SlideShare a Scribd company logo
THE IMPLICATONS OF DESYNCHRONIZED CIRCADIAN
    RHYTHMS IN HUMAN MENTAL HEALTH AND
   SUSCEPTIBILITY IN THE AGING POPULATION


                 Cristina Corlito
                   March 2010



           Supervisor: Dr. Lakin-Thomas
              Advisor: Dr. Unniappan
             Course Director: Dr. Noel




            B.Sc. Honours Thesis

               York University
      Faculty of Science and Engineering
            Department of Biology
2


                                          Abstract
        The organization of the mammalian circadian system relies on temporal order

between behavioural and physiological rhythms that are critical to the normal functioning

of the body and human health. The hypothesis proposed is that a disruption in the sleep-

wake cycle reflects impaired circadian clock functioning, which synergistically leads to

the progression and maintenance of a variety of psychiatric disorders.            The aging

population is most susceptible to the depletion of chronobiological rhythms and sleep

deficits, and thus, the development of psychiatric disorders in the elderly warrants

attention.   By evaluating previous and current literature, it was found that internal

temporal disorder in humans may result from both internal and external factors that

disrupt the coordinated symphony of the SCN and peripheral oscillators. Sleep disorders

and neuropsychiatric illnesses transpire as a result of this chronodisruption. Evidence

suggests that sleep disturbances are a causal factor of psychiatric illness, rather than being

mere complications. It is proposed that senescence not only predisposes the elderly to

chronodisruption and sleep deficits, but also increases their risk for developing frequently

comorbid psychiatric illnesses. Increasing public awareness of the multitude of strategies

available for harmonious synchronization and optimal well-being are profitable to the

elderly in preventing circadian malfunction.
3


                                                         Table of Contents


Introduction ....................................................................................................................... 4
      Introduction and Overview of Biological Rhythms in Mammals: The Origin and Nature of
      Periodicities ..............................................................................................................................4
      The Light-Dark Cycle, Photoreceptors and the Retinohypothalamic Tract ..............................5
      The Suprachiasmatic Nucleus and its Neural Outputs .............................................................7
      Core Clock Molecular Mechanisms ..........................................................................................9
Review of Literature ....................................................................................................... 12
      The Impact of Molecular Clocks on Human Physiology, Behaviour and Neuronal Function:
      Circadian Regulation of Physiological Pathways ................................................................... 12
      Peripheral Oscillators ............................................................................................................ 14
      Clock Mechanism Disruptions and Internal Desynchrony Lead to Disease .......................... 16
      Clocks and Circadian Sleep Disorders .................................................................................... 20
      Clocks and Psychiatric Disorders ........................................................................................... 25
Synthesis and Summary ................................................................................................. 32
      The Coalescence of Circadian Rhythms and Sleep Disorders and Their Synergistic
      Neurobehavioural Consequences ......................................................................................... 32
      Clocks and Aging: The Ensuing Susceptibility to Internal Desynchrony ............................... 33
      The Prevalence of Sleep Disturbances in the Aging Population ............................................ 39
      The Depletion of Chronobiological Rhythms and the Development of Psychiatric Disorders
      with Age ................................................................................................................................. 41
      Possible Treatments and Chronobiotics for Circadian Dysfunction ...................................... 44
Research Proposal........................................................................................................... 50
      Alleviating Sleep Disorders to Alleviate Psychiatric Disturbances ........................................ 50
Acknowledgments ........................................................................................................... 55
References ........................................................................................................................ 56
4


Introduction

Introduction and Overview of Biological Rhythms in Mammals: The
Origin and Nature of Periodicities

       Cellular biology is organized in a temporal manner, with overt circadian

organization pervading all cells of the mammalian system.         Virtually all organisms

exhibit behaviours that follow circadian cycles of rhythmicity, allowing them to operate

in synchrony with the environment. Such rhythmicity is the function of a biological

clock that is endogenous to the organism (Aschoff, 1965). Several lines of evidence

demonstrate that these biological clocks are inherent to living systems.         First and

foremost is the fact that behaviours continue to cycle in the absence of environmental

time cues, negating the idea that rhythmicity is simply a reflexive response to

periodicities in the environment (Aschoff, 1965). Biological oscillations are defined by

periods, measured as the amount of time between two identical phases of behaviour

(Aschoff, 1965).    Additional supporting evidence is observed when an organism is

exposed to an asynchronous environment lacking external time cues, such as continuous

darkness, for example, where it will reveal behavioural rhythms with a periodicity of

approximately twenty-four hours (Aschoff, 1965).        These rhythms are appropriately

termed ‘circadian rhythms,’ derived from the Latin terms circa, meaning approximately,

and dies meaning day (Aschoff, 1965). Therefore, a circadian clock shows endogenous

unremitting oscillations that free-run under constant conditions and displays a periodicity

of about twenty-four hours (Aschoff, 1965).

       In a synchronized state, the circadian rhythm of a mammal is entrained to the

rotation of the earth about its axis through external time cues known as Zeitgebers, of

which the light-dark cycle is dominant, followed by temperature (Aschoff, 1965). One
5


theory on the origin of periodicities postulates that life originated on Earth in the face of

bombardment by cosmic rays (He et al., 2000). Those cells that attempted to replicate

during daylight were destroyed, while those replicating at night when radiation was at a

minimum proliferated (He et al., 2000). An archetypal biological clock arose in order to

confine DNA replication to the dark period as cells exploited these periodic signals for

their survival (He et al., 2000). Zeitgebers, therefore, maintain a sense of harmony

between the periodicity of mammals and that of the environment. Internal biological

timekeeping mechanisms in mammals allow them to anticipate those physiological states

which are best suited to responding to future environmental events (Aschoff, 1965).

Without Zeitgebers, circadian rhythms drift out of phase with the environment (Aschoff,

1978).    In addition to the aforementioned properties of circadian clocks, clocks are

temperature compensated, maintaining constant periodicities despite changes in

physiological temperatures. The mechanisms underlying this process, however, are as of

yet unknown.


The Light-Dark Cycle, Photoreceptors and the Retinohypothalamic Tract

         The sustained cyclical nature of biological systems in spite of a lack of Zeitgebers

demonstrates that these rhythms are produced by an endogenous circadian clock and not

by the daily periodicities of the environment. In mammals, the so-called master circadian

clock is located in the suprachiasmatic nucleus (SCN) of the anterior hypothalamus

(Lucas et al., 2001). The key element of circadian rhythms is the ability of the clock to

be resynchronized by photoentrainment, that is, the ability to coordinate internal

physiological rhythms with external rhythms (Lucas et al., 2001). This is done by

consulting with changes in light luminance and wavelength over the period of a twenty-
6


four hour day and adjusting accordingly with phase delays or advances, allowing

mammals to perform behaviours at suitable times in correlation with different seasons

and different locations (Lucas et al., 2001). In mammals, light is the dominant Zeitgeber

for entrainment, with the retinohypothalamic tract (RHT) as the main circuit through

which light information reaches the SCN (Hattar et al., 2002). Lucas et al. (2001)

demonstrated the lack of contribution by rod and cone photoreceptors to circadian

photosensitivity by generating rodless coneless mice by combining a cl transgene with a

diphtheria-based toxin rdta. These rodless coneless mice continued to display phase

shifts of locomotor activity and suppression of pineal melatonin at night in response to

light pulses (Lucas et al., 2001). As a result, an alternative retinal component was

thought to be necessary for photoreception.

       A subset of retinal ganglion cells (RGCs) containing the photopigment

melanopsin form the retinohypothalamic tract that projects to the suprachiasmatic nucleus

(Fig. 1) (Hattar et al., 2002). Hattar et al. (2002) established melanopsin as a circadian

photoreceptor by injecting RGCs with Lucifer Yellow for fluorescent labeling and then

staining them for melanopsin immunoreactivity. Melanopsin-positive RGCs appeared to

be directly sensitive to light and innervated neurons of the SCN, which then managed the

non-visual lighting information and aligned the circadian clock accordingly (Hattar et al.,

2002). RGCs thus form a large photosensitive receptive field within the outer nuclear

layer of the mammalian retina that detects environmental illumination, permitting

favourable physiological and neurobiological responses (Hattar et al., 2002).
7




Figure 1      The retinohypothalamic tract and photic input pathway in the circadian
timekeeping system of mammals. Intrinsically photosensitive retinal ganglion cells
containing the photopigment melanopsin are located in the outer nuclear layer of the
mammalian retina. The axons of RGCs transmit information on the light-dark cycle in
the environment directly to the SCN via the RHT (Kavakli and Sancar, 2002).

The Suprachiasmatic Nucleus and its Neural Outputs

       Numerous studies have confirmed the suprachiasmatic nucleus of the anterior

hypothalamus as the central circadian oscillator in mammals. Lesion and transplantation

studies performed by Ralph et al. (1990) produced arrhythmic mice through the ablation

of the SCN, resulting in locomotor and metabolic activities that lacked periodicities of

any kind. Subsequent tissue transplantation of a wild-type donor SCN into the host

resulted in the restoration of circadian rhythmicity as seen in overt behaviours. Through

neural projections and hormonal signals the SCN synchronizes other central oscillators in

the brain, whose efferent projections then coordinate the physiological activity of target

organs directly or indirectly (Panda and Hogenesch, 2004; Reiter, 1991). Projections

from the SCN to the subparaventricular zone of the hypothalamus are transmitted to the

medial preoptic region, which conducts thermoregulation in a circadian manner (Moore,
8


1983; Panda and Hogenesch, 2004). The subparaventricular zone also projects onto the

dorsomedial nucleus of the hypothalamus, which monitors fluctuating hormone levels

and the cycle of sleep and wakefulness (Moore, 1983; Panda and Hogenesch, 2004).

       Once external light-dark cues have reached the suprachiasmatic nucleus,

responses are evoked in the dorsomedial nucleus and subsequently the paraventricular

nucleus (PVN) (Klein et al., 1971). Neurons of the PVN synapse onto the preganglionic

sympathetic neurons in the intermediolateral zone of the lateral horns of the thoracic

spinal cord (Klein et al., 1971; Perreau-Lenz et al., 2003). These preganglionic neurons

influence neurons in the superior cervical ganglia whose efferent fibres innervate the

pineal gland (Fig. 2) (Klein et al., 1971). The pineal gland produces the neurohormone

melatonin from tryptophan and secretes it into the bloodstream, allowing it to mediate the

brainstem circuits that control the sleep-wake cycle as it promotes sleep (Perreau-Lenz et

al., 2003). During scotophase, or the dark period of the light-dark cycle, melatonin

secretion reaches a maximum since the SCN has reduced activation, thus removing the

inhibition of sympathetic neurons and promoting melatonin production in the pineal

gland (Perreau-Lenz et al., 2003). It follows that the neuronal and hormonal outputs of

the SCN to autonomous oscillators in the brain and their ensuing projections coordinate

the proper timing of a variety of physiological functions over the twenty-four hour

geophysical day, including but not limited to the sleep-wake control system, hormone

secretion, hunger and satiety and body temperature (Panda and Hogenesch, 2004).
9




Figure 2     Brain anatomy involved in the production of the neurohormone melatonin.
Non-visual light information, or lack thereof, is transmitted to the SCN and induces a
response in the PVN. Neurons of the PVN innervate the preganglionic sympathetic
neurons of the thoracic spinal cord, which then synapse onto the superior cervical ganglia
whose efferent fibres innervate the pineal gland. The pineal gland is the production site
for melatonin and mediates its release into the bloodstream (Lewy, 2010).

Core Clock Molecular Mechanisms

       Mammalian circadian rhythms of sleeping and waking, hormone secretion,

thermoregulation and the like are regulated by the molecular circadian clock mechanism,

intrinsic to every cell of the suprachiasmatic nucleus of the anterior hypothalamus and

every other cell of the human body. Historically, Drosophila melanogaster has been the

model organism in which many of the major circadian clock genes were first identified

(Clayton et al., 2001). Many homologues of the genes and proteins involved in the

generation of biological oscillations in Drosophila have now been cloned in mammals

(Clayton et al., 2001). The actions of these genes and proteins are temporally regulated,

thus giving rise to a mammalian molecular core clock mechanism that consists of
10


transcription-translation autoregulatory feedback loops with both excitatory and

inhibitory components (Fig. 3) (Clayton et al., 2001; Shearman et al., 2000).         The

fundamental factors in this molecular mechanism are the transcription factors CLOCK

and BMAL1, with helix-loop-helix and PAS domains that are indicative of their function

(Shearman et al., 2000). The PAS domains allow these two transcription factors to

dimerize through protein-protein interactions (Clayton et al., 2001).

       In the circadian processes of the core clock molecular mechanism, Clk and Bmal1

genes are transcribed and their protein products heterodimerize when adequate

concentration levels are reached (Shearman et al., 2000).          These dimers bind to

regulatory DNA sequences, or E-boxes, that initiate the transcription of the mammalian

cryptochrome genes Cry1 and Cry2, period genes Per1, Per2, and Per3 and clock-

controlled genes Ccg (Clayton et al., 2001; Shearman et al., 2000). Following post-

translational modifications, cytoplasmic PER2 and CRY heterodimerize and translocate

to the nucleus, where PER2 proteins stimulate the synthesis of BMAL1, in antiphase with

the concentration of the three PER proteins, thereby acting as a positive regulator of the

BMAL1 loop (Shearman et al., 2000). Conversely, CRY proteins bind to CLOCK-

BMAL1 dimers, inhibiting the stimulatory effect they exert on the DNA sequences

encoding PER and CRY, and forming a negative feedback loop (Shearman et al., 2000).

Consequently, as the concentrations of PER and CRY decrease, the inhibition is removed

and CLOCK-BMAL1 dimers recommence transcription and maintain the twenty-four

hour oscillation of the circadian clock mechanism (Clayton et al., 2001).

       The characteristic twenty-four hour periodicity of the circadian clock mechanism

is generated through post-translational modifications, phosphorylation and cellular
11


localization and stability of protein products, leading to time delays (Ikeda et al., 2003).

Cytosolic factors, such as ions and second messengers, exhibit biological oscillations and

support transcription-translation feedback loops (Ikeda et al., 2003). In this fashion, each

standalone circadian clock cell is composed of a molecular oscillator and the rhythm

maintaining effects of post-translational mechanisms. Further investigation of this

process and other clock components is currently underway. Overall, the molecular clock

produces a synchronized rhythmic output from the SCN, which is then conveyed

synaptically and humorally to other brain oscillators and peripheral tissues (Clayton et

al., 2001).   In this manner, the output of the master oscillator coordinates various

physiological and behavioural mammalian systems.




Figure 3 The molecular oscillator within circadian clock cells of the SCN.
Transcription-translation feedback loops with excitatory and inhibitory components
regulate internal rhythmicity within the cell (Clayton et al., 2001).
12


       Mammalian physiological and behavioural systems exhibit biological oscillatory

patterns that harmonize internal conditions with the rhythmic external world. This review

will focus on the implications of desynchronized circadian rhythms in human mental

health. The hypothesis to be proposed is that a disruption in the sleep-wake cycle reflects

impaired circadian clock functioning, which synergistically leads to the progression and

maintenance of psychiatric disorders. The aging population is most susceptible to the

depletion of chronobiological rhythms, and thus, the development of psychiatric disorders

in the elderly merits attention. This will be done by evaluating previous and current

literature and proposing future directions.


Review of Literature

The Impact of Molecular Clocks on Human Physiology, Behaviour and
Neuronal Function: Circadian Regulation of Physiological Pathways

       The endogenous circadian clock permeates through almost every physiological

and behavioural process in the human body and has wide implications for health.

Humans exhibit circadian rhythmicity in such behaviours and physiological processes as

sleep and wakefulness, hormone secretion, thermoregulation, feeding, metabolism,

attentiveness and memory (Clayton et al. 2001; Takahashi et al., 2008). The clock genes

that sustain biological oscillations in the suprachiasmatic nucleus also regulate the

corresponding twenty-four hour human cycle through rhythmic neural and hormonal

output signals to peripheral tissues (Yoo et al., 2004). In this manner, peripheral organs

are able to modify their temporal functioning accordingly (Fig. 4) (Yoo et al., 2004).
13




Figure 4 Circadian rhythmicity in human behaviours and physiological processes.
Rhythmic solar signals entrain the cellular clocks of neurons in the SCN through afferent
retinal innervations. Neural and hormonal outputs from the SCN subsequently adjust the
phase of peripheral organs throughout the body.

       The SCN employs both neural efferents and humoral signals to entrain other brain

oscillators, whose roles in coordinating various physiological processes are crucial (Silver

et al., 1996; Abe et al., 2002). Recall the abovementioned projections from the SCN to

the subparaventricular zone, medial preoptic nucleus, and dorsomedial hypothalamus,

which manage endocrine and autonomic systems, including the hypothalamic-pituitary-

adrenal axis, through supplementary neuronal projections and hormonal signals (Moore et

al., 1983; Girotti et al., 2007). The neurotransmitters glutamate and GABA mediate

synaptic transmission between the SCN and other hypothalamic areas (Hermes et al.,

1996). Using an in vitro postmortem anterograde tracing method, Jiapei and colleagues

(1998) found that the hypothalamic areas innervated by the SCN mediate

parasympathetic and sympathetic signaling centres, the sleep-wake control system, the

arousal system, locomotor activity, body temperature, cardiovascular activity and

hormone secretion. Therefore, the circadian regulation of the synthesis and release of

central nervous system neurotransmitters and neuropeptides, and the ensuing entrainment
14


of brain oscillators, ensures the synchrony of tissue rhythms with the twenty-four hour

geophysical day.

       The outputs of the SCN master clock not only coordinate brain oscillators but also

peripheral organs in order to orchestrate physiological oscillations.       In addition to

hormonal cues, direct neural control of peripheral targets is achieved via the autonomic

nervous system (Cailotto et al., 2005). The SCN may also indirectly control the phase of

peripheral oscillators by regulating the sleep-wake cycle and therefore the rhythm of

feeding behaviour (Cailotto et al., 2005). As a result, an inconsistent feeding schedule

can disturb the harmonious alignment between the SCN and the peripheral organs

involved as it acts as an external entraining agent (Cailotto et al., 2005). Therefore, the

transcription-translation feedback loops of the core clock mechanism not only maintain

rhythmicity in the central oscillator and its clock-controlled genes but also generate

circadian outputs to peripheral targets, the details of which are as of yet not fully

understood (Duffield et al., 2002).


Peripheral Oscillators

       It has been established that the clock genes expressed in the core mechanism of

the master suprachiasmatic nucleus are rhythmically expressed in peripheral circadian

oscillators located throughout the human body (Yamazaki et al., 2000; Duffield et al.,

2002; Yoo et al., 2004).      Balsalobre et al. (2000) examined clock-controlled gene

expression in peripheral mammalian tissues by inducing rhythmicity in immortalized rat-

1 fibroblasts through serum shock.        cDNA microarrays revealed a chronological

production of messenger RNA in response to serum shock and pharmacological

treatment, signifying the circadian regulation of gene expression (Balsalobre et al., 2000).
15


Previously believed to rhythmically dampen after two to seven twenty-four hour cycles

without input from the SCN, Yoo and colleagues (2004) have demonstrated that

peripheral oscillators maintain endogenous rhythmicity whilst displaying desynchrony

amongst themselves without entraining signals (Yamazaki et al., 2000). Yoo et al.

(2004) employed the fusion of the mouse locus mPer2 with a luciferase reporter gene to

reveal strong inherent oscillations of bioluminescence in both the SCN and peripheral

tissues ex vivo. Furthermore, in SCN-lesioned mice, bioluminescence rhythms persisted

for twenty days in peripheral tissues, including the liver, lungs, pituitary and cornea,

however, with an eventual loss of phase coordination (Fig. 5).




Figure 5 Circadian rhythmicity in explanted tissues of the mouse, including the cornea,
liver, pituitary gland, kidney and lung. Luciferase reporter genes revealed inherent
oscillations of bioluminescence in these tissues (Yoo et al., 2004).

       Peripheral tissues isolated in culture, including but not limited to the previously

mentioned lungs, cornea, pituitary gland and liver, express clock-controlled genes that

confer distinctive circadian period and phase properties to those structures (Yoo et al.,

2004). As such, these circadian properties are distinct in different organs and contribute

temporally to their physiological functioning (Yoo et al., 2004). The SCN does not
16


generate but rather coordinates the phase of autonomous peripheral oscillators, thereby

inhibiting internal desynchrony between tissue-specific target clocks and their

synchronized phase relationship with the external environment (Yamazaki et al., 2000;

Yoo et al., 2004).    The phase of each peripheral oscillator induces rhythmic gene

expression, for example that of Per1, resulting in circadian protein product activity,

which in turn regulates rhythmic metabolic events in different tissues throughout the

human body (Ripperger et al., 2000).       The phase of oscillations can be altered by

adjusting the feedback of peripheral clocks characteristic of a tissue to internal and

external Zeitgebers originating from the SCN and environment, respectively (Yamazaki

et al., 2000).


Clock Mechanism Disruptions and Internal Desynchrony Lead to Disease

        The organization of the mammalian circadian system, as reviewed above, relies

on temporal order between behavioural and physiological rhythms that are critical to the

normal functioning of the body and human health. Thus, the concept of the harmful

effects that would ensue as a result of disorder between these phase relationships and the

cyclical expression of clock-controlled genes readily presents itself as there are numerous

avenues through which to disrupt this fragile system (Yamazaki et al., 2000). Internal

temporal disorder in humans may result from both internal and external factors that

disrupt the coordinated symphony of the SCN and peripheral oscillators. Predominant

external factors include light deficiency and irregularity, jet lag, shift work, food intake

and social activities (Skene et al., 1999; Reddy et al., 2002; Solonin et al., 2009; Turner

and Mainster, 2008; Girotti et al., 2009). Internal factors include the disturbance of

proper photoreception, visual loss, decreased melatonin levels and circadian clock gene
17


mutations (Czeisler et al., 1995; Lockley et al., 1997; Turner and Mainster, 2008). As

previously discussed, peripheral oscillators will desynchronize amongst themselves

without temporal adjustments provided by the SCN through neural and hormonal outputs

(Yoo et al., 2004). Proper SCN operations ensure good health by mediating rhythms of

sleep-wake systems, hormone secretion and metabolism, therefore, chronodisruption may

be the cause of a range of diseases (Jiapei et al., 1998; Yamazaki et al., 2000; Cailotto et

al., 2005).

        Light irregularity, or improperly timed ocular light exposure, may result in

chronodisruption by modifying nocturnal melatonin synthesis in the pineal gland

depending on its duration, wavelength, intensity and time of administration (Czeisler et

al., 1995; Skene et al., 1999).     Ocular light exposure in the scotophase decreases

melatonin production (Skene et al., 1999).         Low circulating levels of melatonin

throughout the body may result in numerous diseases as it has been shown to contribute

beneficially to the antioxidant capability of blood plasma (Benot et al., 1999).

Environmental light is the predominant Zeitgeber in circadian timekeeping, and, for that

reason, maintains the greatest influence on human physiological and psychological

health. Photosensitive RGCs best absorb light in the blue sector of the light spectrum at

460 nm, which is quite similar to the wavelength of environmental light (Turner and

Mainster, 2008). Modern artificial lighting, unfortunately, provides only about 1% of

natural light intensity and is distinguished by red spectrum wavelengths, which is

insufficient for suitable photoreception (Turner and Mainster, 2008). Instead, optimal

photoreception requires blue light of high intensity and duration for photoentrainment and

favourable health (Turner and Mainster, 2008).
18


       Light deficiency fails to entrain the SCN to the geophysical day and results in a

subsequent free-running periodicity, as exemplified by blind individuals (Skene et al.,

1999). Blind individuals may be categorized, according to the extent of visual loss, as

having some light perception, and thus photoreception, and those with no light perception

capabilities and no photoreception whatsoever (Skene et al., 1999). In a study by Skene

et al. (1999), 77% of blind subjects capable of photoreception showed normal circadian

rhythmicity, while 67% of those with no light perception showed free-running period

lengths and internal desynchrony. The latter also suffered from daytime somnolence, an

excessive need for sleep during the daytime, and insomnia during the night due to the

temporal disorder of melatonin synthesis and release (Lockley et al., 1997). Further

evidence of the dire consequences of light deficiency, with light as the chief biological

Zeitgeber, is demonstrated by the fact that blind subjects who had no eyes after having

undergone bilateral enucleation showed free-running period lengths ranging from 24.13

to 24.81 hours, albeit in the presence of non-photic signals such as food intake and social

activities (Skene et al., 1999).      On the whole, blind individuals incapable of

photoentrainment exhibit higher levels of circadian disruption and dampened SCN

outputs, thus making them susceptible to diseases, particularly sleeping disorders and

compromised neuropsychiatric conditions (Lockley et al., 1997; Jean-Louis et al., 2005).

        Girotti and colleagues (2009) recently demonstrated the role of food intake as a

non-photic Zeitgeber.     Their study revealed characteristic rhythms of clock gene

expression in each element of the hypothalamic-pituitary-adrenal axis, where a decrease

in food intake in the photophase of the light-dark cycle altered glucocorticoid secretion

and clock gene expression (Girotti et al., 2009).       Physiological processes may be
19


entrained to intermittent feeding schedules, with glucocorticoids synchronizing a

multitude of peripheral organs (Stephan, 1986; Girotti et al., 2009).

       Shift work employees in industrial sectors, medicine and the military show signs

of considerable circadian dysfunction, including such symptoms as biochemical

disturbances, mood disorders, sleeping disorders, metabolic syndrome and an overall

feeling of malaise (Solonin et al., 2009). James et al. (2007) recently investigated the

effects of night shift work on the sleep-wake cycle, outlining desynchrony between the

master circadian clock and the night schedule as subjects maintained day active

entrainment. This was done by comparing oscillatory clock gene expression in peripheral

blood mononuclear cells with the temporally shifted sleep-wake cycle. The exposure to

artificial light at uncharacteristic times communicates odd entraining signals to the SCN

and results in perturbed circadian outputs and melatonin synthesis (James et al., 2007).

The differential responses of peripheral oscillators to the altered phase of input signals

also leads to a lack of internal coordination, hence resulting in feelings of malaise. Shift

work sleep disorder is a circadian rhythm sleep disorder in which the afflicted complain

of daytime sleepiness, insomnia and poor sleep quality (Ursin et al., 2009). Other

circadian rhythm sleep disorders will be addressed in the next section of this review.

       In addition to shift work, jet lag also impairs physical and mental well-being via

circadian desynchrony. The core clock mechanism experiences much more difficulty in

acclimatization to advanced time zones rather than delayed time zones because the

former does not occur as rapidly (Reddy et al., 2002). Reddy and his colleagues (2002)

subjected mice to acute advances or delays in local time and reported that circadian

rhythms of mPer expression in the SCN adjust swiftly to advanced light pulses, while
20


rhythmic mCry1 expression advanced gradually. Conversely, they found that a six hour

delay in local time entailed mPer and mCry adjusting in sequence by the second

oscillation.   This study describes the different effects of traveling east or west, or

advancing or delaying, respectively, on the master clock and the prospective temporal

desynchrony between mPer and mCry expression as a result of jet lag, with ensuing

health complications (Reddy et al., 2002).

        The final factor contributing to clock mechanism disruptions, internal

desynchrony, and thus, disease are genetically mutated circadian clock genes and

polymorphisms. Current research is heavily focused on identifying those clock gene

alterations that result in a variety of disrupted circadian behaviours.      Clock gene

mutations may be implicated in the deterioration of the regimented functioning of both

molecular oscillators and their rhythmic neural and hormonal outputs and their effects

will be discussed in subsequent sections of this review.             Cumulatively, the

aforementioned internal and external factors, particularly insufficient and temporally

displaced environmental light, may induce biological stress and disturb the coordinated

rhythmicity of physiological processes and behaviours necessary for optimal human

health. This review will now turn to those sleep disorders and neuropsychiatric illnesses

that transpire as a result of chronodisruption.


Clocks and Circadian Sleep Disorders

        A profound relationship exists between clock gene variations and changes in

behavioural rhythmicity, most notably sleep parameters in humans. The timing and

amount of sleep are determined by circadian and homeostatic sleep control mechanisms,

respectively (Naylor et al., 2000). The former dictates patterns of sleep and wake at
21


specific phases throughout the twenty-four hour light-dark cycle, while the latter depends

on the need for sleep (Naylor et al., 2000). In a study by Naylor et al. (2000), a mutation

in Clk in the mouse was found to alter not only the timing and length of sleep but sleep

homeostatis as well.    Naylor and colleagues evaluated the effects of the CLOCK

transcription factor mutation by comparing sleep and electroencephalographic (EEG)

activity in homozygous and heterozygous mutants and wild-type mice under conditions

of entrainment, free-running rhythms and recovery from six hours of sleep deprivation.

The results indicated that heterozygotes slept one hour less per day and homozygotes two

hours less per day in contrast to wild-type mice, with lower amounts of non-rapid eye

movement sleep seen. Following periods of sleep deprivation, Clk homozygous mice

displayed 39% less sleep than heterozygotes and wild-type mice. One may attribute these

results to discrepancies of entrainment to the light-dark cycle, however, divergent sleep

behaviours were also seen when mice were free-running in continuous darkness (Naylor

et al., 2000).   A study by Laposky et al. (2005) employed the same investigative

measures in mice with a deletion of Bmal1 and discovered a diminished rhythm of sleep

and wakefulness, a weakened response to sleep deprivation and lengthened sleep periods.

       Whereas mutations in the mammalian cryptochrome genes Cry1 and Cry2 hold

implications for sleep homeostatis, Period genes are not essential for homeostatic sleep

regulation (Wisor et al., 2002; Shiromani et al., 2004). A study by Shiromani and

collaborators (2004) examined the effects of Per1, Per2, Per3 and double Per1-Per2

mutations on sleep factors and found that Per2-mutant and double mutant mice exhibited

longer periods of wakefulness, with less slow-wave sleep (SWS) and rapid-eve

movement (REM) sleep, than wild-type and Per1 deficient mice in states of entrainment.
22


Double mutant strains became arrhythmic in aperiodic conditions, however, the amount

of time spent awake, in SWS and in REM sleep was equivalent to that in an entrained

state even after 36 days, thus signifying the maintenance of total sleep levels. Per genes

are more so involved in altering the phase position of the sleep-wake cycle (Shiromani et

al., 2004). In conclusion, circadian clock gene alterations have profound implications for

both rhythms of sleep and wakefulness and sleep propensity, although knowledge as to

their exclusivity to one or the other is currently unknown.

       Variations in circadian clock genes have a variety of effects on the configuration

of human sleep. Delayed sleep phase syndrome (DSPS) is the most commonly reported

circadian rhythm sleep disorder whose features include sleep periods delayed by 2 to 6

hours, the inability to fall asleep, difficulty waking and a lack of feeling well rested

(Campbell and Murphy, 2007; Chang et al., 2009). One study investigated a 30 year old

graduate student with DSPS whose average bedtime was 3:38 a.m. and usually awoke at

1:47p.m. in order to feel replenished (Campbell and Murphy, 2007). Campbell and

Murphy (2007) examined the sleep and body temperature rhythms of the subject in

comparison to those of 3 normal age-matched subjects with both parties in aperiodic

conditions free from environmental cues. Whereas the time between the core body

temperature minimum and sleep onset in control subjects was 1.63 hours, the graduate

student displayed a phase angle of 3.62 hours. Furthermore, the DSPS patient had a free-

running period length of 25.38 hours compared to an average of 24.44 hours for the

control subjects. One may refute these results by suggesting that the lighting conditions

in temporal isolation contributed to the lengthening of the free-running period in the

DSPS subject, however, the authors noted that illumination was below 50 lux, which is
23


insufficient for optimal photoreception (Campbell and Murphy, 2007). Therefore, DSPS

causes an abnormal endogenous period length and internal desynchrony between sleep

and body temperature rhythms, resulting in poor sleep efficiency and duration.

       The previous findings may be attributed to a polymorphism in the circadian clock

gene Per3 or a missense mutation in the casein kinase I epsilon gene CKI ε. Archer et al.

(2003) have found a correlation between a length polymorphism in Per3 and DSPS,

particularly the shorter allele for which 75% of DSPS patients were homozygous. They

found that the 4-repeat allele, as opposed to the 5-repeat allele, was substantially

prevalent in DSPS patients in contrast to the control group. Recall post-translational

mechanisms, such as phosphorylation, affect the stability of protein products and function

to create time delays in the circadian clock mechanism. PER is targeted for degradation

through phosphorylation by CKI ε, making it unavailable for dimerization and subsequent

nuclear localization and thus causing it to oscillate (Archer et al., 2003).

The shorter variation of PER3 contains fewer phosphorylation sites than its longer

counterpart and may be the cause of polymorphic differences in function and hence

longer endogenous period length seen in DSPS (Archer et al., 2003). A study by Takano

and associates (2004) revealed that a missense mutation in the N408 allele in CKI ε

functions as a safeguard against DSPS by modifying its autophosphorylation activity.

       In contrast to DSPS, advanced sleep phase syndrome (ASPS) dictates human

behaviours marked by early bedtimes, early morning waking and a short endogenous

period length (Xu et al., 2005). ASPS is caused by a mutation in a residue in the casein

kinase I binding site of the Per2 gene and results in attenuated phosphorylation levels

(Archer et al., 2003). The reduction in phosphorylation seen in both DSPS and ASPS is
24


indicative of the different pathological symptoms that may occur as a result of

phosphorylation levels in different PER proteins. In another case, through mutagenesis

screenings of related ASPS patients, Xu et al. (2005) found a threonine to alanine

missense mutation at amino acid 44 in the human CKIδ gene. Subjects under study had

an average bedtime of 6:12 p.m., compared to the control average of 11:24 p.m., and an

average rising time of 4:06 a.m. compared to the control average of 8:00 a.m. Overall,

this T44A mutation decreases CKIδ enzyme activity in ASPS patients and consequently

leads to a shortened activity rhythm and advanced phase of activity in an entrained setting

of 12 hours of light and 12 hours of dark (Xu et al., 2005).

       The etiology of the abovementioned circadian rhythm sleep disorders may be

attributed to circadian clock gene polymorphisms and mutations, whereas obstructive

sleep apnea syndrome (OSAS) and its symptoms produce arrythmicity in clock gene

functioning. This arrythmicity may be credited to fluctuating levels of factors circulating

through the blood (Burioka et al., 2008). Burioka et al. (2008) have measured Per1

mRNA expression in peripheral blood mononuclear cells in those patients with severe

OSAS using polymerase chain reaction analysis over a twenty-four hour period. In

contrast to similar healthy controls, the eight OSAS participants showed no circadian

rhythms of Per1 mRNA expression throughout the day and abnormal elevations of

plasma noradrenaline. Elevated noradrenaline levels and sympathetic activity contributed

to an increase in the transcription of Per1 during sleep (Burioka et al., 2008).

Interestingly, continuous positive airway treatment for a period of three months improved

not only shallow sleep with frequent waking due to hypoxic episodes, but also daily

oscillations of Per1 transcription (Burioka et al., 2008).     The effects of continuous
25


positive airway treatment on clock gene transcription thereby illustrate a mechanism by

which a circadian rhythm sleeping disorder may be managed by improving clock gene

function.

       Fatal familial insomnia (FFI) is a debilitating disorder marked by sleep

deficiency. FFI is a prion disease distinguished by a 178 codon prion protein gene

mutation (Reder et al., 1995). A study by Sforza et al. (1995) studied six subjects with

this disease using twenty-four hour polygraphic recordings in a sleep laboratory. Their

findings revealed severe reductions in total sleep time, impairments in the circadian

regulation of the sleep-wake cycle and abrupt alterations from wakefulness to sleep.

Positron emission topography uncovered atrophy in the thalamus, particularly the antero-

ventral and dorso-medial thalamic nuclei, which take part in regulating the sleep-wake

cycle (Sforza et al., 1995). Over the course of the disease, symptoms of insomnia

progressively worsen and circadian rhythms dampen substantially until the total sleep

time is reduced to about 50 minutes per day and the subject dies (Sforza et al., 1995).

Portaluppi et al. (1994) conducted assays for melatonin in the blood plasma of two FFI

patients and found that concentrations of the hormone decreased in accordance with

disease progression, further compromising circadian rhythmicity.


Clocks and Psychiatric Disorders

       Just as the misalignment of the circadian pacemaker and clock gene mutations and

polymorphisms have been associated with circadian rhythm sleep disorders, these factors

are implicated in psychiatric disorders. A variety of abnormal endogenous circadian

rhythms underlie major depressive disorder (MDD), particularly the sleep-wake cycle

(Gordijn et al., 1994; Emens et al., 2009). Emens and colleagues (2009) set out to
26


demonstrate a correlation between MDD and improper coordination between the

circadian pacemaker and sleeping schedule. Study subjects were comprised of eighteen

females ranging from 19 to 60 years of age who had been diagnosed with MDD

according to the Diagnostic and Statistical Manual of Mental Disorders (DSM),

excluding those with suicidal tendencies, jet lag, shift work positions and medications

that would impede melatonin production.        Emens et al. (2009) calculated circadian

misalignment according to the time difference between melatonin production and the

midpoint of sleep. Those with larger time differences exhibited a phase delay in central

pacemaker rhythmicity in comparison to the timing of sleep and a higher severity of

symptoms (Fig. 6).       These results demonstrate the interaction between circadian

desynchrony, poor sleep and mild to moderate symptoms of depression, though future

studies should be conducted on a more representative sample population (Emens et al.,

2009).




Figure 6 The larger the time difference between melatonin synthesis and the midpoint
of sleep, also known as the phase angle difference (PAD), the higher the severity of
depressive symptoms according to the Hamilton Depression Rating Scale (HAM-D).
Following a clinical assessment by a health professional, a score higher than 7 on the
HAM-D constitutes a diagnosis of MDD (Emens et al., 2009).

         An alternative route by which disrupted circadian oscillations may facilitate MDD

is through the deregulation of mood by the mesolimbic dopaminergic system (Hampp et
27


al., 2008). Hampp and associates (2008) ascertained that Per2 mutant mice have reduced

levels of expression of monoamine oxidase A in the mesolimbic dopaminergic system,

which is an enzyme that mediates dopamine metabolism. The atypical mood behaviours

observed in these mice may be attributed to this clock gene mutation.     Polymorphisms

and mutations in the Clock gene have been connected to bipolar disorder (Benedetti et al.,

2003; Roybal et al., 2007). Roybal and colleagues created Clock mutant mice through

mutagenesis, thereby inhibiting its transcriptional activation of molecular rhythms. The

mice were subjected to tests in which they were able to induce rewarding electrical

stimulation to themselves via electrodes implanted in the medial forebrain bundle. Clock

mutant mice were able to experience euphoria at lower currents than wild type mice and

cocaine decreased these current thresholds substantially in the mutants. This response is

predictive of substance abuse as the mice experienced a greater sense of reward upon

stimulation because of their hypersensitivity, making them more inclined to abuse such

stimulants (Roybal et al., 2007). These states of ecstasy and substance abuse mimic the

condition of bipolar patients (Roybal et al., 2007).

       The mood-related behaviours of Clock mutant mice parallel those humans with

bipolar disorder, including less depression and less anxiety (Roybal et al., 2007). This

was discerned as mice showed little anxiety when subjected to an unprotected arm of a

raised platform. Conversely, when treated with lithium, a mood stabilizer given to

bipolar patients, the mutant mice displayed more wild-type behaviours of high anxiety in

this situation (Roybal et al., 2007). Like the previously mentioned MDD patients, Clock

mutant mice have compromised dopaminergic systems, although with increased firing of

dopaminergic neurons that is diminished through the viral insertion of a gene coding for a
28


wild type CLOCK protein (Roybal et al., 2007). As the name implies, bipolar disorder

alternates between states of mania and depression, with depressive states being

predominant in the winter months (Roybal et al., 2007).

       Winter depression, also known as seasonal affective disorder (SAD), involves

changes in circadian genetic factors, the external environment and circulating melatonin

(Wehr et al., 2001; Johansson et al.., 2003; Partonen et al., 2007). Certain animals

display photoperiodism, that is, the ability to infer the time of year based on the length of

the day. Such information is made available by measuring the duration of melatonin

release during the night, the duration of which is longer in the winter (Wehr et al., 2001).

Wehr et al. (2001) found that variations in season affect patients with SAD but not

similar healthy subjects. Their study measured the duration of melatonin release in dim

light in 55 SAD patients and 55 equivalent healthy subjects throughout the summer and

winter months, with plasma samples being taken every 30 minutes all through the day. In

SAD subjects, melatonin release in the scotophase was much more pronounced in the

winter rather than summer, however, no change was observed in those without SAD

diagnoses.

       In regards to circadian genetic mechanisms, Partonen et al. (2007) surmised the

genes Per2, Bmal1 and Npas2, whose products function mutually in the core circadian

oscillator, are compromised in SAD. As previously mentioned, BMAL1 is a PAS protein

that interacts with other proteins, and as such, dimerizes with NPAS2 and binds to DNA

(Partonen et al., 2007). Single nucleotide polymorphisms were assessed in each of the

three genes in 189 patients and 189 symptom-free controls. Gene-wise logistic regression

analysis revealed SAD to be related to polymorphisms within all three genes and posing
29


the greatest chance of illness due to their magnified cumulative effects (Partonen et al.,

2007). When genetic variations in all three were present, patients had a 10 times greater

chance of developing SAD compared to the controls, while those with less severe allelic

combinations had a 4 times greater chance. Recently, point mutations have been located

in the melanopsin gene Opn4 in retinal ganglion cells, which serve to decrease

photosensitivity (Roecklein et al., 2009). As the contrast between light intensities is

already reduced in the winter, these point mutations aggravate that effect as dusk and

dawn Zeitgebers cannot be detected and proper photoentrainment cannot occur

(Roecklein et al., 2009). Such a hindrance in phototransduction may result in internal

desynchrony, and thus, clinical symptoms of depression in the winter months.

       Deteriorations in mood behaviours are often accompanied by substance abuse,

most likely due to the fact that circadian rhythmicity and dopaminergic systems are

confounded in these patients. Abarca et al. (2002) investigated cocaine addiction in Per

mutant mice in order to ascertain the circadian control of cocaine-induced reward and

behavioural sensitization. A single cocaine injection produced a fivefold increase in

locomotion in Per1 and Per2 knockout and wild type mice compared to saline injections.

During cocaine administration, mice were placed in boxes with two floor divisions in

which one consisted of a rod pattern of texture and the other of circles.        Cocaine

administration was always associated with the same floor division.        After repeated

cocaine injections, wild type mice became sensitized to cocaine-associated factors, Per1

mutants showed no sensitization and Per2 mutants showed an intense sensitized

behavioural reaction. While both wild type and Per2 mutant mice preferred the division

associated with cocaine injections, Per1 mutants did not prefer the side associated with
30


reward. In addition, stronger behavioural responses to the drug were seen in the morning

than at night.   Since all three groups displayed similar levels of locomotor activity in

response to a single cocaine injection, it may be inferred that cocaine addiction, rather

than acute application, is managed by the clock genes Per1 and Per2 with opposing

effects in the circadian system (Abarca et al., 2002).

       Previous discussions highlighted aberrant dopaminergic systems and clock gene

mutations in MDD and bipolar disorder mice models.           These same conditions are

observed in mice addicted to cocaine. McClung and colleagues (2005) found that a loss

of function point mutation in Clock results in the same cocaine sensitization behaviours

outlined by Abarca et al. (2002), with Clock mutants displaying a greater degree of

sensitivity to the rewarding feelings of cocaine. This may be credited to increased levels

of tyrosine hydroxylase activity, an enzyme involved in dopamine metabolism, and

therefore, heightened amounts of dopaminergic transmission in the reward centres of

mice without functional CLOCK proteins (McClung et al., 2005). Alcoholism, another

form of substance abuse possibly under circadian control in humans, may be associated

with excessive levels of glutamate in the extracellular fluid, as suggested by Per2 mutant

mice whose levels of glutamate reuptake transporters in the nervous system are

significantly reduced (Spanagel et al., 2005). These mice display augmented levels of

voluntary alcohol consumption when offered ethanol in comparison to wild type controls

(Spanagel et al., 2005).

       Malformed circadian rhythms of activity, body temperature and sleep are often

prevalent in patients with Alzheimer’s disease, which is common in the elderly. Many

victims of this neurodegenerative illness exhibit what is referred to as ‘sundowning’, or
31


the worsening of Alzheimer’s behaviours in the afternoon and evening (Volicer et al.,

2001).    Volicer et al. (2001) sought to decipher the relationship, if any, between

sundowning and circadian rhythms in a cohort of 25 Alzheimer’s patients and nine

healthy subjects. Their results revealed that those Alzheimer’s patients who undergo

sundowning showed increased nocturnal locomotor activity, with lower amplitudes

during the day, and major phase delays in both activity and body temperature in

comparison to controls. Furthermore, these subjects had severely reduced amplitudes of

body temperature and disrupted sleep parameters. These results imply that patients who

sundown may be suffering from disturbances in their rhythms, however, other

environmental factors must be considered in these habitual states of aggravation (Volicer

et al., 2001). One study that has shed some light on this issue is by Mahlberg and

associates (2008), in which cranial computed tomography revealed significant levels of

pineal calcification in AD patients, thereby limiting melatonin synthesis to sup-optimal

levels and inhibiting the coordination of the circadian system.

         Schizophrenia is a neuropsychiatric illness characterized by distorted cognition,

abnormal affect and social withdrawal (Wulff et al., 2006). Schizophrenic patients show

strong disturbances in their sleep-wake cycles, melatonin patterns and light exposure, yet

research in this area remains vague (Wulff et al., 2006). Wulff et al. (2006) studied a 27-

year old male patient for six weeks, documenting gradually delayed bedtimes and risings,

which culminated in the eventual reversal of night and day activities and low sleep

efficiency. His free-running period length was longer than 24 hours, as well as a free-

running melatonin rhythm of 24.29 hours. Instead of being coordinated with the light-

dark cycle of the environment, the patient’s light exposure was synchronized with his
32


own activity, and the temporal misalignment of these input light signals may exacerbate

his established desynchrony even further (Wulff et al., 2006).


Synthesis and Summary

The Coalescence of Circadian Rhythms and Sleep Disorders and Their
Synergistic Neurobehavioural Consequences

       There is a widely held assumption that sleep deficits are the secondary effects of

psychiatric disorders. While this may certainly be the case under some circumstances,

evidence would lead one to speculate that sleep disturbances are a causal factor of

psychiatric illness, rather than being mere complications. A disruption in the sleep-wake

cycle reflects impaired circadian clock functioning, which synergistically leads to the

progression and maintenance of a variety of psychiatric disorders. Sleep-wake cycles are

perturbed in most if not all of the previously mentioned psychiatric disorders, and many

of the studies discussed implicated irregular clock gene functioning. Recall the study by

Emens et al. (2009), whose results demonstrated a temporal misalignment between the

central circadian pacemaker and the sleep-wake cycle, the degree of which corresponded

to the severity of clinical psychiatric symptoms experienced. Over 10 years earlier,

Boivin and collaborators (1997) released similar findings from a study in which 24

healthy adults were subjected to internal desynchrony by living according to 30 hour and

28 hour sleep-wake cycles. Through the use of psychometric response scales, they too

showed varying mood states based on the degree of displacement between the sleep-wake

cycle and circadian rhythmicity.

       The successful treatment of MDD seems to involve the management of sleep

disturbances, indicating their contribution to this psychiatric disorder (Ohayon and Roth,
33


2003). In 2003, Ohayon and Roth interviewed a representative sample of the population

of the United Kingdom, Germany, Italy and Portugal, totaling 14,915 participants.

Participants were questioned about their sleep habits, sleep symptoms, current mental

health status and history and subsequently diagnosed according to the Diagnostic and

Statistical Manual of Mental Disorders if need be. Symptoms of insomnia arose in 19.1%

of the sample, with 90% of this 19.1% cohort experiencing severe insomnia in excess of

six months. Among those suffering from severe insomnia for six months to five years,

28% held concomitant psychiatric diagnoses, and of those suffering for more than five

years 25.8% held diagnoses. Furthermore, insomnia preceded states of relapse in 56.2%

of interviewees and came about concurrently in 22.1% of states of reversion. Taken as a

whole, subjects afflicted with sleep disturbances show evidence of higher rates of

psychiatric disorders than the general population and these disturbances may be presumed

to be the cause of their onset or recurrence rather than transpiring as symptoms (Ohayon

and Roth, 2003).


Clocks and Aging: The Ensuing Susceptibility to Internal Desynchrony

       The aging population is most susceptible to the depletion of chronobiological

rhythms.   The elderly are predisposed to chronodisruption due to ocular aging and

suboptimal photoreception necessary for circadian photoentrainment (Turner and

Mainster, 2008). Ocular aging consists of the aging of the crystalline lens and the

decreasing size of the pupil, culminating in a significant reduction of phototransduction

by RGCs, with the former blocking the absorption of favourable blue light (Turner and

Mainster, 2008).   In 2008, Turner and Mainster calculated the levels of circadian

photoreception decrease experienced throughout the aging process by multiplying human
34


crystalline lens transmission by pupil diameter and subsequently measuring melatonin

suppression sensitivity from light sources with wavelengths between 350 and 700

nanometres (Table 1). The results illustrated an age-dependent reduction in melatonin

suppression in response to blue light. From Table 1, it may be estimated that a person 95

years of age exhibits one-tenth of the level of photoreception seen in a ten year old.

Likewise, a person who is 85 years old requires 7.58 times brighter light exposures than a

15 year old in order to attain equivalent levels of photoreception.




Table 1 Circadian photoreception at different ages. The numbers in the table indicate
the level of retinal illumination achieved by the ages listed in the top row in contrast to
those in the left column. They also indicate the relative level of light exposure required
by those in the left column to achieve similar levels of effective photoreception as those
in the top row (Turner and Mainster, 2008).

       In addition to ocular aging, the elderly are particularly prone to insufficient light

exposure because of their habitual lifestyles (Turner and Mainster, 2008). Reduced

crystalline lens transmission and pupil diameter require brighter light exposures for the

elderly in order to maintain sufficient photoreception, however, residential lighting is

excessively dim and lacking in blue spectrum wavelengths compared to environmental
35


light (Fig 7) (Turner and Mainster, 2008). The link between circadian desynchrony and

scarce bright light exposure was investigated by Campbell et al. (1988), who recorded

levels of light exposure in 13 Alzheimer’s patients and 10 healthy elderly controls of

similar ages. The data, based on five days of recording subjects in their natural routines

at home, revealed that subjects rarely received exposure to ambient light in excess of

2000 lux. Furthermore, Alzheimer’s patients received 0.5 hours of bright illumination in

comparison to one hour in the control group and, in turn, the control group received one-

third to two-thirds less the amount encountered by healthy younger people. These figures

are even lower in those elderly subjects who are institutionalized in nursing homes and

retirement living centres (Turner and Mainster, 2008). Campbell et al. (1988) noted the

sleep deficits prevalent in both of these groups, therefore, light deficiency may be

implicated in circadian rhythm disturbances in the elderly, as well as the potential

neuropsychiatric illnesses in which they are involved.




Figure 7    Illuminance levels in a variety of settings. Residential lighting typically
ranges from 100 to 500 lux, however, proper physical and especially mental health
36


require environmental light exposures exceeding 1000 to 3000 lux, such as sunlight and
other sources of bright light (Turner and Mainster, 2008).

       Like the aging population, those with cataracts have inadequate levels of circadian

photoreception due to reduced ocular light transmission and smaller pupils (Turner and

Mainster, 2008).   In these cases, the crystalline lens is surgically replaced with an

intraocular lens (IOL), which typically blocks ultra-violet radiation and restores blue

light-inducing photoreception, however, some IOLs block blue light, resorting to

previous ophthalmologic standards (Turner and Mainster, 2008). Although some patients

with IOLs are lacking in circadian light exposures, IOLs have proven to be beneficial for

photoreception in the aging population (Turner and Mainster, 2008). As of yet, research

on the effects of aging on photosensitive RGCs remains controversial. Together, ocular

aging and light deficiency are responsible for the dampening of SCN output signals and

circadian amplitudes in the elderly to some extent, consequently leading to internal

desynchrony.

       An alternative region that appears to be associated with the attenuation of the

biological timekeeping system in the aging population is the SCN. Nygard et al. (2005)

attributed the weakened ability of the central oscillator to synchronize with external

stimuli, the dampening of activity and temperature cycles and disruptions in the sleep-

wake cycle to the altered electrophysiology of the aging SCN. Nygard and colleagues

(2005) used cell-attached and whole cell recordings to study the rhythm of spontaneous

firing and synaptic transmissions in the ventrolateral region of the SCN.             The

ventrolateral portion of the SCN receives input from the RHT and neurons in this region

mediate inhibitory synaptic transmission by expressing vasoactive intestinal polypeptide

(VIP) (Nygard et al., 2005). Single neurons in the ventrolateral region rhythmically
37


alternate between periods of silence and activity (Nygard et al., 2005). Recordings

conducted on slices of the SCN in vitro showed that young mice have a smaller

proportion of silent cells during the day, whereas such rhythmicity appeared to be absent

in older mice as a higher proportion of silent cells fired both during the day and the night.

These results point to an altered response to light in aged mice, with the SCN as a target

of the aging process as seen by the modified firing properties of individual neurons and

subsequently changed SCN output signals (Nygard et al., 2005).

       A recent study by Biello et al. (2009) demonstrated that the aging process alters

the central pacemaker by diminishing its response to phase shifting stimuli. This may be

the reason why the elderly exhibit advanced behavioural rhythms and lose the capacity to

temporally adapt to the environment (Biello et al., 2009). Biello et al. (2009) compared

the phase shifting properties of various neurotransmitters thought to be involved in

entraining the SCN in young and old mice. Glutamate, histamine and NMDA all delayed

the phase of rhythmicity in young mice, and thus are thought to be involved in photic

pathways, however, older mice did not respond as strongly (Biello et al., 2009). The

application of neurotransmitters thought to be involved in non-photic signaling pathways,

Muscimol, a GABA agonist, and 8-OH DPAT, a serotonin agonist, resulted in phase

advances in young mice, whereas older mice showed lesser responses. Gastrin-releasing

peptide and neuropeptide Y induced comparable phase shifts in both young and old mice.

These results are consistent with previous findings and signify the ability of the aging

SCN to phase shift in response to some stimuli, though not all, perhaps implicating the

disruption of particular synaptic pathways and neurotransmitter systems in the
38


dysfunction of the aging SCN (Penev et al., 1995; Palomba et al., 2008; Biello et al.,

2009).

         Since its discovery over twenty years ago, evidence demonstrating that the pineal

production of melatonin declines with aging is now a widely accepted fact. Sack et al.

(1986) demonstrated this concept by performing periodic assays for melatonin’s major

urinary metabolite 6-hydroxymelatonin, and hence measuring the total nocturnal

production of melatonin. Sack and collaborators conducted assays for three consecutive

nights in the summer and winter across a wide range of healthy adults, including medical

students, hospital personnel and retirement home residents.             After adjusting for

demographic variables of height, weight, gender, sleep patterns, smoking, alcohol and

coffee consumption, a significant negative correlation was found between age and

melatonin for both men and women. The same results were attained by Zhou et al.

(2003) by performing assays for melatonin on saliva, which also revealed that the decline

in cyclic melatonin production begins in middle-age, with these subjects having 60% of

the amplitude measured in young controls. Consideration must be given to the possibility

that such weakened melatonin levels in old age may be due to light deficits.

         Collectively, the age-related factors of light deficiency, ocular aging, deteriorated

electrical SCN rhythms, altered neurotransmitter signaling and diminished melatonin

production may be responsible for the vast array of circadian perturbations observed in

the elderly. Such perturbations include the dampening of circadian amplitudes and output

signals, the inability to synchronize with the environment, cognitive impairment,

advanced activity phases, lengthened free-running circadian period lengths, sleep

disturbances and psychiatric disorders (Nygard et al., 2005; Turner and Mainster, 2008;
39


Biello et al., 2009). The most frequently reported sleep-wake cycle alterations in old age

include fragmented sleep with less restoration, an overall phase advance of the sleep-

wake cycle with earlier bedtimes and earlier awakenings, increased daytime drowsiness

and insomnia (Campbell et al., 1988; Biello et al., 2009).


The Prevalence of Sleep Disturbances in the Aging Population

       Recall that the timing and amount of sleep are determined by circadian and

homeostatic sleep control mechanisms, respectively (Naylor et al., 2000). Dijk et al.

(1999) investigated the interplay of the circadian pacemaker and homeostatic factors in

sleep regulation and how they change with aging. The circadian rhythms of 13 older

subjects, ranging from 65 to 75 years old, and 11 younger subjects, ranging from 20 to 30

years old, were assessed using polysomnographic recordings and it was found that older

people wake up one hour earlier than predicted by the endogenous rhythms of core body

temperature and plasma melatonin with which sleep is normally synchronized. The

subjects were placed in states of temporal disorder and the amplitude of the core body

temperature rhythm in the elderly was reduced by 20 to 30% compared to younger

subjects. The older subjects exhibited high levels of sleep fragmentation and shorter

periods of sleeping, with the most fragmentation taking place when body temperature was

on the rise, thus suggesting that they are more sensitive to waking signals from the central

oscillator (Dijk et al., 1999). Following sleep deprivation, Dijk et al. (1999) observed

homeostatic control systems in operation, however, deep slow-wave sleep on EEGs was

markedly less in older subjects. These results indicate that the age-dependent decrease in

sleep quality and earlier sleep onset and wake times are due to the hindered ability of
40


circadian mechanisms to promote sleep during the geophysical morning and the hindered

ability of homeostatic mechanisms in enforcing sleep propensity (Dijk et al., 1999).

       The degeneration of the circadian timing system in the elderly was depicted in a

study by Huang et al. (2002), in which sleep-wake cycles and phases of rest and activity

were measured in routine settings. The study employed the use of wrist actigraphy over

twenty-four hour periods for five to seven consecutive days in young subjects ranging

from 21 to 34 years old, middle-aged subjects of 36 to 44 years old, old subjects of 61 to

79 years old and the oldest subjects of 80 to 91 years old. Those subjects showing

extreme preferences for activity in the morning or evening were excluded from the

investigation. In comparison to the young and middle-aged subjects, the old and oldest

subjects exhibited decreased sleep time, decreased sleep efficiency, longer sleep latency,

a higher number of nocturnal awakenings, a higher number of naps and the highest levels

of sleep fragmentation (Table 2) (Huang et al., 2002). Actigraph readings revealed

attenuated rhythms of rest and activity in the old and oldest subjects, with the lowest

daytime activity and the highest levels of activity during the night. In addition, the data

from these two subject groups are indicative of minimal coupling between sleep and

environmental Zeitgebers (Huang et al., 2002).       These results are suggestive of the

impairment of sleep-wake cycles and rest-activity rhythms with aging.




Table 2 Characteristics of the sleep-wake cycle are presented in the context of the four
age groups under study. The old and oldest subjects spent the most time in bed, however,
41


with the lowest amounts of actual sleep time. The old and oldest subjects took the
longest to fall asleep, had the lowest levels of sleep efficiency, the highest numbers of
nocturnal awakenings, the highest number of naps and the highest sleep fragmentation
indices (Huang et al., 2002).

       Recent research has given credence to the potential involvement of circadian

clock gene alterations and their profound implications for rhythms of sleep and

wakefulness in the aging population. Malatesta et al. (2007) analyzed CLOCK protein

levels in the neurons of the medullary reticular formation, the brain centre that

participates in the regulation of the sleep-wake cycle, in both young and old rats.

Immunocytochemical techniques were applied at different phases of the light-dark

circadian cycle.   Low CLOCK levels were found in the old rats in the nerve cell

compartments under scrutiny, including the cytoplasm, rough endoplasmic reticulum,

nucleus, nucleolus and chromatin. Malatesta and colleagues (2007) speculate that these

depressed levels of CLOCK protein in the neurons of the medullary reticular formation

are associated with significantly disturbed sleep-wake cycles in the elderly.


The Depletion of Chronobiological Rhythms and the Development of
Psychiatric Disorders with Age

       The aging population is most susceptible to the depletion of chronobiological

rhythms, and thus, the development of psychiatric disorders in the elderly merits

attention. Previous sections have outlined the relationships between circadian rhythm

abnormalities, perturbed sleep-wake cycles and aging.        It may be hypothesized that

senescence not only predisposes the elderly to chronodisruption and sleep deficits, but

also increases their risk for developing frequently comorbid psychiatric illnesses (Fig. 8).

The relationship between aging and depression via ocular dysfunction was evaluated by

Jean-Louis and colleagues (2005).          Recall ocular aging results in suboptimal
42


photoreception necessary for circadian photoentrainment. Study subjects’ ages averaged

68.3 years, with 27% being visually impaired according to ophthalmologic assessments.

Low ambient light exposures corresponded with depressed mood states when controlling

for demographic factors and medical complications (Jean-Louis et al., 2005). Ocular

pathologies such as glaucoma, ocular hypertension and cataracts appear to intensify this

relationship by negating light input to the master oscillator, thereby compromising sound

mental health in the elderly (Jean-Louis et al., 2005).




Figure 8 The implications of circadian rhythms in human mental health. A disruption
in the sleep-wake cycle reflects impaired circadian clock functioning, which
synergistically leads to the progression and maintenance of psychiatric disorders. The
aging population is most susceptible to the depletion of chronobiological rhythms and
sleep deficits, and thus, the development of psychiatric disorders in the elderly merits
attention.

       Malformed circadian rhythms and disrupted sleep parameters were previously

discussed in regards to Alzheimer’s disease, which is common in the elderly. A study by
43


Mishima et al. (1999) evaluated fluctuating levels of melatonin, which is thought to

decline with age (Sack et al., 1986), and rest-activity rhythms in elderly patients with

Alzheimer’s disease. Wrist actigraphy was used to assess circadian rest-activity rhythms

and blood samples were assayed for plasma melatonin concentrations. The first study

group consisted of Alzheimer’s patients with ages averaging 75.7 years and the second

group consisted of dementia-free residents of the same nursing-home facility whose ages

averaged 78.3 years, with the latter being free of disturbed sleep-wake cycles. The study

was conducted in conditions of light below 150 lux and minimal physical exercise in

order to prevent the suppression of melatonin.         The Alzheimer’s patients showed

considerably reduced amplitudes of melatonin secretion, with several patients displaying

atypical peak secretion levels during the day, and less total daily secretions in comparison

to the control group (Mishima et al., 1999). In addition, the rest-activity rhythms of the

Alzheimer’s patients proved to be quite erratic. A similar study reported insufficient light

exposure in these patients (Ancoli-Israel et al., 1997). This study establishes a positive

correlation between dampened melatonin rhythms and disturbed sleep-wake patterns and

rest-activity cycles, all of which are characteristic of elderly Alzheimer’s patients

(Mishima et al., 1999; Volicer et al., 2001; Mahlberg et al., 2008).

       As previously described, there is a temporal misalignment of circadian rhythms,

sleep-wake cycles and light exposures observed in schizophrenic patients. Martin et al.

(2001) evaluated these factors in an aged population of schizophrenic patients, consisting

of 14 men and 14 women whose ages averaged 58.3 years. An Actillume wrist monitor

was used to measure both light exposure and activity levels.           The dramatic results

indicated that reduced light exposure was linked to weakened circadian rhythms and
44


sleep fragmentation, especially with age. The mean light exposure among the 28 subjects

was less than 1000 lux, which worsened with age, and correlated with depressed mood

and increased severity of psychiatric symptoms. These patients exhibited an excessive

number of nocturnal awakenings, sometimes leading to insomnia for more than three

hours per night.     These sleep disturbances resulted in more daytime napping, and

therefore, less daytime activity with substandard neuropsychological functioning and

poor cognition.     Actigraph recordings produced one-fifth of the robust amplitude

measured in control participants. Collectively, these findings suggest possible roles for

light deficiency, sleep disturbances, attenuated circadian rhythms, lifestyle and age status

in this psychiatric disorder (Martin et al., 2001). However, the administration of anti-

psychotic medications in many of the subjects may have confounded the results and

further studies are required to delineate their contribution, if any, to these disturbed

behavioural rhythms in comparison to those subjects who are not taking medication

(Martin et al., 2001).


Possible Treatments and Chronobiotics for Circadian Dysfunction

       With circadian rhythm disturbances as characteristic of various sleep disorders

and psychiatric disorders, an assortment of chronobiological therapies has been proposed

to alleviate their symptoms. The major internal Zeitgeber melatonin has been suggested

as an effective pharmacological treatment for a range of circadian disruptions, including

circadian rhythm sleep disorders, jet lag, shift-work maladaptation and free-running

rhythms in the blind (Fischer et al., 2003). Recall that blind individuals are susceptible to

sleep disorders and compromised neuropsychiatric conditions.          Fischer et al. (2003)

investigated whether a single one-time melatonin administration could temporarily
45


entrain blind individuals, thus synchronizing their sleep-wake cycles and melatonin

rhythms and improving sleep conditions.        Twelve men ages 18 to 40, incapable of

photoentrainment, were given 5mg of melatonin one hour before bedtime, with both the

subjects and administrators being unaware if the substance being given was melatonin or

a placebo.    In contrast to the placebo, melatonin increased total sleep time and sleep

efficiency while decreasing the number of nocturnal awakening episodes. In regards to

endocrine processes, adrenocorticotropic (ACTH) hormone and cortisol secretion are

normally inhibited during the first half of sleep and rise in the latter half (Fischer et al.,

2003). With this normally entrained rhythm being desynchronized in blind individuals, a

single dose of melatonin improved sleep by realigning these hormonal rhythms.

       In light of its ability to synchronize circadian rhythms, improve sleep quality and

regulate the hypothalamic-pituitary-adrenal axis, melatonin has the potential to serve as

an anti-depressant (Fischer et al., 2003). This prospect is further supported in a study by

Benedetti et al. (2001), in which similar results as Fischer and colleagues (2003) were

obtained, however, with the additional finding that melatonin decreased the need for the

use of benzodiazepines. In the context of this study, psychoactive benzodiazepine drugs

were being taken by elderly subjects in order to treat insomnia (Benedetti et al., 2001).

Melatonin administration before bedtime eliminated the use of benzodiazepines entirely

in 65% of subjects, while reducing their use by 25 to 66% in 20% of subjects. Anti-

depressants that improve sleep quality could therefore be crucial to treating depressive

disorders.

       Recently, the new chronobiotic agomelatine has been put forward as a potential

anti-depressant because of its coordinating effects on the circadian rest-activity rhythm
46


and its mitigating effects on symptoms of depression and anxiety (Kasper et al., 2010).

Agomelatine acts as an agonist at melatonin receptors MT1 and MT2 and as an

antagonist at serotonin receptors (Kasper et al., 2010). A study by Kasper and colleagues

(2010) compared the effects of agomelatine with sertraline, a selective serotonin reuptake

inhibitor (SSRI) known by its trade name as Zoloft, on patients with MDD. Agomelatine,

in contrast to sertraline, increased the amplitude of rest-activity rhythms within one week.

According to wrist actigraphs and sleep diaries, agomelatine improved sleep quality and

the ease of falling asleep, and relieved feelings of depression and anxiety without any

major adverse effects.     Further research is required to establish agomelatine as an

effective treatment for sleep disorders and affective disorders.

       Selective serotonin reuptake inhibitors are often employed in the treatment of

affective disorders, including major depressive disorder, seasonal affective disorder and

bipolar disorder. Sprouse et al. (2006) demonstrated the ability of fluoxetine, an SSRI, to

alter firing activity of neurons in the SCN, and hence, circadian rhythmicity. At first,

extracellular recordings of spontaneous neuron firing in the hypothalamic SCN slices of

rats in vitro revealed no change in rhythm in response to fluoxetine. This was thought to

be due to the loss of endogenous serotonin levels in culture conditions in vitro (Sprouse et

al., 2006).    When fluoxetine was paired with tryptophan, a serotonin precursor,

microelectrode recordings revealed concentration-dependent phase advances in SCN

rhythms (Sprouse et al., 2006). Further research is being conducted to ascertain the

magnitude and direction of circadian phase shifts in clinical applications of fluoxetine

(Sprouse et al., 2006).
47


       As a mood stabilizer, lithium is used in the treatment of bipolar disorder as it

counteracts both mania and depression (Hafen and Wollnik, 1994). Lithium lengthens

the circadian period of the central pacemaker through direct pharmacological effects

(Iwahana et al., 2004). The drug inhibits the action of a glycogen synthase kinase 3, an

enzyme which targets the transcription factors of the molecular oscillator for degradation

via phosphorylation, thereby slowing the molecular oscillator and relieving the advanced

rhythm disturbances often seen in bipolar patients (Hafen and Wollnik, 1994; Iwahana et

al., 2004). Most antidepressants, including lithium, need 2 to 8 weeks in order to exert

their effects and induce a favourable response in patients (Wu et al., 2009).

Unfortunately, there is a high risk for suicide in patients with bipolar disorder depression

and a treatment regimen that evokes a rapid response and maintains this response is

necessary (Wu et al., 2009).

       Despite its transient effects, sleep deprivation has been discovered as one of the

most prompt and efficient chronotherapeutics, reducing depressive symptoms in 40 to

60% of patients within 24 to 48 hours (Wu et al., 2009). A study by Wu et al. (2009)

evaluated the standard medications lithium and sertraline against a chronotherapeutic

augmentation treatment (CAT), consisting of medications, sleep deprivation, bright light

therapy and sleep phase advances. Forty-nine patients diagnosed with bipolar disorder

according to the DSM were randomly assigned to either the medication group or the CAT

group. CAT subjects were kept awake for 33 hours and then exposed to 5000 lux light

for two hours for three consecutive days following sleep deprivation, as well as three

days of gradual sleep phase advances. The CAT group exhibited a substantial relief of

depressive symptoms as early as two days into treatment, and this effect lasted for seven
48


weeks, at which time 12/19 CAT subjects had gone into remission (Fig. 9).               A

combination of established chronotherapies appears to be most effective in alleviating the

symptoms of this particular psychiatric disorder.




Figure 9 In accordance with the Hamilton Rating Scale for Depression, CAT subjects,
whose treatment consisted of medications, sleep deprivation, bright light therapy and
sleep phase advances, displayed a considerable reduction in depressive symptoms in
comparison to those solely on medications. This difference was seen as early as Day 2
and was maintained for 7 weeks (Wu et al., 2009).

       Bright light therapy in the morning has long been known as an effective treatment

for both MDD and SAD patients. A study by Lewy et al. (1998) demonstrated the

superior efficacy of morning versus evening bright light therapy. This was based on the

prediction that SAD patients would have phase delayed rhythms during winter depression

(Lewy et al., 1998). Forty-five patients with MDD or bipolar disorder with a winter

seasonal pattern and 49 controls participated in the study for six weeks. The participants

were treated with bright light therapy in their homes for two weeks either in the morning,

between 6 to 8 AM, or in the evening, between 7 to 9 PM, then subjected to a week of

light withdrawal, and then treated with light in the opposing time period. Blood samples

were taken once a week in dim light and assayed for melatonin in order to determine
49


circadian phase positions relative to nocturnal melatonin onset. Morning bright light

therapy, which phase advanced patients, proved more effective than that in the evening,

which phase delayed patients, with the former inducing a 27% decrease in depressive

symptoms and intensifying those symptoms during the withdrawal period (Lewy et al.,

1998). The authors of this study suggest the application of bright light therapy without

delay upon awakening for the best results in patients with SAD (Lewy et al., 1998). The

phase advancing and phase delaying effects of bright light therapy may be applied in

treatment regimens for the circadian rhythm sleep disorders DSPS and ASPS as well.

       With light as the dominant Zeitgeber in the human circadian timekeeping system,

it is no surprise that light exposure is one of the most potent treatments for circadian

rhythm disorders. Recall that photosensitive RGCs best absorb light in the blue sector of

the light spectrum at 460 nm, which is quite similar to the wavelength of environmental

light (Turner and Mainster, 2008). A study by Glickman et al. (2006) investigated the

optimal spectral wavelength for phototherapy in 24 SAD patients. Blue light emitting

diode boxes gave off 468 nm light to 11 subjects, while the red light emitting diode boxes

gave off 654 nm to 13 subjects. Light therapy was administered everyday for three

weeks for durations of 45 minutes between 6 to 8 AM. According to the Hamilton

Depression Rating Scale, subjects who had been given short blue wavelength light

treatments scored 7.3 points lower on depressive symptoms than those who had been

exposed to longer red wavelength light, thus confirming blue light as optimal for light

therapy (Glickman et al., 2006). As previously outlined in bright morning light therapy,

the timing of light exposure is critical in light therapy. A recent study by Goel et al.

(2006) applied 10 000 lux light for one hour immediately upon awakening to patients
50


with chronic MDD diagnoses for five weeks.             The results indicated substantial

improvement in symptoms, with depression scores improving by 53.7 % and with

remission rates of 50%. Light therapy is predicted to be most effective in conjunction

with the abovementioned treatment strategies of melatonin and sleep deprivation (Goel et

al., 2006).

        The widespread success of light therapy is derived from its ability to phase shift

the endogenous circadian pacemaker, and so it is used to alleviate the detrimental effects

of jet lag, shift-work, circadian rhythm sleep disorders, Alzheimer’s disease and bipolar

disorder, to name a few.      Optimal photoreception may be achieved in the aging

population with ample natural light exposure, IOLs and bright, appropriately timed

residential lighting (Turner and Mainster, 2008). Structural designs that allow for bright

environmental light exposure during the geophysical day and limited light exposure in the

evening would allow for most advantageous entrainment and internal synchrony (Oren et

al., 1997; Turner and Mainster, 2008). Increasing public awareness of these strategies for

harmonious synchronization and optimal well-being are not only profitable to the elderly

in preventing circadian malfunction but to all age groups.


Research Proposal

Alleviating Sleep Disorders to Alleviate Psychiatric Disturbances

Rationale

        Circadian dysfunction, notably decreased circulating melatonin levels and

disturbed rest-activity rhythms, and the deterioration of the sleep-wake cycle are

characteristic of the aging population (Campbell et al., 1988; Zhou et al., 2003; Nygard et

al., 2005; Turner and Mainster, 2008). In light of those previously mentioned studies that
51


have enhanced sleeping conditions and alleviated depressive symptoms through the use

of melatonin and light therapy, future studies warrant investigating whether correcting

circadian misalignment with these circadian resetting agents will attenuate sleep

disturbances and psychiatric pathologies in the elderly (Lewy et al., 1998; Fischer et al.,

2003; Goel et al., 2006). The proposed study will evaluate the efficacy of a treatment

combining timed bright light exposure and exogenous melatonin, two primary Zeitgebers,

in consolidating circadian rhythms and alleviating sleep disturbances and psychiatric

symptoms in elderly patients with Alzheimer’s disease.


Participants


Participants will be gathered from a geriatric facility and will be comprised of

Alzheimer’s patients ranging from the ages of 60 to 90 years old. Written consent forms

and approval from an institutional ethics board will be attained prior to commencing the

study.


Preliminary Assessment

Prior to commencing treatment, initial assessments of sleep-wake cycles and rest-activity

rhythms will be made for a duration of one week. This information will be gathered

through nurse and staff ratings, subject interviews and with the use of Actillume wrist

monitors (Martin et al., 2001). Actillume recordings will also indicate the amount of

light exposure obtained by the participants. Saliva samples will be taken every 30

minutes during the daytime in dim light conditions within a 24 hour period and assayed

for melatonin concentrations in order to assess the level of circadian misalignment (Zhou

et al., 2003). In addition, the total amount of nocturnal melatonin production will be
52


measured by performing assays for its major urinary metabolite 6-hydroxymelatonin

(Sack et al., 1986). Cognitive tests and psychiatric assessments will be applied to assess

the severity of psychiatric symptoms.


Methods

The study will be conducted in a double-blind manner using placebos. Participants will

randomly be assigned to one of two groups. The first group will receive bright light

exposure of 3000 lux daily between the hours of 6AM and 9AM and 3 mg of melatonin

one hour before bedtime for one month (Lewy et al., 1998; Fischer et al., 2003; Turner

and Mainster, 2008). The use of blue light emitting diodes would be best (Glickman et

al., 2006). The second group will receive light exposure of 100 lux daily between the

hours of 6AM and 9AM and a placebo pill one hour before bedtime for one month. The

use of red light emitting diodes would be best (Glickman et al., 2006). The study will be

conducted in a double-blind manner since the nurses and staff, as well as the participants

themselves, will be unaware as to whether they are administering a treatment regimen or

placebo. A final assessment will be made after one month of treatment by applying the

same procedures as outlined in the preliminary assessment. The healthcare professionals

performing the final cognitive tests and psychiatric assessments will be uninformed as to

which group participants were assigned.


Controls

The only difference between the two groups will be the treatment regimen administered

as they will otherwise be living in the same geriatric facility, be age-matched and have a

diagnosis of dementia. The placebo pills will serve as controls for exogenous melatonin.
53


The use of longer wavelength light will serve as a control as it has been established that

light of this intensity and of the red spectrum are insufficient for the body’s circadian

demands (Glickman et al., 2006; Turner and Mainster, 2008).


Predicted Outcomes

The combined treatment of bright morning light therapy and exogenous melatonin will

most likely result in improved central pacemaker functioning, which will be evident with

decreased activity during the night and increased activity during the day. Treatment will

lead to an increase in sleep efficiency, sleep duration, and deep sleep, with a decrease in

daytime napping, nocturnal awakenings and agitation. Nocturnal melatonin levels will

increase, not only due to the application of exogenous melatonin, but also due to

improved sleeping parameters and light exposure. Cognition and mood will be enhanced

as well. Melatonin rhythms and light exposure are dampened in Alzheimer’s patients,

therefore, this treatment has the potential to restore the compromised rest-activity cycles,

sleep-wake patterns and neuropsychiatric functioning seen in these elderly patients

(Ancoli-Israel et al., 1997; Mishima et al., 1999).


Limitations


The placebo effect, medications and the severity of dementia-related symptoms may

prove to be limitations in this study. This study may be expanded to test for the placebo

effect by including an additional age-matched control group living in the same geriatric

facility with diagnoses of dementia, however, this group would receive neither the

treatment regimen nor placebos. This control group would account for other variables

that may lead to improvements in participants, for example, the context of the facility in
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population
The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population

More Related Content

Viewers also liked

Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture?
Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture? Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture?
Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture?
InsideScientific
 
Powerpoint Presentation Circadian Rhythms
Powerpoint Presentation Circadian RhythmsPowerpoint Presentation Circadian Rhythms
Powerpoint Presentation Circadian Rhythms
wperng
 
050 circadian rhythms biological clocks
050 circadian rhythms   biological clocks050 circadian rhythms   biological clocks
050 circadian rhythms biological clocks
simpletasksgreatconcepts
 
Biological rhythms and sleep
Biological rhythms and sleepBiological rhythms and sleep
Biological rhythms and sleep
Georgie Hartshorne
 
Circadian rhythms
Circadian rhythmsCircadian rhythms
Circadian rhythms
PPRC AYUR
 
Biological clock
Biological clockBiological clock
Biological clock
Vidyasagar University
 
Chronobiolgoy 6378510
Chronobiolgoy 6378510Chronobiolgoy 6378510
Chronobiolgoy 6378510
Suman Sajan
 
Biological clock
Biological clockBiological clock
Biological clock
a-saeed
 
Biological rhythms
Biological rhythmsBiological rhythms
Using Oracle Applications on your iPad
Using Oracle Applications on your iPadUsing Oracle Applications on your iPad
Using Oracle Applications on your iPadOracle Day
 
Hampshire Constabulary Renuneration Planning SFIA
Hampshire Constabulary Renuneration Planning SFIAHampshire Constabulary Renuneration Planning SFIA
Hampshire Constabulary Renuneration Planning SFIASFIA User Forum
 
How to create the life you want
How to create the life you wantHow to create the life you want
How to create the life you want
Self-employed
 
Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...
Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...
Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...
Paul Biya
 
Unit 1.my school
Unit 1.my schoolUnit 1.my school
Unit 1.my school
Alexandre Bárez
 
Best Practice Guideline to Managing On-site Vermiculture Technologies
Best Practice Guideline to Managing On-site Vermiculture TechnologiesBest Practice Guideline to Managing On-site Vermiculture Technologies
Best Practice Guideline to Managing On-site Vermiculture Technologies
x3G9
 
Bai tap trac_nghiem_16_units
Bai tap trac_nghiem_16_unitsBai tap trac_nghiem_16_units
Bai tap trac_nghiem_16_units
Godfrey Tran
 

Viewers also liked (20)

Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture?
Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture? Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture?
Circadian Rhythms of Food Intake: Are You Seeing The Whole Picture?
 
Powerpoint Presentation Circadian Rhythms
Powerpoint Presentation Circadian RhythmsPowerpoint Presentation Circadian Rhythms
Powerpoint Presentation Circadian Rhythms
 
050 circadian rhythms biological clocks
050 circadian rhythms   biological clocks050 circadian rhythms   biological clocks
050 circadian rhythms biological clocks
 
Biological rhythms and sleep
Biological rhythms and sleepBiological rhythms and sleep
Biological rhythms and sleep
 
Circadian rhythms
Circadian rhythmsCircadian rhythms
Circadian rhythms
 
Circadian Rhythms: Sleep-waking cycle
Circadian Rhythms: Sleep-waking cycleCircadian Rhythms: Sleep-waking cycle
Circadian Rhythms: Sleep-waking cycle
 
Biological clock
Biological clockBiological clock
Biological clock
 
Chronobiolgoy 6378510
Chronobiolgoy 6378510Chronobiolgoy 6378510
Chronobiolgoy 6378510
 
Circadian rhythm
Circadian rhythmCircadian rhythm
Circadian rhythm
 
Biological clock
Biological clockBiological clock
Biological clock
 
Biological rhythms
Biological rhythmsBiological rhythms
Biological rhythms
 
Using Oracle Applications on your iPad
Using Oracle Applications on your iPadUsing Oracle Applications on your iPad
Using Oracle Applications on your iPad
 
Hampshire Constabulary Renuneration Planning SFIA
Hampshire Constabulary Renuneration Planning SFIAHampshire Constabulary Renuneration Planning SFIA
Hampshire Constabulary Renuneration Planning SFIA
 
How to create the life you want
How to create the life you wantHow to create the life you want
How to create the life you want
 
Transformator
TransformatorTransformator
Transformator
 
Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...
Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...
Paul Biya - Cameroun - Décret N° 2016355 du 28 juillet 2016 portant inscripti...
 
Migrain tamil
Migrain tamilMigrain tamil
Migrain tamil
 
Unit 1.my school
Unit 1.my schoolUnit 1.my school
Unit 1.my school
 
Best Practice Guideline to Managing On-site Vermiculture Technologies
Best Practice Guideline to Managing On-site Vermiculture TechnologiesBest Practice Guideline to Managing On-site Vermiculture Technologies
Best Practice Guideline to Managing On-site Vermiculture Technologies
 
Bai tap trac_nghiem_16_units
Bai tap trac_nghiem_16_unitsBai tap trac_nghiem_16_units
Bai tap trac_nghiem_16_units
 

Similar to The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population

Circadian and Circannual Rhythms
Circadian and Circannual RhythmsCircadian and Circannual Rhythms
Circadian and Circannual Rhythms
D. B. S. College Kanpur
 
Chronobiology
ChronobiologyChronobiology
Chronobiology
donthuraj
 
Crimson Publishers-More than Meets the Eye: Biological Clocks
 Crimson Publishers-More than Meets the Eye: Biological Clocks Crimson Publishers-More than Meets the Eye: Biological Clocks
Crimson Publishers-More than Meets the Eye: Biological Clocks
CrimsonpublishersMSOR
 
chronobio-161224071722.pdf
chronobio-161224071722.pdfchronobio-161224071722.pdf
chronobio-161224071722.pdf
KavithaaKannan1
 
CHRONO BIOLOGY
CHRONO BIOLOGYCHRONO BIOLOGY
CHRONO BIOLOGY
Vln Sekhar
 
ENEURO.0237-16.2016.full
ENEURO.0237-16.2016.fullENEURO.0237-16.2016.full
ENEURO.0237-16.2016.fullMonica Gonzalez
 
Molecular mechanisms that control circadian rhythms - Mohammed Elreishi
Molecular mechanisms that control circadian rhythms - Mohammed Elreishi Molecular mechanisms that control circadian rhythms - Mohammed Elreishi
Molecular mechanisms that control circadian rhythms - Mohammed Elreishi
Mohammed Elreishi
 
Chronobiology
ChronobiologyChronobiology
Chronobiology
Tessi623
 
Light Deprivation and Its Effects
Light Deprivation and Its EffectsLight Deprivation and Its Effects
Light Deprivation and Its EffectsAlicia Ai Leng
 
BS_Lecture1_Introduction in chronobiology.pptx
BS_Lecture1_Introduction in chronobiology.pptxBS_Lecture1_Introduction in chronobiology.pptx
BS_Lecture1_Introduction in chronobiology.pptx
Bogdan Stana
 
Circadian rhythms
Circadian rhythmsCircadian rhythms
Circadian rhythms
jasleenbrar03
 
Scope of chronobiology
Scope of chronobiologyScope of chronobiology
Scope of chronobiology
Sher Khan
 
CHRONOBIOLOGY
CHRONOBIOLOGYCHRONOBIOLOGY
CHRONOBIOLOGY
RamyaAchar1
 
Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...
Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...
Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...Ligia Westrich
 
138th publication sjodr- 7th name
138th publication  sjodr- 7th name138th publication  sjodr- 7th name
138th publication sjodr- 7th name
CLOVE Dental OMNI Hospitals Andhra Hospital
 
EEG Brain Rhythms
EEG Brain RhythmsEEG Brain Rhythms
EEG Brain Rhythms
Rochelle Schear
 
The Role Of Cytokines On Immune Privilege
The Role Of Cytokines On Immune PrivilegeThe Role Of Cytokines On Immune Privilege
The Role Of Cytokines On Immune Privilege
Katy Allen
 
Anatomy And Physiology
Anatomy And PhysiologyAnatomy And Physiology
Anatomy And Physiology
Jeff Brooks
 
Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)
Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)
Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)
Vall d'Hebron Institute of Research (VHIR)
 

Similar to The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population (20)

Circadian and Circannual Rhythms
Circadian and Circannual RhythmsCircadian and Circannual Rhythms
Circadian and Circannual Rhythms
 
Chronobiology
ChronobiologyChronobiology
Chronobiology
 
Crimson Publishers-More than Meets the Eye: Biological Clocks
 Crimson Publishers-More than Meets the Eye: Biological Clocks Crimson Publishers-More than Meets the Eye: Biological Clocks
Crimson Publishers-More than Meets the Eye: Biological Clocks
 
chronobio-161224071722.pdf
chronobio-161224071722.pdfchronobio-161224071722.pdf
chronobio-161224071722.pdf
 
CHRONO BIOLOGY
CHRONO BIOLOGYCHRONO BIOLOGY
CHRONO BIOLOGY
 
ENEURO.0237-16.2016.full
ENEURO.0237-16.2016.fullENEURO.0237-16.2016.full
ENEURO.0237-16.2016.full
 
Molecular mechanisms that control circadian rhythms - Mohammed Elreishi
Molecular mechanisms that control circadian rhythms - Mohammed Elreishi Molecular mechanisms that control circadian rhythms - Mohammed Elreishi
Molecular mechanisms that control circadian rhythms - Mohammed Elreishi
 
Chronobiology
ChronobiologyChronobiology
Chronobiology
 
Light Deprivation and Its Effects
Light Deprivation and Its EffectsLight Deprivation and Its Effects
Light Deprivation and Its Effects
 
BS_Lecture1_Introduction in chronobiology.pptx
BS_Lecture1_Introduction in chronobiology.pptxBS_Lecture1_Introduction in chronobiology.pptx
BS_Lecture1_Introduction in chronobiology.pptx
 
Circadian rhythms
Circadian rhythmsCircadian rhythms
Circadian rhythms
 
Scope of chronobiology
Scope of chronobiologyScope of chronobiology
Scope of chronobiology
 
CHRONOBIOLOGY
CHRONOBIOLOGYCHRONOBIOLOGY
CHRONOBIOLOGY
 
Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...
Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...
Westrich & Sprouse, 2010 - COID - Circadian rhythm dysregulation in bipolar d...
 
138th publication sjodr- 7th name
138th publication  sjodr- 7th name138th publication  sjodr- 7th name
138th publication sjodr- 7th name
 
EEG Brain Rhythms
EEG Brain RhythmsEEG Brain Rhythms
EEG Brain Rhythms
 
The Role Of Cytokines On Immune Privilege
The Role Of Cytokines On Immune PrivilegeThe Role Of Cytokines On Immune Privilege
The Role Of Cytokines On Immune Privilege
 
Anatomy And Physiology
Anatomy And PhysiologyAnatomy And Physiology
Anatomy And Physiology
 
FULL111711
FULL111711FULL111711
FULL111711
 
Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)
Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)
Stress and the Brain-Gut-Microbiota Axis (Prof. John F. Cryan)
 

Recently uploaded

NVBDCP.pptx Nation vector borne disease control program
NVBDCP.pptx Nation vector borne disease control programNVBDCP.pptx Nation vector borne disease control program
NVBDCP.pptx Nation vector borne disease control program
Sapna Thakur
 
Top-Vitamin-Supplement-Brands-in-India.pptx
Top-Vitamin-Supplement-Brands-in-India.pptxTop-Vitamin-Supplement-Brands-in-India.pptx
Top-Vitamin-Supplement-Brands-in-India.pptx
SwisschemDerma
 
Gram Stain introduction, principle, Procedure
Gram Stain introduction, principle, ProcedureGram Stain introduction, principle, Procedure
Gram Stain introduction, principle, Procedure
Suraj Goswami
 
micro teaching on communication m.sc nursing.pdf
micro teaching on communication m.sc nursing.pdfmicro teaching on communication m.sc nursing.pdf
micro teaching on communication m.sc nursing.pdf
Anurag Sharma
 
Knee anatomy and clinical tests 2024.pdf
Knee anatomy and clinical tests 2024.pdfKnee anatomy and clinical tests 2024.pdf
Knee anatomy and clinical tests 2024.pdf
vimalpl1234
 
Maxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptx
Maxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptxMaxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptx
Maxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptx
Dr. Rabia Inam Gandapore
 
Ocular injury ppt Upendra pal optometrist upums saifai etawah
Ocular injury  ppt  Upendra pal  optometrist upums saifai etawahOcular injury  ppt  Upendra pal  optometrist upums saifai etawah
Ocular injury ppt Upendra pal optometrist upums saifai etawah
pal078100
 
Pharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptx
Pharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptxPharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptx
Pharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptx
Dr. Rabia Inam Gandapore
 
Pharma Pcd Franchise in Jharkhand - Yodley Lifesciences
Pharma Pcd Franchise in Jharkhand - Yodley LifesciencesPharma Pcd Franchise in Jharkhand - Yodley Lifesciences
Pharma Pcd Franchise in Jharkhand - Yodley Lifesciences
Yodley Lifesciences
 
Flu Vaccine Alert in Bangalore Karnataka
Flu Vaccine Alert in Bangalore KarnatakaFlu Vaccine Alert in Bangalore Karnataka
Flu Vaccine Alert in Bangalore Karnataka
addon Scans
 
SURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptx
SURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptxSURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptx
SURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptx
Bright Chipili
 
Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...
Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...
Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...
Oleg Kshivets
 
How STIs Influence the Development of Pelvic Inflammatory Disease.pptx
How STIs Influence the Development of Pelvic Inflammatory Disease.pptxHow STIs Influence the Development of Pelvic Inflammatory Disease.pptx
How STIs Influence the Development of Pelvic Inflammatory Disease.pptx
FFragrant
 
Superficial & Deep Fascia of the NECK.pptx
Superficial & Deep Fascia of the NECK.pptxSuperficial & Deep Fascia of the NECK.pptx
Superficial & Deep Fascia of the NECK.pptx
Dr. Rabia Inam Gandapore
 
Triangles of Neck and Clinical Correlation by Dr. RIG.pptx
Triangles of Neck and Clinical Correlation by Dr. RIG.pptxTriangles of Neck and Clinical Correlation by Dr. RIG.pptx
Triangles of Neck and Clinical Correlation by Dr. RIG.pptx
Dr. Rabia Inam Gandapore
 
Tom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness Journey
Tom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness JourneyTom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness Journey
Tom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness Journey
greendigital
 
Cervical & Brachial Plexus By Dr. RIG.pptx
Cervical & Brachial Plexus By Dr. RIG.pptxCervical & Brachial Plexus By Dr. RIG.pptx
Cervical & Brachial Plexus By Dr. RIG.pptx
Dr. Rabia Inam Gandapore
 
KDIGO 2024 guidelines for diabetologists
KDIGO 2024 guidelines for diabetologistsKDIGO 2024 guidelines for diabetologists
KDIGO 2024 guidelines for diabetologists
د.محمود نجيب
 
Basavarajeeyam - Ayurvedic heritage book of Andhra pradesh
Basavarajeeyam - Ayurvedic heritage book of Andhra pradeshBasavarajeeyam - Ayurvedic heritage book of Andhra pradesh
Basavarajeeyam - Ayurvedic heritage book of Andhra pradesh
Dr. Madduru Muni Haritha
 
Physiology of Special Chemical Sensation of Taste
Physiology of Special Chemical Sensation of TastePhysiology of Special Chemical Sensation of Taste
Physiology of Special Chemical Sensation of Taste
MedicoseAcademics
 

Recently uploaded (20)

NVBDCP.pptx Nation vector borne disease control program
NVBDCP.pptx Nation vector borne disease control programNVBDCP.pptx Nation vector borne disease control program
NVBDCP.pptx Nation vector borne disease control program
 
Top-Vitamin-Supplement-Brands-in-India.pptx
Top-Vitamin-Supplement-Brands-in-India.pptxTop-Vitamin-Supplement-Brands-in-India.pptx
Top-Vitamin-Supplement-Brands-in-India.pptx
 
Gram Stain introduction, principle, Procedure
Gram Stain introduction, principle, ProcedureGram Stain introduction, principle, Procedure
Gram Stain introduction, principle, Procedure
 
micro teaching on communication m.sc nursing.pdf
micro teaching on communication m.sc nursing.pdfmicro teaching on communication m.sc nursing.pdf
micro teaching on communication m.sc nursing.pdf
 
Knee anatomy and clinical tests 2024.pdf
Knee anatomy and clinical tests 2024.pdfKnee anatomy and clinical tests 2024.pdf
Knee anatomy and clinical tests 2024.pdf
 
Maxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptx
Maxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptxMaxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptx
Maxilla, Mandible & Hyoid Bone & Clinical Correlations by Dr. RIG.pptx
 
Ocular injury ppt Upendra pal optometrist upums saifai etawah
Ocular injury  ppt  Upendra pal  optometrist upums saifai etawahOcular injury  ppt  Upendra pal  optometrist upums saifai etawah
Ocular injury ppt Upendra pal optometrist upums saifai etawah
 
Pharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptx
Pharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptxPharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptx
Pharynx and Clinical Correlations BY Dr.Rabia Inam Gandapore.pptx
 
Pharma Pcd Franchise in Jharkhand - Yodley Lifesciences
Pharma Pcd Franchise in Jharkhand - Yodley LifesciencesPharma Pcd Franchise in Jharkhand - Yodley Lifesciences
Pharma Pcd Franchise in Jharkhand - Yodley Lifesciences
 
Flu Vaccine Alert in Bangalore Karnataka
Flu Vaccine Alert in Bangalore KarnatakaFlu Vaccine Alert in Bangalore Karnataka
Flu Vaccine Alert in Bangalore Karnataka
 
SURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptx
SURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptxSURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptx
SURGICAL ANATOMY OF THE RETROPERITONEUM, ADRENALS, KIDNEYS AND URETERS.pptx
 
Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...
Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...
Local Advanced Lung Cancer: Artificial Intelligence, Synergetics, Complex Sys...
 
How STIs Influence the Development of Pelvic Inflammatory Disease.pptx
How STIs Influence the Development of Pelvic Inflammatory Disease.pptxHow STIs Influence the Development of Pelvic Inflammatory Disease.pptx
How STIs Influence the Development of Pelvic Inflammatory Disease.pptx
 
Superficial & Deep Fascia of the NECK.pptx
Superficial & Deep Fascia of the NECK.pptxSuperficial & Deep Fascia of the NECK.pptx
Superficial & Deep Fascia of the NECK.pptx
 
Triangles of Neck and Clinical Correlation by Dr. RIG.pptx
Triangles of Neck and Clinical Correlation by Dr. RIG.pptxTriangles of Neck and Clinical Correlation by Dr. RIG.pptx
Triangles of Neck and Clinical Correlation by Dr. RIG.pptx
 
Tom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness Journey
Tom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness JourneyTom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness Journey
Tom Selleck Health: A Comprehensive Look at the Iconic Actor’s Wellness Journey
 
Cervical & Brachial Plexus By Dr. RIG.pptx
Cervical & Brachial Plexus By Dr. RIG.pptxCervical & Brachial Plexus By Dr. RIG.pptx
Cervical & Brachial Plexus By Dr. RIG.pptx
 
KDIGO 2024 guidelines for diabetologists
KDIGO 2024 guidelines for diabetologistsKDIGO 2024 guidelines for diabetologists
KDIGO 2024 guidelines for diabetologists
 
Basavarajeeyam - Ayurvedic heritage book of Andhra pradesh
Basavarajeeyam - Ayurvedic heritage book of Andhra pradeshBasavarajeeyam - Ayurvedic heritage book of Andhra pradesh
Basavarajeeyam - Ayurvedic heritage book of Andhra pradesh
 
Physiology of Special Chemical Sensation of Taste
Physiology of Special Chemical Sensation of TastePhysiology of Special Chemical Sensation of Taste
Physiology of Special Chemical Sensation of Taste
 

The Implications of Desynchronized Circadian Rhythms in Human Mental Health and Susceptibility in the Aging Population

  • 1. THE IMPLICATONS OF DESYNCHRONIZED CIRCADIAN RHYTHMS IN HUMAN MENTAL HEALTH AND SUSCEPTIBILITY IN THE AGING POPULATION Cristina Corlito March 2010 Supervisor: Dr. Lakin-Thomas Advisor: Dr. Unniappan Course Director: Dr. Noel B.Sc. Honours Thesis York University Faculty of Science and Engineering Department of Biology
  • 2. 2 Abstract The organization of the mammalian circadian system relies on temporal order between behavioural and physiological rhythms that are critical to the normal functioning of the body and human health. The hypothesis proposed is that a disruption in the sleep- wake cycle reflects impaired circadian clock functioning, which synergistically leads to the progression and maintenance of a variety of psychiatric disorders. The aging population is most susceptible to the depletion of chronobiological rhythms and sleep deficits, and thus, the development of psychiatric disorders in the elderly warrants attention. By evaluating previous and current literature, it was found that internal temporal disorder in humans may result from both internal and external factors that disrupt the coordinated symphony of the SCN and peripheral oscillators. Sleep disorders and neuropsychiatric illnesses transpire as a result of this chronodisruption. Evidence suggests that sleep disturbances are a causal factor of psychiatric illness, rather than being mere complications. It is proposed that senescence not only predisposes the elderly to chronodisruption and sleep deficits, but also increases their risk for developing frequently comorbid psychiatric illnesses. Increasing public awareness of the multitude of strategies available for harmonious synchronization and optimal well-being are profitable to the elderly in preventing circadian malfunction.
  • 3. 3 Table of Contents Introduction ....................................................................................................................... 4 Introduction and Overview of Biological Rhythms in Mammals: The Origin and Nature of Periodicities ..............................................................................................................................4 The Light-Dark Cycle, Photoreceptors and the Retinohypothalamic Tract ..............................5 The Suprachiasmatic Nucleus and its Neural Outputs .............................................................7 Core Clock Molecular Mechanisms ..........................................................................................9 Review of Literature ....................................................................................................... 12 The Impact of Molecular Clocks on Human Physiology, Behaviour and Neuronal Function: Circadian Regulation of Physiological Pathways ................................................................... 12 Peripheral Oscillators ............................................................................................................ 14 Clock Mechanism Disruptions and Internal Desynchrony Lead to Disease .......................... 16 Clocks and Circadian Sleep Disorders .................................................................................... 20 Clocks and Psychiatric Disorders ........................................................................................... 25 Synthesis and Summary ................................................................................................. 32 The Coalescence of Circadian Rhythms and Sleep Disorders and Their Synergistic Neurobehavioural Consequences ......................................................................................... 32 Clocks and Aging: The Ensuing Susceptibility to Internal Desynchrony ............................... 33 The Prevalence of Sleep Disturbances in the Aging Population ............................................ 39 The Depletion of Chronobiological Rhythms and the Development of Psychiatric Disorders with Age ................................................................................................................................. 41 Possible Treatments and Chronobiotics for Circadian Dysfunction ...................................... 44 Research Proposal........................................................................................................... 50 Alleviating Sleep Disorders to Alleviate Psychiatric Disturbances ........................................ 50 Acknowledgments ........................................................................................................... 55 References ........................................................................................................................ 56
  • 4. 4 Introduction Introduction and Overview of Biological Rhythms in Mammals: The Origin and Nature of Periodicities Cellular biology is organized in a temporal manner, with overt circadian organization pervading all cells of the mammalian system. Virtually all organisms exhibit behaviours that follow circadian cycles of rhythmicity, allowing them to operate in synchrony with the environment. Such rhythmicity is the function of a biological clock that is endogenous to the organism (Aschoff, 1965). Several lines of evidence demonstrate that these biological clocks are inherent to living systems. First and foremost is the fact that behaviours continue to cycle in the absence of environmental time cues, negating the idea that rhythmicity is simply a reflexive response to periodicities in the environment (Aschoff, 1965). Biological oscillations are defined by periods, measured as the amount of time between two identical phases of behaviour (Aschoff, 1965). Additional supporting evidence is observed when an organism is exposed to an asynchronous environment lacking external time cues, such as continuous darkness, for example, where it will reveal behavioural rhythms with a periodicity of approximately twenty-four hours (Aschoff, 1965). These rhythms are appropriately termed ‘circadian rhythms,’ derived from the Latin terms circa, meaning approximately, and dies meaning day (Aschoff, 1965). Therefore, a circadian clock shows endogenous unremitting oscillations that free-run under constant conditions and displays a periodicity of about twenty-four hours (Aschoff, 1965). In a synchronized state, the circadian rhythm of a mammal is entrained to the rotation of the earth about its axis through external time cues known as Zeitgebers, of which the light-dark cycle is dominant, followed by temperature (Aschoff, 1965). One
  • 5. 5 theory on the origin of periodicities postulates that life originated on Earth in the face of bombardment by cosmic rays (He et al., 2000). Those cells that attempted to replicate during daylight were destroyed, while those replicating at night when radiation was at a minimum proliferated (He et al., 2000). An archetypal biological clock arose in order to confine DNA replication to the dark period as cells exploited these periodic signals for their survival (He et al., 2000). Zeitgebers, therefore, maintain a sense of harmony between the periodicity of mammals and that of the environment. Internal biological timekeeping mechanisms in mammals allow them to anticipate those physiological states which are best suited to responding to future environmental events (Aschoff, 1965). Without Zeitgebers, circadian rhythms drift out of phase with the environment (Aschoff, 1978). In addition to the aforementioned properties of circadian clocks, clocks are temperature compensated, maintaining constant periodicities despite changes in physiological temperatures. The mechanisms underlying this process, however, are as of yet unknown. The Light-Dark Cycle, Photoreceptors and the Retinohypothalamic Tract The sustained cyclical nature of biological systems in spite of a lack of Zeitgebers demonstrates that these rhythms are produced by an endogenous circadian clock and not by the daily periodicities of the environment. In mammals, the so-called master circadian clock is located in the suprachiasmatic nucleus (SCN) of the anterior hypothalamus (Lucas et al., 2001). The key element of circadian rhythms is the ability of the clock to be resynchronized by photoentrainment, that is, the ability to coordinate internal physiological rhythms with external rhythms (Lucas et al., 2001). This is done by consulting with changes in light luminance and wavelength over the period of a twenty-
  • 6. 6 four hour day and adjusting accordingly with phase delays or advances, allowing mammals to perform behaviours at suitable times in correlation with different seasons and different locations (Lucas et al., 2001). In mammals, light is the dominant Zeitgeber for entrainment, with the retinohypothalamic tract (RHT) as the main circuit through which light information reaches the SCN (Hattar et al., 2002). Lucas et al. (2001) demonstrated the lack of contribution by rod and cone photoreceptors to circadian photosensitivity by generating rodless coneless mice by combining a cl transgene with a diphtheria-based toxin rdta. These rodless coneless mice continued to display phase shifts of locomotor activity and suppression of pineal melatonin at night in response to light pulses (Lucas et al., 2001). As a result, an alternative retinal component was thought to be necessary for photoreception. A subset of retinal ganglion cells (RGCs) containing the photopigment melanopsin form the retinohypothalamic tract that projects to the suprachiasmatic nucleus (Fig. 1) (Hattar et al., 2002). Hattar et al. (2002) established melanopsin as a circadian photoreceptor by injecting RGCs with Lucifer Yellow for fluorescent labeling and then staining them for melanopsin immunoreactivity. Melanopsin-positive RGCs appeared to be directly sensitive to light and innervated neurons of the SCN, which then managed the non-visual lighting information and aligned the circadian clock accordingly (Hattar et al., 2002). RGCs thus form a large photosensitive receptive field within the outer nuclear layer of the mammalian retina that detects environmental illumination, permitting favourable physiological and neurobiological responses (Hattar et al., 2002).
  • 7. 7 Figure 1 The retinohypothalamic tract and photic input pathway in the circadian timekeeping system of mammals. Intrinsically photosensitive retinal ganglion cells containing the photopigment melanopsin are located in the outer nuclear layer of the mammalian retina. The axons of RGCs transmit information on the light-dark cycle in the environment directly to the SCN via the RHT (Kavakli and Sancar, 2002). The Suprachiasmatic Nucleus and its Neural Outputs Numerous studies have confirmed the suprachiasmatic nucleus of the anterior hypothalamus as the central circadian oscillator in mammals. Lesion and transplantation studies performed by Ralph et al. (1990) produced arrhythmic mice through the ablation of the SCN, resulting in locomotor and metabolic activities that lacked periodicities of any kind. Subsequent tissue transplantation of a wild-type donor SCN into the host resulted in the restoration of circadian rhythmicity as seen in overt behaviours. Through neural projections and hormonal signals the SCN synchronizes other central oscillators in the brain, whose efferent projections then coordinate the physiological activity of target organs directly or indirectly (Panda and Hogenesch, 2004; Reiter, 1991). Projections from the SCN to the subparaventricular zone of the hypothalamus are transmitted to the medial preoptic region, which conducts thermoregulation in a circadian manner (Moore,
  • 8. 8 1983; Panda and Hogenesch, 2004). The subparaventricular zone also projects onto the dorsomedial nucleus of the hypothalamus, which monitors fluctuating hormone levels and the cycle of sleep and wakefulness (Moore, 1983; Panda and Hogenesch, 2004). Once external light-dark cues have reached the suprachiasmatic nucleus, responses are evoked in the dorsomedial nucleus and subsequently the paraventricular nucleus (PVN) (Klein et al., 1971). Neurons of the PVN synapse onto the preganglionic sympathetic neurons in the intermediolateral zone of the lateral horns of the thoracic spinal cord (Klein et al., 1971; Perreau-Lenz et al., 2003). These preganglionic neurons influence neurons in the superior cervical ganglia whose efferent fibres innervate the pineal gland (Fig. 2) (Klein et al., 1971). The pineal gland produces the neurohormone melatonin from tryptophan and secretes it into the bloodstream, allowing it to mediate the brainstem circuits that control the sleep-wake cycle as it promotes sleep (Perreau-Lenz et al., 2003). During scotophase, or the dark period of the light-dark cycle, melatonin secretion reaches a maximum since the SCN has reduced activation, thus removing the inhibition of sympathetic neurons and promoting melatonin production in the pineal gland (Perreau-Lenz et al., 2003). It follows that the neuronal and hormonal outputs of the SCN to autonomous oscillators in the brain and their ensuing projections coordinate the proper timing of a variety of physiological functions over the twenty-four hour geophysical day, including but not limited to the sleep-wake control system, hormone secretion, hunger and satiety and body temperature (Panda and Hogenesch, 2004).
  • 9. 9 Figure 2 Brain anatomy involved in the production of the neurohormone melatonin. Non-visual light information, or lack thereof, is transmitted to the SCN and induces a response in the PVN. Neurons of the PVN innervate the preganglionic sympathetic neurons of the thoracic spinal cord, which then synapse onto the superior cervical ganglia whose efferent fibres innervate the pineal gland. The pineal gland is the production site for melatonin and mediates its release into the bloodstream (Lewy, 2010). Core Clock Molecular Mechanisms Mammalian circadian rhythms of sleeping and waking, hormone secretion, thermoregulation and the like are regulated by the molecular circadian clock mechanism, intrinsic to every cell of the suprachiasmatic nucleus of the anterior hypothalamus and every other cell of the human body. Historically, Drosophila melanogaster has been the model organism in which many of the major circadian clock genes were first identified (Clayton et al., 2001). Many homologues of the genes and proteins involved in the generation of biological oscillations in Drosophila have now been cloned in mammals (Clayton et al., 2001). The actions of these genes and proteins are temporally regulated, thus giving rise to a mammalian molecular core clock mechanism that consists of
  • 10. 10 transcription-translation autoregulatory feedback loops with both excitatory and inhibitory components (Fig. 3) (Clayton et al., 2001; Shearman et al., 2000). The fundamental factors in this molecular mechanism are the transcription factors CLOCK and BMAL1, with helix-loop-helix and PAS domains that are indicative of their function (Shearman et al., 2000). The PAS domains allow these two transcription factors to dimerize through protein-protein interactions (Clayton et al., 2001). In the circadian processes of the core clock molecular mechanism, Clk and Bmal1 genes are transcribed and their protein products heterodimerize when adequate concentration levels are reached (Shearman et al., 2000). These dimers bind to regulatory DNA sequences, or E-boxes, that initiate the transcription of the mammalian cryptochrome genes Cry1 and Cry2, period genes Per1, Per2, and Per3 and clock- controlled genes Ccg (Clayton et al., 2001; Shearman et al., 2000). Following post- translational modifications, cytoplasmic PER2 and CRY heterodimerize and translocate to the nucleus, where PER2 proteins stimulate the synthesis of BMAL1, in antiphase with the concentration of the three PER proteins, thereby acting as a positive regulator of the BMAL1 loop (Shearman et al., 2000). Conversely, CRY proteins bind to CLOCK- BMAL1 dimers, inhibiting the stimulatory effect they exert on the DNA sequences encoding PER and CRY, and forming a negative feedback loop (Shearman et al., 2000). Consequently, as the concentrations of PER and CRY decrease, the inhibition is removed and CLOCK-BMAL1 dimers recommence transcription and maintain the twenty-four hour oscillation of the circadian clock mechanism (Clayton et al., 2001). The characteristic twenty-four hour periodicity of the circadian clock mechanism is generated through post-translational modifications, phosphorylation and cellular
  • 11. 11 localization and stability of protein products, leading to time delays (Ikeda et al., 2003). Cytosolic factors, such as ions and second messengers, exhibit biological oscillations and support transcription-translation feedback loops (Ikeda et al., 2003). In this fashion, each standalone circadian clock cell is composed of a molecular oscillator and the rhythm maintaining effects of post-translational mechanisms. Further investigation of this process and other clock components is currently underway. Overall, the molecular clock produces a synchronized rhythmic output from the SCN, which is then conveyed synaptically and humorally to other brain oscillators and peripheral tissues (Clayton et al., 2001). In this manner, the output of the master oscillator coordinates various physiological and behavioural mammalian systems. Figure 3 The molecular oscillator within circadian clock cells of the SCN. Transcription-translation feedback loops with excitatory and inhibitory components regulate internal rhythmicity within the cell (Clayton et al., 2001).
  • 12. 12 Mammalian physiological and behavioural systems exhibit biological oscillatory patterns that harmonize internal conditions with the rhythmic external world. This review will focus on the implications of desynchronized circadian rhythms in human mental health. The hypothesis to be proposed is that a disruption in the sleep-wake cycle reflects impaired circadian clock functioning, which synergistically leads to the progression and maintenance of psychiatric disorders. The aging population is most susceptible to the depletion of chronobiological rhythms, and thus, the development of psychiatric disorders in the elderly merits attention. This will be done by evaluating previous and current literature and proposing future directions. Review of Literature The Impact of Molecular Clocks on Human Physiology, Behaviour and Neuronal Function: Circadian Regulation of Physiological Pathways The endogenous circadian clock permeates through almost every physiological and behavioural process in the human body and has wide implications for health. Humans exhibit circadian rhythmicity in such behaviours and physiological processes as sleep and wakefulness, hormone secretion, thermoregulation, feeding, metabolism, attentiveness and memory (Clayton et al. 2001; Takahashi et al., 2008). The clock genes that sustain biological oscillations in the suprachiasmatic nucleus also regulate the corresponding twenty-four hour human cycle through rhythmic neural and hormonal output signals to peripheral tissues (Yoo et al., 2004). In this manner, peripheral organs are able to modify their temporal functioning accordingly (Fig. 4) (Yoo et al., 2004).
  • 13. 13 Figure 4 Circadian rhythmicity in human behaviours and physiological processes. Rhythmic solar signals entrain the cellular clocks of neurons in the SCN through afferent retinal innervations. Neural and hormonal outputs from the SCN subsequently adjust the phase of peripheral organs throughout the body. The SCN employs both neural efferents and humoral signals to entrain other brain oscillators, whose roles in coordinating various physiological processes are crucial (Silver et al., 1996; Abe et al., 2002). Recall the abovementioned projections from the SCN to the subparaventricular zone, medial preoptic nucleus, and dorsomedial hypothalamus, which manage endocrine and autonomic systems, including the hypothalamic-pituitary- adrenal axis, through supplementary neuronal projections and hormonal signals (Moore et al., 1983; Girotti et al., 2007). The neurotransmitters glutamate and GABA mediate synaptic transmission between the SCN and other hypothalamic areas (Hermes et al., 1996). Using an in vitro postmortem anterograde tracing method, Jiapei and colleagues (1998) found that the hypothalamic areas innervated by the SCN mediate parasympathetic and sympathetic signaling centres, the sleep-wake control system, the arousal system, locomotor activity, body temperature, cardiovascular activity and hormone secretion. Therefore, the circadian regulation of the synthesis and release of central nervous system neurotransmitters and neuropeptides, and the ensuing entrainment
  • 14. 14 of brain oscillators, ensures the synchrony of tissue rhythms with the twenty-four hour geophysical day. The outputs of the SCN master clock not only coordinate brain oscillators but also peripheral organs in order to orchestrate physiological oscillations. In addition to hormonal cues, direct neural control of peripheral targets is achieved via the autonomic nervous system (Cailotto et al., 2005). The SCN may also indirectly control the phase of peripheral oscillators by regulating the sleep-wake cycle and therefore the rhythm of feeding behaviour (Cailotto et al., 2005). As a result, an inconsistent feeding schedule can disturb the harmonious alignment between the SCN and the peripheral organs involved as it acts as an external entraining agent (Cailotto et al., 2005). Therefore, the transcription-translation feedback loops of the core clock mechanism not only maintain rhythmicity in the central oscillator and its clock-controlled genes but also generate circadian outputs to peripheral targets, the details of which are as of yet not fully understood (Duffield et al., 2002). Peripheral Oscillators It has been established that the clock genes expressed in the core mechanism of the master suprachiasmatic nucleus are rhythmically expressed in peripheral circadian oscillators located throughout the human body (Yamazaki et al., 2000; Duffield et al., 2002; Yoo et al., 2004). Balsalobre et al. (2000) examined clock-controlled gene expression in peripheral mammalian tissues by inducing rhythmicity in immortalized rat- 1 fibroblasts through serum shock. cDNA microarrays revealed a chronological production of messenger RNA in response to serum shock and pharmacological treatment, signifying the circadian regulation of gene expression (Balsalobre et al., 2000).
  • 15. 15 Previously believed to rhythmically dampen after two to seven twenty-four hour cycles without input from the SCN, Yoo and colleagues (2004) have demonstrated that peripheral oscillators maintain endogenous rhythmicity whilst displaying desynchrony amongst themselves without entraining signals (Yamazaki et al., 2000). Yoo et al. (2004) employed the fusion of the mouse locus mPer2 with a luciferase reporter gene to reveal strong inherent oscillations of bioluminescence in both the SCN and peripheral tissues ex vivo. Furthermore, in SCN-lesioned mice, bioluminescence rhythms persisted for twenty days in peripheral tissues, including the liver, lungs, pituitary and cornea, however, with an eventual loss of phase coordination (Fig. 5). Figure 5 Circadian rhythmicity in explanted tissues of the mouse, including the cornea, liver, pituitary gland, kidney and lung. Luciferase reporter genes revealed inherent oscillations of bioluminescence in these tissues (Yoo et al., 2004). Peripheral tissues isolated in culture, including but not limited to the previously mentioned lungs, cornea, pituitary gland and liver, express clock-controlled genes that confer distinctive circadian period and phase properties to those structures (Yoo et al., 2004). As such, these circadian properties are distinct in different organs and contribute temporally to their physiological functioning (Yoo et al., 2004). The SCN does not
  • 16. 16 generate but rather coordinates the phase of autonomous peripheral oscillators, thereby inhibiting internal desynchrony between tissue-specific target clocks and their synchronized phase relationship with the external environment (Yamazaki et al., 2000; Yoo et al., 2004). The phase of each peripheral oscillator induces rhythmic gene expression, for example that of Per1, resulting in circadian protein product activity, which in turn regulates rhythmic metabolic events in different tissues throughout the human body (Ripperger et al., 2000). The phase of oscillations can be altered by adjusting the feedback of peripheral clocks characteristic of a tissue to internal and external Zeitgebers originating from the SCN and environment, respectively (Yamazaki et al., 2000). Clock Mechanism Disruptions and Internal Desynchrony Lead to Disease The organization of the mammalian circadian system, as reviewed above, relies on temporal order between behavioural and physiological rhythms that are critical to the normal functioning of the body and human health. Thus, the concept of the harmful effects that would ensue as a result of disorder between these phase relationships and the cyclical expression of clock-controlled genes readily presents itself as there are numerous avenues through which to disrupt this fragile system (Yamazaki et al., 2000). Internal temporal disorder in humans may result from both internal and external factors that disrupt the coordinated symphony of the SCN and peripheral oscillators. Predominant external factors include light deficiency and irregularity, jet lag, shift work, food intake and social activities (Skene et al., 1999; Reddy et al., 2002; Solonin et al., 2009; Turner and Mainster, 2008; Girotti et al., 2009). Internal factors include the disturbance of proper photoreception, visual loss, decreased melatonin levels and circadian clock gene
  • 17. 17 mutations (Czeisler et al., 1995; Lockley et al., 1997; Turner and Mainster, 2008). As previously discussed, peripheral oscillators will desynchronize amongst themselves without temporal adjustments provided by the SCN through neural and hormonal outputs (Yoo et al., 2004). Proper SCN operations ensure good health by mediating rhythms of sleep-wake systems, hormone secretion and metabolism, therefore, chronodisruption may be the cause of a range of diseases (Jiapei et al., 1998; Yamazaki et al., 2000; Cailotto et al., 2005). Light irregularity, or improperly timed ocular light exposure, may result in chronodisruption by modifying nocturnal melatonin synthesis in the pineal gland depending on its duration, wavelength, intensity and time of administration (Czeisler et al., 1995; Skene et al., 1999). Ocular light exposure in the scotophase decreases melatonin production (Skene et al., 1999). Low circulating levels of melatonin throughout the body may result in numerous diseases as it has been shown to contribute beneficially to the antioxidant capability of blood plasma (Benot et al., 1999). Environmental light is the predominant Zeitgeber in circadian timekeeping, and, for that reason, maintains the greatest influence on human physiological and psychological health. Photosensitive RGCs best absorb light in the blue sector of the light spectrum at 460 nm, which is quite similar to the wavelength of environmental light (Turner and Mainster, 2008). Modern artificial lighting, unfortunately, provides only about 1% of natural light intensity and is distinguished by red spectrum wavelengths, which is insufficient for suitable photoreception (Turner and Mainster, 2008). Instead, optimal photoreception requires blue light of high intensity and duration for photoentrainment and favourable health (Turner and Mainster, 2008).
  • 18. 18 Light deficiency fails to entrain the SCN to the geophysical day and results in a subsequent free-running periodicity, as exemplified by blind individuals (Skene et al., 1999). Blind individuals may be categorized, according to the extent of visual loss, as having some light perception, and thus photoreception, and those with no light perception capabilities and no photoreception whatsoever (Skene et al., 1999). In a study by Skene et al. (1999), 77% of blind subjects capable of photoreception showed normal circadian rhythmicity, while 67% of those with no light perception showed free-running period lengths and internal desynchrony. The latter also suffered from daytime somnolence, an excessive need for sleep during the daytime, and insomnia during the night due to the temporal disorder of melatonin synthesis and release (Lockley et al., 1997). Further evidence of the dire consequences of light deficiency, with light as the chief biological Zeitgeber, is demonstrated by the fact that blind subjects who had no eyes after having undergone bilateral enucleation showed free-running period lengths ranging from 24.13 to 24.81 hours, albeit in the presence of non-photic signals such as food intake and social activities (Skene et al., 1999). On the whole, blind individuals incapable of photoentrainment exhibit higher levels of circadian disruption and dampened SCN outputs, thus making them susceptible to diseases, particularly sleeping disorders and compromised neuropsychiatric conditions (Lockley et al., 1997; Jean-Louis et al., 2005). Girotti and colleagues (2009) recently demonstrated the role of food intake as a non-photic Zeitgeber. Their study revealed characteristic rhythms of clock gene expression in each element of the hypothalamic-pituitary-adrenal axis, where a decrease in food intake in the photophase of the light-dark cycle altered glucocorticoid secretion and clock gene expression (Girotti et al., 2009). Physiological processes may be
  • 19. 19 entrained to intermittent feeding schedules, with glucocorticoids synchronizing a multitude of peripheral organs (Stephan, 1986; Girotti et al., 2009). Shift work employees in industrial sectors, medicine and the military show signs of considerable circadian dysfunction, including such symptoms as biochemical disturbances, mood disorders, sleeping disorders, metabolic syndrome and an overall feeling of malaise (Solonin et al., 2009). James et al. (2007) recently investigated the effects of night shift work on the sleep-wake cycle, outlining desynchrony between the master circadian clock and the night schedule as subjects maintained day active entrainment. This was done by comparing oscillatory clock gene expression in peripheral blood mononuclear cells with the temporally shifted sleep-wake cycle. The exposure to artificial light at uncharacteristic times communicates odd entraining signals to the SCN and results in perturbed circadian outputs and melatonin synthesis (James et al., 2007). The differential responses of peripheral oscillators to the altered phase of input signals also leads to a lack of internal coordination, hence resulting in feelings of malaise. Shift work sleep disorder is a circadian rhythm sleep disorder in which the afflicted complain of daytime sleepiness, insomnia and poor sleep quality (Ursin et al., 2009). Other circadian rhythm sleep disorders will be addressed in the next section of this review. In addition to shift work, jet lag also impairs physical and mental well-being via circadian desynchrony. The core clock mechanism experiences much more difficulty in acclimatization to advanced time zones rather than delayed time zones because the former does not occur as rapidly (Reddy et al., 2002). Reddy and his colleagues (2002) subjected mice to acute advances or delays in local time and reported that circadian rhythms of mPer expression in the SCN adjust swiftly to advanced light pulses, while
  • 20. 20 rhythmic mCry1 expression advanced gradually. Conversely, they found that a six hour delay in local time entailed mPer and mCry adjusting in sequence by the second oscillation. This study describes the different effects of traveling east or west, or advancing or delaying, respectively, on the master clock and the prospective temporal desynchrony between mPer and mCry expression as a result of jet lag, with ensuing health complications (Reddy et al., 2002). The final factor contributing to clock mechanism disruptions, internal desynchrony, and thus, disease are genetically mutated circadian clock genes and polymorphisms. Current research is heavily focused on identifying those clock gene alterations that result in a variety of disrupted circadian behaviours. Clock gene mutations may be implicated in the deterioration of the regimented functioning of both molecular oscillators and their rhythmic neural and hormonal outputs and their effects will be discussed in subsequent sections of this review. Cumulatively, the aforementioned internal and external factors, particularly insufficient and temporally displaced environmental light, may induce biological stress and disturb the coordinated rhythmicity of physiological processes and behaviours necessary for optimal human health. This review will now turn to those sleep disorders and neuropsychiatric illnesses that transpire as a result of chronodisruption. Clocks and Circadian Sleep Disorders A profound relationship exists between clock gene variations and changes in behavioural rhythmicity, most notably sleep parameters in humans. The timing and amount of sleep are determined by circadian and homeostatic sleep control mechanisms, respectively (Naylor et al., 2000). The former dictates patterns of sleep and wake at
  • 21. 21 specific phases throughout the twenty-four hour light-dark cycle, while the latter depends on the need for sleep (Naylor et al., 2000). In a study by Naylor et al. (2000), a mutation in Clk in the mouse was found to alter not only the timing and length of sleep but sleep homeostatis as well. Naylor and colleagues evaluated the effects of the CLOCK transcription factor mutation by comparing sleep and electroencephalographic (EEG) activity in homozygous and heterozygous mutants and wild-type mice under conditions of entrainment, free-running rhythms and recovery from six hours of sleep deprivation. The results indicated that heterozygotes slept one hour less per day and homozygotes two hours less per day in contrast to wild-type mice, with lower amounts of non-rapid eye movement sleep seen. Following periods of sleep deprivation, Clk homozygous mice displayed 39% less sleep than heterozygotes and wild-type mice. One may attribute these results to discrepancies of entrainment to the light-dark cycle, however, divergent sleep behaviours were also seen when mice were free-running in continuous darkness (Naylor et al., 2000). A study by Laposky et al. (2005) employed the same investigative measures in mice with a deletion of Bmal1 and discovered a diminished rhythm of sleep and wakefulness, a weakened response to sleep deprivation and lengthened sleep periods. Whereas mutations in the mammalian cryptochrome genes Cry1 and Cry2 hold implications for sleep homeostatis, Period genes are not essential for homeostatic sleep regulation (Wisor et al., 2002; Shiromani et al., 2004). A study by Shiromani and collaborators (2004) examined the effects of Per1, Per2, Per3 and double Per1-Per2 mutations on sleep factors and found that Per2-mutant and double mutant mice exhibited longer periods of wakefulness, with less slow-wave sleep (SWS) and rapid-eve movement (REM) sleep, than wild-type and Per1 deficient mice in states of entrainment.
  • 22. 22 Double mutant strains became arrhythmic in aperiodic conditions, however, the amount of time spent awake, in SWS and in REM sleep was equivalent to that in an entrained state even after 36 days, thus signifying the maintenance of total sleep levels. Per genes are more so involved in altering the phase position of the sleep-wake cycle (Shiromani et al., 2004). In conclusion, circadian clock gene alterations have profound implications for both rhythms of sleep and wakefulness and sleep propensity, although knowledge as to their exclusivity to one or the other is currently unknown. Variations in circadian clock genes have a variety of effects on the configuration of human sleep. Delayed sleep phase syndrome (DSPS) is the most commonly reported circadian rhythm sleep disorder whose features include sleep periods delayed by 2 to 6 hours, the inability to fall asleep, difficulty waking and a lack of feeling well rested (Campbell and Murphy, 2007; Chang et al., 2009). One study investigated a 30 year old graduate student with DSPS whose average bedtime was 3:38 a.m. and usually awoke at 1:47p.m. in order to feel replenished (Campbell and Murphy, 2007). Campbell and Murphy (2007) examined the sleep and body temperature rhythms of the subject in comparison to those of 3 normal age-matched subjects with both parties in aperiodic conditions free from environmental cues. Whereas the time between the core body temperature minimum and sleep onset in control subjects was 1.63 hours, the graduate student displayed a phase angle of 3.62 hours. Furthermore, the DSPS patient had a free- running period length of 25.38 hours compared to an average of 24.44 hours for the control subjects. One may refute these results by suggesting that the lighting conditions in temporal isolation contributed to the lengthening of the free-running period in the DSPS subject, however, the authors noted that illumination was below 50 lux, which is
  • 23. 23 insufficient for optimal photoreception (Campbell and Murphy, 2007). Therefore, DSPS causes an abnormal endogenous period length and internal desynchrony between sleep and body temperature rhythms, resulting in poor sleep efficiency and duration. The previous findings may be attributed to a polymorphism in the circadian clock gene Per3 or a missense mutation in the casein kinase I epsilon gene CKI ε. Archer et al. (2003) have found a correlation between a length polymorphism in Per3 and DSPS, particularly the shorter allele for which 75% of DSPS patients were homozygous. They found that the 4-repeat allele, as opposed to the 5-repeat allele, was substantially prevalent in DSPS patients in contrast to the control group. Recall post-translational mechanisms, such as phosphorylation, affect the stability of protein products and function to create time delays in the circadian clock mechanism. PER is targeted for degradation through phosphorylation by CKI ε, making it unavailable for dimerization and subsequent nuclear localization and thus causing it to oscillate (Archer et al., 2003). The shorter variation of PER3 contains fewer phosphorylation sites than its longer counterpart and may be the cause of polymorphic differences in function and hence longer endogenous period length seen in DSPS (Archer et al., 2003). A study by Takano and associates (2004) revealed that a missense mutation in the N408 allele in CKI ε functions as a safeguard against DSPS by modifying its autophosphorylation activity. In contrast to DSPS, advanced sleep phase syndrome (ASPS) dictates human behaviours marked by early bedtimes, early morning waking and a short endogenous period length (Xu et al., 2005). ASPS is caused by a mutation in a residue in the casein kinase I binding site of the Per2 gene and results in attenuated phosphorylation levels (Archer et al., 2003). The reduction in phosphorylation seen in both DSPS and ASPS is
  • 24. 24 indicative of the different pathological symptoms that may occur as a result of phosphorylation levels in different PER proteins. In another case, through mutagenesis screenings of related ASPS patients, Xu et al. (2005) found a threonine to alanine missense mutation at amino acid 44 in the human CKIδ gene. Subjects under study had an average bedtime of 6:12 p.m., compared to the control average of 11:24 p.m., and an average rising time of 4:06 a.m. compared to the control average of 8:00 a.m. Overall, this T44A mutation decreases CKIδ enzyme activity in ASPS patients and consequently leads to a shortened activity rhythm and advanced phase of activity in an entrained setting of 12 hours of light and 12 hours of dark (Xu et al., 2005). The etiology of the abovementioned circadian rhythm sleep disorders may be attributed to circadian clock gene polymorphisms and mutations, whereas obstructive sleep apnea syndrome (OSAS) and its symptoms produce arrythmicity in clock gene functioning. This arrythmicity may be credited to fluctuating levels of factors circulating through the blood (Burioka et al., 2008). Burioka et al. (2008) have measured Per1 mRNA expression in peripheral blood mononuclear cells in those patients with severe OSAS using polymerase chain reaction analysis over a twenty-four hour period. In contrast to similar healthy controls, the eight OSAS participants showed no circadian rhythms of Per1 mRNA expression throughout the day and abnormal elevations of plasma noradrenaline. Elevated noradrenaline levels and sympathetic activity contributed to an increase in the transcription of Per1 during sleep (Burioka et al., 2008). Interestingly, continuous positive airway treatment for a period of three months improved not only shallow sleep with frequent waking due to hypoxic episodes, but also daily oscillations of Per1 transcription (Burioka et al., 2008). The effects of continuous
  • 25. 25 positive airway treatment on clock gene transcription thereby illustrate a mechanism by which a circadian rhythm sleeping disorder may be managed by improving clock gene function. Fatal familial insomnia (FFI) is a debilitating disorder marked by sleep deficiency. FFI is a prion disease distinguished by a 178 codon prion protein gene mutation (Reder et al., 1995). A study by Sforza et al. (1995) studied six subjects with this disease using twenty-four hour polygraphic recordings in a sleep laboratory. Their findings revealed severe reductions in total sleep time, impairments in the circadian regulation of the sleep-wake cycle and abrupt alterations from wakefulness to sleep. Positron emission topography uncovered atrophy in the thalamus, particularly the antero- ventral and dorso-medial thalamic nuclei, which take part in regulating the sleep-wake cycle (Sforza et al., 1995). Over the course of the disease, symptoms of insomnia progressively worsen and circadian rhythms dampen substantially until the total sleep time is reduced to about 50 minutes per day and the subject dies (Sforza et al., 1995). Portaluppi et al. (1994) conducted assays for melatonin in the blood plasma of two FFI patients and found that concentrations of the hormone decreased in accordance with disease progression, further compromising circadian rhythmicity. Clocks and Psychiatric Disorders Just as the misalignment of the circadian pacemaker and clock gene mutations and polymorphisms have been associated with circadian rhythm sleep disorders, these factors are implicated in psychiatric disorders. A variety of abnormal endogenous circadian rhythms underlie major depressive disorder (MDD), particularly the sleep-wake cycle (Gordijn et al., 1994; Emens et al., 2009). Emens and colleagues (2009) set out to
  • 26. 26 demonstrate a correlation between MDD and improper coordination between the circadian pacemaker and sleeping schedule. Study subjects were comprised of eighteen females ranging from 19 to 60 years of age who had been diagnosed with MDD according to the Diagnostic and Statistical Manual of Mental Disorders (DSM), excluding those with suicidal tendencies, jet lag, shift work positions and medications that would impede melatonin production. Emens et al. (2009) calculated circadian misalignment according to the time difference between melatonin production and the midpoint of sleep. Those with larger time differences exhibited a phase delay in central pacemaker rhythmicity in comparison to the timing of sleep and a higher severity of symptoms (Fig. 6). These results demonstrate the interaction between circadian desynchrony, poor sleep and mild to moderate symptoms of depression, though future studies should be conducted on a more representative sample population (Emens et al., 2009). Figure 6 The larger the time difference between melatonin synthesis and the midpoint of sleep, also known as the phase angle difference (PAD), the higher the severity of depressive symptoms according to the Hamilton Depression Rating Scale (HAM-D). Following a clinical assessment by a health professional, a score higher than 7 on the HAM-D constitutes a diagnosis of MDD (Emens et al., 2009). An alternative route by which disrupted circadian oscillations may facilitate MDD is through the deregulation of mood by the mesolimbic dopaminergic system (Hampp et
  • 27. 27 al., 2008). Hampp and associates (2008) ascertained that Per2 mutant mice have reduced levels of expression of monoamine oxidase A in the mesolimbic dopaminergic system, which is an enzyme that mediates dopamine metabolism. The atypical mood behaviours observed in these mice may be attributed to this clock gene mutation. Polymorphisms and mutations in the Clock gene have been connected to bipolar disorder (Benedetti et al., 2003; Roybal et al., 2007). Roybal and colleagues created Clock mutant mice through mutagenesis, thereby inhibiting its transcriptional activation of molecular rhythms. The mice were subjected to tests in which they were able to induce rewarding electrical stimulation to themselves via electrodes implanted in the medial forebrain bundle. Clock mutant mice were able to experience euphoria at lower currents than wild type mice and cocaine decreased these current thresholds substantially in the mutants. This response is predictive of substance abuse as the mice experienced a greater sense of reward upon stimulation because of their hypersensitivity, making them more inclined to abuse such stimulants (Roybal et al., 2007). These states of ecstasy and substance abuse mimic the condition of bipolar patients (Roybal et al., 2007). The mood-related behaviours of Clock mutant mice parallel those humans with bipolar disorder, including less depression and less anxiety (Roybal et al., 2007). This was discerned as mice showed little anxiety when subjected to an unprotected arm of a raised platform. Conversely, when treated with lithium, a mood stabilizer given to bipolar patients, the mutant mice displayed more wild-type behaviours of high anxiety in this situation (Roybal et al., 2007). Like the previously mentioned MDD patients, Clock mutant mice have compromised dopaminergic systems, although with increased firing of dopaminergic neurons that is diminished through the viral insertion of a gene coding for a
  • 28. 28 wild type CLOCK protein (Roybal et al., 2007). As the name implies, bipolar disorder alternates between states of mania and depression, with depressive states being predominant in the winter months (Roybal et al., 2007). Winter depression, also known as seasonal affective disorder (SAD), involves changes in circadian genetic factors, the external environment and circulating melatonin (Wehr et al., 2001; Johansson et al.., 2003; Partonen et al., 2007). Certain animals display photoperiodism, that is, the ability to infer the time of year based on the length of the day. Such information is made available by measuring the duration of melatonin release during the night, the duration of which is longer in the winter (Wehr et al., 2001). Wehr et al. (2001) found that variations in season affect patients with SAD but not similar healthy subjects. Their study measured the duration of melatonin release in dim light in 55 SAD patients and 55 equivalent healthy subjects throughout the summer and winter months, with plasma samples being taken every 30 minutes all through the day. In SAD subjects, melatonin release in the scotophase was much more pronounced in the winter rather than summer, however, no change was observed in those without SAD diagnoses. In regards to circadian genetic mechanisms, Partonen et al. (2007) surmised the genes Per2, Bmal1 and Npas2, whose products function mutually in the core circadian oscillator, are compromised in SAD. As previously mentioned, BMAL1 is a PAS protein that interacts with other proteins, and as such, dimerizes with NPAS2 and binds to DNA (Partonen et al., 2007). Single nucleotide polymorphisms were assessed in each of the three genes in 189 patients and 189 symptom-free controls. Gene-wise logistic regression analysis revealed SAD to be related to polymorphisms within all three genes and posing
  • 29. 29 the greatest chance of illness due to their magnified cumulative effects (Partonen et al., 2007). When genetic variations in all three were present, patients had a 10 times greater chance of developing SAD compared to the controls, while those with less severe allelic combinations had a 4 times greater chance. Recently, point mutations have been located in the melanopsin gene Opn4 in retinal ganglion cells, which serve to decrease photosensitivity (Roecklein et al., 2009). As the contrast between light intensities is already reduced in the winter, these point mutations aggravate that effect as dusk and dawn Zeitgebers cannot be detected and proper photoentrainment cannot occur (Roecklein et al., 2009). Such a hindrance in phototransduction may result in internal desynchrony, and thus, clinical symptoms of depression in the winter months. Deteriorations in mood behaviours are often accompanied by substance abuse, most likely due to the fact that circadian rhythmicity and dopaminergic systems are confounded in these patients. Abarca et al. (2002) investigated cocaine addiction in Per mutant mice in order to ascertain the circadian control of cocaine-induced reward and behavioural sensitization. A single cocaine injection produced a fivefold increase in locomotion in Per1 and Per2 knockout and wild type mice compared to saline injections. During cocaine administration, mice were placed in boxes with two floor divisions in which one consisted of a rod pattern of texture and the other of circles. Cocaine administration was always associated with the same floor division. After repeated cocaine injections, wild type mice became sensitized to cocaine-associated factors, Per1 mutants showed no sensitization and Per2 mutants showed an intense sensitized behavioural reaction. While both wild type and Per2 mutant mice preferred the division associated with cocaine injections, Per1 mutants did not prefer the side associated with
  • 30. 30 reward. In addition, stronger behavioural responses to the drug were seen in the morning than at night. Since all three groups displayed similar levels of locomotor activity in response to a single cocaine injection, it may be inferred that cocaine addiction, rather than acute application, is managed by the clock genes Per1 and Per2 with opposing effects in the circadian system (Abarca et al., 2002). Previous discussions highlighted aberrant dopaminergic systems and clock gene mutations in MDD and bipolar disorder mice models. These same conditions are observed in mice addicted to cocaine. McClung and colleagues (2005) found that a loss of function point mutation in Clock results in the same cocaine sensitization behaviours outlined by Abarca et al. (2002), with Clock mutants displaying a greater degree of sensitivity to the rewarding feelings of cocaine. This may be credited to increased levels of tyrosine hydroxylase activity, an enzyme involved in dopamine metabolism, and therefore, heightened amounts of dopaminergic transmission in the reward centres of mice without functional CLOCK proteins (McClung et al., 2005). Alcoholism, another form of substance abuse possibly under circadian control in humans, may be associated with excessive levels of glutamate in the extracellular fluid, as suggested by Per2 mutant mice whose levels of glutamate reuptake transporters in the nervous system are significantly reduced (Spanagel et al., 2005). These mice display augmented levels of voluntary alcohol consumption when offered ethanol in comparison to wild type controls (Spanagel et al., 2005). Malformed circadian rhythms of activity, body temperature and sleep are often prevalent in patients with Alzheimer’s disease, which is common in the elderly. Many victims of this neurodegenerative illness exhibit what is referred to as ‘sundowning’, or
  • 31. 31 the worsening of Alzheimer’s behaviours in the afternoon and evening (Volicer et al., 2001). Volicer et al. (2001) sought to decipher the relationship, if any, between sundowning and circadian rhythms in a cohort of 25 Alzheimer’s patients and nine healthy subjects. Their results revealed that those Alzheimer’s patients who undergo sundowning showed increased nocturnal locomotor activity, with lower amplitudes during the day, and major phase delays in both activity and body temperature in comparison to controls. Furthermore, these subjects had severely reduced amplitudes of body temperature and disrupted sleep parameters. These results imply that patients who sundown may be suffering from disturbances in their rhythms, however, other environmental factors must be considered in these habitual states of aggravation (Volicer et al., 2001). One study that has shed some light on this issue is by Mahlberg and associates (2008), in which cranial computed tomography revealed significant levels of pineal calcification in AD patients, thereby limiting melatonin synthesis to sup-optimal levels and inhibiting the coordination of the circadian system. Schizophrenia is a neuropsychiatric illness characterized by distorted cognition, abnormal affect and social withdrawal (Wulff et al., 2006). Schizophrenic patients show strong disturbances in their sleep-wake cycles, melatonin patterns and light exposure, yet research in this area remains vague (Wulff et al., 2006). Wulff et al. (2006) studied a 27- year old male patient for six weeks, documenting gradually delayed bedtimes and risings, which culminated in the eventual reversal of night and day activities and low sleep efficiency. His free-running period length was longer than 24 hours, as well as a free- running melatonin rhythm of 24.29 hours. Instead of being coordinated with the light- dark cycle of the environment, the patient’s light exposure was synchronized with his
  • 32. 32 own activity, and the temporal misalignment of these input light signals may exacerbate his established desynchrony even further (Wulff et al., 2006). Synthesis and Summary The Coalescence of Circadian Rhythms and Sleep Disorders and Their Synergistic Neurobehavioural Consequences There is a widely held assumption that sleep deficits are the secondary effects of psychiatric disorders. While this may certainly be the case under some circumstances, evidence would lead one to speculate that sleep disturbances are a causal factor of psychiatric illness, rather than being mere complications. A disruption in the sleep-wake cycle reflects impaired circadian clock functioning, which synergistically leads to the progression and maintenance of a variety of psychiatric disorders. Sleep-wake cycles are perturbed in most if not all of the previously mentioned psychiatric disorders, and many of the studies discussed implicated irregular clock gene functioning. Recall the study by Emens et al. (2009), whose results demonstrated a temporal misalignment between the central circadian pacemaker and the sleep-wake cycle, the degree of which corresponded to the severity of clinical psychiatric symptoms experienced. Over 10 years earlier, Boivin and collaborators (1997) released similar findings from a study in which 24 healthy adults were subjected to internal desynchrony by living according to 30 hour and 28 hour sleep-wake cycles. Through the use of psychometric response scales, they too showed varying mood states based on the degree of displacement between the sleep-wake cycle and circadian rhythmicity. The successful treatment of MDD seems to involve the management of sleep disturbances, indicating their contribution to this psychiatric disorder (Ohayon and Roth,
  • 33. 33 2003). In 2003, Ohayon and Roth interviewed a representative sample of the population of the United Kingdom, Germany, Italy and Portugal, totaling 14,915 participants. Participants were questioned about their sleep habits, sleep symptoms, current mental health status and history and subsequently diagnosed according to the Diagnostic and Statistical Manual of Mental Disorders if need be. Symptoms of insomnia arose in 19.1% of the sample, with 90% of this 19.1% cohort experiencing severe insomnia in excess of six months. Among those suffering from severe insomnia for six months to five years, 28% held concomitant psychiatric diagnoses, and of those suffering for more than five years 25.8% held diagnoses. Furthermore, insomnia preceded states of relapse in 56.2% of interviewees and came about concurrently in 22.1% of states of reversion. Taken as a whole, subjects afflicted with sleep disturbances show evidence of higher rates of psychiatric disorders than the general population and these disturbances may be presumed to be the cause of their onset or recurrence rather than transpiring as symptoms (Ohayon and Roth, 2003). Clocks and Aging: The Ensuing Susceptibility to Internal Desynchrony The aging population is most susceptible to the depletion of chronobiological rhythms. The elderly are predisposed to chronodisruption due to ocular aging and suboptimal photoreception necessary for circadian photoentrainment (Turner and Mainster, 2008). Ocular aging consists of the aging of the crystalline lens and the decreasing size of the pupil, culminating in a significant reduction of phototransduction by RGCs, with the former blocking the absorption of favourable blue light (Turner and Mainster, 2008). In 2008, Turner and Mainster calculated the levels of circadian photoreception decrease experienced throughout the aging process by multiplying human
  • 34. 34 crystalline lens transmission by pupil diameter and subsequently measuring melatonin suppression sensitivity from light sources with wavelengths between 350 and 700 nanometres (Table 1). The results illustrated an age-dependent reduction in melatonin suppression in response to blue light. From Table 1, it may be estimated that a person 95 years of age exhibits one-tenth of the level of photoreception seen in a ten year old. Likewise, a person who is 85 years old requires 7.58 times brighter light exposures than a 15 year old in order to attain equivalent levels of photoreception. Table 1 Circadian photoreception at different ages. The numbers in the table indicate the level of retinal illumination achieved by the ages listed in the top row in contrast to those in the left column. They also indicate the relative level of light exposure required by those in the left column to achieve similar levels of effective photoreception as those in the top row (Turner and Mainster, 2008). In addition to ocular aging, the elderly are particularly prone to insufficient light exposure because of their habitual lifestyles (Turner and Mainster, 2008). Reduced crystalline lens transmission and pupil diameter require brighter light exposures for the elderly in order to maintain sufficient photoreception, however, residential lighting is excessively dim and lacking in blue spectrum wavelengths compared to environmental
  • 35. 35 light (Fig 7) (Turner and Mainster, 2008). The link between circadian desynchrony and scarce bright light exposure was investigated by Campbell et al. (1988), who recorded levels of light exposure in 13 Alzheimer’s patients and 10 healthy elderly controls of similar ages. The data, based on five days of recording subjects in their natural routines at home, revealed that subjects rarely received exposure to ambient light in excess of 2000 lux. Furthermore, Alzheimer’s patients received 0.5 hours of bright illumination in comparison to one hour in the control group and, in turn, the control group received one- third to two-thirds less the amount encountered by healthy younger people. These figures are even lower in those elderly subjects who are institutionalized in nursing homes and retirement living centres (Turner and Mainster, 2008). Campbell et al. (1988) noted the sleep deficits prevalent in both of these groups, therefore, light deficiency may be implicated in circadian rhythm disturbances in the elderly, as well as the potential neuropsychiatric illnesses in which they are involved. Figure 7 Illuminance levels in a variety of settings. Residential lighting typically ranges from 100 to 500 lux, however, proper physical and especially mental health
  • 36. 36 require environmental light exposures exceeding 1000 to 3000 lux, such as sunlight and other sources of bright light (Turner and Mainster, 2008). Like the aging population, those with cataracts have inadequate levels of circadian photoreception due to reduced ocular light transmission and smaller pupils (Turner and Mainster, 2008). In these cases, the crystalline lens is surgically replaced with an intraocular lens (IOL), which typically blocks ultra-violet radiation and restores blue light-inducing photoreception, however, some IOLs block blue light, resorting to previous ophthalmologic standards (Turner and Mainster, 2008). Although some patients with IOLs are lacking in circadian light exposures, IOLs have proven to be beneficial for photoreception in the aging population (Turner and Mainster, 2008). As of yet, research on the effects of aging on photosensitive RGCs remains controversial. Together, ocular aging and light deficiency are responsible for the dampening of SCN output signals and circadian amplitudes in the elderly to some extent, consequently leading to internal desynchrony. An alternative region that appears to be associated with the attenuation of the biological timekeeping system in the aging population is the SCN. Nygard et al. (2005) attributed the weakened ability of the central oscillator to synchronize with external stimuli, the dampening of activity and temperature cycles and disruptions in the sleep- wake cycle to the altered electrophysiology of the aging SCN. Nygard and colleagues (2005) used cell-attached and whole cell recordings to study the rhythm of spontaneous firing and synaptic transmissions in the ventrolateral region of the SCN. The ventrolateral portion of the SCN receives input from the RHT and neurons in this region mediate inhibitory synaptic transmission by expressing vasoactive intestinal polypeptide (VIP) (Nygard et al., 2005). Single neurons in the ventrolateral region rhythmically
  • 37. 37 alternate between periods of silence and activity (Nygard et al., 2005). Recordings conducted on slices of the SCN in vitro showed that young mice have a smaller proportion of silent cells during the day, whereas such rhythmicity appeared to be absent in older mice as a higher proportion of silent cells fired both during the day and the night. These results point to an altered response to light in aged mice, with the SCN as a target of the aging process as seen by the modified firing properties of individual neurons and subsequently changed SCN output signals (Nygard et al., 2005). A recent study by Biello et al. (2009) demonstrated that the aging process alters the central pacemaker by diminishing its response to phase shifting stimuli. This may be the reason why the elderly exhibit advanced behavioural rhythms and lose the capacity to temporally adapt to the environment (Biello et al., 2009). Biello et al. (2009) compared the phase shifting properties of various neurotransmitters thought to be involved in entraining the SCN in young and old mice. Glutamate, histamine and NMDA all delayed the phase of rhythmicity in young mice, and thus are thought to be involved in photic pathways, however, older mice did not respond as strongly (Biello et al., 2009). The application of neurotransmitters thought to be involved in non-photic signaling pathways, Muscimol, a GABA agonist, and 8-OH DPAT, a serotonin agonist, resulted in phase advances in young mice, whereas older mice showed lesser responses. Gastrin-releasing peptide and neuropeptide Y induced comparable phase shifts in both young and old mice. These results are consistent with previous findings and signify the ability of the aging SCN to phase shift in response to some stimuli, though not all, perhaps implicating the disruption of particular synaptic pathways and neurotransmitter systems in the
  • 38. 38 dysfunction of the aging SCN (Penev et al., 1995; Palomba et al., 2008; Biello et al., 2009). Since its discovery over twenty years ago, evidence demonstrating that the pineal production of melatonin declines with aging is now a widely accepted fact. Sack et al. (1986) demonstrated this concept by performing periodic assays for melatonin’s major urinary metabolite 6-hydroxymelatonin, and hence measuring the total nocturnal production of melatonin. Sack and collaborators conducted assays for three consecutive nights in the summer and winter across a wide range of healthy adults, including medical students, hospital personnel and retirement home residents. After adjusting for demographic variables of height, weight, gender, sleep patterns, smoking, alcohol and coffee consumption, a significant negative correlation was found between age and melatonin for both men and women. The same results were attained by Zhou et al. (2003) by performing assays for melatonin on saliva, which also revealed that the decline in cyclic melatonin production begins in middle-age, with these subjects having 60% of the amplitude measured in young controls. Consideration must be given to the possibility that such weakened melatonin levels in old age may be due to light deficits. Collectively, the age-related factors of light deficiency, ocular aging, deteriorated electrical SCN rhythms, altered neurotransmitter signaling and diminished melatonin production may be responsible for the vast array of circadian perturbations observed in the elderly. Such perturbations include the dampening of circadian amplitudes and output signals, the inability to synchronize with the environment, cognitive impairment, advanced activity phases, lengthened free-running circadian period lengths, sleep disturbances and psychiatric disorders (Nygard et al., 2005; Turner and Mainster, 2008;
  • 39. 39 Biello et al., 2009). The most frequently reported sleep-wake cycle alterations in old age include fragmented sleep with less restoration, an overall phase advance of the sleep- wake cycle with earlier bedtimes and earlier awakenings, increased daytime drowsiness and insomnia (Campbell et al., 1988; Biello et al., 2009). The Prevalence of Sleep Disturbances in the Aging Population Recall that the timing and amount of sleep are determined by circadian and homeostatic sleep control mechanisms, respectively (Naylor et al., 2000). Dijk et al. (1999) investigated the interplay of the circadian pacemaker and homeostatic factors in sleep regulation and how they change with aging. The circadian rhythms of 13 older subjects, ranging from 65 to 75 years old, and 11 younger subjects, ranging from 20 to 30 years old, were assessed using polysomnographic recordings and it was found that older people wake up one hour earlier than predicted by the endogenous rhythms of core body temperature and plasma melatonin with which sleep is normally synchronized. The subjects were placed in states of temporal disorder and the amplitude of the core body temperature rhythm in the elderly was reduced by 20 to 30% compared to younger subjects. The older subjects exhibited high levels of sleep fragmentation and shorter periods of sleeping, with the most fragmentation taking place when body temperature was on the rise, thus suggesting that they are more sensitive to waking signals from the central oscillator (Dijk et al., 1999). Following sleep deprivation, Dijk et al. (1999) observed homeostatic control systems in operation, however, deep slow-wave sleep on EEGs was markedly less in older subjects. These results indicate that the age-dependent decrease in sleep quality and earlier sleep onset and wake times are due to the hindered ability of
  • 40. 40 circadian mechanisms to promote sleep during the geophysical morning and the hindered ability of homeostatic mechanisms in enforcing sleep propensity (Dijk et al., 1999). The degeneration of the circadian timing system in the elderly was depicted in a study by Huang et al. (2002), in which sleep-wake cycles and phases of rest and activity were measured in routine settings. The study employed the use of wrist actigraphy over twenty-four hour periods for five to seven consecutive days in young subjects ranging from 21 to 34 years old, middle-aged subjects of 36 to 44 years old, old subjects of 61 to 79 years old and the oldest subjects of 80 to 91 years old. Those subjects showing extreme preferences for activity in the morning or evening were excluded from the investigation. In comparison to the young and middle-aged subjects, the old and oldest subjects exhibited decreased sleep time, decreased sleep efficiency, longer sleep latency, a higher number of nocturnal awakenings, a higher number of naps and the highest levels of sleep fragmentation (Table 2) (Huang et al., 2002). Actigraph readings revealed attenuated rhythms of rest and activity in the old and oldest subjects, with the lowest daytime activity and the highest levels of activity during the night. In addition, the data from these two subject groups are indicative of minimal coupling between sleep and environmental Zeitgebers (Huang et al., 2002). These results are suggestive of the impairment of sleep-wake cycles and rest-activity rhythms with aging. Table 2 Characteristics of the sleep-wake cycle are presented in the context of the four age groups under study. The old and oldest subjects spent the most time in bed, however,
  • 41. 41 with the lowest amounts of actual sleep time. The old and oldest subjects took the longest to fall asleep, had the lowest levels of sleep efficiency, the highest numbers of nocturnal awakenings, the highest number of naps and the highest sleep fragmentation indices (Huang et al., 2002). Recent research has given credence to the potential involvement of circadian clock gene alterations and their profound implications for rhythms of sleep and wakefulness in the aging population. Malatesta et al. (2007) analyzed CLOCK protein levels in the neurons of the medullary reticular formation, the brain centre that participates in the regulation of the sleep-wake cycle, in both young and old rats. Immunocytochemical techniques were applied at different phases of the light-dark circadian cycle. Low CLOCK levels were found in the old rats in the nerve cell compartments under scrutiny, including the cytoplasm, rough endoplasmic reticulum, nucleus, nucleolus and chromatin. Malatesta and colleagues (2007) speculate that these depressed levels of CLOCK protein in the neurons of the medullary reticular formation are associated with significantly disturbed sleep-wake cycles in the elderly. The Depletion of Chronobiological Rhythms and the Development of Psychiatric Disorders with Age The aging population is most susceptible to the depletion of chronobiological rhythms, and thus, the development of psychiatric disorders in the elderly merits attention. Previous sections have outlined the relationships between circadian rhythm abnormalities, perturbed sleep-wake cycles and aging. It may be hypothesized that senescence not only predisposes the elderly to chronodisruption and sleep deficits, but also increases their risk for developing frequently comorbid psychiatric illnesses (Fig. 8). The relationship between aging and depression via ocular dysfunction was evaluated by Jean-Louis and colleagues (2005). Recall ocular aging results in suboptimal
  • 42. 42 photoreception necessary for circadian photoentrainment. Study subjects’ ages averaged 68.3 years, with 27% being visually impaired according to ophthalmologic assessments. Low ambient light exposures corresponded with depressed mood states when controlling for demographic factors and medical complications (Jean-Louis et al., 2005). Ocular pathologies such as glaucoma, ocular hypertension and cataracts appear to intensify this relationship by negating light input to the master oscillator, thereby compromising sound mental health in the elderly (Jean-Louis et al., 2005). Figure 8 The implications of circadian rhythms in human mental health. A disruption in the sleep-wake cycle reflects impaired circadian clock functioning, which synergistically leads to the progression and maintenance of psychiatric disorders. The aging population is most susceptible to the depletion of chronobiological rhythms and sleep deficits, and thus, the development of psychiatric disorders in the elderly merits attention. Malformed circadian rhythms and disrupted sleep parameters were previously discussed in regards to Alzheimer’s disease, which is common in the elderly. A study by
  • 43. 43 Mishima et al. (1999) evaluated fluctuating levels of melatonin, which is thought to decline with age (Sack et al., 1986), and rest-activity rhythms in elderly patients with Alzheimer’s disease. Wrist actigraphy was used to assess circadian rest-activity rhythms and blood samples were assayed for plasma melatonin concentrations. The first study group consisted of Alzheimer’s patients with ages averaging 75.7 years and the second group consisted of dementia-free residents of the same nursing-home facility whose ages averaged 78.3 years, with the latter being free of disturbed sleep-wake cycles. The study was conducted in conditions of light below 150 lux and minimal physical exercise in order to prevent the suppression of melatonin. The Alzheimer’s patients showed considerably reduced amplitudes of melatonin secretion, with several patients displaying atypical peak secretion levels during the day, and less total daily secretions in comparison to the control group (Mishima et al., 1999). In addition, the rest-activity rhythms of the Alzheimer’s patients proved to be quite erratic. A similar study reported insufficient light exposure in these patients (Ancoli-Israel et al., 1997). This study establishes a positive correlation between dampened melatonin rhythms and disturbed sleep-wake patterns and rest-activity cycles, all of which are characteristic of elderly Alzheimer’s patients (Mishima et al., 1999; Volicer et al., 2001; Mahlberg et al., 2008). As previously described, there is a temporal misalignment of circadian rhythms, sleep-wake cycles and light exposures observed in schizophrenic patients. Martin et al. (2001) evaluated these factors in an aged population of schizophrenic patients, consisting of 14 men and 14 women whose ages averaged 58.3 years. An Actillume wrist monitor was used to measure both light exposure and activity levels. The dramatic results indicated that reduced light exposure was linked to weakened circadian rhythms and
  • 44. 44 sleep fragmentation, especially with age. The mean light exposure among the 28 subjects was less than 1000 lux, which worsened with age, and correlated with depressed mood and increased severity of psychiatric symptoms. These patients exhibited an excessive number of nocturnal awakenings, sometimes leading to insomnia for more than three hours per night. These sleep disturbances resulted in more daytime napping, and therefore, less daytime activity with substandard neuropsychological functioning and poor cognition. Actigraph recordings produced one-fifth of the robust amplitude measured in control participants. Collectively, these findings suggest possible roles for light deficiency, sleep disturbances, attenuated circadian rhythms, lifestyle and age status in this psychiatric disorder (Martin et al., 2001). However, the administration of anti- psychotic medications in many of the subjects may have confounded the results and further studies are required to delineate their contribution, if any, to these disturbed behavioural rhythms in comparison to those subjects who are not taking medication (Martin et al., 2001). Possible Treatments and Chronobiotics for Circadian Dysfunction With circadian rhythm disturbances as characteristic of various sleep disorders and psychiatric disorders, an assortment of chronobiological therapies has been proposed to alleviate their symptoms. The major internal Zeitgeber melatonin has been suggested as an effective pharmacological treatment for a range of circadian disruptions, including circadian rhythm sleep disorders, jet lag, shift-work maladaptation and free-running rhythms in the blind (Fischer et al., 2003). Recall that blind individuals are susceptible to sleep disorders and compromised neuropsychiatric conditions. Fischer et al. (2003) investigated whether a single one-time melatonin administration could temporarily
  • 45. 45 entrain blind individuals, thus synchronizing their sleep-wake cycles and melatonin rhythms and improving sleep conditions. Twelve men ages 18 to 40, incapable of photoentrainment, were given 5mg of melatonin one hour before bedtime, with both the subjects and administrators being unaware if the substance being given was melatonin or a placebo. In contrast to the placebo, melatonin increased total sleep time and sleep efficiency while decreasing the number of nocturnal awakening episodes. In regards to endocrine processes, adrenocorticotropic (ACTH) hormone and cortisol secretion are normally inhibited during the first half of sleep and rise in the latter half (Fischer et al., 2003). With this normally entrained rhythm being desynchronized in blind individuals, a single dose of melatonin improved sleep by realigning these hormonal rhythms. In light of its ability to synchronize circadian rhythms, improve sleep quality and regulate the hypothalamic-pituitary-adrenal axis, melatonin has the potential to serve as an anti-depressant (Fischer et al., 2003). This prospect is further supported in a study by Benedetti et al. (2001), in which similar results as Fischer and colleagues (2003) were obtained, however, with the additional finding that melatonin decreased the need for the use of benzodiazepines. In the context of this study, psychoactive benzodiazepine drugs were being taken by elderly subjects in order to treat insomnia (Benedetti et al., 2001). Melatonin administration before bedtime eliminated the use of benzodiazepines entirely in 65% of subjects, while reducing their use by 25 to 66% in 20% of subjects. Anti- depressants that improve sleep quality could therefore be crucial to treating depressive disorders. Recently, the new chronobiotic agomelatine has been put forward as a potential anti-depressant because of its coordinating effects on the circadian rest-activity rhythm
  • 46. 46 and its mitigating effects on symptoms of depression and anxiety (Kasper et al., 2010). Agomelatine acts as an agonist at melatonin receptors MT1 and MT2 and as an antagonist at serotonin receptors (Kasper et al., 2010). A study by Kasper and colleagues (2010) compared the effects of agomelatine with sertraline, a selective serotonin reuptake inhibitor (SSRI) known by its trade name as Zoloft, on patients with MDD. Agomelatine, in contrast to sertraline, increased the amplitude of rest-activity rhythms within one week. According to wrist actigraphs and sleep diaries, agomelatine improved sleep quality and the ease of falling asleep, and relieved feelings of depression and anxiety without any major adverse effects. Further research is required to establish agomelatine as an effective treatment for sleep disorders and affective disorders. Selective serotonin reuptake inhibitors are often employed in the treatment of affective disorders, including major depressive disorder, seasonal affective disorder and bipolar disorder. Sprouse et al. (2006) demonstrated the ability of fluoxetine, an SSRI, to alter firing activity of neurons in the SCN, and hence, circadian rhythmicity. At first, extracellular recordings of spontaneous neuron firing in the hypothalamic SCN slices of rats in vitro revealed no change in rhythm in response to fluoxetine. This was thought to be due to the loss of endogenous serotonin levels in culture conditions in vitro (Sprouse et al., 2006). When fluoxetine was paired with tryptophan, a serotonin precursor, microelectrode recordings revealed concentration-dependent phase advances in SCN rhythms (Sprouse et al., 2006). Further research is being conducted to ascertain the magnitude and direction of circadian phase shifts in clinical applications of fluoxetine (Sprouse et al., 2006).
  • 47. 47 As a mood stabilizer, lithium is used in the treatment of bipolar disorder as it counteracts both mania and depression (Hafen and Wollnik, 1994). Lithium lengthens the circadian period of the central pacemaker through direct pharmacological effects (Iwahana et al., 2004). The drug inhibits the action of a glycogen synthase kinase 3, an enzyme which targets the transcription factors of the molecular oscillator for degradation via phosphorylation, thereby slowing the molecular oscillator and relieving the advanced rhythm disturbances often seen in bipolar patients (Hafen and Wollnik, 1994; Iwahana et al., 2004). Most antidepressants, including lithium, need 2 to 8 weeks in order to exert their effects and induce a favourable response in patients (Wu et al., 2009). Unfortunately, there is a high risk for suicide in patients with bipolar disorder depression and a treatment regimen that evokes a rapid response and maintains this response is necessary (Wu et al., 2009). Despite its transient effects, sleep deprivation has been discovered as one of the most prompt and efficient chronotherapeutics, reducing depressive symptoms in 40 to 60% of patients within 24 to 48 hours (Wu et al., 2009). A study by Wu et al. (2009) evaluated the standard medications lithium and sertraline against a chronotherapeutic augmentation treatment (CAT), consisting of medications, sleep deprivation, bright light therapy and sleep phase advances. Forty-nine patients diagnosed with bipolar disorder according to the DSM were randomly assigned to either the medication group or the CAT group. CAT subjects were kept awake for 33 hours and then exposed to 5000 lux light for two hours for three consecutive days following sleep deprivation, as well as three days of gradual sleep phase advances. The CAT group exhibited a substantial relief of depressive symptoms as early as two days into treatment, and this effect lasted for seven
  • 48. 48 weeks, at which time 12/19 CAT subjects had gone into remission (Fig. 9). A combination of established chronotherapies appears to be most effective in alleviating the symptoms of this particular psychiatric disorder. Figure 9 In accordance with the Hamilton Rating Scale for Depression, CAT subjects, whose treatment consisted of medications, sleep deprivation, bright light therapy and sleep phase advances, displayed a considerable reduction in depressive symptoms in comparison to those solely on medications. This difference was seen as early as Day 2 and was maintained for 7 weeks (Wu et al., 2009). Bright light therapy in the morning has long been known as an effective treatment for both MDD and SAD patients. A study by Lewy et al. (1998) demonstrated the superior efficacy of morning versus evening bright light therapy. This was based on the prediction that SAD patients would have phase delayed rhythms during winter depression (Lewy et al., 1998). Forty-five patients with MDD or bipolar disorder with a winter seasonal pattern and 49 controls participated in the study for six weeks. The participants were treated with bright light therapy in their homes for two weeks either in the morning, between 6 to 8 AM, or in the evening, between 7 to 9 PM, then subjected to a week of light withdrawal, and then treated with light in the opposing time period. Blood samples were taken once a week in dim light and assayed for melatonin in order to determine
  • 49. 49 circadian phase positions relative to nocturnal melatonin onset. Morning bright light therapy, which phase advanced patients, proved more effective than that in the evening, which phase delayed patients, with the former inducing a 27% decrease in depressive symptoms and intensifying those symptoms during the withdrawal period (Lewy et al., 1998). The authors of this study suggest the application of bright light therapy without delay upon awakening for the best results in patients with SAD (Lewy et al., 1998). The phase advancing and phase delaying effects of bright light therapy may be applied in treatment regimens for the circadian rhythm sleep disorders DSPS and ASPS as well. With light as the dominant Zeitgeber in the human circadian timekeeping system, it is no surprise that light exposure is one of the most potent treatments for circadian rhythm disorders. Recall that photosensitive RGCs best absorb light in the blue sector of the light spectrum at 460 nm, which is quite similar to the wavelength of environmental light (Turner and Mainster, 2008). A study by Glickman et al. (2006) investigated the optimal spectral wavelength for phototherapy in 24 SAD patients. Blue light emitting diode boxes gave off 468 nm light to 11 subjects, while the red light emitting diode boxes gave off 654 nm to 13 subjects. Light therapy was administered everyday for three weeks for durations of 45 minutes between 6 to 8 AM. According to the Hamilton Depression Rating Scale, subjects who had been given short blue wavelength light treatments scored 7.3 points lower on depressive symptoms than those who had been exposed to longer red wavelength light, thus confirming blue light as optimal for light therapy (Glickman et al., 2006). As previously outlined in bright morning light therapy, the timing of light exposure is critical in light therapy. A recent study by Goel et al. (2006) applied 10 000 lux light for one hour immediately upon awakening to patients
  • 50. 50 with chronic MDD diagnoses for five weeks. The results indicated substantial improvement in symptoms, with depression scores improving by 53.7 % and with remission rates of 50%. Light therapy is predicted to be most effective in conjunction with the abovementioned treatment strategies of melatonin and sleep deprivation (Goel et al., 2006). The widespread success of light therapy is derived from its ability to phase shift the endogenous circadian pacemaker, and so it is used to alleviate the detrimental effects of jet lag, shift-work, circadian rhythm sleep disorders, Alzheimer’s disease and bipolar disorder, to name a few. Optimal photoreception may be achieved in the aging population with ample natural light exposure, IOLs and bright, appropriately timed residential lighting (Turner and Mainster, 2008). Structural designs that allow for bright environmental light exposure during the geophysical day and limited light exposure in the evening would allow for most advantageous entrainment and internal synchrony (Oren et al., 1997; Turner and Mainster, 2008). Increasing public awareness of these strategies for harmonious synchronization and optimal well-being are not only profitable to the elderly in preventing circadian malfunction but to all age groups. Research Proposal Alleviating Sleep Disorders to Alleviate Psychiatric Disturbances Rationale Circadian dysfunction, notably decreased circulating melatonin levels and disturbed rest-activity rhythms, and the deterioration of the sleep-wake cycle are characteristic of the aging population (Campbell et al., 1988; Zhou et al., 2003; Nygard et al., 2005; Turner and Mainster, 2008). In light of those previously mentioned studies that
  • 51. 51 have enhanced sleeping conditions and alleviated depressive symptoms through the use of melatonin and light therapy, future studies warrant investigating whether correcting circadian misalignment with these circadian resetting agents will attenuate sleep disturbances and psychiatric pathologies in the elderly (Lewy et al., 1998; Fischer et al., 2003; Goel et al., 2006). The proposed study will evaluate the efficacy of a treatment combining timed bright light exposure and exogenous melatonin, two primary Zeitgebers, in consolidating circadian rhythms and alleviating sleep disturbances and psychiatric symptoms in elderly patients with Alzheimer’s disease. Participants Participants will be gathered from a geriatric facility and will be comprised of Alzheimer’s patients ranging from the ages of 60 to 90 years old. Written consent forms and approval from an institutional ethics board will be attained prior to commencing the study. Preliminary Assessment Prior to commencing treatment, initial assessments of sleep-wake cycles and rest-activity rhythms will be made for a duration of one week. This information will be gathered through nurse and staff ratings, subject interviews and with the use of Actillume wrist monitors (Martin et al., 2001). Actillume recordings will also indicate the amount of light exposure obtained by the participants. Saliva samples will be taken every 30 minutes during the daytime in dim light conditions within a 24 hour period and assayed for melatonin concentrations in order to assess the level of circadian misalignment (Zhou et al., 2003). In addition, the total amount of nocturnal melatonin production will be
  • 52. 52 measured by performing assays for its major urinary metabolite 6-hydroxymelatonin (Sack et al., 1986). Cognitive tests and psychiatric assessments will be applied to assess the severity of psychiatric symptoms. Methods The study will be conducted in a double-blind manner using placebos. Participants will randomly be assigned to one of two groups. The first group will receive bright light exposure of 3000 lux daily between the hours of 6AM and 9AM and 3 mg of melatonin one hour before bedtime for one month (Lewy et al., 1998; Fischer et al., 2003; Turner and Mainster, 2008). The use of blue light emitting diodes would be best (Glickman et al., 2006). The second group will receive light exposure of 100 lux daily between the hours of 6AM and 9AM and a placebo pill one hour before bedtime for one month. The use of red light emitting diodes would be best (Glickman et al., 2006). The study will be conducted in a double-blind manner since the nurses and staff, as well as the participants themselves, will be unaware as to whether they are administering a treatment regimen or placebo. A final assessment will be made after one month of treatment by applying the same procedures as outlined in the preliminary assessment. The healthcare professionals performing the final cognitive tests and psychiatric assessments will be uninformed as to which group participants were assigned. Controls The only difference between the two groups will be the treatment regimen administered as they will otherwise be living in the same geriatric facility, be age-matched and have a diagnosis of dementia. The placebo pills will serve as controls for exogenous melatonin.
  • 53. 53 The use of longer wavelength light will serve as a control as it has been established that light of this intensity and of the red spectrum are insufficient for the body’s circadian demands (Glickman et al., 2006; Turner and Mainster, 2008). Predicted Outcomes The combined treatment of bright morning light therapy and exogenous melatonin will most likely result in improved central pacemaker functioning, which will be evident with decreased activity during the night and increased activity during the day. Treatment will lead to an increase in sleep efficiency, sleep duration, and deep sleep, with a decrease in daytime napping, nocturnal awakenings and agitation. Nocturnal melatonin levels will increase, not only due to the application of exogenous melatonin, but also due to improved sleeping parameters and light exposure. Cognition and mood will be enhanced as well. Melatonin rhythms and light exposure are dampened in Alzheimer’s patients, therefore, this treatment has the potential to restore the compromised rest-activity cycles, sleep-wake patterns and neuropsychiatric functioning seen in these elderly patients (Ancoli-Israel et al., 1997; Mishima et al., 1999). Limitations The placebo effect, medications and the severity of dementia-related symptoms may prove to be limitations in this study. This study may be expanded to test for the placebo effect by including an additional age-matched control group living in the same geriatric facility with diagnoses of dementia, however, this group would receive neither the treatment regimen nor placebos. This control group would account for other variables that may lead to improvements in participants, for example, the context of the facility in