the study.<br />

All animals were sacrificed on day 4 at 1200h, approximately 24h after seizure onset.<br />

Animals processed for plasma CORT determination and in situ hybridization experiments<br />

were killed by rapid decapitation. Brains were rapidly removed, frozen in isopentane cooled to –<br />

40 o C on dry ice, then stored at –80 o C. Brains were later warmed to -20 o C, sectioned to 16 um<br />

with a Bright-Hacker cryostat, thaw-mounted onto Ultrastick Slides (Becton Dickinson, Franklin<br />

42


Lakes, NJ), and stored at -20 o C until processing for in situ hybridization. Trunk blood samples<br />

were collected into heparinized tubes on ice, and plasma was separated by centrifugation and<br />

stored at –20 o C.<br />

Animals processed for cell counts received a lethal dose of sodium pentobarbitol (Butler,<br />

Columbus, OH)(150mg/kg) before undergoing transcardial perfusion with 0.9% saline followed<br />

by 4% paraformaldehyde. Brains were removed, post-fixed in 4% paraformaldehyde overnight,<br />

then placed in 30% sucrose at 4 o C for 24h. Brains were then stored in 70% ethanol until<br />

embedded in paraffin.<br />

Radioimmunoassay<br />

Radioimmunoassay was performed to measure plasma CORT concentration. A doubleantibody<br />

rat corticosterone kit (ICN Biochemicals, Costa Mesa, CA) was used to determine 125 I-<br />

labeled CORT levels. Due to a skewed distribution of CORT values, we performed a log<br />

transformation of the data before analysis by two-way ANOVA. Data shown in Table 3.1<br />

represent actual CORT values listed as mean ± standard error. Two-way ANOVA was used to<br />

analyze log-transformed CORT values for main effects (of antagonist groups and KA treatment)<br />

and their interactions. Pairwise comparisons were made to determine significance of individual<br />

group effects within treatments and vice versa using the Student-Newman-Keuls test, with<br />

significance set at p ≤ 0.05.<br />

In situ hybridization<br />

Antisense riboprobes specific for rat MR (550 bp, complementary to the coding region and 3’<br />

untranslated region of rat MR mRNA) and rat GR (456 bp, complementary to the coding region<br />

and 3’ untranslated region of rat GR mRNA) were labeled according to previously published<br />

methods (Seroogy and Herman, 1996) by standard in vitro transcription using [ 35 S]UTP. Tissue<br />

sections removed from a –20 o C freezer were fixed in 4% paraformaldehyde for 10 min. Slides<br />

were then rinsed sequentially in the following solutions: 5 mM KPBS, 5 mM KPBS with 0.2%<br />

glycine, 5mM KPBS, 0.1 M triethanolamine with 0.25% acetic anhydride, 0.2X SSC, and graded<br />

alcohols. Labeled probes were added to a hybridization buffer containing 50% formamide, and<br />

diluted probe (50 ul containing 1x 10 6 c.p.m.) was applied to each slide. Slides were<br />

coverslipped, placed in moistened chambers, and incubated overnight at 55 o C. Following<br />

43


hybridization, slides were rinsed in 2X SSC and incubated in 0.2X SSC at 60 o C for 1 hr. Slides<br />

were then dehydrated and exposed for 14 days to Kodak BioMAX xray film.<br />

Autoradiographic images were digitized with NIH Image 1.62 software for Macintosh. Based<br />

on morphological criteria, regions CA1, CA3, and DG of the hippocampal formation were<br />

manually sampled for densitometric analysis. <strong>The</strong> mean gray level of an unhybridized<br />

(background) area from each section was subtracted from sampled regions to arrive at corrected<br />

gray level. <strong>The</strong> data were then expressed as a percentage of control (vehicle/saline) and analyzed<br />

by two-way ANOVA within each subregion. Pairwise comparisons were made using the<br />

Student-Neuman-Keuls test to determine significant effects of groups within treatments and vice<br />

versa, with significance set at p ≤ 0.05. Data in Figures 3.2 and 3.3 are presented as percentage<br />

of vehicle/saline control ± standard error.<br />

Cell counts<br />

Nissl-stained sections of perfused tissue were examined by light microscopy for undamaged<br />

neurons in the CA1 and CA3 pyramidal layers of dorsal hippocampus approximately 3mm<br />

posterior to bregma. Cell counts were not performed in dentate gyrus because no damaged<br />

granule cells were observed and because of the general insensitivity of this subregion to KA<br />

(Sperk et al., 1983). Neurons identified by round, healthy-looking somata and prominent<br />

nucleoli were counted by an observer blinded to the experimental design. In preparation for cell<br />

counts, perfused, paraffin-embedded, slide-mounted tissue was deparaffinized in xylene for 10<br />

min, hydrated through graded alcohols, and immersed in cresyl violet Nissl stain for 5 min.<br />

Slides were then rinsed briefly in dH 2 O, dehydrated through graded alcohols, cleared in xylene<br />

for 5 min, and coverslipped with DPX mountant.<br />

To obtain cell counts, numbers of neurons with identifiable nucleoli were recorded from a<br />

100 um 2 field. Counts were performed in three sections per subject, and within each section,<br />

three fields were counted from CA1 and from CA3 then averaged to calculate subregional means<br />

per section. Subregional means from each section were then averaged to determine mean<br />

numbers of live CA1 and CA3 neurons for each animal. Cell counts were analyzed for<br />

significant effects using two-way ANOVA. Post hoc comparisons were made using the Student-<br />

Newman-Keuls test, with significance set at p ≤ 0.05. Data in Table 3.2 are presented as mean<br />

cell number per 100 um 2 field ± standard error.<br />

44


Results<br />

Effects of ACR blockade and KA on CORT levels<br />

Two-way ANOVA indicates that 24h post-insult, KA significantly increased CORT levels<br />

[F(1,42) = 65.42; p ≤ 0.05) (Table 3.1). Post hoc comparisons indicate that KA-induced<br />

elevations in CORT levels were significant in all pretreatment groups. ACR antagonists also<br />

produced a significant main effect on CORT levels [F(2,42) = 5.96; p ≤ 0.05]. In saline-treated<br />

animals, MR blockade with SPIRO increased basal CORT levels, while GR antagonist RU486<br />

decreased basal CORT levels relative to vehicle-pretreated animals.<br />

Table 3.1. Effects of adrenocorticosteroid receptor blockade and kainic acid on plasma<br />

corticosterone levels (ng/ml)<br />

Group Treatment N Mean CORT Level<br />

VEHICLE SALINE 11 78.35 ± 20.89<br />

SPIRO SALINE 7 137.01 ± 23.24 a<br />

RU486 SALINE 6 23.93 ± 5.07 a<br />

VEHICLE KA 11 332.03 ± 59.42 b<br />

SPIRO KA 6 275.43 ± 53.35 b<br />

RU486 KA 7 325.85 ± 69.87 b<br />

SPIRO = MR antagonist spironolactone<br />

RU486 = GR antagonist mifepristone<br />

KA = excitotoxin kainic acid<br />

a p ≤ 0.05 as compared to vehicle/saline subjects (effect of antagonist)<br />

b p ≤ 0.05 as compared to saline subjects within the same pretreatment<br />

group (effect of kainic acid)<br />

45


Effects of ACR blockade and KA on MR and GR mRNA levels<br />

Relative changes in ACR mRNA levels were determined by film densitometry following in<br />

situ hybridization with MR and GR-specific riboprobes (see Methods, Figure 3.1).<br />

Figure 3.1. Photomicrographs illustrating the effects of saline (A,C) and kainic acid (B,D) on<br />

MR (A,B) and GR (C,D) mRNA levels in hippocampus of vehicle-pretreated rats. Panels A,B:<br />

Kainic acid decreases MR mRNA levels in CA1 and CA3. Panels C,D: Kainic acid decreases<br />

GR mRNA levels in dentate gyrus.<br />

46


Pretreatment with ACR antagonists significantly affected MR mRNA levels in hippocampal<br />

subregions CA1 [F(2,42) = 5.51; p ≤ 0.05], CA3 [F(2,42) = 6.58; p ≤ 0.05], and DG [F(2,42) =<br />

4.80; p ≤ 0.05] (Figure 3.2). MR antagonist SPIRO increased MR mRNA signal in CA1, CA3,<br />

and DG of saline-treated animals, while GR antagonist RU486 had no significant effects on basal<br />

MR mRNA levels. KA effects on MR mRNA levels were significant in CA1 [F(1,42) = 83.43; p<br />

≤ 0.05] and CA3 [F(1,42) = 60.62; p ≤ 0.05]. KA decreased MR mRNA levels in regions CA1<br />

and CA3 of all pretreatment groups.<br />

Figure 3.2. Effects of ACR blockade and KA treatments on MR mRNA levels in rat<br />

hippocampus. Semi-quantitative analysis expressed as percentage of control (vehicle/saline)<br />

corrected gray level. SPIRO significantly increased basal MR mRNA levels in CA1, CA3, and<br />

DG of saline-treated animals. KA significantly decreased MR mRNA levels in CA1 and CA3 of<br />

all pretreatment groups. [ a p ≤ 0.05 as compared to vehicle/saline animals (effect of antagonist<br />

group). b p ≤ 0.05 as compared to saline animals within same pretreatment group (effect of KA<br />

treatment).]<br />

47


Two-way ANOVA demonstrates a significant group/treatment interaction effect on GR<br />

mRNA levels in region CA3 [F(2,42) = 5.16; p ≤ 0.05] (Figure 3.3). SPIRO and RU486<br />

significantly increased basal GR mRNA in region CA3, but no significant interaction effects<br />

within post hoc comparisons were observed. Main effects of KA treatment on GR mRNA levels<br />

were significant in regions CA1 [F(1,42) = 7.27; p ≤ 0.05], CA3 [F(2,42) = 24.72; p ≤ 0.05], and<br />

DG [F(2,42) = 91.25; p ≤ 0.05]. KA induced significant decreases in GR mRNA signal in CA1<br />

of RU486 animals, in CA3 of SPIRO and RU486-treated animals, and in DG of all pretreatment<br />

groups.<br />

Figure 3.3. Effects of ACR blockade and KA treatments on GR mRNA levels in rat<br />

hippocampus. Semi-quantitative analysis expressed as percentage of control (vehicle/saline)<br />

corrected gray level. SPIRO and RU486 significantly increased basal GR mRNA levels in CA3<br />

of saline-treated animals. KA decreased GR mRNA levels in CA1 of the RU486 pretreatment<br />

group, in CA3 of SPIRO and RU486 pretreatment groups, and in DG of all groups. [ a p ≤ 0.05 as<br />

compared to vehicle/saline animals (effect of antagonist group). b p ≤ 0.05 as compared to saline<br />

animals within same pretreatment group (effect of KA treatment).]<br />

48


Effects of ACR blockade and KA on hippocampal neuron viability<br />

Neurons were counted to determine the effects of ACR antagonists on hippocampal neuron<br />

viability in CA1 and CA3 following KA challenge (see Methods) (Table 3.2). Neither ACR<br />

blockade nor KA significantly affected the number of CA1 pyramidal neurons. Kainic acid<br />

produced a significant main effect on the number of viable neurons in CA3 [F(1,22) = 6.35; p ≤<br />

0.05]. <strong>The</strong> excitotoxin decreased live cell numbers in CA3 of SPIRO-pretreated subjects, but<br />

post hoc comparisons indicate no significant effects of antagonist pretreatment.<br />

Table 3.2. Effects of adrenocorticosteroid receptor blockade and kainic acid<br />

on hippocampal neuron viability<br />

Region Group Treatment N<br />

Mean LIVE neurons<br />

per 100 um 2<br />

CA1 VEHICLE SALINE 4 63.54 ± 3.06<br />

SPIRO SALINE 4 62.46 ± 2.51<br />

RU486 SALINE 3 65.00 ± 2.25<br />

VEHICLE KA 8 60.79 ± 2.16<br />

SPIRO KA 4 47.25 ± 14.81<br />

RU486 KA 5 48.90 ± 8.36<br />

CA3 VEHICLE SALINE 4 39.67 ± 1.69<br />

SPIRO SALINE 4 34.63 ± 1.52<br />

RU486 SALINE 3 31.39 ± 6.54<br />

VEHICLE KA 8 30.38 ± 2.82<br />

SPIRO KA 4 22.60 ± 3.92 a<br />

RU486 KA 5 28.10 ± 5.20<br />

SPIRO = MR antagonist spironolactone<br />

RU486 = GR antagonist mifepristone<br />

KA = excitotoxin kainic acid<br />

a p ≤ 0.05 as compared to SPIRO/saline subjects (effect of kainic acid)<br />

49


Discussion<br />

<strong>The</strong> present study characterizes the effects of MR and GR antagonists on plasma CORT<br />

levels and hippocampal ACR mRNA regulation under basal conditions and after excitotoxic<br />

challenge with kainic acid. We report increases in basal MR mRNA levels in CA1, CA3, and<br />

DG following MR blockade and demonstrate increases in basal GR mRNA levels in CA3<br />

following both MR and GR blockade. Further, we report that administration of the ionotropic<br />

glutamate receptor agonist kainic acid reduces MR and GR mRNA levels in specific<br />

hippocampal subfields at 24h following insult, indicating differential regulation of ACR mRNAs<br />

during excitotoxic injury. Cell counts following KA administration provide evidence that the<br />

magnitude of ACR mRNA downregulation exceeds that of cell loss.<br />

Effects of ACR blockade and KA on CORT levels<br />

In this study, pretreatment with SPIRO significantly increased basal CORT levels (Table<br />

3.1). <strong>The</strong>re is evidence that MR blockade with antagonist RU28318 increases CORT levels in<br />

rat (Ratka et al., 1989; van Haarst et al., 1997), suggesting that the SPIRO-induced CORT<br />

increase is due to MR blockade and a resultant disinhibition of adrenocortical secretion.<br />

However, other studies using SPIRO have reported no significant effects on plasma CORT levels<br />

(Chabert et al., 1984; Herman and Spencer, 1998). <strong>The</strong> reason for this disagreement may be<br />

related to differing treatment protocols and durations.<br />

Glucocorticoid receptor blockade with RU486 decreased basal CORT levels in this study.<br />

This effect concurs with data demonstrating RU486-mediated suppression of CORT induction 15<br />

minutes following introduction to a novel environment (Ratka et al., 1989). Considering the<br />

relatively high CORT levels observed in our saline/vehicle-treated (control) subjects, the<br />

observed decrease in CORT levels may reflect RU486-mediated reduction of CORT levels<br />

subsequent to stress associated with either the euthanasia procedure or intraperitoneal injections<br />

received the previous day. Despite its efficacy as a GR antagonist, RU486 can also act as a<br />

partial GR agonist (Nordeen et al., 1995), which might explain a reduced stress response in<br />

saline animals. However, CORT response to KA is not attenuated in RU486-treated animals,<br />

suggesting that effects of RU486 on delayed negative feedback are minimal.<br />

50


Kainic acid induced a massive increase in CORT secretion observable 24h after injection.<br />

Acute effects of KA include hyperactivity, “wet dog shakes”, and prolonged tonic-clonic<br />

convulsions. Dehydration, brain edema, and perivenous hemorrhages are also common 24-48h<br />

after KA injection (Sperk et al., 1983). It is therefore likely that the excitotoxin is a potent<br />

stressor on multiple levels and thus increases CORT secretion for an extended period after<br />

injection.<br />

ACR antagonists modulate MR and GR mRNA levels<br />

Treatment with the MR antagonist SPIRO increased basal MR mRNA levels in all observed<br />

hippocampal subregions by approximately 30-40% (Figures 3.1 and 3.2). <strong>The</strong>se changes<br />

complement reports of MR mRNA upregulation by adrenalectomy (Herman et al., 1989a; Reul<br />

et al., 1989; Herman et al., 1993; Chao et al., 1998a) and provide additional evidence for<br />

homologous feedback regulation of MR. <strong>The</strong> MR also appears to regulate expression of GR.<br />

<strong>The</strong> present data demonstrate a significant SPIRO-induced increase in GR mRNA in CA3; GR<br />

mRNA levels are similarly elevated in CA1 and DG of SPIRO-pretreated subjects but do not<br />

reach our criteria for statistical significance. <strong>The</strong>se data add to previous results from our<br />

laboratory showing significant increases in GR mRNA levels in CA1 and DG with SPIRO<br />

pretreatment (Herman and Spencer, 1998). In addition, the MR agonist aldosterone decreases<br />

GR binding in primary hippocampal cell cultures (O'Donnell and Meaney, 1994), and<br />

administration of aldosterone or low, MR-activating doses of CORT reverses ADX-induced<br />

increases in GR mRNA (Chao et al., 1998a). Together, these results and the current data suggest<br />

that MR functions as a tonic inhibitor of GR expression in hippocampus.<br />

Preceding evidence also suggests that the activated glucocorticoid receptor negatively<br />

regulates ACR expression. MR mRNA levels decrease during periods of high CORT secretion<br />

(Herman et al., 1993) when the GR is most extensively activated (Reul and de Kloet, 1985).<br />

Further, glucocorticoid receptor agonist dexamethasone reverses ADX-induced increases in GR<br />

(Herman et al., 1989a; Reul et al., 1989) and MR (Herman et al., 1989a) mRNA levels and<br />

reduces GR mRNA expression in culture (Rosewicz et al., 1988). In the present experiment,<br />

RU486 significantly increased GR mRNA levels in CA3 (Figures 3.1 and 3.3), suggesting that<br />

glucocorticoid receptor blockade increases GR biosynthesis by homologous feedback<br />

mechanisms. However, RU486 did not increase MR mRNA levels, implying that glucocorticoid<br />

51


eceptor regulation of MR is limited to suppressing MR levels during periods of extensive GR<br />

activation.<br />

SPIRO and RU486 are commonly used to achieve ACR blockade and are among the most<br />

specific antagonists available, but like many other antagonist drugs they may produce some<br />

nonspecific effects. SPIRO is considered relatively specific for the MR but has been<br />

demonstrated a weak GR antagonist in a transformed murine fibroblast cell line (Couette et al.,<br />

1992). RU486 is an extremely potent GR antagonist in vitro, binding the GR with approximately<br />

18 times the affinity of CORT, but also has strong anti-progestin activity (Baulieu, 1989).<br />

RU486 may also act as a GR agonist under certain circumstances (Nordeen et al., 1995) and may<br />

not be sufficient to fully block GR activation in vivo. However, it should be noted that the<br />

currently reported antagonist effects correspond very well with the effects of adrenalectomy<br />

(Herman et al., 1989a; Reul et al., 1989; Herman et al., 1993; Chao et al., 1998a); the antagonist<br />

drugs increase hippocampal ACR mRNA expression in saline animals for the duration of the<br />

experiment.<br />

Excitotoxic challenge with KA decreases MR and GR mRNA levels<br />

<strong>The</strong> present data show dramatic reductions in hippocampal MR and GR mRNA levels<br />

following kainic acid insult (Figures 3.1, 3.2, and 3.3). KA-induced decreases in MR and GR<br />

binding have previously been demonstrated using whole dissected hippocampus from<br />

adrenalectomized rats (Lowy, 1992). Reductions in MR and GR binding are similar in<br />

magnitude to the currently reported mRNA decreases, are evident as early as one hour after<br />

systemic KA administration, and persist for at least 24 hours (Lowy, 1992). <strong>The</strong>refore, the<br />

current data provide further evidence that MR and GR availability is reduced by KA and indicate<br />

that this reduction is likely achieved through genomic downregulation, rather than by protein<br />

degradation or receptor inactivation.<br />

ACR downregulation by KA does not appear to be MR or GR-dependent. In the current<br />

study, KA reduces ACR mRNA expression in the presence of MR or GR blockade. In addition,<br />

a previous study has demonstrated that KA decreases ACR binding in adrenalectomized rats<br />

(Lowy, 1992); corticosterone depletion in adrenalectomized rats prevents ACR activation and<br />

their potential participation in the regulation of receptor expression. Furthermore, the pattern of<br />

ACR downregulation produced by KA is inconsistent with the pattern of regulation associated<br />

52


with high CORT levels. High CORT levels decrease MR and GR mRNA levels by equal<br />

proportions in CA1, CA3, and DG (Herman et al., 1993). In contrast, KA regulation of ACR<br />

expression differs substantially by hippocampal subfield. Mineralocorticoid receptor expression<br />

is downregulated in KA-sensitive regions CA1 and CA3, while GR expression is reduced in the<br />

relatively KA-insensitive DG. <strong>The</strong>se results suggest that KA regulation of ACR expression is<br />

not mediated by KA induction of CORT, but that differential ACR regulation by KA is achieved<br />

through other mechanisms related to KA susceptibility, such as glutamate receptor activation or<br />

Ca 2+ influx.<br />

Interestingly, increases in MR and GR mRNA that occur with ACR antagonist pretreatment<br />

are reversed by KA insult. Although KA does not affect MR mRNA levels in DG of vehiclepretreated<br />

animals, the SPIRO-induced increase in this region is negated by KA treatment.<br />

Similarly, KA does not reduce GR mRNA levels in the pyramidal layers of vehicle-pretreated<br />

animals, but SPIRO and RU486-induced increases in CA3 are abolished with KA injection.<br />

<strong>The</strong>se results suggest that homologous regulation of the receptors is suppressed or overridden by<br />

KA, functionally disconnecting autoregulatory feedback.<br />

Hippocampal pyramidal neurons are extremely sensitive to KA toxicity; CA3 neurons are<br />

particularly vulnerable, while CA1 neurons are slightly less susceptible, to KA-induced damage<br />

and cell degeneration (Nadler et al., 1978). <strong>The</strong> potential impact of cell loss on mRNA levels as<br />

measured by in situ hybridization is an important consideration that is presently undetermined.<br />

However, it is clear that in the hippocampal pyramidal layers, KA-induced decreases in MR<br />

mRNA levels exceed the proportion of neuronal loss indicated by cell counts – by approximately<br />

25 to 50% (see Figures 3.1, 3.2, and Table 3.2). Furthermore, DG granule cells are resistant to<br />

KA toxicity and exhibit no histological damage after systemic KA administration (Sperk et al.,<br />

1983). <strong>The</strong>refore, the approximately 40-55% reduction in GR mRNA levels observed in DG is<br />

not attributable to a decline in cell number.<br />

KA-induced repression of ACR mRNA levels might influence neuronal viability during<br />

injury. <strong>The</strong> receptors have potent cellular effects, some of which could endanger neurons under<br />

compromising conditions. For example, MR-modulated excitation of pyramidal neurons<br />

(Reiheld et al., 1984; Joels and de Kloet, 1990, 1992) could be permissive to excitotoxic damage.<br />

MR blockade with SPIRO has been shown to inhibit CORT-mediated increases in susceptibility<br />

to KA-induced convulsions in mice (Roberts and Keith, 1994a, b). Thus, downregulation of the<br />

53


MR in pyramidal cells may reduce excitability and vulnerability to excitotoxic challenge.<br />

Moreover, the GR has long been recognized as a potentially compromising influence on<br />

hippocampal neurons. GR activation has been shown to increase intracellular calcium levels in<br />

CA1 pyramidal cells (Kerr et al., 1992; Karst et al., 1994) and exacerbate the damaging effects of<br />

toxic insults including KA, oxidative stress, and ischemia (Sapolsky, 1985; Sapolsky and<br />

Pulsinelli, 1985; Sapolsky, 1986a; Sapolsky et al., 1988; Goodman et al., 1996; McIntosh and<br />

Sapolsky, 1996a). Downregulation of the GR in dentate gyrus might protect granule cells from<br />

effects of excitotoxicity such as extracellular glutamate accumulation, Ca 2+ influx, and oxidative<br />

stress.<br />

MR blockade increases CA3 vulnerability to KA<br />

Kainic acid decreased the number of live CA3 neurons counted in SPIRO-pretreated<br />

subjects; however, KA had no significant effect on numbers of viable neurons in any other<br />

region or group (Table 3.2). <strong>The</strong>se data indicate that MR blockade increases susceptibility to<br />

KA-induced degeneration in CA3. We have previously demonstrated that MR blockade with<br />

SPIRO decreases basal mRNA levels of the anti-apoptotic gene bcl-2 in the hippocampal<br />

pyramidal layers (<strong>McCullers</strong> and Herman, 1998) (Chapter Two). Thus, MR appears to play a<br />

neuroprotective role during neuronal injury that may occur through maintenance of viabilityenhancing<br />

genes such as bcl-2.<br />

<strong>The</strong> results of this study demonstrate that MR and GR mRNA levels respond to stress as<br />

conveyed not only by differential receptor activation, but also by ACR-independent signals<br />

generated by excitotoxic stress. Further, kainic acid regulation of ACR mRNA levels appears to<br />

override ACR-dependent feedback mechanisms. <strong>The</strong> present evidence for modulation of ACR<br />

mRNA expression during injury emphasizes the importance of ACR function for maintenance of<br />

neuronal viability, particularly considering the prevalence of HPA dysregulation in<br />

neurodegenerative conditions such as Alzheimer’s disease and in normal aging.<br />

54


Chapter Four: Mifepristone protects CA1 hippocampal neurons following<br />

traumatic brain injury in rat<br />

Introduction<br />

Glucocorticoids modulate many cellular processes relevant to injury, including metabolism,<br />

ion balance, immune function, inflammation, neurotrophin expression, and cell death (Munck et<br />

al., 1984; McEwen et al., 1986; Reagan and McEwen, 1997; Joels and Vreugdenhil, 1998). Two<br />

steroid hormone receptors mediate glucocorticoid action: the mineralocorticoid receptor (MR)<br />

and the glucocorticoid receptor (GR). Glucocorticoids bind GR with lower affinity but higher<br />

capacity relative to MR; thus, MR is extensively bound at very low (e.g. circadian nadir)<br />

glucocorticoid levels, while GR mediates the effects of high glucocorticoid levels (Reul and de<br />

Kloet, 1985; Reul et al., 1987). <strong>The</strong> MR and GR are abundantly expressed in hippocampus<br />

(Reul and de Kloet, 1985; Aronsson et al., 1988; Arriza et al., 1988; Reul et al., 1989), making<br />

this region a prime target for glucocorticoid action. Excessive activation of GR has been shown<br />

to increase hippocampal neuronal vulnerability to several forms of neuronal insult, including<br />

excitotoxicity, oxidative stress, and ischemia (Sapolsky, 1985; Sapolsky and Pulsinelli, 1985;<br />

Sapolsky, 1986a; Sapolsky et al., 1988; Goodman et al., 1996; McIntosh and Sapolsky, 1996a).<br />

Moderate to severe traumatic brain injury (TBI) in rat produces neuronal damage and cell<br />

death in hippocampus (Cortez et al., 1989; Lowenstein et al., 1992; Goodman et al., 1994; Hicks<br />

et al., 1996). <strong>The</strong> significant cell death that occurs following TBI is believed to evolve in a<br />

biphasic sequence consisting of the primary mechanical insult coupled with a progressive<br />

secondary necrosis. Results of experimental TBI research have proposed several injury<br />

mechanisms including excitotoxicity (Choi, 1992; Hayes et al., 1992), ischemia (Okiyama et al.,<br />

1994), free radical formation (Braughler and Hall, 1992), elevated intracellular calcium (Young,<br />

1992), edema (Wahl et al., 1988), proteolysis (Kampfl et al., 1997), and mitochondrial<br />

dysfunction (Sullivan et al., 1998; Sullivan et al., 1999). Glucocorticoids likely modulate some<br />

or all of the processes mediating secondary injury. One possible mechanism whereby<br />

glucocorticoids could influence neuronal vulnerability to injury is through regulation of the antiapoptotic<br />

B-cell leukemia/lymphoma-2 gene, bcl-2. Several studies indicate that bcl-2 is induced<br />

by TBI and has neuroprotective effects after traumatic injury (Clark et al., 1997; Raghupathi et<br />

55


al., 1998; Nakamura et al., 1999). Our laboratory has demonstrated that MR blockade with<br />

spironolactone decreases basal bcl-2 messenger RNA (mRNA) levels (<strong>McCullers</strong> and Herman,<br />

1998) (Chapter Two), suggesting that MR activation maintains neuroprotective bcl-2 expression<br />

and may participate in regulation of neuronal viability after TBI.<br />

<strong>The</strong> present study is designed to test the hypotheses that GR blockade will attenuate and MR<br />

blockade will exacerbate hippocampal cell loss following controlled cortical impact (CCI), an<br />

experimental model of TBI. Animals were pretreated with vehicle, the MR antagonist<br />

spironolactone (SPIRO), or the GR antagonist mifepristone (RU486) and subsequently subjected<br />

to either sham operation or injury by CCI. Using the optical fractionator method, CA1, CA3,<br />

and dentate gyrus (DG) hippocampal neurons were counted twenty-four hours after surgery to<br />

determine SPIRO and RU486 effects on cell survival after CCI. In addition, mRNA levels of the<br />

anti-apoptotic gene bcl-2, the pro-apoptotic gene bax, and the tumor-suppressor gene p53 (a<br />

transcriptional regulator of bcl-2 and bax) (Miyashita et al., 1994) were examined after MR or<br />

GR blockade and CCI to determine whether MR or GR regulation of these genes correlates with<br />

neuronal viability after traumatic brain injury.<br />

Materials and Methods<br />

Subjects<br />

Young adult male Sprague Dawley rats (n=60) (Harlan Sprague Dawley, Inc., Indianapolis,<br />

IN) weighing 200-250g were maintained on a 12 hr light/dark cycle in an environment with<br />

constant temperature, humidity, and ad libitum access to food and water. All animal procedures<br />

were performed in accordance with the National Institutes of Health guidelines and approved by<br />

the University of Kentucky Institutional Animal Care and Use Committee. All efforts were<br />

made to minimize animal suffering and the number of animals used.<br />

In vivo protocols<br />

Subjects received subcutaneous injections of propylene glycol vehicle, MR antagonist<br />

spironolactone (SPIRO) (50 mg/kg) (Herman and Spencer, 1998) (Sigma, St. Louis, MO), or GR<br />

antagonist RU486 (25 mg/kg) (Ratka et al., 1989) (Sigma) twice daily, in morning and afternoon.<br />

Pretreatment was conducted for forty-eight hours to ensure adequate time for the altered<br />

56


expression of ACR-regulated genes (Herman and Spencer, 1998). After a final injection was<br />

administered at 900h on day three, subjects were transported to a surgical procedure room.<br />

Between 1100h-1200h, animals were subjected to craniotomy only (sham operation) or a<br />

moderate unilateral cortical contusion (2mm). All animals were sacrificed on day four at 1200h,<br />

approximately twenty-four hours after surgery. 8-12 subjects were included in each<br />

pretreatment/surgery group. Half of the subjects in each condition were perfused for cell counts,<br />

and half were decapitated for in situ hybridization analysis, yielding n = 4-6 subjects per<br />

statistical group. Surgical procedures are briefly described below.<br />

<strong>The</strong> moderate CCI injury used in these experiments results in significant loss of cortical<br />

tissue, blood brain barrier disruption, and loss of hippocampal neurons as previously described<br />

(Baldwin et al., 1997; Scheff and Sullivan, 1999), thus mimicking aspects of the pathology of<br />

human closed-head injury. Animals were anesthetized with isoflurane (2%) and placed in a<br />

stereotaxic frame (Kopf Instr., Tujunga, CA) prior to CCI injury. Using sterile procedures, the<br />

skin was retracted and a 6mm unilateral craniotomy was centered between bregma and lambda.<br />

<strong>The</strong> skull cap was removed without disruption of the underlying dura. <strong>The</strong> exposed brain was<br />

injured using a pneumatically controlled impacting device with a 5mm beveled tip that<br />

compressed the cortex 3.5 m/sec to a depth of 2mm. Following injury, Surgicel (Johnson and<br />

Johnson, Arlington, TX) was laid upon the dura, and the skull cap was replaced. A thin coat of<br />

dental acrylic was then spread over the craniotomy site and allowed to dry before the wound was<br />

stapled closed. Body temperature was monitored continuously throughout the surgical/recovery<br />

procedures and maintained at 37 o C with a heating pad.<br />

Animals processed for in situ hybridization experiments were killed by decapitation. Brains<br />

were rapidly removed, frozen in isopentane cooled to –40 o C on dry ice, then stored at –80 o C<br />

until sectioned. Trunk blood samples were collected on ice, and plasma was stored at -20 o C<br />

until plasma corticosterone levels were measured using commercial radioimmunoassay kits per<br />

manufacturers’ instructions (ICN Biochemicals, Costa Mesa, CA). Animals processed for cell<br />

counts received a lethal dose of sodium pentobarbitol (Butler, Columbus, OH) (150mg/kg)<br />

before undergoing transcardial perfusion with 0.9% saline followed by 4% paraformaldehyde.<br />

Brains were removed and post-fixed in 4% paraformaldehyde overnight before further<br />

processing.<br />

57


Tissue processing<br />

Frozen brains were removed from a –80 o C freezer, warmed to –20 o C, and sectioned to 16 um<br />

with a Bright-Hacker cryostat. Sections were then thaw-mounted onto Ultrastick Slides (Becton<br />

Dickinson, Franklin Lakes, NJ) and stored at –20 o C until processing for in situ hybridization.<br />

Perfused brains were processed as previously described (Baldwin et al., 1997). Briefly, brains<br />

were placed in a graded series of celloidin (Mallinckrodt Baker, Inc., Paris. KY) dissolved in<br />

50:50 ETOH:ethyl ether. After removal from 12% celloidin, the brains were placed in Peel-A-<br />

Way (Polyscience, Warrington, PA) disposable histology embedding molds with fresh 12%<br />

celloidin and exposed to chloroform vapors for 18 hours. Once the celloidin hardened, molds<br />

were removed. <strong>The</strong> embedded tissue was then attached to wooden blocks and stored in 70%<br />

EtOH until sectioning. <strong>The</strong> brains were cut in coronal sections (50 um) with a Zeiss HM440E<br />

sliding microtome. Sections were collected from anterior to posterior beginning at the<br />

appearance of anterior hippocampus. In order to select a series in an unbiased manner such that<br />

every section had an equal probability of being selected, a die was thrown to determine the first<br />

section saved, after which every twelfth section was kept for staining with Giemsa. Six drops of<br />

1% glacial acetic acid were added per 100 ml of Giemsa stain (Sigma, St. Louis, MO) diluted 1:4<br />

with water. Following staining the sections were placed in cedar wood oil for 5 min to reduce<br />

background and subsequently washed in Hemo-D (Fisher Scientific, Pittsburgh, PA). <strong>The</strong><br />

sections were then mounted on slides, covered with permount, and coverslipped.<br />

Optical fractionator method<br />

Bioquant Image Analysis software (R and M Biometrics, Memphis, TN) was used to<br />

estimate total cell number in regions of interest using the optical fractionator method of cell<br />

counting. <strong>The</strong> optical fractionator method involves the combining of optical dissector counting<br />

with fractionator sampling. This method of cell counting has several desirable attributes, such as<br />

being unaffected by tissue shrinkage and not requiring rigorous definitions of structural<br />

boundaries. For detailed explanation and discussion please see (West et al., 1991). <strong>The</strong> optical<br />

fractionator involves counting neurons with optical dissectors in a uniform sample that<br />

constitutes a known fraction of the structure being analyzed; the method systematically samples<br />

a known fraction of the section thickness, of a known fraction of sectional area, of a known<br />

58


fraction of the sections that contain the region of interest. <strong>The</strong> equation is as follows (our<br />

specific parameters):<br />

N = ∑ Q • t 1 1<br />

• •<br />

h asf ssf<br />

Where N is the total number of neurons, and Σ Q is the total number of neurons counted in each<br />

region of interest in each animal. t is the measured section thickness (42-45 um), h is the height<br />

of the dissector (20 um), 1/asf is the counting grid area (100 um X 100 um)/ the dissector area<br />

(20 um X 20 um), and 1/ssf is the sampling section fraction (12). Criteria for counting neuronal<br />

nuclei were observed as previously described (Baldwin et al., 1997). Briefly, the plane of focus<br />

was moved 5 um down from the top of the section to obviate the problem of an uneven counting<br />

surface. Neuronal nuclei in focus at level –5 um, the top level of the dissector, were not counted,<br />

but nuclei in focus at –25 um, the bottom level of the dissector, were counted. Through the<br />

dissector, nuclei completely within the counting frame or touching the upper or right side of the<br />

frame were counted. Nuclei touching the left or lower sides of the frame were not counted<br />

(Baldwin et al., 1997). Coefficient of error (CE) and coefficient of variation (CV) were<br />

calculated to determine intra-animal variation and inter-animal variation, respectively (Tables 4.1<br />

and 4.2).<br />

Table 4.1. Mean coefficient of error (CE)<br />

CA1 CA3 Dentate gyrus<br />

Vehicle 0.082 0.144 0.059<br />

Spironolactone 0.090 0.143 0.093<br />

RU486 0.071 0.122 0.080<br />

Table 4.2. Mean coefficient of variation (CV)<br />

CA1 CA3 Dentate gyrus<br />

Vehicle 0.241 0.365 0.156<br />

Spironolactone 0.162 0.304 0.094<br />

RU486 0.372 0.218 0.086<br />

Cell count data were analyzed by two-way ANOVA within each subregion, with hemisphere<br />

and drug designated as independent variables. Pairwise comparisons were made using the<br />

Student-Neuman-Keuls test with significance set at p ≤ 0.05. Data in Figure 4.1 are presented as<br />

neurons x 10 4 ± standard error.<br />

59


In situ hybridization<br />

Antisense riboprobes complementary to coding regions for bcl-2 (550 bp) (courtesy of Dr. D.<br />

Vaux, <strong>The</strong> Walter and Eliza Hall Institute, <strong>The</strong> Royal Melbourne Hospital, Victoria, Australia),<br />

bax (100 bp) (courtesy of Dr. J. Springer, University of Kentucky, KY, USA), and p53 (950 bp)<br />

(courtesy of Dr. G. Lozano, MD Anderson Cancer Center, Houston, TX, USA) mRNAs were<br />

generated for in situ hybridization. Probes were labeled according to previously published<br />

methods (Seroogy and Herman, 1996) by standard in vitro transcription using [ 35 S]UTP. Tissue<br />

sections removed from a –20 o C freezer were fixed in 4% paraformaldehyde for 10 min. Slides<br />

were then rinsed sequentially in the following solutions: 5 mM KPBS, 5 mM KPBS with 0.2%<br />

glycine, 5mM KPBS, 0.1 M triethanolamine with 0.25% acetic anhydride, 0.2X SSC, and graded<br />

alcohols. Labeled probes were added to a hybridization buffer containing 50% formamide, and<br />

diluted probe (50 ul containing 1x 10 6 c.p.m.) was applied to each slide. Slides were<br />

coverslipped, placed in moistened chambers, and incubated overnight at 55 o C. Following<br />

hybridization, slides were rinsed in 2X SSC and incubated in 0.2X SSC at 60 o C for 1 hr. Slides<br />

were then dehydrated and exposed for 14-18 days to Kodak BioMAX xray film.<br />

Autoradiographic images were digitized with NIH Image 1.62 software for Macintosh. A<br />

standard curve was generated prior to image capture to ensure linearity of the signal. Based on<br />

morphological criteria, regions CA1, CA3, and DG of the hippocampal formation were manually<br />

sampled from both hemispheres for densitometric analysis. <strong>The</strong> mean gray level of an<br />

unhybridized (background) area from each hemisphere was subtracted from sampled regions to<br />

arrive at corrected gray level. Ipsilateral mRNA data were normalized as a percent of control<br />

(sham/vehicle) values from the ipsilateral hemisphere, and contralateral mRNA data were<br />

normalized as a percent of control (sham/vehicle) values from the contralateral hemisphere, with<br />

sham/vehicle levels expressed as one hundred percent. In situ hybridization data were then<br />

analyzed by two-way ANOVA within each hemisphere and subregion with drug and injury<br />

designated as independent variables. Post hoc comparisons were made using the Student-<br />

Newman-Keuls test, with significance set at p ≤ 0.05.<br />

60


Results<br />

Effects of adrenocorticosteroid receptor blockade and traumatic brain injury on hippocampal<br />

neuron viability<br />

<strong>The</strong> numbers of hippocampal neurons surviving CCI at 24 hours in hippocampal subregions<br />

CA1, CA3, and DG were assessed using the optical fractionator method of cell counting (see<br />

Methods). Because sham operation does not cause cell loss and damage is confined to the<br />

ipsilateral hemisphere of injured rats, the numbers of neurons in sham/ipsilateral,<br />

sham/contralateral, and injured/contralateral hemispheres were statistically indistinguishable (i.e.<br />

unchanged by surgery) within each subregion (data not shown). <strong>The</strong>refore, cell counts from<br />

injured/contralateral hemispheres were used as control values against which to compare<br />

injured/ipsilateral hemisphere values to determine the effects of CCI on neuronal survival.<br />

Effects of antagonist pretreatments were assessed as compared to vehicle within the ipsilateral<br />

and contralateral hemispheres.<br />

Two-way ANOVA indicates a significant effect of hemisphere on numbers of neurons in<br />

CA1 [F(1,24) = 9.32; p ≤ 0.05], CA3 [F(1,24) = 38.17; p ≤ 0.05], and DG [F(1,24) = 5.12; p ≤<br />

0.05] (Figure 4.1). In CA1 of RU486-pretreated animals, numbers of surviving neurons were<br />

statistically indistinguishable in the ipsilateral and contralateral hemispheres. In contrast,<br />

numbers of hippocampal neurons in CA1 of vehicle and SPIRO-pretreated animals were<br />

significantly decreased by injury in the ipsilateral as compared to contralateral hemisphere. In<br />

CA3, injury significantly reduced the number of viable pyramidal neurons in the ipsilateral<br />

hemisphere in all pretreatment groups. Injury also significantly decreased the number of<br />

ipsilateral DG granule neurons counted in SPIRO and RU486-pretreated animals. Post-injury<br />

corticosterone levels fell within the range of approximately 15-80 ng/ml in all drug pretreatment<br />

groups, indicating that excessive corticosterone levels were not likely to contribute to the<br />

differential neuronal survival between groups.<br />

61


Figure 4.1. Effects of MR blockade, GR blockade, and controlled cortical impact injury on<br />

hippocampal neuron viability as measured by the optical fractionator method (see Methods).<br />

Injury significantly decreased the number of viable neurons in CA1 and CA3 of vehiclepretreated<br />

animals (n = 6). Pretreatment with RU486 protected CA1 pyramidal neurons from<br />

injury-induced cell loss (n = 5). Injury also decreased the number of viable neurons in CA3 and<br />

dentate gyrus of spironolactone (n = 4) and RU486-pretreated animals. [ B p ≤ 0.05 as compared<br />

to contralateral hemisphere within same pretreatment group (effect of injury).]<br />

Effects of adrenocorticosteroid receptor blockade and traumatic brain injury on bcl-2, bax, and<br />

p53 mRNA levels<br />

<strong>The</strong> effects of MR or GR blockade and CCI on hippocampal bcl-2, bax, and p53 mRNA<br />

levels were determined using densitometric analysis following in situ hybridization with genespecific<br />

radiolabeled riboprobes (see Methods, Figure 4.2). Unlike measures of cell loss, mRNA<br />

levels may be affected by injury both in the contralateral (unoperated) and ipsilateral (shamoperated<br />

or injured) hemispheres. <strong>The</strong>refore, all injury-induced changes were assessed as<br />

compared to sham values from the same hemisphere. To this end, separate two-way ANOVAs<br />

were conducted comparing sham/contralateral to injured/contralateral values and comparing<br />

sham/ipsilateral to injured/ipsilateral values, with drug and injury as independent variables. In<br />

each hemisphere, effects of drug were assessed as compared to vehicle within the sham or<br />

injured treatment groups.<br />

62


IPSI<br />

CONTRA<br />

Figure 4.2. Photomicrographs illustrating (A) bcl-2, (B) bax, and (C) p53 mRNA levels in<br />

hippocampus of vehicle-pretreated, injured rats at 24h. Injury decreases bcl-2 mRNA levels in<br />

ipsilateral dentate gyrus (panel A) but has no significant effect on bax or p53 mRNA levels in<br />

vehicle-pretreated rats (panels B,C).<br />

63


Two-way ANOVA demonstrates significant drug/injury interaction effects on contralateral<br />

bcl-2 mRNA levels in CA1 [F(2,27) = 3.47; p ≤ 0.05] and DG [F(2,27) = 5.67; p ≤ 0.05] (Figure<br />

4.3). Post hoc comparisons indicate that bcl-2 signal was significantly decreased by RU486 and<br />

by SPIRO in contralateral CA1 and DG of sham-operated animals. Injury increased contralateral<br />

bcl-2 mRNA levels in DG of RU486 animals.<br />

Drug/injury interaction effects on bcl-2 mRNA levels are observed in CA1 [F(2,27) = 5.01; p<br />

≤ 0.05], CA3 [F(2,27) = 3.94; p ≤ 0.05], and DG [F(2,27) = 6.92; p ≤ 0.05] of the ipsilateral<br />

hemisphere (Figure 4.3). In sham-operated animals, ipsilateral bcl-2 mRNA levels were<br />

significantly decreased by SPIRO in DG and by RU486 in CA1 and DG. CCI decreased<br />

ipsilateral bcl-2 signal in DG of vehicle animals and increased ipsilateral bcl-2 signal in DG of<br />

RU486 animals.<br />

Overall, bax mRNA levels were increased in DG [F(1,28) = 4.49; p ≤ 0.05] (Figure 4.4)<br />

contralateral to CCI. However, post hoc comparisons demonstrate no significant effects of drug<br />

or injury on bax mRNA levels, in any region or hemisphere.<br />

Significant effects of injury on p53 mRNA levels are detected in contralateral CA1 [F(1,29)<br />

= 6.97; p ≤ 0.05] and DG [F(1,29) = 4.57] (Figure 4.5). Post hoc comparisons indicate that CCI<br />

decreased contralateral p53 signal in CA1 and DG of SPIRO-pretreated animals. No significant<br />

effects of drug or injury on p53 mRNA levels were observed in the ipsilateral hemisphere.<br />

64


Figure 4.3. Effects of MR blockade, GR blockade, and controlled cortical impact injury on bcl-2<br />

mRNA levels in hippocampus. Semi-quantitative analysis expressed as percentage of control<br />

(sham/vehicle; see methods) corrected gray level; n = 5-6 per group. Contralateral bcl-2 mRNA<br />

levels were decreased by spironolactone and RU486 in CA1 and dentate gyrus of sham-operated<br />

animals. Ipsilateral bcl-2 levels were decreased by spironolactone in dentate gyrus and by<br />

RU486 in CA1 and dentate gyrus of sham-operated animals. Injury increased contralateral bcl-2<br />

mRNA levels in dentate gyrus of RU486 animals. CCI decreased ipsilateral bcl-2 signal in<br />

dentate gyrus of vehicle animals and increased ipsilateral bcl-2 signal in dentate gyrus of RU486<br />

animals. [ A p ≤ 0.05 as compared to vehicle animals within the same surgery condition (effect of<br />

drug). B p ≤ 0.05 as compared to sham-operated animals within same drug pretreatment group<br />

(effect of injury).]<br />

65


Figure 4.4. Effects of MR blockade, GR blockade, and controlled cortical impact injury on bax<br />

mRNA levels in hippocampus. Semi-quantitative analysis expressed as percentage of control<br />

(sham/vehicle; see methods) corrected gray level; n = 4-6 per group. Post hoc comparisons<br />

indicate no significant effects of antagonist pretreatment or injury on bax mRNA levels in any<br />

region or hemisphere.<br />

66


Figure 4.5. Effects of MR blockade, GR blockade, and controlled cortical impact injury on p53<br />

mRNA levels in hippocampus. Semi-quantitative analysis expressed as percentage of control<br />

(sham/vehicle; see methods) corrected gray level; n = 5-6 per group. Injury decreased<br />

contralateral p53 mRNA levels in CA1 and dentate gyrus of SPIRO-pretreated animals. Post hoc<br />

comparisons indicate no significant effects of antagonist pretreatments. [ B p ≤ 0.05 as compared<br />

to sham-operated animals within same drug pretreatment group (effect of injury).]<br />

67


Discussion<br />

<strong>The</strong> present study demonstrates that GR blockade with RU486 prevents CA1 neuron loss 24h<br />

after CCI injury. <strong>The</strong> data do not suggest that regulation of viability-related genes bcl-2, bax, or<br />

p53 is involved in the selective preservation of CA1 neurons by RU486. Furthermore, MR<br />

blockade with SPIRO did not exacerbate CCI-induced neuron loss in hippocampus.<br />

Hippocampal neuron loss 24 hours after controlled cortical impact injury<br />

<strong>The</strong> present study uses the optical fractionator method of cell counting, a procedure that<br />

yields an unbiased estimation of total cell number, to evaluate survival of hippocampal neurons<br />

24h after CCI injury (Figure 4.1). Counts revealed that CCI significantly decreased the number<br />

of ipsilateral CA1 pyramidal neurons in vehicle and SPIRO-pretreated rats by approximately<br />

35%. In contrast, no significant cell loss occurred in RU486-pretreated animals, demonstrating<br />

protection of CA1 neurons by RU486. Approximately 40-55% of CA3 neurons were lost at 24h.<br />

We have previously reported a similar amount of cell loss in CA3 after the same level of injury<br />

using the optical dissector method of cell counting. <strong>The</strong> same report demonstrated that no<br />

significant cell loss occurs in CA3 beyond 24 hours post-injury, indicating that neuronal cell<br />

death in this region is rapid following TBI (Baldwin et al., 1997). In the current study, CCI also<br />

reduced the number of granule cells in DG by approximately 20-30%. While injury did not<br />

significantly decrease the number of DG neurons in the vehicle-pretreated group, post-hoc<br />

comparisons indicate that the amount of cell loss in vehicle-pretreated subjects was not<br />

statistically different from that observed in SPIRO and RU486-pretreated subjects. <strong>The</strong><br />

appearance of the data suggests that the actual extent of cell loss in vehicle-pretreated animals<br />

may have been obscured by high variability in the number of contralateral cells counted in this<br />

region.<br />

RU486 protection is specific to CA1<br />

<strong>The</strong> selective defense of CA1 neurons suggests either that RU486 preferentially protects this<br />

neuronal population or that CCI injury in CA1 is more responsive to treatment. Differential<br />

action of RU486 might be achieved through the heterogeneous distribution of GR across<br />

subfields (Herman, 1993). <strong>The</strong> high density of GR expressed in CA1 relative to other<br />

68


hippocampal regions might explain, in part, the efficacy of RU486. Relatively little GR is<br />

expressed in CA3, perhaps lessening the impact of RU486 on this subregion. However, GR<br />

levels are substantial in DG, a region not protected by RU486, suggesting that additional factors<br />

are involved in determining neuronal susceptibility to injury.<br />

Previous studies predict that CCI may affect neurons differently according to subregion,<br />

allowing for selective intervention. Subpopulations of hippocampal neurons are selectively<br />

vulnerable to different forms of insult; CA1 neurons are most severely damaged by ischemia<br />

(Kirino, 1982; Pulsinelli et al., 1982; Schmidt-Kastner and Freund, 1991), while CA3 neurons<br />

are most vulnerable, and DG granule cells least vulnerable, to kainate excitotoxicity (Nadler et<br />

al., 1978; Kohler and Schwarcz, 1983). Ischemia (Okiyama et al., 1994) and excitotoxicity<br />

(Choi, 1992; Hayes et al., 1992) are among the many components of TBI which can lead to<br />

neuronal death. Differential vulnerability in hippocampal neuron populations could therefore<br />

govern the severity of CCI injury or the length of time over which cell death occurs, allowing for<br />

more successful treatment in specific regions. CCI results in very rapid cell loss in CA3<br />

(Baldwin et al., 1997). Death of CA1 neurons as induced by ischemia is slower in comparison,<br />

occuring over a period of hours to days (Kirino, 1982; Pulsinelli et al., 1982). It is possible that<br />

CCI injury is either delayed or less severe in CA1 as compared to CA3, such that intervention<br />

with RU486 is able to increase or prolong neuronal survival. Clarification of this issue will<br />

require further elucidation of the mechanisms governing CCI-induced cell death and a time<br />

course characterization of CA1 neuron survival after CCI.<br />

A variety of factors could account for differential vulnerability to TBI by region. For<br />

example, CA3 neurons do not express neuronal calcium-binding proteins calbindin-D 28k or<br />

parvalbumin (Sloviter, 1989), which may reduce CA3 Ca 2+ buffering capacity as compared to<br />

CA1. In addition, ionotropic glutamate receptors are heterogeneously distributed in<br />

hippocampus; NMDA and AMPA subunits are preferentially expressed in CA1 and DG, while<br />

kainate receptors are mainly found in CA3 (Greenamyre et al., 1985; Tarazi et al., 1996).<br />

Glutamate receptor configuration may dictate sensitivity to the large amounts of glutamate<br />

released during injury. Subregional differences in afferent input, intracellular signaling, control<br />

of ion homeostasis, antioxidant defenses, and neurotrophic factor expression are among other<br />

factors that could also differentially influence cell viability.<br />

69


Glucocorticoid receptor blockade and neuronal viability<br />

GR-activating concentrations of glucocorticoids exacerbate many types of neuronal insult<br />

including kainic acid toxicity, oxidative stress, amyloid-beta peptide toxicity, and ischemia<br />

(Sapolsky, 1985; Sapolsky and Pulsinelli, 1985; Sapolsky, 1986a; Sapolsky et al., 1988;<br />

Goodman et al., 1996; McIntosh and Sapolsky, 1996a). <strong>The</strong> GR may heighten neuronal<br />

sensitivity to insult through several mechanisms. Glucocorticoids can inhibit glucose transport<br />

(Horner et al., 1990) and may increase neuronal vulnerability to insult through energy depletion<br />

(Sapolsky, 1986b). Elevated glucocorticoid concentrations can also result in excitatory amino<br />

acid accumulation (Stein-Behrens et al., 1992; Stein-Behrens et al., 1994b), increased<br />

intracellular Ca 2+ concentrations (Kerr et al., 1992; Elliott and Sapolsky, 1993; Karst et al.,<br />

1994), and free radical formation (McIntosh and Sapolsky, 1996a, b), all of which may<br />

compromise cells by triggering Ca 2+ -dependent proteolysis, lipid peroxidation, oxidative stress,<br />

and mitochondrial dysfunction.<br />

Inactivation of GR by glucocorticoid removal or GR blockade has been shown to improve<br />

survival of rodent hippocampal neurons following several types of neurotoxic challenge.<br />

Metyrapone, an inhibitor of adrenal glucocorticoid synthesis, reduces loss of CA1 neurons after<br />

ischemia (Smith-Swintosky et al., 1996) and loss of CA1 and CA3 neurons after insult with<br />

kainic acid, an excitotoxic glutamate analog (Stein and Sapolsky, 1988; Smith-Swintosky et al.,<br />

1996). RU486 reverses glucocorticoid-mediated vulnerability to glutamate in a cultured murine<br />

hippocampal cell line (Behl et al., 1997b). Further, RU486 (Antonawich et al., 1999) and<br />

removal of endogenous glucocorticoids by adrenalectomy (Sapolsky and Pulsinelli, 1985; Morse<br />

and Davis, 1990) attenuate loss of CA1 neurons following ischemia. However, studies have<br />

indicated that while adrenalectomy and RU486 improve CA1 viability in gerbil at 4 days<br />

following ischemia, the protection is lost at 7 days (Morse and Davis, 1990; Antonawich et al.,<br />

1999). Thus it is possible that RU486 protection of CA1 neurons after CCI is a transient effect.<br />

Additional time points will be required in order to determine if RU486 completely prevents CA1<br />

cell loss or merely delays neuronal loss after injury.<br />

Metyrapone and adrenalectomy studies suggest that RU486 may exert its neuroprotective<br />

effects by blocking the deleterious effects of GR. RU486 is a potent inhibitor of GR activation<br />

that both impedes and competes with glucocorticoid binding to the GR. Complexes of RU486<br />

bound to the GR can bind glucocorticoid response elements found in target gene promoter<br />

70


egions but do not facilitate transcription, c.f.(Cadepond et al., 1997). While RU486 generally<br />

acts as an anti-glucocorticoid, it may also act as a GR agonist under certain circumstances<br />

(Nordeen et al., 1995). It is therefore possible that RU486 influences GR activation in an<br />

unpredicted manner after CCI. However, we have consistently observed in the current subjects<br />

(<strong>McCullers</strong>, 2001a) (Chapter Five) and in a previous study (<strong>McCullers</strong> and Herman, 2001)<br />

(Chapter Three) that in vivo pretreatment with RU486 increases GR mRNA levels. This effect<br />

concurs with GR mRNA upregulation after adrenalectomy (Herman et al., 1989a; Reul et al.,<br />

1989) and provides evidence for the functional blockade of GR in this study. In addition to its<br />

anti-glucocorticoid effects, RU486 is also a potent progesterone receptor antagonist. However,<br />

modest hippocampal expression of progesterone receptor (Kato et al., 1994) and evidence that<br />

progesterone is neuroprotective after TBI (Roof and Hall, 2000) suggest that progesterone<br />

receptor blockade is unlikely to underlie RU486 protection of CA1 neurons after CCI.<br />

Potential mechanisms of RU486 neuroprotection<br />

Subfield-specific modulation of Ca 2+ homeostasis may influence neuronal viability after TBI.<br />

Studies of CA1 neurons in hippocampal slices demonstrate enhanced afterhyperpolarization<br />

(Joels and de Kloet, 1989; Kerr et al., 1989), Ca 2+ action potentials (Kerr et al., 1989), and Ca 2+<br />

currents (Kerr et al., 1992; Karst et al., 1994) with increased GR activation. <strong>The</strong>se reports<br />

indicate that GR influences Ca 2+ influx in CA1 neurons and suggest that excessive GR activation<br />

could lead to Ca 2+ neurotoxicity. In agreement with this hypothesis, prolonged exposure to high<br />

glucocorticoid concentrations increases basal intracellular Ca 2+ levels (Elliott and Sapolsky,<br />

1993) and prolongs intracellular Ca 2+ elevation after kainic acid treatment (Elliott and Sapolsky,<br />

1992) in hippocampal neuron cultures. Characterizations of Ca 2+ channel subunit mRNA<br />

expression at low and high glucocorticoid levels are also consistent with the hypothesis that<br />

predominant MR occupation leads to restricted Ca 2+ influx, while additional GR occupation<br />

enhances Ca 2+ influx (Nair et al., 1998). In the present study, RU486 blockade of GR should<br />

result in predominant MR activation. RU486 may therefore shift expression of Ca 2+ channel<br />

subunits and other Ca 2+ regulatory genes to a neuroprotective configuration that could prevent<br />

excessive intracellular Ca 2+ accumulation in CA1 neurons.<br />

RU486 could also improve CA1 neuronal viability via non-genomic action. RU486 is<br />

reported to have antioxidant effects that may mediate neuroprotection in CA1 pyramidal cells.<br />

71


Previous studies have shown that RU486 prevents oxidation of the low-density lipoprotein<br />

(Parthasarathy et al., 1994) and inhibits oxidative modification of proteins by pre-existing lipid<br />

peroxides in cell-free systems (Carpenter et al., 1996). Furthermore, Behl, et al. (1997a) have<br />

demonstrated neuroprotective antioxidant effects of RU486 in cellular and organotypic slice<br />

cultures. Pretreatment with 10 -5 M RU486 protects rat primary hippocampal cultures from beta<br />

amyloid toxicity, mouse clonal hippocampal HT22 cells from beta amyloid, H 2 O 2 , and glutamate<br />

insults, and organotypic rat hippocampal slices from H 2 O 2 -induced cell death (Behl et al.,<br />

1997a). Intracellular peroxide accumulation and DNA degradation were also prevented by<br />

RU486 in HT22 cells. Experiments in monkey kidney CV 1 cells suggest that the protective<br />

effects of RU486 are independent of GR or PR antagonism or activation (Behl et al., 1997a).<br />

Structural similarity to other antioxidant compounds suggests that the dimethylaminophenyl side<br />

chain moiety of RU486 is likely responsible for the antioxidant properties of RU486<br />

(Parthasarathy et al., 1994; Carpenter et al., 1996; Behl et al., 1997a).<br />

Expression of apoptosis-related genes after corticosteroid receptor blockade and injury<br />

<strong>The</strong> present data demonstrate decreased hippocampal bcl-2 mRNA levels in SPIRO and<br />

RU486-pretreated rats 24h after sham-operation, as compared to vehicle-pretreated subjects<br />

(Figure 4.3). <strong>The</strong>se results extend our previous finding that SPIRO decreases bcl-2 mRNA<br />

levels in hippocampus (<strong>McCullers</strong> and Herman, 1998) (Chapter Two). This evidence for<br />

reduction of bcl-2 mRNA levels by MR and GR antagonists suggests that both receptor types,<br />

perhaps functioning as heterodimers (Trapp et al., 1994), are required for optimal maintenance of<br />

bcl-2 expression. However, the antagonist-induced decreases in bcl-2 mRNA levels observed in<br />

sham animals do not appear to predict increased hippocampal neuron vulnerability to injury.<br />

Injury-induced cell loss is not increased in antagonist-injected subjects in any region (as<br />

compared to vehicle-injected controls); on the contrary, RU486 treated subjects clearly show<br />

protection despite decreased bcl-2 mRNA levels observed in sham animals. <strong>The</strong>se data fail to<br />

support the hypothesis that decreased bcl-2 expression subsequent to ACR blockade will increase<br />

neuronal vulnerability to CCI. It should be noted, however, that the effects of ACR blockade on<br />

bcl-2 mRNA levels are very different in sham versus injured animals. Among injured animals,<br />

ACR blockade did not reduce bcl-2 mRNA levels (as compared to vehicle-injected controls) 24<br />

hours post-injury in any hippocampal region. Thus, CCI injury abrogates the differential<br />

72


decrease of bcl-2 mRNA levels in antagonist-pretreated groups, perhaps accounting for the lack<br />

of correlation between sham bcl-2 levels and injury induced cell loss. Nonetheless, it is possible<br />

that a shorter period of antagonist pretreatment may be sufficient for RU486 protection of CA1<br />

neurons while avoiding the potential undesired effect of bcl-2 mRNA downregulation.<br />

<strong>The</strong> present study demonstrates decreased ipsilateral bcl-2 mRNA levels in DG of vehiclepretreated<br />

subjects following CCI injury. Decreased bcl-2 expression has also been<br />

demonstrated in fluid percussion (Dong et al., 2001) and in vitro (Morrison et al., 2000) models<br />

of TBI. In contrast, Clark et al. have reported increases in hippocampal bcl-2 mRNA levels 24h<br />

following CCI with the addition of 30 minutes moderate hypoxemia following injury (Clark et<br />

al., 1997). Increased bcl-2 mRNA levels appear to be associated with neuroprotection after CCI<br />

with hypoxia (Clark et al., 1997), and Bcl-2 overexpression in transgenic mice decreases<br />

hippocampal tissue loss after CCI (Nakamura et al., 1999). In addition, decreased bcl-2 mRNA<br />

levels have been associated with increased kainate-induced hippocampal cell loss (Gillardon et<br />

al., 1995; <strong>McCullers</strong> and Herman, 2001) (Chapter Three). However, decreased bcl-2 mRNA<br />

levels observed in DG of injured animals in the current study were not associated with<br />

significant neuronal loss following CCI injury. As previously noted, a high degree of variability<br />

in contralateral DG cell counts may have obscured significant effects of CCI on DG neuron loss.<br />

Even so, the process of secondary neuronal degeneration is very complex and involves numerous<br />

factors, and decreased bcl-2 expression, though perhaps an endangering factor, need not dictate<br />

neuronal loss following injury.<br />

Hippocampal bax mRNA levels were unchanged by MR/GR antagonist pretreatments or CCI<br />

(Figure 4.4). <strong>The</strong>se findings agree with results obtained using an in vitro model of TBI<br />

(Morrison et al., 2000). <strong>The</strong> ratio of anti-apoptotic proteins such as Bcl-2 to pro-apoptotic<br />

proteins such as Bax is believed to determine cell sensitivity to apoptotic signals (Oltvai et al.,<br />

1993). <strong>The</strong> current data suggest that transcriptional regulation of this ratio would be<br />

accomplished by modulation of bcl-2, and not bax, following traumatic brain injury.<br />

Levels of p53 mRNA in the present study were not significantly elevated in vehiclepretreated<br />

animals at the 24 hour time point; however, our data do not preclude changes earlier in<br />

time. For instance, another group has demonstrated increases in p53 mRNA levels 6 hours<br />

following fluid percussion brain injury which return to sham levels by 24 hours (Napieralski et<br />

al., 1999). <strong>The</strong> current report also demonstrates that injury reduced contralateral p53 signal in<br />

73


CA1 and DG of SPIRO animals (Figure 4.5). We have previously shown that SPIRO attenuates<br />

kainic acid-induced increases in p53 mRNA levels (<strong>McCullers</strong> and Herman, 1998) (Chapter<br />

Two), but the effects of p53 downregulation on neuronal survival are presently unclear.<br />

Hippocampal cell death is associated with functional (Lowenstein et al., 1992) and<br />

neurobehavioral (Hicks et al., 1993; Smith et al., 1994) deficits following TBI. By preventing<br />

CA1 neuron loss after CCI, RU486 could potentially improve behavioral outcome after injury.<br />

RU486 may be particularly effective in promoting neuron survival due to its combined antiglucocorticoid<br />

effects and antioxidant properties. While glucocorticoids have been used<br />

clinically to control edema after TBI, the value of this treatment is unclear. Some evidence<br />

indicates that glucocorticoids can improve recovery after central nervous system injury (Hall et<br />

al., 1992), but many studies show no improvement with glucocorticoid administration (Tornheim<br />

and McLaurin, 1978; Cooper et al., 1979; Giannotta et al., 1984; Brain, 1996). In contrast, many<br />

studies have demonstrated exacerbation of neuronal insult by glucocorticoids (Sapolsky, 1985;<br />

Sapolsky and Pulsinelli, 1985; Sapolsky, 1986a; Sapolsky et al., 1988; Goodman et al., 1996;<br />

McIntosh and Sapolsky, 1996a) and improved neuronal viability with antiglucocorticoid<br />

treatments (Sapolsky and Pulsinelli, 1985; Stein and Sapolsky, 1988; Morse and Davis, 1990;<br />

Smith-Swintosky et al., 1996; Behl et al., 1997b; Antonawich et al., 1999). Considering the<br />

benefits of glucocorticoid removal and blockade in other models of neuronal injury, the<br />

antioxidant properties of RU486, and the current data demonstrating CA1 neuron protection,<br />

RU486 appears to have potential as a useful treatment for head injury.<br />

74


Chapter Five: Traumatic brain injury regulates adrenocorticosteroid<br />

receptor mRNA levels in rat hippocampus<br />

Introduction<br />

<strong>The</strong> mineralocorticoid receptor (MR) and the glucocorticoid receptor (GR), collectively<br />

referred to as adrenocorticosteroid receptors (ACRs), are ligand-activated transcription factors<br />

that mediate glucocorticoid action in brain. <strong>The</strong> MR binds glucocorticoids with high affinity (kD<br />

~ 0.5-1 nM) and is extensively occupied at basal glucocorticoid levels (Reul and de Kloet, 1985).<br />

<strong>The</strong> GR binds glucocorticoids with a five-fold lower affinity, resulting in greater GR activation<br />

with increasing glucocorticoid levels (i.e. during stress and at the peak of circadian hormone<br />

secretion)(Reul and de Kloet, 1985; Spencer et al., 1990). This two-receptor system allows for<br />

differential regulation of MR and GR-coordinated gene expression in response to changes in<br />

circulating glucocorticoid concentration. <strong>The</strong> binding of glucocorticoid ligand allows receptor<br />

translocation to the cell nucleus, where the MR and GR can regulate transcription either by direct<br />

DNA binding to hormone response elements in gene promoter regions or by protein-protein<br />

interactions with other transcription factors [see (Yamamoto, 1985; De Kloet et al., 1998)].<br />

Considerable evidence suggests that excessive GR activation by CORT increases<br />

hippocampal vulnerability to several forms of neuronal insult including excitotoxicity, oxidative<br />

stress, and ischemia (Sapolsky, 1985; Sapolsky and Pulsinelli, 1985; Sapolsky, 1986a; Sapolsky<br />

et al., 1988; Goodman et al., 1996; McIntosh and Sapolsky, 1996a). Furthermore, blockade of<br />

GR reduces hippocampal neuron vulnerability to traumatic brain injury (TBI) (<strong>McCullers</strong> et al.,<br />

2001b) (Chapter Four), indicating the potential deleterious effects of GR activation with this type<br />

of injury. <strong>The</strong> degree of GR activation experienced by cells is largely determined by circulating<br />

glucocorticoid concentration and the amount of available GR; therefore, regulation of GR<br />

expression and the magnitude of HPA response are crucial determinants of glucocorticoid<br />

influence on neuronal viability after injury.<br />

To examine the respective roles of MR and GR in regulating ACR expression and<br />

glucocorticoid secretion following TBI, adult male Sprague Dawley rats were pretreated with<br />

vehicle, the MR antagonist spironolactone (SPIRO), or the GR antagonist mifepristone (RU486).<br />

Subjects were subsequently sham-operated or injured using the controlled cortical impact (CCI)<br />

75


model of TBI. Twenty-four hours post-injury, ACR mRNA levels were measured to determine<br />

if glucocorticoid signaling is altered through regulation of ACR expression following injury. As<br />

the acute hypothalamic-pituitary-adrenal (HPA) response to CCI has not been characterized,<br />

additional studies were performed to assess CORT and ACTH secretion over the 24-hour period<br />

following injury.<br />

Materials and Methods<br />

Subjects<br />

Young adult male Sprague Dawley rats (Harlan Sprague Dawley, Inc., Indianapolis, IN)<br />

weighing 200-250g were maintained in an environment with constant temperature, humidity, ad<br />

libitum access to food and water, and a 12-hour light/dark cycle, with lights on between 600h-<br />

1800h. All measures were taken to minimize animal pain or discomfort. All animal procedures<br />

were performed in accordance with the National Institutes of Health guidelines and approved by<br />

the University of Kentucky Institutional Animal Care and Use Committee.<br />

In vivo protocols<br />

Subjects (n = 5-7 per design group) received subcutaneous injections of propylene glycol<br />

vehicle, MR antagonist spironolactone (SPIRO) (50 mg/kg) (Herman and Spencer, 1998)<br />

(Sigma, St. Louis, MO), or GR antagonist RU486 (25 mg/kg) (Ratka et al., 1989) (Sigma) in<br />

morning (900h) and afternoon (1600h) for two days. On day three, animals were administered a<br />

final pretreatment injection at 900h and transported to a surgical procedure room. Between<br />

1100h-1200h, animals were subjected to craniotomy (sham operation) or a moderate unilateral<br />

cortical contusion, described below. Subjects were killed on day 4 at 1200h, 24h after surgery.<br />

A separate set of animals (n = 6 per group) was used for analysis of plasma ACTH and<br />

CORT levels during the 24 hours following surgery. Baseline blood samples were drawn by tail<br />

nick between 900-1000 hours, after which animals were transported to a surgical procedure<br />

room. Subjects then underwent sham operation or moderate unilateral cortical contusion<br />

between 1200-1300 hours. Sham and CCI surgeries were performed under identical conditions<br />

on separate days. Blood samples were collected at 1.5, 6, 12, 18, and 24 hours after the start of<br />

surgery. All samples were collected by tail nick, with the exception of the 24-hour point, which<br />

76


was obtained by collection of trunk blood at the time of sacrifice. Blood sampling was initiated<br />

1.5 hours after the start of surgery to allow all subjects to wake from anesthesia, and six hour<br />

intervals were observed thereafter to assess circadian hormone levels at times including lights off<br />

(1800 h) and lights on (600h).<br />

Surgical procedures were performed as previously described (Baldwin et al., 1997). Briefly,<br />

animals were anesthetized with isoflurane (2%) and placed in a stereotaxic frame (Kopf Instr.,<br />

Tujunga, CA). Sterile procedures were used to retract the skin and perform a 6mm unilateral<br />

craniotomy midway between bregma and lambda. <strong>The</strong> underlying dura was not disrupted during<br />

removal of the portion of skull. A pneumatically controlled impacting device with a 5mm<br />

beveled tip was used to injure the exposed brain, compressing the cortex at 3.5m/sec to a depth<br />

of 2mm. Following injury, Surgicel (Johnson and Johnson, Arlington, TX) was laid upon the<br />

dura, and the skull cap was replaced. After a thin coat of dental acrylic was spread over the<br />

craniotomy site and allowed to dry, the wound was stapled closed. Body temperature was<br />

monitored continuously throughout the surgical/recovery procedures and maintained at 37 o C<br />

with a heating pad.<br />

Animals processed for these experiments were killed by decapitation. Brains were rapidly<br />

removed, frozen in isopentane cooled to –40 o C on dry ice, then stored at –80 o C until sectioned.<br />

Trunk blood samples were collected on ice, and plasma was stored at –20 o C. Frozen brains were<br />

removed from a –80 o C freezer, warmed to –20 o C, and sectioned to 16 um with a Bright-Hacker<br />

cryostat. Sections were then thaw-mounted onto Ultrastick Slides (Becton Dickinson, Franklin<br />

Lakes, NJ) and stored at –20 o C until processing for in situ hybridization.<br />

In situ hybridization<br />

Antisense riboprobes complementary to coding and 3’ untranslated regions for rat MR<br />

mRNA (550 bp, courtesy of P. Patel and S. Watson, Univ. Michigan) and rat GR mRNA (456<br />

bp, courtesy of K. Yamamoto, UCSF) were labeled according to previously published methods<br />

(Seroogy and Herman, 1996) by standard in vitro transcription using [ 35 S]UTP. Slide-mounted<br />

tissue sections were removed from a -20 o C freezer and fixed in 4% paraformaldehyde for 10<br />

min. Slides were then rinsed in 5 mM KPBS, 5 mM KPBS with 0.2% glycine, 5mM KPBS, 0.1<br />

M triethanolamine with 0.25% acetic anhydride, and 0.2X SSC, then dehydrated in graded<br />

alcohols. Labeled probes were diluted in hybridization buffer containing 50% formamide, and<br />

77


50 ul of diluted probe (containing 1x 10 6 c.p.m.) was applied to each slide. Slides were then<br />

coverslipped, placed in moistened chambers, and incubated overnight at 55 o C. Following<br />

hybridization, slides were rinsed in 2X SSC and incubated in 0.2X SSC at 60 o C for 1 hr. After<br />

dehydration in graded alcohols, slides were exposed for 14-18 days to Kodak BioMAX xray<br />

film.<br />

Densitometric analysis was performed using NIH Image 1.62 software for Macintosh. A<br />

standard curve was generated using 14 C standards prior to image capture to ensure signal<br />

linearity. Based on morphological criteria, regions CA1, CA3, and DG of the hippocampal<br />

formation were manually sampled from the ipsilateral (operated) and contralateral (unoperated)<br />

hemispheres. <strong>The</strong> mean gray level of an unhybridized area (background) from each hemisphere<br />

was subtracted from sampled regions to compute corrected gray level. Ipsilateral mRNA data<br />

were normalized as a percent of control (sham/vehicle) values from the ipsilateral hemisphere,<br />

and contralateral mRNA data were normalized as a percent of control (sham/vehicle) values<br />

from the contralateral hemisphere, with sham/vehicle levels expressed as one hundred percent.<br />

Each region of each hemisphere was then analyzed by two-way analysis of variance with drug<br />

and injury designated as independent variables. Post hoc comparisons were made using the<br />

Student-Newman-Keuls test, with significance set at p ≤ 0.05.<br />

Radioimmunoassay<br />

Plasma ACTH and corticosterone concentrations were determined using ( 125 I-labeled) rat<br />

ACTH (Diasorin, Stillwater, MN, USA) and corticosterone (ICN Biochemicals, Costa Mesa,<br />

CA) kits per manufacturers’ instructions. <strong>The</strong> limits of detection for each kit are 15 pg/ml and 8<br />

ng/ml, respectively.<br />

A skewed distribution of CORT levels observed in subjects pretreated with antagonist drugs<br />

was normalized by log transformation of the data prior to analysis for significant main effects by<br />

two-way analysis of variance. Post hoc comparisons were made using the Student-Newman-<br />

Keuls test, with significance set at p ≤ 0.05. Data presented in Table 5.1 are actual CORT values<br />

listed as mean concentration ± standard error. Time course values for ACTH and CORT were<br />

analyzed for significant main effects using a repeated measures analysis of variance. Post hoc<br />

comparisons were made using the Student-Newman-Keuls test, with significance set at p ≤ 0.05.<br />

Data graphed in Figures 5.1 and 5.2 represent mean hormone concentration ± standard error.<br />

78


Results<br />

Effects of ACR blockade and CCI on plasma ACTH and corticosterone levels<br />

ACTH and CORT levels were measured 24 hours following surgery to assess the effects of<br />

ACR blockade and CCI on HPA activation. Two-way ANOVA demonstrates a significant<br />

drug/injury interaction effect on plasma CORT levels at 24h [F(2,29) = 8.02; p ≤ 0.05] (Table<br />

5.1). Post hoc comparisons indicate that SPIRO significantly increased CORT secretion in<br />

sham-operated animals. Controlled cortical impact significantly decreased CORT levels in<br />

SPIRO and RU486-pretreated animals as compared to sham-operated animals within the same<br />

drug pretreatment groups. RU486 animals also had significantly lower CORT levels after injury<br />

as compared to vehicle/injured animals.<br />

Table 5.1. Effects of adrenocorticosteroid receptor blockade and controlled cortical impact on<br />

plasma corticosterone levels (ng/ml)<br />

Group Treatment n Mean CORT Level<br />

VEHICLE SHAM 7 80.5 ± 29.9<br />

SPIRO SHAM 6 324.5 ± 77.2 a<br />

RU486 SHAM 6 121.0 ± 16.2<br />

VEHICLE INJURED 5 81.7 ± 33.6<br />

SPIRO INJURED 6 51.9 ± 26.0 b<br />

RU486 INJURED 5 14.7 ± 6.8 a,b<br />

SPIRO = MR antagonist spironolactone<br />

RU486 = GR antagonist (mifepristone)<br />

a p < 0.05 compared to vehicle-pretreated subjects in same surgery group (effect of antagonist)<br />

b p < 0.05 compared to sham-operated subjects in same pretreatment group (effect of injury)<br />

79


Plasma ACTH and corticosterone levels: time course after injury<br />

Blood samples taken 24 hours after surgery suggest the dynamic regulation of CORT<br />

secretion over the post-injury period. To examine the kinetics of HPA activation following CCI,<br />

additional animals were subjected to sham operation or injury, after which sequential blood<br />

samples were drawn by tail nick over the 24-hour post-surgery period. Plasma ACTH levels<br />

exhibited a significant injury/time interaction as determined by repeated measures analysis of<br />

variance [F(5,41) = 7.89; p ≤ 0.05] (Figure 5.1). Plasma ACTH concentrations were increased to<br />

approximately 200 pg/ml at 1.5 hours and at 18 hours post-surgery in injured animals. ACTH<br />

concentrations in sham-operated animals were indistinguishable from baseline at all time points.<br />

Plasma CORT concentrations also demonstrate a significant injury/time interaction as<br />

indicated by repeated measures analysis of variance [F(5,50) = 5.29; p ≤ 0.05] (Figure 5.2).<br />

Corticosterone levels are significantly increased in injured rats at 1.5 hours (109 ± 12 ng/ml) and<br />

at 6 hours (53 ± 4 ng/ml) following surgery, but returned to baseline by 12 hours and remained<br />

unchanged for the rest of the 24-hour period. Corticosterone concentrations in sham-operated<br />

animals were indistinguishable from baseline at all time points.<br />

80


Figure 5.1. Twenty-four hour time course analysis of plasma ACTH levels (pg/ml) following<br />

controlled cortical impact. Injury increased ACTH levels at 1.5 and 18 hours after surgery (p ≤<br />

0.05). ACTH levels in sham-operated animals were indistinguishable from baseline at all time<br />

points.<br />

Figure 5.2. Twenty-four hour time course analysis of plasma corticosterone levels (ng/ml)<br />

following controlled cortical impact. Injury increased corticosterone secretion at 1.5 and 6 hours<br />

after surgery (p ≤ 0.05); levels returned to baseline by 12 hours after injury. Corticosterone<br />

levels in sham-operated animals were indistinguishable from baseline at all time points.<br />

81


Effects of ACR blockade and CCI on ACR mRNA levels<br />

Relative changes in mRNA levels were measured by film densitometry following in situ<br />

hybridization with specific complementary antisense riboprobes (see Methods) (Figure 5.3).<br />

Figure 5.3. Photomicrographs illustrating the effects of sham operation (A,C) and controlled<br />

cortical impact (B,D) on MR (A,B) and GR (C,D) mRNA levels in hippocampus of vehiclepretreated<br />

rats. Panels A,B: Controlled cortical impact has no significant effects on MR mRNA<br />

levels in vehicle-pretreated animals. Panels C,D: Controlled cortical impact decreases ipsilateral<br />

GR mRNA levels in dentate gyrus and increases contralateral GR mRNA levels in CA1 and DG<br />

of vehicle-pretreated subjects.<br />

82


In the hemisphere contralateral to injury, two-way ANOVA demonstrates significant main<br />

effects of injury on MR mRNA levels in CA1 [F(1,28) = 13.10; p ≤ 0.05], CA3 [F(1,28) = 32.39;<br />

p ≤ 0.05], and DG [F(1,28) = 4.31; p ≤ 0.05] (Figure 5.4). Post hoc comparisons indicate that<br />

injury induces significant contralateral increases in MR mRNA levels in CA1 of SPIROpretreated<br />

animals and in CA3 of SPIRO and RU486 animals; comparisons in DG did not reach<br />

statistical significance. No significant effects of antagonist pretreatment were observed in<br />

contralateral MR mRNA levels.<br />

In the injured hemisphere, two-way ANOVA demonstrates significant main effects of injury<br />

on MR mRNA levels in CA1 [F(1,28) = 6.55; p ≤ 0.05] and CA3 [F(1,28) = 42.18; p ≤ 0.05]<br />

(Figure 5.4). Post hoc comparisons indicate that cortical impact significantly increased<br />

ipsilateral MR mRNA levels in CA1 of RU486-pretreated animals and in CA3 of SPIRO and<br />

RU486 animals. A significant main effect of drug on MR mRNA levels is observed in ipsilateral<br />

CA1 [F(2,28) = 3.68; p ≤ 0.05], but post hoc comparisons did not distinguish between individual<br />

pretreatment groups.<br />

Contralateral GR mRNA levels exhibited significant drug/injury interactions in CA1 [F(2,28)<br />

= 7.75; p ≤ 0.05] and DG [F(2,28) = 5.70; p ≤ 0.05] (Figure 5.5). In sham-operated animals,<br />

SPIRO and RU486 increased contralateral GR signal in CA1 and DG. Cortical injury increased<br />

contralateral GR mRNA levels in CA1 and DG of vehicle-pretreated animals and decreased<br />

contralateral GR mRNA levels in CA1 of RU486 animals.<br />

Main effects of drug/injury interaction were observed in ipsilateral GR mRNA levels in<br />

subregions CA1 [F(2,28) = 6.53; p ≤ 0.05], CA3 [F(2,28) = 3.71; p ≤ 0.05], and DG [F(2,28) =<br />

4.29; p ≤ 0.05] (Figure 5.5). RU486 significantly increased ipsilateral GR signal in CA1, CA3,<br />

and DG of sham-operated animals, and SPIRO increased ipsilateral GR signal in CA1. Cortical<br />

impact significantly decreased ipsilateral GR mRNA levels in CA1 and CA3 of SPIRO and<br />

RU486-pretreated animals and in DG of all antagonist pretreatment groups.<br />

83


Figure 5.4. Effects of MR blockade (SPIRO), GR blockade (RU486), and controlled cortical<br />

impact (CCI) injury on MR mRNA levels in hippocampus. Semi-quantitative analysis expressed<br />

as percentage of control (sham/vehicle) corrected gray level. Injury and antagonist pretreatments<br />

interact to increase contralateral and ipsilateral MR mRNA levels in CA1 and CA3. [ B p ≤ 0.05<br />

as compared to sham-operated animals within same pretreatment group (effect of CCI).]<br />

84


Figure 5.5. Effects of MR blockade, GR blockade, and controlled cortical impact (CCI) injury<br />

on GR mRNA levels in hippocampus. Semi-quantitative analysis expressed as percentage of<br />

control (sham/vehicle) corrected gray level. Antagonist pretreatment increases GR mRNA levels<br />

in sham-operated animals, but antagonist-induced increases in GR mRNA levels are prevented<br />

by injury. Contralateral GR mRNA levels are increased in CA1 and DG of vehicle-pretreated<br />

animals, and ipsilateral GR mRNA levels are decreased in DG of all pretreatment groups after<br />

injury. [ A p ≤ 0.05 as compared to vehicle-pretreated animals within same surgical treatment<br />

(effect of antagonist pretreatment). B p ≤ 0.05 as compared to sham-operated animals within<br />

same pretreatment group (effect of CCI).]<br />

85


Discussion<br />

CCI modulates HPA activation<br />

<strong>The</strong> extent of HPA activation after TBI may influence neuronal survival, given evidence for<br />

glucocorticoid exacerbation of several forms of neuronal injury (Sapolsky, 1985; Sapolsky and<br />

Pulsinelli, 1985; Sapolsky, 1986a; Sapolsky et al., 1988; Goodman et al., 1996; McIntosh and<br />

Sapolsky, 1996a). In the first experiment of this study, vehicle-pretreated animals demonstrated<br />

equivalent plasma CORT levels 24 hours after sham operation and injury. <strong>The</strong> levels were<br />

slightly higher than would be expected in unoperated rats but were not of a magnitude to saturate<br />

GR binding at the 24-hour time point. Another study assessing ACTH and CORT levels in rats<br />

following closed head injury reported similar results at 24 hours (Shohami et al., 1995).<br />

Clinical studies have reported increased HPA activation in humans after head injury (King et<br />

al., 1970; Pentelenyi and Kammerer, 1984). To determine if differential activation of the HPA<br />

axis occurs during the first 24 hours after experimental CCI, a time course analysis of plasma<br />

CORT and ACTH levels was performed in sham-operated versus injured rats. Blood sampling<br />

over the 24-hour period following surgeries confirmed the differential activation of the HPA axis<br />

in animals injured by CCI. Increased levels of ACTH and CORT were observed 1.5 hours<br />

following injury, after which a prolonged, low level of CORT secretion was observed for at least<br />

6 hours after injury.<br />

Plasma ACTH and CORT levels six hours after surgery (time of day = 1800h) indicate a<br />

disruption of normal circadian rhythms. Diurnal peaks in rodent ACTH and CORT secretion<br />

normally occur near the onset of darkness (Ruhmann-Wennhold and Nelson, 1977; D'Agostino<br />

et al., 1982; Kwak et al., 1992). ACTH levels in CCI-injured animals were indistinguishable<br />

from baseline at 1800h, the time of lights-off, and plasma CORT concentrations, although<br />

slightly elevated at 1800h, did not approach typical peak levels of 150-200 ng/ml (D'Agostino et<br />

al., 1982; Kwak et al., 1992). Sham-operated animals also lacked ACTH and CORT rhythms,<br />

exhibiting unchanged plasma hormone levels across the 24-hour period after surgery. Alteration<br />

of circadian rhythms has been reported after surgery in rats (Levine et al., 1980) and in humans<br />

(King et al., 1970; McIntosh et al., 1981; Pentelenyi and Kammerer, 1984), but the exact cause<br />

for disruption in these subjects, whether by anesthesia or the trauma of craniotomy, remains to be<br />

determined.<br />

86


Injured animals demonstrated an unusual peak in ACTH levels 18h after the onset of surgery<br />

(time of day = 600h). While the significance of this increase is unclear, we speculate that<br />

increased ACTH levels could result from decreased sensitivity to glucocorticoid-mediated<br />

feedback. Decreased sensitivity to feedback normally occurs at the peak of the circadian rhythm<br />

and after stress (Dallman, 1992) and could be precipitated by the stress of injury or alteration of<br />

circadian rhythms.<br />

ACR antagonists modulate plasma corticosterone levels<br />

Blockade of MR-mediated feedback inhibition of the HPA axis may account for increased<br />

plasma CORT levels observed in SPIRO/sham subjects (Table 5.1). Our laboratory has<br />

previously demonstrated increased basal CORT secretion with SPIRO administration (<strong>McCullers</strong><br />

and Herman, 2001) (Chapter Three), although some studies report no effect of SPIRO on<br />

morning CORT levels (Chabert et al., 1984; Herman and Spencer, 1998; Yau et al., 1999).<br />

Activation of MR is required to maintain CORT secretion at the low basal levels normally<br />

observed in rats in the morning (Spencer et al., 1998). Furthermore, studies using the MR<br />

antagonist RU28318 have consistently demonstrated that MR blockade increases adrenocortical<br />

secretion (Ratka et al., 1989; van Haarst et al., 1997; Spencer et al., 1998). <strong>The</strong> mechanism<br />

whereby CCI interacts with ACR antagonists to reduce CORT secretion is presently unknown.<br />

<strong>The</strong> inhibitory influence of the hippocampus over HPA activation (Herman et al., 1989b;<br />

Jacobson and Sapolsky, 1991) invites speculation that this structure may be involved. It is<br />

possible that aberrant hippocampal activation subsequent to injury in combination with altered<br />

ACR signaling might culminate in decreased HPA secretion, but support for this hypothesis will<br />

require further study.<br />

ACR antagonists regulate GR mRNA levels<br />

SPIRO pretreatment did not alter MR mRNA levels in sham-operated animals (Figure 5.4).<br />

Previously we have observed that SPIRO pretreatment (followed by i.p. saline injection)<br />

increases MR mRNA levels by approximately 30-40% in CA1, CA3, and DG (<strong>McCullers</strong> and<br />

Herman, 2001) (Chapter Three), and many preceding studies suggest that MR is subject to autoregulation<br />

(Herman et al., 1989a; Reul et al., 1989; Herman et al., 1993; Chao et al., 1998a). <strong>The</strong><br />

lack of MR mRNA upregulation observed in the SPIRO/sham-operated group may be<br />

87


attributable to significantly increased CORT secretion in these subjects (Table 5.1). MR mRNA<br />

levels are negatively regulated by GR activation at high glucocorticoid levels (Herman et al.,<br />

1989a; Herman et al., 1993). Thus, enhanced GR activation due to elevated CORT levels in<br />

SPIRO/sham animals may attenuate the transcriptional effects of MR blockade.<br />

Both MR antagonist SPIRO and GR antagonist RU486 increased GR mRNA levels in shamoperated<br />

animals (Figure 5.5). We have observed similar effects with administration of these<br />

antagonists in other experiments (Herman and Spencer, 1998; <strong>McCullers</strong> and Herman, 2001)<br />

(Chapter Three). <strong>The</strong>se data, together with reports that GR agonist dexamethasone (Herman et<br />

al., 1989a; Reul et al., 1989), MR agonist aldosterone, or low, MR binding doses of CORT<br />

(Chao et al., 1998a) each reverse ADX-induced increases in GR mRNA provide compelling<br />

evidence for regulation of hippocampal GR mRNA expression by both MR and GR.<br />

CCI increases MR mRNA and decreases GR mRNA levels<br />

Injury increased MR mRNA levels in CA1 and CA3 of MR and GR antagonist-pretreated<br />

animals (Figure 5.4). Increased MR mRNA levels in SPIRO-pretreated subjects may be<br />

attributable to MR auto-regulation (Herman et al., 1989a; Reul et al., 1989; Herman et al., 1993;<br />

Chao et al., 1998a); however, previous experiments in our laboratory have not indicated MR<br />

mRNA upregulation with RU486 pretreatment (<strong>McCullers</strong> and Herman, 2001) (Chapter Three).<br />

<strong>The</strong>se results suggest that TBI influences ACR-mediated regulation of MR expression. It is<br />

interesting to note that CORT levels are significantly lower in these groups as compared to<br />

sham-operated animals (Table 5.1). <strong>The</strong>se data raise the possibility that injury-induced CORT<br />

secretion could influence MR mRNA levels through auto-regulatory processes.<br />

Injury did not alter ipsilateral GR mRNA expression in CA1 and CA3 of vehicle-pretreated<br />

animals (Figure 5.5). However, injury abolished autoregulatory effects of SPIRO and RU486,<br />

sharply decreasing ipsilateral GR mRNA levels (as compared to sham) by approximately 50-<br />

80% in antagonist-pretreated subjects. <strong>The</strong>se results indicate that CCI suppresses or interferes<br />

with GR autoregulatory processes in hippocampus. Our laboratory has reported similar effects<br />

of excitotoxic insult with kainic acid (<strong>McCullers</strong> and Herman, 2001) (Chapter Three).<br />

While it is clear that GR auto-regulation (Herman et al., 1989a; Reul et al., 1989) is disrupted<br />

after injury, the extent to which GR mRNA levels may be further decreased due to cell loss is<br />

presently ambiguous. Nonetheless, decreases in GR signal in the ipsilateral DG of all<br />

88


pretreatment groups surpass the magnitude of cell loss experienced in this region (<strong>McCullers</strong> et<br />

al., 2001b) (Chapter Four). We have observed similar GR mRNA decreases of approximately<br />

40-55% in DG of vehicle, SPIRO, and RU486 pretreated animals 24h after excitotoxic insult<br />

with kainic acid, at a dose which does not cause cell loss in DG (<strong>McCullers</strong> and Herman, 2001)<br />

(Chapter Three). Thus, multiple types of neuronal injury result in the significant downregulation<br />

of GR mRNA levels in DG.<br />

<strong>The</strong> extent to which cell loss affects measurement of mRNA levels by in situ hybridization is<br />

an important consideration that is presently undefined. In parallel subjects, CCI resulted in the<br />

loss of approximately 40% of CA1 and CA3 pyramidal neurons and 10% of DG granule neurons<br />

24 hours after injury, while no cell loss was observed in the contralateral hemisphere of injured<br />

rats or in sham-operated subjects (<strong>McCullers</strong> et al., 2001b) (Chapter Four). If comparable loss is<br />

assumed in this study, the lack of demonstrable GR mRNA decrement in injured, vehicle-treated<br />

groups indicates that the in situ hybridization technique underestimates cell loss. <strong>The</strong> lack of<br />

correlation between cell loss and mRNA level is also seen for MR, where levels are significantly<br />

higher on the injured side. It is possible that GR mRNA is upregulated in CA1 and CA3 to<br />

compensate for signal that would otherwise be decreased by cell loss; however, the increases in<br />

MR mRNA argue against this explanation, as greater induction of MR mRNA would have to be<br />

assumed. In conjunction with the differential effects of antagonist pretreatment, it is likely that<br />

the observed levels of GR and MR are due to injury-related gene regulation rather than simple<br />

cell loss. This notion is supported by the large and disproportionate loss of GR mRNA in DG,<br />

the region showing least cell death following CCI.<br />

Traumatic brain injury initiates a multitude of cellular and molecular cascades that influence<br />

gene expression. Although it is difficult to isolate specific factors of the many that are likely to<br />

influence neuronal GR expression, results from the current study and others (Lowy, 1990, 1992;<br />

<strong>McCullers</strong> and Herman, 2001) (Chapter Three) suggest a role for glutamate in the regulation of<br />

hippocampal GRs. Extracellular glutamate levels increase in hippocampus after CCI injury<br />

(Faden et al., 1989) and kainic acid injection (Liu et al., 1997), and similar patterns of GR<br />

mRNA downregulation are observed in hippocampus after both types of injury (<strong>McCullers</strong> and<br />

Herman, 2001) (Chapter Three). Thus, excessive glutamate receptor activation may lead to<br />

interference with antagonist-mediated autoregulatory effects and to the preferential<br />

downregulation of GR mRNA in DG. Glutamate receptor activation and alterations in calcium<br />

89


signaling that accompany CCI (McIntosh et al., 1998c) could potentially regulate GR mRNA<br />

levels via activation of transcription factors (Cabral et al., 2001), expression of neurotrophic<br />

factors (Sarrieau et al., 1996), or altered neurotransmitter release (Seckl et al., 1990). <strong>The</strong> results<br />

of this study also indicate that glucocorticoid secretion is increased immediately following CCI,<br />

potentially exposing hippocampal neurons to endangering levels of hormone. Considering the<br />

negative effects of excessive GR activation on injured hippocampal neurons, injury modulation<br />

of HPA activation and ACR expression may influence hippocampal neuron viability after<br />

traumatic injury.<br />

90


Chapter Six: Discussion<br />

Adrenocorticosteriod receptor regulation of hippocampal neuron viability<br />

<strong>The</strong> present thesis examines the respective roles of MR and GR on hippocampal neuron<br />

vulnerability to injury. To isolate individual receptor effects on neuronal survival, MR and GR<br />

antagonists were administered in subjects prior to excitotoxic challenge with KA or traumatic<br />

injury by CCI. Cell counts performed twenty-four hours following insult tested the hypotheses<br />

that MR blockade would increase and GR blockade would decrease hippocampal neuron<br />

vulnerability to each form of injury. In addition, expression of bcl-2 and other viability-related<br />

genes was examined following injury to determine if ACR antagonist effects on viability<br />

correlate with regulation of the selected genes. Finally, MR and GR mRNAs were measured<br />

after injury to test the hypothesis that glucocorticoid signaling is altered following injury via<br />

regulation of hippocampal ACR expression.<br />

Adrenocorticosteriod receptor regulation of hippocampal neuron survival<br />

<strong>The</strong> results of these experiments support the hypotheses that MR protects and GR endangers<br />

challenged hippocampal neurons. In support of the latter, GR antagonist RU486 prevented the<br />

loss of CA1 neurons twenty-four hours following CCI injury (Chapter Four). <strong>The</strong> selective<br />

protection of CA1 neurons may be due to regional differences in CCI injury or density of GR<br />

expression (see Chapter Four). <strong>The</strong> potential for RU486 neuroprotection against KA<br />

excitotoxicity could not be evaluated due to the absence of measurable KA-induced injury in<br />

vehicle and RU486-pretreated subjects (Chapter Three). <strong>The</strong> current work predicts that RU486<br />

would protect hippocampal neurons subjected to a more severe KA lesion.<br />

Pretreatment with SPIRO, an MR antagonist, increased vulnerability to KA excitotoxicity in<br />

hippocampal field CA3 (Chapter Three). While these results may suggest selective MR<br />

protection of this region, the comparable distribution of MR in CA1 and CA3 suggests that<br />

differential cell loss is more likely attributable to the selective vulnerability of CA3 neurons to<br />

KA insult (Nadler et al., 1978). Whether MR blockade increases susceptibility to KA by<br />

increasing injury severity or by lowering the neuronal threshold for cell death remains unclear.<br />

MR blockade might exacerbate neuronal injury by disrupting cellular Ca2+ homeostasis.<br />

91


Adrenalectomy, which prevents endogenous CORT production and subsequent MR activation,<br />

increases Ca 2+ currents in CA1 neurons, while selective activation of MR results in very small<br />

Ca 2+ currents (Karst et al., 1994). Predominant MR activation may therefore reduce harmful<br />

Ca 2+ influx (Karst et al., 1994), whereas MR blockade may contribute to pathologic rises in<br />

intracellular Ca 2+ . Alternatively, MR blockade may lower the neuronal “set point” for cell death<br />

by modulating expression of pro- or anti- apoptotic factors, such as members of the bcl-2 family.<br />

Following TBI, SPIRO did not exacerbate neuronal loss in CA1 and CA3. However, these<br />

regions suffered substantial (35-40%) injury-induced neuronal loss in vehicle and SPIROpretreated<br />

subjects. Because CCI alone was sufficient to destroy vulnerable cells, compromising<br />

effects of MR blockade may not have been evident at this level of injury.<br />

Adrenocorticosteriod receptor regulation of viability-related gene expression<br />

<strong>The</strong> current data indicate potential for ACR influence on neuronal survival via regulation of<br />

the anti-apoptotic gene bcl-2. MR blockade with SPIRO decreased basal mRNA levels of the<br />

survival-promoting gene bcl-2 in CA1 (Chapter Two), suggesting that SPIRO would increase<br />

hippocampal vulnerability to insult in this region. However, subjects receiving KA injection<br />

following SPIRO pretreatment exhibited no differential cell loss in CA1 (Chapter Three). In<br />

addition, RU486 decreased bcl-2 levels in CA1 and DG of animals subjected to sham-CCI, yet<br />

RU486 had a dramatic neuroprotective effect on CA1 neurons after CCI injury (Chapter Four).<br />

Thus, ACR-induced decreases in bcl-2 mRNA expression observed in uninjured (saline-injected<br />

or sham-operated) animals did not consistently predict increased vulnerability following injury.<br />

Experiments in the present thesis do not provide support for bax, p53, BDNF, or NT-3<br />

involvement in ACR modulation of hippocampal survival. Adrenocorticosteriod regulation of<br />

hippocampal bax, BDNF, or NT-3 mRNA levels was not observed in any of the current<br />

experiments (note that BDNF and NT-3 mRNA levels were not assessed in the CCI model). In<br />

addition, ACR blockade had no significant effects on hippocampal p53 mRNA levels in CCIinjured<br />

hemispheres (Chapter Four). Kainic acid-induced increases in p53 mRNA levels were<br />

attenuated by MR blockade in CA3, suggesting the potential for enhanced resistance to apoptosis<br />

in this region, yet MR blockade increased CA3 vulnerability to KA excitotoxicity (Chapter<br />

Three). <strong>The</strong>refore, ACR-mediated effects on p53 expression do not appear to have significant<br />

influence on hippocampal neuron survival.<br />

92


Adrenocorticosteriod receptor regulation of hippocampal neuron survival - potential mechanisms<br />

Adrenocorticosteriod receptor modulation of calcium homeostasis<br />

Altered Ca 2+ homeostasis is proposed to be a key cellular mechanism underlying<br />

glucocorticoid neurotoxicity. Voltage-dependent Ca 2+ currents are strongly regulated by<br />

corticosteroids with U-shaped dose dependency; in CA1 of hippocampus, Ca 2+ current amplitude<br />

is large in the absence of corticosteroids, small under conditions of predominant MR activation,<br />

and large with the additional occupation of GR at higher glucocorticoid levels (Kerr et al., 1992;<br />

Karst et al., 1994; Werkman et al., 1997). Evidence suggests that neuronal Ca 2+ influx is<br />

enhanced by glucocorticoid modulation of L-type Ca 2+ currents (Landfield and Eldridge, 1994;<br />

Joels, 2001). In combination with depolarizing stimuli such as KA-induced seizures or excessive<br />

levels of extracellular glutamate, GR-mediated increases in intracellular Ca 2+ levels may enhance<br />

neuronal vulnerability to Ca 2+ -related injury mechanisms (Landfield and Eldridge, 1991; Joels,<br />

2001).<br />

In addition to increasing Ca 2+ influx, GR activation reduces Ca 2+ extrusion in hippocampal<br />

neurons (Elliott and Sapolsky, 1993; Joels, 2001). Consistent with these findings, high CORT<br />

levels reduce expression of PMCA1 (Bhargava et al., 2000), a plasma membrane Ca 2+ ATPase<br />

responsible for translocating calcium from the cytosol to the extracellular space (Garcia and<br />

Strehler, 1999). Overexpression of PMCA isoform 4 in pheochromocytoma 12 cells reduces<br />

vulnerability to Ca 2+ -mediated cell death, suggesting that regulation of PMCA expression may<br />

influence cellular survival of insults involving pathological increases in intracellular calcium<br />

(Garcia et al., 2001). In agreement with this hypothesis, KA administration decreases<br />

hippocampal expression of PMCA isoforms 1 and 2 (Garcia et al., 1997b). Further reduction in<br />

PMCA levels prompted by GR activation would be expected to intensify injury-induced rises in<br />

intracellular Ca 2+ levels and exacerbate neuronal degeneration.<br />

Adrenocorticosteriod receptor blockade with MR or GR antagonists may influence<br />

hippocampal neuronal viability by altering Ca 2+ conductances or extrusion mechanisms.<br />

Blocking MR activation with SPIRO should result in the predominant activation of GR. It is<br />

well established that voltage-dependent Ca 2+ currents increase with GR activation at high CORT<br />

levels (Kerr et al., 1992; Karst et al., 1994; Werkman et al., 1997). However, it is currently<br />

93


unclear whether exclusive GR activation is sufficient to induce large Ca 2+ currents in<br />

hippocampal neurons. Electrophysiological studies in hippocampal slices have reported mixed<br />

results; one study indicates that GR agonist RU28362 increases Ca 2+ currents in CA1 neurons<br />

(Kerr et al., 1992) while another indicates that it does not (Karst et al., 1994). If exclusive GR<br />

activation is indeed sufficient to increase Ca 2+ currents, then SPIRO blockade of MR may<br />

endanger hippocampal neurons by enhancing Ca 2+ influx, particularly under conditions of GRsaturating<br />

CORT levels. In terms of the current experiments, this scenario is consistent with the<br />

increased vulnerability of CA3 neurons to KA toxicity following SPIRO pretreatment (Chapter<br />

Three).<br />

Glucocorticoid receptor blockade with RU486 should result in the predominant activation of<br />

MR. Exclusive MR occupation yields relatively small Ca 2+ currents (Karst et al., 1994) and may<br />

thus protect hippocampal neurons from exposure to excessive Ca 2+ influx. In addition, GR<br />

blockade should prevent GR-mediated repression of PMCA isoforms, improving Ca 2+ extrusion<br />

mechanisms. <strong>The</strong>se beneficial effects of GR blockade on neuronal Ca 2+ homeostasis are<br />

consistent with the neuroprotective effects of RU486 on CA1 neurons following TBI (Chapter<br />

Four).<br />

Adrenocorticosteriod receptor suppression of inflammation<br />

Glucocorticoids inhibit synthesis, release, and/or function of many factors involved in the<br />

inflammatory response, including: cytokines, such as interleukin (IL)-1β, IL-2, IL-3, IL-5, IL-6,<br />

IL-12, and tumor necrosis factor alpha (TNF-α); chemokines such as IL-8; and inflammatory<br />

mediators and enzymes such as histamine, bradykinin, and nitric oxide (Cato and Wade, 1996;<br />

Steer et al., 1997; Sapolsky et al., 2000). Glucocorticoids also repress expression of<br />

cyclooxygenase 2 (COX-2) (Cato and Wade, 1996; Sapolsky et al., 2000), an enzyme that<br />

catalyzes prostaglandin formation (Rivest, 1999). Further, glucocorticoids inhibit expression of<br />

adhesion molecules, alter lymphocyte trafficking, and reduce activation and proliferation of B<br />

and T cells (Cato and Wade, 1996; Sapolsky et al., 2000). Glucocorticoid inhibition of KAinduced<br />

AP-1 (Unlap and Jope, 1995a) and NF-κB (Unlap and Jope, 1995b) DNA binding has<br />

been demonstrated, suggesting that the potent anti-inflammatory and immunomodulatory action<br />

of glucocorticoids may be mediated in part through repression of AP-1 and NF-κB activity (De<br />

Bosscher et al., 2000).<br />

94


Inflammation has been proposed to influence neuronal survival following several forms of<br />

insult including KA-mediated excitotoxicity and traumatic brain injury. Kainic acid increases<br />

glial expression of several cytokines in hippocampus and cortex including IL-1β (Minami et al.,<br />

1991; Yabuuchi et al., 1993; Eriksson et al., 2000). Administration of IL-1ra, an endogenous<br />

antagonist of IL-1, protects neurons against KA-induced degeneration in CA1 and CA3,<br />

suggesting a deleterious role for IL-1β in excitotoxic neuronal death (Panegyres and Hughes,<br />

1998). Hippocampal and cortical mRNA levels of IL-1β also increase following experimental<br />

TBI (Fan et al., 1995). Administration of IL-1 receptor antagonist IL-1ra has been shown to<br />

reduce cortical lesion volume (Toulmond and Rothwell, 1995), attenuate hippocampal neuron<br />

loss, and improve cognitive function following lateral fluid percussion (Sanderson et al., 1999),<br />

providing further evidence that IL-1β may contribute to injury-mediated neurodegeneration. As<br />

glucocorticoids decrease IL-1 production (McEwen et al., 1997), ACR regulation of this<br />

cytokine should enhance neuronal survival after injury and would not appear to explain the<br />

endangering effects of GR.<br />

Other studies raise the possibility that mediators of inflammation may play a neuroprotective<br />

role after neuronal injury. Mice lacking IL-6 demonstrate increased susceptibility to KAinduced<br />

damage in hippocampus, suggesting that IL-6 can improve survival after excitotoxic<br />

injury (Penkowa et al., 2001). In addition, selective inhibitors of COX-2 exacerbate KA-induced<br />

neuronal degeneration in hippocampus (Baik et al., 1999) and worsen motor performance<br />

following TBI (Dash et al., 2000), indicating a protective role for COX-2. Glucocorticoids<br />

decrease IL-6 production (McEwen et al., 1997) and inhibit COX-2 expression (Masferrer et al.,<br />

1992), raising the possibility that the anti-inflammatory effects of glucocorticoids may contribute<br />

to neuronal vulnerability to injury.<br />

Glucocorticoids may also increase neuronal vulnerability to injury via suppression of TNF-α,<br />

a pro-inflammatory cytokine that is rapidly induced after several types of brain injury including<br />

KA excitotoxicity (Minami et al., 1991; de Bock et al., 1996) and TBI (Taupin et al., 1993; Fan<br />

et al., 1996; Knoblach et al., 1999). <strong>The</strong> role for TNF-α in mediating neuronal survival is<br />

controversial; several studies indicate that inhibition of TNF-α reduces neurological deficits and<br />

hippocampal damage following TBI (Shohami et al., 1996; Shohami et al., 1997; Knoblach et al.,<br />

1999). However, increasing evidence suggests that TNFα may have neuroprotective effects<br />

under certain circumstances following central nervous system insult. Pretreatment with TNF-α<br />

95


protects cultured hippocampal neurons against oxidative insults (Barger et al., 1995; Mattson et<br />

al., 1997), excitatory amino acid toxicity (Cheng et al., 1994), and injury induced by glucose<br />

deprivation (Cheng et al., 1994). In addition, TNF-deficient mice exhibit long-lasting motor<br />

deficits following CCI brain injury as compared to wild-type (despite increased memory and<br />

motor deficits within the acute posttraumatic period), suggesting that TNF may facilitate longterm<br />

recovery after brain injury (Scherbel et al., 1999). Further, mice lacking TNF receptors are<br />

more susceptible to neuronal loss induced by focal cerebral ischemia and KA administration<br />

(Bruce et al., 1996). TNF-α appears to protect neurons by stabilizing cellular Ca 2+ homeostasis<br />

and stimulating antioxidant pathways via activation of NF-κB (Cheng et al., 1994; Barger et al.,<br />

1995; Bruce et al., 1996; Mattson et al., 1997; Furukawa and Mattson, 1998). Glucocorticoids<br />

may counter these potentially neuroprotective effects through suppression of TNF-α release from<br />

monocytes and macrophages (Joyce et al., 1997; Steer et al., 1997) and inhibition of NF-κB<br />

activation (De Bosscher et al., 2000).<br />

Glucocorticoids have been proposed to protect the organism by limiting inflammatory<br />

responses that, left unchecked, might threaten cellular homeostasis (Munck et al., 1984). In this<br />

view, it is possible that glucocorticoids prevent excessive activation of mediators of<br />

inflammation that would otherwise contribute to neuronal degeneration after injury. However, it<br />

is also possible that glucocorticoid suppression of the beneficial effects of inflammation, such as<br />

Ca 2+ regulation and the expression of antioxidants (Wong and Goeddel, 1988; Cheng et al.,<br />

1994), may contribute to neuronal degeneration. Administration of RU486 would be expected to<br />

attenuate the anti-inflammatory effects of GR activation, as RU486 has been shown to reverse<br />

GR suppression of TNFα release from monocytes (Steer et al., 1997). Thus, RU486 repression<br />

of anti-inflammatory glucocorticoid effects may contribute to the improved survival of<br />

hippocampal neurons after injury.<br />

Adrenocorticosteriod receptor regulation of hippocampal neurotrophic factor expression<br />

Despite negative findings reported in the current work, the possibility for glucocorticoid<br />

modulation of neuronal viability via regulation of neurotrophic factors can not be dismissed.<br />

Due to the twenty-six hour time period between antagonist injection and decapitation, ACR<br />

antagonist effects on BDNF or NT-3 mRNA levels may not have been detected. Glucocorticoid<br />

regulation of BDNF expression is rapid and transient; hippocampal BDNF mRNAs are reduced<br />

96


within one hour after the onset of stress, recover substantially by four hours, and return towards<br />

baseline levels within twenty-four hours (Smith et al., 1995b). Similarly, hippocampal nerve<br />

growth factor mRNA levels peak approximately three hours after CORT injection, fall below<br />

baseline levels by six hours, and return close to control levels by 24 hours (Lindholm et al.,<br />

1994), emphasizing the transitory nature of growth factor induction. In addition, neurotrophin<br />

levels were not assessed in sham and CCI-injured animals in the current work, and ACR effects<br />

on numerous other trophic factors remain to be investigated.<br />

Several previous studies validate the possibility that ACR regulation of neurotrophic factors<br />

may influence hippocampal vulnerability to injury. Although reports are not entirely consistent,<br />

there is general agreement that in specific hippocampal subfields, CORT and stress decrease<br />

hippocampal BDNF mRNA levels, while adrenalectomy increases BDNF mRNAs and decreases<br />

NT-3 mRNAs in hippocampus (Chao and McEwen, 1994; Smith et al., 1995a, b; Schaaf et al.,<br />

1997; Chao et al., 1998b). Futher, CORT and dexamethasone have been shown to increase nerve<br />

growth factor mRNA levels (Lindholm et al., 1994; Scully and Otten, 1995). <strong>The</strong>re is also<br />

evidence to indicate that glucocorticoids regulate neurotrophic factor expression following<br />

injury. Adrenalectomy attenuates BDNF mRNA induction and nearly abrogates NGF mRNA<br />

induction by KA, and GR agonist dexamethasone enhances NGF mRNA induction by KA<br />

(Barbany and Persson, 1993). In cultured hippocampal neurons, KA-induced increases in BDNF<br />

mRNAs are negated by dexamethasone (Cosi et al., 1993). Furthermore, serum deprivationinduced<br />

death of cultured hippocampal neurons is accelerated by CORT and associated with<br />

decreases in BDNF mRNA levels and protein (Nitta et al., 1999). Addition of BDNF to serumdeprived<br />

cultures attenuates negative CORT effects, supporting the hypothesis that reduced<br />

synthesis of BDNF contributes to CORT-mediated endangerment of hippocampal neurons (Nitta<br />

et al., 1999). A recent study has also demonstrated that glucocorticoids influence neurotrophic<br />

factor expression following TBI. Adrenalectomy enhances fluid percussion injury-induced<br />

increases in BDNF mRNA levels, an effect that is reversed with administration of CORT<br />

(Grundy et al., 2000). Adrenalectomy also prevents and CORT replacement permits the<br />

induction of NGF mRNA levels following fluid-percussion injury, indicating the essential role of<br />

CORT in the NGF response (Grundy et al., 2001).<br />

Despite evidence for glucocorticoid regulation of neurotrophic factor expression following<br />

injury, the potential impact of this regulation on hippocampal neuronal viability is presently<br />

97


unclear. High glucocorticoid levels decrease BDNF mRNAs and increase pyramidal neuron<br />

vulnerability to injury (Sapolsky, 1994; Smith et al., 1995b), suggesting reduced BDNF<br />

expression as a potential mechanism for glucocorticoid toxicity. Reports demonstrating an<br />

association between decreased BDNF mRNA levels and increased neuronal degeneration support<br />

this hypothesis (Hicks et al., 1999; Nitta et al., 1999). In contrast, the attenuation of KA-induced<br />

increases in BDNF expression by adrenalectomy is not consistent with a protective role for<br />

glucocorticoid removal (Barbany and Persson, 1993). <strong>The</strong> neuroprotective value of BDNF is<br />

also uncertain, as exogenous administration of BDNF has been shown to exacerbate CA3<br />

neuronal loss following KA administration (Rudge et al., 1998). Furthermore, glucocorticoids<br />

potentiate NGF expression and maintain NT-3 mRNA levels, in seeming disagreement with the<br />

profile for neuronal endangerment by glucocorticoids. <strong>The</strong> literature suggests that blockade of<br />

GR with RU486 should enhance KA- and TBI-mediated induction of BDNF expression, while<br />

decreasing injury-induced NGF expression. However, the potential effects MR blockade on<br />

injury-induced neurotrophin expression are less clear because it has not been established whether<br />

GR activation alone is sufficient for neurotrophin regulation. Clarification of these issues will<br />

require further study.<br />

Summary<br />

<strong>The</strong> present thesis provides evidence supporting the hypotheses that MR protects and GR<br />

endangers hippocampal neurons following specific types of neuronal injury. <strong>The</strong> current<br />

experiments did not support the hypotheses that ACR effects on neuronal survival involve the<br />

regulation of bcl-2, bax, or p53 expression. Rather, the widespread effects of glucocorticoids on<br />

several other systems crucial to neuronal survival suggest that glucocorticoids influence<br />

hippocampal neuron viability via more global mechanisms. Glucocorticoids exert direct effects<br />

on cellular metabolism, Ca 2+ homeostasis, inflammatory processes, and neurotrophic factor<br />

expression; in addition, these targets are largely interrelated and capable of altering one another<br />

(Figure 6). For instance, glucocorticoid effects on neurotrophin expression, cytokine production,<br />

and cellular metabolism can influence intracellular Ca 2+ levels (Sapolsky, 1990; Mattson and<br />

Scheff, 1994; Tymianski and Tator, 1996; Jiang and Guroff, 1997). In turn, Ca 2+ influx can<br />

modulate neurotrophin expression (Finkbeiner, 2000), regulate cytokine production (Hotchkiss<br />

and Karl, 1996), and disrupt mitochondrial production of ATP (Green and Reed, 1998). It<br />

98


ecomes clear that glucocorticoid effects on hippocampal neuron viability are not likely carried<br />

out via isolated signaling pathways, but rather they result from the integration of many factors<br />

associated with the injury response.<br />

Figure 6. Potential mechanisms for glucocorticoid modulation of hippocampal neuron survival<br />

after injury. Excessive activation of the glucocorticoid receptor may impair antioxidant<br />

defenses, suppress neuroprotective inflammatory responses, decrease neuronal energy stores,<br />

increase voltage-dependent Ca 2+ currents, and compromise ATP-dependent processes<br />

responsible for maintaining membrane potential. All of these factors may culminate to increase<br />

oxidative stress and intracellular Ca 2+ levels, potentially leading to neuronal death.<br />

99


Injury regulation of adrenocorticosteroid receptor mRNA levels<br />

Experiments in the current thesis reveal that hippocampal ACR mRNA levels are<br />

dramatically regulated by excitotoxic and traumatic injury (Chapters 6 and 8). Kainic acid<br />

decreased MR mRNA levels by approximately 40% in CA1 and CA3 and decreased GR mRNA<br />

levels by approximately 40% in DG (Chapter Three). Traumatic injury did not influence MR<br />

mRNA levels but had powerful effects on GR levels; CCI decreased GR levels by approximately<br />

30% in DG of the injured hemisphere and increased GR levels by approximately 20% in CA1<br />

and DG of the contralateral hemisphere (Chapter Five). <strong>The</strong>se data support the hypothesis that<br />

glucocorticoid signaling is altered through regulation of ACR expression following injury.<br />

Injury regulation of adrenocorticosteroid receptor mRNA levels – potential mechanisms<br />

As discussed in Chapters Three and Five, altered ACR expression following KA<br />

administration and CCI is not attributable to CORT-mediated auto-regulatory mechanisms. Still,<br />

the downregulation of dentate gyrus GR mRNA levels by both types of injury suggests the<br />

possibility of a common regulatory mechanism. Potential candidates for this role include<br />

glutamate receptor activation and Ca 2+ signaling, as both KA-induced seizures and TBI lead to<br />

the accumulation of extracellular glutamate and increased intracellular Ca 2+ levels (Ben-Ari,<br />

1985; Faden et al., 1989; Sun et al., 1992; McIntosh et al., 1998c).<br />

Glutamate regulation of adrenocorticosteriod receptor mRNA levels<br />

Receptor expression patterns provide no evidence for differential action of ionotropic<br />

glutamate receptors in DG, as NMDA and AMPA glutamate receptor subtypes are preferentially<br />

expressed in CA1, and KA receptors are mainly located in stratum lucidum of CA3 (Foster et al.,<br />

1981; Represa et al., 1987). However, metabotropic glutamate receptor subtype 1 (mGluR1) is<br />

strongly expressed in granule neuron cell bodies and the molecular layer of dentate gyrus<br />

(Shigemoto et al., 1997), and mGluR1 expression is increased in dentate gyrus following<br />

systemic KA injection (Blumcke et al., 2000). <strong>The</strong> literature provides support for mGluR<br />

activation as a viable candidate mechanism for regulation of GR expression after injury. Group<br />

1 mGluRs, which include mGluR1 and mGluR5, stimulate formation of cyclic adenosine<br />

monophosphate (cAMP) (Aramori and Nakanishi, 1992), which has been shown to decrease GR<br />

100


expression in multiple cell lines. <strong>The</strong> cAMP analog 8-bromo-cAMP decreases GR binding in<br />

mouse AtT20 anterior pituitary corticotrope tumor cells, and forskolin, a drug that stimulates<br />

adenylate cyclase, decreases GR binding and mRNA levels (Sheppard et al., 1991). In addition,<br />

treatment with dibutyryl cAMP, another cAMP analog, decreases GR protein levels in rat C6<br />

glioma cells, B104 neuroblastoma cells, and in rat primary hippocampal and cortical cultures<br />

(Sipe, 2000). Thus, activation of mGluR1 represents a possible mechanism whereby injury may<br />

decrease GR expression in DG via cAMP-dependent pathways.<br />

Calcium regulation of adrenocorticosteriod receptor mRNA levels<br />

Intracellular Ca 2+ -dependent signaling may also mediate injury-induced changes in ACR<br />

expression. Calcium influx (Young, 1992; Fineman et al., 1993; Tymianski and Tator, 1996) has<br />

been hypothesized to regulate changes in gene expression induced by both kainate (Pennypacker<br />

and Hong, 1995) and TBI (McIntosh et al., 1997; Morrison et al., 2000). Furthermore, Ca 2+<br />

regulation of GR expression has been demonstrated in vitro in AtT-20 cells (Sheppard, 1994).<br />

However, the lack of Ca 2+ accumulation in DG following KA injection (in contrast to significant<br />

Ca 2+ accumulation in CA3) (Friedman, 1997) and the high density of Ca 2+ -binding proteins in<br />

DG (Sloviter, 1989) indicate that GR downregulation does not simply correspond to overall Ca 2+<br />

load. Rather, Ca 2+ -mediated regulation of GR expression might occur via distinct Ca 2+ signaling<br />

pathways linked to specific routes of Ca 2+ influx, such as L-type voltage-dependent Ca 2+<br />

channels (Bading et al., 1993). In support of this hypothesis, treatment with BAY K8644, which<br />

selectively promotes Ca 2+ influx through L-type Ca 2+ channels, decreases GR binding in AtT-20<br />

cells (Sheppard, 1994).<br />

Injury regulation of adrenocorticosteroid receptor mRNA levels – potential influence on<br />

hippocampal neuron viability<br />

Potential effects of GR downregulation on hippocampal circuitry<br />

Ample evidence supports the hypothesis that GR compromises survival of CA1 and CA3<br />

pyramidal neurons after injury, yet GR activation does not appear to compromise DG granule<br />

neuron survival after administration of KA or TBI. Nonetheless, alterations in dentate neuron<br />

excitability may influence downstream synaptic transmission and the survival of pyramidal<br />

101


neurons, as DG is the major recipient of perforant path projections from the entorhinal cortex and<br />

delivers the majority of excitatory input into CA3 (Amaral and Witter, 1989) (see Figure 1.5).<br />

Electrophysiological studies demonstrating GR attenuation of CA1 neuron excitability (Kerr et<br />

al., 1989) raise the possibility that decreased GR expression in DG may enhance KA and TBIinduced<br />

granule neuron hyperexcitability (Lowenstein et al., 1992; Sloviter, 1992; Santhakumar<br />

et al., 2000), potentially contributing to excitotoxic stress in CA3 and CA1. However, a recent<br />

report demonstrates that in contrast to CA1, GR activation does not suppress synaptically evoked<br />

field potentials in dentate gyrus (Stienstra and Joels, 2000), indicating that decreased GR<br />

expression may not influence excitability in this region. In addition, GR blockade, which should<br />

have a maximal effect if any on glucocorticoid-modulated synaptic transmission, does not<br />

compromise CA3 or CA1 survival after TBI (Chapter Four). It therefore does not appear likely<br />

that altered GR expression influences pyramidal neuron survival via modulation of DG synaptic<br />

transmission.<br />

Alternatively, downregulation of GR expression may facilitate rearrangement of DG circuitry<br />

after excitotoxic and traumatic brain injury. This hypothesis is supported by studies<br />

investigating reactive synaptogenesis in response to entorhinal cortex lesion in the rat.<br />

Glucocorticoids inhibit axon sprouting (Scheff and Cotman, 1982) and influence synaptic<br />

replacement (Scheff et al., 1986) in DG following delivery of entorhinal cortex lesion. In<br />

addition, an approximately 33% bilateral decrease in GR mRNA levels has been observed in DG<br />

one day following unilateral entorhinal cortex lesion (O'Donnell et al., 1993). <strong>The</strong> authors of<br />

this study proposed that decreased GR expression might serve to attenuate glucocorticoid<br />

inhibition of lesion-induced synaptogenesis (O'Donnell et al., 1993). Strikingly, the patterns of<br />

GR downregulation following entorhinal cortex lesion, kainic acid insult, and TBI are very<br />

similar (Chapters Three and Five), raising speculation that decreased expression of GR in dentate<br />

gyrus may have a permissive effect on restructuring of DG circuitry following multiple types of<br />

injury. Kainic acid is known to induce mossy fiber sprouting (Davenport et al., 1990), but the<br />

extent of DG remodeling induced by TBI remains to be investigated.<br />

Potential effects of GR downregulation on neurogenesis<br />

In addition to facilitating rearrangement of residual DG circuitry, downregulation of GR<br />

expression may influence DG neurogenesis after injury. Neurogenesis of DG granule cells<br />

102


continues well into adulthood in the rat and is inhibited by corticosterone in a concentrationdependent<br />

manner (Cameron and Gould, 1994). Kainic acid (Gray and Sundstrom, 1998) and<br />

TBI (Dash et al., 2001) have both been shown to enhance DG neurogenesis. GR downregulation<br />

in dentate gyrus after KA and traumatic brain injury may therefore facilitate injury-induced<br />

neurogenesis and the subsequent restoration of functional DG circuitry.<br />

103


Summary<br />

<strong>The</strong> present thesis supports the hypotheses that MR activation enhances and GR activation<br />

endangers neuronal survival in specific hippocampal subregions following excitotoxic or<br />

traumatic brain injury. Experiments in the current work demonstrate a neuroprotective effect of<br />

GR blockade after experimental TBI in vivo, as pretreatment with RU486 prevented CCIinduced<br />

neuronal loss in hippocampal subregion CA1 (Chapter Four). <strong>The</strong> present data also<br />

provide evidence that MR protects hippocampal neurons following excitotoxic injury, as MR<br />

blockade with SPIRO increased CA3 cell loss following systemic KA injection (Chapter Three).<br />

In addition, this work demonstrates KA and TBI-induced downregulation of GR mRNA levels in<br />

dentate gyrus, which may facilitate the recovery of functional DG circuitry after injury (Chapters<br />

Three and Five). Adrenocorticosteroid receptors clearly influence hippocampal neuron survival<br />

after injury, and future study of ACR regulation under challenging conditions could contribute to<br />

the improved treatment of seizure, stroke, traumatic injury, and neurodegenerative diseases.<br />

104


List of Abbreviations<br />

Appendix<br />

mineralocorticoid receptor (MR), glucocorticoid receptor (GR), adrenocorticosteroid receptor<br />

(ACR), mifepristone (RU486), hypothalamic-pituitary-adrenal (HPA), corticotropin-releasing<br />

hormone (CRH), adrenocorticotropin-releasing hormone (ACTH), paraventricular nucleus of the<br />

hypothalamus (PVN), heat shock protein (HSP), glucocorticoid response element (GRE),<br />

activator protein-1 (AP-1), cyclic AMP response element binding protein (CREB), nuclear<br />

factor-κB (NF-κB), negative glucocorticoid response element (nGRE), calcium (Ca 2+ ),<br />

corticosterone (CORT), plasma membrane calcium pump isoform 1 (PMCA1), brain-derived<br />

neurotrophic factor (BDNF), neurotrophin-3 (NT-3), excitatory postsynaptic potential (EPSP),<br />

entorhinal cortex (EC), subicular complex (S), nerve growth factor (NGF), apoptosis-inducing<br />

factor (AIF), apoptotic protease activating factor-1 (Apaf-1), inhibitor of CAD (ICAD), caspaseactivated<br />

deoxyribonuclease (CAD), traumatic brain injury (TBI), controlled cortical impact<br />

(CCI), anti-apoptotic B-cell leukemia/lymphoma-2 gene (bcl-2), messenger RNA (mRNA),<br />

interleukin (IL), tumor necrosis factor alpha (TNF-α), cyclooxygenase 2 (COX-2), metabotropic<br />

glutamate receptor subtype 1 (mGluR1), cyclic adenosine monophosphate (cAMP)<br />

105


References<br />

Adams JH, Doyle D, Graham DI, Lawrence AE, McLellan DR, Gennarelli TA, Pastuszko M,<br />

and Sakamoto T (1985) <strong>The</strong> contusion index: a reappraisal in human and experimental<br />

non-missile head injury. Neuropathol Appl Neurobiol 11:299-308.<br />

Ahima RS and Harlan RE (1990) Charting of type II glucocorticoid receptor-like<br />

immunoreactivity in the rat central nervous system. Neuroscience 39:579-604.<br />

Amaral DG and Witter MP (1989) <strong>The</strong> three-dimensional organization of the hippocampal<br />

formation: a review of anatomical data. Neuroscience 31:571-591.<br />

Antakly T, Mercille S and Cote JP (1987) Tissue-specific dopaminergic regulation of the<br />

glucocorticoid receptor in the rat pituitary. Endocrinology 120:1558-1562.<br />

Antonawich FJ, Miller G, Rigsby DC, and Davis JN (1999) Regulation of ischemic cell death by<br />

glucocorticoids and adrenocorticotropic hormone. Neuroscience 88:319-325.<br />

Antoni FA (1986) Hypothalamic control of adrenocorticotropin secretion: advances since the<br />

discovery of 41-residue corticotropin-releasing factor. Endocr Rev 7:351-378.<br />

Aramori I and Nakanishi S (1992) Signal transduction and pharmacological characteristics of a<br />

metabotropic glutamate receptor, mGluR1, in transfected CHO cells. Neuron 8:757-765.<br />

Armanini MP, Hutchins C, Stein BA, and Sapolsky RM (1990) Glucocorticoid endangerment of<br />

hippocampal neurons is NMDA-receptor dependent. Brain Res 532:7-12.<br />

Aronsson M, Fuxe K, Dong Y, Agnati LF, Okret S, and Gustafsson JA (1988) Localization of<br />

glucocorticoid receptor mRNA in the male rat brain by in situ hybridization. Proc Natl<br />

Acad Sci U S A 85:9331-9335.<br />

Arriza JL, Simerly RB, Swanson LW, and Evans RM (1988) <strong>The</strong> neuronal mineralocorticoid<br />

receptor as a mediator of glucocorticoid response. Neuron 1:887-900.<br />

Arriza JL, Weinberger C, Cerelli G, Glaser TM, Handelin BL, Housman DE, and Evans RM<br />

(1987) Cloning of human mineralocorticoid receptor complementary DNA: structural and<br />

functional kinship with the glucocorticoid receptor. Science 237:268-275.<br />

Auphan N, DiDonato JA, Rosette C, Helmberg A, and Karin M (1995) Immunosuppression by<br />

glucocorticoids: inhibition of NF-kappa B activity through induction of I kappa B<br />

synthesis. Science 270:286-290.<br />

106


Axelson DA, Doraiswamy PM, McDonald WM, Boyko OB, Tupler LA, Patterson LJ, Nemeroff<br />

CB, Ellinwood EH, Jr., and Krishnan KR (1993) Hypercortisolemia and hippocampal<br />

changes in depression. Psychiatry Res 47:163-173.<br />

Bading H, Ginty DD, and Greenberg ME (1993) Regulation of gene expression in hippocampal<br />

neurons by distinct calcium signaling pathways. Science 260:181-186.<br />

Baik EJ, Kim EJ, Lee SH, and Moon C (1999) Cyclooxygenase-2 selective inhibitors aggravate<br />

kainic acid induced seizure and neuronal cell death in the hippocampus. Brain Res<br />

843:118-129.<br />

Baldwin SA, Gibson T, Callihan CT, Sullivan PG, Palmer E, and Scheff SW (1997) Neuronal<br />

cell loss in the CA3 subfield of the hippocampus following cortical contusion utilizing<br />

the optical disector method for cell counting. J Neurotrauma 14:385-398.<br />

Ballarin M, Ernfors P, Lindefors N, and Persson H (1991) Hippocampal damage and kainic acid<br />

injection induce a rapid increase in mRNA for BDNF and NGF in the rat brain. Exp<br />

Neurol 114:35-43.<br />

Barbany G and Persson H (1993) Adrenalectomy attenuates kainic acid-elicited increases of<br />

messenger RNAs for neurotrophins and their receptors in the rat brain. Neuroscience<br />

54:909-922.<br />

Barde YA (1989) Trophic factors and neuronal survival. Neuron 2:1525-1534.<br />

Barger SW, Horster D, Furukawa K, Goodman Y, Krieglstein J, and Mattson MP (1995) Tumor<br />

necrosis factors alpha and beta protect neurons against amyloid beta-peptide toxicity:<br />

evidence for involvement of a kappa B-binding factor and attenuation of peroxide and<br />

Ca2+ accumulation. Proc Natl Acad Sci U S A 92:9328-9332.<br />

Bar-Peled O, Knudson M, Korsmeyer SJ, and Rothstein JD (1999) Motor neuron degeneration is<br />

attenuated in bax-deficient neurons in vitro. J Neurosci Res 55:542-556.<br />

Baulieu EE (1989) Contragestion and other clinical applications of RU 486, an antiprogesterone<br />

at the receptor. Science 245:1351-1357.<br />

Beer R, Franz G, Schopf M, Reindl M, Zelger B, Schmutzhard E, Poewe W, and Kampfl A<br />

(2000) Expression of Fas and Fas ligand after experimental traumatic brain injury in the<br />

rat. J Cereb Blood Flow Metab 20:669-677.<br />

Behl C, Trapp T, Skutella T, and Holsboer F (1997a) Protection against oxidative stress-induced<br />

neuronal cell death--a novel role for RU486. Eur J Neurosci 9:912-920.<br />

107


Behl C, Lezoualc'h F, Trapp T, Widmann M, Skutella T, and Holsboer F (1997b)<br />

Glucocorticoids enhance oxidative stress-induced cell death in hippocampal neurons in<br />

vitro. Endocrinology 138:101-106.<br />

Ben-Ari Y (1985) Limbic seizure and brain damage produced by kainic acid: mechanisms and<br />

relevance to human temporal lobe epilepsy. Neuroscience 14:375-403.<br />

Ben-Ari Y and Gho M (1988) Long-lasting modification of the synaptic properties of rat CA3<br />

hippocampal neurones induced by kainic acid. J Physiol 404:365-384.<br />

Ben-Ari Y and Cossart R (2000) Kainate, a double agent that generates seizures: two decades of<br />

progress. Trends Neurosci 23:580-587.<br />

Bertorelli G, Bocchino V, and Olivieri D (1998) Heat shock protein interactions with the<br />

glucocorticoid receptor. Pulm Pharmacol <strong>The</strong>r 11:7-12.<br />

Bettler B and Mulle C (1995) Review: neurotransmitter receptors. II. AMPA and kainate<br />

receptors. Neuropharmacology 34:123-139.<br />

Bhargava A, Meijer OC, Dallman MF, and Pearce D (2000) Plasma membrane calcium pump<br />

isoform 1 gene expression is repressed by corticosterone and stress in rat hippocampus. J<br />

Neurosci 20:3129-3138.<br />

Blaha GR, Raghupathi R, Saatman KE, and McIntosh TK (2000) Brain-derived neurotrophic<br />

factor administration after traumatic brain injury in the rat does not protect against<br />

behavioral or histological deficits. Neuroscience 99:483-493.<br />

Blaustein MP (1988) Calcium transport and buffering in neurons. Trends Neurosci 11:438-443.<br />

Blumcke I, Becker AJ, Klein C, Scheiwe C, Lie AA, Beck H, Waha A, Friedl MG, Kuhn R,<br />

Emson P, Elger C, and Wiestler OD (2000) Temporal lobe epilepsy associated upregulation<br />

of metabotropic glutamate receptors: correlated changes in mGluR1 mRNA<br />

and protein expression in experimental animals and human patients. J Neuropathol Exp<br />

Neurol 59:1-10.<br />

Bracken MB, Shepard MJ, Collins WF, Holford TR, Young W, Baskin DS, Eisenberg HM,<br />

Flamm E, Leo-Summers L, and Maroon J (1990) A randomized, controlled trial of<br />

methylprednisolone or naloxone in the treatment of acute spinal-cord injury. Results of<br />

the Second National Acute Spinal Cord Injury Study. N Engl J Med 322:1405-1411.<br />

108


Brady LS, <strong>Lynn</strong> AB, Whitfield HJ, Jr., Kim H, and Herkenham M (1992) Intrahippocampal<br />

colchicine alters hypothalamic corticotropin-releasing hormone and hippocampal steroid<br />

receptor mRNA in rat brain. Neuroendocrinology 55:121-133.<br />

Brain TF (1996) <strong>The</strong> role of glucocorticoids in the treatment of severe head injury. Brain Trauma<br />

Foundation. J Neurotrauma 13:715-718.<br />

Braughler JM and Hall ED (1992) Involvement of lipid peroxidation in CNS injury. J<br />

Neurotrauma 9:S1-7.<br />

Bruce AJ, Boling W, Kindy MS, Peschon J, Kraemer PJ, Carpenter MK, Holtsberg FW, and<br />

Mattson MP (1996) Altered neuronal and microglial responses to excitotoxic and<br />

ischemic brain injury in mice lacking TNF receptors. Nat Med 2:788-794.<br />

Bruner KL, Derfoul A, Robertson NM, Guerriero G, Fernandes-Alnemri T, Alnemri ES, and<br />

Litwack G (1997) <strong>The</strong> unliganded mineralocorticoid receptor is associated with heat<br />

shock proteins 70 and 90 and the immunophilin FKBP-52. Recept Signal Transduct 7:85-<br />

98.<br />

Bullock R and Fujisawa H (1992) <strong>The</strong> role of glutamate antagonists for the treatment of CNS<br />

injury. J Neurotrauma 9:S443-462.<br />

Cabral AL, Hays AN, Housley PR, Brentani MM and Martins VR (2001) Repression of<br />

glucocorticoid receptor gene transcription by c-Jun. Mol Cell Endocrinol 175:67-79.<br />

Cadepond F, Ulmann A, and Baulieu EE (1997) RU486 (mifepristone): mechanisms of action<br />

and clinical uses. Annu Rev Med 48:129-156.<br />

Cameron HA and Gould E (1994) Adult neurogenesis is regulated by adrenal steroids in the<br />

dentate gyrus. Neuroscience 61:203-209.<br />

Carpenter SE, Santanam N, Murphy AA, Rock JA, and Parthasarathy S (1996) Inhibition of<br />

oxidative modification of proteins by RU486. Fertil Steril 66:90-94.<br />

Cato AC and Wade E (1996) Molecular mechanisms of anti-inflammatory action of<br />

glucocorticoids. Bioessays 18:371-378.<br />

Chabert PR, Guelpa-Decorzant C, Riondel AM, and Vallotton MB (1984) Effect of<br />

spironolactone on electrolytes, renin, ACTH and corticosteroids in the rat. J Steroid<br />

Biochem 20:1253-1259.<br />

109


Chao HM and McEwen BS (1994) Glucocorticoids and the expression of mRNAs for<br />

neurotrophins, their receptors and GAP-43 in the rat hippocampus. Brain Res Mol Brain<br />

Res 26:271-276.<br />

Chao HM, Choo PH, and McEwen BS (1989) Glucocorticoid and mineralocorticoid receptor<br />

mRNA expression in rat brain. Neuroendocrinology 50:365-371.<br />

Chao HM, Ma LY, McEwen BS, and Sakai RR (1998a) Regulation of glucocorticoid receptor<br />

and mineralocorticoid receptor messenger ribonucleic acids by selective agonists in the<br />

rat hippocampus. Endocrinology 139:1810-1814.<br />

Chao HM, Sakai RR, Ma LY, and McEwen BS (1998b) Adrenal steroid regulation of<br />

neurotrophic factor expression in the rat hippocampus. Endocrinology 139:3112-3118.<br />

Cheng B and Mattson MP (1994) NT-3 and BDNF protect CNS neurons against<br />

metabolic/excitotoxic insults. Brain Res 640:56-67.<br />

Cheng B, Christakos S, and Mattson MP (1994) Tumor necrosis factors protect neurons against<br />

metabolic-excitotoxic insults and promote maintenance of calcium homeostasis. Neuron<br />

12:139-153.<br />

Chiolero R and Berger M (1994) Endocrine response to brain injury. New Horiz 2:432-442.<br />

Choi DW (1992) Excitotoxic cell death. J Neurobiol 23:1261-1276.<br />

Clark RS, Chen J, Watkins SC, Kochanek PM, Chen M, Stetler RA, Loeffert JE, and Graham SH<br />

(1997) Apoptosis-suppressor gene bcl-2 expression after traumatic brain injury in rats. J<br />

Neurosci 17:9172-9182.<br />

Clark RS, Kochanek PM, Chen M, Watkins SC, Marion DW, Chen J, Hamilton RL, Loeffert JE,<br />

and Graham SH (1999) Increases in Bcl-2 and cleavage of caspase-1 and caspase-3 in<br />

human brain after head injury. Faseb J 13:813-821.<br />

Cooper PR, Moody S, Clark WK, Kirkpatrick J, Maravilla K, Gould AL, and Drane W (1979)<br />

Dexamethasone and severe head injury. A prospective double-blind study. J Neurosurg<br />

51:307-316.<br />

Cortez SC, McIntosh TK, and Noble LJ (1989) Experimental fluid percussion brain injury:<br />

vascular disruption and neuronal and glial alterations. Brain Res 482:271-282.<br />

Cosi C, Spoerri PE, Comelli MC, Guidolin D, and Skaper SD (1993) Glucocorticoids depress<br />

activity-dependent expression of BDNF mRNA in hippocampal neurones. Neuroreport<br />

4:527-530.<br />

110


Couette B, Marsaud V, Baulieu EE, Richard-Foy H, and Rafestin-Oblin ME (1992)<br />

Spironolactone, an aldosterone antagonist, acts as an antiglucocorticosteroid on the<br />

mouse mammary tumor virus promoter. Endocrinology 130:430-436.<br />

Coyle JT, Bird SJ, Evans RH, Gulley RL, Nadler JV, Nicklas WJ, and Olney JW (1981)<br />

Excitatory amino acid neurotoxins: selectivity, specificity, and mechanisms of action.<br />

Based on an NRP one-day conference held June 30, 1980. Neurosci Res Program Bull<br />

19:1-427.<br />

D'Agostino J, Vaeth GF, and Henning SJ (1982) Diurnal rhythm of total and free concentrations<br />

of serum corticosterone in the rat. Acta Endocrinol (Copenh) 100:85-90.<br />

Dallman MF, Akana, S.F., Scribner, K.A., Bradbury, M.J., Walker, C.D., Strack, A.M., and<br />

Cascio, C.S. (1992) Mortyn Jones Memorial Lecture: Stress, feedback and facilitation in<br />

the hypothalamo-pituitary-adrenal axis. J Neuroendocrinol 4:517-526.<br />

Dash PK, Mach SA, and Moore AN (2000) Regional expression and role of cyclooxygenase-2<br />

following experimental traumatic brain injury. J Neurotrauma 17:69-81.<br />

Dash PK, Mach SA, and Moore AN (2001) Enhanced neurogenesis in the rodent hippocampus<br />

following traumatic brain injury. J Neurosci Res 63:313-319.<br />

Davenport CJ, Brown WJ, and Babb TL (1990) Sprouting of GABAergic and mossy fiber axons<br />

in dentate gyrus following intrahippocampal kainate in the rat. Exp Neurol 109:180-190.<br />

de Bock F, Dornand J, and Rondouin G (1996) Release of TNF alpha in the rat hippocampus<br />

following epileptic seizures and excitotoxic neuronal damage. Neuroreport 7:1125-1129.<br />

De Bosscher K, Vanden Berghe W, and Haegeman G (2000) Mechanisms of anti-inflammatory<br />

action and of immunosuppression by glucocorticoids: negative interference of activated<br />

glucocorticoid receptor with transcription factors. J Neuroimmunol 109:16-22.<br />

De Kloet ER and Reul JM (1987) Feedback action and tonic influence of corticosteroids on brain<br />

function: a concept arising from the heterogeneity of brain receptor systems.<br />

Psychoneuroendocrinology 12:83-105.<br />

De Kloet ER, Vreugdenhil E, Oitzl MS, and Joels M (1998) Brain corticosteroid receptor<br />

balance in health and disease. Endocr Rev 19:269-301.<br />

Diamond MI, Miner JN, Yoshinaga SK, and Yamamoto KR (1990) Transcription factor<br />

interactions: selectors of positive or negative regulation from a single DNA element.<br />

Science 249:1266-1272.<br />

111


Dietrich WD, Alonso O, and Halley M (1994) Early microvascular and neuronal consequences<br />

of traumatic brain injury: a light and electron microscopic study in rats. J Neurotrauma<br />

11:289-301.<br />

Dingledine R and McBain CJ (1999) Glutamate and Aspartate, 6th Edition. Philadelphia, PA:<br />

Lippincott Williams and Wilkins.<br />

Dixon CE, Clifton GL, Lighthall JW, Yaghmai AA, and Hayes RL (1991) A controlled cortical<br />

impact model of traumatic brain injury in the rat. J Neurosci Methods 39:253-262.<br />

Dixon CE, Lyeth BG, Povlishock JT, Findling RL, Hamm RJ, Marmarou A, Young HF, and<br />

Hayes RL (1987) A fluid percussion model of experimental brain injury in the rat. J<br />

Neurosurg 67:110-119.<br />

Dong G, Singh DK, Dendle P, and Prasad RM (2001) Regional expression of Bcl-2 mRNA and<br />

mitochondrial cytochrome c release after experimental brain injury in the rat. Brain Res<br />

903:45-52.<br />

Drouin J, Sun YL, Chamberland M, Gauthier Y, De Lean A, Nemer M, and Schmidt TJ (1993)<br />

Novel glucocorticoid receptor complex with DNA element of the hormone-repressed<br />

POMC gene. Embo J 12:145-156.<br />

Dugich-Djordjevic MM, Tocco G, Lapchak PA, Pasinetti GM, Najm I, Baudry M, and Hefti F<br />

(1992) Regionally specific and rapid increases in brain-derived neurotrophic factor<br />

messenger RNA in the adult rat brain following seizures induced by systemic<br />

administration of kainic acid. Neuroscience 47:303-315.<br />

Eldadah BA and Faden AI (2000) Caspase pathways, neuronal apoptosis, and CNS injury. J<br />

Neurotrauma 17:811-829.<br />

Elliott EM and Sapolsky RM (1992) Corticosterone enhances kainic acid-induced calcium<br />

elevation in cultured hippocampal neurons. J Neurochem 59:1033-1040.<br />

Elliott EM and Sapolsky RM (1993) Corticosterone impairs hippocampal neuronal calcium<br />

regulation. Brain Res 602:84-90.<br />

Elliott EM, Mattson MP, Vanderklish P, Lynch G, Chang I, and Sapolsky RM (1993)<br />

Corticosterone exacerbates kainate-induced alterations in hippocampal tau<br />

immunoreactivity and spectrin proteolysis in vivo. J Neurochem 61:57-67.<br />

112


Enari M, Sakahira H, Yokoyama H, Okawa K, Iwamatsu A, and Nagata S (1998) A caspaseactivated<br />

DNase that degrades DNA during apoptosis, and its inhibitor ICAD. Nature<br />

391:43-50.<br />

Eriksson C, Tehranian R, Iverfeldt K, Winblad B, and Schultzberg M (2000) Increased<br />

expression of mRNA encoding interleukin-1beta and caspase-1, and the secreted isoform<br />

of interleukin-1 receptor antagonist in the rat brain following systemic kainic acid<br />

administration. J Neurosci Res 60:266-279.<br />

Evan G and Littlewood T (1998) A matter of life and cell death. Science 281:1317-1322.<br />

Evans RM (1988) <strong>The</strong> steroid and thyroid hormone receptor superfamily. Science 240:889-895.<br />

Faden AI (1993) Experimental neurobiology of central nervous system trauma. Crit Rev<br />

Neurobiol 7:175-186.<br />

Faden AI, Demediuk P, Panter SS, and Vink R (1989) <strong>The</strong> role of excitatory amino acids and<br />

NMDA receptors in traumatic brain injury. Science 244:798-800.<br />

Fan L, Young PR, Barone FC, Feuerstein GZ, Smith DH, and McIntosh TK (1995) Experimental<br />

brain injury induces expression of interleukin-1 beta mRNA in the rat brain. Brain Res<br />

Mol Brain Res 30:125-130.<br />

Fan L, Young PR, Barone FC, Feuerstein GZ, Smith DH, and McIntosh TK (1996) Experimental<br />

brain injury induces differential expression of tumor necrosis factor-alpha mRNA in the<br />

CNS. Brain Res Mol Brain Res 36:287-291.<br />

Feuerstein GZ, Wang X, and Barone FC (1997) Inflammatory gene expression in cerebral<br />

ischemia and trauma. Potential new therapeutic targets. Ann N Y Acad Sci 825:179-193.<br />

Fineman I, Hovda DA, Smith M, Yoshino A, and Becker DP (1993) Concussive brain injury is<br />

associated with a prolonged accumulation of calcium: a 45Ca autoradiographic study.<br />

Brain Res 624:94-102.<br />

Finkbeiner S (2000) Calcium regulation of the brain-derived neurotrophic factor gene. Cell Mol<br />

Life Sci 57:394-401.<br />

Foster AC, Mena EE, Monaghan DT, and Cotman CW (1981) Synaptic localization of kainic<br />

acid binding sites. Nature 289:73-75.<br />

Frerking M and Nicoll RA (2000) Synaptic kainate receptors. Curr Opin Neurobiol 10:342-351.<br />

113


Friedman LK (1997) Developmental switch in phenotypic expression of preproenkephalin<br />

mRNA and 45Ca2+ accumulation following kainate-induced status epilepticus. Brain<br />

Res Dev Brain Res 101:287-293.<br />

Fujikawa DG, Shinmei SS, and Cai B (2000) Kainic acid-induced seizures produce necrotic, not<br />

apoptotic, neurons with internucleosomal DNA cleavage: implications for programmed<br />

cell death mechanisms. Neuroscience 98:41-53.<br />

Furukawa K and Mattson MP (1998) <strong>The</strong> transcription factor NF-kappaB mediates increases in<br />

calcium currents and decreases in NMDA- and AMPA/kainate-induced currents induced<br />

by tumor necrosis factor-alpha in hippocampal neurons. J Neurochem 70:1876-1886.<br />

Fuxe K, Wikstrom AC, Okret S, Agnati LF, Harfstrand A, Yu ZY, Granholm L, Zoli M, Vale W,<br />

and Gustafsson JA (1985) Mapping of glucocorticoid receptor immunoreactive neurons<br />

in the rat tel- and diencephalon using a monoclonal antibody against rat liver<br />

glucocorticoid receptor. Endocrinology 117:1803-1812.<br />

Gall C, Murray K, and Isackson PJ (1991) Kainic acid-induced seizures stimulate increased<br />

expression of nerve growth factor mRNA in rat hippocampus. Brain Res Mol Brain Res<br />

9:113-123.<br />

Garcia ML and Strehler EE (1999) Plasma membrane calcium ATPases as critical regulators of<br />

calcium homeostasis during neuronal cell function. Front Biosci 4:D869-882.<br />

Garcia ML, Garcia VB, Isackson PJ, and Windebank AJ (1997a) Long-term alterations in growth<br />

factor mRNA expression following seizures. Neuroreport 8:1445-1449.<br />

Garcia ML, Murray KD, Garcia VB, Strehler EE, and Isackson PJ (1997b) Seizure-induced<br />

alterations of plasma membrane calcium ATPase isoforms 1, 2 and 3 mRNA and protein<br />

in rat hippocampus. Brain Res Mol Brain Res 45:230-238.<br />

Garcia ML, Usachev YM, Thayer SA, Strehler EE, and Windebank AJ (2001) Plasma membrane<br />

calcium ATPase plays a role in reducing Ca(2+)-mediated cytotoxicity in PC12 cells. J<br />

Neurosci Res 64:661-669.<br />

Giannotta SL, Weiss MH, Apuzzo ML, and Martin E (1984) High dose glucocorticoids in the<br />

management of severe head injury. Neurosurgery 15:497-501.<br />

Gillardon F, Wickert H, and Zimmermann M (1995) Up-regulation of bax and down-regulation<br />

of bcl-2 is associated with kainate-induced apoptosis in mouse brain. Neurosci Lett<br />

192:85-88.<br />

114


Gomez F, Lahmame A, de Kloet ER, and Armario A (1996) Hypothalamic-pituitary-adrenal<br />

response to chronic stress in five inbred rat strains: differential responses are mainly<br />

located at the adrenocortical level. Neuroendocrinology 63:327-337.<br />

Goodman JC, Cherian L, Bryan RM, Jr., and Robertson CS (1994) Lateral cortical impact injury<br />

in rats: pathologic effects of varying cortical compression and impact velocity. J<br />

Neurotrauma 11:587-597.<br />

Goodman Y, Bruce AJ, Cheng B, and Mattson MP (1996) Estrogens attenuate and corticosterone<br />

exacerbates excitotoxicity, oxidative injury, and amyloid beta-peptide toxicity in<br />

hippocampal neurons. J Neurochem 66:1836-1844.<br />

Gould E (1994) <strong>The</strong> effects of adrenal steroids and excitatory input on neuronal birth and<br />

survival. Ann N Y Acad Sci 743:73-92.<br />

Gould E, Woolley CS, and McEwen BS (1990) Short-term glucocorticoid manipulations affect<br />

neuronal morphology and survival in the adult dentate gyrus. Neuroscience 37:367-375.<br />

Gould E, Woolley CS, Cameron HA, Daniels DC, and McEwen BS (1991) Adrenal steroids<br />

regulate postnatal development of the rat dentate gyrus: II. Effects of glucocorticoids and<br />

mineralocorticoids on cell birth. J Comp Neurol 313:486-493.<br />

Gray WP and Sundstrom LE (1998) Kainic acid increases the proliferation of granule cell<br />

progenitors in the dentate gyrus of the adult rat. Brain Res 790:52-59.<br />

Green DR and Reed JC (1998) Mitochondria and apoptosis. Science 281:1309-1312.<br />

Greenamyre JT, Olson JM, Penney JB, Jr., and Young AB (1985) Autoradiographic<br />

characterization of N-methyl-D-aspartate-, quisqualate- and kainate-sensitive glutamate<br />

binding sites. J Pharmacol Exp <strong>The</strong>r 233:254-263.<br />

Grundy PL, Patel N, Harbuz MS, Lightman SL, and Sharples PM (2000) Glucocorticoids<br />

modulate BDNF mRNA expression in the rat hippocampus after traumatic brain injury.<br />

Neuroreport 11:3381-3384.<br />

Grundy PL, Patel N, Harbuz MS, Lightman SL, and Sharples PM (2001) Glucocorticoids<br />

modulate the NGF mRNA response in the rat hippocampus after traumatic brain injury.<br />

Brain Res 892:386-390.<br />

Hall ED, Yonkers PA, Andrus PK, Cox JW, and Anderson DK (1992) Biochemistry and<br />

pharmacology of lipid antioxidants in acute brain and spinal cord injury. J Neurotrauma<br />

9:S425-442.<br />

115


Hashimoto K, Watanabe K, Nishimura T, Iyo M, Shirayama Y, and Minabe Y (1998) Behavioral<br />

changes and expression of heat shock protein hsp-70 mRNA, brain-derived neurotrophic<br />

factor mRNA, and cyclooxygenase-2 mRNA in rat brain following seizures induced by<br />

systemic administration of kainic acid. Brain Res 804:212-223.<br />

Hayes RL, Jenkins LW, and Lyeth BG (1992) Neurotransmitter-mediated mechanisms of<br />

traumatic brain injury: acetylcholine and excitatory amino acids. J Neurotrauma 9:S173-<br />

187.<br />

Hayes RL, Yang K, Raghupathi R, and McIntosh TK (1995) Changes in gene expression<br />

following traumatic brain injury in the rat. J Neurotrauma 12:779-790.<br />

He H, Lam M, McCormick TS, and Distelhorst CW (1997) Maintenance of calcium homeostasis<br />

in the endoplasmic reticulum by Bcl-2. J Cell Biol 138:1219-1228.<br />

Henshall DC, Skradski SL, Lan JQ, Ren T, and Simon RP (2001) Increased Bcl-w expression<br />

following focally evoked limbic seizures in the rat. Neurosci Lett 305:153-156.<br />

Herman JP (1993) Regulation of adrenocorticosteroid receptor mRNA expression in the central<br />

nervous system. Cell Mol Neurobiol 13:349-372.<br />

Herman JP and Watson SJ (1995) Stress regulation of mineralocorticoid receptor heteronuclear<br />

RNA in rat hippocampus. Brain Res 677:243-249.<br />

Herman JP and Spencer R (1998) Regulation of hippocampal glucocorticoid receptor gene<br />

transcription and protein expression in vivo. J Neurosci 18:7462-7473.<br />

Herman JP, Patel PD, Akil H, and Watson SJ (1989a) Localization and regulation of<br />

glucocorticoid and mineralocorticoid. Mol Endocrinol 3:1886-1894.<br />

Herman JP, Watson SJ, Chao HM, Coirini HM, and McEwen BS (1993) Diurnal regulation of<br />

glucocorticoid receptor and mineralocorticoid receptor mRNAs in the rat hippocampus.<br />

Mol Cell Neurosci 4:181-190.<br />

Herman JP, Schafer MK, Young EA, Thompson R, Douglass J, Akil H, and Watson SJ (1989b)<br />

Evidence for hippocampal regulation of neuroendocrine neurons of the hypothalamopituitary-adrenocortical<br />

axis. J Neurosci 9:3072-3082.<br />

Herrlich P (2001) Cross-talk between glucocorticoid receptor and AP-1. Oncogene 20:2465-<br />

2475.<br />

116


Hicks R, Soares H, Smith D, and McIntosh T (1996) Temporal and spatial characterization of<br />

neuronal injury following lateral fluid-percussion brain injury in the rat. Acta<br />

Neuropathol (Berl) 91:236-246.<br />

Hicks RR, Smith DH, Lowenstein DH, Saint Marie R, and McIntosh TK (1993) Mild<br />

experimental brain injury in the rat induces cognitive deficits associated with regional<br />

neuronal loss in the hippocampus. J Neurotrauma 10:405-414.<br />

Hicks RR, Numan S, Dhillon HS, Prasad MR, and Seroogy KB (1997) Alterations in BDNF and<br />

NT-3 mRNAs in rat hippocampus after experimental brain trauma. Brain Res Mol Brain<br />

Res 48:401-406.<br />

Hicks RR, Li C, Zhang L, Dhillon HS, Prasad MR, and Seroogy KB (1999) Alterations in BDNF<br />

and trkB mRNA levels in the cerebral cortex following experimental brain trauma in rats.<br />

J Neurotrauma 16:501-510.<br />

Hollenberg SM, Weinberger C, Ong ES, Cerelli G, Oro A, Lebo R, Thompson EB, Rosenfeld<br />

MG, and Evans RM (1985) Primary structure and expression of a functional human<br />

glucocorticoid receptor cDNA. Nature 318:635-641.<br />

Hollmann M and Heinemann S (1994) Cloned glutamate receptors. Annu Rev Neurosci 17:31-<br />

108.<br />

Hollmann M, Hartley M, and Heinemann S (1991) Ca2+ permeability of KA-AMPA--gated<br />

glutamate receptor channels depends on subunit composition. Science 252:851-853.<br />

Honkaniemi J and Sharp FR (1999) Prolonged expression of zinc finger immediate-early gene<br />

mRNAs and decreased protein synthesis following kainic acid induced seizures. Eur J<br />

Neurosci 11:10-17.<br />

Hori N, Ffrench-Mullen JM, and Carpenter DO (1985) Kainic acid responses and toxicity show<br />

pronounced Ca2+ dependence. Brain Res 358:380-384.<br />

Horner HC, Packan DR, and Sapolsky RM (1990) Glucocorticoids inhibit glucose transport in<br />

cultured hippocampal neurons and glia. Neuroendocrinology 52:57-64.<br />

Hotchkiss RS and Karl IE (1996) Calcium: a regulator of the inflammatory response in<br />

endotoxemia and sepsis. New Horiz 4:58-71.<br />

Hovda DA, Becker DP, and Katayama Y (1992) Secondary injury and acidosis. J Neurotrauma<br />

9:S47-60.<br />

117


Jacobson L and Sapolsky R (1991) <strong>The</strong> role of the hippocampus in feedback regulation of the<br />

hypothalamic-pituitary-adrenocortical axis. Endocr Rev 12:118-134.<br />

Jenster G, Spencer TE, Burcin MM, Tsai SY, Tsai MJ, and O'Malley BW (1997) Steroid<br />

receptor induction of gene transcription: a two-step model. Proc Natl Acad Sci USA<br />

94:7879-7884.<br />

Jiang H and Guroff G (1997) Actions of the neurotrophins on calcium uptake. J Neurosci Res<br />

50:355-360.<br />

Joels M (2001) Corticosteroid actions in the hippocampus. J Neuroendocrinol 13:657-669.<br />

Joels M and de Kloet ER (1989) Effects of glucocorticoids and norepinephrine on the excitability<br />

in the hippocampus. Science 245:1502-1505.<br />

Joels M and de Kloet ER (1990) Mineralocorticoid receptor-mediated changes in membrane<br />

properties of rat CA1 pyramidal neurons in vitro. Proc Natl Acad Sci USA 87:4495-4498.<br />

Joels M and de Kloet ER (1992) Control of neuronal excitability by corticosteroid hormones.<br />

Trends Neurosci 15:25-30.<br />

Joels M and Vreugdenhil E (1998) Corticosteroids in the brain. Cellular and molecular actions.<br />

Mol Neurobiol 17:87-108.<br />

Johren O, Flugge G, and Fuchs E (1994) Hippocampal glucocorticoid receptor expression in the<br />

tree shrew: regulation by psychosocial conflict. Cell Mol Neurobiol 14:281-296.<br />

Jones MT and Gillham B (1988) Factors involved in the regulation of adrenocorticotropic<br />

hormone/beta-lipotropic hormone. Physiol Rev 68:743-818.<br />

Jordan J, Galindo MF, Prehn JH, Weichselbaum RR, Beckett M, Ghadge GD, Roos RP, Leiden<br />

JM, and Miller RJ (1997) p53 expression induces apoptosis in hippocampal pyramidal<br />

neuron cultures. J Neurosci 17:1397-1405.<br />

Joyce DA, Steer JH, and Abraham LJ (1997) Glucocorticoid modulation of human<br />

monocyte/macrophage function: control of TNF-alpha secretion. Inflamm Res 46:447-<br />

451.<br />

Jurgensmeier JM, Xie Z, Deveraux Q, Ellerby L, Bredesen D, and Reed JC (1998) Bax directly<br />

induces release of cytochrome c from isolated mitochondria. Proc Natl Acad Sci USA<br />

95:4997-5002.<br />

Kadekaro M, Ito M, and Gross PM (1988) Local cerebral glucose utilization is increased in<br />

acutely adrenalectomized rats. Neuroendocrinology 47:329-334.<br />

118


Kaminska B, Filipkowski RK, Zurkowska G, Lason W, Przewlocki R, and Kaczmarek L (1994)<br />

Dynamic changes in the composition of the AP-1 transcription factor DNA-binding<br />

activity in rat brain following kainate-induced seizures and cell death. Eur J Neurosci<br />

6:1558-1566.<br />

Kampfl A, Posmantur RM, Zhao X, Schmutzhard E, Clifton GL, and Hayes RL (1997)<br />

Mechanisms of calpain proteolysis following traumatic brain injury: implications for<br />

pathology and therapy: implications for pathology and therapy: a review and update. J<br />

Neurotrauma 14:121-134.<br />

Kanai Y, Smith CP, and Hediger MA (1993) A new family of neurotransmitter transporters: the<br />

high-affinity glutamate transporters. Faseb J 7:1450-1459.<br />

Karin M (1998) New twists in gene regulation by glucocorticoid receptor: is DNA binding<br />

dispensable Cell 93:487-490.<br />

Karst H, Wadman WJ, and Joels M (1994) Corticosteroid receptor-dependent modulation of<br />

calcium currents in rat hippocampal CA1 neurons. Brain Res 649:234-242.<br />

Kato J, Hirata S, Nozawa A, and Yamada-Mouri N (1994) Gene expression of progesterone<br />

receptor isoforms in the rat brain. Horm Behav 28:454-463.<br />

Katoh-Semba R, Takeuchi IK, Inaguma Y, Ito H, and Kato K (1999) Brain-derived neurotrophic<br />

factor, nerve growth and neurotrophin-3 selected regions of the rat brain following kainic<br />

acid-induced seizure activity. Neurosci Res 35:19-29.<br />

Kaya SS, Mahmood A, Li Y, Yavuz E, Goksel M, and Chopp M (1999) Apoptosis and<br />

expression of p53 response proteins and cyclin D1 after cortical impact in rat brain. Brain<br />

Res 818:23-33.<br />

Keinanen K, Wisden W, Sommer B, Werner P, Herb A, Verdoorn TA, Sakmann B, and Seeburg<br />

PH (1990) A family of AMPA-selective glutamate receptors. Science 249:556-560.<br />

Keller-Wood ME and Dallman MF (1984) Corticosteroid inhibition of ACTH secretion. Endocr<br />

Rev 5:1-24.<br />

Kerr DS, Campbell LW, Hao SY, and Landfield PW (1989) Corticosteroid modulation of<br />

hippocampal potentials: increased effect with aging. Science 245:1505-1509.<br />

Kerr DS, Campbell LW, Thibault O, and Landfield PW (1992) Hippocampal glucocorticoid<br />

receptor activation enhances voltage-dependent Ca2+ conductances: relevance to brain<br />

aging. Proc Natl Acad Sci USA 89:8527-8531.<br />

119


King LR, McLaurin RL, Lewis HP, and Knowles HC (1970) Plasma cortisol levels after head<br />

injury. Ann Surg 172:975-984.<br />

Kirino T (1982) Delayed neuronal death in the gerbil hippocampus following ischemia. Brain<br />

Res 239:57-69.<br />

Kluck RM, Bossy-Wetzel E, Green DR, and Newmeyer DD (1997) <strong>The</strong> release of cytochrome c<br />

from mitochondria: a primary site for Bcl-2 regulation of apoptosis. Science 275:1132-<br />

1136.<br />

Knoblach SM, Fan L, and Faden AI (1999) Early neuronal expression of tumor necrosis factoralpha<br />

after experimental brain injury contributes to neurological impairment. J<br />

Neuroimmunol 95:115-125.<br />

Kohler C and Schwarcz R (1983) Comparison of ibotenate and kainate neurotoxicity in rat brain:<br />

a histological study. Neuroscience 8:819-835.<br />

Korsmeyer SJ (1995) Regulators of cell death. Trends Genet 11:101-105.<br />

Kotapka MJ, Graham DI, Adams JH, and Gennarelli TA (1992) Hippocampal pathology in fatal<br />

non-missile human head injury. Acta Neuropathol (Berl) 83:530-534.<br />

Kroemer G, Zamzami N, and Susin SA (1997) Mitochondrial control of apoptosis. Immunol<br />

Today 18:44-51.<br />

Kwak SP, Young EA, Morano I, Watson SJ, and Akil H (1992) Diurnal corticotropin-releasing<br />

hormone mRNA variation in the hypothalamus exhibits a rhythm distinct from that of<br />

plasma corticosterone. Neuroendocrinology 55:74-83.<br />

Kwak SP, Patel PD, Thompson RC, Akil H, and Watson SJ (1993) 5'-Heterogeneity of the<br />

mineralocorticoid receptor messenger ribonucleic acid: differential expression and<br />

regulation of splice variants within the rat hippocampus. Endocrinology 133:2344-2350.<br />

Landfield PW and Eldridge JC (1991) <strong>The</strong> glucocorticoid hypothesis of brain aging and<br />

neurodegeneration: recent modifications. Acta Endocrinol (Copenh) 125:54-64.<br />

Landfield PW and Eldridge JC (1994) <strong>The</strong> glucocorticoid hypothesis of age-related hippocampal<br />

neurodegeneration: role of dysregulated intraneuronal calcium. Ann NY Acad Sci<br />

746:308-321.<br />

Landfield PW, Waymire JC, and Lynch G (1978) Hippocampal aging and adrenocorticoids:<br />

quantitative correlations. Science 202:1098-1102.<br />

120


Landfield PW, Baskin RK, and Pitler TA (1981) Brain aging correlates: retardation by<br />

hormonal-pharmacological treatments. Science 214:581-584.<br />

Landfield PW, Thibault O, Mazzanti ML, Porter NM, and Kerr DS (1992) Mechanisms of<br />

neuronal death in brain aging and Alzheimer's disease: role of endocrine-mediated<br />

calcium dyshomeostasis. J Neurobiol 23:1247-1260.<br />

Lee S, Williamson J, Lothman EW, Szele FG, Chesselet MF, Von Hagen S, Sapolsky RM,<br />

Mattson MP, and Christakos S (1997) Early induction of mRNA for calbindin-D28k and<br />

BDNF but not NT-3 in rat hippocampus after kainic acid treatment. Brain Res Mol Brain<br />

Res 47:183-194.<br />

Lerma J, Morales M, Vicente MA, and Herreras O (1997) Glutamate receptors of the kainate<br />

type and synaptic transmission. Trends Neurosci 20:9-12.<br />

Levine RL, McIntosh TK, Lothrop DA, and Jackson BT (1980) Circadian periodicity of plasma<br />

corticosterone levels in rats subjected to hemorrhagic shock and surgical trauma. Horm<br />

Res 13:385-395.<br />

Lewen A, Matz P, and Chan PH (2000) Free radical pathways in CNS injury. J Neurotrauma<br />

17:871-890.<br />

Li P, Nijhawan D, Budihardjo I, Srinivasula SM, Ahmad M, Alnemri ES, and Wang X (1997)<br />

Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an<br />

apoptotic protease cascade. Cell 91:479-489.<br />

Lieberman SA, Oberoi AL, Gilkison CR, Masel BE, and Urban RJ (2001) Prevalence of<br />

neuroendocrine dysfunction in patients recovering from traumatic brain injury. J Clin<br />

Endocrinol Metab 86:2752-2756.<br />

Lindholm D, Dechant G, Heisenberg CP, and Thoenen H (1993) Brain-derived neurotrophic<br />

factor is a survival factor for cultured rat cerebellar granule neurons and protects them<br />

against glutamate-induced neurotoxicity. Eur J Neurosci 5:1455-1464.<br />

Lindholm D, Castren M, Hengerer B, Leingartner A, Castren E, and Thoenen H (1994)<br />

Glucocorticoids and neurotrophin gene regulation in the nervous system. Ann N Y Acad<br />

Sci 746:195-202; discussion 202-193.<br />

Lindsay RM, Wiegand SJ, Altar CA, and DiStefano PS (1994) Neurotrophic factors: from<br />

molecule to man. Trends Neurosci 17:182-190.<br />

121


Liu Z, Stafstrom CE, Sarkisian MR, Yang Y, Hori A, Tandon P, and Holmes GL (1997) Seizureinduced<br />

glutamate release in mature and immature animals: an in vivo microdialysis<br />

study. Neuroreport 8:2019-2023.<br />

Lopez E, Pozas E, Rivera R, and Ferrer I (1999) Bcl-2, Bax and Bcl-x expression following<br />

kainic acid administration at convulsant doses in the rat. Neuroscience 91:1461-1470.<br />

Lowenstein DH, Thomas MJ, Smith DH, and McIntosh TK (1992) Selective vulnerability of<br />

dentate hilar neurons following traumatic brain injury: a potential mechanistic link<br />

between head trauma and disorders of the hippocampus. J Neurosci 12:4846-4853.<br />

Lowy MT (1990) MK-801 antagonizes methamphetamine-induced decreases in hippocampal<br />

and striatal corticosteroid receptors. Brain Res 533:348-352.<br />

Lowy MT (1992) Kainic acid-induced decrease in hippocampal corticosteroid receptors. J<br />

Neurochem 58:1561-1568.<br />

Lowy MT, Wittenberg L, and Novotney S (1994) Adrenalectomy attenuates kainic acid-induced<br />

spectrin proteolysis and heat shock protein 70 induction in hippocampus and cortex. J<br />

Neurochem 63:886-894.<br />

Lu J, Moochhala S, Kaur C, and Ling E (2000) Changes in apoptosis-related protein (p53, Bax,<br />

Bcl-2 and Fos) expression with DNA fragmentation in the central nervous system in rats<br />

after closed head injury. Neurosci Lett 290:89-92.<br />

Luisi BF, Xu WX, Otwinowski Z, Freedman LP, Yamamoto KR, and Sigler PB (1991)<br />

Crystallographic analysis of the interaction of the glucocorticoid receptor with DNA.<br />

Nature 352:497-505.<br />

Magarinos AM and McEwen BS (1995) Stress-induced atrophy of apical dendrites of<br />

hippocampal CA3c neurons: involvement of glucocorticoid secretion and excitatory<br />

amino acid receptors. Neuroscience 69:89-98.<br />

Malva JO, Carvalho AP, and Carvalho CM (1998) Kainate receptors in hippocampal CA3<br />

subregion: evidence for a role in regulating neurotransmitter release. Neurochem Int<br />

32:1-6.<br />

122


Marin MC, Fernandez A, Bick RJ, Brisbay S, Buja LM, Snuggs M, McConkey DJ, von<br />

Eschenbach AC, Keating MJ, and McDonnell TJ (1996) Apoptosis suppression by bcl-2<br />

is correlated with the regulation of nuclear and cytosolic Ca2+. Oncogene 12:2259-2266.<br />

Martignoni E, Costa A, Sinforiani E, Liuzzi A, Chiodini P, Mauri M, Bono G, and Nappi G<br />

(1992) <strong>The</strong> brain as a target for adrenocortical steroids: cognitive implications.<br />

Psychoneuroendocrinology 17:343-354.<br />

Martin LJ, Kaiser A, Yu JW, Natale JE, and Al-Abdulla NA (2001) Injury-induced apoptosis of<br />

neurons in adult brain is mediated by p53-dependent and p53-independent pathways and<br />

requires Bax. J Comp Neurol 433:299-311.<br />

Masferrer JL, Seibert K, Zweifel B, and Needleman P (1992) Endogenous glucocorticoids<br />

regulate an inducible cyclooxygenase enzyme. Proc Natl Acad Sci U S A 89:3917-3921.<br />

Mattson MP and Scheff SW (1994) Endogenous neuroprotection factors and traumatic brain<br />

injury: mechanisms of action and implications for therapy. J Neurotrauma 11:3-33.<br />

Mattson MP, Goodman Y, Luo H, Fu W, and Furukawa K (1997) Activation of NF-kappaB<br />

protects hippocampal neurons against oxidative stress-induced apoptosis: evidence for<br />

induction of manganese superoxide dismutase and suppression of peroxynitrite<br />

production and protein tyrosine nitration. J Neurosci Res 49:681-697.<br />

McCormick JA, Lyons V, Jacobson MD, Noble J, Diorio J, Nyirenda M, Weaver S, Ester W,<br />

Yau JL, Meaney MJ, Seckl JR, and Chapman KE (2000) 5'-heterogeneity of<br />

glucocorticoid receptor messenger RNA is tissue specific: differential regulation of<br />

variant transcripts by early-life events. Mol Endocrinol 14:506-517.<br />

<strong>McCullers</strong> DL and Herman JP (1998) Mineralocorticoid receptors regulate bcl-2 and p53 mRNA<br />

expression in hippocampus. Neuroreport 9:3085-3089.<br />

<strong>McCullers</strong> DL and Herman JP (2001) Adrenocorticosteroid receptor blockade and excitotoxic<br />

challenge regulate adrenocorticosteroid receptor mRNA levels in hippocampus. J<br />

Neurosci Res 64:277-283.<br />

<strong>McCullers</strong> DL, Sullivan, P.G., Scheff, S.W., and Herman, J.P. (2001a) Traumatic brain injury<br />

regulates adrenocorticosteroid receptor mRNA levels in rat hippocampus. Submitted.<br />

123


<strong>McCullers</strong> DL, Sullivan, P.G., Scheff, S.W., and Herman, J.P. (2001b) Mifepristone protects<br />

CA1 hippocampal neurons following traumatic brain injury in rat. Neuroscience, in press.<br />

McEwan IJ, Wright AP, and Gustafsson JA (1997) Mechanism of gene expression by the<br />

glucocorticoid receptor: role of protein-protein interactions. Bioessays 19:153-160.<br />

McEwen BS (1992) Re-examination of the glucocorticoid hypothesis of stress and aging. Prog<br />

Brain Res 93:365-381; discussion 382-363.<br />

McEwen BS and Stellar E (1993) Stress and the individual. Mechanisms leading to disease. Arch<br />

Intern Med 153:2093-2101.<br />

McEwen BS, De Kloet ER, and Rostene W (1986) Adrenal steroid receptors and actions in the<br />

nervous system. Physiol Rev 66:1121-1188.<br />

McEwen BS, Biron CA, Brunson KW, Bulloch K, Chambers WH, Dhabhar FS, Goldfarb RH,<br />

Kitson RP, Miller AH, Spencer RL, and Weiss JM (1997) <strong>The</strong> role of adrenocorticoids as<br />

modulators of immune function in health and disease: neural, endocrine and immune<br />

interactions. Brain Res Brain Res Rev 23:79-133.<br />

McIntosh LJ and Sapolsky RM (1996a) Glucocorticoids may enhance oxygen radical-mediated<br />

neurotoxicity. Neurotoxicology 17:873-882.<br />

McIntosh LJ and Sapolsky RM (1996b) Glucocorticoids increase the accumulation of reactive<br />

oxygen species and enhance adriamycin-induced toxicity in neuronal culture. Exp Neurol<br />

141:201-206.<br />

McIntosh LJ, Cortopassi KM, and Sapolsky RM (1998a) Glucocorticoids may alter antioxidant<br />

enzyme capacity in the brain: kainic acid studies. Brain Res 791:215-222.<br />

McIntosh TK, Saatman KE, and Raghupathi R (1997) Calcium and the pathogenesis of traumatic<br />

CNS injury: cellular and molecular mechanisms. <strong>The</strong> Neuroscientist 3:169-175.<br />

McIntosh TK, Juhler M, and Wieloch T (1998b) Novel pharmacologic strategies in the treatment<br />

of experimental traumatic brain injury: 1998. J Neurotrauma 15:731-769.<br />

McIntosh TK, Lothrop DA, Lee A, Jackson BT, Nabseth D, and Egdahl RH (1981) Circadian<br />

rhythm of cortisol is altered in postsurgical patients. J Clin Endocrinol Metab 53:117-<br />

122.<br />

124


McIntosh TK, Saatman KE, Raghupathi R, Graham DI, Smith DH, Lee VM, and Trojanowski<br />

JQ (1998c) <strong>The</strong> Dorothy Russell Memorial Lecture. <strong>The</strong> molecular and cellular sequelae<br />

of experimental traumatic brain injury: pathogenetic mechanisms. Neuropathol Appl<br />

Neurobiol 24:251-267.<br />

Meijer OC and de Kloet ER (1998) Corticosterone and serotonergic neurotransmission in the<br />

hippocampus: functional implications of central corticosteroid receptor diversity. Crit<br />

Rev Neurobiol 12:1-20.<br />

Merry DE and Korsmeyer SJ (1997) Bcl-2 gene family in the nervous system. Annu Rev<br />

Neurosci 20:245-267.<br />

Miesfeld R, Rusconi S, Godowski PJ, Maler BA, Okret S, Wikstrom AC, Gustafsson JA, and<br />

Yamamoto KR (1986) Genetic complementation of a glucocorticoid receptor deficiency<br />

by expression of cloned receptor cDNA. Cell 46:389-399.<br />

Minami M, Kuraishi Y, and Satoh M (1991) Effects of kainic acid on messenger RNA levels of<br />

IL-1 beta, IL-6, TNF alpha and LIF in the rat brain. Biochem Biophys Res Commun<br />

176:593-598.<br />

Mitchell JB, Betito K, Rowe W, Boksa P, and Meaney MJ (1992) Serotonergic regulation of<br />

type II corticosteroid receptor binding in hippocampal cell cultures: evidence for the<br />

importance of serotonin-induced changes in cAMP levels. Neuroscience 48:631-639.<br />

Miyashita T and Reed JC (1995) Tumor suppressor p53 is a direct transcriptional activator of the<br />

human bax gene. Cell 80:293-299.<br />

Miyashita T, Krajewski S, Krajewska M, Wang HG, Lin HK, Liebermann DA, Hoffman B, and<br />

Reed JC (1994) Tumor suppressor p53 is a regulator of bcl-2 and bax gene expression in<br />

vitro and in vivo. Oncogene 9:1799-1805.<br />

Mocchetti I and Wrathall JR (1995) Neurotrophic factors in central nervous system trauma. J<br />

Neurotrauma 12:853-870.<br />

Morrison B, 3rd, Eberwine JH, Meaney DF, and McIntosh TK (2000) Traumatic injury induces<br />

differential expression of cell death genes in organotypic brain slice cultures determined<br />

by complementary DNA array hybridization. Neuroscience 96:131-139.<br />

Morrison RS, Wenzel HJ, Kinoshita Y, Robbins CA, Donehower LA, and Schwartzkroin PA<br />

(1996) Loss of the p53 tumor suppressor gene protects neurons from kainate-induced cell<br />

death. J Neurosci 16:1337-1345.<br />

125


Morse JK and Davis JN (1990) Regulation of ischemic hippocampal damage in the gerbil:<br />

adrenalectomy alters the rate of CA1 cell disappearance. Exp Neurol 110:86-92.<br />

Munck A, Guyre PM, and Holbrook NJ (1984) Physiological functions of glucocorticoids in<br />

stress and their relation to pharmacological actions. Endocr Rev 5:25-44.<br />

Nadler JV, Perry BW, and Cotman CW (1978) Intraventricular kainic acid preferentially<br />

destroys hippocampal pyramidal cells. Nature 271:676-677.<br />

Nair SM, Werkman TR, Craig J, Finnell R, Joels M, and Eberwine JH (1998) Corticosteroid<br />

regulation of ion channel conductances and mRNA levels in individual hippocampal CA1<br />

neurons. J Neurosci 18:2685-2696.<br />

Nakamura M, Raghupathi R, Merry DE, Scherbel U, Saatman KE, and McIntosh TK (1999)<br />

Overexpression of Bcl-2 is neuroprotective after experimental brain injury in transgenic<br />

mice. J Comp Neurol 412:681-692.<br />

Napieralski JA, Raghupathi R, and McIntosh TK (1999) <strong>The</strong> tumor-suppressor gene, p53, is<br />

induced in injured brain regions following experimental traumatic brain injury. Brain Res<br />

Mol Brain Res 71:78-86.<br />

Nichols NR and Finch CE (1994) Gene products of corticosteroid action in hippocampus. Ann N<br />

Y Acad Sci 746:145-154; discussion 154-146.<br />

Nilsson P, Hillered L, Olsson Y, Sheardown MJ, and Hansen AJ (1993) Regional changes in<br />

interstitial K+ and Ca2+ levels following cortical compression contusion trauma in rats. J<br />

Cereb Blood Flow Metab 13:183-192.<br />

Nitta A, Ohmiya M, Sometani A, Itoh M, Nomoto H, Furukawa Y, and Furukawa S (1999)<br />

Brain-derived neurotrophic factor prevents neuronal cell death induced by corticosterone.<br />

J Neurosci Res 57:227-235.<br />

Nordeen SK, Bona BJ, Beck CA, Edwards DP, Borror KC, and DeFranco DB (1995) <strong>The</strong> two<br />

faces of a steroid antagonist: when an antagonist isn't. Steroids 60:97-104.<br />

O'Donnell D and Meaney MJ (1994) Aldosterone modulates glucocorticoid receptor binding in<br />

hippocampal cell cultures via the mineralocorticoid receptor. Brain Res 636:49-54.<br />

O'Donnell D, Baccichet A, Seckl JR, Meaney MJ, and Poirier J (1993) Entorhinal cortex lesions<br />

transiently alter glucocorticoid but not mineralocorticoid receptor gene expression in the<br />

rat hippocampus. J Neurochem 61:356-359.<br />

126


Okiyama K, Rosenkrantz TS, Smith DH, Gennarelli TA, and McIntosh TK (1994) (S)-emopamil<br />

attenuates acute reduction in regional cerebral blood flow following experimental brain<br />

injury. J Neurotrauma 11:83-95.<br />

Oltvai ZN, Milliman CL, and Korsmeyer SJ (1993) Bcl-2 heterodimerizes in vivo with a<br />

conserved homolog, Bax, that accelerates programmed cell death. Cell 74:609-619.<br />

Orti E, Bodwell JE, and Munck A (1992) Phosphorylation of steroid hormone receptors. Endocr<br />

Rev 13:105-128.<br />

Otto C, Reichardt HM, and Schutz G (1997) Absence of glucocorticoid receptor-beta in mice. J<br />

Biol Chem 272:26665-26668.<br />

Packan DR and Sapolsky RM (1990) Glucocorticoid endangerment of the hippocampus: tissue,<br />

steroid and receptor specificity. Neuroendocrinology 51:613-618.<br />

Pan G, O'Rourke K, and Dixit VM (1998a) Caspase-9, Bcl-XL, and Apaf-1 form a ternary<br />

complex. J Biol Chem 273:5841-5845.<br />

Pan G, Humke EW, and Dixit VM (1998b) Activation of caspases triggered by cytochrome c in<br />

vitro. FEBS Lett 426:151-154.<br />

Panegyres PK and Hughes J (1998) <strong>The</strong> neuroprotective effects of the recombinant interleukin-1<br />

receptor antagonist rhIL-1ra after excitotoxic stimulation with kainic acid and its<br />

relationship to the amyloid precursor protein gene. J Neurol Sci 154:123-132.<br />

Parthasarathy S, Morales AJ, and Murphy AA (1994) Antioxidant: a new role for RU-486 and<br />

related compounds. J Clin Invest 94:1990-1995.<br />

Paskitti ME, McCreary BJ, and Herman JP (2000) Stress regulation of adrenocorticosteroid<br />

receptor gene transcription and mRNA expression in rat hippocampus: time-course<br />

analysis. Brain Res Mol Brain Res 80:142-152.<br />

Patel PD, Sherman TG, Goldman DJ, and Watson SJ (1989) Molecular cloning of a<br />

mineralocorticoid (type I) receptor complementary DNA from rat hippocampus. Mol<br />

Endocrinol 3:1877-1885.<br />

Pavlides C, Watanabe Y, and McEwen BS (1993) Effects of glucocorticoids on hippocampal<br />

long-term potentiation. Hippocampus 3:183-192.<br />

Pavlides C, Watanabe Y, Magarinos AM, and McEwen BS (1995) Opposing roles of type I and<br />

type II adrenal steroid receptors in hippocampal long-term potentiation. Neuroscience<br />

68:387-394.<br />

127


Penkowa M, Molinero A, Carrasco J, and Hidalgo J (2001) Interleukin-6 deficiency reduces the<br />

brain inflammatory response and increases oxidative stress and neurodegeneration after<br />

kainic acid-induced seizures. Neuroscience 102:805-818.<br />

Pennypacker KR and Hong JS (1995) Kainate-induced changes in gene expression in the rat<br />

hippocampus. Prog Brain Res 105:105-116.<br />

Pentelenyi T and Kammerer L (1984) Alterations of the serum cortisol and blood glucose in<br />

brain-injured patients. Injury 15:397-402.<br />

Phillips RG, Lawrence MS, Ho DY, and Sapolsky RM (2000) Limitations in the neuroprotective<br />

potential of gene therapy with Bcl-2. Brain Res 859:202-206.<br />

Picard D and Yamamoto KR (1987) Two signals mediate hormone-dependent nuclear<br />

localization of the glucocorticoid receptor. Embo J 6:3333-3340.<br />

Povlishock JT and Christman CW (1995) <strong>The</strong> pathobiology of traumatically induced axonal<br />

injury in animals and humans: a review of current thoughts. J Neurotrauma 12:555-564.<br />

Pratt WB (1993) <strong>The</strong> role of heat shock proteins in regulating the function, folding, and<br />

trafficking of the glucocorticoid receptor. J Biol Chem 268:21455-21458.<br />

Pulsinelli WA, Brierley JB, and Plum F (1982) Temporal profile of neuronal damage in a model<br />

of transient forebrain ischemia. Ann Neurol 11:491-498.<br />

Qiao X, Hughes PE, Venero JL, Dugich-Djordjevic MM, Nichols NR, Hefti F, and Knusel B<br />

(1996) NT-4/5 protects against adrenalectomy-induced apoptosis of rat hippocampal<br />

granule cells. Neuroreport 7:682-686.<br />

Raghupathi R, Graham DI, and McIntosh TK (2000) Apoptosis after traumatic brain injury. J<br />

Neurotrauma 17:927-938.<br />

Raghupathi R, Fernandez SC, Murai H, Trusko SP, Scott RW, Nishioka WK, and McIntosh TK<br />

(1998) BCL-2 overexpression attenuates cortical cell loss after traumatic brain injury in<br />

transgenic mice. J Cereb Blood Flow Metab 18:1259-1269.<br />

Ratka A, Sutanto W, Bloemers M, and de Kloet ER (1989) On the role of brain<br />

mineralocorticoid (type I) and glucocorticoid (type II) receptors in neuroendocrine<br />

regulation. Neuroendocrinology 50:117-123.<br />

Reagan LP and McEwen BS (1997) Controversies surrounding glucocorticoid-mediated cell<br />

death in the hippocampus. J Chem Neuroanat 13:149-167.<br />

128


Reiheld CT, Teyler TJ, and Vardaris RM (1984) Effects of corticosterone on the<br />

electrophysiology of hippocampal CA1 pyramidal cells in vitro. Brain Res Bull 12:349-<br />

353.<br />

Represa A, Tremblay E, and Ben-Ari Y (1987) Kainate binding sites in the hippocampal mossy<br />

fibers: localization and plasticity. Neuroscience 20:739-748.<br />

Reul JM and de Kloet ER (1985) Two receptor systems for corticosterone in rat brain:<br />

microdistribution and differential occupation. Endocrinology 117:2505-2511.<br />

Reul JM, van den Bosch FR, and de Kloet ER (1987) Relative occupation of type-I and type-II<br />

corticosteroid receptors in rat brain following stress and dexamethasone treatment:<br />

functional implications. J Endocrinol 115:459-467.<br />

Reul JM, Pearce PT, Funder JW, and Krozowski ZS (1989) Type I and type II corticosteroid<br />

receptor gene expression in the rat: effect of adrenalectomy and dexamethasone<br />

administration. Mol Endocrinol 3:1674-1680.<br />

Rivest S (1999) What is the cellular source of prostaglandins in the brain in response to systemic<br />

inflammation Facts and controversies. Mol Psychiatry 4:500-507.<br />

Roberts AJ and Keith LD (1994a) Sensitivity of the circadian rhythm of kainic acid-induced<br />

convulsion susceptibility to manipulations of corticosterone levels and mineralocorticoid<br />

receptor binding. Neuropharmacology 33:1087-1093.<br />

Roberts AJ and Keith LD (1994b) Mineralocorticoid receptors mediate the enhancing effects of<br />

corticosterone on convulsion susceptibility in mice. J Pharmacol Exp <strong>The</strong>r 270:505-511.<br />

Roof RL and Hall ED (2000) Gender differences in acute CNS trauma and stroke:<br />

neuroprotective effects of estrogen and progesterone. J Neurotrauma 17:367-388.<br />

Rosewicz S, McDonald AR, Maddux BA, Goldfine ID, Miesfeld RL, and Logsdon CD (1988)<br />

Mechanism of glucocorticoid receptor down-regulation by glucocorticoids. J Biol Chem<br />

263:2581-2584.<br />

Ross DT, Graham DI, and Adams JH (1993) Selective loss of neurons from the thalamic<br />

reticular nucleus following severe human head injury. J Neurotrauma 10:151-165.<br />

Rothman SM and Olney JW (1987) Excitotoxicity and the NMDA receptor. Trends Neurosci<br />

10:299-302.<br />

129


Rudge JS, Mather PE, Pasnikowski EM, Cai N, Corcoran T, Acheson A, Anderson K, Lindsay<br />

RM, and Wiegand SJ (1998) Endogenous BDNF protein is increased in adult rat<br />

hippocampus after a kainic acid induced excitotoxic insult but exogenous BDNF is not<br />

neuroprotective. Exp Neurol 149:398-410.<br />

Ruhmann-Wennhold A and Nelson DH (1977) Plasma ACTH levels in stressed and nonstressed<br />

adrenalectomized rats. Ann N Y Acad Sci 297:498-508.<br />

Sakahira H, Enari M, and Nagata S (1998) Cleavage of CAD inhibitor in CAD activation and<br />

DNA degradation during apoptosis. Nature 391:96-99.<br />

Sakhi S, Sun N, Wing LL, Mehta P, and Schreiber SS (1996) Nuclear accumulation of p53<br />

protein following kainic acid-induced seizures. Neuroreport 7:493-496.<br />

Sakhi S, Bruce A, Sun N, Tocco G, Baudry M, and Schreiber SS (1994) p53 induction is<br />

associated with neuronal damage in the central nervous system. Proc Natl Acad Sci USA<br />

91:7525-7529.<br />

Sanderson KL, Raghupathi R, Saatman KE, Martin D, Miller G, and McIntosh TK (1999)<br />

Interleukin-1 receptor antagonist attenuates regional neuronal cell death and cognitive<br />

dysfunction after experimental brain injury. J Cereb Blood Flow Metab 19:1118-1125.<br />

Santhakumar V, Bender R, Frotscher M, Ross ST, Hollrigel GS, Toth Z, and Soltesz I (2000)<br />

Granule cell hyperexcitability in the early post-traumatic rat dentate gyrus: the 'irritable<br />

mossy cell' hypothesis. J Physiol 524:117-134.<br />

Sapolsky RM (1985) A mechanism for glucocorticoid toxicity in the hippocampus: increased<br />

neuronal vulnerability to metabolic insults. J Neurosci 5:1228-1232.<br />

Sapolsky RM (1986a) Glucocorticoid toxicity in the hippocampus. Temporal aspects of synergy<br />

with kainic acid. Neuroendocrinology 43:440-444.<br />

Sapolsky RM (1986b) Glucocorticoid toxicity in the hippocampus: reversal by supplementation<br />

with brain fuels. J Neurosci 6:2240-2244.<br />

Sapolsky RM (1990) Glucocorticoids, hippocampal damage and the glutamatergic synapse. Prog<br />

Brain Res 86:13-23.<br />

Sapolsky RM (1994) <strong>The</strong> physiological relevance of glucocorticoid endangerment of the<br />

hippocampus. Ann N Y Acad Sci 746:294-304; discussion 304-297.<br />

Sapolsky RM and Pulsinelli WA (1985) Glucocorticoids potentiate ischemic injury to neurons:<br />

therapeutic implications. Science 229:1397-1400.<br />

130


Sapolsky RM, Krey LC, and McEwen BS (1984) Stress down-regulates corticosterone receptors<br />

in a site-specific manner in the brain. Endocrinology 114:287-292.<br />

Sapolsky RM, Krey LC, and McEwen BS (1985) Prolonged glucocorticoid exposure reduces<br />

hippocampal neuron number: implications for aging. J Neurosci 5:1222-1227.<br />

Sapolsky RM, Krey LC, and McEwen BS (1986) <strong>The</strong> neuroendocrinology of stress and aging:<br />

the glucocorticoid cascade hypothesis. Endocr Rev 7:284-301.<br />

Sapolsky RM, Packan DR, and Vale WW (1988) Glucocorticoid toxicity in the hippocampus: in<br />

vitro demonstration. Brain Res 453:367-371.<br />

Sapolsky RM, Romero LM, and Munck AU (2000) How do glucocorticoids influence stress<br />

responses Integrating permissive, suppressive, stimulatory, and preparative actions.<br />

Endocr Rev 21:55-89.<br />

Sarrieau A, O'Donnell D, Alonso R, and Meaney MJ (1996) Regulation of glucocorticosteroid<br />

receptor expression in rat hippocampal cell cultures by nerve growth factor. Neurosci<br />

Lett 206:207-211.<br />

Schaaf MJ, Hoetelmans RW, de Kloet ER, and Vreugdenhil E (1997) Corticosterone regulates<br />

expression of BDNF and trkB but not NT-3 and trkC mRNA in the rat hippocampus. J<br />

Neurosci Res 48:334-341.<br />

Scheff SW and Cotman CW (1982) Chronic glucocorticoid therapy alters axon sprouting in the<br />

hippocampal dentate gyrus. Exp Neurol 76:644-654.<br />

Scheff SW and Sullivan PG (1999) Cyclosporin A significantly ameliorates cortical damage<br />

following experimental traumatic brain injury in rodents. J Neurotrauma 16:783-792.<br />

Scheff SW, Hoff SF, and Anderson KJ (1986) Altered regulation of lesion-induced<br />

synaptogenesis by adrenalectomy and corticosterone in young adult rats. Exp Neurol<br />

93:456-470.<br />

Scheff SW, Baldwin SA, Brown RW, and Kraemer PJ (1997) Morris water maze deficits in rats<br />

following traumatic brain injury: lateral controlled cortical impact. J Neurotrauma<br />

14:615-627.<br />

Scheinman RI, Cogswell PC, Lofquist AK, and Baldwin AS, Jr. (1995) Role of transcriptional<br />

activation of I kappa B alpha in mediation of immunosuppression by glucocorticoids.<br />

Science 270:283-286.<br />

131


Scherbel U, Raghupathi R, Nakamura M, Saatman KE, Trojanowski JQ, Neugebauer E, Marino<br />

MW, and McIntosh TK (1999) Differential acute and chronic responses of tumor necrosis<br />

factor-deficient mice to experimental brain injury. Proc Natl Acad Sci USA 96:8721-<br />

8726.<br />

Schmidt-Kastner R and Freund TF (1991) Selective vulnerability of the hippocampus in brain<br />

ischemia. Neuroscience 40:599-636.<br />

Scully JL and Otten U (1995) Neurotrophin expression modulated by glucocorticoids and<br />

oestrogen in immortalized hippocampal neurons. Brain Res Mol Brain Res 31:158-164.<br />

Seckl JR and Fink G (1992) Antidepressants increase glucocorticoid and mineralocorticoid<br />

receptor mRNA expression in rat hippocampus in vivo. Neuroendocrinology 55:621-626.<br />

Seckl JR, Dickson KL, and Fink G (1990) Central 5,7-dihydroxytryptamine lesions decrease<br />

hippocampal glucocorticoid and mineralocorticoid receptor messenger ribonucleic acid<br />

expression. J Neuroendocrinol 2:911-916.<br />

Seroogy KB and Herman JP (1996) In situ hybridization approaches to the study of the nervous<br />

system. In: Neurochemistry - A Practical Approach (2nd ed.) (Turner AJ, Bachelard HS,<br />

eds), pp 121-150: Oxford University Press.<br />

Shapira Y, Shohami E, Sidi A, Soffer D, Freeman S, and Cotev S (1988) Experimental closed<br />

head injury in rats: mechanical, pathophysiologic, and neurologic properties. Crit Care<br />

Med 16:258-265.<br />

Sheppard KE (1994) Calcium and protein kinase C regulation of the glucocorticoid receptor in<br />

mouse corticotrope tumor cells. J Steroid Biochem Mol Biol 48:337-345.<br />

Sheppard KE, Roberts JL, and Blum M (1991) Adrenocorticotropin-releasing factor downregulates<br />

glucocorticoid receptor expression in mouse corticotrope tumor cells via an<br />

adenylate cyclase-dependent mechanism. Endocrinology 129:663-670.<br />

Shigemoto R, Kinoshita A, Wada E, Nomura S, Ohishi H, Takada M, Flor PJ, Neki A, Abe T,<br />

Nakanishi S, and Mizuno N (1997) Differential presynaptic localization of metabotropic<br />

glutamate receptor subtypes in the rat hippocampus. J Neurosci 17:7503-7522.<br />

Shohami E, Bass R, Trembovler V, and Weidenfeld J (1995) <strong>The</strong> effect of the adrenocortical<br />

axis upon recovery from closed head injury. J Neurotrauma 12:1069-1077.<br />

132


Shohami E, Bass R, Wallach D, Yamin A, and Gallily R (1996) Inhibition of tumor necrosis<br />

factor alpha (TNFalpha) activity in rat brain is associated with cerebroprotection after<br />

closed head injury. J Cereb Blood Flow Metab 16:378-384.<br />

Shohami E, Gallily R, Mechoulam R, Bass R, and Ben-Hur T (1997) Cytokine production in the<br />

brain following closed head injury: dexanabinol (HU-211) is a novel TNF-alpha inhibitor<br />

and an effective neuroprotectant. J Neuroimmunol 72:169-177.<br />

Siman R, Noszek JC, and Kegerise C (1989) Calpain I activation is specifically related to<br />

excitatory amino acid induction of hippocampal damage. J Neurosci 9:1579-1590.<br />

Sipe KJ, Hynds DL, Snow D, and Herman JP (2000) Temporal correlation between cAMPinduced<br />

changes in calcium release and decreased glucocorticoid receptor expression in<br />

C6 glioma cells. Society for Neuroscience Abstracts 26:414.<br />

Sloviter RS (1989) Calcium-binding protein (calbindin-D28k) and parvalbumin<br />

immunocytochemistry: localization in the rat hippocampus with specific reference to the<br />

selective vulnerability of hippocampal neurons to seizure activity. J Comp Neurol<br />

280:183-196.<br />

Sloviter RS (1992) Possible functional consequences of synaptic reorganization in the dentate<br />

gyrus of kainate-treated rats. Neurosci Lett 137:91-96.<br />

Sloviter RS, Valiquette G, Abrams GM, Ronk EC, Sollas AL, Paul LA, and Neubort S (1989)<br />

Selective loss of hippocampal granule cells in the mature rat brain after adrenalectomy.<br />

Science 243:535-538.<br />

Smith DH, Lowenstein DH, Gennarelli TA, and McIntosh TK (1994) Persistent memory<br />

dysfunction is associated with bilateral hippocampal damage following experimental<br />

brain injury. Neurosci Lett 168:151-154.<br />

Smith MA, Makino S, Kvetnansky R, and Post RM (1995a) Stress and glucocorticoids affect the<br />

expression of brain-derived neurotrophic factor and neurotrophin-3 mRNAs in the<br />

hippocampus. J Neurosci 15:1768-1777.<br />

Smith MA, Makino S, Kvetnansky R, and Post RM (1995b) Effects of stress on neurotrophic<br />

factor expression in the rat brain. Ann NY Acad Sci 771:234-239.<br />

133


Smith-Swintosky VL, Pettigrew LC, Sapolsky RM, Phares C, Craddock SD, Brooke SM, and<br />

Mattson MP (1996) Metyrapone, an inhibitor of glucocorticoid production, reduces brain<br />

injury induced by focal and global ischemia and seizures. J Cereb Blood Flow Metab<br />

16:585-598.<br />

Spencer RL, Young EA, Choo PH, and McEwen BS (1990) Adrenal steroid type I and type II<br />

receptor binding: estimates of in vivo receptor number, occupancy, and activation with<br />

varying level of steroid. Brain Res 514:37-48.<br />

Spencer RL, Kim PJ, Kalman BA, and Cole MA (1998) Evidence for mineralocorticoid receptor<br />

facilitation of glucocorticoid receptor-dependent regulation of hypothalamic-pituitaryadrenal<br />

axis activity. Endocrinology 139:2718-2726.<br />

Sperk G, Lassmann H, Baran H, Kish SJ, Seitelberger F, and Hornykiewicz O (1983) Kainic acid<br />

induced seizures: neurochemical and histopathological changes. Neuroscience 10:1301-<br />

1315.<br />

Steer JH, Vuong Q, and Joyce DA (1997) Suppression of human monocyte tumour necrosis<br />

factor-alpha release by glucocorticoid therapy: relationship to systemic monocytopaenia<br />

and cortisol suppression. Br J Clin Pharmacol 43:383-389.<br />

Stein BA and Sapolsky RM (1988) Chemical adrenalectomy reduces hippocampal damage<br />

induced by kainic acid. Brain Res 473:175-180.<br />

Stein-Behrens B, Mattson MP, Chang I, Yeh M, and Sapolsky R (1994a) Stress exacerbates<br />

neuron loss and cytoskeletal pathology in the hippocampus. J Neurosci 14:5373-5380.<br />

Stein-Behrens BA, Lin WJ, and Sapolsky RM (1994b) Physiological elevations of<br />

glucocorticoids potentiate glutamate accumulation in the hippocampus. J Neurochem<br />

63:596-602.<br />

Stein-Behrens BA, Elliott EM, Miller CA, Schilling JW, Newcombe R, and Sapolsky RM (1992)<br />

Glucocorticoids exacerbate kainic acid-induced extracellular accumulation of excitatory<br />

amino acids in the rat hippocampus. J Neurochem 58:1730-1735.<br />

Stienstra CM and Joels M (2000) Effect of corticosteroid treatment in vitro on adrenalectomyinduced<br />

impairment of synaptic transmission in the rat dentate gyrus. J Neuroendocrinol<br />

12:199-205.<br />

Sullivan PG, Thompson MB, and Scheff SW (1999) Cyclosporin A attenuates acute<br />

mitochondrial dysfunction following traumatic brain injury. Exp Neurol 160:226-234.<br />

134


Sullivan PG, Keller JN, Mattson MP, and Scheff SW (1998) Traumatic brain injury alters<br />

synaptic homeostasis: implications for impaired mitochondrial and transport function. J<br />

Neurotrauma 15:789-798.<br />

Sun AY, Cheng Y, and Sun GY (1992) Kainic acid-induced excitotoxicity in neurons and glial<br />

cells. Prog Brain Res 94:271-280.<br />

Sutanto W and de Kloet ER (1991) Mineralocorticoid receptor ligands: biochemical,<br />

pharmacological, and clinical aspects. Med Res Rev 11:617-639.<br />

Sutton RL, Lescaudron L, and Stein DG (1993) Unilateral cortical contusion injury in the rat:<br />

vascular disruption and temporal development of cortical necrosis. J Neurotrauma<br />

10:135-149.<br />

Takahashi M, Billups B, Rossi D, Sarantis M, Hamann M, and Attwell D (1997) <strong>The</strong> role of<br />

glutamate transporters in glutamate homeostasis in the brain. J Exp Biol 200:401-409.<br />

Talmi M, Carlier E, Bengelloun W, and Soumireu-Mourat B (1995) Synergistic action of<br />

corticosterone on kainic acid-induced electrophysiological alterations in the<br />

hippocampus. Brain Res 704:97-102.<br />

Tarazi FI, Florijn WJ, and Creese I (1996) Regulation of ionotropic glutamate receptors<br />

following subchronic and chronic treatment with typical and atypical antipsychotics.<br />

Psychopharmacology (Berl) 128:371-379.<br />

Taupin V, Toulmond S, Serrano A, Benavides J, and Zavala F (1993) Increase in IL-6, IL-1 and<br />

TNF levels in rat brain following traumatic lesion. Influence of pre- and post-traumatic<br />

treatment with Ro5 4864, a peripheral-type (p site) benzodiazepine ligand. J<br />

Neuroimmunol 42:177-185.<br />

Timsit S, Rivera S, Ouaghi P, Guischard F, Tremblay E, Ben-Ari Y, and Khrestchatisky M<br />

(1999) Increased cyclin D1 in vulnerable neurons in the hippocampus after ischaemia and<br />

epilepsy: a modulator of in vivo programmed cell death Eur J Neurosci 11:263-278.<br />

Tomasevic G, Raghupathi R, Oga M, Scherbel U, Wieloch T, and McIntosh TK (1999)<br />

Experimental TBI in mice lacking the tumor suppressor p53 gene. J Neurotrauma 16:999.<br />

Tornheim PA and McLaurin RL (1978) Effect of dexamethasone on cerebral edema from cranial<br />

impact in the cat. J Neurosurg 48:220-227.<br />

Toulmond S and Rothwell NJ (1995) Interleukin-1 receptor antagonist inhibits neuronal damage<br />

caused by fluid percussion injury in the rat. Brain Res 671:261-266.<br />

135


Trapp T, Rupprecht R, Castren M, Reul JM, and Holsboer F (1994) Heterodimerization between<br />

mineralocorticoid and glucocorticoid receptor: a new principle of glucocorticoid action in<br />

the CNS. Neuron 13:1457-1462.<br />

Truettner J, Schmidt-Kastner R, Busto R, Alonso <strong>OF</strong>, Loor JY, Dietrich WD, and Ginsberg MD<br />

(1999) Expression of brain-derived neurotrophic factor, nerve growth factor, and heat<br />

shock protein HSP70 following fluid percussion brain injury in rats. J Neurotrauma<br />

16:471-486.<br />

Tsunashima K, Schwarzer C, Kirchmair E, Sieghart W, and Sperk G (1997) GABA(A) receptor<br />

subunits in the rat hippocampus III: altered messenger RNA expression in kainic acidinduced<br />

epilepsy. Neuroscience 80:1019-1032.<br />

Tymianski M and Tator CH (1996) Normal and abnormal calcium homeostasis in neurons: a<br />

basis for the pathophysiology of traumatic and ischemic central nervous system injury.<br />

Neurosurgery 38:1176-1195.<br />

Unlap T and Jope RS (1995a) Diurnal variation in kainate-induced AP-1 activation in rat brain:<br />

influence of glucocorticoids. Brain Res Mol Brain Res 28:193-200.<br />

Unlap T and Jope RS (1995b) Inhibition of NFkB DNA binding activity by glucocorticoids in rat<br />

brain. Neurosci Lett 198:41-44.<br />

van Eekelen JA, Kiss JZ, Westphal HM, and de Kloet ER (1987) Immunocytochemical study on<br />

the intracellular localization of the type 2 glucocorticoid receptor in the rat brain. Brain<br />

Res 436:120-128.<br />

Van Eekelen JA, Jiang W, De Kloet ER, and Bohn MC (1988) Distribution of the<br />

mineralocorticoid and the glucocorticoid receptor mRNAs in the rat hippocampus. J<br />

Neurosci Res 21:88-94.<br />

van Haarst AD, Oitzl MS, and de Kloet ER (1997) Facilitation of feedback inhibition through<br />

blockade of glucocorticoid receptors in the hippocampus. Neurochem Res 22:1323-1328.<br />

Vreugdenhil E, de Jong J, Schaaf MJ, Meijer OC, Busscher J, Vuijst C, and de Kloet ER (1996)<br />

Molecular dissection of corticosteroid action in the rat hippocampus. Application of the<br />

differential display techniques. J Mol Neurosci 7:135-146.<br />

Wahl M, Unterberg A, Baethmann A, and Schilling L (1988) Mediators of blood-brain barrier<br />

dysfunction and formation of vasogenic brain edema. J Cereb Blood Flow Metab 8:621-<br />

634.<br />

136


Werkman TR, Van der Linden S, and Joels M (1997) Corticosteroid effects on sodium and<br />

calcium currents in acutely dissociated rat CA1 hippocampal neurons. Neuroscience<br />

78:663-672.<br />

West MJ, Slomianka L, and Gundersen HJ (1991) Unbiased stereological estimation of the total<br />

number of neurons in thesubdivisions of the rat hippocampus using the optical<br />

fractionator. Anat Rec 231:482-497.<br />

White-Gbadebo D and Hamm RJ (1993) Chronic corticosterone treatment potentiates deficits<br />

following traumatic brain injury in rats: implications for aging. J Neurotrauma 10:297-<br />

306.<br />

Won SJ, Ko HW, Kim EY, Park EC, Huh K, Jung NP, Choi I, Oh YK, Shin HC, and Gwag BJ<br />

(1999) Nuclear factor kappa B-mediated kainate neurotoxicity in the rat and hamster<br />

hippocampus. Neuroscience 94:83-91.<br />

Wong GH and Goeddel DV (1988) Induction of manganous superoxide dismutase by tumor<br />

necrosis factor: possible protective mechanism. Science 242:941-944.<br />

Woolley CS, Gould E, and McEwen BS (1990) Exposure to excess glucocorticoids alters<br />

dendritic morphology of adult hippocampal pyramidal neurons. Brain Res 531:225-231.<br />

Woolley CS, Gould E, Sakai RR, Spencer RL, and McEwen BS (1991) Effects of aldosterone or<br />

RU28362 treatment on adrenalectomy-induced cell death in the dentate gyrus of the adult<br />

rat. Brain Res 554:312-315.<br />

Xiang H, Hochman DW, Saya H, Fujiwara T, Schwartzkroin PA, and Morrison RS (1996)<br />

Evidence for p53-mediated modulation of neuronal viability. J Neurosci 16:6753-6765.<br />

Xu J, Qu ZX, Hogan EL, and Perot PL (1992) Protective effect of methylprednisolone on<br />

vascular injury in rat spinal cord injury. J Neurotrauma 9:245-253.<br />

Yabuuchi K, Minami M, Katsumata S, and Satoh M (1993) In situ hybridization study of<br />

interleukin-1 beta mRNA induced by kainic acid in the rat brain. Brain Res Mol Brain<br />

Res 20:153-161.<br />

Yamamoto KR (1985) Steroid receptor regulated transcription of specific genes and gene<br />

networks. Annu Rev Genet 19:209-252.<br />

Yang J, Liu X, Bhalla K, Kim CN, Ibrado AM, Cai J, Peng TI, Jones DP, and Wang X (1997)<br />

Prevention of apoptosis by Bcl-2: release of cytochrome c from mitochondria blocked.<br />

Science 275:1129-1132.<br />

137


Yang K, Perez-Polo JR, Mu XS, Yan HQ, Xue JJ, Iwamoto Y, Liu SJ, Dixon CE, and Hayes RL<br />

(1996) Increased expression of brain-derived neurotrophic factor but not neurotrophin-3<br />

mRNA in rat brain after cortical impact injury. J Neurosci Res 44:157-164.<br />

Yau JL and Seckl JR (1992) Central 6-hydroxydopamine lesions decrease mineralocorticoid, but<br />

not glucocorticoid receptor gene expression in the rat hippocampus. Neurosci Lett<br />

142:159-162.<br />

Yau JL, Kelly PA, and Seckl JR (1994) Increased glucocorticoid receptor gene expression in the<br />

rat hippocampus following combined serotonergic and medial septal cholinergic lesions.<br />

Brain Res Mol Brain Res 27:174-178.<br />

Yau JL, Noble J, and Seckl JR (1999) Continuous blockade of brain mineralocorticoid receptors<br />

impairs spatial learning in rats. Neurosci Lett 277:45-48.<br />

Yonish-Rouach E, Resnitzky D, Lotem J, Sachs L, Kimchi A, and Oren M (1991) Wild-type p53<br />

induces apoptosis of myeloid leukaemic cells that is inhibited by interleukin-6. Nature<br />

352:345-347.<br />

Young W (1992) Role of calcium in central nervous system injuries. J Neurotrauma 9:S9-25.<br />

Zennaro MC, Keightley MC, Kotelevtsev Y, Conway GS, Soubrier F, and Fuller PJ (1995)<br />

Human mineralocorticoid receptor genomic structure and identification of expressed<br />

isoforms. J Biol Chem 270:21016-21020.<br />

Zhang X, Gelowitz DL, Lai CT, Boulton AA, and Yu PH (1997) Gradation of kainic acidinduced<br />

rat limbic seizures and expression of hippocampal heat shock protein-70. Eur J<br />

Neurosci 9:760-769.<br />

Zhu RL, Graham SH, Jin J, Stetler RA, Simon RP, and Chen J (1997) Kainate induces the<br />

expression of the DNA damage-inducible gene, GADD45, in the rat brain. Neuroscience<br />

81:707-720.<br />

Zipfel GJ, Babcock DJ, Lee JM, and Choi DW (2000) Neuronal apoptosis after CNS injury: the<br />

roles of glutamate and calcium. J Neurotrauma 17:857-869.<br />

Zou H, Henzel WJ, Liu X, Lutschg A, and Wang X (1997) Apaf-1, a human protein homologous<br />

to C. elegans CED-4, participates in cytochrome c-dependent activation of caspase-3.<br />

Cell 90:405-413.<br />

138


<strong>Deanna</strong> <strong>Lynn</strong> <strong>McCullers</strong><br />

Date of Birth: September 24, 1974<br />

Place of Birth: Newport News, Virginia<br />

Education: B.S. Psychology, 1995<br />

Virginia Polytechnic Institute and State University<br />

Research and Professional Experience:<br />

1995 - 1996 Undergraduate research assistant<br />

Virginia Polytechnic Institute and State University<br />

1996 - 2001 Graduate student<br />

University of Kentucky<br />

Department of Anatomy and Neurobiology<br />

1998 - 1999 Teaching assistant, MD 817, Neurosciences<br />

University of Kentucky College of Medicine<br />

1999 Teaching assistant and guest lecturer<br />

ANA 636, Advanced Neuroanatomy<br />

University of Kentucky<br />

Department of Anatomy and Neurobiology<br />

2000 – 2001 Guest lecturer, ANA 206/209, Principles of Human Anatomy<br />

University of Kentucky<br />

Department of Anatomy and Neurobiology<br />

Awards:<br />

1996-1997 Hughes Predoctoral Fellowship<br />

1997-1998 Hughes Predoctoral Fellowship<br />

1998-1999 Hughes Predoctoral Fellowship<br />

1998-1999 Kentucky Research Challenge Trust Fund Fellowship<br />

1999-2001 NIH Training Grant:<br />

Molecular and Cellular Aspects of Brain Aging<br />

139


Publications and Abstracts<br />

<strong>McCullers</strong>, D.L., Sullivan, P.G., Scheff, S.W., and Herman, J.P. (2001) RU486 protects CA1<br />

hippocampal neurons following traumatic brain injury in rat. In press, Neuroscience.<br />

<strong>McCullers</strong>, D.L., Sullivan, P.G., Scheff, S.W., and Herman, J.P. (2001) Traumatic brain injury<br />

regulates adrenocorticosteroid receptor mRNA levels in rat hippocampus. Submitted.<br />

<strong>McCullers</strong> D. L. and Herman, J.P. (2001) Adrenocorticosteroid receptor blockade and<br />

excitotoxic challenge regulate adrenocorticosteroid receptor mRNA levels in hippocampus.<br />

Journal of Neuroscience Research 64: 277-283.<br />

Pedersen, W.A., <strong>McCullers</strong>, D.L., Culmsee, C., Haughey, N.J., Herman, J.P., and Mattson, M.P.<br />

(2001) Corticotropin-releasing hormone protects neurons against insults relevant to the<br />

pathogenesis of Alzheimer’s disease. Neurobiology of Disease 8(3): 492-503.<br />

<strong>McCullers</strong>, D.L., and Herman, J.P. (1998) Mineralocorticoid receptors regulate bcl-2 and p53<br />

mRNA expression in hippocampus. Neuroreport 9(13), 3085-3089.<br />

<strong>McCullers</strong>, D.L. and Herman, J.P. (2001) Adrenalectomy with corticosterone replacement<br />

prevents restraint stress-induced downregulation of glucocorticoid receptor mRNA levels in<br />

CA1 of rat hippocampus. Society for Neuroscience Abstracts 27.<br />

<strong>McCullers</strong>, D.L., Sullivan, P.G., Scheff, S.W., and Herman, J.P. (2000) Effects of<br />

adrenocorticosteroid receptor blockade and traumatic brain injury on hippocampal neuron<br />

viability. Society for Neuroscience Abstracts 26: 419.<br />

<strong>McCullers</strong>, D.L. and Herman, J.P. (1999) Mineralocorticoid receptor blockade<br />

and kainic acid-induced changes in viability-related gene expression. Society for<br />

Neuroscience Abstracts 25: 615.<br />

140


<strong>McCullers</strong>, D.L., and Herman, J.P. (1998) Blockade of mineralocorticoid receptors increases<br />

hippocampal cell death following kainic acid treatment. Society for Neuroscience Abstracts<br />

24: 1380.<br />

<strong>McCullers</strong>, D.L., Watson, R.E., and Herman, J.P. (1997) Blockade of mineralocorticoid<br />

receptors alters transcription of the apoptosis-related genes bcl-2 and bax. Society For<br />

Neuroscience Abstracts 23: 1242.<br />

<strong>Deanna</strong> L. <strong>McCullers</strong><br />

141

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!