<<

Health Science Campus

FINAL APPROVAL OF DISSERTATION Doctor of Philosophy in Biomedical Sciences

The Role of Calcium and Mitochondria in the Etiology and Treatment of Three Different Disease Paradigms

Submitted by: Christine Brink

In partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biomedical Sciences

Examination Committee

Major Advisor: David Giovannucci, Ph.D.

Academic Linda Dokas, Ph.D. Advisory Committee: Joseph Margiotta, Ph.D.

Andrew Beavis, Ph.D.

Ana Marie Oyarce, Ph.D.

L. John Greenfield, M.D., Ph.D.

Senior Associate Dean College of Graduate Studies Michael S. Bisesi, Ph.D.

Date of Defense: December 7, 2007

The Role of Calcium and Mitochondria in the Etiology and Treatment of Three

Different Disease Paradigms

Christine A. Brink

University of Toledo, Health Science Campus

2007

Acknowledgements

First, I would like to thank my advisor, Dr. David Giovannucci, for his

encouragement and support throughout my graduate training. His advice and

understanding have been essential to my success over the past 5 years. I will

always be grateful for the education he has provided to me, and the compassion

he has shown toward me.

Next, I would like to thank Jenny Giovannucci for her illustrations and

encouragement. I would also like to thank Rebecca Pierson for providing the

primary cortical neuronal cultures, and countless words of advice. Additionally, I

would like to thank Christian Peters for his counsel and assistance throughout

our training. In addition, I would like to thank my committee members, Dr.

Greenfield, Dr. Dokas, Dr. Margiotta, Dr. Beavis, and Dr. Oyarce, for their

contributions and guidance.

Lastly, I would like to express my unending gratitude toward my family. My

husband’s patience and encouragement have been indispensable during this time. Also, the love and support of my parents were critical in preparing me to take on this incredible and fascinating journey.

i Abbreviations Used

IP3 1,4,5,triphosphate

GABA γ-aminobutyric acid

Allo

AD Alzheimer’s disease

Aβ Amyloid-β

APP Amyloid precursor protein

ALS Amyotrophic lateral sclerosis

Ca2+ Calcium

CRAC Calcium release activated channel

CCE Capacitative calcium entry

FCCP Carbonylcyanide-4-(trifluoromethoxy)-phenylhydrazone

CAI Carboxyamidotriazole

Icrac Current from CRAC channel

CIF Cytochrome c efflux inducing factor

DMEM Delbeco’s Modified Eagle Medium

EC50 Effective concentration at 50%

ER Endoplasmic reticulum

EDTA Ethylenediamine-tetraacetic acid disodium salt

EGTA Ethleneglycol-O, O’ bis(2aminoethyl)N,N,N’,N’-tetraacetic acid

GFAP Glial fibrillary acidic protein

HVA High voltage activated

ii HN Hypothalamic neurohypophysis

HPA Hypothalamic pituitary axis

ISS Intercellular saline solution

LDP Long term depression

LTP Long term potentiation

LVA Low voltage activated

HEDTA N-(2-hydroxyethlenediamine-N, N’, N’-triacetic acid trisodium salt

PIP2 Phosphatidylinositol biphosphate

PLC Phospholipase C

PSS Physiological saline solution

PKC Protein kinase C

SR Sarcoplasmic reticulum

SERCA Sarcoplasmic/endoplasmic reticulum calcium ATPase

SCID Severe combined immunodeficency syndrome

SNARES Soluble N-ethylmalemide sensitive factor protein receptor

SOC Store operated channel

TMRM Tetramethyl-rhodamine methyl ester perchlorate

TRP Transient receptor protein

VGCC Voltage gated

iii

Table of Contents

Preliminaries

Page

Acknowledgments i

Abbreviations Page ii

Table of Contents iv

Chapters

Page

I Introduction 1

II Methods 6

III Background and Literature Review 15

IV Results 39

V Results 56

VI Results 78

VII Conclusions 103

Bibliography 106

Abstract 139

iv Chapter I: Introduction

Calcium (Ca2+) is an important cellular messenger involved in a variety of critical messaging pathways. From cell proliferation to apoptosis, Ca2+ is involved in multiple facets of cell health. Even in the most primitive prokaryotes maintaining the Ca2+ gradient is important, and cytosolic free Ca2+ concentrations are controlled (Case, 2007) . In eukaryotes, Ca2+’s importance is demonstrated early in development, as it controls fertilization and cellular differentiation (Lory et al., 2006). After development, Ca2+ plays a critical role through its control of gene transcription, and is responsible for a myriad of messages including: control of movement, synaptic plasticity, wound healing, insulin secretion and a host of other important functions (Berridge et al., 1998). Consequently, to control all these signaling pathways, Ca2+ uses a diverse range of signaling mechanisms to convey the proper signal for the desired outcome.

To achieve specificity of the signals, cells have devised a multitude of means to separate the messages. This is accomplished by a variety of types, locations, amplitudes, kinetics and specific sources of Ca2+ signals. For example,

Ca2+ can enter the cytosol through the plasma membrane from a variety of types of Ca2+ channels or transporters, or Ca2+ can be released from internal stores through the action of second messengers like inositol 1,4,5-triphosphate (IP3) on the intracellular release channel, the IP3 Receptor (Broad et al., 1999).

Localization of Ca2+ signals is often manifested as microdomains classified by the types of channels involved. Specific examples of these are: sparklets from voltage-operated channels, sparks from ryanodine receptors, blinks which are

1 decreases in ER Ca2+ due to release, and puffs which are responsible for Ca2+ waves induced by IP3 receptors (Berridge, 2006). In addition, there are two types

of Ca2+ waves, intracellular and intercellular, and each of these wave types is

responsible for different signals (Bootman et al., 2001). There are also a

multitude of types of Ca2+ channels that help with specificity of the Ca2+

message. These include: voltage-gated channels (L-type, T-type, R-type, N-

type), pumps and exchangers, Ca2+ ATPases, receptor operated channels,

second messenger operated channels, and channels on mitochondria and

internal stores. Additionally, each of these sources of Ca2+ signaling may be

linked to specific physiological outcomes.

However, because Ca2+ signaling is so critical and utilizes such diverse pathways, disturbing the critical balance of Ca2+ levels can be devastating to cell health and physiological cell function. Correspondingly, many diseases involve an interruption or unbalance in a Ca2+-signaling pathway. For example,

Alzheimer’s Disease (AD), cancer, diabetes, and Parkinson’s disease, have all

been linked to a possible deregulation of Ca2+; and in some cases altering the

Ca2+ signaling has been considered a possible treatment mechanism. Ca2+ is

known to be involved in regulation of the cell cycle. For example, Kao et al (1990)

demonstrated that chelating cytosolic Ca2+ blocked the transition from

metaphase to anaphase in three different cell types, showing that altering

physiological Ca2+ can alter the cell-cycle. Additionally, Guderman and Roelle

(2006) found evidence suggesting that disruption of Ca2+ signaling was a key

event in the abnormal growth of some small cell lung cancers. Furthermore,

2 many neurodegenerative disorders have focused on Ca2+ dysregulation having a

role in the death of . Ca2+ is a major player in the pathways of apoptosis,

having a role in both the intrinsic and extrinsic pathways of apoptosis, including

those induced by ROS, UV-fragmentation of DNA, and cytotoxic drugs

(Hajnoczky et al., 2003). Apoptosis is suspected to play a role in neuronal death

in diseases such as AD. Additionally, there are multiple diseases and disorders that involve defects of specific Ca2+ channels. For example, genetic defects in

voltage gated Ca2+ channels can result in disorders involving the musculature,

neurological, cardiac and vision systems. However, in order to exploit Ca2+

signaling pathways for treatment, a better understanding of the pathways must

be established.

One important element in understanding Ca2+ signaling is the role of

mitochondria. The ability of mitochondria to uptake Ca2+ was first discovered in

the 1960’s (Babcock et al., 1997). Mitochondria sequester Ca2+ through the

mitochondrial uniporter, a highly selective pore on the mitochondria that utilizes

the membrane potential of mitochondria to drive the uptake of positively charged

Ca2+ ions (Kirichok et al., 2004). The mitochondrial membrane potential is established through the electron gradient created by the proton-motive force, a mechanism in which positively charged hydrogen ions are moved from inside the inner mitochondrial membrane to outside the membrane. Thus, this process creates a negative charge on the inside of the mitochondrial membrane, which is the driving force for influx of positively charged Ca2+ ions. Therefore, the

3 mitochondrial membrane potential is essential for the maintenance of Ca2+

signaling in mitochondria.

Furthermore, mitochondria are intimately involved in shaping many Ca2+

signals (Duchen, 2000; Rizzuto et al., 2000). However, the role of mitochondria in

Ca2+ signaling pathways was not fully appreciated until the discovery of Ca2+ microdomains, discussed previously. For example, the role of mitochondria in capacitative Ca2+ entry has recently been identified. When Ca2+ is released from

internal stores, store operated channels (SOC) on the plasma membrane are

activated, causing them to open and influx more Ca2+. High levels of Ca2+ around the SOC can deactivate the channels, causing them to close. Mitochondria, however, can take up the excess Ca2+ around the SOC keeping the levels low

enough for the channel to remain open, allowing more Ca2+ to enter the cell.

Furthermore, mitochondrial Ca2+ can also be used to signal cellular events like apoptosis. Activation of the mitochondrial permeability transition pore (PTP) from

Ca2+ overload or other apoptotic signals can cause the initiation of the

mitochondrial apoptosis pathway, which will be further discussed later in this

paper.

In this body of work, I investigated to role of Ca2+ and mitochondria in

three different disease paradigms. The first is capacitative Ca2+ entry and the

mechanism of a drug that alters this pathway to treat some types of cancer. The

second is the toxic effects of a peptide, Aβ, thought to be involved in AD. Finally,

I investigated the mechanisms of Ca2+ entry in primary cortical neurons after

acute response to the allopregnanolone and the neurotransmitter

4 GABA, and the possible role of this pathway in catamenial epilepsy. Recent evidence has suggested a possible role of Ca2+ in the neurosteroid regulation of

GABA receptors indicating a role for Ca2+ in modulating epilepsy. Tus I dissected

the role of Ca2+and mitochondria in three separate disease states in order to

develop a better understanding of Ca2+ signaling and mitochondrial health in

disease etiology and treatment. I hypothesize that in each of these three disease

models, mitochondrial health and calcium signaling are ultimately involved in the

disease etiology.

5 Chapter II – Materials and Methods

This section contains material and methods for all reported data for this body of

work. Please refer to this section for all information regarding experimental

techniques.

Cell Preparations

HEK Cell Culture

HEK-293 cells stably expressing the M3 muscarinic receptor were cultured in Delbecco’s modified eagle medium (DMEM) supplemented with 10% fetal calf

serum and (200,000 U/mL) and streptomycin (200,000 U/mL). These

are non-malignant cells.

Preparation of single nerve endings

Isolated nerve terminals were prepared from male Sprague Dawley rats

(~250 g) following CO2 asphyxiation and decapitation. The neural and intermediate lobes were separated from the anterior lobe of the pituitary under physiological saline solution (PSS) (for ingredients see solution section). The neurointermediate lobe was removed from the neural lobe, and isolated terminals were prepared by trituration in 150 µl of buffer that contained (in mM) 270 sucrose, 0.01 Ethleneglycol-O, O’ bis(2aminoethyl)N,N,N’,N’-tetraacetic acid

(EGTA), 10 Tris-HEPES, pH 7.2. Terminals were allowed to adhere to clean

6 glass coverslips that formed the bottom of a culture/recording dish and used or

maintained in culture for up to 48hrs.

Primary culture of pituicytes

The isolated posterior pituitary was cut into approximately 6-8 small pieces

in PSS. Equal parts of thrombin and chicken plasma were used to make clots on

clean glass coverslips placed in a 6 well dish and a tissue piece was inserted into

each clot. After 3 to 5 minutes, DMEM containing 1% penicillin/streptomycin, and

10% fetal calf serum were added. Pituicytes grew out from the clot, and were maintained for 2 weeks at 37˚C, changing the media every 2-3 days. To confirm that our cultures were primarily pituicytes we fixed the cells and preformed immunocytochemistry with an antibody against glial fibrillary acidic protein

(GFAP) in combination with trypan blue staining (for total cell number) and found

>99% of the viable cells stained with trypan blue also had the GFAP marker.

Cortical Cell Culture

Primary cortical cells were prepared by Rebecca Pierson. Cortices from

E16 rat pups were removed and placed in Hank’s Balanced Salt Solution

(HBSS). All non-cortical tissue was removed and cortices were kept on ice.

Tissue was spun for 5 minutes, supernatant was removed and 0.25% trypsin-

EDTA was added and incubated for 11 minutes to break up the tissue. Trypsin

was removed and tissue was rinsed 3 times with S minimal essential media

(SMEM). After the third rinse cells were suspended in complete SMEM (SMEM

7 with 10% horse serum and 10% fetal bovine serum). Cells were plated at 1 x 106 cells/mL, then placed in an incubator and kept at 37ºC and 5% CO2. Cells were

grown for 2 weeks before imaging experiments were preformed. For neurons chronically treated with allopregnanolone, 1nM allopregnanolone was applied to

the cell SMEM at day 12 in culture and imaging experiments were preformed at

day 14 in culture.

Imaging

Wide Field Fluorescence Microscopy

Ca2+ measurements were obtained using a pressure driven perfusion

system. Wide fluorescence imaging was preformed in a Nikon TE2000-S inverted

fluorescence microscope (Nikon, Melville NY) equipped with a monochrometer-

based imaging system (TILL Photonics, Martinsried Germany) and a Nikon 40X

SuperFluor oil-immersion objective lens, NA = 1.3.

Confocal microscopy

Time lapsed 4-D images were acquired using a Leica TCS SP5

broadband confocal microscope (Leica, Mannheim, Germany) system coupled to

a DMI 6000CS inverted microscope equipped with Argon-488 and diode pumped

solid state-561 laser sources. Confocal z-series were obtained with a 20x

objective NA = 0.07 excited with 488 nm or 561 nm lasers. Acquisition of emitted

light was optimized using a tunable SP detector. Pituicytes and nerve endings

8 were loaded with tetramethylrhodamine methyl ester perchlorate (TMRM) at 1μM for 15 minutes at room temperature in the dark. Fluorescently tagged Aβ was added to media 15 minutes before imaging began. Cells were imaged every 15 minutes for 12 hours. During imaging, cells were kept at 37°C in a 5% CO2

environment using a stage top environmental Luden chamber. Analysis was achieved using Leica software and IGOR PRO.

Dyes and Loading

Cytosolic Calcium Measurements

2+ 2+ The ratiometric [Ca ]c indicator dye Fura-2 AM was used to measure Ca

influx in pituicytes and nerve endings. Both were loaded in PSS containing 1μM

Fura for 30 minutes at room temperature in the dark. Following loading cells were

placed back in dye free PSS and placed in a perfusion chamber on the same

inverted microscope as previously described. (Various stimuli were applied to the

cells to induce Ca2+ influx.) The cells were excited with 340nm and 380nm light

and emission light was collected at 540nm. The ratio of 340/380 was determined

using T.I.L.L. Vision software.

Mitochondrial Calcium measurements

Cells were grown on clean 25 mm glass coverslips, which formed the

bottom of a perfusion chamber. Cells were loaded with 2 μM Rhod-2 AM, Rhod-

FF AM, Rhod-5N AM mixture for 30 minutes in the dark at room temperature.

9 Cells were then permeabilized for 3 minutes at 37˚C in an intracellular saline solution (ISS) containing 10 μM digitonin, but with no added Ca2+ (for ingredients

of ISS see solutions section). Intracellular Ca2+ solutions were obtained by

adding N-(2-hydroxyethlenediamine-N, N’, N’-triacetic acid trisodium salt

(HEDTA)/ Ca2+. EGTA (250μM) was added to the “zero” Ca2+ solution, pH 6.8.

The intracellular Ca2+ solutions contained between 3-3000 mM Ca2+. All fluorescence data was converted to ΔF/F0=100[(F-F0)/F0], were F is the recorded

fluorescence and F0 is the average of the first 15 frames of data. Full frame

images were collected at 1-second intervals for at least 400 seconds and

changes are expressed as the percentage of increase compared with F0.

TMRM digital imaging analysis

Cells were loaded with either 12.5 nM or 1.5 μM TMRM in PSS for 15

minutes at 37˚C in the dark. Cells were washed with PSS and kept at room

temperature for 30 minutes before measuring to allow the dye to concentrate in

the mitochondria. Following loading, coverslips were mounted in chambers and

monitored by digital fluorescence imaging at 1Hz.

JC-1 optical measurement

Cells plated on 25 mM coverglass were incubated in phosphate-buffered

saline (PBS) supplemented with 2μg/mL JC-1 (5,5’,6,6’-tetrachloro-1,1’,3,3’-

tetraethylbenzimidazolylcarbocyanine iodide) for 30 minutes at room temperature

in the dark. Cells were then rinsed with PSS and kept at room temperature for 30

10 minutes to allow for dye to accumulate in the mitochondria. Coverslips were then

mounted in chambers and monitored by digital fluorescence imaging.

Solutions

Physiological Saline Solution: (PSS) containing 140 mM NaCl, 5mM KCl, 1mM

MgCl2, 2.2 mM CaCl2, 10mM HEPES, 10mM glucose, pH 7.2.

Internal Saline Solutions: (ISS) contained 130mM KCl, 10mM NaCl, 1mM

K3PO4, 1mM ATP, 0.02 mM ADP, 2mM succinate, 20mM HEPES, 2mM MgCl2

(adjusted to buffer Ca2+). Intracellular Ca2+ solutions were obtained by adding

HEDTA/ Ca2+. EGTA (250μM) was added to the “zero” Ca2+ solution, pH 6.8.

Hank’s Balanced Salt Solution: (HBSS) contained 7.5% Sodium Bicarbonate,

1% 10mg/mL Gentamicin and 3.574g HEPES, pH 7.3.

SMEM: contained Minimum Essential Medium with 2.703g glucose, 10mM l- , and 5% penicillin/streptomycin.

Preparation of beta-amyloid peptides

Aβ oligomers (Aβ25-35, Aβ35-25 from Sigma or Fluorescent Aβ1-42 from

AnaSpec) were prepared by reconstituting the lyophilized power with ddH20,

11 aliquoted and left at room temperature overnight (10-12 hrs) before being stored

at –20°C in the dark.

CAI: CAI (carboxyamidotriazole) stock was made up in DMSO and was diluted in

PSS to final concentration. CAI was a gift from the Drug Synthesis and Chemistry

Branch, National Cancer Institute (Bethesda, MD).

Neurosteroids: Allopregnanolone and inactive allopregnanolone (Sigma) were

made up at a stock concentration of 1μM in DMSO and was diluted to final concentration in PSS.

GABA: γ-aminobutyric acid (GABA) was made up in DMSO at 100μM and was

diluted in PSS to final concentration.

Plate reader

Mitochondrial Calcium Analysis

HEK-293 cells were plated in 24 well plates and grown to confluency, then

loaded with 2 μM Rhod-2 AM, Rhod-FF AM, Rhod-5N AM, and 15 nM

Mitotracker Green. Cells were then permeabilized with 10 μM digitonin in 0 Ca2+

ISS for 3 minutes at 37ºC. These conditions were optimized using digital fluorescence imaging (See previous section). Cells were then rinsed and placed in an ISS solution with a known Ca2+ amount from 3-3000 μM Ca2+. A Fluostar

12 Optima multi-well plate reader (BMG Labtechnologies) was used to detect the

fluorescence. Cells were excited at 550 and 485 and emissions were monitored

at 590 and 520.

TMRM Plate Reader analysis

HEK-293 cells were loaded with 1.5 µM TMRM in PSS for 15 minutes at

room temperature in the dark. Cells were then washed with fresh dye-free PSS

and spun down at 300 g for 5 minutes. Cells were the resuspended in 24 mL

PSS and aliquoted into a 24 well plate at 1 mL per well and allowed to rest for 30 minutes prior to monitoring by fluorescence plate reader. Cells were excited with

544nM light and emission was measured at 595 nM. Wells were measured at 1- minute intervals for 10 minutes prior to treatment. Wells were treated with 0 μM

CAI (control) 10 μM CAI of 20 μM carbonyl cyanide 4-

trifluoromethoxyphenylhydrazone (FCCP) and monitoring was resumed for an

additional 30 minutes.

JC-1 Plate reader analysis

Cells were incubated in phosphate-buffered saline (PBS) supplemented

with 2μg/mL JC-1 for 30 minutes at room temperature in the dark. Cells were

then centrifuged at 1200 g for 5 minutes and resuspended in dye free

physiological saline solutions (PSS) containing 140 mM NaCl, 5mM KCl, 1mM

MgCl2, 2.2 mM CaCl2, 10mM HEPES, 10mM glucose, pH 7.2. The cells were

then placed in 48 well culture plates (Corning plates 3548). A Fluostar Optima

13 multi-well plate reader (BMG Labtechnologies) was used to detect the

fluorescence emission at 590 nM and 520 nM in response to alternating 550 nM and 485 nM excitation, respectively.

Statistics

All data was analyzed with student t-test using GraphPad Prism 4, expressed as mean ± SEM. Significance was set as p<0.05

14 Chapter III: Background and Significance

This work focuses on three different Ca2+ signaling mechanisms, with an emphasis on their role in disease etiologies and mitochondrial involvement. First,

I reviewed capacitative Ca2+ entry and the role of this pathway in cell proliferation and the corresponding relationship to cancer. Secondly, I have discussed Ca2+ signaling at the synapse and the role of glial cells in modulating Ca2+ signals in neurons, as the synapse is the primary target in AD. I have also discussed the role of Ca2+ in neurodegeneration through excitotoxicity and apoptosis.

Additionally, I addressed voltage gated Ca2+ channels and their role in long-term potentiation and long-term depression in neurons. Thirdly, I have discussed the possible role of voltage gated Ca2+ channels in modulating GABA receptor subunit expression and the relationship of this to disease as they may play a role in catamenial epilepsy. Lastly, I have discussed the involvement of mitochondria in each of these signaling pathways or disease pathologies.

Capacitative Calcium Entry

Capacitative Ca2+ entry (CCE) was first described by Putney in 1986 as the continuous loading and discharge of an intracellular Ca2+ store (Putney,

2007). Currently CCE is believed to be a mechanism used by cells to replenish

Ca2+ stores when levels get too low. The Ca2+ store (endoplasmic reticulum, ER or sarcoplasmic reticulum, SR for example) extends some unknown signal to a channel on the plasma membrane (referred to as a store operated channel or

SOC), which allows the entry of Ca2+ into the cytosol from the extracellular

15 space. Entering Ca2+ is bound to mobile and immobile buffers or sequestered

into the ER or other Ca2+ sinks like the mitochondria in the area of the SOC.

When Ca2+ levels within the ER or Ca2+ store are replenished, the cytosolic Ca2+

levels rise and deactivate the SOC via Ca2+ dependent inactivation.

Evidence supporting this theory began when Takemura (1989) showed

that IP3 is not necessary for CCE by introducing the use of thapsigargin.

Thapsigargin allows the release of Ca2+ from internal stores without activating

IP3. Thapsigargin poisons the SERCA (sarcoplasmic/endoplasmic reticulum

calcium ATPase) pump of the Ca2+ store. The SERCA pump is a Ca2+ ATPase

on the ER/SR membrane, which binds calcium and transports it into the lumen of

the ER/SR. When it is poisoned by thapsigargin, it cannot pump Ca2+ into the

ER/SR. Thus, by treating cells with thapsigargin and observing CCE, Takemura

demonstrated that the release of Ca2+ from internal stores alone was enough to

2+ activate CCE; IP3 was not necessary. Additionally, this revealed that the Ca

entry was from across the plasma membrane (as experiments done in the

absence of external Ca2+ did not show SOC) and not a cycle of uptake and

release of Ca2+ from internal stores, as originally proposed.

The next step was to identify how the Ca2+ was entering the cell. The Ca2+

2+ current created by CCE is known as Icrac (Ca release activated current) and was discovered by Hoth and Penner (1992). Hoth and Penner showed a highly selective Ca2+ current that was vastly different from other Ca2+ currents

suggesting the molecular structure was also different from other known Ca2+

16 channels. Once the current was identified, investigations looking at the proteins

responsible for the SOC and mechanisms of activation of this channel began.

Initial investigations pointed to the TRP family of proteins. Hardie and

Minke (1993) were first to identify that the TRP channel was a Ca2+ channel

through a Drosophila mutant. In Drosophilia the TRP channel is involved in Ca2+

influx in response to light entering the retina. In the mutant Drosophilia, it was

found that a defect in the TRP gene caused a deficiency in the Ca2+ influx that

lead to a transient and unsustained response to light (Parekh and Putney, 2005).

Because it was discovered that TRP was activated downstream of PLC, and SOC entry was also known to be activated downstream of PLC, investigation of a possible link between SOC and TRP channels began (Parekh and Putney

2005). Three major subfamilies of TRP channels have been identified in mammals: TRPC (canonical), TRPV (vanniliod), TRPM (melastatin) and a

diverse list of functions for these channels have been established including a role

as thermal and noxious receptors and mechanoreceptors (Huang, 2004). TRPC has the greatest homology to Drosophilia and have been the most studied in

terms of a relationship to SOC and Icrac. TRP channels have 6 transmembrane

domains and probably assemble in homo or hetero tetramers to form cation

selective channels (Nilius and Voets, 2005). Most are permeable to Ca2+ and provide Ca2+ influx. Some have been shown to reside on intracellular membranes

(Huang, 2004).

TRPC have been shown to be activated by stimulation of isoforms of PLC.

They have been shown to be responsible for agonist-stimulated Ca2+ influx in

17 most cell types (Kiselyov et al., 2005). Most experiments studying CCE have

used either over expression of TRP channels or blocking endogenous TRP

channels. Expression in Sf9 insect cells demonstrated a greater Ca2+ influx after

thapsigargin exposure. However, it was eventually shown that TRP channels in

Drosophila are not related to SOC, but require downstream lipid products of PLC, like PIP2 (Parekh and Putney 2005). Consequently, their role in CCE is not fully understood. However, new evidence suggests a role for another protein, Orai1.

Orai was discovered by Feske et al (2006) when one case of hereditary severe combined immune deficiency (SCID) was found to have a mutation in

Orai1. Furthermore, Feske et al (2006) showed that this mutation affected store-

operated Ca2+ entry in T-cells, thereby affecting the immune response. Orai1 is

located in the plasma membrane, making it an ideal candidate for the store

operated channel. In a study by Yeromin et al (2006), point mutations in the S1-

S2 loop of Orai transformed the Icrac current into an outward rectifying current or a

non-conduction channel depending on the mutation. Similarly, Prakriya (2006)

showed that point mutations in Orai in the highly conserved transmembrane

region caused a decrease in Ca2+ influx and an increase in monocovalent cation

current for Icrac. These studies indicated Orai has a role in the current involved in

CCE, and perhaps may be the SOC channel itself.

Several theories for activation of the SOC have been proposed including a

diffusible message, conformational coupling or exocytosis (Putney, 2007). One

factor that has been identified as a possible diffusible messenger for CCE is CIF

(Ca2+ influx factor). CIF, discovered by Randriamampita and Tsien (1993) may be

18 a viable candidate for this job; however, the identity of the protein is still

unknown. Additionally, there are several studies pointing to a role of Stim1 in

CCE. First, knock-down experiments by Liou et al (2005) and Roos et al (2005)

of Stim1 (not Stim2) resulted in a reduction (or nearly complete inhibition) of Ca2+ influx by CCE. Immunocytochemistry experiments have shown that Stim1 co- localizes with both ER and plasma membrane. Additionally, Liou et al (2005) have demonstrated that Stim1 can rearrange into puncta accumulated near the plasma membrane after Ca2+ store depletion, while mutant Stim1 did not respond

to Ca2+ depletion. This indicates that Stim1 may be functioning as the Ca2+ sensor activating the SOC at the plasma membrane.

Further evidence that Stim is the Ca2+ sensor involves the EF hand region of Stim. This region of the protein is predicted to be in the lumen of the ER and is therefore hypothesized to be the Ca2+ sensor region, monitoring the Ca2+ levels

in the Ca2+ stores (Soboloff et al., 2006). Mutations studies in the EF hand region

caused Stim to be localized into puncta before CCE activation, and Ca2+

depletion had no further effect on localization. Additionally, the mutation caused

an increase in Ca2+ influx if the stores were filled, but failed to promote SOC influx in response to store depletion (Roos et al 2005).

Moreover, Soboloff et al (2006) and Zhang et al (2006) have identified

Stim and Orai to be necessary for CRAC channel function, and co-expression of

these proteins causes a synergistic effect to form a large Icrac current. Yeromin

(2006) showed that thapsigargin treatment enhanced interaction between Orai and Stim. Additionally, Prakriya et al. (2006) demonstrated that over expression

19 of both increased Icrac. Currently, the mechanism of CCE is believed to be that

Stim1 is activated when ER Ca2+ is low; Ca2+ is then released from internal

stores and Stim1 redistributes to interact directly with Orai1, which allows Ca2+

influx through the Orai channel at the plasma membrane allowing Ca2+ to

accumulate in the cytosol and to be taken up by the internal stores.

CCE and Proliferation

Calcium has the ability to regulate both cell death and cell proliferation in

many cell types. As cancer involves cells that are growing unnaturally fast,

targeting Ca2+ pathways involved in cell proliferation or apoptosis may have a

potential therapeutic benefit in the treatment of cancer. Many studies have

demonstrated that CCE plays a role in regulating cell proliferation. First, Ca2+’s

2+ importance in mitosis was demonstrated when caged Ca and caged IP3 were used to elevate cytosolic Ca2+ artificially in some cell types (Lu and Means,

1993). The induction of anaphase was observed as confirmed by the presence of nuclear membrane breakdown and condensation of chromosomes. Conversely, when chelators such as EGTA reduced intracellular Ca2+, onset of anaphase was

blocked, as noted by lack of nuclear membrane breakdown and condensation of chromosomes (Kao et al., 1990). These results support the idea that Ca2+ may have a role in both normal and pathological cell cycle events. Additionally, the

2+ use of caged IP3 suggests the possible role of CCE specifically in Ca ’s control

of cell proliferation.

20 CCE has been shown to affect proliferation in various cell types. Waldron

et al (1994) used thapsigargin (known to trigger CCE) in DDT-MF2, a hamster

smooth muscle cell, to cause growth-arrest in these cells and then used serum to

restore cell growth. Waldron showed that SERCA pump expression was

necessary for cell re-entry into cell cycle and cell division. Golovina (1999;

Golovina et al., 2001) has shown that maintenance of the sarcoplasmic reticulum

Ca2+ levels are essential for cell proliferation and enhancement of CCE increased proliferation in human pulmonary artery smooth muscle cells. Chao (1992)

2+ 2+ showed that release of Ca from IP3 dependent Ca stores in fibroblasts by thapsigargin could activate MAP kinase activity, known to be involved in cell proliferation. Additionally, Ciapa et al. (1994) has shown that bursts of IP3 resulting in Ca2+ oscillations are necessary for final cell division.

T-lymphocytes, or T-cells, are a highly proliferating cell type. Domletsch

and Lewis (1994) were the first to show that Ca2+ stores and store operated Ca2+

channels were involved in generating Ca2+ oscillations in T-cells. Subsequently,

Dolmetsch et al. (1998) showed that these oscillations controlled gene

expression and cell differentiation by altering the frequency of the oscillations.

Under normal conditions, T-cells are activated by CCE when an antigen binds to

2+ a T-cell antigen receptor and causes the generation of IP3, which releases Ca from internal stores and causes Ca2+ influx through the plasma membrane

(Dolmetsch et al. 1998). Increases in cytosolic Ca2+ concentration can activate

transcription factors such as NFAT, NFkB, JNK1, MEK2 and CREB, which

control several diverse pathways including cell proliferation, cell differentiation,

21 and cell death (Feske et al., 2001). Feske et al showed that in T-cells from a

SCID patient, there was a defect in transmembrane Ca2+ influx after thapsigargin

treatment. These cells also exhibited defects in NFAT translocation to the nuclear

envelope as well as gene expression changes compared to T-cells from a normal

patient. Artificially increasing cytosolic Ca2+ increased NFAT translocation in

patient T-cells. Additionally, effects of CCE as measured by thapsigargin were

also seen in fibroblasts from the SCID patient.

Mitochondria and CCE

In addition to other homeostatic mechanisms, mitochondria have been

shown to be an integral part of Ca2+ signaling. Moreover, mitochondria are an

important modulator of the CCE pathway. As mentioned before, high amounts of

Ca2+ at the SOC can shut down the channel. Under normal conditions, this may

mean that the internal stores have been replenished. However, evidence has

shown that mitochondria are involved in CCE signaling. For example, Park

(2001) has shown in pancreatic acinar cells, that there is a population of

mitochondria that respond to store operated Ca2+ influx. Lawrie (1996) also

showed that the location of mitochondria is important in their regulation of Ca2+

influx in endothelial cells.

Glitch (2002) used RBL-1 cells, a rat basophil leukemia cell line, to show that IP3 failed to activate SOC entry unless mitochondria were in an energetic

state. In this same cell line, Gilabert et al. (2001) also showed that Ca2+

dependent slow-inactivation of Ca2+ influx in CCE is regulated by mitochondria.

22 This ability of mitochondria to regulate Ca2+ influx was also shown by Malli et al

(2003). Malli also demonstrated that when mitochondrial Ca2+ influx was inhibited, ER Ca2+ stores refilling was prevented in an endothelial cell line.

Hajnoczky et al (1999) used imaging experiments in hepatocytes to show similarly that blocking mitochondrial Ca2+ uptake effected ER Ca2+ uptake after

IP3 activation. Furthermore, Bakowski and Parekh (2007) found that pyruvic acid, which is a rate-limiting substrate in mitochondrial respiration, increased SOC entry by decreasing the inactivation of SOC. These experiments illustrate that mitochondria affect two distinct aspects of CCE, sensitivity to Ca2+ release from internal stores and Ca2+ dependent inactivation of Ca2+ influx through SOC channels.

Additionally, the role of mitochondria in CCE Ca2+ signaling has been directly demonstrated. Hoth (2000) used patch-clamping techniques in T-cells to show the role of mitochondria in controlling Icrac. Additionally, they demonstrated that inhibiting mitochondria reduced intracellular Ca2+ and NFAT translocation.

Nunez (2006) showed that blocking mitochondria with FCCP inhibited not only

SOC entry, but also inhibited cell proliferation in Jurkat cells and a human colon cancer cell line. Additionally, Nunez showed that treating the cells with salicylate decreased mitochondrial membrane potential and decreased cell proliferation in the cancer cell line. This evidence clearly shows a potential therapeutic target for both CCE and mitochondria involved in CCE.

23

Calcium Signaling at the Synapse

Calcium signaling at the synapse is unique, yet maintains the diverse

range of signals previously discussed in non-excitable cells. The synapse

contains three main parts, a pre-synaptic element, a post-synaptic element and

the surrounding glial cells. Here I will discuss two types of Ca2+ signaling specific

to the synapse, neurotransmitter release and long-term potentiation (LTP). Next,

I will discuss neurodegeneration and the role of glia in excitotoxicity and Ca2+ signaling in neurons. Then I will review another type of Ca2+ regulated cell death

related to neurodegeneration, apoptosis. Lastly, I will review the role of

mitochondria in regulating neurotransmitter release and apoptosis.

Calcium and Neurotransmitter Release

To begin, one of the unique Ca2+ signaling pathways in neurons is

neurotransmitter release. Neurotransmitter release is regulated by a family of

proteins known as SNARES (soluble N-ethylmalemide sensitive factor protein

receptor). There are two subgroups in this protein family, t-snares (or q-snares),

which are on the plasma membrane or intracellular organelle target membranes,

and v-snares (or r-snares), which are on the vesicle to be released. A group of v-

snares and t-snares will interact to form a SNARE protein complex. Different cells and compartments within cells have different groups of these proteins. For example, specific sets of snares are important for the release of different types of neurotransmitters (Sudhof, 1995), while others are important for transporting

24 GLUT-4 in skeletal muscle cells (Slot et al., 1997). While it was thought that Ca2+

was somehow involved in this signaling, it was not known how until Walch-

Solimena et al (1993) discovered the protein synaptotagmin. Synaptotagmin has

conserved C2A and C2B domains that can bind up to 5 Ca2+ ions (Walch-

Solimena et al., 1993). When Ca2+ enters the presynaptic axon, it binds to

synaptotagmin and the SNARE protein complex initiates the fusion of the vesicle

with the plasma membrane and the contents of the vesicle are released into the

synaptic cleft and subsequently activate receptors at the post-synaptic

membrane.

Other Unique Aspects of Calcium Signaling in Neurons

Neurons, like many other cells, utilize patterned Ca2+ signaling to activate

transcription factors. The Ca2+ dependent activation or deactivation of specific neuronal genes contributes to long-term potentiation (LTP) and long-term

depression (LTD) respectively. During these processes Ca2+ enters the cell at

specific microdomains or with specific kinetics to activate transcription factors

that can modulate LTP and LTD to regulate synaptic plasticity. Calcium entry

through NMDA receptors promotes CREB signaling by initiating phosphorylation

of CREB and CREB dependent gene-expression (Riccio and Ginty, 2002). CREB

is a nuclear transcription factor that is responsible for expression of genes with a

CRE promoter (Carrasco and Hidalgo, 2006). CREB activation from Ca2+ entry through NMDA receptors at post-synaptic microdomains has been implicated in synaptic plasticity. Calcium entry at NMDA receptors has also been shown to

25 activate ERK, and this has been implicated as a requirement for some forms of

LTP (Carrasco et al., 2002). For example, certain long term spatial memory tasks

are ERK dependent as demonstrated by Thomas and Huganir (2004). NMDA is

not the only place for Ca2+ entry that activates transcription factors; Ca2+ entry at

L-type Ca2+ channels, a voltage gated Ca2+ channel has also been shown to

activate CREB, and also ERK and NFAT (Carrasco et al. 2002). NFAT is a

cytoplasmic transcription factor that is Ca2+ dependent via calcinurin (Carrasco et

al. 2002). Unlike CREB signaling, Ca2+ activates NFAT by removing the phosphorylation, which allows translocation of NFAT to the nucleus. NFAT can

2+ also be activated by Ca released from internal stores after IP3 stimulation. In

the next section I will review voltage gated Ca2+ channels and their relationship to

GABA receptors.

Voltage Gated Calcium Channels

There are three main groups of voltage gated Ca2+ channels (VGCC):

Cav1 (L-type), Cav2 (P, N, and R-type), and Cav3 (T-type). The letter

designations are based on the channels biophysical and pharmacological

features (Bidaud et al., 2006). Channels are also identified by their voltage

sensitivity as high voltage activated (HVA) or low voltage activated (LVA). We will

be focusing on the L-type channel. L-type channels were initially identified by

their sensitivity to dihydropyridine (Kanngiesser et al., 1988). There are four main

groups of L-type channels: Cav1.1 (found in skeletal muscle), Cav1.2 and Cav1.3

(found in neurons, cardiac muscle, smooth muscle, fibroblasts, and pancreatic

26 cells), and Cav1.4 (found in retina) (Lipscombe et al., 2004). The Cav1.1 are

essential for muscular control and movement, while the Cav1.4 are important in

vision and visual discrimination. I will focus on Cav1.2 and Cav1.3 and their roles

in neuronal Ca2+ signaling.

The Cav1.2 and Cav1.3 differ in their activation sensitivity: Cav1.2 is an

HVA, while Cav1.3 is LVA (Lipscombe et al., 2002). Tanabe et al. (1987) cloned

the first channel and this information was then used to screen for other subunits.

The L-type channel is best known for it’s role in excitation coupling to muscle

contraction in skeletal muscle (Beam et al., 1989). However, important functions

of the L-type VGCC in neurons include initiation of neurotransmitter release,

control of neuronal firing patterns, neurite outgrowth, and gene expression

(Lipscombe et al., 2002); (Kim et al., 1998).

The initial criteria for identification of L-type channels included activation

by strong depolarization, high sensitivity to dihydropyridine and its agonists and

antagonists, slow activation kinetics, Ca2+ dependent inactivation, and large

single channel conductance. However, many deviations from this criteria have

since been discovered including low voltage activation channels (Cav1.3),

differences in inhibition by certain antagonists, variation in activation kinetics, and others (Lipscombe et al., 2002; Lipscombe et al., 2004). These differences may be due to splice variants and post-translational modifications.

The L-type channel is composed of four subunits, α-1, α-2δ, β, and γ (Kim

et al., 1998). There are 10 genes that encode for the α1 subunit with over 1000

theoretical variants (Gray et al., 2007). There are 6 known isoforms (Kim , Chang

27 et al. 1998). The α1 subunit expression controls temporal and spatial expression

of the channel (Kim, Chang et al. 1998). It is the pore forming unit of the channel

(Bidaud et al., 2006). Additionally, the α1 subunit is the binding site for dihydropyridine (Kim, Chang et al. 1998). Alternative splicing occurs at sites that are important for controlling channel activity and, therefore, can affect channel biophysics, density, targeting, and coupling to downstream pathways (Gray et al.,

2007). The other subunits are encoded by 4 genes for Cavβ, 3 for Cavαδ, and

eight for Cavγ (Lipscombe et al., 2002); although not all the γ variants associate

with the other VGCC subunits.

Calcium Signaling Modulation of GABA Receptor Expression

As previously discussed, Ca2+ signaling in neurons can affect gene

transcription and expression. Recent evidence has shown that Ca2+ and

depolarization of neurons can affect the expression of GABA receptor subunits.

For example, Gault (1997) demonstrated that treatment of cerebellar granular

neurons in depolarizing media (elevated potassium) caused an eight-fold

increase in the expression of the GABA receptor subunit delta, and this effect

was reduced when cells were returned to non-depolarizing media. Depolarization can occur by activation of L-type VGCC. Additionally, studies using modulators of

PKC (Ca2+-phospolipid-dependent protein kinase) demonstrated that increase in

PKC activity caused a decrease in GABA receptor activity. In addition,

Leidenheimer and Chapell (1997) showed that the effects of PKC could modulate

28 the effects of the positive GABA receptor modulator, THDOC, a neurosteroid.

The PKC pathway can be activated by Ca2+ entry at L-type VGCC

More recent studies have linked both GABA and L-type VGCC to these

effects. GABA has been shown to cause an increase in intracellular Ca2+ in various cell types and different points in development. For example, Rivera et al.

(2005) demonstrated that early in development intracellular chloride is increased and GABA is depolarizing. Chavas et al. (2004) used cerebellar intraneurons to demonstrate that GABA agonists could induce a Ca2+ rise until postnatal day 20.

Pellistri et al. (2005) then showed that the Ca2+ rise in these cells was blocked by both and , a GABA receptor blocker and L-type VGCC

blocker respectively. Additionally, in whole brain cultures, Lyons et al. (2001)

demonstrated that when blocked the GABA induced Ca2+ increase, the

down regulation of GABA receptors was also blocked. Additionally they showed

that this down regulation of GABA receptors was not seen with application of

high potassium alone, indicating that both GABA and Ca2+ influx were required

for the down regulation of GABA receptors.

Calcium and Neurodegeneration

Whereas Ca2+ has the ability to regulate cell growth and function, Ca2+

also has been linked to various forms of neurodegeneration (neuronal death).

Calcium can activate oxygenases, increasing oxidative stress altering

mitochondrial Ca2+ levels and energy metabolism (Mattson, 2007). Ca2+ can also

activate either directly or indirectly, calpains and caspases leading to degradation

29 of cytoskeletal proteins, membrane receptors and metabolic proteins (Stefanis,

2005). Additionally, Ca2+ can trigger apoptosis altering mitochondrial Ca2+ signaling by a process that will be discussed later. Calcium is also involved in excitotoxicity. Excitotoxicity is a process by which excessive glutamate release and glutamate receptor overactivation induce excessive cytosolic Ca2+ and

neuronal death (Mattson, 2007). Mitochondria are involved in excitotoxicity

because they help to buffer Ca2+ in the cytosol. Pivovarova et al (2004) has

shown that NMDA stimulation leads to an increase in cytosolic Ca2+ leading to a

subpopulation of mitochondria undergoing swelling and eventually releasing

apoptotic proteins. The role of mitochondria in neurodegeneration will be

discussed in more detail later. Astrocytes, however, may have a more significant

role in regulating glutamate and mediating excitotoxicity.

Calcium and Apoptosis

Apoptosis was first described by Carl Voght in 1842, but the name was introduced by John Kerr in 1972 to describe the formation of “apoptotic bodies” from a cell (Lawen, 2003). Characteristics of apoptosis include cell shrinkage, rounding up and loss of contacts with neighboring cells, ER swelling and formation of vesicles and vacuoles, chromatic condensation and fragmentation and cells breaking up to form “apoptotic bodies” containing the cellular contents

(Lawen 2003). While apoptosis occurs without the leakage of cellular contents, necrosis, another type of cell death, can cause leakage of the cellular contents

30 and inflammation. Excitotoxicity can lead to both necrotic and apoptotic death in neurons (Haeberlein, 2004).

Apoptotic cell death is involved in many different types of diseases including neurodegenerative disorders like AD, Parkinson’s disease, AIDS, DNA mutations in p53 and Bcl-2, auto-immune diseases, and cellular stressors and toxins such as oxidative stress and (Lawen 2003). There are two major pathways for apoptosis, the extrinsic pathway and the intrinsic pathway. The extrinsic pathway is activated by channels and proteins on the plasma membrane, which may or may not also activate the intrinsic pathway. The intrinsic pathway involves the mitochondria and will be discussed in more detail later. Calcium is involved in both pathways and there is evidence suggesting that apoptosis induced by reactive oxygen species, UV-irradiation, drugs, cytokines, and ischemia-reperfusion require Ca2+ (Hajnoczky et al., 2003).

Glial-Neuronal Crosstalk

Astrocytes have been shown to have roles in neurogenesis, neuroprotection and neurodegeneration. For example, Song et al (2002) demonstrated that adult astrocytes in hippocampus promote neurogenesis by directing stem cells to become neurons. Watts (2004) used electron microscopy to show glial cells engulfing axons during development, demonstrating a role in axon pruning. Additionally, under physiological conditions, astrocytes have a neuroprotective effect by controlling glutamate transmission and regulating glial glutamate transporters (Gegelashvili et al., 2007). Much of what is known about

31 the bi-directional signaling of glia and neurons in the CNS has been learned by

monitoring Ca2+ in glial cells (Metea and Newman, 2006). Recent evidence has

shown that glial cells play an integral role in synaptic signaling. For example,

Muller cells, a specialized glial cell in the mammalian retina, generate transient increases in Ca2+ in response to light, and these are modulated by neuronal

activity (Newman, 2005).

Additionally, Pfrieger (1997) demonstrated that glial cells are important in

the development of fully functional neuronal synapses in culture. Furthermore,

glial cells of the posterior pituitary are involved in the modulation of hormone

release from nerve endings (Hatton 2002). Moreover, hormones released by the

nerve endings of the posterior pituitary can have effects on glial cells with

consequences for neuronal activity. In addition, in the supraoptic nucleus, glia

surrounding the synaptic area appear to control glutamate uptake and activation of presynaptic glutamate receptors, having a role in neurotransmitter release

(Piet et al., 2004). The role of glia in modulation of neuronal glutamate uptake may also have an important role in neuronal excitotoxicity, as will be discussed later.

Glial Role in Excitotoxicity

Glutamate transporters are abundant on astrocytes and can regulate the

concentration of glutamate present in the synaptic cleft (Tilleux and Hermans,

2007). However, under pathological conditions, impaired glutamate uptake by

astrocytes can lead to overactivation of post-synaptic glutamate receptors

32 leading to uncontrolled Ca2+ influx and eventually neuronal death (Tilleux and

Hermans 2007). Additionally, Zhao et al. (2004) has shown that microglia can

influence the efficiency of glial glutamate uptake. Astrocytes, which are highly sensitive to changes in the local environment, can have glutamate uptake affected by neighboring cells or diffusible factors (Tilleux and Hermans 2007).

This is obviously a factor in the process of neurodegeneration in ALS

(amyotrophic lateral sclerosis). For example, Rao and Weiss (2004) showed that

ROS (reactive oxygen species) produced in neurons affected the surrounding

astrocytes causing deficiencies in their glutamate uptake causing an increase in

stress to neurons leading to a viscous cycle ending in neuronal death. Van

Damme et al (2007) showed more specifically that the SOD1 mutation, a

common mutation in familial ALS, in astrocytes caused a reduction in the GluR2

receptor subunit in motor neurons making them more susceptible to Ca2+ influx. It

has been demonstrated that ALS patients have elevated glutamate levels in their

cerebrospinal fluid in addition to a severe reduction of GLT1, an astrocytic

glutamate transporter (Barbeito et al., 2004). Many ALS patients have been

shown to have gliosis surrounding the motor neurons (Barbetio et al. 2004).

Inflammatory changes in reactive astrocytes can modulate the release of

glutamate, and the use of COX-2 inhibitors to reduce inflammation has been

shown to delay onset of disease (Drachman et al., 2002), indicating a role for

reactive astrocytes in the disease mechanism. Excitotoxicity has been indicated

in the pathology of several other neurodegenerative disorders; however, it is not the only pathway for cell death that involves Ca2+.

33

Mitochondria and Apoptosis

Mitochondria play a role in both the activation and prevention of apoptosis.

Additionally, they may also be involved in determining if cell death will occur via

apoptosis or necrosis. Apoptosis is characterized by maintenance of ATP levels,

as it is required for activation of some caspases and DNA fragmentation.

Therefore, if there is acute disruption of mitochondrial ATP production, necrosis

will occur; however, if there is a chronic decrease is ATP production, as is

common in neurodegenerative disorders and aging, apoptosis is the more

common path for cell death (Haeberlein 2004). There are several proteins

involved in the intrinsic apoptotic pathway. The most well known family of

proteins associated with mitochondria and apoptosis are the Bcl family of

proteins, which provide both pro-apoptotic and anti-apoptotic signals.

The Bcl proteins are separated into subgroups based on their structure:

the Bcl-2 like survival factors, the Bax like death factors, and the BH-3 only death

proteins (Borner, 2003). Bcl-2 and Bcl-xL are the most important anti-apoptotic

proteins (Lawen, 2003). It has been shown that these proteins do not increase

proliferation, but actually prevent apoptosis by delaying the cells entry into the s

phase of the cell cycle (Vaux et al., 1988). Additionally, Bcl-2 has been shown to

decrease endoplasmic reticulum Ca2+ levels (Vanden Berghe et al., 2003) and

Bcl-xL has been identified as a requirement for neuronal survival (Motoyama in

(Cory and Adams, 2002). The two other subgroups of Bcl proteins are Bax and

BH-3 only proteins, and they are both pro-apoptotic protein groups. These

34 proteins are thought to act by binding to the pro-survival proteins and rendering

them inactive (Cory and Adams 2002). Bid, a BH-3 only protein, is a protein that

promotes cell death by activating Bax and Bak, which are thought to provoke or

contribute to the permeabilization of the mitochondrial membrane (Cory and

Adams 2002).

Once the mitochondrial membrane is permeabilized other, pro-apoptotic

proteins can be released into the cytosol where they can initiate caspases and promote more downstream pathways of apoptosis. Among these proteins are: cytochrome c, Diablo/Smac, Omi/HtrA2, AIF, and endonuclease G (Haeberlein

2004, Hajnozky 2003). Cytochrome C release is important in neuronal apoptosis activation from a variety of cellular stressors. Additionally, is it associated with

Fig. 1 Mitochondria and Apoptosis

Fig. 1 Model of part of the intrinsic apoptosis pathway. This depicts

some of the proteins associated with mitochondrial apoptosis pathways.

35 striatal damage in Huntington’s disease (Kiechle et al., 2002). Once cytochrome

c is released it interacts with apaf-1 and caspase-9 to create the apoptosome,

which can then activate caspase-3, an effector caspase (Cory and Adams 2002).

Endonuclease G, once released, can translocate to the nucleus and initiate caspase-independent DNA fragmentation. AIF (apoptosis inducing factor) is also translocated to the nucleus and initiates caspase-independent nuclear damage.

AIF has been shown to be induced after various neurotoxic stimuli and models of traumatic brain injury (Haeberlein 2004). The release of these proteins may work in concert to activate apoptosis (Haeberlein 2004). Smac/Diablo and Omi/HtrA2 are believed to bind to inhibitors of apoptosis (IAP’s) thus promoting caspase auto-activation (Haeberlein 2004 and Cory and Adams 2002).

Mitochondria and Calcium Signaling and the Synapse

Mitochondria play an important role in regulating Ca2+ signaling in

neurons. In a study by Wang et al (2003) disrupting the distribution of

mitochondria in neurons caused an increase in cytosolic Ca2+ and a subsequent decrease in mitochondrial Ca2+ levels, thus, indicating that mitochondria respond to Ca2+ microdomains in neurons and can shape specific neuronal Ca2+

signaling. The involvement of mitochondria in neurotransmission was initially

suggested by Parducz and Joo (1976), when they used electron microscopy to

show mitochondria from “stimulated” nerve endings had an increase in

accumulated Ca2+ compared to mitochondria from “unstimulated” nerve endings.

36 Giovannucci et al. (1999) provided support for the idea that mitochondria can modulate neurosecretion when they showed that poisoning mitochondria enhanced the secretory response in chromaffin cells. Additionally, a genetic mutation discovered in the nerve terminals of Milton Drosophilia, showed impaired synaptic transmission. Stowers et al. (2002) demonstrated that this mutation was responsible for the disruption of mitochondrial translocation to the synapse. Furthermore, the Bcl-2 family protein, Bcl-xL, a mitochondrial associated protein, has been demonstrated to modulate synaptic transmission

(Jonas et al., 2003).

The Role of Mitochondria in LTP/LTD

While I have previously discussed that L-type VGCC have a role in LTP and LTD, mitochondria also have demonstrated a role in LTP/LTD and synaptic plasticity. For example, Calabresi et al. (2001) used inhibitors of mitochondrial complex II to produce LTP of NMDA mediated synapse excitation. Additionally,

Weeber et al. (2002) and Levy et al. (2003) showed that inhibiting mitochondrial porins (ion conducting pores on the mitochondrial membrane) and mitochondrial membrane potential caused deficits in long and short term synaptic plasticity.

Specifically, deficits in fear conditioning and spatial learning tasks were observed. Also, Gincel et al. (2000) demonstrated that glutamate could modulate function of VDAC (a mitochondrial porin protein).

37

Role of Dysfunctional Mitochondria in Neurological Disorders

Modulation of mitochondria complexes and porins has a role in diseases

including Huntington’s disease and epilepsy. The mitochondrial toxin 3-

nitroprotionic acid (inhibitor on mitochondrial complex II) has been shown to

evoke seizures and can mimic the pathology of Huntington’s disease in rats and

humans (Calabresi et al., 2001). AMPA receptor antagonists have been shown to

inhibit the occurrence of seizures induced by 3-nitroprotionic acid in rats

(Urbanska et al., 1999), indicating a possible role for Ca2+. Additionally, the mitochondrial porin VDAC has been demonstrated to be altered after seizure

(Jiang et al., 2007) and mitochondrial dysfunction and ultrastructural damage was observed after kainic induced seizure in rats (Chuang et al., 2004). Currently mitochondria and mitochondrial-associated proteins are being considered as a possible therapeutic target for seizure disorders (Henshall et al., 2000).

38 Chapter IV

What is the mechanism by which CAI (Carboxyamidotriazole) alters Capacitative

Calcium Entry?

Background and Lit Review

In this section, I investigated the mechanism of carboxyamidotriazole’s

(CAI) effect on CCE using digital fluorescence imaging and fluorescent dyes. CAI

is a drug currently in Phase II clinical trials for treatment of various types of

cancer (Hussain et al., 2003). I investigated the effects of CAI on mitochondrial

membrane potential using two different fluorescent indicators of mitochondrial

membrane potential. I have shown that CAI has an effect on mitochondrial

membrane potential in a time and concentration dependent manner. Next, I

examined CAI’s effect on mitochondrial Ca2+ uptake. Again I used fluorescent

dyes to show that CAI causes mitochondria to uptake less Ca2+. Thus, I have shown that CAI’s effect on CCE involves, at least in part, an increase in mitochondrial membrane potential and consequent decrease in mitochondrial

Ca2+ uptake.

Cancer and CCE

In cancer, the uncontrolled growth of cells can be caused by an increase

in proliferation of cells, or a decrease or defect in the ability of cells to undergo

apoptosis. Due to the importance of Ca2+ in both proliferation and apoptosis, Ca2+ signaling pathways have been the target of many treatment possibilities.

39 Two approaches to treatment via Ca2+ pathways have been utilized. One

is termed “prodrug” treatment. In this case a non-specific Ca2+ pump inhibitor is

expressed with a tumor-specific enzyme to target the drug to the cancer

(Monteith et al., 2007). An example of this type of treatment option is the use of

thapsigargin, an inhibitor of the SERCA pump combined with prostate-specific

antigen (PSA) a prostate cancer specific protease (Denmeade and Isaacs,

2005). In this case the thapsigargin is inactive until it reaches the target cancer

cells where it is activated and can potentially affect Ca2+ signaling. The second

approach involves direct targeting of specific isoforms of Ca2+ channels or pumps

that have been associated with a particular cancer. One potential area for this

type of treatment is the TRP channels.

Though the role of TRP channels in CCE is not fully understood, there is a

plethora of evidence that they are involved in Ca2+ deregulation in many types of

cancer. Specifically, the over-expression of TRPV6, and TRPM8 have been demonstrated in cancers including prostate, breast, colon, lung, thyroid, and ovarian carcinomas (Prevarskaya et al., 2007). Additionally, the down regulation of TRPM1 has been indicated as a prognostic marker for melanoma metastasis

(Bodding, 2007). TRPV6 has been indicated as being highly selective for Ca2+

(Luo et al., 2001); however patch-clamp recordings have not been able to confirm these findings (Bodding 2007). Never the less, the TRP family of proteins

remains a possible therapeutic target for many types of cancer.

Mitochondria are another potential target for cancer treatment. This is

because of the role mitochondria play in apoptosis. Though the known

40 involvement of mitochondria in malignancies is rare, all of the anti-apoptotic Bcl-2

family members could potentially have oncogene potential (Cory et al., 2003).

Additionally, some pro-apoptotic proteins have potential as possible therapeutic

targets. Thus, a drug or treatment that could affect both Ca2+ signaling pathways

involved in proliferation and mitochondrial pathways of apoptosis would have a

higher probability of being effective for multiple causes of cancer.

CAI (Carboxyamidotriazole)

CAI (carboxyamidotriazole) is a substituted carboxyamido- with a

halogenated benzophenone tail. It was first discovered by Kohn and Liotta (1990)

and identified as a novel antiproliferative and antimetastasis agent. Kohn and

others then began to examine the mechanism behind CAI’s beneficial effects.

They found that CAI inhibited Ca2+ influx, release and cellular

proliferation (Kohn et al., 1994). Additionally studies by Enfissi et al. (2004)

showed that CAI blocked Ca2+ entry due to release of Ca2+ from internal stores.

In addition to the evidence of CAI’s effect on Ca2+ signaling, Perabo et al. (2004)

demonstrated that CAI induced apoptosis and decreased Bcl-2 expression in

bladder cancer cells. Bcl-2 is an anti-apoptotic member of the Bcl protein family, and is associated with mitochondria. Guo et al (2006) also showed an increase in

apoptosis and down regulation of Bcl-2 in human breast cancer cells.

The induction of apoptosis and alterations in expression of Bcl-2 indicate a

possible role for the mitochondria. As previously mentioned, mitochondria are a

major player in apoptosis, and they have a role in regulating CCE. Clearly this

41 evidence indicates a possible role of mitochondria in the mechanism of action of

CAI. Additionally, evidence form Glitch et al. (2002) showed that mitochondria

need to be in an energetic state or IP3 fails to activate SOC entry. Malli et al.

(2003) and Gilabert and Parekh (2000) also showed that mitochondria have the

ability to regulate Ca2+ influx in CCE. In fact, mitochondria affect both sensitivity

to Ca2+ release from internal stores and Ca2+ dependent inactivation of Ca2+ influx through SOC channels. Therefore, it is possible to manipulate CCE by altering the release of Ca2+ from the internal stores, blocking the CCE channels,

or inhibit the ability of mitochondria to uptake Ca2+, thereby causing the channel to shut down early.

As previously discussed, CCE appears to be involved in cellular

proliferation (Golovina, 1999). This means that manipulating CCE may allow scientists to manipulate cellular proliferation in disease. CAI is a compound currently in phase II clinical trial as an anti-cancer compound. It has been shown

to inhibit cellular proliferation in human cancer patients, with fewer side effects than other anti-cancer compounds (Haverstick et al., 2000). However, the mechanism by which it has an effect is not known. CAI has many effects on Ca2+

regulation, not all involved with CCE, however, NCCE is not thought to be

involved in cellular proliferation, and CCE has been shown to have involvement

in cellular proliferation. It is important to determine what the effects of CAI are on

CCE and what aspect of the pathway CAI affects.

In this chapter I will investigate CAI’s effect on CCE, and what aspect of

CCE CAI may be affected. First, initial work by Mignen shows that CAI does

42 cause a dose dependent decrease in CCE after CAI treatment. He also shows that CAI inhibits [3H] thymidine incorporation in HEK-293 cells indicating a role in proliferation (Mignen et al., 2005). Based on this information, the question remains is CAI having an affect directly on the SOC or is there an affect on the mitochondria, causing the SOC to shut off early (See Model Fig 2).

Fig. 2 Model of CAI’s Possible Effects on CCE

Fig. 2 Model of CAI’s possible effects on CCE. I hypothesize that CAI will have an effect on mitochondria, causing them to be less polarized and take up less calcium. This would cause a build up of calcium and deactivation of the SOC.

I will show that CAI affects mitochondria, causing them to have a less negative membrane potential, as measured with JC-1 and TMRE. Additionally, the mitochondria treated with CAI take up less Ca2+ than control mitochondria. Thus,

CAI inhibits the ability of mitochondria to uptake Ca2+, allowing Ca2+ to build up around the store-operated channels and causing them to shut down prematurely.

43 Results

Effects of CAI on amplitude of CCE

First, to show that CCE had an effect on CCE, Mignen used HEK-293

cells loaded with Fura-2 AM to measure changes in cytosolic Ca2+. To initiate

CCE, Ca2+ stores were depleted with Thapsigargin (1 μM) in an external solution

containing no Ca2+. After the internal stores were depleted, an external solution

with 2 mM Ca2+ was added back to the cells. The change in cytosolic Ca2+ after the addition of external Ca2+ was attributed to the store-operated channels.

Control cells were compared to those treated with CAI for time points between 10

seconds and 5 minutes at doses ranging between 0.1 μM and 10 μM. A dose

dependent effect was observed with the CAI causing a decrease in the amount of

Ca2+ entering the cytosol after application of external Ca2+. After 5 minutes of CAI

treatment, the half-maximal inhibition was observed with 0.5 μM CAI and

maximal inhibition with 2 μM CAI, while after 10 seconds 10 μM CAI reduced

CCE to 40% of control. Concentrations over 10 μM were not used due to

potentially toxic effects (Enfissi 2004).

CAI’s effect on HEK-293 cell proliferation

Additionally, to show that CAI had an effect on cell proliferation in HEK-

293 cells, Mignen assessed cell proliferation by measuring the incorporation of

[3H] thymidine. Cells were incubated for 24 hours in the presence of 10% FCS.

External medium was then supplemented with various concentrations of CAI and

incubated for a further 48 hours. CAI blocked the incorporation of [3H] thymidine

into HEK-293 cells with an IC50 of 1.6 μM whereas vehicle had no effect. This is

44 consistent with what has been previously shown in other cell types, that CAI

blocks proliferation of cells (Mignen et al. 2005).

CAI depolarizes the mitochondrial inner-membrane potential in a time

dependent manner.

As previously described, mitochondria play an integral role in the

mechanism of CCE. By buffering the amount of cytosolic Ca2+ available around

the store-operated channels mitochondria prolong the open state of the channels

and delay the Ca2+ dependent deactivation of the channel. Therefore, I monitored

CAI’s effect on mitochondrial health and their ability to uptake Ca2+. To determine the time dependent effects of CAI, I used a multi-well fluorescence plate reader in combination with the mitochondrial selective ratiometric dye JC-1. A description of the dye can be found in chapter two. The ratio of the red to green fluorescence of JC-1 is used to monitor mitochondrial membrane potential without effects due to mitochondrial swelling, which can occur when mitochondria become unhealthy.

45 Fig. 3 Effects of CAI on Mitochondrial Membrane Potential

CAI Dose response for JC-1 A B Effects of CAI on Mitochondrial Membrane Potential 2.5 1.0

0.9 2.0

0.8 Control 1.5 0.7 CAI

FCCP 0.6 F590/F520 F590/F520 1.0

0.5 0.5 0.4 0.0

-1 1 3 C 10 10 10 0 10 20 30 [CAI] (µM) Confocal Images of JC-1 Loaded Cells Time (min)

control CAI FCCP

Fig. 3 A. HEK-293 cells were loaded with JC-1 and treated with various concentrations of CAI. JC-1 fluorescence levels were collected with a multi- well plate reader and a dose response was calculated. B. A time course of the effects of CAI on mitochondrial membrane potential. Cells were loaded with JC-1 and analyzed with the plate reader. C. Visualization of mitochondrial membrane potential with confocal microscopy. Cells were loaded with JC-1 then treated with DMSO (control) CAI (10μM) or FCCP (20μM).

HEK-293 cells were loaded with JC-1 and then treated with 10 μM CAI in suspension. The ratio at 590-520 was calculated using the plate reader at intervals of 5 minutes for up to 35 minutes following application of vehicle (0.1%

46 DMSO), 20 μM FCCP or 10 μM CAI (Fig 3B). CAI significantly altered

mitochondrial membrane potential in a time dependent manner. These results

were confirmed using confocal imaging in HEK-293 cells loaded with JC-1 and

treated for 20 minutes with vehicle, FCCP or CAI. The distinct red punctate

signaling seen in the vehicle treated cells is replaced with a diffuse green staining

in cells treated with FCCP, and this is mimicked in cells treated with CAI (Fig.

3C), consistent with the plate reader results. Additionally, the plate reader was

used to determine a dose response for CAI’s effect of mitochondrial membrane

potential (Fig 3A). These results are consistent with the dose response for CAI’s

effect on CCE preformed by Mignen.

While the ratio-metric nature of JC-1 provides several qualitative

advantages of this dye, there is some evidence that JC-1 is not as quantitative as other mitochondrial membrane potential dyes. Therefore, to obtain quantitative information of the effects of CAI on mitochondrial membrane potential, I used the dye TMRM. TMRM provides information on mitochondrial membrane potential in two ways. First, for digital fluorescence imaging experiments, a low concentration

(12.5 nM) of TMRM was used to observe the spontaneous changes in mitochondrial membrane potential as well as the possible redistribution of dye

(Fig 4A). In this case, ROI’s were placed on structures identified as mitochondria.

Additionally, these experiments were performed using a high concentration of

47

Fig. 5 Effects of CAI as Measured by TMRM A B A1 A2 B1 B2 C

4500

4000

A3 TMRM Fluorescence 3500 B3 200 300

TMRM Fluorescence 280 TMRM Fluorescence 3000 180 260

240 2500 220 160

200 2000 180 140

160

1500 140 120

0 20 40 60 80 100 120 0 50 100 150 200 0 10 20 30 40 Time (Sec) Time (Sec) Time (min)

Fig. 4 A. Cells loaded with TMRM at 12.5nM and treated with FCCP. Images of loaded cells before (A1) and after (A2) FCCP treatment. A3 Example traces of ROI’s of mitochondria. FCCP was administered at 60 seconds. B. Cells loaded with 1μM TMRM. At this concentration dye quenches in the mitochondria and as they become depolarized the dye leaks out and the cytosolic signal increases. Images before (B1) and after (B2) FCCP treatment. B3 Example trace of cytosolic signal. FCCP was administered at 125 seconds. C. Plate reader experiement in which cells loaded with 1μM TMRM and treated with CAI at time 10 min. Open circles are CAI treated.

TMRM. In this case the dye is taken up into the mitochondria in larger quantities where it then quenches (gives off less fluorescence due to interference). As the mitochondria loose their negative membrane potential, the dye leaks out and the cytosol gives off a higher fluorescence (Fig 4B). Both of these methods indicated that CAI altered mitochondrial membrane potential significantly different from control, but not significantly different from the positive control, 20 μM FCCP. I performed the plate reader experiments (similar to those previously described 48 with JC-1) using a high concentration (1µM) of TMRM. These results also

confirmed the information obtained with the JC-1, demonstrating that CAI caused

a significant increase in the depolarization of mitochondria (Fig 4C).

CAI inhibits mitochondrial calcium import.

Because of the observed effects of CAI on mitochondrial membrane

potential, and the fact that mitochondrial Ca2+ uptake depends on the

mitochondrial membrane potential, I examined the effects of CAI on

mitochondrial Ca2+ influx. To determine if the effects of CAI on mitochondrial

membrane potential were in fact reducing the ability of mitochondria to uptake

Ca2+, HEK-293 cells were permeabilized with digitonin. This was done to bypass any possible effects of CAI directly on the store-operated channels or any other

aspect of Ca2+ signaling. Cells were loaded with a mixture of Rhod dyes, (Rhod-2

AM, Rhod-FF AM, and Rhod-5N AM) to be able to measure Ca2+ changes with a

greater dynamic range. To negate the possible effects of CAI on other aspects of

CCE or Ca2+ signaling, we permeabilized the cells with 10 μM digitonin for 3

minutes at 37ºC in a intracellular solution containing no added Ca2+ (Bruce 2004)

(See Fig 5).

49 Fig. 5 Model of Digitonin Permeabilized HEK-293 Cells

e d

c

b

300 Ca

20% a δF

50 s

abc d e

Fig. 5 Example of cells loaded with Rhod dyes and permeabilized with digitonin. After permeabilization, an ISS with 300μM calcium was applied. Trace represents an ROI (region of interest) placed punctuate fluorescent area thought to be a mitochondria or group of mitochondria.

Under these conditions the plasma membrane was permeablized, while

the mitochondrial membrane remained intact, as determined by JC-1.

Additionally, the permeabilization reduced the background cytosolic Rhod signal,

indicating that the plasma membrane was in fact permeabilized (Fig. 5a). (Cells loaded with Rhod had a strong fluorescence signal in the mitochondria and a minimal fluorescence signal in the cytosol. Cells that continued to exhibit cytosolic fluorescence after digitonin treatment were not included in the data.)

50 ROI’s were placed on fluorescent punctate within permeabilized cells.

Identification of these puncta as mitochondria was confirmed by co-labeling with

MitoTracker Green-FM, a dye that selectively loads into mitochondria.

MitoTracker dyes are not Ca2+ sensitive and the fluorescence is not altered by

Ca2+ changes. An added advantage of this approach was that Rhod signal could be normalized to the MitoTracker signal. The validity of the technique was confirmed using FCCP and observing a consequent reduction in fluorescence.

After permeabilization was established, cells were exposed to intracellular solutions containing a set amount of Ca2+ ranging from 3 μM to 3 mM. The change in Rhod fluorescence was monitored and maximal fluorescence change was used to determine significance.

For multi-well plate reader experiments, cells were co-loaded with Mito-

Tracker Green and Rhod dyes and the ratio of 590/520 nM fluorescence was

determined (Fig 6A). I chose to focus on 30, 100 and 300 μM Ca2+ due to the

established EC50 for control cells. Control cells showed an EC50 of ~190 μM for

2+ mitochondrial Ca uptake. Cells pretreated with 10 μM CAI showed an EC50 of

~180 μM Ca2+ (Fig 6C). Though this change is small, it indicates an attenuation in

the dose response curve for mitochondrial Ca2+ influx in cells treated with CAI.

Next, I returned to the digital fluorescence imaging experiments to observe the kinetics of mitochondria in cells treated with CAI. I found a significant effect of

CAI at 30 and 300 μM Ca2+, but not 100 μM Ca2+ in maximal fluorescence

change. However, as can be seen in sample traces, I also observed two different

51 Fig. 6 Effect of CAI on Mitochondrial Calcium Uptake

2+ 2+ 2+ 2+ A Effects of CAI on [Ca ]m B Change in [Ca ]m in Response to Applied Ca [Ca ]o 1.00 types of kinetic effects of CAI. In some cases, there is3000 a decrease in fluorescence

0.95 after initial Ca2+ uptake. This decrease in fluorescence may indicate that Control 1000 0.90 CAI mitochondria are undergoing activation of the mitochondrial300 transition pore. The F590/F520

100 mitochondrial0.85 transition pore is a large20 % pore that is thought to open30 in response to dF 25 s 3 0 5 10 15 20 25 30 35 Time (min) Time (min) Relationship of [Ca2+] to Change in [Ca2+] Effect of CAI Treatment on Ca2+ Response C o m D 140

120 Control CAI

100 CAI 10 % 80 dF 200 s 30 60

40 20

% Change in Fluorescence 20 10 0

0 1 2 3 4 % Change in Fluorescence 0 10 10 10 10 10 Control CAI [Ca2+] (µM) m

Fig. 6 A Cells were co-loaded with Rhod dyes and MitoTracker Green and permeabilized (as described in Methods). Analysis of fluorescence was preformed by a multi well plate reader. Cells treated with CAI showed a decrease in mitochondrial calcium uptake after treatment. Cells were treated with CAI and calcium at time 0. B Sample traces of mitochondrial calcium uptake for treatment with ISS with indicated amount of calcium. C Dose response for mitochondria after treatment with indicated amount of calcium and difference for CAI treated cells. D. Statistical analysis of difference in calcium uptake by mitochondria treated with CAI and sample traces. Two different type of responses were seen in cells treated with CAI, the last trace suggest some cells may be undergoing activation of the permeability transition pore. * indicated p<0.001.

52 mitochondrial apoptosis activation. Thus, if the pore is opening in response to

CAI treatment, one might observe leaking of the dye and a decrease in

fluorescence. More studies are needed to confirm this possible observation.

Discussion

CCE has been shown to be involved in cell proliferation in a number of cell

types. CAI, a compound in phase II clinical trial as an anti-cancer drug, has been

shown to decrease cell proliferation. Because this drug has been shown to be

useful in several types of cancer with fewer side effects than other drugs, it is

important to understand the mechanism of the drugs action. Because CCE is a

complicated pathway that involves several important and diverse areas,

unraveling the mechanism of CAI is complicated.

Mignen et al (2005) has shown that CAI had an affect on cell proliferation in HEK-293 cells causing a decrease in proliferation. Additionally, I have shown

that CAI had a negative effect on mitochondrial membrane potential which

correlated to a decrease in mitochondrial Ca2+ uptake compared to control. This

evidence indicates that CAI has an effect on mitochondria causing them to

become depolarized. This depolarization decreases the amount of Ca2+ the mitochondria can buffer around the store-operated channels causing them to deactivate early, causing a reduction in CCE. This reduction in CCE is thought to be responsible for the decrease in cell proliferation that Mignen and others have shown.

53 I have clearly shown that CAI has an effect on mitochondria causing a

significant effect on membrane potential and Ca2+ uptake. This effect is

somewhat unexpected given that CAI has few side effects and mitochondria are

so ubiquitous. This raises the question that there might be a difference between

mitochondria of cancer cells, and mitochondria of healthy cells. Are mitochondria

in cancer cells more susceptible to the effects of CAI? Could this make

mitochondria a potential target for cancer treatment? Additionally, the effects on

mitochondria membrane potential indicate that ATP levels could be disrupted

after CAI treatment. ATP levels are important in determining cell death pathways

(Haeberlein 2004). If ATP levels are affected, do they play a role in CAI’s ability

to decrease proliferation? Or are cell death pathways activated? Percabo and

Guo (2004) demonstrated that CAI caused an increase in apoptosis and

alterations in mitochondrial associated apoptosis proteins. May CAI’s effect on

mitochondria activate apoptosis in addition to disruption of CCE? Obviously more

studies are needed to assess these questions.

Lastly, due to the possible effect of CAI on ATP it is also important to

determine in the decrease in proliferation is truly a decrease in the number of cells proliferating or an effect of cell death via apoptosis or necrosis. As previously discussed, others have shown that CAI can increase apoptosis and

decrease mitochondrial associated anti-apoptotic cells. This indicates that it is

possible that mitochondria may play an additional role in the mechanism of CAI

by inducing apoptosis. I have shown that there seems to be two types of kinetic

responses in cells treated with CAI. In the second type of response,

54 mitochondrial fluorescence had a dramatic decrease after an initial increase. This may indicate that mitochondria are undergoing permeability transition, or opening of the mitochondrial permeability transition pore. This event has been indicated in the mitochondrial apoptotic pathway and represents an area for further investigation.

55 Chapter V:

Are There Differential Effects of Amyloid-Beta on Cytosolic Calcium and

Mitochondrial Health on Glial Cells and Nerve Terminals of the Hypothalamic-

Neurohypophysis?

In this section, I examined the differential effects of Alzheimer’s disease

(AD) on the hypothalamo-neurohypophysial axis (HN), an area of the brain not affected by AD. While many neurodegenerative diseases have demonstrated a deregulation of Ca2+ signaling, the mechanism of Ca2+ signaling in neurodegeneration is not fully understood. I have investigated the role of Ca2+ signaling in AD by comparing the effects of the toxic peptide Aβ on the HN to what has been shown in the literature for other brain areas that are affected in

AD. Additionally, I have focused on the effects at the synapse, because the synapse has been shown to be preferentially affected in AD (Selkoe 2002).

Alzheimer’s Disease

Alzheimer’s disease (AD) was first described by Dr. Alzheimer in 1906. It is a fatal disease that is the seventh leading cause of death in America today

(www.alz.org). AD is the most common form of dementia affecting memory, cognitive abilities, social behavior, and sometimes causing paranoia and depression. AD is the most common neurodegenerative disorder; there are two types of AD, early onset occurs before age 65, and is typically caused by a hereditary genetic mutation and traditional AD occurring in adults over 65. The

56 cause of the disease is still unknown, and while there are treatments for many of

the symptoms of AD, there is no cure, and AD is still a terminal disease.

While the cause of AD is still unknown, research has centered around the

two hallmarks of AD, plaques and tangles. Tangles are found within the and composed of a hyperphosphorylated tau protein. Genetic evidence,

however, has identified the protein involved in plaques to be a more likely

candidate for causing, or being involved in the cause of AD. Plaques are seen in

most adults over 40; however, AD patients have an unusually large amount of

plaques, combined with other symptoms (Okamura et al., 2004). The main

constituent of plaques is a 40-42 amino acid peptide called amyloid-beta (Aβ). Aβ

comes from the native protein amyloid precursor protein (APP).

APP Processing

APP is a type 1 membrane bound protein found in brain and other tissues

(Canevari et al., 2004). The physiological function of APP is still unknown

(Kamenetz et al., 2003) though there is evidence for roles in G protein coupled receptor signaling, neurogenesis and regulation of excitatory synapses (Caille et al., 2004). Mouse knockouts of APP appear without noticeable phenotype, however it is thought that APP like proteins (APLP1 and APLP2) may compensate for the loss of APP (Caille et al., 2004). There are two main pathways for APP processing, a non-amyloidgenic pathway, and an amyloidgenic pathway. In the non-amyloidgenic pathway APP is cleaved by α-

secretase within the region of the Aβ peptide. The c-terminal fragment can be

57

Fig. 7 APP

β α γ

α β

γ γ

Fig. 7 Model of APP processing by the α, β, and γ secretases. The red box represents the β-Amyloid peptide. Cleavage by β and γ secretases gives the peptide.

further cleaved by β-secretase releasing a p3 fragment. The role of these protein fragments is not known, but they are not known to be toxic or involved in AD. The amyloidgenic pathway is responsible for releasing Aβ from APP by β-secretase and γ-secretase activity. The γ-secretase can cleave the Aβ at either the 40 or 42 residues (Canevari et al., 2004). The Aβ1-42 peptide is thought to be the more toxic form and is most abundant in plaques. The Aβ peptide is sticky and forms monomers, oligomers and fibrils. While the fibrils comprise the main component

58 of plaques, there is new evidence that oligomers of Aβ may be involved in earlier toxicity in AD (Takahashi et al., 2004).

Recently proteins responsible for secretase activity or modifying secretase activity have been identified. Although little information regarding α-secretase activity is known, the proteins responsible for β-secretase have been identified.

BACE1 and BACE2 are aspartic proteases that have been shown to be responsible for β-secretase activity. While the physiological function of BACE is still unknown, BACE has been shown to be increased in AD. However, BACE knockout mice show a normal phenotype (Canevari et al., 2004). The identity of

γ-secretase is also still unknown; however, it is believed that it is an enzymatic complex, which includes presenilins 1 and 2, nicastrin and possibly other proteins

(Li et al., 2003). Presenilin-1 is an integral membrane protein with 6-8 membrane-spanning regions (Begley et al., 1999). Mutations causing familial AD have been associated with the presenilin proteins as well as with the APP protein itself. Many transgenic mouse models of AD are based on mutations of APP. The most common is the “Swiss” mutation, TG2576, with a mutated APP 670/671

(Takahashi et al., 2004). Other mouse models of AD include multiple mutations on either APP, presenilin proteins, or tau associated mutations or combinations thereof.

Another important aspect of APP processing concerns the clearance of

Aβ. The only protein to be shown to cleave Aβ in vivo models (Li et al., 2003) is

Neprilysin. Neprilysin is a 90-110 kDa plasma membrane glycoprotein protein from the M13 family (Wang et al., 2003). Neprilysin (also known as CALLA10

59 because it was previously identified as Common Acute Lymphoblastic Leukemia

Antigen 10) is present both pre and post synaptically in brain regions vulnerable

to plaque formation in AD (Wang et al., 2003). Neprilysin can also hydrolyze and

inactive several other proteins, including enkephalin, ANP, endothelin and

substance P among others (Shirotani et al., 2001). Neprilysin has also been

shown to be damaged by 4-hydroxylnonenal, a by-product of lipid peroxidation, and has been proposed by Wang (2003) as a possible factor for Aβ accumulation

in AD. This is based on evidence that neprilysin mRNA and protein amount are

decreased in AD patients in areas of high plaque accumulation (hippocampus

and temporal gyrus) compared to controls (Yasojima et al., 2001). Other proteins

have been implicated as possible players in Aβ clearance. Insulysin, for example,

has been shown to cleave Aβ in vitro (Mukherjee et al., 2000) but not in vivo.

Aβ at the Synapse

Aβ has been shown to be toxic to neurons, but the nature of that toxicity

has yet to be elucidated. Evidence indicates that the synapse may be involved in

the earliest events in AD (Selkoe 2002). The synapse is the primary site of Ca2+ dysregulation in AD (Mattson 2003). In a study by Lacor et al. (2007) hippocampal cultures treated with Aβ were found to have a significant decrease in spine density and abnormal spine morphology. Takahashi et al. (2002) also saw abnormal synaptic morphology in AD brain. Stern et al. (2004) showed that

Aβ plaque formation in Tg2576 mice (an AD mouse model) caused significant impairment in synaptic function. Furthermore, Kamenetz et al. (2003) showed

60 that neuronal activity modulates formation and secretion of Aβ in hippocampus

and that Aβ depressed excitatory synaptic transmission.

Previous studies in cortical and hippocampal cultures have determined Aβ

treatment to be toxic to neurons, astrocytes, and synaptosomes. Fu et al. (2006)

used MTT assay to identify a loss of viability in primary cortical neurons after Aβ

treatment. Kelly and Ferreira (2006) and Ekinci et al. (1999) have determined

that Aβ treatment can cause dysfunction of Ca2+ regulation in both cortical and

hippocampal neurons. Aβ treatment has also been shown to affect both

mitochondrial dysfunction and Ca2+ deregulation in hippocampal astrocytes

(Abramov et al., 2003). Mantha et al. (2006) demonstrated that Aβ treatment

caused an increase in resting Ca2+ levels in synaptosomes. In addition,

mitochondrial respiration (Selkoe, 2002) and assays of mitochondrial enzyme

activity (Mantha et al., 2006) have been shown to be altered by Aβ treatment.

Based on this evidence I decided to look at the effects of Aβ on two elements of the synapse, astrocytes and nerve terminals of the posterior pituitary.

Posterior Pituitary as a model system to Study AD

One interesting aspect of AD that has not been investigated is the

differential effects of the disease on different areas of the brain. The

hippocampus appears to be affected early in the disease, while other areas remain unaffected until the late stages in AD. For example, although changes in the hypothalamic-pituitary-adrenal axis (HPA) have been reported in some late stage AD patients, Petrie et al (1999) and Hatzinger et al. (2007) both indicate

61 these changes resulted from upstream-mediated stressors, primarily from the hippocampus. Additionally, Touma et al. (2004) concluded that hypersecretion of glucocorticoid from the pituitary in TgCRND8 mice (an APP transgenic model of

AD) was also caused by changes in hippocampus. Consequently, the HN, unlike the hippocampus, is not a major area of neurodegeneration in AD. However, why the HN does not seem to be affected has not been investigated. In this work, I examined the possible role of Aβ in the differential effects of AD at the HN utilizing the unique architecture of the posterior pituitary.

The posterior pituitary is largely comprised of astrocytic glial cells called

pituicytes, neurosecretory terminals of neurons from the supraoptic nucleus, and

endothelial cells of the vasculature system. These nerve terminals release either

oxytocin (OT) or vasopressin (VP) into the bloodstream. OT is involved in uterine

contractions during labor and milk ejection during lactation; it also has

implications in aggression (Huber et al., 2005). VP is involved in regulation of

blood volume during dehydration and hemorrhagic shock; and was also recently

implicated in aggression (Huber et al., 2005). Correspondingly, OT and VP have

effects on tissues in mammilary glands, uterus, kidneys and vascular beds

(Hatton, 2002) and the amygdala (Huber et al., 2005).

The astrocytes of the posterior pituitary have roles in both maintenance of

nerve endings under basal conditions and maintenance of active states (Hatton,

2002). To sustain normal conditions, the pituicytes are flat and engulf the nerve

terminals separating them from the vasculature. The astrocytes may also release

large amounts of onto the neurons, which inhibits the release of VP via

62 glycine receptors on the neurons (Rosso et al., 2004). During lactation or

dehydration, when the neurons are active, the glial cells retract from the basal

lamina and the nerve terminals, exposing the terminals to the vasculature

system. Subsequently, terminals become enlarged as they release their peptide

into the blood stream (Hatton, 2002). The release of VP can also cause a rise in

Ca2+ in astrocytes from internal Ca2+ stores that cause the release of glutamate,

which can prolong the activity of the terminals (Hatton, 2002).

To facilitate these studies, we used the neural lobe of the pituitary as a

tractable simplified model of hypothalamic tissue. The neural lobe contains

primarily, astrocytic glial cells called pituicytes, nerve terminals of supraoptic

magnocellular neurosecretory neurons that have their cell bodies in the

hypothalamus, and endothelial cells of the micro-vasculature. Therefore utilizing cultures from the posterior pituitary allows one to look at neuronal endings and

glia in a more simplified environment (Fig. 8A). In the current study I exploited the

unique architecture of the posterior pituitary to prepare pure cultures of isolated

central nerve terminals or their associated astrocytes to clarify the earliest Aβ-

mediated modifications of neuronal and glial health. In an effort to begin to

address the underlying causes of the HN resistance to early manifestation of AD,

I investigated differences, primarily focusing on Aβ-mediated mitochondrial

dysfunction and Ca2+ deregulation, in the HN in contrast to that reported by

others from cortex and hippocampus. I hypothesize that Aβ will preferentially

effect the glial cells of the HN sparing the nerve endings from the toxic effects of

Aβ (Fig. 8B).

63 Fig. 8 Possible Effects of Aβ on Cells of the HPA

1 Norm a l re d ox 2 Ph a se I 3 Ph a se II microenvironment Compromised Neuro- redox balance degeneration AB AB THI O LS Ca 2+

NO, RNS, ROS APP, neurotoxic peptides?

Fig. 8 Model of possible effects of Aβ on glia and nerve endings of the posterior pituitary. 1. Under normal conditions the glial cells are supportive of the synapse. 2. I hypothesize that during Aβ exposure the pituicytes will be preferentially affected. 3. Is there neurodegeneration of this model of a HPA synapse? I hypothesize that Aβ will have a differential effect on the cells of the posterior pituitary compared to cells from other brain areas, and that this differential effect may explain the differential effect of AD on this brain area.

In this study, I began by investigating the effects of Aβ treatment on mitochondrial membrane potential at several time points. I found a differential effect of Aβ on pituicytes and nerve terminals. Next, I examined the effect of Aβ on cytosolic Ca2+ levels. Consistent with the mitochondrial membrane potential results, I found a differential effect of Aβ on nerve endings and pituicytes.

However, when measuring the change in cytosolic Ca2+ after stimulus, I found both nerve endings and pituicytes displayed a higher influx of Ca2+ in response to stimulus after exposure to Aβ for 24 hours. Lastly, I observed the distribution of

Aβ using a fluorescently tagged Aβ peptide. Again, I found a differential effect of

Aβ on nerve endings and pituicytes. While the nerve endings showed no signs of taking up the peptide, the pituicytes exhibited uptake of the peptide after 3 hours.

64 I hypothesized that these differential effects of Aβ may play a role in the lack of

toxicity observed in the HN in AD.

Results

Mitochondrial Membrane Potential

As previously discussed, Aβ has been shown to affect mitochondrial

membrane potential in both isolated mitochondria and mitochondria of intact

neurons and glia. Therefore, I explored Aβ’s effect on mitochondrial membrane

potential in neurosecretosomes and pituicytes following 6, 12, 24 or 48 hours of

continuous Aβ treatment. To measure mitochondrial membrane potential I used

the potentiometric dye called JC-1, a positively charged dye that accumulates in

the mitochondria. For an explanation of how the dye works see methods section.

Pituicytes cultures were prepared by removing the posterior pituitary, placing it in

a clot on a clean coverslip and grown in complete DMEM for 14 days (see

methods). At this time the pituicytes grew out from the explants and were

identified by immunocytochemistry for GFAP (glial fibulary acidic protein as

shown in Fig 9A). Pituicytes grown 14 days in culture, were then treated with

25μM Aβ25-35 in DMEM for the indicated time point. Pituicytes were loaded with

JC-1 2μg/μL for 30 min in the dark at room temperature. As shown in Fig 9B, control cells demonstrated pituicytes with punctate red signal and minimal green background signal. However, after 24 hours Aβ treatment, pituicytes

65 Fig. 9 Effects of Aβ on Mitochondrial Membrane Potential of Pituicytes

A B E 125

100

75 *

C D 50 * * % of Control 25 *

0 Con 6hrs 12hrs 24hrs 48hrs FCCP

Fig. 9 Effects of Aß on mitochondria membrane potential in pituicytes. Differential interference contrast image of live pituicytes (A) and Immunocytochemistry of cultures for GFAP (B). Representative pictures of control (C) Aß 25 µM treated for 24 hrs (D) and FCCP (E). Membrane potentials as measured by JC-1 for pituicytes compared to control values. FCCP was used as a positive control. Significance indicated by * p< 0.001.

demonstrated much less punctate red signal and significantly higher green signal as shown in Fig 9C, thus indicating that the mitochondrial membrane potential was less depolarized after 24 hour treatment. The protonophore FCCP, a mitochondrial uncoupler, also caused a decrease in red signal and increase in diffuse green signal as can be seen in Fig 9D. The time course of Aβ treatment is illustrated in Fig 9E. As shown, Aβ treatment caused a significant biphasic decrease in JC-1 ratio values in a time dependent manner compared to control, indicating a significant decrease in mitochondrial membrane potential after 12 hours (0.2214 ± 0.01) 24 hours (0.35 ± 0.03) and 48 hours (0.64 ± 0.01) compared to control (p<0.0001). The levels at 24 hours were comparable to

66 those induced by FCCP (p< 0.001, 0.49 ± 0.05). The 6 hour treatment time point

was not significantly different from control (p= 0.08, 0.96 ± 0.01).

Next, to determine the effects of Aβ on mitochondrial membrane potential

in nerve endings, I utilized the same JC-1 dye in the neurosecretosome

preparation. In this case Aβ treatment was examined after 24 and 48 hours. As

shown in Fig. 10A, control nerve endings, in culture for 24 hours exhibited

punctate red signaling, indicative of healthy mitochondria. However, as shown in

Fig. 10C and 10D, nerve endings treated with 25 μM Aβ for 24 or 48 hours

respectively, also demonstrated punctate red signaling, contrary to what was

seen in the pituicytes. Fig 10D is again an example of FCCP treated nerve endings, illustrating uncoupled mitochondria. The comparison of JC-1 ratios in control versus Aβ treated nerve endings for 24 and 48 hours is demonstrated in

Fig 10E. As shown in Fig 10E, control nerve endings at 24 hours showed similar fluorescence intensity as treated nerve endings (1.11 ± 0.08, p= 0.37) as did control and treated at 48 hours (0.91 ± 0.07 p= 0.45). Similar to pituicytes, as shown in Figure 10D, FCCP treatment caused a rapid and significant decrease in mitochondrial membrane potential (0.26 ± 0.01, p< 0.001). Because of the semi- quantitative nature of JC-1 (as explained in the previous chapter), these experiments were repeated using TMRM (data not shown). TMRM is a fluorescent indicator of mitochondrial membrane potential; similar to JC-1, but it is a more quantitative dye. Results with this dye indicated a significant increase

67 Fig. 10 Effects of Aβ on Mitochondrial Membrane Potential in Nerve Endings Control Treated 125 A B FCCP 100

75

CD 50

% of% Control * 25

0

Fig. 10 Effects of Aß on mitochondrial membrane potential in nerve endings. Representative images of control (B) and Aß 25 µM treated nerve endings at 24 hrs (C) and 48 hrs (D). E. Relative measure measurements of membrane potentials by JC-1 for treated nerve endings compared to their time matched controls. FCCP was used as a positive control. * indicates p< 0.001.

in fluorescent signal in treated nerve endings. This indicates that the Aβ treated mitochondria retained a more negative membrane potential, in contrast to what was determined in the Aβ treated pituicytes and what others have shown in cortical synaptosomes (Begley et al., 1999).

Cytosolic Calcium Measurements

Mitochondria are known to buffer physiologically relevant cytosolic Ca2+

levels in the pituitary nerve endings (Giovannucci 1997). However, the ability of

mitochondria to uptake Ca2+ is dependent on the depolarization state of the

mitochondrial membrane. Therefore, I hypothesized that the decrease in

mitochondrial membrane potential in pituicytes following 24 hours Aβ treatment

would deregulate cytosolic Ca2+ signaling. Again, we treated pituicytes with 25μM

68 Aβ for 24 hours and found that treated pituicytes had a significant increase in

resting Ca2+ levels (686.9 ± 6 ratio units) compared to control pituicytes (658.6 ±

8 ratio units, p= 0.004 ) as measured with Fura-2 AM. To investigate the effects of Aβ on evoked rises in Ca2+ in pituicytes, the physiological agonist vasopressin

(AVP) was used at 100nM and 1μM. AVP was applied for 50 seconds to activate cytosolic Ca2+ increase in pituicytes; two concentrations of agonist were used to

activate varying levels of Ca2+. The maximal increase in fluorescence from baseline (resting fluorescence) after stimulus was used to estimate the increase

Ca2+. Surprisingly, none of the treatments showed a significant difference from control in the amplitude of Ca2+ influx. However, others have reported

heterogeneity of signal for pituicyte clusters versus isolated pituicytes (Venance

et al., 1997), (Troadec et al., 1999). I also observed two distinct populations that

differed in responses to stimulus, although in the current study we found no

consistent correlation of signal amplitude and cell confluency.

To distinguish between the two populations, the data was binned in

increments of 20 ratio units change (Ca2+ increase from baseline after stimulus),

plotted as histograms and Gaussian peak detection algorithm functions were

applied to assess the distribution of responses to AVP stimulus. This analysis

demonstrated that there were two populations of responses to AVP treatment.

These populations were identified as “low responders” and “high responders”.

69 Fig. 11 Effect of Aβ on Cytosolic Calcium in Pituicytes

A a1 C Control Treated 2000 100 μM AVP

1000

B b1 Fura Ratio

0 100 μM Low 100 μM High 1 μM Low 1 μM High

1 μM AVP

Fig. 11 Sample traces of cytosolic calcium signal in pituicytes loaded with Fura- 2 after indicated stimulus for 50 seconds (A-B). (All resting fluorescence was normalized to 0.) Bar graph of low responders and high responders for AVP stimulated pituicytes (C).

Cells with maximal changes fewer than 850 units were termed “low responders”,

and cells with maximal changes above 850 units were termed “high responders”.

We chose this cut off point based on the distribution of responses indicated by

Gaussian fits. Examples of these groups are shown in sample traces in Fig 11 A-

B. We compared “low responders” versus “high responders” for each AVP

stimulation condition using student t-tests. It was determined that three of the

four conditions showed a significant difference in maximal Ca2+ increase

following AVP stimulus. Both high and low responders were significantly

increased in treated pituicytes compared to control after 1 μM AVP stimulus (low

459.6 ± 18 ratio units, high 1246 ± 27 ratio units). In contrast, following 100 nM

70 AVP stimulus, only the high responders showed a significant difference in treated

versus control (high 100 μM 1214 ± 60 ratio units, low 1 373.3 ± 29 ratio units,

high 1 987.1 ± 52 ratio units, low 1630 ± 61ratio units; p<0.01). (Control 405.2 ±

23 ratio units, Treated 405.2 ± 22 ratio units p= 0.9) Data is shown in Fig. 11C.

These same experiments were repeated in the nerve endings. I

hypothesized that because there was no change in mitochondrial membrane potential, there would be no change in cytosolic Ca2+ handling. Surprisingly, there was an observed significant difference in the resting levels of Ca2+. Treated

nerve endings actually had lower resting Ca2+ consistent with the mitochondrial

membrane potential determined by TMRM (Control 718.4 ± 25 ratio units,

Treated 617.7 ± 25 ratio units, p= 0.01). Next, I wanted to assess Aβ’s effect on

evoked Ca2+ rise in the nerve endings. To accomplish this I used a PSS solution supplemented with elevated potassium (50mM) to activate voltage gated Ca2+ channels. High potassium solution was applied for 50 seconds to induce a rise in cytosolic Ca2+. In nerve endings treated with Aβ a significantly greater rise in

Ca2+ was induced compared to control (Control 800.6 ± 68 ratio units, Treated

1020 ± 56 ratio units, p= 0.01). Example traces are shown in Fig. 12A; data is

shown in Fig. 12B. Thus, in contrast to pituicytes the resting Ca2+ was decreased

in nerve endings, and the amplitude of the Ca2+ rise after stimulus was

significantly increased in treated nerve endings. These results were similar to

that previously demonstrated for cortical synaptosomes (Begley et al., 1999),

(MacManus et al., 2000).

71

Fig. 12 Effects of Aβ on Cytosolic Calcium in Nerve Endings

A 1250 Control B * Treated 1000

750 100 * 25 500

250 Fura Ratio 0 Resting Change

Fig. 12 Sample traces of calcium influx in nerve endings loaded with Fura-2 after high K PSS for 50 sec (A). (Resting fluorescence normalized to 0.) Bar graph of data for resting fluorescence and maximal change in fluorescence after stimulus with high K in control and treated nerve endings (B).

Fluorescent Aβ Uptake

As a result of these differential effects of Aβ, I wanted to investigate the pituicytes and nerve endings handling of Aβ. Fluorescently tagged Aβ peptide

(Fl-Aβ) (AnaSpec) was used to treat the cells. Fl-Aβ was applied to the culture media, just as the cells had been treated with Aβ in the previous experiments.

Confocal microscopy was preformed in the continuous presence of the Fl-Aβ.

Internalization of the peptide was observed in pituicytes over a 12 hour period.

The fluorescent signal was maintained for at least 48 hours. The time course of

Fl-Aβ internalization was obtained using a Luden Chamber and 4-D confocal imaging. We observed a rise in fluorescent signal

72 Fig. 13 Uptake of Fluorescent Aβ A B

15 min 45 min CD

2 hrs 3 hrs

Fig. 13 Uptake of fluorescent AB in pituicytes loaded with TMRM (red). Projected confocal images at indicated time after Fl-AB treatment (green) (A-D). Image stacks were collected every 15 minutes for 12 hours.

within 1 hour (Fig. 13A) and over a period of 3 hours (Fig 13D) after which the signal stabilized. To visualize the cytosol, pituicytes were loaded with a high concentration of TMRM. (For an explanation of high concentration loading of

TMRM versus low concentration loading see Methods sections.) Images were 73 taken every 15 minutes. There was no significant effect of Aβ seen with TMRM,

however photobleaching was observed within the first hour of analysis making

the results difficult to interpret (Data not shown). Additionally, I attempted to

locate subcellular compartmentalization of the peptide. To accomplish this I used

fluorescent markers for mitochondria and lysosomes; however, no co-localization

was found after 48 hours.

Subsequently, I applied Fl-Aβ peptide to the nerve endings, again in the

same manner as previously described. In contrast to what was observed in the

pituicytes, the nerve endings showed no indication of internalization of the peptide at time point up to 48 hours. Images were collected using wide field fluorescence imaging (Fig 14). This differential uptake may explain the differential effects of Aβ on mitochondrial membrane potential and cytosolic Ca2+.

Fig. 14 Nerve Endings Treated with Fluorescent Aβ

Fig. 14 Bright field image (A) and fluorescence image of NE treated with Fl-AB. No uptake was found in NE.

74 Discussion

In these experiments, I have shown that Aβ treatment differentially affects

nerve endings and pituicytes. Where as pituicytes showed a time dependent

biphasic decrease in mitochondrial membrane potential, the nerve endings

exhibited no change as measured by JC-1 and a trend to be more polarized as

measured by TMRM. Consistent with these results, Aβ treatment was shown to

alter cytosolic resting Ca2+ levels. While the pituicytes showed an increase in

resting Ca2+ after Aβ treatment, the nerve endings showed a decrease in resting

Ca2+ levels after Aβ treatment. Surprisingly, when the change in cytosolic Ca2+ after stimulus was measured, I observed an elevated Ca2+ increase in both nerve

endings and pituicytes after Aβ treatment. Lastly, using a fluorescently tagged Aβ

peptide I demonstrated that pituicytes could take up the peptide after 3 hours, while nerve endings did not show any indication of uptaking of the peptide up to

48 hours. This evidence suggests two important findings: first the differential effect of Aβ on the pituicytes and the nerve endings of the HN, second the contrasting effects of Aβ on nerve endings of the HN compared to what has been shown in other brain areas.

For example, the effect of Aβ on pituicytes was not unexpected given that

others have shown similar effect. For example, Abramov et al. (2004)

demonstrated that glia cells in co-culture demonstrated a less polarized

mitochondrial membrane potential after Aβ treatment. However, the effect on

nerve terminals was unexpected. Studies with synaptosomes from cortical tissue

have shown mitochondria to be affected by Aβ (Begley et al., 1999). Because

75 synaptosome preparations and the nerve terminal preparation of the HN are similar, it is surprising to see differential results between these two preparations.

One possible explanation for this dissimilarity is that mitochondria of nerve

endings from specific brain areas are more or less susceptible to toxicity by Aβ.

Additionally, the cytosolic Ca2+ results are similarly surprising. While

resting Ca2+ levels are consistent with the mitochondrial membrane potential

data, they again differ from what has been demonstrated in synaptosomes.

(Mungarro-Menchaca et al., 2002). To further demonstrate the role of

mitochondria in the change in cytosolic Ca2+, I measured mitochondrial Ca2+

changes using Rhod dyes. No significant changes were found for nerve endings

treated with Aβ, while pituicytes treated with Aβ showed a significant decrease in

mitochondrial Ca2+ levels (data not shown). This indicated that the effects of Aβ

on cytosolic Ca2+ may not be related to mitochondria, but may be caused by

another effect of Aβ.

This possibility was further demonstrated by the effects of Aβ on the

elevation of cytosolic Ca2+ after stimulation. In both pituicytes and nerve endings,

there is a significantly higher increase in cytosolic Ca2+ in treated cells and nerve

endings compared to control. These results are clearly unexpected and may

indicate that Aβ is affecting the nerve endings in a way that does not directly involve the mitochondria.

A second possible explanation for the differential affects of Aβ is that the

nerve endings of the HN handle Aβ differently than other neurons. To further

explore this possibility, I implemented a fluorescently tagged Aβ peptide to

76 investigate how the peptide may be affecting the pituicytes and nerve endings.

Again, I observed a differential effect of Aβ. While the pituicytes exhibited uptake

of the peptide after 3 hours, which was sustained for at least 96 hours. The nerve endings did not show any indication that there was any uptake of the peptide up

to 48 hours. Additionally, if it is shown that Aβ is taken up by cortical or

hippocampal neurons this may help to explain why the neurons of the HN are not

affect in AD. These observations may be a key to explaining the differential effect

of AD on the HN.

77 Chapter VI:

How Does Allopregnanolone Treatment Affect Calcium Signaling in Cortical

Neurons?

Allopregnanolone, a progesterone metabolite, has been demonstrated to

have an anti-convulsant effect in women (Herzog, 1995). Changes in

allopregnanolone levels during menses may be involved in the etiology of a type

of epilepsy called catamenial epilepsy which will be discussed later in this

section. Recent evidence in NT2-N cells suggests that allopregnanolone may

alter the expression of proteins involved in Ca2+ signaling pathways (Pierson

2005), suggesting that altered Ca2+ homeostasis may be involved. Consequently,

due to the role of mitochondria on seizure activity, and mitochondria’s role in the

shaping of Ca2+ signals involving LTP and LTD, assessing the effects of

allopregnanolone on mitochondrial function is also important. Because

allopregnanolone is a known GABAR modulator and GABAR have been a

primary site for epilepsy treatment, it is also important to investigate the possible role of GABAR in the anti-convulsant effects of allopregnanolone. I hypothesized

that GABA and allopregnanolone activate an efflux of chloride through the

2+ GABAAR initiating a depolarization and Ca influx through VGCC. This excess

cytosolic calcium may lead to the changes in GABAAR subunit composition

demonstrated by Pierson (2005) and may alter seizure susceptibility.

Mitochondria may play a role in this process by altering the Ca2+ signaling. I have

examined the ability of allopregnanolone to cause a Ca2+ influx in primary cortical

neurons. Additionally, I have examined to effects of chronic allopregnanolone

78 treatment on mitochondrial membrane potential and cytosolic calcium levels in

order to tease out the role of Ca2+ and mitochondrial health in catamenial epilepsy.

Catamenial Epilepsy

The word catamenial is derived from the Greek word, “katamenious” meaning monthly; and originally, catamenial epilepsy referred to the cyclical nature of seizures attributed to the cycles of the moon (Foldvary-Schaefer and

Falcone, 2003). Now, catamenial epilepsy refers to a woman with epilepsy whose seizures are in some way influenced by sex hormones secreted during

Fig. 15 Model of Hormone Fluctuations During the Menstral Cycle

Mensus Follicular Phase Luteal Phase

Ovulation

Fig 15 Model of hormone fluctuations during the mentrual cycle. The black line represents estradiol, the blue line represents progesterone, and the red line represents allopregnanolone.

79 the menstrual cycle. Sir Charles Locock first described the relationship in 1857

(Foldvary-Schaefer and Falcone, 2003). The percentage of women whom

experience catamenial epilepsy is thought to be between 10 and 70% (Reddy,

2004). This large range is likely because the basis for determining catamenial

epilepsy is often unclear as some women experience an increase of seizures at

different times during their cycle (Pierson et al., 2005).

The menstrual cycle consists of 3 main stages, with specific changes in

hormone levels (Fig. 15). First is menses, a 1-4 day period when the body rids

itself of the unfertilized ovum. At this stage, estradiol (a form of estrogen) is low, progesterol (a form of progesterone) is low and allopregnanolone (a progesterone metabolite) is also low. Next is the follicular phase, an 8-10 day period when the ovum is becoming mature and preparing to be released from the ovaries. During this period, estradiol levels rise where as progesterol and allopregnanolone levels remain low. At this point there is ovulation and the luteal phase begins. This is typically a 14 day period when estradiol levels decrease, but not to the low levels observed during menses. Conversely, progesterol and allopregnanolone levels increase (Reddy 2004, (Scharfman and MacLusky,

2006).

The three common patterns of catamenial epilepsy described by Herzog

et al. (1997) are classified by changes in seizure frequency during these three

stages of the menstrual cycle. These are: 1. perimenstural, an increase in

seizure frequency during menses, 2. periovulatory, an increase in seizures during

ovulation, and 3. inadequate luteal phase, which occurs in women who

80 experience an inadequate luteal phase (one lasting less then 14 days). In these

women, seizures are common in all phases except the mid-follicular phase. It is

thought that the cyclical changes in hormones, both increases and withdrawals

are responsible for these patterns of catamenial epilepsy (Reddy 2004).

Hormones

Estrogen

The three main types of estrogens are esterone, estradiol and esteriol

(Reddy 2004, Foldvary-Schaefer and Falcone, 2003). As previously mentioned, estradiol increases in the follicular phase and decreases slightly in the luteal phase then decreases more during menses. Estrogens have been shown to be

proconvulsant in both animals and humans, and are thought to act through

NMDA receptors (Reddy 2004). Estradiol was shown to facilitate kindling

seizures in animal models (Reddy 2004). Estrogen has been shown to increase

or exacerbate seizures when given intravenously to woman premenstrually

(Reddy 2004). Chronic exposure in rats to estradiol increased the number and

density of dendritic spines and excitatory synapses in the hippocampus of rats,

likely through activation of the NMDA receptors (Reddy, 2004). Clearly, estrogen

increases may play a role in seizure frequency in women.

Progesterone and Allopregnanolone

The three main types or progesterones are progestins, progestagens, and

progesterone (Scharfman and MacLusky, 2006). Progesterones have been

shown to be anti-convulsant, but the mechanism responsible for this effect is

unknown. One of the issues involved is that progesterone can be converted into

81 dihydroprogesterone, which is further metabolized into allopregnanolone (5α- pregnan-3α-ol-20-one). Allopregnanolone is an endogenous neurosteroid and a potent positive allosteric modulator of the GABAA receptor (GABAAR).

Progesterone has been linked to the etiology of catamenial epilepsy since

Laidlaw (1956) first suggested that seizure exacerbation was seen after rapid decline in progesterone, and natural and synthetic progestrin therapy have been beneficial for some women with catamenial epilepsy (Herzog, 1995), is it unknown if these effects are from progesterone or the metabolite allopregnanolone.

There is evidence that the anti-seizure effects or progesterone are not related to the progesterone receptor. First, the effects are rapid. This is not consistent with the actions of the progesterone receptor which acts by altering gene transcription (Reddy 2004). Additionally, the effects are not blocked by

RU486 a progesterone (Mohammad et al., 1998). Also,

Herzog found that progesterone’s therapeutic activity required to conversion of progesterone to a 5α reduced metabolite, such and allopregnanolone (Herzog and Frye, 2003). Allopregnanolone was also shown to block the seizure facilitating effects of sulfate (Kapur, 2003). Allopregnanolone is converted from progesterone in both peripheral tissue and in the brain

(Corpechot 1993). The physiological concentration during the luteal phase is 2-

4nM (Reddy, 2003) Allopregnanolone is a specific modulator of the GABAAR; it does not act through the progesterone receptor.

82 Neurosteroid Effects on the GABAAR

As previously mentioned, allopregnanolone is a known GABAR

modulator. GABA is the major inhibitory neurotransmitter in the central nervous

system. It binds to GABAAR, GABAB receptors (GABABR) and GABAC receptors

(GABACR) usually leading to an inhibition of cell depolarization. In developing

neurons GABAAR, and GABACR have been shown to cause depolarization. The

GABAAR and GABACR are ionotropic channels; they form a chloride specific

channel (Macdonald and Olsen, 1994). They are inhibitory when chloride goes

into the cell and excitatory when chloride leaves the cell through the channel.

GABABR are metabotropic receptors, acting through a G-protein coupled second messenger or the receptor can be coupled to Ca2+ of potassium channels (Bettler

et al., 2004). GABA dysfunction at both receptors is thought to be involved in disease states, including epilepsy, addiction, depression, and schizophrenia

(Bettler et al. 2004).

Some research into how allopregnanolone and may affect

seizure susceptibility has indicated a role for GABAAR subunit subtype

expression. First, Pierson et al. (2005) used rt-PCR to show that progesterone

treatment caused an increase in α2 subunits of the GABAAR at both 2 and 7 days

of chronic treatment, while γ3 was decreased after 2 days, and α5 and γ3 were

both increased after 7 days. Smith et al. (2007) showed that allopregnanolone caused an increase in α4 and δ subunits. The importance of changes in GABAR subunits is discussed later in this section. Pierson et al. (2005) also showed changes in α4 after both progesterone and estrogen treatments. Both α4 and δ

83 subunits have been shown to alter GABAAR pharmacology and synaptic currents

(Smith et al. 2007). Wafford et al. (1996) showed that withdrawal of progesterone

caused an increase in α4 subunit expression and that this related to a

pharmacological change. Most of these studies, however, were done in cell lines,

and the effects of neurosteroids vary within brain areas, and even receptor

subunit configuration (Belelli and Lambert, 2005). Thus, while there is evidence

that these neurosteroids can have an effect on GABAAR, more information is

needed.

GABAAR

GABAAR are composed of 5 subunits making up a chloride pore. Each subunit has 4 transmembrane domains and different subunits have binding sites

for various agonists, antagonists and modulators. The channel typically has 3

open states, 10 closed sites and a desensitized state (Macdonald and Olsen,

1994; Enz and Cutting, 1999). However, these gating properties vary based on

subunit subtype composition, which will be discussed later. The desensitized

state is important in drug tolerance. There are several agonists and antagonists

as well as positive and negative modulators of the GABAAR. There are specific

binding sites for , , , and anesthetic

steroids. Selective agonists include GABA and , and bicuculline is a

competitive antagonist at this same site. Non-competitive antagonists include

picrotoxin, PTZ and TBPS (Macdonald and Olsen 1994). Benzodiazepines,

barbiturates, and anesthetic steroids are all agonists of the channel with there

84 own binding sites. There are also other modulators of the GABAAR that do not

have known binding sites. Allopregnanolone, for example, is a positive allosteric

modulator of the GABAAR, which will be discussed later in this section (Reddy,

2003).

As previously mentioned, there are several subunits of the GABAAR, and there are different subtypes of the subunits. The different subtypes of subunits give GABAAR different physiological and pharmacological properties. The

different subunits include: α1-6, β1-4, γ1-3, δ, ε, π, and σ (Pierson et al., 2005).

GABAAR are typically composed of 2α subunits, 2β subunits, and 1γ subunit. The subunits differ through development, in various cell types and brain regions, and with environmental exposure. Different combinations of subunit subtypes may be responsible for susceptibility to diseases and drug sensitivity and tolerance. For example, the α1 subtype confers binding of type 1 benzodiazepines, while α2, 3, and 5 confer type 2 binding (Pritchett and Seeburg, 1991). There are many other examples of the specificity of subtypes. Αlpha subunits typically confer benzodiazapine and steroid modulation, while β and γ subunits have only been shown to affect benzodiazepine efficacy.

GABABBR

GABABR induce a slow inhibitory post-synaptic current because, unlike

GABAAR which open a chloride sensitive pore, GABABR are coupled to either G

2+ proteins, Ca channels or potassium channels. GABABR has only two subunits,

GABAB1 and GABAB2 and both are required for a functional channel. In fact,

85 without the GABAB2 subunit the GABAB1 subunit will remain at the endoplasmic

reticulum; GABAB2 is required to traffic the receptor to the plasma membrane

were it can interact with a G-protein or Ca2+ or (Bettler et al.,

2004). GABAB1 has 7 transmembrane domains like other G-protein coupled

receptors. (β-chlorophenyl-GABA) is the only available for

GABABR, but other antagonists include saclofen. Transcription factors like

CREB2 and ATF-4 are also able to interact directly with the GABABR (Bettler,

Kaupmann et al. 2004). There are isoforms of the GABAB1 subunit; the most

abundant are GABAB1 and GABAB2 (Bettler, Kaupmann et al. 2004). There are no

known splice variants of the GABAB2R.

Relationship between GABAR and L-type VGCC

Given the amount of evidence pointing to a role for the GABAR in epilepsy

and allopregnanole’s known activity at the GABAR, the question remained if the

GABAR and calcium signaling could be related. In fact, a bidirectional

relationship between GABAR and L-type VGCC has been recently suggested. In

a study by Katsura et al. (2007) chronic exposure to benzodiazepines, GABAR

modulators, resulted in a change in L-type VGCC subunit expression and ultimately caused a change in Ca2+ influx in response to depolarization in these

cells. Conversely, L-type VGCC blockers have been shown to alter the effects of

withdrawal from GABAR modulators. El Ganouni (2004) has shown that

nifendipine, an L-type VGCC blocker, prevented anxiogenic-like effects caused

by (a GABAR modulator) withdrawal when administered both acutely

86 or concurrently. Gupta et al. (1996) also showed that nifendipine, as well as

nimodipine and (also L-type VGCC blockers) showed dose dependent blockade of all withdrawal symptoms (including audiogenic seizures) of

, another benzodiazepine. Clearly, these data indicate that GABAR

modulators and L-type VGCC have some relationship that should be further

investigated.

Fig. 16 Model of Possible Mechanism for Ca2+ Entry After Allopregnanole Treatment Cl- Allo

VGCC GABAA R

Ca2+ Ca2+ signaling Alteration of pathway GABAR subunit

Nucleus

Fig. 16 A model of the hypothesized effect of allopregnanole treatment on cortical neurons. First allopreganone activates an efflux of chloride from 2+ the GABAAR causing a depolarization and influx of Ca and after chronic allopregananole exposure mitochondria may alter Ca2+ signaling ultimately changing transcription and GABAR subunit composition.

I hypothesized that allopregnanolone activates an efflux of chloride

2+ through the GABAAR initiating a depolarization and Ca influx through VGCC.

This excess cytosolic calcium may lead to the changes in GABAAR subunit

87 composition demonstrated by Peirson (2005) which may alter seizure

succeptibility or drug efficacy. Mitochondria may play a role in this process by altering the Ca2+ signaling (Fig 16). In this study, I have examined the ability of

allopregnanolone to cause a Ca2+ influx in primary coritical neurons. Additionally,

I have examined the effects of chronic allopregnanolone treatment on mitochondrial membrane potential and cytosolic calcium levels in order to tease

out the role of Ca2+ and mitochondrial health in catamenial epilepsy.

Results

Dose Response for Allopregnanolone and GABA

To perform these and subsequent experiments, I used cortical neurons

prepared from embryonic rats and kept in culture for 14 days. Cells were loaded

with 2μM Fura-2 in the dark at room temperature for 30 minutes. Neurons were

identified by morphology (large cell body with processes) as determined by

Fig. 17 Immunocytochemistry of Cortical Neuron Cultures

MAP-2 GFAP

Fig. 17 Immunocytochemistry of cortical cultures. Both neurons and glia are present in the cultures. Data was collected only from cells with morphology similar to what was identified by MAP-2 staining. 88 immunocytochemistry of MAP-2 (Fig 17A) and other cells types were excluded from analysis in this study. Specifically, glial cells were identified by morphology as demonstrated by GFAP staining (Fig 17B).

After cells were loaded they were mounted on the bottom of a perfusion chamber. (See Methods for more details of this procedure.) Cells were then acutely treated with allopregnanolone or GABA for 5 seconds and Ca2+ changes were monitored by changes in the Fura-2 fluorescence ratio. The acute treatment of GABA or allopregnanolone caused a rapid increase in cytosolic Ca2+ that

Fig. 18 Effect of Acute Alloprenanolone or GABA on Cytosolic Calcium

A 1600 1500

1400

1300

1200

Fura-2 Ratio Fura-2 1100

1000

0 50 100 150 600 B Allo C 700 500

600 400 500

400 300

300 200 Fura-2 Ratio Fura-2 Ratio Fura-2 200 100 100 0

-9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 Gaba (μM) Allo (μM)

Fig. 18 A sample trace of cortical neuron response to acute allopregnanolone treatment. Primary cortical neurons were loaded with Fura-2AM and treated with a 5 second application of indicated compound. B Dose response curve for GABA. EC50 was determined to be 30µM. C Dose response curve for allopregnanolone. EC50 was determined to be 100nM.

89 typically returned to baseline within 100 seconds (Fig 18A). To quantify this

change, I measured the maximal increase in fluorescence ratio induced by GABA

or allopregnanolone for a variety of concentrations (GABA: 1nM, 30nM, 1μM, 10

μM, 30 μM, 300 μM, 1mM and 3mM. Allopregnanolone: 1pM, 100pM, 1nM,

100nM, 1 μM, 10 μM, and 30 μM) and used IGOR softwear to create a dose response curve for condition. I then used the dose response curve to determine

the EC50 for each treatment (GABA: 30μM Allopregnanolone: 100nM ) (Fig 18)

which were both within physiological ranges for each compound. Unless

otherwise stated, the EC50 concentrations were used for all acute treatment

experiments.

To be sure that the GABA and allopregnanolone evoked Ca2+ rise was not

a non-specific effect, I used two controls, and an inactive form of

allopregnanolone. Cholesterol was chosen because the structure is similar to

allopregnanolone, and the inactive allopregnanolone is a common control for

allopregnanolone. As shown in Fig. 19 acute treatment with cholesterol had no

effect on cytosolic Ca2+ levels (N=20), nor did the inactive allopregnanolone

(N=19). These results indicate that the evoked Ca2+ rise is specific for

allopregnanolone and GABA.

90

Fig. 19 Cortical Neurons Calcium Response to Cholesterol AB

1020

1000 1050 980 1000 960 950 940 Fura Ratio 900

920 Fura Ratio 850 900

0 20 40 60 80 100 120 800

Time (min) 0 20 40 60 80 100 Cholesterol Allo Time (min) Inactive allo Allo

Fig. 19 Rat cortical neurons were acutely treated with cholesterol or inactive allopregnanolone to test for non-specific affects on calcium uptake. No responses were seen with either compound. Active allopregnanolone was used as a control at (A) 90 second and (B) 75 seconds.

The Role of Internal Calcium Stores

After determining that acute treatment with allopregnanolone or GABA could induce a rise in Ca2+, I next wanted to identify the source of the cytosolic

Ca2+ increase. First, I investigated the possible role of internal Ca2+ stores,

2+ because GABABR are coupled to G proteins that can induce the release of Ca from internal stores and activate store operated Ca2+ entry. I began by using an inhibitor of internal Ca2+ store release 2-APB. This was shown to inhibit the rise of Ca2+ induced by ATP, which is known to activate Ca2+ release from internal stores. Cells were pre-treated for 30 minutes with 2-APB and then acutely treated with allopregnanolone or GABA while in the presence of 2-APB. For both conditions there was no effect of 2-APB on Ca2+ increase (GABA control 224.8 ±

91 29.46 treated 249.4 ± 54.59; Allo control 236.4 ± 29.46 treated 277.7 ± 18.31)

(Fig. 20C). Additionally, to determine if external Ca2+ was required, cells were

stimulated in the presence of 0 Ca2+ external solutions (0 Ca2+ PSS). (See

methods for complete description of 0 Ca2+ PSS.) Cells were put in 0 Ca2+ PSS for 2-5 minutes before acute treatment with GABA or allopregnanolone. In both cases there was no observed Ca2+ increase when cells were in 0 Ca2+ PSS

(GABA N=28; Allo N=18) (Fig 20A-B). Cells were treated with 100μM ATP as a control as it does not require external Ca2+ to cause cytosolic Ca2+ increase. This

evidence indicated that the Ca2+ increase was likely from an external source.

Fig. 20 Role of Internal Calcium Stores A B 1600 1200

1400 1100 1200 1000 Fura Ratio Fura Ratio 1000 900

800

0 50 100 150 ATP Data 1 0 50 100 150 200 TIme TIme 450 400 ATP 350 300 250 200 Fura Ratio 150 100 50 0 Allo Allo 2-APB GABA GABA 2 APB

Fig. 20 Cultered cortical neurons were acutely treated with GABA at 25 sec (A) or allopregnanolone at 25 sec (B) in the presence of 0 calcium. No response was observed. Cells were treated with ATP as a control at (A) 125 seconds (B) 100 seconds. C. Cells were treated with 2-APB. No effect on calcium entry was observed.

92 Role of L-Type Voltage Gated Calcium Channels

To determine if L-type Ca2+ channels could be the source of Ca2+ influx, I

utilized the L-type VGCC blocker nifedipine. I investigated the role of these

channels because they are fast activating and abundant on neurons, making

them a good candidate for the rapid Ca2+ increase seen after acute treatment.

Additionally, L-type VGCC blockers have been shown to affect changes in Ca2+ signaling associated with effects of GABA modulators (Rabbani and Little, 1999), and allopregnanolone is a GABA modulator. Cells were pretreated with nifedipine for 3 minutes before stimulation with acute allopregnanolone or GABA. In the presence of nifedipine, no Ca2+ increase was observed (GABA N=23; Allo N=42)

(Fig 21). This indicates that the L-type VGCCs may be involved in the Ca2+ influx induced by GABA or allopregnanolone acute treatment.

Fig. 21 Effect of Nifedipine and on Calcium Influx

B A 1200

1100 500

450 1000

400 900

350 Fura-2 Ratio Fura-2 Ratio 800 300

0 50 100 150 TIme 0 50 100 150 200 TIme GABA ATP GABA ATP

Fig. 21 Nifedipine, a blocker of L-type VGCC successfully blocked the calcium increase induced by acute GABA or allopregnanolone treatment. (A) Cortical Neurons in the presence of nifedipine were treated with 30μM GABA at 20 seconds. Cells were then treated with 100μM ATP at 110 seconds as a control. (B) Bumetanide, a blocker of KCCl1 transporter also blocked the acute effects of allopregnanolone or GABA treatment. In this figure, cells were treated with 30μM GABA in the presence of bumetanide, and with 100μM ATP as a control at 70 seconds. 93 GABA Depolarization

These results were puzzling, because allopregnanolone is not known to

have a direct effect on L-type VGCCs; but somehow these channels appeared to

be activated after acute treatment. To address this question we used a blocker

for the NKCCl1 chloride exchanger, bumetanide. If GABA and allopregnanolone

acute treatment were causing an efflux of chloride and subsequent depolarization

of the cell to activate L-type VGCCs, then disrupting the chloride gradient would

inhibit this effect. Bumetanide disrupts the chloride gradient by inhibiting the

NKCCl1 chloride exchanger. Cells were pretreated with bumetanide for 30

minutes after being loaded with Fura-2. In the continued presence of

bumetanide, neither GABA nor allopregnanolone were able to induce a Ca2+ increase in cortical neuron cultures (GABA N=30; Allo N=21) (Fig. 21B). This indicates that the acute treatment of GABA or allopregnanolone causes a depolarization through the efflux of chloride, inducing a calcium influx through the

VGCC, however because bumetanide diminishes the chloride gradient, this does not indicate what channel is involved in the chloride efflux.

GABA Channel Blockers

Because allopregnanolone is a GABAAR modulator, and previous data

has shown affects of allopregnanolone on GABAAR subunit expression (Peirson

et al 2005) I first looked at the possible role of GABAAR in the depolarization of

the neurons. As described previously, there are three types of GABA channels,

GABAA, GABAB, and GABAC. To investigate the role of the GABAAR, I used a

94 GABAAR antagonist bicuculine. Bicuculine requires an initial activation of the

GABAAR to block the channel, so we pretreated the cells for either 5 or 100 seconds. In both cases we saw no effect of bicuculine to block the Ca2+ increase

induced by GABA or allopregnanolone (GABA 224.8 ± 14.22 control 176.1 ±

21.43 treated; Allo 236.4 ± 29.46 control 308.4 ± 43.60 treated) (Fig 22A). Next, I

looked at the GABAB channel using saclofen, a GABAB antagonist. Again cells

were pretreated for 30 minutes before acute treatment with GABA or

allopregnanolone. Again, there was no effect of saclofen on the Ca2+ increase

(GABA 224.8 ± 14.22 control 220.6 ± 28.9 treated; Allo 236.4 ± 29.46 control

324.2 ± 64.97 treated) (Fig 22B). Lastly, I used the GABAAR and GABACR blocker picrotoxin. Once more there was no change in Ca2+ increase after acute

treatment (data not shown). Because the channel often has to be opened to see

an effect of picrotoxin, I also tried acute treatment of GABA before application of

picrotoxin, then a second acute treatment of GABA or allopregnanolone. This

also had no effect on Ca2+ increase (data not shown).

Fig. 22 Role of GABA Channel Blockers Data 1 B A Data 1 400 400

300 300

200 200

100 Fura-2 Ratio Fura-2 Ratio 100

0 Allo Con Allo Bic Gaba Con Gaba Bic 0 Allo Allo SAC GABA GABA SAC

Fig. 22 Blockers of GABAAR (A) bicucline or GABABR (B) saclofen had no effect on the increase in cytosolic calcium induced by either allopregnanolone or GABA. 95 Effect of Chronic Treatment

Once I established that allopregnanolone can induce a Ca2+ increase, I next wanted to examine the effects of chronic allopregnanolone treatment on

Ca2+ signaling. Based on previous work by Pierson (2005), chronic treatment was established to be 48 hours treatment with 1nM allopregnanolone. After 48 hours the allopregnanolone was washed off and cells were loaded with Fura-2 just as in the previous experiments. Because of the role of mitochondria in Ca2+ signaling and synaptic plasticity, I first looked at the effects of chronic allopregnanolone treatment on mitochondria. We used the mitochondrial membrane potential dye

JC-1 to observe differences in mitochondrial membrane potential after chronic treatment. I found that the chronically treated neurons exhibited a significantly different membrane potential compared to control (Fig 23) (p<0.001, 2910 ±

72.49 control 872.4. ± 58.96 treated) This effect could be involved in altering the

Ca2+ signaling induced by acute GABA or allopregnanolone treatment.

Fig. 23 Effect of Chronic Allopregnanolone on Mitochondrial Membrane PotentialData 1 3000

2000 TMRM

1000

0 Control Chronic Treated

Fig. 23 Cortical cells treated with 0.1% DMSO (Control) or 1μM allopregnanolonefor 48 hours then loaded with TMRM 12.5nM. Cells treated with allopregnanolone showed a significacntly lower TMRM fluorescence then control cells. * indicates p<0.001. 96 I next utilized the Ca2+ indicator dye Fura-2 to determine changes in resting Ca2+ levels, and induced Ca2+ levels. First, I demonstrated a significant

change in resting Ca2+ levels in chronically treated cells. A significantly higher

resting Ca2+ was observed in neurons chronically treated with allopregnanolone.

This effect was expected given that mitochondrial membrane potential was

significantly less negative. Because there are two channels involved in the Ca2+ increase after allopregnanolone stimulation, I used a high potassium stimulation to investigate changes in VGCC. I found no significant difference in the Ca2+ influx induced by high potassium. Next, I used Ca2+ increase induced by acute

allopregnanolone or GABA stimulation. I chose three different concentrations of

GABA or allopregnanolone for the acute treatment: a low concentration (1pM

allopregnanolone, 10 μM GABA), middle (100nM allopregnanolone, 30μM

GABA), and high concentration (10μM allopregnanolone, 300μM GABA). These concentrations were chosen based on data from the dose response curves, the low concentration was at the bottom of the curve, the high concentration is at the top of the curve, and the middle concentration is the EC50. Chronically treated

cells showed a significantly lower Ca2+ increase after acute GABA treatment (Fig

24B) (GABA: low p=0.0042, 303.6 ± 42.41, 78.44 ± 13.29; middle p<0.0001,

224.8 ± 14.22, 74.67 ± 14.73; high p<0.0001, 680.0 ± 61.90 222.0. ± 42.93;

allopregnanolone: middle p<0.0001, 236.4 ± 29.46, 27.88 ± 2.083; high p<0.001,

459.6 ± 42.71, 33.14 ± 5.352). Additionally, chronically treated cells exhibited a

significantly lower Ca2+ rise induced by acute allopregnanolone treatment (Fig

24A); and the low concentration had no indication of a rise in Ca2+ in the

97 chronically treated cells. These results indicate that chronic allopregnanolone

treatment caused a change in some aspect of the mechanism of Ca2+ increase

induced by acute GABA or allopregnanolone treatment.

Fig. 24 Effect of Chronic Allopregnanolone Treatment on Calcium Influx

A Data 1 B Allo GABAData 1 600 750 500

400 500 300 * 200 Fura-2 Ratio Fura-2 Ratio 250 100 * *** 0 0 1pM 100nM 10μM 10μM 30μM 300μM 100 nM Chronic 1 pM Chronic 10 nM Chronic 10 µM Chronic 30 µM Chronic 300 µM Chronic Fig. 24 Chronic Treatment of allopregnanolone caused significantly less calcium increase after GABA or allopregnanolone stimulation. Primary cortical neurons were treated for 48 hours with 1μM allopregnanolone, then loaded with Fura-2 AM as before and calcium influx was induced by acute treatment of GABA or allorepgnanolone at the indicated concentration. Chronically treated cells had no calcium increase with 1pM acute allopregnanolone stimulation. * indicates p<0.0001.

Discussion

In this study, I have shown that GABA and allopregnanolone can induce a

cytosolic Ca2+ increase in a dose dependent manner in neurons after 14 days in

culture. I have shown that this effect is specific, as cholesterol does not cause

the same Ca2+ increase. However, chronic (48 hour) treatment with allopregnanolone causes the amount of Ca2+ increase to be diminished for both

98 acute GABA and acute allopregnanolone treatment. Additionally, the Ca2+ is not coming from internal stores, but from external solution. Nifedipine blocks the Ca2+

increase indicating that the L-type VGCC are likely the plasma membrane

channel that Ca2+ is entering the cell through. Lastly, I have shown that inhibiting

the NKCCl1 transporter also eliminates the Ca2+ increase induced by GABA and

allopregnanolone. This indicates that these compounds are causing chloride to

exit the cell leading to depolarization of the cell and activation of the VGCC.

In the first experiments, I treated cortical neurons with various

concentrations of GABA or allopregnanolone and measured the increase in

cytosolic Ca2+ with Fura-2. I used this information to create dose response curves

using IGOR software and identified the EC50 for both compounds. The EC50 was used for all future experiments unless were otherwise indicated. To test for any non-specific response to steroids, I determined whether cholesterol could induce a Ca2+ increase. Cholesterol had no effect on cytosolic Ca2+ levels as indicated

by Fura-2. This indicates that GABA and allopregnanolone are having a stero-

specific effect on the cortical neurons to induce a cytosolic Ca2+ increase. While

GABA has been shown to be depolarizing in some immature neurons (Grobin et

al., 2006), this has not been shown in “mature” neurons. Cortical neurons in

culture for 14 days are considered “mature”, so this response is novel.

Additionally, the mechanism involved in the cytosolic Ca2+ increase, including the

location of Ca2+ entry was previously unknown.

Next, I looked at the effect that chronic allopregnanolone treatment had on the Ca2+ increase induced by acute treatment. I compared three different

99 concentrations, a low middle and high concentration, for each acute treatment

conditions after chronic allopregnanolone treatment. I found that the chronic

treatment caused a decrease in Ca2+ influx after acute GABA or allopregnanolone treatment, except in the case of the chronically treated low

dose acute allopregnanolone treatment, which had no Ca2+ influx. These results

indicate that chronic exposure to allopregnanolone, as may be the case for a

woman during her menstrual cycle, may alter GABA's effect on chronic neurons.

This may alter a woman’s seizure susceptibility at certain times during her cycle.

In order to investigate this further, we must first understand the mechanism of the

induced Ca2+ increase.

Therefore, I set out to identify the source of the Ca2+ increase by first

investigating the role of internal Ca2+ stores versus Ca2+ influx from external

solution. I used 2-APB, an inhibitor of internal Ca2+ release, and found that it had

no effect on acute GABA or allopregnanolone treatment. Because 2-APB has

been shown to have different effect in different cell types, I also examined the

location of Ca2+ entry using cells in 0 Ca2+ external solutions. In both cases,

when cells were in 0 Ca2+ external solution there was no increase in cytosolic

Ca2+ after acute treatment. This indicates that the Ca2+ increase is coming from

the external solutions through a plasma membrane channel and not from internal

Ca2+ stores.

There are several sources for Ca2+ entry at the plasma membrane;

however, the knowledge that GABA can be depolarizing led us to begin our

search with the L-type VGCC. I used the L-type VGCC blocker nifedipine to

100 investigate the role of these channels after the acute treatment of GABA or

allopregnanolone. I found that in both cases nifedipine blocked all Ca2+ entry. To

be sure the cells would still respond to stimulus I used ATP to induce Ca2+ release from the internal stores and observed an increase in intracellular Ca2+.

This data combined with the 0 external Ca2+ data indicated that in fact the Ca2+ entry was from external solution through the VGCC.

This indicated that the chronic treatment of GABA and allopregnanolone

was causing the GABA channels to allow chloride to exit the cell leading to a

depolarization. To test this I used bumetanide, a blocker of the NKCC1

transporter. This transporter is electro-neutral and is responsible for entry of

sodium, potassium and chloride into the cell. Its activity is believed to be responsible for maintaining the chloride gradient (Pieraut et al., 2007). Use of

bumetanide disrupts the chloride gradient so that it will not move through plasma

membrane channels in either direction. Bumetanide, like nifedipine, blocked the

Ca2+ influx induced by acute GABA and allopregnanolone treatment. However,

because bumetanide disrupts the chloride gradient, the effect could be from any

number of chloride selective ion channels.

To further investigate the source of chloride efflux, I began treating cells

with various GABA channel blockers. The GABAbR blocker saclofen had no

effect on Ca2+ influx; however, because of the fast kinetics of the response

GABAbR are an unlikely source of the chloride efflux, due to the fact that

GABAbR are metabotropic receptors and have slower kinetics. More interesting,

2+ however, is that the GABAaR blocker bicuculine also had no effect on the Ca

101 increase. This was surprising given the evidence that allopregnanolone can

effect many of the GABAaR subunits, including subunits that have known effects

on drug susceptibility and synaptic currents. It is possible that the effects are due

to GABAcR, which are similar in function and structure to the GABAaR, but have different pharmacology including that they are not blocked by bicuculline (Enz and Cutting, 1999). Furthermore, there is evidence that allopregnanolone may have a direct effect on GABAcR (Li et al., 2006). Additionally there are other

chloride sensitive channels and transporters on the plasma membrane of cortical

neurons. Conversely, it is not known if allopregnanolone has any effects on these

channels.

102 Chapter VII

With these experiments I have demonstrated that Ca2+ signaling and

mitochondrial health are important players in the etiology of three different

disease states. While the importance of Ca2+ signaling in cell growth and

apoptosis was known, the first set of experiments illustrated that modulating Ca2+ signaling by altering mitochondrial health may be an effective treatment in some types of cancer. Furthermore, identifying a drug that can specifically target mitochondria in cancer cells may be an effective treatment with minimal side effects for some cancer types. In the second set of experiments, a known neurotoxin, Aβ, exhibited differential effects on mitochondria and Ca2+ signaling

in different cell types. While Ca2+ and mitochondria have been thought to be

involved in pathology of AD, this differential effect and differential uptake of the

peptide may explain why the hypothalamus is not affected in early AD. More importantly, this may indicate a possible treatment or preventative mechanisms

for AD. In the last set of experiments, I demonstrated that chronic

allopregnanolone treatment can alter Ca2+ signaling and mitochondrial health.

These changes may be responsible for the changes observed by Pierson (2005)

in GABAR subunit composition and ultimately effect seizure susceptibility. More

studies are needed to fully understand the Ca2+ signaling pathway involved and how this pathway may be exploited for possible treatment of catamenial epilepsy.

Through these examples, I have shown that understanding the role of calcium

signaling and mitochondrial function in disease are important tools in

understanding the underlying mechanism of disease pathology.

103

Conclusions:

1. CAI acute treatment causes mitochondria in HEK-293 cells to be less

polarized in a time and concentration dependent manner.

2. CAI acute treatment affects the ability of mitochondria to sequester Ca2+;

and this effect is independent of any effects of CAI on Ca2+ channels at

the plasma membrane.

3. Amyloid-ß has an effect on mitochondrial membrane potential in pituicytes

causing them to become less polarized; however, there is no effect of Aβ

on the mitochondrial membrane potential in nerve terminals of the

posterior pituitary. .

4. Amyloid-ß causes a significant decrease in the resting Ca2+ levels of nerve

endings, and a significant increase in the resting Ca2+ levels of pituicytes

5. Amyloid-ß treatment causes a significantly higher maximal change in Ca2+

after stimulus in both pituicytes and nerve endings.

6. Fluorescently tagged Aβ was taken up by pituicytes in culture; however, it

was not taken up by nerve endings within 48 hours.

7. Acute treatment of primary rat cortical neurons with either GABA or

allopregnanolone cause a rapid rise in cytosolic Ca2+ levels in a

concentration dependent manner.

8. The Ca2+ rise induced by acute GABA or allopregnanolone treatment was

blocked by both nifedipine and bumetanide, but not by bicuculine,

saclofen, picrotoxin or 2-APB.

104 9. Chronic treatment of cortical neurons with 1μM allopregnanolone caused a

significant decrease in the maximal rise in cytosolic Ca2+ after stimulus

with GABA or allopregnanolone acutely.

10. Chronic treatment of cortical neurons with 1μM allopregnanolone caused a

significant decrease in mitochondrial membrane polarization compared to

control.

105 References

Abramov, A. Y., L. Canevari and M. R. Duchen. (2003). "Changes in intracellular

calcium and glutathione in astrocytes as the primary mechanism of

amyloid neurotoxicity." J Neurosci 23: 5088-95.

Abramov, A. Y., L. Canevari and M. R. Duchen. (2004). "Calcium signals induced

by amyloid beta peptide and their consequences in neurons and

astrocytes in culture." Biochim Biophys Acta 1742: 81-7.

Babcock, D. F., J. Herrington, P. C. Goodwin, Y. B. Park and B. Hille. (1997).

"Mitochondrial participation in the intracellular Ca2+ network." J Cell Biol

136: 833-44.

Bakowski, D. and A. B. Parekh. (2007). "Regulation of store-operated calcium

channels by the intermediary metabolite pyruvic acid." Curr Biol 17: 1076-

81.

Barbeito, L. H., M. Pehar, P. Cassina, M. R. Vargas, H. Peluffo, L. Viera, A. G.

Estevez and J. S. Beckman. (2004). "A role for astrocytes in motor neuron

loss in amyotrophic lateral sclerosis." Brain Res Brain Res Rev 47: 263-

74.

106 Beam, K. G., T. Tanabe and S. Numa. (1989). "Structure, function, and

regulation of the skeletal muscle dihydropyridine receptor." Ann N Y Acad

Sci 560: 127-37.

Begley, J. G., W. Duan, S. Chan, K. Duff and M. P. Mattson. (1999). "Altered

calcium homeostasis and mitochondrial dysfunction in cortical synaptic

compartments of presenilin-1 mutant mice." J Neurochem 72: 1030-9.

Belelli, D. and J. J. Lambert. (2005). "Neurosteroids: endogenous regulators of

the GABA(A) receptor." Nat Rev Neurosci 6: 565-75.

Berridge, M. J. (2006). "Calcium microdomains: organization and function." Cell

Calcium 40: 405-12.

Berridge, M. J., M. D. Bootman and P. Lipp. (1998). "Calcium--a life and death

signal." Nature 395: 645-8.

Bettler, B., K. Kaupmann, J. Mosbacher and M. Gassmann. (2004). "Molecular

structure and physiological functions of GABA(B) receptors." Physiol Rev

84: 835-67.

107 Bidaud, I., A. Mezghrani, L. A. Swayne, A. Monteil and P. Lory. (2006). "Voltage-

gated calcium channels in genetic diseases." Biochim Biophys Acta 1763:

1169-74.

Bodding, M. (2007). "TRP proteins and cancer." Cell Signal 19: 617-24.

Bootman, M. D., P. Lipp and M. J. Berridge. (2001). "The organisation and

functions of local Ca(2+) signals." J Cell Sci 114: 2213-22.

Borner, C. (2003). "The Bcl-2 protein family: sensors and checkpoints for life-or-

death decisions." Mol Immunol 39: 615-47.

Broad, L. M., D. L. Armstrong and J. W. Putney, Jr. (1999). "Role of the inositol

1,4,5-trisphosphate receptor in Ca(2+) feedback inhibition of calcium

release-activated calcium current (I(crac))." J Biol Chem 274: 32881-8.

Caille, I., B. Allinquant, E. Dupont, C. Bouillot, A. Langer, U. Muller and A.

Prochiantz. (2004). "Soluble form of amyloid precursor protein regulates

proliferation of progenitors in the adult subventricular zone." Development

131: 2173-81.

Calabresi, P., P. Gubellini, B. Picconi, D. Centonze, A. Pisani, P. Bonsi, P.

Greengard, R. A. Hipskind, E. Borrelli and G. Bernardi. (2001). "Inhibition

108 of mitochondrial complex II induces a long-term potentiation of NMDA-

mediated synaptic excitation in the striatum requiring endogenous

dopamine." J Neurosci 21: 5110-20.

Canevari, L., A. Y. Abramov and M. R. Duchen. (2004). "Toxicity of amyloid beta

peptide: tales of calcium, mitochondria, and oxidative stress." Neurochem

Res 29: 637-50.

Carrasco, M. A. and C. Hidalgo. (2006). "Calcium microdomains and gene

expression in neurons and skeletal muscle cells." Cell Calcium 40: 575-83.

Carrasco, M. P., J. M. Jimenez-Lopez, J. L. Segovia and C. Marco. (2002).

"Comparative study of the effects of short- and long-term ethanol

treatment and alcohol withdrawal on phospholipid biosynthesis in rat

hepatocytes." Comp Biochem Physiol B Biochem Mol Biol 131: 491-7.

Case, R. M., D. Eisner, A. Gurney, O. Jones, S. Muallem and A. Verkhratsky.

(2007). "Evolution of calcium homeostasis: from birth of the first cell to an

omnipresent signalling system." Cell Calcium 42: 345-50.

Chao, T. S., K. L. Byron, K. M. Lee, M. Villereal and M. R. Rosner. (1992).

"Activation of MAP kinases by calcium-dependent and calcium-

109 independent pathways. Stimulation by thapsigargin and epidermal growth

factor." J Biol Chem 267: 19876-83.

Chavas, J., M. E. Forero, T. Collin, I. Llano and A. Marty. (2004). "Osmotic

tension as a possible link between GABA(A) receptor activation and

intracellular calcium elevation." Neuron 44: 701-13.

Chuang, Y. C., A. Y. Chang, J. W. Lin, S. P. Hsu and S. H. Chan. (2004).

"Mitochondrial dysfunction and ultrastructural damage in the hippocampus

during kainic acid-induced status epilepticus in the rat." Epilepsia 45:

1202-9.

Ciapa, B., D. Pesando, M. Wilding and M. Whitaker. (1994). "Cell-cycle calcium

transients driven by cyclic changes in inositol trisphosphate levels." Nature

368: 875-8.

Cory, S. and J. M. Adams. (2002). "The Bcl2 family: regulators of the cellular life-

or-death switch." Nat Rev Cancer 2: 647-56.

Cory, S., D. C. Huang and J. M. Adams. (2003). "The Bcl-2 family: roles in cell

survival and oncogenesis." Oncogene 22: 8590-607.

110 Denmeade, S. R. and J. T. Isaacs. (2005). "The SERCA pump as a therapeutic

target: making a "smart bomb" for prostate cancer." Cancer Biol Ther 4:

14-22.

Dolmetsch, R. E. and R. S. Lewis. (1994). "Signaling between intracellular Ca2+

stores and depletion-activated Ca2+ channels generates [Ca2+]i

oscillations in T lymphocytes." J Gen Physiol 103: 365-88.

Dolmetsch, R. E., K. Xu and R. S. Lewis. (1998). "Calcium oscillations increase

the efficiency and specificity of gene expression." Nature 392: 933-6.

Drachman, D. B., K. Frank, M. Dykes-Hoberg, P. Teismann, G. Almer, S.

Przedborski and J. D. Rothstein. (2002). "Cyclooxygenase 2 inhibition

protects motor neurons and prolongs survival in a transgenic mouse

model of ALS." Ann Neurol 52: 771-8.

Duchen, M. R. (2000). "Mitochondria and calcium: from cell signalling to cell

death." J Physiol 529 Pt 1: 57-68.

Ekinci, F. J., K. U. Malik and T. B. Shea. (1999). "Activation of the L voltage-

sensitive calcium channel by mitogen-activated protein (MAP) kinase

following exposure of neuronal cells to beta-amyloid. MAP kinase

111 mediates beta-amyloid-induced neurodegeneration." J Biol Chem 274:

30322-7.

El Ganouni, S., N. Hanoun, C. Boni, A. Tazi, F. Hakkou and M. Hamon. (2004).

"Prevention of diazepam withdrawal syndrome by nifedipine-behavioural

and neurochemical studies." Pharmacol Biochem Behav 79: 269-77.

Enfissi, A., S. Prigent, P. Colosetti and T. Capiod. (2004). "The blocking of

capacitative calcium entry by 2-aminoethyl diphenylborate (2-APB) and

carboxyamidotriazole (CAI) inhibits proliferation in Hep G2 and Huh-7

human hepatoma cells." Cell Calcium 36: 459-67.

Enz, R. and G. R. Cutting. (1999). "GABAC receptor rho subunits are

heterogeneously expressed in the human CNS and form homo- and

heterooligomers with distinct physical properties." Eur J Neurosci 11: 41-

50.

Feske, S., J. Giltnane, R. Dolmetsch, L. M. Staudt and A. Rao. (2001). "Gene

regulation mediated by calcium signals in T lymphocytes." Nat Immunol 2:

316-24.

Feske, S., Y. Gwack, M. Prakriya, S. Srikanth, S. H. Puppel, B. Tanasa, P. G.

Hogan, R. S. Lewis, M. Daly and A. Rao. (2006). "A mutation in Orai1

112 causes immune deficiency by abrogating CRAC channel function." Nature

441: 179-85.

Foldvary-Schaefer, N. and T. Falcone. (2003). "Catamenial epilepsy:

pathophysiology, diagnosis, and management." Neurology 61: S2-15.

Fu, H., W. Li, Y. Lao, J. Luo, N. T. Lee, K. K. Kan, H. W. Tsang, K. W. Tsim, Y.

Pang, Z. Li, D. C. Chang, M. Li and Y. Han. (2006). "Bis(7)-tacrine

attenuates beta amyloid-induced neuronal apoptosis by regulating L-type

calcium channels." J Neurochem 98: 1400-10.

Gault, L. M. and R. E. Siegel. (1997). "Expression of the GABAA receptor delta

subunit is selectively modulated by depolarization in cultured rat cerebellar

granule neurons." J Neurosci 17: 2391-9.

Gegelashvili, M., A. Rodriguez-Kern, L. Sung, K. Shimamoto and G.

Gegelashvili. (2007). "Glutamate transporter GLAST/EAAT1 directs cell

surface expression of FXYD2/gamma subunit of Na, K-ATPase in human

fetal astrocytes." Neurochem Int 50: 916-20.

Gilabert, J. A., D. Bakowski and A. B. Parekh. (2001). "Energized mitochondria

increase the dynamic range over which inositol 1,4,5-trisphosphate

activates store-operated calcium influx." Embo J 20: 2672-9.

113

Gilabert, J. A. and A. B. Parekh. (2000). "Respiring mitochondria determine the

pattern of activation and inactivation of the store-operated Ca(2+) current

I(CRAC)." Embo J 19: 6401-7.

Gincel, D., S. D. Silberberg and V. Shoshan-Barmatz. (2000). "Modulation of the

Voltage-Dependent Anion Channel (VDAC) by Glutamate1." J Bioenerg

Biomembr 32: 571-83.

Giovannucci, D. R., M. D. Hlubek and E. L. Stuenkel. (1999). "Mitochondria

regulate the Ca(2+)-exocytosis relationship of bovine adrenal chromaffin

cells." J Neurosci 19: 9261-70.

Glitsch, M. D., D. Bakowski and A. B. Parekh. (2002). "Store-operated Ca2+

entry depends on mitochondrial Ca2+ uptake." Embo J 21: 6744-54.

Golovina, V. A. (1999). "Cell proliferation is associated with enhanced

capacitative Ca(2+) entry in human arterial myocytes." Am J Physiol 277:

C343-9.

Golovina, V. A., O. Platoshyn, C. L. Bailey, J. Wang, A. Limsuwan, M. Sweeney,

L. J. Rubin and J. X. Yuan. (2001). "Upregulated TRP and enhanced

114 capacitative Ca(2+) entry in human pulmonary artery myocytes during

proliferation." Am J Physiol Heart Circ Physiol 280: H746-55.

Gray, A. C., J. Raingo and D. Lipscombe. (2007). "Neuronal calcium channels:

splicing for optimal performance." Cell Calcium 42: 409-17.

Grobin, A. C., S. Gizerian, J. A. Lieberman and A. L. Morrow. (2006). "Perinatal

allopregnanolone influences prefrontal cortex structure, connectivity and

behavior in adult rats." Neuroscience 138: 809-19.

Gudermann, T. and S. Roelle. (2006). "Calcium-dependent growth regulation of

small cell lung cancer cells by neuropeptides." Endocr Relat Cancer 13:

1069-84.

Guo, L., Z. S. Li, H. L. Wang, C. Y. Ye and D. C. Zhang. (2006). "Carboxyamido-

triazole inhibits proliferation of human breast cancer cells via G(2)/M cell

cycle arrest and apoptosis." Eur J Pharmacol 538: 15-22.

Gupta, M. B., C. Nath, G. K. Patnaik and R. C. Saxena. (1996). "Effect of calcium

channel blockers on withdrawal syndrome of lorazepam in rats." Indian J

Med Res 103: 310-4.

115 Haeberlein, S. L. (2004). "Mitochondrial function in apoptotic neuronal cell

death." Neurochem Res 29: 521-30.

Hajnoczky, G., E. Davies and M. Madesh. (2003). "Calcium signaling and

apoptosis." Biochem Biophys Res Commun 304: 445-54.

Hajnoczky, G., R. Hager and A. P. Thomas. (1999). "Mitochondria suppress local

feedback activation of inositol 1,4, 5-trisphosphate receptors by Ca2+." J

Biol Chem 274: 14157-62.

Hardie, R. C. and B. Minke. (1993). "Novel Ca2+ channels underlying

transduction in Drosophila photoreceptors: implications for

phosphoinositide-mediated Ca2+ mobilization." Trends Neurosci 16: 371-

6.

Hatton, G. I. (2002). "Glial-neuronal interactions in the mammalian brain." Adv

Physiol Educ 26: 225-37.

Hatzinger, M., S. Brand, S. Perren, A. von Wyl, K. von Klitzing and E. Holsboer-

Trachsler. (2007). "Hypothalamic-pituitary-adrenocortical (HPA) activity in

kindergarten children: importance of gender and associations with

behavioral/emotional difficulties." J Psychiatr Res 41: 861-70.

116 Haverstick, D. M., T. N. Heady, T. L. Macdonald and L. S. Gray. (2000).

"Inhibition of human prostate cancer proliferation in vitro and in a mouse

model by a compound synthesized to block Ca2+ entry." Cancer Res 60:

1002-8.

Henshall, D. C., R. S. Clark, P. D. Adelson, M. Chen, S. C. Watkins and R. P.

Simon. (2000). "Alterations in bcl-2 and caspase gene family protein

expression in human temporal lobe epilepsy." Neurology 55: 250-7.

Herzog, A. G. (1995). "Progesterone therapy in women with complex partial and

secondary generalized seizures." Neurology 45: 1660-2.

Herzog, A. G. and C. A. Frye. (2003). "Seizure exacerbation associated with

inhibition of progesterone metabolism." Ann Neurol 53: 390-1.

Herzog, A. G., P. Klein and B. J. Ransil. (1997). "Three patterns of catamenial

epilepsy." Epilepsia 38: 1082-8.

Hoth, M., D. C. Button and R. S. Lewis. (2000). "Mitochondrial control of calcium-

channel gating: a mechanism for sustained signaling and transcriptional

activation in T lymphocytes." Proc Natl Acad Sci U S A 97: 10607-12.

117 Hoth, M. and R. Penner. (1992). "Depletion of intracellular calcium stores

activates a calcium current in mast cells." Nature 355: 353-6.

Huang, C. L. (2004). "The transient receptor potential superfamily of ion

channels." J Am Soc Nephrol 15: 1690-9.

Huber, D., P. Veinante and R. Stoop. (2005). "Vasopressin and oxytocin excite

distinct neuronal populations in the central amygdala." Science 308: 245-

8.

Hussain, M. M., H. Kotz, L. Minasian, A. Premkumar, G. Sarosy, E. Reed, S.

Zhai, S. M. Steinberg, M. Raggio, V. K. Oliver, W. D. Figg and E. C. Kohn.

(2003). "Phase II trial of carboxyamidotriazole in patients with relapsed

epithelial ovarian cancer." J Clin Oncol 21: 4356-63.

Jiang, W., B. Du, Z. Chi, L. Ma, S. Wang, X. Zhang, W. Wu, X. Wang, G. Xu and

C. Guo. (2007). "Preliminary explorations of the role of mitochondrial

proteins in refractory epilepsy: Some Findings From Comparative

Proteomics." J Neurosci Res 85: 3160-70.

Jonas, E. A., D. Hoit, J. A. Hickman, T. A. Brandt, B. M. Polster, Y. Fannjiang, E.

McCarthy, M. K. Montanez, J. M. Hardwick and L. K. Kaczmarek. (2003).

118 "Modulation of synaptic transmission by the BCL-2 family protein BCL-xL."

J Neurosci 23: 8423-31.

Kamenetz, F., T. Tomita, H. Hsieh, G. Seabrook, D. Borchelt, T. Iwatsubo, S.

Sisodia and R. Malinow. (2003). "APP processing and synaptic function."

Neuron 37: 925-37.

Kanngiesser, U., P. Nalik and O. Pongs. (1988). "Purification and affinity labeling

of dihydropyridine receptor from rabbit skeletal muscle membranes." Proc

Natl Acad Sci U S A 85: 2969-73.

Kao, J. P., J. M. Alderton, R. Y. Tsien and R. A. Steinhardt. (1990). "Active

involvement of Ca2+ in mitotic progression of Swiss 3T3 fibroblasts." J

Cell Biol 111: 183-96.

Kapur, J. (2003). "Role of GABA(A) receptor-mediated inhibition in the

pathogenesis of generalized seizures." Exp Neurol 184: 1-2.

Katsura, M., M. Shibasaki, K. Kurokawa, A. Tsujimura and S. Ohkuma. (2007).

"Up-regulation of L-type high voltage-gated calcium channel subunits by

sustained exposure to 1,4- and 1,5-benzodiazepines in cerebrocortical

neurons." J Neurochem.

119 Kelly, B. L. and A. Ferreira. (2006). "beta-Amyloid-induced dynamin 1

degradation is mediated by N-methyl-D-aspartate receptors in

hippocampal neurons." J Biol Chem 281: 28079-89.

Kiechle, T., A. Dedeoglu, J. Kubilus, N. W. Kowall, M. F. Beal, R. M. Friedlander,

S. M. Hersch and R. J. Ferrante. (2002). "Cytochrome C and caspase-9

expression in Huntington's disease." Neuromolecular Med 1: 183-95.

Kim, H. L., Y. J. Chang, S. M. Lee and Y. S. Hong. (1998). "Genomic structure of

the regulatory region of the voltage-gated calcium channel alpha 1D." Exp

Mol Med 30: 246-51.

Kirichok, Y., G. Krapivinsky and D. E. Clapham. (2004). "The mitochondrial

calcium uniporter is a highly selective ." Nature 427: 360-4.

Kiselyov, K., J. Y. Kim, W. Zeng and S. Muallem. (2005). "Protein-protein

interaction and functionTRPC channels." Pflugers Arch 451: 116-24.

Kohn, E. C., C. C. Felder, W. Jacobs, K. A. Holmes, A. Day, R. Freer and L. A.

Liotta. (1994). "Structure-function analysis of signal and growth inhibition

by carboxyamido-triazole, CAI." Cancer Res 54: 935-42.

120 Kohn, E. C. and L. A. Liotta. (1990). "L651582: a novel antiproliferative and

antimetastasis agent." J Natl Cancer Inst 82: 54-60.

Lacor, P. N., M. C. Buniel, P. W. Furlow, A. S. Clemente, P. T. Velasco, M.

Wood, K. L. Viola and W. L. Klein. (2007). "Abeta oligomer-induced

aberrations in synapse composition, shape, and density provide a

molecular basis for loss of connectivity in Alzheimer's disease." J Neurosci

27: 796-807.

Laidlaw, J. (1956). "Catamenial epilepsy." Lancet 271: 1235-7.

Lawen, A. (2003). "Apoptosis-an introduction." Bioessays 25: 888-96.

Lawrie, A. M., R. Rizzuto, T. Pozzan and A. W. Simpson. (1996). "A role for

calcium influx in the regulation of mitochondrial calcium in endothelial

cells." J Biol Chem 271: 10753-9.

Leidenheimer, N. J. and R. Chapell. (1997). "Effects of PKC activation and

receptor desensitization on neurosteroid modulation of GABA(A)

receptors." Brain Res Mol Brain Res 52: 173-81.

121 Levy, M., G. C. Faas, P. Saggau, W. J. Craigen and J. D. Sweatt. (2003).

"Mitochondrial regulation of synaptic plasticity in the hippocampus." J Biol

Chem 278: 17727-34.

Li, G., L. Percontino, Q. Sun, A. S. Qazi and P. H. Frederikse. (2003). "Beta-

amyloid secretases and beta-amloid degrading enzyme expression in

lens." Mol Vis 9: 179-83.

Li, W., D. F. Covey, J. M. Alakoskela, P. K. Kinnunen and J. H. Steinbach.

(2006). "Enantiomers of neuroactive steroids support a specific interaction

with the GABA-C receptor as the mechanism of steroid action." Mol

Pharmacol 69: 1779-82.

Liou, J., M. L. Kim, W. D. Heo, J. T. Jones, J. W. Myers, J. E. Ferrell, Jr. and T.

Meyer. (2005). "STIM is a Ca2+ sensor essential for Ca2+-store-depletion-

triggered Ca2+ influx." Curr Biol 15: 1235-41.

Lipscombe, D., T. D. Helton and W. Xu. (2004). "L-type calcium channels: the

low down." J Neurophysiol 92: 2633-41.

Lipscombe, D., J. Q. Pan and A. C. Gray. (2002). "Functional diversity in

neuronal voltage-gated calcium channels by alternative splicing of

Ca(v)alpha1." Mol Neurobiol 26: 21-44.

122

Lory, P., I. Bidaud and J. Chemin. (2006). "T-type calcium channels in

differentiation and proliferation." Cell Calcium 40: 135-46.

Lu, K. P. and A. R. Means. (1993). "Regulation of the cell cycle by calcium and

calmodulin." Endocr Rev 14: 40-58.

Luo, D., L. M. Broad, G. S. Bird and J. W. Putney, Jr. (2001). "Mutual antagonism

of calcium entry by capacitative and arachidonic acid-mediated calcium

entry pathways." J Biol Chem 276: 20186-9.

Lyons, H. R., M. B. Land, T. T. Gibbs and D. H. Farb. (2001). "Distinct signal

transduction pathways for GABA-induced GABA(A) receptor down-

regulation and uncoupling in neuronal culture: a role for voltage-gated

calcium channels." J Neurochem 78: 1114-26.

Macdonald, R. L. and R. W. Olsen. (1994). "GABAA receptor channels." Annu

Rev Neurosci 17: 569-602.

MacManus, A., M. Ramsden, M. Murray, Z. Henderson, H. A. Pearson and V. A.

Campbell. (2000). "Enhancement of (45)Ca(2+) influx and voltage-

dependent Ca(2+) channel activity by beta-amyloid-(1-40) in rat cortical

123 synaptosomes and cultured cortical neurons. Modulation by the

proinflammatory cytokine interleukin-1beta." J Biol Chem 275: 4713-8.

Malli, R., M. Frieden, K. Osibow, C. Zoratti, M. Mayer, N. Demaurex and W. F.

Graier. (2003). "Sustained Ca2+ transfer across mitochondria is Essential

for mitochondrial Ca2+ buffering, sore-operated Ca2+ entry, and Ca2+

store refilling." J Biol Chem 278: 44769-79.

Mantha, A. K., K. Moorthy, S. M. Cowsik and N. Z. Baquer. (2006). "Membrane

associated functions of neurokinin B (NKB) on Abeta (25-35) induced

toxicity in aging rat brain synaptosomes." Biogerontology 7: 19-33.

Mantha, A. K., K. Moorthy, S. M. Cowsik and N. Z. Baquer. (2006).

"Neuroprotective role of neurokinin B (NKB) on beta-amyloid (25-35)

induced toxicity in aging rat brain synaptosomes: involvement in oxidative

stress and excitotoxicity." Biogerontology 7: 1-17.

Mattson, M. P. (2007). "Calcium and neurodegeneration." Aging Cell 6: 337-50.

Metea, M. R. and E. A. Newman. (2006). "Calcium signaling in specialized glial

cells." Glia 54: 650-5.

124 Mignen, O., C. Brink, A. Enfissi, A. Nadkarni, T. J. Shuttleworth, D. R.

Giovannucci and T. Capiod. (2005). "Carboxyamidotriazole-induced

inhibition of mitochondrial calcium import blocks capacitative calcium entry

and cell proliferation in HEK-293 cells." J Cell Sci 118: 5615-23.

Mohammad, S., A. Abolhassan and M. H. Pourgholami. (1998). "Evaluation of

the profile of progesterone in male amygdala-kindled rats."

Epilepsy Res 30: 195-202.

Monteith, G. R., D. McAndrew, H. M. Faddy and S. J. Roberts-Thomson. (2007).

"Calcium and cancer: targeting Ca2+ transport." Nat Rev Cancer 7: 519-

30.

Mukherjee, A., E. Song, M. Kihiko-Ehmann, J. P. Goodman, Jr., J. S. Pyrek, S.

Estus and L. B. Hersh. (2000). "Insulysin hydrolyzes amyloid beta

peptides to products that are neither neurotoxic nor deposit on amyloid

plaques." J Neurosci 20: 8745-9.

Mungarro-Menchaca, X., P. Ferrera, J. Moran and C. Arias. (2002). "beta-

Amyloid peptide induces ultrastructural changes in synaptosomes and

potentiates mitochondrial dysfunction in the presence of ryanodine." J

Neurosci Res 68: 89-96.

125 Newman, E. A. (2005). "Calcium increases in retinal glial cells evoked by light-

induced neuronal activity." J Neurosci 25: 5502-10.

Nilius, B. and T. Voets. (2005). "TRP channels: a TR(I)P through a world of

multifunctional cation channels." Pflugers Arch 451: 1-10.

Nunez, L., R. A. Valero, L. Senovilla, S. Sanz-Blasco, J. Garcia-Sancho and C.

Villalobos. (2006). "Cell proliferation depends on mitochondrial Ca2+

uptake: inhibition by salicylate." J Physiol 571: 57-73.

Okamura, N., T. Suemoto, H. Shimadzu, M. Suzuki, T. Shiomitsu, H. Akatsu, T.

Yamamoto, M. Staufenbiel, K. Yanai, H. Arai, H. Sasaki, Y. Kudo and T.

Sawada. (2004). "Styrylbenzoxazole derivatives for in vivo imaging of

amyloid plaques in the brain." J Neurosci 24: 2535-41.

Parducz, A. and F. Joo. (1976). "Visualization of stimulated nerve endings by

preferential calcium accumulation of mitochondria." J Cell Biol 69: 513-7.

Parekh, A. B. and J. W. Putney, Jr. (2005). "Store-operated calcium channels."

Physiol Rev 85: 757-810.

Park, M. K., M. C. Ashby, G. Erdemli, O. H. Petersen and A. V. Tepikin. (2001).

"Perinuclear, perigranular and sub-plasmalemmal mitochondria have

126 distinct functions in the regulation of cellular calcium transport." Embo J

20: 1863-74.

Pellistri, F., S. Casagrande, M. Raimondo, A. Cupello and M. Robello. (2005).

"Different chloride electrochemical gradients across the plasma membrane

in subcellular compartments of rat cerebellum granules." Neurosci Lett

381: 139-43.

Perabo, F. G., A. Wirger, S. Kamp, H. Lindner, D. H. Schmidt, S. C. Muller and E.

C. Kohn. (2004). "Carboxyamido-triazole (CAI), a signal transduction

inhibitor induces growth inhibition and apoptosis in bladder cancer cells by

modulation of Bcl-2." Anticancer Res 24: 2869-77.

Petrie, E. C., C. W. Wilkinson, S. Murray, C. Jensen, E. R. Peskind and M. A.

Raskind. (1999). "Effects of Alzheimer's disease and gender on the

hypothalamic-pituitary-adrenal axis response to lumbar puncture stress."

Psychoneuroendocrinology 24: 385-95.

Pfrieger, F. W. and B. A. Barres. (1997). "Synaptic efficacy enhanced by glial

cells in vitro." Science 277: 1684-7.

Pieraut, S., V. Laurent-Matha, C. Sar, T. Hubert, I. Mechaly, C. Hilaire, M.

Mersel, E. Delpire, J. Valmier and F. Scamps. (2007). "NKCC1

127 phosphorylation stimulates neurite growth of injured adult sensory

neurons." J Neurosci 27: 6751-9.

Pierson, R. C., A. M. Lyons and L. J. Greenfield, Jr. (2005). "Gonadal steroids

regulate GABAA receptor subunit mRNA expression in NT2-N neurons."

Brain Res Mol Brain Res 138: 105-15.

Piet, R., D. A. Poulain and S. H. Oliet. (2004). "Contribution of astrocytes to

synaptic transmission in the rat supraoptic nucleus." Neurochem Int 45:

251-7.

Pivovarova, N. B., H. V. Nguyen, C. A. Winters, C. A. Brantner, C. L. Smith and

S. B. Andrews. (2004). "Excitotoxic calcium overload in a subpopulation of

mitochondria triggers delayed death in hippocampal neurons." J Neurosci

24: 5611-22.

Prakriya, M., S. Feske, Y. Gwack, S. Srikanth, A. Rao and P. G. Hogan. (2006).

"Orai1 is an essential pore subunit of the CRAC channel." Nature 443:

230-3.

Prevarskaya, N., L. Zhang and G. Barritt. (2007). "TRP channels in cancer."

Biochim Biophys Acta 1772: 937-46.

128 Pritchett, D. B. and P. H. Seeburg. (1991). "gamma-Aminobutyric acid type A

receptor point mutation increases the affinity of compounds for the

benzodiazepine site." Proc Natl Acad Sci U S A 88: 1421-5.

Putney, J. W., Jr. (2007). "Recent breakthroughs in the molecular mechanism of

capacitative calcium entry (with thoughts on how we got here)." Cell

Calcium 42: 103-10.

Rabbani, M. and H. J. Little. (1999). "Increases in neuronal Ca2+ flux after

withdrawal from chronic treatment." Eur J Pharmacol 364: 221-

7.

Randriamampita, C. and R. Y. Tsien. (1993). "Emptying of intracellular Ca2+

stores releases a novel small messenger that stimulates Ca2+ influx."

Nature 364: 809-14.

Rao, S. D. and J. H. Weiss. (2004). "Excitotoxic and oxidative cross-talk between

motor neurons and glia in ALS pathogenesis." Trends Neurosci 27: 17-23.

Reddy, D. S. (2003). "Pharmacology of endogenous neuroactive steroids." Crit

Rev Neurobiol 15: 197-234.

129 Reddy, D. S. (2004). "Pharmacology of catamenial epilepsy." Methods Find Exp

Clin Pharmacol 26: 547-61.

Reddy, D. S. (2004). "Role of neurosteroids in catamenial epilepsy." Epilepsy

Res 62: 99-118.

Riccio, A. and D. D. Ginty. (2002). "What a privilege to reside at the synapse:

NMDA receptor signaling to CREB." Nat Neurosci 5: 389-90.

Rivera, C., J. Voipio and K. Kaila. (2005). "Two developmental switches in

GABAergic signalling: the K+-Cl- cotransporter KCC2 and carbonic

anhydrase CAVII." J Physiol 562: 27-36.

Rizzuto, R., P. Bernardi and T. Pozzan. (2000). "Mitochondria as all-round

players of the calcium game." J Physiol 529 Pt 1: 37-47.

Roos, J., P. J. DiGregorio, A. V. Yeromin, K. Ohlsen, M. Lioudyno, S. Zhang, O.

Safrina, J. A. Kozak, S. L. Wagner, M. D. Cahalan, G. Velicelebi and K. A.

Stauderman. (2005). "STIM1, an essential and conserved component of

store-operated Ca2+ channel function." J Cell Biol 169: 435-45.

130 Rosso, L., B. Peteri-Brunback, P. Poujeol, N. Hussy and J. M. Mienville. (2004).

"Vasopressin-induced taurine efflux from rat pituicytes: a potential

negative feedback for hormone secretion." J Physiol 554: 731-42.

Scharfman, H. E. and N. J. MacLusky. (2006). "The influence of gonadal

hormones on neuronal excitability, seizures, and epilepsy in the female."

Epilepsia 47: 1423-40.

Selkoe, D. J. (2002). "Alzheimer's disease is a synaptic failure." Science 298:

789-91.

Shirotani, K., S. Tsubuki, N. Iwata, Y. Takaki, W. Harigaya, K. Maruyama, S.

Kiryu-Seo, H. Kiyama, H. Iwata, T. Tomita, T. Iwatsubo and T. C. Saido.

(2001). "Neprilysin degrades both amyloid beta peptides 1-40 and 1-42

most rapidly and efficiently among thiorphan- and phosphoramidon-

sensitive endopeptidases." J Biol Chem 276: 21895-901.

Slot, J. W., G. Garruti, S. Martin, V. Oorschot, G. Posthuma, E. W. Kraegen, R.

Laybutt, G. Thibault and D. E. James. (1997). "Glucose transporter

(GLUT-4) is targeted to secretory granules in rat atrial cardiomyocytes." J

Cell Biol 137: 1243-54.

131 Smith, S. S., H. Shen, Q. H. Gong and X. Zhou. (2007). "Neurosteroid regulation

of GABA(A) receptors: Focus on the alpha4 and delta subunits."

Pharmacol Ther 116: 58-76.

Soboloff, J., M. A. Spassova, X. D. Tang, T. Hewavitharana, W. Xu and D. L. Gill.

(2006). "Orai1 and STIM reconstitute store-operated calcium channel

function." J Biol Chem 281: 20661-5.

Song, H., C. F. Stevens and F. H. Gage. (2002). "Astroglia induce neurogenesis

from adult neural stem cells." Nature 417: 39-44.

Stefanis, L. (2005). "Caspase-dependent and -independent neuronal death: two

distinct pathways to neuronal injury." Neuroscientist 11: 50-62.

Stern, E. A., B. J. Bacskai, G. A. Hickey, F. J. Attenello, J. A. Lombardo and B. T.

Hyman. (2004). "Cortical synaptic integration in vivo is disrupted by

amyloid-beta plaques." J Neurosci 24: 4535-40.

Stowers, R. S., L. J. Megeath, J. Gorska-Andrzejak, I. A. Meinertzhagen and T.

L. Schwarz. (2002). "Axonal transport of mitochondria to synapses

depends on milton, a novel Drosophila protein." Neuron 36: 1063-77.

132 Sudhof, T. C. (1995). "The synaptic vesicle cycle: a cascade of protein-protein

interactions." Nature 375: 645-53.

Takahashi, R. H., C. G. Almeida, P. F. Kearney, F. Yu, M. T. Lin, T. A. Milner and

G. K. Gouras. (2004). "Oligomerization of Alzheimer's beta-amyloid within

processes and synapses of cultured neurons and brain." J Neurosci 24:

3592-9.

Takahashi, R. H., T. A. Milner, F. Li, E. E. Nam, M. A. Edgar, H. Yamaguchi, M.

F. Beal, H. Xu, P. Greengard and G. K. Gouras. (2002). "Intraneuronal

Alzheimer abeta42 accumulates in multivesicular bodies and is associated

with synaptic pathology." Am J Pathol 161: 1869-79.

Takemura, H., A. R. Hughes, O. Thastrup and J. W. Putney, Jr. (1989).

"Activation of calcium entry by the tumor promoter thapsigargin in parotid

acinar cells. Evidence that an intracellular calcium pool and not an inositol

phosphate regulates calcium fluxes at the plasma membrane." J Biol

Chem 264: 12266-71.

Tanabe, T., H. Takeshima, A. Mikami, V. Flockerzi, H. Takahashi, K. Kangawa,

M. Kojima, H. Matsuo, T. Hirose and S. Numa. (1987). "Primary structure

of the receptor for calcium channel blockers from skeletal muscle." Nature

328: 313-8.

133

Thomas, G. M. and R. L. Huganir. (2004). "MAPK cascade signalling and

synaptic plasticity." Nat Rev Neurosci 5: 173-83.

Tilleux, S. and E. Hermans. (2007). "Neuroinflammation and regulation of glial

glutamate uptake in neurological disorders." J Neurosci Res 85: 2059-70.

Touma, C., O. Ambree, N. Gortz, K. Keyvani, L. Lewejohann, R. Palme, W.

Paulus, K. Schwarze-Eicker and N. Sachser. (2004). "Age- and sex-

dependent development of adrenocortical hyperactivity in a transgenic

mouse model of Alzheimer's disease." Neurobiol Aging 25: 893-904.

Troadec, J. D., S. Thirion, D. Petturiti, M. T. Bohn and P. Poujeol. (1999). "ATP

acting on P2Y receptors triggers calcium mobilization in primary cultures

of rat neurohypophysial astrocytes (pituicytes)." Pflugers Arch 437: 745-

53.

Urbanska, E. M., P. Blaszczak, T. Saran, Z. Kleinrok and W. A. Turski. (1999).

"AMPA/kainate-related mechanisms contribute to convulsant and

proconvulsant effects of 3-nitropropionic acid." Eur J Pharmacol 370: 251-

6.

134 Van Damme, P., E. Bogaert, M. Dewil, N. Hersmus, D. Kiraly, W. Scheveneels, I.

Bockx, D. Braeken, N. Verpoorten, K. Verhoeven, V. Timmerman, P.

Herijgers, G. Callewaert, P. Carmeliet, L. Van Den Bosch and W.

Robberecht. (2007). "Astrocytes regulate GluR2 expression in motor

neurons and their vulnerability to excitotoxicity." Proc Natl Acad Sci U S A

104: 14825-30.

Vanden Berghe, T., M. Kalai, G. van Loo, W. Declercq and P. Vandenabeele.

(2003). "Disruption of HSP90 function reverts tumor necrosis factor-

induced necrosis to apoptosis." J Biol Chem 278: 5622-9.

Vaux, D. L., S. Cory and J. M. Adams. (1988). "Bcl-2 gene promotes

haemopoietic cell survival and cooperates with c-myc to immortalize pre-B

cells." Nature 335: 440-2.

Venance, L., N. Stella, J. Glowinski and C. Giaume. (1997). "Mechanism involved

in initiation and propagation of receptor-induced intercellular calcium

signaling in cultured rat astrocytes." J Neurosci 17: 1981-92.

Wafford, K. A., S. A. Thompson, D. Thomas, J. Sikela, A. S. Wilcox and P. J.

Whiting. (1996). "Functional characterization of human gamma-

aminobutyric acidA receptors containing the alpha 4 subunit." Mol

Pharmacol 50: 670-8.

135

Walch-Solimena, C., K. Takei, K. L. Marek, K. Midyett, T. C. Sudhof, P. De

Camilli and R. Jahn. (1993). "Synaptotagmin: a membrane constituent of

neuropeptide-containing large dense-core vesicles." J Neurosci 13: 3895-

903.

Waldron, R. T., A. D. Short, J. J. Meadows, T. K. Ghosh and D. L. Gill. (1994).

"Endoplasmic reticulum calcium pump expression and control of cell

growth." J Biol Chem 269: 11927-33.

Wang, D. S., N. Iwata, E. Hama, T. C. Saido and D. W. Dickson. (2003).

"Oxidized neprilysin in aging and Alzheimer's disease brains." Biochem

Biophys Res Commun 310: 236-41.

Wang, G. J., J. G. Jackson and S. A. Thayer. (2003). "Altered distribution of

mitochondria impairs calcium homeostasis in rat hippocampal neurons in

culture." J Neurochem 87: 85-94.

Watts, R. J., O. Schuldiner, J. Perrino, C. Larsen and L. Luo. (2004). "Glia engulf

degenerating axons during developmental axon pruning." Curr Biol 14:

678-84.

136 Weeber, E. J., M. Levy, M. J. Sampson, K. Anflous, D. L. Armstrong, S. E.

Brown, J. D. Sweatt and W. J. Craigen. (2002). "The role of mitochondrial

porins and the permeability transition pore in learning and synaptic

plasticity." J Biol Chem 277: 18891-7.

Whittington, M. A. and H. J. Little. (1993). "Changes in voltage-operated calcium

channels modify ethanol withdrawal hyperexcitability in mouse

hippocampal slices." Exp Physiol 78: 347-70.

Yasojima, K., H. Akiyama, E. G. McGeer and P. L. McGeer. (2001). "Reduced

neprilysin in high plaque areas of Alzheimer brain: a possible relationship

to deficient degradation of beta-amyloid peptide." Neurosci Lett 297: 97-

100.

Yeromin, A. V., S. L. Zhang, W. Jiang, Y. Yu, O. Safrina and M. D. Cahalan.

(2006). "Molecular identification of the CRAC channel by altered ion

selectivity in a mutant of Orai." Nature 443: 226-9.

Zhang, S. L., A. V. Yeromin, X. H. Zhang, Y. Yu, O. Safrina, A. Penna, J. Roos,

K. A. Stauderman and M. D. Cahalan. (2006). "Genome-wide RNAi screen

of Ca(2+) influx identifies genes that regulate Ca(2+) release-activated

Ca(2+) channel activity." Proc Natl Acad Sci U S A 103: 9357-62.

137 Zhao, W., W. Xie, W. Le, D. R. Beers, Y. He, J. S. Henkel, E. P. Simpson, A. A.

Yen, Q. Xiao and S. H. Appel. (2004). "Activated microglia initiate motor

neuron injury by a nitric oxide and glutamate-mediated mechanism." J

Neuropathol Exp Neurol 63: 964-77.

138 Abstract

Calcium signals are integral steps in pathways for a large variety of important

cellular functions ranging from cell growth and proliferation to apoptotic cell

death. The mechanism and subcellular location of calcium entry or rise in

cytosolic calcium is also important in determining what type of message the

calcium is activating. Additionally, the control of calcium levels and signaling

pathways is being investigated in treatments several types of diseases. Another

area of investigation in importance in studying calcium signaling is mitochondria.

Mitochondria are vital organelles with essential roles including the production of

ATP, regulation of calcium signals and are the main intracellular initiator of

apoptosis. Monitoring mitochondrial calcium and membrane potential can provide clues in overall cell health and mechanism of cytosolic calcium entry. In this

project I have used both imaging techniques and high through put assays to

monitor calcium and mitochondrial membrane potential in two models of disease

states and to investigate the mechanism of a possible treatment.

139