altered expression of cellular markers in molecular brain aging · 2017-03-07 · brain and their...

75
Altered Expression of Cellular Markers in Molecular Brain Aging by Anulika Nwakaeze A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Pharmacology and Toxicology University of Toronto © Copyright by Anulika Nwakaeze 2016

Upload: others

Post on 26-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Altered Expression of Cellular Markers in Molecular Brain Aging

by

Anulika Nwakaeze

A thesis submitted in conformity with the requirements for the degree of Master of Science

Graduate Department of Pharmacology and Toxicology

University of Toronto

© Copyright by Anulika Nwakaeze 2016

 

ii  

 

Altered Expression of Cellular Markers in Molecular Brain Aging

Anulika Nwakaeze

Master of Science

Graduate Department of Pharmacology and Toxicology

University of Toronto

2016 ABSTRACT The molecular aging of the human brain encompasses pervasive transcriptome changes

associated with “normal” brain aging that occur in local cortical circuits comprised of GABA

neurons, pyramidal cells, and astrocytes. Cortical samples obtained from a novel postmortem

human cohort were analyzed using RT-qPCR technology to assess changes in the expression of

nine markers representative of neuropeptide signalling, synaptic function, calcium regulation,

and glial activation within cortical cellular networks that occur in congruence with a pervasive

molecular program mediating brain aging. Presynaptic interneuron markers displayed decreases

in expression in aged subjects compared to young controls, whilst postsynaptic interneuron

marker expression was increased, as was the expression of glial markers. Although changes in

extrasynaptic glutamate marker expression were not significantly different, there was a nominal

decrease in their expression in aged subjects. Altogether, these results replicate findings of

pervasive molecular changes in the aged cortex, and suggest consequences for cortical network

function.

iii  

 

ACKNOWLEDGEMENTS

I would like to take this opportunity to express my sincere gratitude to my supervisor, Dr.

Etienne Sibille, for accepting me into his prestigious lab and for providing me with counselling

and guidance throughout my degree. Additionally, I am grateful to my advisor, Dr. Peter

McPherson, for his understanding and patience with me during difficult moments of my graduate

studies, and to my committee members for their valuable insights into my project.

I would also like to extend my appreciation to my colleagues in the Sibille lab, particularly to

Drs. Brad Rocco, Hyunjung Oh, and Yuliya Nikolova, who mentored me in this exciting project.

Their knowledge and expertise in molecular biological and data analytical techniques were

instrumental to my success in the program.

Finally, I am endlessly grateful to my parents for their unwavering support, love, and

encouragement.

iv  

 

TABLE OF CONTENTS Abstract ........................................................................................................................................... II

Acknowledgements ....................................................................................................................... III

List of Tables ................................................................................................................................ VI

List of Figures .............................................................................................................................. VII

List of Abbreviations .................................................................................................................. VIII

List of Appendices ........................................................................................................................ IX

Chapter 1: Introduction and Background ........................................................................................ 1

Research objectives .................................................................................................................... 2

Literature review ........................................................................................................................ 2

An introduction to aging ..................................................................................................... 2

Normal brain aging ............................................................................................................. 3

Definition of normal brain aging ............................................................................ 3

Functional changes in normal brain aging .............................................................. 4

Molecular changes in normal brain aging ............................................................... 6

Altered neuropeptide gene expression in GABAergic neurons ................ 10

Altered intracellular calcium signalling in local cortical circuits ............. 12

Altered GABAergic and glutamatergic transmission across synapses ..... 14

Altered activity of glial cells ..................................................................... 16

Molecular aging of the prefrontal cortex .......................................................................... 18

Age-by-disease interaction hypothesis ............................................................................. 18

Rationale ................................................................................................................................... 20

Hypothesis ................................................................................................................................ 20

v  

   

Predictions ................................................................................................................................ 20

Research tools ........................................................................................................................... 21

Chapter 2: Materials and Methods ................................................................................................ 22

Summary of research methods ................................................................................................. 23

RT-qPCR experimental protocol ............................................................................................... 24

Summary of RT-qPCR protocol ....................................................................................... 24

Sample collection ............................................................................................................. 25

RNA isolation and conversion to cDNA .......................................................................... 28

Primer design, preparation, and testing ............................................................................. 30

RT-qPCR program and assay ............................................................................................ 31

Fold change calculation ................................................................................................... 32

Statistical analyses ............................................................................................................ 33

Chapter 4: Results ......................................................................................................................... 34

The majority of gene expression changes are age-dependent .................................................. 35

Altered GABAergic marker expression with age ..................................................................... 35

Increased glial marker expression with age .............................................................................. 36

Unchanged glutamatergic marker expression with age ............................................................. 36

Principle component of variance in gene expression is captured by age .................................. 40

Gene expression changes in circuitry are concerted ................................................................. 41

Cell markers group together irrespective of age ....................................................................... 41

Chapter 5: Discussion ................................................................................................................... 45

Summary ................................................................................................................................... 46

Role of cell markers in GABA-mediated local circuit function ................................................ 46

vi  

 

Role of observed changes in aging ............................................................................................ 47

Age-by-disease interaction model ............................................................................................. 49

Study limitations ...................................................................................................................... 52

Conclusions ............................................................................................................................... 53

Appendices .................................................................................................................................... 54

RNA extraction data .................................................................................................................. 54

Standardized cDNA volumes .................................................................................................... 54

Primer sequences ....................................................................................................................... 55

UNIANOVA analysis of variance ............................................................................................. 55

Correlation matrix of gene expression ...................................................................................... 56

References ..................................................................................................................................... 57

vii  

 

LIST OF TABLES

Description Page no. Table 1:sample demographics 27

Table 2:cDNA synthesis reagent volumes

Table 3: cDNA synthesis program

28

29

Table 4:RT-qPCR program 31

viii  

 

LIST OF FIGURES

Description Page no. Figure 1: select changes in cortical tripartite synapses with age

Figure 2: synaptic interactions in local cortical circuit

Figure 3: demographic characteristics of age-cohort

8

11

27

Figure 4: gene expression changes in aged vs. comparison subjects 37

Figure 5: altered GABAergic marker expression with age

Figure 6: increased glial marker expression with age

Figure 7: unchanged glutamatergic marker expression with age

Figure 8: principle component analysis of cell marker gene expression

Figure 9: gene expression changes in circuitry are concerted

Figure 10: cell markers group together irrespective of age

38

39

40

42

43

44

ix  

 

LIST OF ABBREVIATIONS

Abbreviation Description AC

ACTB

adenylate cyclase

beta-actin

AD

ALDH1L1

BMI

alzheimer’s disease

aldehyde dehydrogenase 1 family member L1

body mass index

BBB

BPD

CALB1

cAMP

CBP

CNS

blood brain barrier

bipolar disorder

calbindin

cyclic AMP

calcium binding protein

central nervous system

CYCLO

EAAT2

fMRI

GABA

GAPDH

GFAP

GLUL

GPCR

GPHN

GRM2

cyclophilin g

excitatory amino acid transporter 2

functional magnetic resonance imaging

gamma-amino butyric acid

glyceraldehyde-3-phosphate dehydrogenase

glial fibrillary acidic protein

glutamine synthase

g protein-coupled receptor

gephyrin

glutamate receptor metabotropic 2

x  

   

GRM3

MDD

mRNA

PD

PET

PFC

PVALB

RT-qPCR

SCZ

SST

glutamate receptor metabotropic 3

major depressive disorder

messenger ribonucleic acid

parkinson’s disease

positron emission tomography

prefrontal cortex

parvalbumin

reverse-transcription quantitative polymerase chain reaction

schizophrenia

somatostatin

xi  

 

LIST OF APPENDICES

Content Page no. Appendix 1:RNA extraction data

Appendix 2: standardized cDNA volumes

53

53

Appendix 3:primer sequences 54

Appendix 4:UNIANOVA analysis of variance

Appendix 5: correlation matrix of gene expression

54

55

1  

 

CHAPTER 1: INTRODUCTION AND BACKGROUND

2  

 

RESEARCH OBJECTIVES

The goal of this thesis is to assess a relationship between gene expression changes of cellular

markers in the orbitoventral prefrontal cortex across aging:

Aim I: Examine age-dependent changes in the gene expression profiles of cortical

neurons (pyramidal cells and GABAergic interneurons) and glia (astrocytes).

Aim II: Examine a correlation between advancing age and the accumulation of gene

expression changes.

Aim III: Examine a concerted movement of gene expression changes that occur in

interneurons, pyramidal cells, and astrocytes.

LITERATURE REVIEW

An Introduction to Aging

Aging is broadly defined as the time-dependent functional decline that affects most living

organisms1. This deterioration is the predominant risk factor for developing human

pathologies, including neurological disease 1.The number of older persons, i.e. those aged 60

years or over, has substantially increased globally, and that growth is projected to accelerate.

This demographic is projected to grow by 56% by 2030 and to double in size by 2050,

reaching nearly 2.1 billion 2.The number of people aged 80 years or over, defined as the

“oldest-old”, is showing the fastest acceleration in growth2. The global proportion of people

aged 80 years or over is expected to rise to more than 20% in 20502. The number of older

persons is growing faster than any other age demographic worldwide 2.

3  

 

The aging process is most advanced in urban areas and in high-income countries 2. In Canada,

the number of persons aged 65 and over outnumbered children under 15 years for the first time

in the country’s history 3. 16.1% of Canadians were aged 65 years and over 3. The share of

Canadians in this demographic is projected to account for 20.1% of the national population by

2024 3.

The global aging phenomenon is largely attributed to improvements in medical care and health

practices4. However, with the increase in longevity comes the higher risk of developing

diseases associated with aging, such as neurodegenerative and neuropsychiatric disorders5.

Thus, it is essential to extensively characterize the process of brain aging, beginning with

elucidating the molecular mechanisms and genetic underpinnings of age-related changes in the

brain and their overlap with neurological diseases.

Normal Brain Aging

Definition of Normal Brain Aging

Normal brain aging refers to the characteristic molecular changes of the central nervous system

(CNS) that occur with age in the absence of clinically diagnosed neurodegenerative and

neuropsychiatric pathology6. The aging process is associated with patterns of molecular

changes which mediate widespread effects in cognitive, motor, and mood systems that

typically occur with age6.

Normal aging of the brain is a largely understudied area of research6.Brain aging is associated

with progressive transcriptome changes regulated by molecular signals within neurons and

4  

 

glial cells, which underlie age-related cognitive, motor, and mood changes and susceptibility to

neurological disease 6. Studies have revealed that a conserved molecular brain aging program

may contribute to functional declines along the lifespan and to gating of age-related diseases,

and is thus regarded as a framework for scientific investigation of the aging brain, and for

clinical extensions of the model.

Functional Changes in Normal Brain Aging

Normal declines in neurological function in the absence of associated diseases occur with age

and are well documented, as well as proceed in a conserved manner. Data obtained from meta-

analytical studies have highlighted significant declines in cognitive speed on the order of ~40-

60% in individuals aged 80 compared to age 20, without dementia 7. Moreover, cognitive

processes are differentially affected by aging, such that certain abilities exhibit more

pronounced declines in function compared to others7. So-called “fluid abilities” reliant on

processing speed, problem solving, inhibitory function, working and long-term memory, as

well as spatial ability, decline with age7. Conversely, “crystallized abilities” reliant on

knowledge or expertise, including world and general knowledge, implicit memory, vocabulary,

and occupational expertise have not been shown to decline, and may even improve, with age7.

Thus, the effects of the aging process on cognitive abilities have been reliably shown to affect

brain function in a distinctive manner, and in the absence of neurological disease, such that

these effects are defined as “normal”. It is important to note that functional changes in

cognitive ability may be explained by a loss of grey matter volume and molecular disturbances;

loss of dendritic arborization and transcriptomic changes in gene expression have consistently

been reported to be associated with the aging phenotype8,9.

5  

 

In addition to cognitive indices, aging also compromises motor ability. Reaction time, speed of

movement, as well as hand and foot coordination, consistently decline with age10. This

progressive and continuous phenomenon is most likely mediated by a reduction in

dopaminergic receptor (D2) levels in the dorsal striatum of the basal ganglia and the frontal

cortex, which are correlated with both motor speed and cognitive function, respectively11.

Lastly, meta-analytical studies demonstrate that mood and affective perception are also altered

with age, in addition to cognitive and motor parametres. However, these changes are generally

more positively skewed in contrast to declining cognitive and motor function with age. Older

adults perform equally as well as younger adults in recognizing happy emotions, and are

proven to be better at recognizing disgust; although, they are worse at sad and angry emotional

recognition12. Additionally, older adults demonstrate a positivity bias in attention and memory,

recover more quickly from negative events, and are less likely to engage in destructive

interpersonal behaviours13. Thus, mood and affective perception are additional functional

indices that are altered by aging.

An important caveat to these documented changes in cognitive, motor, and mood/affective

processing is the large individual variability in the rates of age-related functional decline14. For

instance, cognitive capacity, i.e., the processing of mental information, in youth accounts for

~50% of the variance of cognitive capacity in old age15,16. Lifestyle is also a robust predictor of

cognitive ability along the lifespan. Exercise, body mass index (BMI), balanced nutrition,

higher education, cardiovascular integrity, caloric restriction, and cigarette avoidance

positively correlate with slower cognitive decline with age6.

6  

 

Molecular Changes in Normal Brain Aging

Cellular networks comprised of interneurons, pyramidal cells, and astrocytes mediate the

transmission of brain activity in the cortex17.

Cortical interneurons are responsible for mediating GABAergic neurotransmission and possess

a varied range of molecular, structural, and electrophysiological properties18.They are

delineated into distinct subtypes that differ in their distribution across the cortical layers, as

well as in the cellular compartment with which they exert their effects on pyramidal cells. y-

aminobutyric acid (GABA) is synthesized in presynaptic terminals by glutamic acid

decarboxylase (GAD), using the neurotransmitter glutamate as a precursor19. It is packaged

into synaptic vesicles and is released in an activity-dependent manner when action potentials

depolarize the synaptic terminal, causing an influx of calcium ions and subsequent exocytosis

19. During activity-dependent synaptic release, GABA acts at chloride-permeant synaptic

receptors, as well as at extrasynaptic GABAergic receptor sites 19.Extrasynaptic ionotropic

GABAARs and metabotropic GABABRs mediate tonic inhibition19. Accumulating data

implicate tonic inhibition at GABAergic synapses as a particularly robust regulator of neuronal

activity and network dynamics in cortical circuits 19. Recent evidence highlights changes in

tonic inhibitory tone in the aged prefrontal cortex (PFC) 19. Deficits in GABAergic regulation

of excitatory cell input and output may underlie changes in cellular processing that contribute

to aging19.

In contrast to GABA, the predominant excitatory neurotransmitter in the mammalian central

nervous system is glutamate. Adequate supplies of glutamate are maintained in the CNS via the

glutamate-GABA-glutamine biochemical cycle 20. A portion of glutamate and GABA released

7  

 

from neurons is taken up by astrocytes, which in turn release glutamine to be taken up by

neurons for use as a precursor for glutamate or GABA synthesis 20. Pyramidal cells utilize

glutamate to transmute excitatory signals across cortical layers and brain regions 21.

Astrocytes are integral components of proper cortical circuit function 22. They were originally

characterized solely as cells that provided nutritional and structural support for electrically

active neurons 22. However, it was later discovered that astrocytes also possessed intrinsic

electrical activity and were capable of releasing “gliotransmitters”, which mediated

intercellular communication via gap junctions and regulated the excitability of neurons 23,24,26.

Astrocytes are responsive to a variety of neurotransmitters released by neuronal cells and in

turn modulate the release of gliotransmitters22,23. The feedback loop between astrocytes and

neurons influences the functioning of the cortical microenvironment 23,26. They are capable of

releasing glutamate, are actively involved in its uptake, and are also responsive to calcium

transmission 23,27. Astrocytes supply the metabolic needs of neurons via contacts with blood

vessels, through which they play a vital role in the formation of the blood brain barrier (BBB)

23,28. Lastly, astrocytes secrete an array of neurotrophins, including BDNF, which are crucial

for neuronal survival, regeneration, and differentiation 23,29.

8  

 

Genome-wide studies have identified approximately 5-10% of genes expressed in the human

brain that are regulated by age30. Data generated from gene microarray technology in human

prefrontal cortex samples identified 7.5% of tested genes exhibited age-dependent regulation31.

This set of age-dependent genes displays a high degree of conservation across cohorts and

Figure 1. Select alterations in molecular signature of interneurons, pyramidal cells and astrocytes that occur age. Synaptic interactions and signalling between interneurons, pyramidal cells, and astrocytes are altered during the aging process. Typical biological changes associated with aging are highlighted. Purple, interneuron; Blue, pyramidal cell; Orange, astrocyte. SST, somatostatin; CALB-1, calbindin; GFAP, glial fibrillary acidic protein. Adapted from Sibille, 2013. Copyright © Pergamon Press 2011.

9  

 

cortical brain regions, regardless of different sample sizes, expression platforms, and analytical

methods30–35. The consistency and specificity of age effects on gene expression fulfill criteria

for use as biomarkers of brain aging 36. Accordingly, the predicted age of an individual is

highly correlated with chronological age, as evidenced by regression analyses of age-

dependent gene expression across the lifespan 30,31. Thus, “molecular age” can be used as a

functional assay to measure biological aging of the brain as well as to assess individual

deviations from expected trajectories 30,31.

Altogether, the conserved and restricted scope of transcript changes implies that specific

cellular populations and biological processes are selectively affected during aging 5.

Expression of genes that mediate glial inflammation, synaptic function and plasticity, calcium

regulation, and neuropeptide signalling, have consistently been shown to be affected by the

aging process across multiple studies 1,5,6,23,31,32,37,38. Overall, genes upregulated by age are

largely of glial origin, and are related to inflammation and cellular defenses, whilst

downregulated genes are mostly enriched in neuronal transcripts relating to cellular

communication and signalling (Figure 1) 30. Changes in the expression of age-regulated genes

involved in these biological processes are integral components of normal brain aging. As such,

cellular markers of neuropeptide signalling (somatostatin, SST), intracellular calcium

regulation (parvalbumin, PVALB; calbindin, CALB1), synaptic function and plasticity

(gephyrin, GPHN; metabotropic glutamate receptors 2/3, GRM2/3) and astrocyte reactivity

(glial fibrillary acidic protein, GFAP; aldehyde dehydrogenase 1 family member L1,

ALDH1L1; glutamine synthetase, GLUL), were chosen to assess the molecular correlates of

aging in a novel human postmortem cohort.

10  

 

Altered neuropeptide gene expression in inhibitory GABAergic neurons

Neuropeptides are a large class of messenger molecules that are responsible for modulating

neuronal activity and intercellular communication 39. Neuropeptides are packaged separately

from neurotransmitters and are released by different mechanisms, including from nerve

endings, soma, and dendrites40–42. Following synaptic and extrasynaptic release, neuropeptides

exert their effects via G protein-coupled receptors (GPCRs) 39. Neuropeptides are actively

degraded by extracellular peptidases soon after being released, thereby limiting their biological

effects39. Neuropeptides differ in their regional expression, mechanism and range of action, and

excitatory/inhibitory effects on target neurons39–41. Neuropeptides transfer molecular messages

between populations of neurons and can evoke changes in network function40,41.

Somatostatin (SST)

SST is an inhibitory neuromodulatory peptide43. It is expressed in a subtype of

GABAergic neurons 43.

SST-expressing interneurons inhibit the dendritic compartments of principal excitatory

glutamatergic, i.e., pyramidal, cells, in a manner that is inversely proportional to their

intervening distance 40,41. SST cells represent ~18-20% of interneurons in the cortex,

with differential expression across cortical layers such that layer 2/3/5 SST neurons,

which send abundant projections to layer 1, inhibit pyramidal cells, whilst layer 4 SST

neurons, which do not project to layer 1, disinhibit excitatory cells 43. They are

recruited in a feedforward mechanism by activated pyramidal cells, for which they also

provide feedback inhibition42. SST+ interneurons are characterized by delayed,

sustained, and adapting electrophysiological properties 42. SST+ interneurons also

11  

 

display low firing synchrony due to the absence of reciprocal inhibition between SST

cells (Figure 2) 43. Thus, SST neurons are specialized in customized and targeted

regulation of incoming local excitatory signals and may also provide significant

inhibition to most other GABA neuron subtypes, thereby regulating overall inhibitory

tone within local cortical circuits 44.

The expression of SST has been reported to decrease with age in the amygdala, in the

dorsolateral and orbitoventral prefrontal cortices, as well as in the subgenual anterior

cingulate cortex30,49.This decline in expression appears to be due to a downregulation in

transcript levels rather than to a loss of SST-positive cells43.

In summary, the decline in the expression of SST associated with aging suggests changes to the

overall inhibitory tone within cortical circuits and subsequent alterations in circuit activity and

pyramidal cell output.

Figure 2. Synaptic interactions between SST+ and PV+ interneurons with pyramidal cells in local cortical circuit. Pyramidal cells are inhibited in the perisomatic region by PV cells and in the dendritic compartment by SST cells (not shown). PV cells display reciprocal inhibition. PYR, pyramidal cell; SST, somatostatin; PV, parvalbumin. Adapted from French et al., 2014. Copyright © Springer New York 2014.

Cortical  Layers  

12  

 

Altered intracellular calcium signalling in local cortical circuits

Calcium signalling is central to the function of neurons, mediating neurotransmitter and

neuropeptide release, neuronal membrane excitability, gene induction, neuronal plasticity and

growth, as well as cellular processes that underlie learning and memory 50,51. Deregulation of

calcium signalling has been suggested to play a role in brain aging as a result of an increase in

resting calcium ion concentrations and a substantial prolongation of stimulus-evoked calcium

signals 38,50,52. Calcium-binding peptides (CBPs) regulate the cytosolic concentration of

calcium ions and can function as calcium transporters or sensors53. Observations in the aging

human central nervous system imply that an increase in intracellular calcium is associated with

the phenotype 38,50.

Parvalbumin (PVALB)

PVALB is a calcium-binding peptide 21. It is expressed in a subset of GABAergic

interneurons that nonspecifically target the perisomatic compartment of pyramidal cells

in a manner dependent on the distance between the excitatory and inhibitory cell 40.

Neurons expressing PVALB are directly activated by thalamic and corticocortical

projections 44. They are characterized by fast-spiking and non-adapting

electrophysiological properties 21. In contrast to SST interneurons, PVALB+ cells are

highly synchronized through dense reciprocal inhibition (Figure 2) 49,50. Thus, PVALB-

expressing interneurons are specialized in regulating the output and synchronization of

pyramidal cells 49,50.

13  

 

The association of PVALB expression with age is contested. Some studies indicate no

significant changes in the neuropeptide expression in aged subjects compared to young

controls, whilst others report an increase or a decrease in its expression in the

experimental group 50,56–58. These findings also appear to differ depending on the

cortical area examined. Notwithstanding, PVALB expression has been shown to

significantly decline in the ACC and in the orbitoventral PFC31.

Calbindin (CALB1)

CALB1 is a high-affinity cytosolic calcium binding protein (CBP)50,58. It is a member

of the calcium-binding superfamily that includes calmodulin and troponin C. It is

expressed in a subset of interneurons that overlap with SST expression, as well as in

pyramidal cells59–61.

CALB1 buffers the entry of calcium ions following glutamate receptor stimulation,

thereby regulating intracellular calcium concentrations in neurons 50,58. CALB1 also

protects neurons from cellular insults that induce spikes in intracellular calcium ions 50.

CALB1 expression decreases across cortical areas with age 50. These include the visual

association cortex, primary visual cortex, posterior cingulate cortex, hippocampus, and

striatum 50,58.

Altogether, the changes in calcium-binding protein expression observed with PVALB and

CALB1 may pose serious consequences for unbound intracellular calcium concentrations,

14  

 

thereby affecting aspects of neuronal function, such as neuropeptide release and gene

expression.

Altered inhibitory GABAergic and excitatory glutamatergic transmission across synapses

The appropriate transfer of information within the brain requires a homeostatic balance

between excitatory and inhibitory transmission across synapses62. Changes in the biological

components of excitatory and inhibitory neurotransmission are hypothesized to contribute to

normal aging 5.

Gephyrin (GPHN)

GPHN encodes a neuronal assembly protein 63. It is expressed in postsynaptic

inhibitory membranes where it anchors inhibitory GABAA neurotransmitter receptors to

the postsynaptic cytoskeleton via a high-affinity binding subunit and tubulin dimers 64.

GPHN expression levels provide an indication of the total amount of GABAA receptors

present at a given time 65. GPHN expression is robust in the cortex, hippocampus, and

dorsal striatum (caudate-putamen) 66. GPHN is localized to the soma and dendrites of

pyramidal cells.

GPHN has been reported to show significant age-related increases in transcript

expression in several brain regions, including the dorsal cochlear nucleus (DCN) and

the visual cortex 67,68.

15  

 

Metabotropic glutamate receptor 2/3 (GRM2/3)

GRM2/3 are a family of G protein-coupled receptors (GPCRs) in the metabotropic

glutamate receptor family. Metabotropic glutamate receptors are classified into three

subfamilies based on sequence homology, signal transduction mechanisms, and

pharmacologic properties69. GRM2/3 are classified into Group II of this superfamily69.

GRM2/3 inhibit cyclic AMP (cAMP) signal transduction and adenylate cyclase (AC)

activity69. GRM2/3 are expressed in pyramidal cells and in astroglia, where they

regulate the expression of the glial glutamate transporter, excitatory amino acid

transporter 2 (EAAT2), which is essential in clearing synaptic glutamate 70,71. EAAT2-

mediated glutamate reuptake is a critical component in regulating glutamate

neurotransmission 71,72.

GRM2/3 gene expression is associated with prefrontal physiology and cognition71.

Expression is also related to glial function and neuroprotection 73.

Studies report a decline of GRM2/3 expression in the prefrontal cortex across

adulthood 73,74.

Altogether, changes in the balance of excitatory and inhibitory synaptic transmission is

associated with the normal aging process and may be affected by alterations to GABAergic and

glutamatergic signalling within cellular networks.

16  

 

Altered activity of glial cells

Astrocytes are essential contributors to the microenvironment of neuronal networks75. They are

capable of responding to neuronal health and activity and release an assortment of pro-

inflammatory cyto- and chemokines in response to cellular insult and injury 75,76. Inflammation

is a prominent feature of the aging process, which may be mediated in part by exaggerated glial

responses.

Glial fibrillary acidic protein (GFAP)

GFAP encodes an intermediate filament in astrocytes77. It is used as a marker to

distinguish astrocytes from other types of glial cells.

GFAP contributes to the modulation of astrocyte motility and shape by conferring

structural stability to astrocytic processes 23. Overexpression of GFAP results in fatal

encephalopathy in transgenic mice 23,78.

Elevated levels of GFAP during molecular brain aging lead to glial dystrophy and

inflammation48.

Aldehyde dehydrogenase 1L1 (ALDH1L1)

ALDH1L1 belongs to the aldehyde dehydrogenase family of genes79. It is expressed in

astrocytes79. It catalyzes the conversion of 10-formyltetrahydrofolate, NADP+, and

water to tetrahydrofolate, NADPH, and carbon dioxide for use in downstream

biochemical reactions79.

17  

 

Unique functions of ALDH1L1 in the central nervous system are infrequently reported;

however, it may be linked to neural tube defects during early CNS development 79,80.

ALDH1L1 may also be responsible for regulating cell division in astrocytes 79.

Alterations of ALDH1L1 in reactive astrocytes in association with neural injury or

neurodegenerative diseases are also understudied, which is also true of molecular brain

aging research 79.

Glutamine synthetase (GLUL)

GLUL belongs to the glutamine synthetase family of genes81. It is expressed in

astrocytes81. It catalyzes the conversion of glutamate and ammonia to glutamine in an

ATP-dependent reaction for downstream biochemical reactions81. GLUL also catalyzes

the production of GABA from L-glutamate81.

Upregulation of GLUL is an indicator of astrogliosis, that is, morphological and

functional changes in astrocytes, including hypertrophy and increased proliferation 23,82.

Astrogliosis occurs with aging 23.

Thus, changes in the expression of astrocytic markers may be suggestive of an increase in

cellular reactivity and a subsequent rise in pro-inflammatory mediators that are associated with

molecular brain aging.

18  

 

Molecular Aging of the Prefrontal Cortex

The aging process affects brain regions in a selective manner. Executive functions, such as

working memory, planning, and goal directed behaviour, most often exhibit age-related

cognitive declines before others 83. The prefrontal cortex (PFC) is primarily responsible for

effecting these complex executive functions and has thus been postulated to be particularly

vulnerable to normal aging processes; it appears that neurons and circuits of the PFC become

selectively altered across the lifespan 8,83. Functional magnetic resonance imaging (fMRI) and

positron emission tomography (PET) studies suggest that aged adults typically exhibit reduced

activation of the PFC compared to young adults84–86. Additionally, aged adults activate a

broader area of the PFC when performing executive tasks, and also activate the contralateral

hemisphere, thereby indicating a loss of hemispheric asymmetry84–86.Thus, the orbitoventral

PFC was selected for the purpose of this study because of its association with aging processes

and neurological diseases.

Age-by-Disease Interaction Hypothesis

Normal brain aging shares mechanistic similarities with disease pathways that promote

neurodegenerative and neuropsychiatric diseases 87. Age-related gene expression changes

overlap both in terms of affected genes and in direction of change, i.e. up- or downregulation,

with those observed in neurological disorders 30,31. For example, SST expression progressively

decreases in normal brain aging and is also downregulated in individuals with major depressive

disorder (MDD) 49. Furthermore, molecular age deviates from chronological age in individuals

with neurodegenerative and neuropsychiatric disorders30,31. Although normal brain aging has

been associated with positive effects in certain functional indices, such as improvements in

19  

 

linguistic ability and emotional control, it also imposes a robust drive towards the onset and

development of many neurological diseases, including Alzheimer’s (AD) and Parkinson’s (PD)

diseases, as well as major depressive (MDD) and bipolar (BPD) disorders, and schizophrenia

(SCZ)6,88.

A framework has been suggested in which molecular aging of the brain promotes biological

changes that in turn places the system at a higher susceptibility to developing age-related

neurological diseases6,37,89.

The mechanisms underlying normal aging process and neurological disorders associated with

age are still being characterized87. The similar trends in gene expression changes common to

aging and neurological diseases suggest that the aging process elicits global transcriptome

changes that confers molecular susceptibility to developing neuropsychiatric disease, as

proposed by an age-by-disease interaction model36,89.

20  

 

RATIONALE

A molecular program that alters the transcriptomic profile of local cortical circuits is an

emergent theory of normal brain aging. To further validate this theory, it is prudent to replicate

findings of altered cellular marker expression in the aged cortex by analyzing a novel cohort

using RT-qPCR technology.

HYPOTHESIS

Gene expression changes observed in subjects from a novel age-cohort will exhibit consistency

with previous reports due to a pervasive molecular program that mediates the aging process in

the human prefrontal cortex.

PREDICTIONS

Cellular markers of local cortical circuits are differentially affected during aging. Expression

forGABA neuron markers SST, PVALB, and CALB1 are predicted to be lower in the aged

subjects, as well as the glutamate signalling markers GRM2 and GRM3. Conversely, astrocyte

markers GFAP, ALDH1L1, and GLUL are predicted to be higher in the aged subjects.

Additionally, postsynaptic clustering peptide marker GPHN is predicted to be upregulated in

aged subjects. These expected changes are predicted based on predominant theories of altered

intracellular calcium signalling and synaptic function, reduced inhibitory tone, and increased

glial reactivity in association with aging as a result of an intrinsic molecular program that

elicits widespread effects in cortical gene expression.

21  

 

RESEARCH TOOLS

Validating altered gene expression in a novel age-cohort using RT-qPCR

To study systematic differences in the gene expression profiles of aged postmortem subjects in

comparison to younger controls, reverse transcription quantitative polymerase chain reaction

(RT-qPCR) is a useful and powerful tool because it facilitates the quantification of nucleic acid

expression in biological samples90–92.

RT-qPCR is a molecular biology technique used to quantitatively monitor the amplification of

targeted nucleic acids in real time after each amplification cycle. Real-time detection of PCR

products is enabled by the introduction of a fluorescent reporter dye within reaction vessels,

which increases its intensity of fluorescence in proportion to increasing amounts of product

DNA.

22  

 

CHAPTER 2: MATERIALS AND METHODS

23  

 

SUMMARY OF RESEARCH METHODS

Ninemarkers representing neuropeptide signalling, synaptic function, calcium regulation, and

astrocyte activation in GABA neurons, pyramidal cells, and astrocyteswere selected based on

literature evidence of age-regulated expression for RT-qPCR analysis. The corresponding

genes are SST (somatostatin), parvalbumin (PVALB), calbindin (CALB1), GPHN (gephyrin),

GRM2 (glutamate receptor metabotropic 2), GRM3 (glutamate receptor, metabotropic 3),

GFAP (glial fibrillary acidic protein), ALDH1L1 (aldehyde dehydrogenase 1 family member

L1), and GLUL (glutamine synthase). Thirty samples of human grey matter from prefrontal

cortex tissue of Brodmann areas 11 and 47 (BA11 and BA47; orbitoventral prefrontal cortex)

that were obtained from the McGill University brain bank at the Douglas Institute were

processed into 100 µm thick sections using the Leica CM1950 cryostat system. Two grey

matter sections of 100 µm thickness were immersed in 350 µL of Buffer RLT (Qiagen

miRNeasy) for homogenization. Total RNA was extracted using the Qiagen RNeasy® Micro

protocol. Concentration and purity of RNA were determined by absorbance ratios at 260, 280,

and 230 nm using the Implen P360 NanoPhotometer. RNA integrity values were generated

using the Agilent 2100 BioAnalyzer microfluidics system. Subsequently, cDNA was

synthesized from 1 µg of total RNA using the Superscript VILOTM kit and gene expression was

quantified by real time polymerase chain reaction using the Bio-Rad Universal SYBR green kit

and Bio-Rad C1000 touch thermal cycler. Finally, gene enrichment was analyzed as a function

of the quantification cycle (Chq) values obtained from the real-time reaction and expression of

cellular markers were presented as a geometric mean of the ratio of the target mRNA copies to

mRNA copies of the reference genes beta-actin, GAPDH, and cyclophilin G. Data was

processed to determine differential expression of the target genes in aged vs. control subjects,

24  

 

and to demonstrate a linear effect of age on gene expression, as well as to suggest a concerted

movement of gene expression changes with age.

RT-QPCR EXPERIMENTAL PROTOCOL

Summary of RT-qPCR technology

Fluorescence detection modules contained within the RT-qPCR instrument monitor the

fluorescence signal as amplification proceeds. The measured fluorescent signal is proportional

to the total amount of amplicon, i.e., the amplified sequence of interest. The change in

fluorescence over time is used to determine the amount of amplicon produced in each PCR

cycle. RT-qPCR allows the user to assess the initial copy number of the target sequence with

accuracy and sensitivity and without the need for gel electrophoresis.

Data generated from RT-qPCR is displayed as an amplification plot consisting of the number

of PCR cycles on the x-axis, usually ranging from 0-40 cycles, and relative fluorescence units

on the y-axis. Amplification plots are typically divided into two phases – an exponential phase,

during which the amount of PCR product approximately doubles in each cycle, and a non-

exponential phase, in which the reaction slows and reaches a plateau due to reagent exhaustion.

Initially, fluorescence remains at background levels during the exponential phase until enough

amplified product accumulates to yield a detectable fluorescence signal. The cycle number at

which fluorescence crosses past the threshold is known as the quantification cycle, or Cq. The

Cq of a reaction is largely determined by the initial amount of nucleic acid template present at

the start of the amplification. The Cq of a sample is inversely proportional to the initial amount

of template available in the reaction, such that a low, or early, Cq is indicative of larger

25  

 

amounts of starting material present in the reaction because relatively fewer amplification

cycles are required to accumulate enough product to breach threshold and produce

fluorescence above background levels. This relationship forms the quantitative aspect of RT-

qPCR.

The RT-qPCR reagents utilized for this experiment were obtained from Bio-Rad and consisted

of the SsoAdvancedTM Universal SYBR® Green supermix, including buffer, DNA polymerase,

dNTPs, and a dsDNA-binding dye. Specific primer pairs were designed to optimize RNA

detection using the web-based Primer3Plus software, and cDNA was generated from extracted

RNA to serve as the template for the reaction. Specificity of the RT-qPCR assay was assessed

by confirming a single product peak in the melt curve analysis provided by the real-time

instrument.

Sample Collection for RT-qPCR

Fresh-frozen cortical blocks of Brodmann Areas BA11 and BA47, corresponding to the

orbitoventral prefrontal cortex (PFC), were obtained from the Douglas Institute Brain Bank

(Canada) from male subjects without a diagnosed neurodegenerative or neuropsychiatric

disorder. Thirty samples were used in this study and were previously delineated into two

groups on the basis of age, such that control subjects (n = 13) were grouped under age 45 and

aged subjects (n = 17) were grouped above age 60. Subject groups differed in mean age (p =

1.55E-15) but did not significantly differ in other demographic parametres of postmortem

interval (PMI; p = 0.24), brain pH (p = 0.06), or RNA integrity number (RIN; p = 0.84) as

determined by UNIANOVA (Figure 2).

26  

 

0.0#

20.0#

40.0#

60.0#

80.0#

Age#

Age$(years)$

Demographic$Differences$in$Age$

Comparison#

Aged#

p#=#1.55E915#

0.0#

5.0#

10.0#

15.0#

20.0#

25.0#

30.0#

PMI#

PMI$(ho

urs)$

Demographic$Differences$in$PMI$

Comparison#

Aged#

p#=#0.24##

1.0$

3.0$

5.0$

7.0$

9.0$

11.0$

13.0$

pH$

pH#

Demographic#Differences#in#pH#

Comparison$

Aged$

p$=$0.06$$

2.0$

4.0$

6.0$

8.0$

10.0$

RIN$

RIN$

Demographic$Differences$in$RIN$

Comparison$

Aged$

p$=$0.84$

Figure 3. Demographic characteristics of age-cohort.

Comparison and aged subjects were significantly different in mean age (p = 1.55E-15) but did

not significantly differ in PMI (p = 0.24), pH (p = 0.06), or RIN values (p = 0.84). Error bars

represent standard error.

27  

 

Grey matter was identified in the cortical blocks on the basis of gross anatomy and was

selectively harvested into 100 µm sections using a Leica CM1950 cryostat set to -20°C in an

RNAse-free environment. Cortical blocks were approximately 1.5 cm3 in volume, thus the

dimensions for the cryosections were comparable across subjects. Two grey matter

cryosections of 100 µm thickness and approximately 0.75 cm2area were collected for each

subject and immediately immersed in 350 µL of chilled Buffer RLT (Qiagen miRNeasy) for

manual homogenization using RNase-free disposable pellet pestles (Fisherbrand™). Control

and aged samples were collected in a semi-randomized (alternating) manner to minimize order

effects in sample collection93. Following homogenization, samples were transferred on dry ice

and stored at -80°C until further processing.

Demographic and postmortem characteristics of human subjects      Comparison subjects Aged subjects

Case Sex Age (yrs)

PMI (hrs) pH RIN

Cause of death Case Sex

Age (yrs)

PMI (hrs) pH RIN

Cause of death

DH1025 M 41 14.5 5.89 5 Natural DH411 M 76 10.75 5.95 4.4 COPD DH787 M 43 23.75 6.25 2.8 Thrombosis DH476 M 73 27.5 6.17 4.4 COPD S11 M 19 32 6.35 2.3 Natural

DH505 M 69 23.5 6.04 2.9 CA

S15 M 30 30 6.22 4.4 Accident

DH530 M 67 24.75 6.23 6.5 AP S16 M 28 27 6.59 3.3 Natural

DH580 M 69 34.12 5.85 4.2 N/A

S17 M 41 24 5.95 2.4 Accident

DH598 M 79 21.92 6.4 4.3 PE S173 M 20 12 6.22 2.4 Natural

DH650 M 73 10 5.46 2.4 N/A

S20 M 31 29.5 6.49 5.6 Accident

DH724 M 70 32.75 6.02 2.8 HC S215 M 43 27 6.33 6.4 Accident

DH745 M 63 16.75 6.02 5 MI

S250 M 26 12 6.75 3.2 Natural

DH776 M 70 26 5.76 2.4 N/A S31 M 21 24 6.27 4.3 Accident

DH796 M 69 27.58 5.6 4.5 MI

S36 M 27 20.5 6.18 5 Natural

DH880 M 62 6.17 6.15 4.2 VH S94 M 15 27 5.95 4.7 Natural

DH974 M 72 24.5 6.36 6.1 MI

            DH988 M 66 20.5 6.09 4.3 TD

            S101 M 63 13 6.84 3.9 N/A

S187 M 71 17 6.2 3.8 N/A

Table 1. Demographic and postmortem information of study sample. Total of 30 subjects divided into 13 young comparison subjects and 17 aged subjects. M, male; PMI, postmortem interval; RIN, RNA integrity number; COPD, chronic obstructive pulmonary disease; CA, cardiac arrest; AP, acute peritonitis; PE, pulmonary edema; HC, hemorrhagic colitis; MI, myocardial infarction; VH, ventricular hypertrophy; TD, thromboembolic disease; N/A, not available

28  

 

RNA Isolation and Conversion to cDNA

Total RNA was isolated according to the Qiagen RNeasy® micro protocol. Briefly, tissue

samples were homogenized and subjected to centrifugation to separate the supernatant

containing total RNA. Genomic DNA (gDNA) was separated from the preparation using a

gDNA spin column provided in the kit. Subsequent steps purified the RNA, which was eluted

using RNase-free water. Approximately 12 µL of RNA was collected per sample.

Concentration and purity of RNA were determined using a nanospectrophotometer to assess

absorbance ratios. Samples with A260/A280 ratios greater than 1.9 were accepted for

downstream processing, as commonly acceptable A260/A280 ratios for nucleic acids is ~2.

A260/A230 ratios were lower than the commonly acceptable range of 2.0-2.2 (Appendix 1).

This may be attributed to residual RNA extraction buffers in the eluents. This did not,

however, impact the extraction of high concentration RNA to sufficiently create enough

reagents for cDNA synthesis. RNA integrity was assessed to determine the extent of RNA

degradation within the samples using an Agilent 2100 Bioanalyzer System. RNA integrity

numbers (RIN) were generated for each sample. The generally high level of degradation found

within the samples was noted for downstream measures to counteract possible processing

difficulties from degraded RNA.

1 µg of RNA per sample was used to synthesize cDNA using the Superscript VILOTM kit. The

required volumes of reagents were calculated according to the following paradigm:

Components Volume (µL) Nuclease-free water x

Reaction mix 4 Enzyme mix 2

RNA (1 µg) TOTAL 20

29  

 

The reaction mix and the enzyme mix are set to 4 µL and 2 µL per 20 µL reaction, respectively.

The required volume of RNA that amounted to 1 µg of product was added to the reaction tube

and the volume of nuclease-free water was adjusted such that the final volume was 20 µL. A

no-reverse transcriptase control was included to control for genomic DNA (gDNA)

contamination. Reaction tubes were placed in a Bio-Rad T100 thermal cycler set to an adapted

Superscript VILOTM protocol. The program consisted of an initial incubation at 25˚C for 10

minutes, followed by incubation at 42˚C for 120 minutes to allow cDNA synthesis. Reaction

termination was set at 85˚C for 5 minutes, and the samples were held at 4˚C. The program is

summarized below.

Step Temp (˚C)

Time

Incubation 25 10 min Incubation 42 2 h

Termination 85 5 min Hold 4 ∞

Newly synthesized cDNA was diluted 1:4 with nuclease-free water by adding 80 µL into the

reaction tubes to bring the final volume to 100 µL. Samples were stored at -20˚C until further

processing.

Samples were then tested for viability with a validated in-house actin primer. Average Cq

values were calculated and standard errors less than 0.2 standard deviations from the mean of

the biological sample were used for analyses 94,95. Samples were standardized to an actin Cq

value of 20 by adding nuclease-free water according to the following formula(Appendix 2):

30  

 

 [ 2 !"!!"!"# − 1 𝑥  𝑉𝑜𝑙𝑢𝑚𝑒

Newly standardized samples were then stored at -20˚C until further processing.

RT-qPCR Primer Design, Preparation, and Testing

Primers were designed using the Primer3Plus software(Appendix 3). Primers were restricted to

80 base pairs to target short RNA fragments as a preemptive measure to account for RNA

degradation96,97. Sequences were selected that encoded the longest isoform of the transcript

variant. Optimal Tm was set to 60˚C. The first three available forward and reverse primer

sequence pairs representing non-overlapping base pairs were selected for testing and

validation.

100 µM stock solutions were prepared for all forward and reverse primers by adding the

amount of nuclease-free water equal to 10 times the nanomolar amount of the primer.

Lastly, 10 µL of the forward and 10 µL of the reverse primers were added to 980 µL of

nuclease-free water in a 1.5 mL Eppendorf tube to create a 1 mL working solution of each

primer. The working solutions were stored at -20˚C and at4˚C after thawing.

Primer sequences were tested for greatest affinity with an in-house validated actin primer and

stock human cDNA to determine which of the sequence pairs functioned most optimally. Each

primer pair was tested in triplicate to calculate an average actin Cq and standard errors less

than 0.2 standard deviations from the mean of the biological sample were used for analyses

94,95. No template controls were included to control for the presence of contaminants and

primer-dimer formation. Primer pairs that produced earlier amplifications of actin in the human

31  

 

sample, and thus lower Cq values, and also tested negative for contaminants or primer-dimers,

were selected for downstream use.

RT-qPCR Program and Assay

The RT-qPCR program used for the Bio-Rad C1000 touch thermal cycler was set according to

the Bio-Rad Universal SYBR green protocol. The program consisted of an initial denaturation

step carried out at 95˚C for 30 seconds. This was followed by 40 qPCR cycles consisting of

denaturation at 95˚C for 15 seconds and annealing/extension at 60˚C for 30 seconds, followed

by a plate read to generate a melting curve. Lastly, a final round of denaturation at 95˚C for 30

seconds, followed by a final extension at 65˚C for 30 seconds and a hold at 20˚C for 5 minutes.

The program is summarized below.

Step Temp (˚C)

Time

Initial denaturation 95 30 sec

Denature 95 15 sec Anneal/extension 60 30 sec

Plate read Denature 95 30 sec Final extension 65 30 sec

Hold 20 5 min

Samples for aged and comparison subjects were prepared in a semi-random (alternating)

manner to minimize order effects in sample preparation and subsequent confounds in data

analysis 93. 96-well plates supplied by Bio-Rad Laboratories were used to assay the samples.

Two GABAergic markers (SST, PVALB) and one glial marker (GFAP) were chosen to pilot

the study based on evidence of robust differential expression with age 5,31. Reference genes

beta-actin (ACTN), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and cyclophilin G

40 cycles

32  

 

(CYCLO) were chosen as internal controls 98. Each reference gene and pilot gene of interest

was tested in quadruplicate for each subject using a 96-well plate supplied by Bio-Rad

Laboratories. Average Cq values were calculated and standard errors less than 0.1 standard

deviations from the mean of the biological sample were used for analyses94,95. The remaining

cell markers (CALB1, GPHN, ALDH1L1, GLUL, GRM2, GRM3) were tested in triplicate

along with two reference genes (ACTN, GAPDH) to accommodate more biological replicates

within the plate design. CYCLO was excluded from the subsequent plate design and future

analyses because it correlated poorly with both ACTN (r = 0.7) and GAPDH (r = 0.7) gene

expression in the pilot study, whereas ACTN and GAPDH expression correlated strongly with

each other (r = 0.9).

RT-qPCR Fold Change Calculation

All RT-qPCR curves were set to a threshold of 150 RFU to standardize data analysis across

biological samples. Delta Cq values to determine the quantification cycle for each target gene

were calculated for each subject by subtracting the mean Cq value of the target gene from that

of ACTN and GAPDH. The difference in quantification cycle threshold represents the log2-

transformed expression ratio of each gene of interest to the geometric mean of the two

reference genes. Relative expression levels of the genes of interest were calculated according

to the following formula:

2!!"#$%&'

Statistical Analyses

Statistical significance threshold was set at p ≤ 0.05. Expression values were found to display

left-handed skewness and were subjected to a square root transformation to achieve a more

33  

 

normalized distribution of values across all subjects. Expression values greater than 3 standard

deviations from the mean were considered to be outliers and were excluded from analysis

where possible, or were replaced by the mean expression value for that gene when a value was

mandated. A univariate analysis of variance (UNIANOVA) model was performed using SPSS

to identify any demographic effect of age, PMI, pH and RIN on expression level of the target

genes(Appendix 4). Results were adjusted for multiple comparisons. Age was retained as a

significant covariate for expression levels of all target genes. pH was found to be a covariate

for ALDH1L1 and GPHN expression.

34  

 

CHAPTER 3: RESULTS

35  

 

The Majority of Gene Expression Changes are Age-Dependent

A two-group analysis was performed to determine the main effect of age on gene

expression between aged and comparison subjects. Expression of SST was significantly

reduced in aged subjects (p = 1.95E-04), as was expression of PVALB (p = 0.021).

Expression of CALB1 showed a nominal reduction in aged subjects (p = 0.124). GPHN

expression was significantly increased in aged subjects (p = 0.008). Expression of

GRM2 (p = 0.291) and GRM3 (0.112) were not significantly reduced in aged subjects.

Expression of ALDH1L1 was significantly increased (p = 0.011) as well as the

expression of GFAP (p = 0.004). Lastly, the expression of GLUL was nominally

increased in aged subjects (p = 0.095). Altogether, the discrepancies in gene expression

between aged and comparison subjects from this novel cohort are largely consistent

with previous findings in the literature, thereby supporting the model of a conserved

and pervasive molecular program that mediates the aging process.

Altered GABAergic Marker Expression with Age

Correlation analyses were performed to reveal the linear effect of age on gene

expression levels. The expression of SST was moderately negatively correlated with

age across the lifespan (r = -0.5). Expression of PVALB was also moderately

negatively correlated with age across the lifespan (r = 0.4). CALB1 expression was

weakly negatively correlated with age across the lifespan (r = -0.3). GPHN expression

was weakly positively correlated with age across the lifespan (r = 0.3). Thus, the

expression of SST and PVALB displayed a moderate linear decrease due to the effect

of age.

36  

 

Increased Glial Marker Expression with Age

The expression of ALDH1L1 was moderately positively correlated with age across the

lifespan (r = 0.4). Expression of GFAP was also moderately positively correlated with

age across the lifespan (r = 0.4). GLUL expression was weakly positively correlated

with age across the lifespan (r = 0.2). Thus, ALDH1L1 and GFAP exhibited moderate

linear increases in gene expression with advancing age.

Unchanged Glutamatergic Marker Expression with Age

Expression of GRM2 was weakly positively correlated with age across the lifespan (r =

-0.2), as was GRM3 expression (r = -0.3). Thus, GRM2/3 expression did not show

significant changes in linear expression across the lifespan.

37  

 

Figure 4. Main effect of gene expression levels in aged versus comparison subjects.

Expression of SST was significantly reduced in aged subjects (p = 1.95E-04), as was

expression of PVALB (p = 0.021). Expression of CALB1 was not significantly reduced in

aged subjects (p = 0.124). GPHN expression was significantly increased (p = 0.008).

Expression of GRM2 (p = 0.291) and GRM3 (0.112) were not significantly reduced in aged

subjects. Expression of ALDH1L1 was significantly increased (p = 0.011) as well as the

expression of GFAP (p = 0.004). Lastly, the expression of GLUL was nominally increased in

aged subjects (p = 0.095). Error bars represent standard error.

0.0%  

50.0%  

100.0%  

150.0%  

200.0%  

250.0%  

300.0%  

350.0%  

Rel

ativ

e E

xpre

ssio

n (%

) Gene expression levels in aged vs. comparison subjects

Comparison

Aged *

***

***

*** ***

#

38  

 

Figure 5a-d. Linear effect of age on

SST (a), PVALB (b), CALB1 (c), and

GPHN (d) expression.

The expression of SST was significantly

decreased and moderately negatively

correlated with age across the lifespan

(p = 1.95E-04; r = -0.5). Expression of

PVALB was also significantly

decreased and moderately negatively

correlated with age across the lifespan

(p = 0.02; r = 0.4). CALB1 expression

was non-significantly decreased and

weakly negatively correlated with age

across the lifespan (p = 0.12; r = -0.3).

GPHN expression was significantly

increased and weakly positively

correlated with age across the lifespan

(p = 0.008; r = 0.3).

0  

2  

4  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Decreased SST Expression with Age

R  =  -­‐0.5    p  =  1.95E-­‐4  

0  

1  

2  

3  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Decreased PVALB Expression with Age R  =  -­‐0.4   p  =  0.02

0  

1  

2  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Decreased CALB1 Expression with Age

R  =  -­‐0.3    p  =  0.12  

0  1  2  3  4  5  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Increased GPHN Expression with Age

R  =  0.3   p  =  0.008

39  

 

Figure 6a-c. Linear effect of age on

ALDH1L1 (a), GFAP (b), and

GLUL (c) expression.

The expression of ALDH1L1 was

significantly increased and moderately

positively correlated with age across

the lifespan (p = 0.01; r = 0.4).

Expression of GFAP was also

significantly increased and moderately

positively correlated with age across

the lifespan (p = 0.04; r = 0.4). GLUL

expression was nominally increased

and weakly positively correlated with

age across the lifespan (p = 0.09; r =

0.2).

0  

0.5  

1  

1.5  

2  

2.5  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Increased ALDH1L1 Expression with Age

R  =  0.4   p  =  0.01

0  

10  

20  

30  

40  

50  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Increased GFAP Expression with Age R  =  0.4 p  =  0.04

0  

20  

40  

60  

80  

100  

120  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Increased GLUL Expression with Age

R  =  0.2   p  =  0.09

40  

 

Principle Component of Variance in Gene Expression is Captured by Age

A principle component analysis was performed to extract the primary variable that was

most attributable to differences in gene expression between subject groups. Principle

component 1 (PC1) was determined as the principle component of variance in gene

expression at a threshold of 40% and greater and accounted for 77.1% of the variance

in the gene expression data. Principle component 2 (PC2) accounted for 13.6% of the

variance. The distribution of genes according to PC1 largely paralleled the direction of

reported expression changes due to age and was thus attributed as the primary variable

mediating differences in gene expression between subject groups.

[CAPTION]

0  

0.05  

0.1  

0.15  

0.2  

0.25  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Unchanged GRM2 Expression with Age

R  =  -­‐0.2    p  =  0.29  

0  

0.5  

1  

1.5  

2  

0   20   40   60   80   100  

Exp

ress

ion

Valu

e

Age (years)

Unchanged GRM3 Expression with Age

R  =  -­‐0.3    p  =  0.11  

Figure 7a-b. Linear effect of age on

GRM2 (a) and GRM3 (b) expression.

Expression of GRM2 was weakly and

correlated with age across the lifespan (p

= 0.29; r = -0.2). GRM3 expression was

also weakly and non-significantly

correlated with age across the lifespan (p

= 0.11; r = -0.3).

41  

 

Gene Expression Changes in Circuitry are Concerted

A heat map was generated for expression values to determine clustering effects of gene

expression as a function of age. Strong correlations between GLUL, GPHN,

ALDH1L1, and GFAP age-dependent expression formed a cluster group, as did strong

correlations between SST, CALB1, GRM2, PVALB, and GRM3 expression. These

cluster groups were representative of age-dependent gene expression changes in

functional units of the cortex that played a role in glial activity and in

excitatory/inhibitory/calcium transmission. These results suggest that age-regulated

genes are part of a transcriptomic program that elicits expression changes occurring in

parallel in groups of specific genes.

Cell Markers Group Together Irrespective of Age

A heat map was also generated to determine gene expression clustering effects in the

absence of age as a statistical covariate of the data. Strong correlations between GLUL,

GPHN, ALDH1L1, GFAP, and GRM3 expression formed a cluster group, as did SST,

CALB1, GRM2, and PVALB expression. These cluster groups were mostly similar to

those generated with age as a covariate of the expression data, which is suggestive of

intrinsic networks of cells and biological systems that are programed to undergo

concerted changes associated with the aging phenotype.

42  

 

Figure 8. Analysis of the principle components of variance attributed to cell marker gene expression.

The principle component matrix is depicted as (PC1, PC2) values in (x, y) coordinates.

Principle component 1 (PC1) was determined as the principle component of variance in gene

expression at a threshold of 40% and greater. PC1 accounted for 77.1% of the variance in the

gene expression data. Principle component 2 (PC2) accounted for 13.6% of the variance. The

distribution of genes according to PC1 largely parallels expression changes due to result of

age.

SST  

GFAP  

PVALB  ALDH1L1  

CALB1  GLUL  

GPHN  GRM2  

GRM3  

.000  

.100  

.200  

.300  

.400  

.500  

.600  

.700  

.800  

.900  

-­‐.800   -­‐.600   -­‐.400   -­‐.200   .000   .200   .400   .600   .800   1.000  

Principle Component Analysis of Cell Marker Gene Expression

(-­‐0.588,  0.575)  

(-­‐0.524,  0.448) (-­‐0.613,  0.411)

(-­‐0.564,  0.196)

(0.083,  0.834)

(0.780,  0.472) (0.835,  0.451)

(0.737,  0.023)

(0.915,  0.113)

Principle component 1 (PC1)

Prin

cipl

e co

mpo

nent

2 (P

C2)

43  

 

Figure 9. Heat map of cellular marker expression due to age.

Correlation matrix composed of pairwise correlations between cellular markers accounted for

by age. GLUL, GPHN, ALDH1L1, and GFAP expression were strongly correlated with each

other and less so with SST, CALB1, GRM2, PVALB, and GRM3 expression, and vice versa

(Appendix 5).

ALDH1L1  

ALDH

1L1  

44  

 

Figure 10. Heat map of cellular marker expression in absence of age effect.

Correlation matrix composed of pairwise correlations between cellular markers in the absence

of age effects. GLUL, GPHN, ALDH1L1, GFAP, and GRM3 expression were strongly

correlated with each other and less so with SST, CALB1, GRM2, and PVALB expression, and

vice versa (Appendix 5).

 

GLUL  

GPHN  ALDH1L1  

GFAP  

GRM3  

SST  

CALB1  

GRM2  

PVALB  

CALB1  

GRM

2  

PVAL

B  

SST  

GRM

3  

GFA

P  

ALDH

1L1  

GPH

N  

GLU

L  

45  

 

CHAPTER 4: DISCUSSION

46  

 

SUMMARY

Expression of SST (somatostatin) was significantly reduced in aged subjects, as well as with

PVALB (parvalbumin) (p = 1.95E-04; p = 0.021). Expression of CALB1 (calbindin) was

nominally decreased in aged subjects (p = 0.124). Conversely, GFAP (glial fibrillary acidic

protein), ALDH1L1 (aldehyde dehydrogenase 1 family member L1), and GPHN (gephyrin)

expression were significantly upregulated in aged subjects (p = 0.004; p = 0.011; p = 0.008).

Expression of GLUL (glutamine synthase) trended towards increased expression in aged

subjects (p = 0.095).Expression of group II metabotropic glutamate receptors GRM2 and

GRM3 were nominally reduced in aged subjects (p = 0.291; p = 0.112). Altogether, these

results indicate changes in the gene expression profile of the aged prefrontal cortex in a novel

cohort that are largely consistent with previous findings.

ROLE OF CELL MARKERS IN GABA-MEDIATED LOCAL CIRCUIT FUNCTION

The significantly reduced expression of interneuron markers SST and PVALB, and the

nominal reduction of CALB1 expression observed in aged subjects suggests a disturbance, if

not a reduction, in inhibitory GABAergic tone within cortical neurocircuits. The significantly

increased expression of GPHN may be indicative of a cellular attempt to increase the density of

inhibitory receptors in postsynaptic membranes. Additionally, the significant upregulation of

glial markers GFAP and ALDH1L1, and the nominal upregulation of GLUL expression may

represent an increase in glial reactivity, which may have profound effects on the cortical

microenvironment and on circuit function due to the release of cytotoxic factors. The trend of

reduced GRM2/3 expression in aged subjects might be indicative of subtle alterations to

glutamatergic signalling within cortical neuronal populations. Ultimately, the molecular

47  

 

changes that occur with aging may engender an imbalance between excitatory and inhibitory

drives in local cortical circuits, and thus precipitate neurological disease.

The results obtained from this study are largely consistent with previously reported data.

Inhibitory neuron markers SST, PVALB, and CALB1 have been shown to be downregulated in

older subjects compared to young controls, and in the absence of diagnosed neuropsychiatric

disorder. The postsynaptic marker GPHN exhibited an upregulation in expression during the

aging process, which is consistent with prior literature. Additionally, the change in glial marker

expression of GFAP, ALDH1L1, and GLUL were congruent with previous cohorts. Lastly,

although changes GRM2/3 expression failed to reach statistical significance, their nominal

reduction is indicative of a trend towards downregulation across the lifespan.

Altogether, these findings validate previous reports in the literature and evince the importance

of replication to support newly emerging theories in relation to the molecular correlates of

aging. Furthermore, these findings also validate the novel age-cohort as appropriate subjects

with which to investigate the molecular mechanisms of aging in future investigations.

ROLE OF OBSERVED CHANGES IN AGING

Data generated from this study support that signalling within local cortical circuits is altered

during the aging process. These alterations appear to involve altered inhibitory GABAergic

neuropeptide expression, excitatory glutamatergic signalling, and intracellular calcium

transmission, and an accompanying increase in glial reactivity. The significant downregulation

of SST and PVALB, and nominal reduction in CALB1 expression suggests reduced inhibitory

markerexpression of GABAergic interneurons in the aging cortex. Furthermore, this

48  

 

phenomenon may pose consequences for cytosolic calcium signalling in neuronal cells, due to

the downregulation of cytosolic binding proteins PVALB and CALB1, and may alter

homeostatic levels of cellular function. An increase in the pool of unbound calcium ions could

have widespread and ultimately deleterious range of effects on cell, network, and brain,

integrity. Interestingly, the significantly increased expression of GPHN may indicate a

compensatory mechanism to increase in the density of inhibitory receptors in postsynaptic

membranes to account for a reduction in inhibitory neuropeptide transmission. The nominal

downregulation of GRM2 and GRM3suggests a decrease in glutamatergic signal transmission

within cortical circuits, possibly as a result of altered pyramidal cell function. Expression

changes in GRM2/3 may be secondary to primary deficits in inhibitory neuronal populations,

which appear to be particularly vulnerable to age effects, and would thus be less significantly

affected across the lifespan 99. The reactivity of astrocytes appears to be increased in aged

subjects, as suggested by the significant upregulation of ALDH1L1 and GFAP, and the

observed trend in GLUL. This increase in reactivity may indirectly impact the environment of

cortical circuits as well as directly affect the activity of interneurons and pyramidal cells. The

trend towards increased expression of GLUL, which mediates the metabolism of glutamate to

GABA in astrocytes, may also represent a compensatory mechanism to increase the overall

pool of GABAergic signalling within cortical circuits by increasing the catalysis GABA

biosynthesis. Overall, these results suggest that in cortical neurocircuits, altered gene

expression changes the balance between GABAergic and glutamatergic signalling, as well as

intracellular calcium transmission and the reactivity of glial cells, and is associated with the

aging phenotype.

49  

 

These results lend credence to the theory of a transcriptomic program regulating the aging

phenotype. Analysis of the principle component of variance attributed to cellular marker

expression suggested that the magnitude and direction of gene expression change could likely

be attributed to age. The first principle component (PC1) accounted for most of the variance in

the expression data (77.1%) and resulted in a distribution that mirrored the effect of age on

gene expression of the selected markers. Additionally, the generation of heat maps highlighted

strong correlations between the gene expression of cellular markers that could be

conceptualized as a belonging to a similar functional group, such as involvement in

GABAergic/glutamate/calcium regulation or astrocyte reactivity. Furthermore, these functional

groups were primarily maintained after the contribution of age was excluded from the model,

suggesting a concerted change in neurocircuits that shifted cells towards a different pattern of

gene expression across the lifespan. In other words, the data suggests correlated plasticity

within functional cell circuits in aging.

AGE-BY-DISEASE INTERACTION MODEL

Normal molecular brain aging is an integral aspect and modulator of disease onset and

progression, ranging from late onset neurodegenerative diseases, such as AD and PD, to earlier

onset neuropsychiatric diseases, including SCZ and BPD58,59.

Molecular aging preferentially affects a large proportion of genes related to neurological

diseases60.Approximately 5-10% of genes expressed by the human genome are regulated by

age60. From this set, ~34% of age-regulated genes in the human brain have been shown to be

associated to at least one of six age-gated neurological diseases, including AD, PD, HD, ALS,

BPD, and SCZ60. Moreover, the majority of genes associated with these diseases have

50  

 

expression levels that significantly vary with chronological age60. Additionally, the direction of

expression changes that occur in disease-related genes is almost unanimously toward pro-

disease directions along the lifespan60. These findings suggest that aging may promote

disease60.

It appears that genes related to developmental transcriptional programs may have a dual role in

promoting the aging process, especially considering that certain developmental phenomena,

such as synaptic pruning, mirror aging phenotypes (i.e., synaptic loss), and thus could be

extensions from similar biological programs6,31. This may also explain, at least in part, why a

variety of genes associated with developmental diseases, such as BPD and SCZ, are also

regulated by age6,31.

Whilst evidence suggests that normal aging may modulate disease, the converse does not

appear to be true, i.e., not all genes associated with disease susceptibility have been proven to

modulate aging6,15,102. For instance, neuronal loss, which is an integral feature of

neurodegenerative diseases, does not occur during normal aging6,103.

The remarkably consistent and selective gene expression changes that have been reported by

numerous groups using a variety of platforms could be attributed to a tightly regulated age-

gated transcriptional program6. Furthermore, alternative hypotheses, such as cellular wear and

tear and oxidative stress, only partially account for the selectivity of the observed changes in

gene expression6.

51  

 

A robust example of this proposed age-by-disease interaction model can be found with MDD.

The etiology of MDD shares some molecular similarities with aging. For example, the

expression of SST is reduced in various brain regions in subjects with MDD, as is also true of

cortical SST expression in aging 49,87,98. Additionally, the absence of SST transmission or cell

function induces changes to behavioural emotionality in mice 18,104. Although the diagnoses of

mood disorders are not necessarily higher in the older human population, studies do

demonstrate a rise in the numbers of depressive symptoms in the elderly6,105. This

incongruence may likely suggest issues with diagnostic structure and procedure6. The

misattribution of psychomotor retardation, anhedonia, memory loss, and reduced cognitive

speed to aging per se leads to under diagnosing of depression in older adults6,106,107.

Interestingly, adults with poorer performance on cognitive tests have been shown to defer less

to positivity-bias, suggesting decreased emotional control, and thus introducing a cause-effect

confound6,13. Moreover, the presence of low mood and major depressive symptoms in the

context of AD and PD, further imply dysregulation of emotional control6,105. Altogether, it

seems that the loss of emotional regulation with age in people with more accelerated cognitive

decline may represent a major risk factor for depression in old age and is thus a robust example

of a proposed age-by-disease interaction model6.

52  

 

STUDY LIMITATIONS

Although postmortem studies are useful in providing the resolution required to characterize the

aging phenotype at the level of populations of neurons and neural circuits, the evidence

provided by this study is correlative and cannot be used to make causal claims about the

molecular mechanisms of aging. Genetic manipulation studies in animal models would be

necessary to determine the cause-effect relationship between gene expression changes and the

aging phenotype. In addition to this, the analysis of postmortem samples can be problematic

for reasons that preclude the standardization of data, including variation in postmortem

intervals, which can impact the structural integrity of target molecules and, thus, their accurate

quantification. Furthermore, the heterogeneity of postmortem cohorts can be confounding to

the emergence of unifying theories associated with the neurobiology of aging, due to

discrepancies in gene expression results. The subject groups in this age-cohort were notably

different in their causes of death, and as such it cannot be conclusively determined whether the

observed differences in gene expression were attributable to age or to the cause of death.

Another caveat is that the conclusions generated from this study are limited by the restricted

scope of cellular markers that were analyzed. Additional research would be required with a

broader range of genes representing GABAergic/glutamatergic/calcium signalling, synaptic

function, and astrocyte reactivity to strengthen theories about the molecular correlates of aging

and their relation to neurological diseases. Lastly, studies of postmortem mRNA transcript

levels represent an isolated time period and may not necessarily be reflective of cellular

function at the protein level. Supplementation with proteomic analysis would be beneficial in

assessing the dynamic patterns of functional protein networks across the lifespan.

53  

 

CONCLUSIONS

The modulation of an array of neurological diseases by normal molecular brain aging is a

promising framework for the creation of anti-aging therapeutics. This emerging knowledge

notwithstanding, it is still largely unknown how aging gates neurological diseases, and which

elements contribute to individual variability in age of onset. A transcriptional program that

progressively constrains genes towards disease-promoting directions may represent the integral

feature of normal brain aging that is necessary for disease onset. This knowledge will be

essential in moving the field forward in that it forms the basis for a dimensional definition of

diseases that progresses beyond the traditional categorical system of neurological diseases.

54  

 

APPENDIX

1. Absorbance ratios and concentration yields from RNA extraction Comparison subjects Aged subjects

Case A260/A280

A260/A230 Conc. (ng/uL) Total Yield

(ng) Case A260/A280

A260/A230 Conc. (ng/uL) Total Yield

(ng) DH1025 2.097 1.667 319 3828 DH411 1.987 1.619 195 2340

DH787 2.032 0.887 76 912 DH476 1.984 1.951 241 2892

S11 1.97 1.13 195 2340 DH505 1.969 2.033 100 1200

S15 1.99 2.114 163 1956 DH580 1.972 2.069 84 1008

S16 1.976 1.723 97 1164 DH598 2.029 0.767 92 1104

S17 2.007 1.443 121 1452 DH650 2 1.908 210 2520

S173 2.007 1.443 121 1452 DH724 2 2.077 86 1032

S20 1.99 2.135 400 4800 DH745 1.95 0.279 493 5916

S215 2.008 1.84 96 1152 DH776 1.988 1.283 129 1548

S31 2.009 1.225 373 4476 DH796 1.988 1.283 129 1548

S36 1.978 1.69 144 1728 DH880 1.98 2.126 390 4680

S94 1.97 1.791 79 948 DH965 2 1.86 170 2040

DH974 1.971 1.093 188 2256

DH988 1.991 1.957 92 1104

DH989 1.947 2.846 148 1776

S101 2.031 0.975 79 948

S187 2 1.415 116 1392

2. Standardized cDNA volumes

Sample Cqavg Cq SE dH2O to dilute to ACTN Cq 20 Sample Cqavg Cq SE dH2O to dilute to

ACTN Cq 20 S215 16.93 0.08 719 S16 16.37 0.03 1104

DH530 16.49 0.02 1008 DH598 16.36 0.04 1109

S20 16.61 0.18 918 S250 16.89 0.06 741

DH974 17.67 0.18 390 DH880 16.81 0.05 790

DH1025 16.50 0.04 997 DH787 16.89 0.12 741

DH745 18.08 0.03 269 DH580 18.10 0.02 264

S36 16.66 0.03 886 S173 17.31 0.01 528

DH796 19.75 0.05 18 DH965 17.42 0.01 481

S46 23.96 0.08 -91 S173 15.67 0.05 1848

DH411 16.36 0.03 1112 S101 16.36 0.04 1108

S94 16.39 0.06 1086 S11 17.62 0.04 407

DH989 19.70 0.05 22 S187 18.62 0.02 155

S15 17.48 0.01 460 DH724 19.17 0.11 75

DH476 16.94 0.03 712 DH505 18.17 0.01 248

S31 16.78 0.03 808 DH650 19.85 0.09 10

DH988 17.67 0.03 389 DH776 20.36 0.02 -21

55  

 

APPENDIX 3. Forward and reverse primer sequences

Gene Accession no. Sequence Start pos. (bp) Size (bp) Conc. (nmol) Prod. Size (bp) ACTB NM_001101 GCAAAGACCTGTACGCCAAC 953 20 30.8 80

CTCCTTCTGCATCCTGTCG 1032 19 26.8 GAPDH NM_002046 GGGCTCTCCAGAACATCATC 790 20 27 80

CTTCCCGTTCAGCTCAGG 869 18 29.4 CYCLO NM_004792 GTTATCGAACTCCTTCCAGATCC 1239 23 24.3 80

CATCTCTTGCCTCCAATGTG 1318 20 37 SST NM_001048 AACCCAACCAGACGGAGA 298 18 32.8 80

CTCAAGCCTCATTTCATCCTG 377 21 34.7 PVALB NM_002854 CGCTACCGACTCCTTCGAC 110 19 30.9 80

CTTCTTCACATCATCCGCACT 188 21 27.4 CALB1 NM_004929 CGAGAAGAATAAACAGGATCTGG 841 23 26.9 80

TCCCTCCATCCGACAAAG 920 18 25.2 GPHN NM_000310 TCAGAATCGCCTCGTGCT 2022 18 31.8 80

CTGCACCTGGACTGGACTTT 2101 20 31.3 GFAP NM_002055 GATGGCGCTAGGCATACAA 2360 19 28.1 80

GTCACTTCCTTAATTCCCACAATC 2439 24 30.5 ALDH1L1 XM_006713481 CATCTATCACCCGTCACTGC 398 20 27.5 80

CTTTCTTATCTCCGTGAATGAGG 477 23 26.5 GLUL NM_002065 GCGGGAGGAGAATGGTCT 1680 18 28.7 80

AGGCACGGATGTGGTACTG 1759 19 30.3 GRM2 NM_000839 CATTCCTGCCCATCTTCTATGT 2563 20 27.1 80

GCTGAGGCTGACTGACACG 2642 22 23.1 GRM3 NM_000840 TTAGTCTGTTCCTCCCTTATTTGAAG 980 26 25.7 80

GCCAATGTGTCATGCTAGTCC 1059 21 31.2

4. UNIANOVA analysis of variance SST PVALB CALB1 GPHN GFAP ALDH1L1 GLUL GRM2 GRM3 Age Pearson

Correlation -.511 -.429 -.279 .276 .383 .395 .222 -.2 -.321

Sig. (2-tailed) .000 .021 .124 .008 .004 .011 .095 .291 .112

N 30 30 30 30 30 30 30 30 30 PMI Pearson

Correlation .124 -.117 .06 -.169 -.318 -.176 -.158 .061 .134

Sig. (2-tailed) .904 .232 .550 .551 .259 .887 .541 .822 .794

N 30 30 30 30 30 30 30 30 30 pH Pearson

Correlation .164 .25 .082 .468 -.163 .179 .21 .188 .343

Sig. (2-tailed) .720 .753 .364 .002 .765 .017 .136 .780 .239

N 30 30 30 30 30 30 30 30 30 RIN Pearson

Correlation .209 -.109 .13 .114 -.154 -.042 .08 .279 .204

Sig. (2-tailed) .172 .766 .206 .958 .318 .355 .850 .156 .405

N 30 30 30 30 30 30 30 30 30

56  

 

APPENDIX

5. Gene expression correlation matrix SST PVALB CALB1 GPHN GFAP ALDH1L1 GLUL GRM2 GRM3

SST 1 0.4506 0.5389 -0.2259 -0.4412 -0.4283 -0.1620 0.3492 0.3291

PVALB 0.4506 1 0.3117 -0.3367 -0.2982 -0.4210 -0.3101 0.2210 0.0460

CALB1 0.5389 0.3117 1 -0.3051 -0.3215 -0.4456 -0.2929 0.4478 0.1446

GPHN -0.2259 -0.3367 -0.3051 1 0.5241 0.7777 0.8028 -0.0713 0.3639

GFAP -0.4412 -0.2982 -0.3215 0.5241 1 0.5681 0.5988 -0.3088 0.0255

ALDH1L1 -0.4283 -0.4210 -0.4456 0.7777 0.5681 1 0.8570 -0.4636 0.0622

GLUL -0.1620 -0.3101 -0.2929 0.8028 0.5988 0.8570 1 -0.3400 0.3957

GRM2 0.3492 0.2210 0.4478 -0.0713 -0.3088 -0.4636 -0.3400 1 0.2504

GRM3 0.3291 0.0460 0.1446 0.3639 0.0255 0.0622 0.3957 0.2504 1

57  

 

REFERENCES

1. López-Otín C, Blasco MA, Partridge L, Serrano M, Kroemer G. The hallmarks of aging. Cell. 2013;153(6):1194-1217. doi:10.1016/j.cell.2013.05.039.

2. World Population Ageing 2015.

3. Canada’s Population Estimates: Age and Sex.; 2015. http://www.statcan.gc.ca/daily-quotidien/150929/dq150929b-eng.htm.

4. Vaupel JW. Biodemography of Human Ageing. Nature. 2010;464(7288):536-542. doi:10.1038/nature08984.

5. Sibille E. Molecular aging of the brain, neuroplasticity, and vulnerability to depression and other brain-related disorders. Dialogues Clin Neurosci. 2013;15(1):53-65. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3622469&tool=pmcentrez&rendertype=abstract. Accessed January 18, 2016.

6. Glorioso C, Sibille E. Between destiny and disease: Genetics and molecular pathways of human central nervous system aging. Prog Neurobiol. 2011;93(2):165-181. doi:10.1016/j.pneurobio.2010.11.006.

7. Salthouse TA. The Processing-Speed Theory of Adult Age Differences in Cognition. Psychol Rev. 1996;103(3):403-428.

8. Dickstein DL, Kabaso D, Rocher AB, Luebke JI, Wearne SL, Hof PR. Changes in the structural complexity of the aged brain. Aging Cell. 2007;6(3):275-284. doi:10.1111/j.1474-9726.2007.00289.x.

9. Feldman ML, Dowd C. Loss of dendritic spines in aging cerebral cortex. Anat Embryol (Berl). 1975;148(3):279-301. doi:10.1007/BF00319848.

10. Kauranen K, Vanharanta H. Influences of aging, gender, and handedness on motor performance of upper and lower extremities. Percept Mot Skills. 1996;82(2):515-525. doi:10.2466/pms.1996.82.2.515.

11. Volkow N, Gur R, Wang G, et al. Association between decline in brain dopamine activity with age and. Am J Psychiatry. 1998;155(3).

12. Ruffman T, Henry JD, Livingstone V, Phillips LH. A meta-analytic review of emotion recognition and aging: Implications for neuropsychological models of aging. Neurosci Biobehav Rev. 2008;32:863-881. doi:10.1016/j.neubiorev.2008.01.001.

13. Mather M, Carstensen LL, Carstensen LL, et al. Aging and motivated cognition: the positivity effect in attention and memory. Trends Cogn Sci. 2005;9(10):496-502. doi:10.1016/j.tics.2005.08.005.

14. Wilson RS, Barnes LL, Mendes de Leon CF, et al. Depressive symptoms, cognitive decline, and risk of AD in older persons. Neurology. 2002;59(3):364-370. http://www.ncbi.nlm.nih.gov/pubmed/12177369. Accessed June 26, 2016.

15. Deary IJ, Strand S, Smith P, Fernandes C. Intelligence and educational achievement. Intelligence. 2007. doi:10.1016/j.intell.2006.02.001.

58  

 

16. John B. Carroll. Human Cognitive Abilities. Cambridge: Cambridge University Press; 1993.

17. Araque A, Navarrete M. Glial cells in neuronal network function. Philos Trans R Soc Lond B Biol Sci. 2010;365(1551):2375-2381. doi:10.1098/rstb.2009.0313.

18. Soumier A, Sibille E. Opposing effects of acute versus chronic blockade of frontal cortex somatostatin-positive inhibitory neurons on behavioral emotionality in mice. Neuropsychopharmacology. 2014;39(9):2252-2262. doi:10.1038/npp.2014.76.

19. McQuail JA, Frazier CJ, Bizon JL. Molecular aspects of age-related cognitive decline: the role of GABA signaling. Trends Mol Med. 2015;21(7):450-460. doi:10.1016/j.molmed.2015.05.002.

20. Bak LK, Schousboe A, Waagepetersen HS. The glutamate/GABA-glutamine cycle: aspects of transport, neurotransmitter homeostasis and ammonia transfer. J Neurochem. 2006;98(3):641-653. doi:10.1111/j.1471-4159.2006.03913.x.

21. Northoff G, Sibille E. Why are cortical GABA neurons relevant to internal focus in depression? A cross-level model linking cellular, biochemical and neural network findings. Mol Psychiatry. 2014;19(9):966-977. doi:10.1038/mp.2014.68.

22. Hoogland TM, Parpura V. Editorial: The role of glia in plasticity and behavior. Front Cell Neurosci. 2015;9. doi:10.3389/fncel.2015.00356.

23. Lynch AM, Murphy KJ, Deighan BF, et al. The impact of glial activation in the aging brain. Aging Dis. 2010;1(3):262-278. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3295033&tool=pmcentrez&rendertype=abstract. Accessed March 1, 2016.

24. Volterra A, Meldolesi J. Astrocytes, from brain glue to communication elements: the revolution continues. Nat Rev Neurosci. 2005;6(8):626-640. doi:10.1038/nrn1722.

25. Kielian T. Glial connexins and gap junctions in CNS inflammation and disease. J Neurochem. 2008;106(3):1000-1016. doi:10.1111/j.1471-4159.2008.05405.x.

26. Halassa MM, Fellin T, Haydon PG, et al. The tripartite synapse: roles for gliotransmission in health and disease. Trends Mol Med. 2007;13(2):54-63. doi:10.1016/j.molmed.2006.12.005.

27. Ni Y, Malarkey EB, Parpura V. Vesicular release of glutamate mediates bidirectional signaling between astrocytes and neurons. J Neurochem. 2007;103(4):1273-1284. doi:10.1111/j.1471-4159.2007.04864.x.

28. Abbott NJ, Rönnbäck L, Hansson E. Astrocyte-endothelial interactions at the blood-brain barrier. Nat Rev Neurosci. 2006;7(1):41-53. doi:10.1038/nrn1824.

29. Kuno R, Yoshida Y, Nitta A, et al. The role of TNF-alpha and its receptors in the production of NGF and GDNF by astrocytes. 2006. doi:10.1016/j.brainres.2006.07.120.

30. Erraji-Benchekroun L, Underwood M, Arango V, et al. Molecular aging of human prefrontal cortex is selective and continuous throughout adult life. Biol Psychiatry. 2005. http://ac.els-cdn.com/S0006322304011187/1-s2.0-S0006322304011187-main.pdf?_tid=c5c53c64-be1a-11e5-a8e9-00000aab0f02&acdnat=1453145970_aec5df533cb3c4df2c8363736ce0c4eb. Accessed January 18, 2016.

31. Glorioso C, Oh S, Douillard GG, Sibille E. Brain molecular aging, promotion of neurological disease and modulation by Sirtuin5 longevity gene polymorphism. Neurobiol Dis. 2010.

59  

 

doi:10.1016/j.nbd.2010.09.016.

32. Lee CK, Weindruch R, Prolla T a. Gene-expression profile of the ageing brain in mice. Nat Genet. 2000;25(3):294-297. doi:10.1038/77046.

33. Lu T, Pan Y, Kao SY, et al. Gene regulation and DNA damage in the ageing human brain. Nature. 2004;429(6994):883-891. doi:10.1038/nature02661.

34. Jiang CH, Tsien JZ, Schultz PG, Hu Y. The effects of aging on gene expression in the hypothalamus and cortex of mice. Proc Natl Acad Sci U S A. 2001;98(4):1930-1934. doi:10.1073/pnas.98.4.1930.

35. Blalock EM, Chen K-C, Sharrow K, et al. Gene microarrays in hippocampal aging: statistical profiling identifies novel processes correlated with cognitive impairment. J Neurosci. 2003;23(9):3807-3819. doi:23/9/3807 [pii].

36. McKinney B, Oh H, Sibille E. Age-by-disease biological interactions: implications for late-life depression. Front Genet. 2012. http://www.sibille.pitt.edu/Publications/2012_McKinneyOhSibille.pdf. Accessed February 9, 2016.

37. Yankner BA, Lu T, Loerch P. The Aging Brain. Annu Rev Pathol. 2008.

38. Kirischuk S, Verkhratsky A. Calcium homeostasis in aged neurones. Life Sci. 1996;59(5-6):451-459. http://www.ncbi.nlm.nih.gov/pubmed/8761333. Accessed July 5, 2016.

39. van den Pol AN. Neuropeptide transmission in brain circuits. Neuron. 2012;76(1):98-115. doi:10.1016/j.neuron.2012.09.014.

40. Leng G, Ludwig M. Neurotransmitters and peptides: whispered secrets and public announcements. J Physiol. 2008;586(Pt 23):5625-5632. doi:10.1113/jphysiol.2008.159103.

41. Ludwig M, Leng G. Dendritic peptide release and peptide-dependent behaviours. Nat Rev Neurosci. 2006;7(2):126-136. doi:10.1038/nrn1845.

42. Ludwig M, Pittman QJ. Talking back: Dendritic neurotransmitter release. Trends Neurosci. 2003;26(5):255-261. doi:10.1016/S0166-2236(03)00072-9.

43. Seney ML, Tripp A, McCune S, A. Lewis D, Sibille E. Laminar and cellular analyses of reduced somatostatin gene expression in the subgenual anterior cingulate cortex in major depression. Neurobiol Dis. 2015;73:213-219. doi:10.1016/j.nbd.2014.10.005.

44. Packer AM, McConnell DJ, Fino E, Yuste R. Axo-Dendritic Overlap and Laminar Projection Can Explain Interneuron Connectivity to Pyramidal Cells. Cereb Cortex. 2012;23(12):2790-2802. doi:10.1093/cercor/bhs210.

45. Fino E, Yuste R. Dense inhibitory connectivity in neocortex. Neuron. 2011;69(6):1188-1203. doi:10.1016/j.neuron.2011.02.025.

46. Ma Y, Hu H, Berrebi AS, Mathers PH, Agmon A. Distinct subtypes of somatostatin-containing neocortical interneurons revealed in transgenic mice. J Neurosci. 2006;26(19):5069-5082. doi:10.1523/JNEUROSCI.0661-06.2006.

47. Pfeffer CK, Xue M, He M, Huang ZJ, Scanziani M. Inhibition of inhibition in visual cortex: the logic of connections between molecularly distinct interneurons. Nat Neurosci. 2013;16(8):1068-1076. doi:10.1038/nn.3446.

60  

 

48. Fino E, Packer AM, Yuste R. The logic of inhibitory connectivity in the neocortex. Neuroscientist. 2013;19(3):228-237. doi:10.1177/1073858412456743.

49. Tripp A, Kota RS, Lewis DA, Sibille E. Reduced somatostatin in subgenual anterior cingulate cortex in major depression. Neurobiol Dis. 2011;42(1):116-124. doi:10.1016/j.nbd.2011.01.014.

50. Bu J, Sathyendra V, Nagykery N, Geula C. Age-related changes in calbindin-D 28k , calretinin, and parvalbumin- immunoreactive neurons in the human cerebral cortex. Exp Neurol. 2002. doi:10.1016/S0014-4886(03)00094-3.

51. Ang LC, Shul DD. Peptidergic neurons of subcortical white matter in aging and Alzheimer’s brain. Brain Res. 1995;674(2):329-335. doi:10.1016/0006-8993(95)00034-N.

52. Khachaturian ZS. Diagnosis of Alzheimer’s Disease. Arch Neurol. 1985;42(11):1097-1105. doi:10.1001/archneur.1985.04060100083029.

53. Yáñez M, Gil-Longo J, Campos-Toimil M. Calcium binding proteins. Adv Exp Med Biol. 2012;740:461-482. doi:10.1007/978-94-007-2888-2_19.

54. Sohal VS, Zhang F, Yizhar O, Deisseroth K. Parvalbumin neurons and gamma rhythms enhance cortical circuit performance. Nature. 2009;459(7247):698-702. doi:10.1038/nature07991.

55. Cardin JA, Carlén M, Meletis K, et al. Driving fast-spiking cells induces gamma rhythm and controls sensory responses. Nature. 2009;459(7247):663-667. doi:10.1038/nature08002.

56. Sibille E, French B. Biological substrates underpinning diagnosis of major depression. Int J Neuropsychopharmacol. 2013. doi:10.1017/S1461145713000436.

57. Kerman IA. New insights into BDNF signaling: relevance to major depression and antidepressant action. Am J Psychiatry. 2012;169(11):1137-1140. doi:10.1176/appi.ajp.2012.12081053.

58. Kishimoto J, Tsuchiya T, Cox H, Emson PC, Nakayama Y. Age-related changes of calbindin-D28k, calretinin, and parvalbumin mRNAs in the hamster brain. Neurobiol Aging. 1998. doi:10.1016/S0197-4580(97)00166-8.

59. Rocco BR, Sweet RA, Lewis DA, Fish KN. GABA-Synthesizing Enzymes in Calbindin and Calretinin Neurons in Monkey Prefrontal Cortex. Cereb Cortex. 2015. doi:10.1093/cercor/bhv051.

60. Fung SJ, Fillman SG, Webster MJ, Weickert CS. Schizophrenia and bipolar disorder show both common and distinct changes in cortical interneuron markers. Schizophr Res. 2014;155:26-30. doi:10.1016/j.schres.2014.02.021.

61. Chow A, Erisir A, Farb C, et al. K(+) channel expression distinguishes subpopulations of parvalbumin- and somatostatin-containing neocortical interneurons. J Neurosci. 1999;19(21):9332-9345. http://www.ncbi.nlm.nih.gov/pubmed/10531438. Accessed July 27, 2016.

62. Maffei A, Nelson SB, Turrigiano GG. Selective reconfiguration of layer 4 visual cortical circuitry by visual deprivation. Nat Neurosci. 2004;7(12):1353-1359. doi:10.1038/nn1351.

63. Lionel AC, Vaags AK, Sato D, et al. Rare exonic deletions implicate the synaptic organizer Gephyrin (GPHN) in risk for autism, schizophrenia and seizures. Hum Mol Genet. 2013;22(10):2055-2066. doi:10.1093/hmg/ddt056.

61  

 

64. Zurek AA, Kemp SWP, Aga Z, et al. α 5GABA A receptor deficiency causes autism-like behaviors. Ann Clin Transl Neurol. 2016;3(5):392-398. doi:10.1002/acn3.303.

65. Pinto L, Dan Y. Cell-Type-Specific Activity in Prefrontal Cortex during Goal-Directed Behavior. Neuron. 2015;87(2):437-450. doi:10.1016/j.neuron.2015.06.021.

66. Waldvogel HJ, Baer K, Snell RG, During MJ, Faull RLM, Rees MI. Distribution of gephyrin in the human brain: an immunohistochemical analysis. Neuroscience. 2003;116(1):145-156. doi:10.1016/S0306-4522(02)00550-X.

67. Wang H, Turner JG, Ling L, Parrish JL, Hughes LF, Caspary DM. Age-related changes in glycine receptor subunit composition and binding in dorsal cochlear nucleus. Neuroscience. 2009;160(1):227-239. doi:10.1016/j.neuroscience.2009.01.079.

68. Pinto JGA, Hornby KR, Jones DG, Murphy KM. Developmental changes in GABAergic mechanisms in human visual cortex across the lifespan. Front Cell Neurosci. 2010;4:16. doi:10.3389/fncel.2010.00016.

69. Li M-L, Hu X-Q, Li F, Gao W-J. Perspectives on the mGluR2/3 agonists as a therapeutic target for schizophrenia: Still promising or a dead end? Prog Neuro-Psychopharmacology Biol Psychiatry. 2015;60:66-76. doi:10.1016/j.pnpbp.2015.02.012.

70. Aronica E, Gorter JA, Ijlst-Keizers H, et al. Expression and functional role of mGluR3 and mGluR5 in human astrocytes and glioma cells: opposite regulation of glutamate transporter proteins. Eur J Neurosci. 2003;17(10):2106-2118. http://www.ncbi.nlm.nih.gov/pubmed/12786977. Accessed July 5, 2016.

71. Egan MF, Straub RE, Goldberg TE, et al. Variation in GRM3 affects cognition, prefrontal glutamate, and risk for schizophrenia. Proc Natl Acad Sci U S A. 2004;101(34):12604-12609. doi:10.1073/pnas.0405077101.

72. Rothstein JD, Dykes-Hoberg M, Pardo CA, et al. Knockout of glutamate transporters reveals a major role for astroglial transport in excitotoxicity and clearance of glutamate. Neuron. 1996;16(3):675-686. doi:10.1016/s0896-6273(00)80086-0.

73. Harrison PJ, Lyon L, Sartorius LJ, Burnet PWJ, Lane TA. The group II metabotropic glutamate receptor 3 (mGluR3, mGlu3, GRM3): expression, function and involvement in schizophrenia. J Psychopharmacol. 2008;22(3):308-322. doi:10.1177/0269881108089818.

74. Crook J, Akil M, Law B, Hyde T, Kleinman J. Comparative analysis of group II metabotropic glutamate receptor immunoreactivity in Brodmann’s area 46 of the dorsolateral prefrontal cortex from patients with schizophrenia and normal subjects. Mol Psychiatry. 2002;7:157-164. doi:10.1038/sj/mp/4000966.

75. Carson MJ, Thrash JC, Walter B. The cellular response in neuroinflammation: The role of leukocytes, microglia and astrocytes in neuronal death and survival. Clin Neurosci Res. 2006;6(5):237-245. doi:10.1016/j.cnr.2006.09.004.

76. Farina C, Aloisi F, Meinl E. Astrocytes are active players in cerebral innate immunity. Trends Immunol. 2007;28(3):138-145.

77. Gomes FCA, Paulin D, Moura Neto V. Glial fibrillary acidic protein (GFAP): modulation by growth factors and its implication in astrocyte differentiation. Brazilian J Med Biol Res. 1999;32(5):619-631. doi:10.1590/S0100-879X1999000500016.

62  

 

78. Messing A, Head MW, Galles K, Galbreath EJ, Goldman JE, Brenner M. Fatal encephalopathy with astrocyte inclusions in GFAP transgenic mice. Am J Pathol. 1998;152(2):391-398. http://www.ncbi.nlm.nih.gov/pubmed/9466565. Accessed July 5, 2016.

79. Yang Y, Vidensky S, Jin L, et al. Molecular comparison of GLT1+ and ALDH1L1+ astrocytes in vivo in astroglial reporter mice. Glia. 2011;59(2):200-207. doi:10.1002/glia.21089.

80. Anthony TE, Heintz N. The folate metabolic enzyme ALDH1L1 is restricted to the midline of the early CNS, suggesting a role in human neural tube defects. J Comp Neurol. 2007;500(2):368-383. doi:10.1002/cne.21179.

81. Niciu MJ, Henter ID, Sanacora G, Zarate CA. Glial abnormalities in substance use disorders and depression: Does shared glutamatergic dysfunction contribute to comorbidity? World J Biol Psychiatry. 2014;15(1):2-16. doi:10.3109/15622975.2013.829585.

82. Colangelo AM, Alberghina L, Papa M. Astrogliosis as a therapeutic target for neurodegenerative diseases. Neurosci Lett. 2014;565:59-64. doi:10.1016/j.neulet.2014.01.014.

83. Salthouse TA, Atkinson TM, Berish DE. Executive Functioning as a Potential Mediator of Age-Related Cognitive Decline in Normal Adults. J Exp Psychol. 2003. doi:10.1037/0096-3445.132.4.566.

84. Persson J, Sylvester C-YC, Nelson JK, Welsh KM, Jonides J, Reuter-Lorenz PA. Selection requirements during verb generation: differential recruitment in older and younger adults. Neuroimage. 2004;23(4):1382-1390. doi:10.1016/j.neuroimage.2004.08.004.

85. Logan JM, Sanders AL, Snyder AZ, et al. Under-recruitment and nonselective recruitment: dissociable neural mechanisms associated with aging. Neuron. 2002;33(5):827-840. doi:10.1016/s0896-6273(02)00612-8.

86. Yankner B a, Lu T, Loerch P. The aging brain. Annu Rev Pathol. 2008;3:41-66. doi:10.1146/annurev.pathmechdis.2.010506.092044.

87. Douillard-Guilloux G, Guilloux J-P, Lewis DA, Sibille E. Anticipated brain molecular aging in major depression. Am J Geriatr Psychiatry. 2013;21(5):450-460. doi:10.1097/JGP.0b013e318266b7ad.

88. Christensen H. What cognitive changes can be expected with normal ageing? Aust N Z J Psychiatry. 2001;35(6):768-775. doi:10.1046/j.1440-1614.2001.00966.x.

89. McKinney BC, Sibille E. The age-by-disease interaction hypothesis of late-life depression. Am J Geriatr Psychiatry. 2013;21(5):418-432. doi:10.1097/JGP.0b013e31826ce80d.

90. Taylor S, Wakem M, Dijkman G, Alsarraj M, Nguyen M. A practical approach to RT-qPCR-Publishing data that conform to the MIQE guidelines. Methods. 2010;50(4):S1-S5. doi:10.1016/j.ymeth.2010.01.005.

91. Lefever S, Hellemans J, Pattyn F, et al. RDML: structured language and reporting guidelines for real-time quantitative PCR data. Nucleic Acids Res. 2009;37(7):2065-2069. doi:10.1093/nar/gkp056.

92. Bustin SA, Benes V, Garson JA, et al. The MIQE guidelines: minimum information for publication of quantitative real-time PCR experiments. Clin Chem. 2009;55(4):611-622. doi:10.1373/clinchem.2008.112797.

63  

 

93. Altman DG, Bland JM. How to randomise. BMJ. 1999;319(7211):703-704. http://www.ncbi.nlm.nih.gov/pubmed/10480833. Accessed July 23, 2016.

94. Altman DG, Bland JM. Standard deviations and standard errors. BMJ. 2005;331(7521):903. doi:10.1136/bmj.331.7521.903.

95. Yuan JS, Reed A, Chen F, Stewart CN. Statistical analysis of real-time PCR data. BMC Bioinformatics. 2006;7:85. doi:10.1186/1471-2105-7-85.

96. Thornton B, Basu C. Real-time PCR (qPCR) primer design using free online software. Biochem Mol Biol Educ. 2011;39(2):145-154. doi:10.1002/bmb.20461.

97. Zhang J, Schindler T, Küng E, et al. Highly sensitive amplicon-based transcript quantification by semiconductor sequencing. BMC Genomics. 2014;15(1):565. doi:10.1186/1471-2164-15-565.

98. Tripp A, Oh H, Guilloux J-P, Martinowich K, Lewis DA, Sibille E. Brain-Derived Neurotrophic Factor Signaling and Subgenual Anterior Cingulate Cortex Dysfunction in Major Depressive Disorder. Am J Psychiatry. 2012;169(11):1194-1202. doi:10.1176/appi.ajp.2012.12020248.

99. Loerch PM, Lu T, Dakin KA, et al. Evolution of the Aging Brain Transcriptome and Synaptic Regulation. Orban L, ed. PLoS One. 2008;3(10):e3329. doi:10.1371/journal.pone.0003329.

100. Van Den Eeden SK, Tanner CM, Bernstein AL, et al. Incidence of Parkinson’s disease: variation by age, gender, and race/ethnicity. Am J Epidemiol. 2003;157(11):1015-1022. http://www.ncbi.nlm.nih.gov/pubmed/12777365. Accessed June 25, 2016.

101. Tohen M, Bromet E, Murphy JM, Tsuang MT. Psychiatric epidemiology. Harv Rev Psychiatry. 2000;8(3):111-125. doi:10.1191/096228098675787079.

102. Mattay VS, Goldberg TE, Sambataro F, Weinberger DR. Neurobiology of cognitive aging: insights from imaging genetics. Biol Psychol. 2008;79(1):9-22. doi:10.1016/j.biopsycho.2008.03.015.

103. Gómez-Isla T, Hollister R, West H, et al. Neuronal loss correlates with but exceeds neurofibrillary tangles in Alzheimer’s disease. Ann Neurol. 1997;41(1):17-24. doi:10.1002/ana.410410106.

104. Lin LC, Sibille E. Somatostatin, neuronal vulnerability and behavioral emotionality. Mol Psychiatry. 2015;20(3):377-387. doi:10.1038/mp.2014.184.

105. Fiske A, Wetherell JL, Gatz M. Depression in Older Adults. Annu Rev Clin Psychol. 2009. doi:10.1146/annurev.clinpsy.032408.153621.

106. Alexopoulos GS. Depression in the elderly. Lancet. 2005;365(9475):1961-1970. doi:10.1016/S0140-6736(05)66665-2.

107. Alexopoulos G, Schultz S, Lebowitz B. Late-life depression: a model for medical classification. Biol Psychiatry. 2005. http://ac.els-cdn.com/S0006322305005780/1-s2.0-S0006322305005780-main.pdf?_tid=48867c5c-be1c-11e5-afd8-00000aab0f6c&acdnat=1453146619_dfa3385d7f7f0cac6e5d7f4c901ec5e4. Accessed January 18, 2016.

64  

 

BLANK