ArticlePDF Available

Dehydration and Cognition in Geriatrics: A Hydromolecular Hypothesis

Authors:
  • Patton State Hospital

Abstract and Figures

Dehydration is one of the ten most frequent diagnoses responsible for the hospital admission of elderly in the United States. It is associated with increased mortality, morbidity and an estimated cost of 1.14 billion per year (1-4). Older individuals are predisposed to dehydration encephalopathy as a result of decreased total body water (TBW) and diminished sensation of thirst. We hypothesize that thirst blunting in older individuals is the result of a defective microRNA-6842-3p failing to silence the expression of the vesicular GABA transporters (VGAT) and alpha 7 cholinergic nicotinic receptors in the subfornical organ (SFO) of the hypothalamus. We hypothesize further that resultant dehydration facilitates protein misfolding and aggregation, predisposing to neurocognitive disorders. We completed a search of predicted microRNA targets, utilizing the public domain tool miRDB and found that microRNA-6842-3p modulates the SLC6A1 and CHRNA7 genes both of which were previously hypothesized to inhibit the thirst sensation by their action on SFO. The primary aim of this article is to answer two questions: Can prevention and correction of dehydration in elderly lower age-related cognitive deterioration? Can exosomal miR-6842 in the peripheral blood predict dehydration encephalopathy in elderly?
Content may be subject to copyright.
HYPOTHESIS AND THEORY
published: 12 May 2016
doi: 10.3389/fmolb.2016.00018
Frontiers in Molecular Biosciences | www.frontiersin.org 1May 2016 | Volume 3 | Article 18
Edited by:
Megha Agrawal,
University of Illinois at Chicago, USA
Reviewed by:
Ravi C. Kalathur,
New York Structural Biology Center,
USA
Enno Klussmann,
Max Delbrück Center for Molecular
Medicine, Germany
*Correspondence:
Adonis Sfera
dr.sfera@gmail.com;
Michael Cummings
Michael.Cummings@DSH.ca.gov
Specialty section:
This article was submitted to
Molecular Diagnostics,
a section of the journal
Frontiers in Molecular Biosciences
Received: 28 January 2016
Accepted: 25 April 2016
Published: 12 May 2016
Citation:
Sfera A, Cummings M and Osorio C
(2016) Dehydration and Cognition in
Geriatrics: A Hydromolecular
Hypothesis. Front. Mol. Biosci. 3:18.
doi: 10.3389/fmolb.2016.00018
Dehydration and Cognition in
Geriatrics: A Hydromolecular
Hypothesis
Adonis Sfera 1, 2*, Michael Cummings 2*and Carolina Osorio 1
1Department of Psychiatry, Loma Linda University, Loma Linda, USA, 2Patton State Hospital, Patton, USA
Dehydration is one of the ten most frequent diagnoses responsible for the hospital
admission of elderly in the United States. It is associated with increased mortality,
morbidity and an estimated cost of 1.14 billion per year (Xiao et al., 2004; Schlanger et al.,
2010; Pretorius et al., 2013; Frangeskou et al., 2015). Older individuals are predisposed
to dehydration encephalopathy as a result of decreased total body water (TBW) and
diminished sensation of thirst. We hypothesize that thirst blunting in older individuals is the
result of a defective microRNA-6842-3p failing to silence the expression of the vesicular
GABA transporters (VGAT) and alpha 7 cholinergic nicotinic receptors in the subfornical
organ (SFO) of the hypothalamus. We hypothesize further that resultant dehydration
facilitates protein misfolding and aggregation, predisposing to neurocognitive disorders.
We completed a search of predicted microRNA targets, utilizing the public domain tool
miRDB and found that microRNA-6842-3p modulates the SLC6A1 and CHRNA7 genes
both of which were previously hypothesized to inhibit the thirst sensation by their action
on SFO. The primary aim of this article is to answer two questions: Can prevention
and correction of dehydration in elderly lower age-related cognitive deterioration? Can
exosomal miR-6842 in the peripheral blood predict dehydration encephalopathy in
elderly?
Keywords: dehydration, aquaporins, extracellular space, protein folding, protein conformational dynamics
HYDRATION AND COGNITION
Dehydration is one of the most common medical problems in seniors diagnosed in 6.7% of
hospitalized patients over the age of 65 (Warren et al., 1994). It leads to poor outcomes and
increased health care expenditures. Novel studies reveal that if not prevented or treated promptly,
dehydration results in longer intensive care unit (ICU) stay, higher hospital readmission rates and
placement in long term facilities (Xiao et al., 2004; Frangeskou et al., 2015). On the other hand,
preventing dehydration not only reduces healthcare expenditures, but also improves outcomes and
the elderly patients’ quality of life.
Dehydration is a contributing factor for delirium, a neurobehavioral syndrome recently
demonstrated to be a strong risk factor for dementia (Inouye, 1998; Davis et al., 2012). It is therefore
crucial to recognize and diagnose dehydration quickly, however at the present time there are no
specific biological markers for this condition. Clinical signs, plasma osmolality and urine markers
have poor specificity in elderly (George and Rockwood, 2004). For this reason potential epigenetic
markers such as microRNA-6842-3p obtained from peripheral blood exosomes may contribute not
only to early diagnosis, but also to prevention of dehydration.
Sfera et al. Biomarkers of Dehydration in Geriatrics
Water is an essential body nutrient and its homeostasis is
crucial for life. Early in the evolution, marine animals were
surrounded by water, but survival on dry land required built-in,
“portable” water (Warren et al., 1994). In humans, the muscle
tissue is a genuine fluid reservoir, carrying over 80% of TBW
(George and Rockwood, 2004).
The brain, in spite of being a highly lipophilic organ consists
of 80% water (Tait et al., 2008). Most of the CNS intracellular
water is stored in astrocytes. These cells are characterized by high
aquaporin (AQP) expression which makes them four times more
permeable to water than other brain cells, therefore true “brain
cisterns” for times of water scarcity (Thrane et al., 2014). With
the same token, because of their high AQP content, astrocytes are
prone to pathological water retention and swelling. Novel studies
demonstrate that astrocytes respond to peripheral dehydration
by up-regulation of AQP-4 proteins on their end-feet processes
probably in order to preserve water. For example, preclinical
studies demonstrate that a hyperosmotic milieu induces AQP
expression in astrocytes (Yang et al., 2013).
Overexpression of AQP-4 channels and augmented water
intake transforms these cells into genuine “sponges” resulting in
extracellular dehydration, extracellular space (ECS) hypovolemia.
If severe enough this condition may turn into a medical
emergency, dehydration encephalopathy or delirium.
The process of aging seems to undo the evolutionary
advantage of “portable water” as elderly individuals are known
to lose their fluid reservoirs by age-related decrease in both
muscle mass and astrocyte density. For example, dehydration was
demonstrated to accelerate the progression of AD which is also
known to be associated with loss of astrocytes (Ogawa et al., 2011;
Reyes-Haro et al., 2015; Rodríguez-Arellano et al., 2016).
It is well known that aging is associated with reduced
acetylcholine (ACh) in the brain, but it is perhaps less emphasized
that aging contributes to down-regulation of alpha7 nicotinic
acetylcholine receptors (alpha7nAChR) (Utsugisawa et al., 1999;
Akhmedov et al., 2013), rendering the CNS less responsive to
ACh. This is significant for the sensation of thirst which is
physiologically activated by ACh. Lower cholinergic activation
predisposes to inflammation which is also involved in cognitive
impairment. We discussed inflammation in the aging brain
elsewhere and this subject will not be brought here (Sfera and
Osorio, 2014). The alpha7nAChR are encoded by CHRNA7 gene
which is subject to microRNA epigenetic regulation, including
miR-6842.
Concerning the relationship between dehydration and
impaired cognition nutrition studies demonstrate that a
loss of only 1–2% of TBW may result in impaired cognitive
performance; in elderly this percentage was shown to be
even lower (Han and Wilber, 2013; Riebl and Davy, 2013).
Furthermore, the link between hydration and cognition can
be demonstrated by the neurocognitive disorders associated
with up-regulation of AQP-4 expression primarily on asyrocytic
end-feet (Table 1).
Novel studies demonstrate that both dehydration and
aging were associated with AQP-4 up-regulation, therefore it
should not come as a surprise that aging and water loss go
hand in hand (Trinh-Trang-Tan et al., 2003). Interestingly,
several amyloid-binding, neuroprotective compounds were
demonstrated to down-regulate AQP-4 expression, further
demonstrating the role of water in amyloid pathology (Table 2).
In addition, neuroimaging studies in dehydrated elderly, show
decrease in gray and white matter volume (Streitbürger et al.,
2012). However, it is important to keep in mind that most brain
volumetric studies rely on diffusion tensor imaging (DTI) which
detects water anisotropy and is therefore highly dependent on the
brain fluid dynamics (Meng et al., 2004; Nakamura et al., 2014).
WATER AND PROTEIN MISFOLDING
DISORDERS
Misfolded protein aggregates were shown to be involved in
many human diseases, including neurocognitive disorders and
diabetes type 2, but in spite of the increasing prevalence of
these conditions, the reason proteins misfold is not completely
understood.
Water has been known to play a major role in protein
conformational dynamics (Lemieux, 1996; Phillips, 2002; Zhao
et al., 2013). In order to become biologically active newly
transcribed proteins must fold along specific axes like paper
in the ancient Japanese art of origami (Collet, 2011; Chong
TABLE 1 | Disorders associated with cognitive deficit and AQP-4
up-regulation.
AQP-4/Cognitive deficit disorders References
Cerebral amyloid angiopathy Foglio and Fabrizio, 2010;
Moftakhar et al., 2010
Alzheimer’s disease Nagelhus and Ottersen, 2013;
Lan et al., 2015
Parkinson’s disease Subburaman and Vanisree, 2011;
Zhang et al., 2016
Multiple sclerosis Tanaka et al., 2007
Neuromyelitis optica Saji et al., 2013; Zhang et al.,
2015
Traumatic brain injury Hu et al., 2005
Cerebral ischemia Zador et al., 2009
Epilepsy Binder et al., 2012; Alvestad
et al., 2013
HIV encephalitis St. Hillaire et al., 2005
Progressive multifocal leukoencephalopathy Aoki-Yoshino et al., 2005;
Florence et al., 2012
TABLE 2 | Neuroprotective compounds associated with AQP-4
down-regulation.
AQP-4 down-regulation References
Rapamycin Guo et al., 2014
Erythropoietin Gunnarson et al., 2009; McCook et al., 2012
Curcumin Laird et al., 2010; Wang et al., 2015
Purines Morelli et al., 2010; Lee et al., 2013
Progesteron He et al., 2014
Melatonin Dehghan et al., 2013; Lin et al., 2013;
Bhattacharyaa et al., 2014
Frontiers in Molecular Biosciences | www.frontiersin.org 2May 2016 | Volume 3 | Article 18
Sfera et al. Biomarkers of Dehydration in Geriatrics
and Ham, 2015). Recently it was demonstrated that water
plays a crucial role in this process as it forms hydrogen
bonds with the amino acid chains, facilitating their collapse
into three dimensional molecular structures. In the presence
of water, folding occurs almost instantly (140 ns), resulting in
biologically active molecules available for chemical reactions
at the opportune time (Sen and Voorheis, 2014; Vajda and
Perczel, 2014). In the absence of hydration the folding process
is significantly slower and the biomolecules may miss the timing
of their reactions. This results in molecular overcrowding which
predisposes to misfolding (Gregersen et al., 2006; Stoppini
et al., 2009). Indeed, it was hypothesized by others that
biomolecular crowding relative to the fluid volume is inductive of
misfolding and aggregation (Tokuriki et al., 2004; Yerbury et al.,
2005).
Novel studies in protein conformational dynamics
demonstrate that both protein misfolding, their repair and
removal can take place in the intra and the extracellular
compartment. The chance of protein misfolding is higher in
the extracellular space (ECS) which is a rougher environment
exposing these biomolecules to a higher degree of shear and tear
(Ker and Chen, 1998; Genereux and Wiseman, 2015). For this
reason, we focus our study on the ECS where hypovolemia may
facilitate protein misfolding and aggregation.
It was hypothesized that adequate water circulation via
aquaporin (AQP) channels is essential for clearing beta amyloid
and for preventing its build-up characteristic for Alzheimer’s
disease (AD) (Figure 1). The glymphatic system paradigm
suggests that insufficient amyloid clearance and its subsequent
aggregation is the result of impaired water movement (Xie et al.,
2013). This model, however pays less attention as to why proteins
misfold in the first place.
The hydromolecular hypothesis is therefore complementary
to the glymphatic model, but also differs from it by elevating
water from an inert medium to an active participant in cognition
(via protein folding) (Levy and Onuchic, 2006). This hypothesis
raises another interesting question: do proteins participate in
information processing directly?
Novel studies in neuroscience demonstrate that proteins
participate in cognition by their ability to access logic
gates, the elementary building blocks of digital circuits (Qi
et al., 2013). These molecules are endowed with abilities to
adaptively change their shapes in Transformers-like fashion,
assembling and disassembling in response to electronic signals or
electromagnetic fields (Kidd et al., 2009; Ausländer et al., 2012).
For example, proteins were shown to assemble in the neuronal
post-synaptic membrane into heteroreceptor complexes which
may engender memory “bar codes” (Fuxe et al., 2007; Chen et al.,
2012). Calcium-calmodulin-dependent kinase III, a component
of neuronal microtubules, was hypothesized to store long term
memory by reorganizing its spatial structure in response to
synaptic activity (Smythies, 2015). Interestingly, water plays a
FIGURE 1 | Astrocyte swelling as a result of AQP-4 channels up-regulation with interstitial fluid hypovolemia and beta-amyloid misfolding.
Frontiers in Molecular Biosciences | www.frontiersin.org 3May 2016 | Volume 3 | Article 18
Sfera et al. Biomarkers of Dehydration in Geriatrics
major role in this model. Several studies revealed that dendritic
spine biomolecules may play a crucial role in associative memory
as they endow the neural circuits with Boolean logic (Craddock
et al., 2012; De Ronde et al., 2012; Qi et al., 2013). Furthermore,
proteins are endowed with Lego-like abilities to interlink,
engendering large intra and extracellular biomolecular networks
with hypothesized roles in cognition (Chen et al., 2012; Mancuso
et al., 2014). In light of this data we believe that alteration of
the normal protein conformation may impair cognition directly,
rather than indirectly by damaging synapses and neurons which
is the traditional view.
EPIGENOMIC REGULATION OF THE
SUBFORNICAL ORGAN (SFO)
Elderly individuals are prone to dehydration as a result of blunted
thirst sensation and loss of TBW as discussed above (Cowen
et al., 2013; Hooper et al., 2014). Recent preclinical data reveal
that the subfornical organ (SFO) of the hypothalamus functions
as a “thirst center” in the mammalian brain, regulating the
basic instinct of water intake (Oka et al., 2015). Since the SFO
lacks a blood-brain-barrier (BBB) it may be well positioned to
detect peripheral dehydration and respond to it by increasing
the sensation of thirst lowering water output. The SFO contains
sensitive osmoreceptors which convert peripheral changes in
osmolality into an excitatory neuronal signal, triggering both the
sensation of thirst and the release of arginine vasopressin (AVP)
by the posterior pituitary (Azizi et al., 2008).
FIGURE 2 | Water channels expressed on the SFO cells. Inhibitory
GABAergic neurons express VGAT, astrocytes AQP-9 and tanycytes AQP-4.
It was recently demonstrated that the SFO contains both
excitatory and inhibitory neurons which can be activated by
the ECS water volume and osmolality (Oka et al., 2015).
ECS hypovolemia activates the SFO excitatory neurons (which
express ETV-1 transcription factor), triggering thirst. ECS
normovolemia, on the other hand activates the SFO inhibitory
neurons (which express the vesicular GABA transporter
(VGAT)], inhibiting the sensation of thirst.
These genetically distinct neuronal groups may explain both
dehydration and psychotic polydipsia. For example, excessive
activation of excitatory, or failure to activate inhibitory SFO
neurons may result in psychotic polydipsia. The opposite may be
true in dehydration.
Several prior studies revealed that the sensation of thirst
may also be activated by the stimulation of SFO neuronal
cholinergic receptors. The SFO neurons express both nicotinic
and muscarinic cholinergic receptors, while the SFO astrocytes
express only alpha 7- nAChRs (Honda et al., 2003; Tanaka,
2003; Ono et al., 2008). Age-related paucity of these receptors
interferes with ACh activation of the thirst sensation. The glial
water channels consist of AQP-9 expressed by astrocytes and
AQP-4 expressed by tanycytes (Figure 2).
In addition to decreasing the expression of alpha 7 nAChRs,
the aging process was documented to augment the expression
of AQP channels on astrocytic end-feet as part of an
age-related senescence-associated secretory phenotype (SASP).
SASP is characterized by low grade inflammation, increased
accumulation of misfolded protein aggregates and astrocyte
swelling induced by AQP up-regulation (Picciotto and Zoli, 2002;
Salminen et al., 2011; Akhmedov et al., 2013).
FIGURE 3 | Physiologically, microRNA-6842 silences SLC6A1 and
CHRNA7 genes, activating the sensation of thirst.
Frontiers in Molecular Biosciences | www.frontiersin.org 4May 2016 | Volume 3 | Article 18
Sfera et al. Biomarkers of Dehydration in Geriatrics
Peripheral dehydration was demonstrated to alter the
expression of several SFO- related genes (Hindmarch et al.,
2008). One of these genes is SLC6A1 which expresses
VGAT on the cellular membranes of the SFO inhibitory
neurons.
Method: we conducted a search of miRDB, a public
online database for microRNA target prediction and functional
annotations. The targets in miRDB are predicted by the
bioinformatics tool, MirTarget. MirTarget was developed by
analyzing thousands of miRNA-target interactions from high-
throughput sequencing experiments. We searched the human
database for the genes of interest SLC 32A1 and CHRNA 7,
coding for VGAT and alpha 7 nicotinic cholinergic receptors
respectively. We conducted a separate search for each of the two
genes by utilizing the gene symbol SLC 32A1 and CHRNA 7.
The results revealed that 131 microRNAs modulate the SLC 32A1
gene and 57 microRNAs the CHRNA 7 gene. Analyzing this data,
miR by miR we found one common microRNA modulating both
genes, the miR-6842 (Figure 3).
A dysfunctional miR-6842 may fail to silence the SLC6A1
gene, preventing inhibition of the SFO GABAergic neurons with
resultant thirst blocking. The same is achieved via failure to
inhibit the CHRNA-7 gene, thus preventing ACh-induced thirst.
CONCLUSIONS
The hydromolecular hypothesis endeavors to explain the
relationship between dehydration and decreased cognition in
elderly as resulting from protein misfolding and aggregation in
the context of low interstitial fluid volume (ECS hypovolemia).
Defective proteins may affect cognition either directly via
impaired information processing in the brain biomolecular
networks, or indirectly via neuronal and synaptic damage, or
both.
MicroRNA-6842 may constitute a biological marker with
predictive value for dehydration encephalopathy in elderly as it
regulates two genes involved in the sensation of thirst.
AUTHOR CONTRIBUTIONS
All authors listed, have made substantial, direct and intellectual
contribution to the work, and approved it for publication.
REFERENCES
Akhmedov, K., Rizzo, V., Kadakkuzha, B. M., Carter, C. J., Magoski, N. S., Capo,
T. R., et al. (2013). Decreased response to acetylcholine during aging of Aplysia
neuron R15. PLoS ONE 8:e84793. doi: 10.1371/journal.pone.0084793
Alvestad, S., Hammer, J., Hoddevik, E. H., Skare, Ø., Sonnewald, U., Amiry-
Moghaddam, M., et al. (2013). Mislocalization of AQP4 precedes chronic
seizures in the kainate model of temporal lobe epilepsy. Epilepsy Res. 105,
30–41. doi: 10.1016/j.eplepsyres.2013.01.006
Aoki-Yoshino, K., Uchihara, T., Duyckaerts, C., Nakamura, A., Hauw, J. J.,
and Wakayama, Y. (2005). Enhanced expression of aquaporin 4 in human
brain with inflammatory diseases. Acta Neuropathol. 110, 281–288. doi:
10.1007/s00401-005-1052-2
Ausländer, S., Ausländer, D., Müller, M., Wieland, M., and Fussenegger, M. (2012).
Programmable single-cell mammalian biocomputers. Nature 487, 123–127. doi:
10.1038/nature11149
Azizi, M., Iturrioz, X., Blanchard, A., Peyrard, S., De Mota, N., Chartrel, N., et al.
(2008). Reciprocal regulation of plasma apelin and vasopressin by osmotic
stimuli. J. Am. Soc. Nephrol. 19, 1015–1024. doi: 10.1681/ASN.2007070816
Bhattacharyaa, P., Kumar, A., Paulb, P. S., and Patnaikb, R. (2014). Melatonin
renders neuroprotection by protein kinase C mediated aquaporin-4 inhibition
in animal model of focal cerebral ischemia. Life Sci. 100, 97–109. doi:
10.1016/j.lfs.2014.01.085
Binder, D. K., Nagelhus, E. A., and Ottersen, O.P. (2012). Aquaporin-4 and
epilepsy. Glia 60, 1203–1214. doi: 10.1002/glia.22317
Chen, Y.-S., Hon, M.-Y., and Huang, G. S. (2012). A protein transistor made of an
antibody molecule and two gold nanoparticles. Nat. Nanotechnol. 7, 197–203.
doi: 10.1038/nnano.2012.7
Chong, S. H., and Ham, S. (2015). Distinct role of hydration water in protein
misfolding and aggregation revealed by fluctuating thermodynamics analysis.
Acc. Chem. Res. 48, 956–965. doi: 10.1021/acs.accounts.5b00032
Collet, O. (2011). How does the first water shell fold proteins so fast? J. Chem. Phys.
134, 085107. doi: 10.1063/1.3554731
Cowen, L. E., Hodak, S. P., and Verbalis, J. G. (2013). Age-associated abnormalities
of water homeostasis. Endocrinol. Metab. Clin. North Am. 42, 349–370. doi:
10.1016/j.ecl.2013.02.005
Craddock, T. J. A., Tuszynski, J. A., and Hameroff, S. (2012). Cytoskeletal signaling:
is memory encoded in microtubule lattices by CaMKII phosphorylation? PLoS
Comput. Biol. 8:e1002421. doi: 10.1371/journal.pcbi.1002421
Davis, D. H. J., Terrera, G. M., Tuomo, P., Sulkava, R., MacLullich, A. M.
J., and Carol, B. (2012). Delirium is a strong risk factor for dementia in
the oldest-old: a population-based cohort study. Brain 135, 2809–2816. doi:
10.1093/brain/aws190
Dehghan, F., Khaksari Hadad, M., Asadikram, G., Najafipour, H., and Shahrokhi,
N. (2013). Effect of melatonin on intracranial pressure and brain edema
following traumatic brain injury: role of oxidative stresses. Arch. Med. Res. 44,
251–258. doi: 10.1016/j.arcmed.2013.04.002
De Ronde, W., Rein ten Wolde, P., and Mugler, A. (2012). Protein logic: a statistical
mechanical study of signal integration at the single-molecule level. Biophys. J.
103, 1097–1107. doi: 10.1016/j.bpj.2012.07.040
Florence, C. M., Baillie, L. D., and Mulligan, S. J. (2012). Dynamic volume changes
in astrocytes are an intrinsic phenomenon mediated by bicarbonate ion flux.
PLoS ONE 7:e51124. doi: 10.1371/journal.pone.0051124
Foglio, E., and Fabrizio, R. E. (2010). Aquaporins and neurodegenerative diseases.
Curr. Neuropharmacol. 8, 112–121. doi: 10.2174/157015910791233150
Frangeskou, M., Lopez-Valcarcel, B., and Serra-Majem, L. (2015). Dehydration in
the elderly: a review focused on economic burden. J. Nutr. Health Aging. 19,
619–627. doi: 10.1007/s12603-015-0491-2
Fuxe, K., Canals, M., Torvinen, M., Marcellino, D., Terasmaa, A., Genedani, S.,
et al. (2007). Intramembrane receptor-receptor interactions: a novel principle
in molecular medicine. J. Neural Transm. 114, 49–75. doi: 10.1007/s00702-006-
0589-0
Genereux, J. C., and Wiseman, R. L. (2015). Regulating extracellular proteostasis
capacity through the unfolded protein response. Prion 9, 10–21. doi:
10.1080/19336896.2015.1011887
George, J., and Rockwood, K. (2004). Dehydration and delirium–not a
simple relationship. J. Gerontol. A Biol. Sci. Med. Sci. 59, 811–812. doi:
10.1093/gerona/59.8.M811
Gregersen, N., Bross, P., Vang, S., and Christensen, J. H. (2006). Protein misfolding
and human disease. Annu. Rev. Genomics Hum. Genet. 7, 103–124. doi:
10.1146/annurev.genom.7.080505.115737
Gunnarson, E., Song, Y., Kowalewski, J. M., Brismar, H., Brines, M., Cerami, A.,
et al. (2009). Erythropoietin modulation of astrocyte water permeability as a
component of neuroprotection. Proc. Natl. Acad. Sci. U.S.A. 106, 1602–1607.
doi: 10.1073/pnas.0812708106
Guo, W., Feng, G., Miao, Y., Liu, G., and Xu, C. (2014). Rapamycin alleviates brain
edema after focal cerebral ischemia reperfusion in rats. Immunopharmacol.
Immunotoxicol. 36, 211–223. doi: 10.3109/08923973.2014.913616
Frontiers in Molecular Biosciences | www.frontiersin.org 5May 2016 | Volume 3 | Article 18
Sfera et al. Biomarkers of Dehydration in Geriatrics
Han, J. H., and Wilber, S. T. (2013). Altered mental status in older
emergency department patients. Clin. Geriatr. Med. 29, 101–136. doi:
10.1016/j.cger.2012.09.005
He, L., Zhang, X., Wei, X., and Li, Y. (2014). Progesterone attenuates aquaporin-4
expression in an astrocyte model of ischemia/reperfusion. Neurochem. Res. 39,
2251–2261. doi: 10.1007/s11064-014-1427-7
Hindmarch, C., Fry, M., Yao, S. T., Smith, P. M., Murphy, D., and Ferguson, A.
V. (2008). Microarray analysis of the transcriptome of the subfornical organ in
the rat: regulation by fluid and food deprivation. Am. J. Physiol. Regul. Integr.
Comp. Physiol. 295, R1914–R1920. doi: 10.1152/ajpregu.90560.2008
Honda, E., Ono, K., Toyono, T., Kawano, H., Masuko, S., and Inenaga, K.
(2003). Activation of muscarinic receptors in rat subfornical organ neurones.
J. Neuroendocrinol. 15, 770–777. doi: 10.1046/j.1365-2826.2003.01057.x
Hooper, L., Bunn, D., Jimoh, F. O., and Fairweather-Tait, S. J. (2014).
Water-loss dehydration and aging. Mech. Ageing Dev. 136–137, 50–58. doi:
10.1016/j.mad.2013.11.009
Hu, H., Yao, H. T., Zhang, W. P., Zhang, L., Ding, W., Zhang, S. H., et al. (2005).
Increased expression of aquaporin-4 in human traumatic brain injury and brain
tumors. J. Zhejiang Univ. Sci. B. 6, 33–37. doi: 10.1631/jzus.2005.B0033
Inouye, S. K. (1998). Delirium in hospitalized older patients. Clin. Geriatr. Med.
14, 754–764.
Ker, Y. C., and Chen, R. H. (1998). Shear-induced conformational changes and
gelation of soy protein isolate suspensions. Food Sci. Technol. 31, 107–113. doi:
10.1006/fstl.1997.0306
Kidd, B. A., Baker, D., and Thomas, W. E. (2009). Computation of conformational
coupling in allosteric proteins. PLoS Comput. Biol. 5:e1000484. doi: 10.1371/
journal.pcbi.1000484
Laird, M. D., Sukumari-Ramesh, S., Swift, A. E., Meiler, S. E., Vender, J. R., and
Dhandapani, K. M. (2010). Curcumin attenuates cerebral edema following
traumatic brain injury in mice: a possible role for aquaporin-4? J. Neurochem.
113, 637–648. doi: 10.1111/j.1471-4159.2010.06630.x
Lan, Y. L., Zhao, J., Ma, T., and Li, S. (2015). The potential roles of aquaporin 4
in alzheimer’s disease. Mol. Neurobiol. doi: 10.1007/s12035-015-9446-1. [Epub
ahead of print].
Lee, M. R., Ruby, C. L., Hinton, D. J., Choi, S., Adams, C. A., Young Kang,
N., et al. (2013). Striatal adenosine signaling regulates EAAT2 and astrocytic
AQP4 expression and alcohol drinking in mice. Neuropsychopharmacology 38,
437–445. doi: 10.1038/npp.2012.198
Lemieux, R. M. (1996). How water provides the impetus for molecular recognition
in aqueous solution. Acc. Chem. Res. 29, 373–380. doi: 10.1021/ar9600087
Levy, Y., and Onuchic, J. N. (2006). Water mediation in protein folding and
molecular recognition. Annu. Rev. Biophys. Biomol. Struct. 35, 389–415. doi:
10.1146/annurev.biophys.35.040405.102134
Lin, L., Huang, Q.-X., Yang, S.-S., Chu, J., Wang, J.-Z., and Tian, Q. (2013).
Melatonin in Alzheimer’s disease. Int. J. Mol. Sci. 14, 14575–14593. doi:
10.3390/ijms140714575
Mancuso, J. J., Cheng, J., Yin, Z., Gilliam, J. C., Xia, X., Li, X., et al.
(2014). Integration of multiscale dendritic spine structure and function
data into systems biology models. Front. Neuroanat. 8:130. doi:
10.3389/fnana.2014.00130
McCook, O., Georgieff, M., Scheuerle, A., Möller, P., Thiemermann, C., and
Radermacher, P. (2012). Erythropoietin in the critically ill: do we ask the right
questions? Crit. Care 16, 319. doi: 10.1186/cc11430
Meng, S., Qiao, M., Lin, L., Del Bigio, M. R., Tomanek, B., and Tuor, U. I. (2004).
Correspondence of AQP4 expression and hypoxic-ischaemic brain oedema
monitored by magnetic resonance imaging in the immature and juvenile rat.
Eur. J. Neurosci. 19, 2261–2269. doi: 10.1111/j.0953-816X.2004.03315.x
Moftakhar, P., Lynch, M. D., Pomakian, J. L., and Vinters, H. V. (2010).
Aquaporin expression in the brains of patients with or without cerebral
amyloid angiopathy. J. Neuropathol. Exp. Neurol. 69, 1201–1209. doi:
10.1097/NEN.0b013e3181fd252c
Morelli, M., Carta, A. R., Kachroo, A., and Schwarzschild, M. A. (2010).
Pathophysiological roles for purines: adenosine, caffeine and urate. Prog. Brain
Res. 183, 183–208. doi: 10.1016/S0079-6123(10)83010-9
Nagelhus, E. A., and Ottersen, O. P. (2013). Physiological roles of aquaporin-4 in
brain. Physiol. Rev. 93, 1543–1562. doi: 10.1152/physrev.00011.2013
Nakamura, K., Brown, R. A., Araujo, D., Narayanan, S., and Arnold, D. L. (2014).
Correlation between brain volume change and T2 relaxation time induced by
dehydration and rehydration: implications for monitoring atrophy in clinical
studies. Neuroimage 6, 166–170. doi: 10.1016/j.nicl.2014.08.014
Ogawa, E., Sakakibara, R., Endo, K., Tateno, F., Matsuzawa, Y., Hosoe, N.,
et al. (2011). Incidence of dehydration encephalopathy among patients with
disturbed consciousness at a hospital emergency unit. Clin. Prac. 1:e9. doi:
10.4081/cp.2011.e9
Oka, Y., Ye, M., and Zuker, C. S. (2015). Thirst driving and suppressing signals
encoded by distinct neural populations in the brain. Nature 520, 349–352. doi:
10.1038/nature14108
Ono, K., Toyono, T., and Inenaga, K. (2008). Nicotinic receptor subtypes in rat
subfornical organ neurons and glial cells. Neuroscience 154, 994–1001. doi:
10.1016/j.neuroscience.2008.04.028
Phillips, R. S. (2002). How does active site water affect enzymatic stereo
recognition? J. Mol. Cat. B. 19–20, 103–107. doi: 10.1016/S1381-
1177(02)00156-X
Picciotto, M. R., and Zoli, M. (2002). Nicotinic receptors in aging and dementia. J.
Neurobiol. 53, 641–655. doi: 10.1002/neu.10102
Pretorius, R. W., Gataric, G., Swedlund, S. K., and Miller, J. R. (2013). Reducing the
risk of adverse drug events in older adults. Am. Fam. Physician. 87, 331–336.
Qi, H., Qiu, X., Wang, C., Gaoa, Q., and Zhang, C. (2013). Digital electrogenerated
chemiluminescence biosensor for the determination of multiple proteins based
on Boolean logic gate. Anal. Methods 5, 612–615. doi: 10.1039/c2ay26054a
Reyes-Haro, D., Labrada-Moncada, F. E., Miledi, R., and Martínez-Torres, A.
(2015). Dehydration-induced anorexia reduces astrocyte density in the rat
corpus callosum. Neural Plast. 2015:474917. doi: 10.1155/2015/474917
Riebl, S. K., and Davy, B. M. (2013). The hydration equation: update on water
balance and cognitive performance. ACSM’s Health Fit. J. 17, 21–28. doi:
10.1249/FIT.0b013e3182a9570f
Rodríguez-Arellano, J. J., Parpura, V., Zorec, R., and Verkhratsky, A. (2016).
Astrocytes in physiological aging and Alzheimer’s disease. Neuroscience 323,
170–182. doi: 10.1016/j.neuroscience.2015.01.007
Saji, E., Arakawa, M., Yanagawa, K., Toyoshima, Y., Yokoseki, A., Okamoto, K.,
et al. (2013). Cognitive impairment and cortical degeneration in neuromyelitis
optica. Ann. Neurol. 73, 65–76. doi: 10.1002/ana.23721
Salminen, A., Ojala, J., Kaarniranta, K., Haapasalo, A., Hiltunen, M., and Soininen,
H. (2011). Astrocytes in the aging brain express characteristics of senescence-
associated secretory phenotype. Eur. J. Neurosci. 34, 3–11. doi: 10.1111/j.1460-
9568.2011.07738.x
Schlanger, L. E., Bailey, J. L., and Sands, J. M. (2010). Electrolytes in the aging. Adv.
Chronic Kidney Dis. 17, 308–319. doi: 10.1053/j.ackd.2010.03.008
Sen, S., and Voorheis, H. P. (2014). Protein folding: understanding the role
of water and the low Reynolds number environment as the peptide chain
emerges from the ribosome and folds. J. Theor. Biol. 363, 169–187. doi:
10.1016/j.jtbi.2014.07.025
Sfera, A., and Osorio, C. (2014). Water for thought: is there a role for aquaporin
channels in delirium? Front. Psychiatry 5:57. doi: 10.3389/fpsyt.2014.00057
Smythies, J. (2015). On the possible role of protein vibrations in information
processing in the brain: three Russian dolls. Front. Mol. Neurosci. 8:38. doi:
10.3389/fnmol.2015.00038
St. Hillaire, C., Vargas, D., Pardo, C. A., Gincel, D., Mann, J., Rothstein, J. D., et al.
(2005). Aquaporin 4 is increased in association with human immunodeficiency
virus dementia: implications for disease pathogenesis. J. Neurovirol. 11,
535–543. doi: 10.1080/13550280500385203
Stoppini, M., Obici, L., Lavatelli, F., Giorgetti, S., Marchese, L., Moratti, R., et al.
(2009). Proteomics in protein misfolding diseases. Clin. Chem. Lab. Med. 47,
627–635. doi: 10.1515/CCLM.2009.164
Streitbürger, D.-P., Möller, H. E., Tittgemeyer, M., Hund-Georgiadis, M., Schroeter,
M. L., and Mueller, K. (2012). Investigating structural brain changes of
dehydration using voxel-based morphometry. PLoS ONE 7:e44195. doi:
10.1371/journal.pone.0044195
Subburaman, T. T., and Vanisree, A. J. (2011). Oxidative pathology and AQP4
mRNA expression in patients of Parkinson’s disease in Tamil Nadu. Ann.
Neurosci. 18, 109–112. doi: 10.5214/ans.0972.7531.1118306
Tait, M. J, Saadoun, S., Bell, B. A, and Papadopoulos, M. C. (2008). Water
movements in the brain: role of aquaporins. Trends Neurosci. 31, 37–43. doi:
10.1016/j.tins.2007.11.003
Tanaka, J. (2003). Activation of cholinergic pathways from the septum
to the subfornical organ area under hypovolemic condition in
Frontiers in Molecular Biosciences | www.frontiersin.org 6May 2016 | Volume 3 | Article 18
Sfera et al. Biomarkers of Dehydration in Geriatrics
rats. Brain Res. Bull. 61, 497–504. doi: 10.1016/S0361-9230(03)
00186-2
Tanaka, M., Tanaka, K., Komori, M., and Saida, T. (2007). Anti-aquaporin
4 antibody in Japanese multiple sclerosis: the presence of optic spinal
multiple sclerosis without long spinal cord lesions and anti-aquaporin 4
antibody. J. Neurol. Neurosurg. Psychiatry 78, 990–992. doi: 10.1136/jnnp.2006.
114165
Thrane, A. S., Thrane, V. R., and Nedergaard, M. (2014). Drowning stars:
reassessing the role of astrocytes in brain edema. Trends Neurosci. 37, 620–628.
doi: 10.1016/j.tins.2014.08.010
Tokuriki, N., Kinjo, M., Negi, S., Hoshino, M., Goto, Y., Urabe, I., et al. (2004).
Protein folding by the effects of macromolecular crowding. Protein Sci. 13,
125–133. doi: 10.1110/ps.03288104
Trinh-Trang-Tan, M. M., Geelen, G., Teillet, L., and Corman, B. (2003).
Urea transporter expression in aging kidney and brain during dehydration.
Am. J. Physiol. Regul. Integr. Comp. Physiol. 285, R1355–R1365. doi:
10.1152/ajpregu.00207.2003
Utsugisawa, K., Nagane, Y., Tohgi, H., Yoshimura, M., Ohba, H., and Genda, Y.
(1999). Changes with aging and ischemia in nicotinic acetylcholine receptor
subunit alpha7 mRNA expression in postmortem human frontal cortex and
putamen. Neurosci. Lett. 270, 145–148. doi: 10.1016/S0304-3940(99)00473-5
Vajda, T., and Perczel, A. (2014). Role of water in protein folding, oligomerization,
amyloidosis and miniprotein. J. Pept. Sci. 20, 747–759. doi: 10.1002/psc.2671
Wang, B. F., Cui, Z. W., Zhong, Z. H., Sun, Y. H., Sun, Q. F., Yang,
G. Y., et al. (2015). Curcumin attenuates brain edema in mice with
intracerebral hemorrhage through inhibition of AQP4 and AQP9 expression.
Acta Pharmacol. Sin. 36, 939–948. doi: 10.1038/aps.2015.47
Warren, J. L., Bacon, W. E., Harris, T., McBean, A. M., Foley, D. J., and Phillips,
C. (1994). The burden and outcomes associated with dehydration among US
elderly, 1991. Am. J. Public Health. 84, 1265–1269. doi: 10.2105/AJPH.84.
8.1265
Xiao, H., Barber, J., and Campbell, E. S. (2004). Economic burden of
dehydration among hospitalized elderly patients. Am. J. Health Syst. Pharm. 61,
2534–2540.
Xie, L., Kang, H., Xu, Q., Chen, M. J., Liao, Y., Thiyagarajan, M., et al. (2013).
Sleep drives metabolite clearance from the adult brain. Science 342, 373–377.
doi: 10.1126/science.1241224
Yang, M., Gao, F., Liu, H., Yu, W. H., Zhuo, F., Qiu, G. P., et al. (2013).
Hyperosmotic induction of aquaporin expression in rat astrocytes through a
different MAPK pathway. J. Cell Biochem. 114, 111–119. doi: 10.1002/jcb.24308
Yerbury, J. J., Stewart, E. M., Wyatt, A. R., and Wilson, M. R. (2005). Quality
control of protein folding in extracellular space. EMBO Rep. 6, 1131–1136. doi:
10.1038/sj.embor.7400586
Zador, Z., Stiver, S., Wang, V., and Manley, G. T. (2009). Role of aquaporin-4 in
cerebral edema and stroke. Handb. Exp. Pharmacol. 159–170. doi: 10.1007/978-
3-540-79885-9_7
Zhang, J., Yang, B., Sun, H., Zhou, Y., Liu, M., Ding, J., et al. (2016). Aquaporin-4
deficiency diminishes the differential degeneration of midbrain dopaminergic
neurons in experimental Parkinson’s disease. Neurosci. Lett. 614, 7–15. doi:
10.1016/j.neulet.2015.12.057
Zhang, N., Li, Y. J., Fu, Y., Shao, J. H., Luo, L. L., Yang, L., et al. (2015). Cognitive
impairment in Chinese neuromyelitis optica. Mult. Scler. 21, 1839–1846. doi:
10.1177/1352458515576982
Zhao, L., Li, W., and Tian, P. (2013). Reconciling mediating and slaving roles
of water in protein conformational dynamics. PLoS ONE 8:e60553. doi:
10.1371/journal.pone.0060553
Conflict of Interest Statement: The authors declare that the research was
conducted in the absence of any commercial or financial relationships that could
be construed as a potential conflict of interest.
Copyright © 2016 Sfera, Cummings and Osorio. This is an open-access article
distributed under the terms of the Creative Commons Attribution License (CC
BY). The use, distribution or reproduction in other forums is permitted, provided
the original author(s) or licensor are credited and that the original publication
in this journal is cited, in accordance with accepted academic practice. No
use, distribution or reproduction is permitted which does not comply with these
terms.
Frontiers in Molecular Biosciences | www.frontiersin.org 7May 2016 | Volume 3 | Article 18
... Additionally, it was indicated that impairment was greater in studies reporting >2% body mass loss in comparison with 2% [6]. There is also evidence, showing that dehydration can accelerate cognitive decline in people with dementia [7]. ...
Article
Purpose of review The aim of this study was to conduct a review of the literature published over the past 18 months and present the latest findings on hydration in individuals with dementia. Recent findings A systematic review identified sarcopenia, polypharmacy, delayed oral transit, and poor rinsing ability as markers of eating-drinking-swallowing difficulties in early-stage dementia. A cross-sectional study found a high prevalence of dehydration (57–68%) among dementia patients, associated with hypertension, diabetes, chronic kidney disease, dysphagia, and cognitive decline. An analysis of national dementia care guidelines showed that only the UK and Switzerland addressed assisted nutrition and hydration. “Jelly Drops,” a hydrating product for dementia patients, received innovation awards. A study on US Physician Orders for Life Sustaining Treatment forms highlighted inconsistent terminology for end-of-life nutrition and hydration, calling for clearer language to aid decision-making. Summary The new hydration-related recommendations in the ESPEN 2024 guidelines for dementia reflect a more individualized, proactive, and comprehensive approach to managing hydration. These guidelines emphasize the importance of early detection, personalized interventions, and consistent monitoring to ensure that dehydration is identified and treated promptly. Furthermore, the current literature supports the need for a specific terminology for dementia management using nutrition and hydration to improve patients’ health outcomes.
... One of the problems said to result from dehydration is that amyloid protein is more likely not to be cleared from the brain and for it to misfold to form neurotoxic amyloid plaques (Sfera et al., 2016). Both dehydration and sleeplessness are likely to be marked by symptoms of fatigue, confusion, "brain fog," and memory lapses. ...
... One of the problems said to result from dehydration is that amyloid protein is more likely not to be cleared from the brain and for it to misfold to form neurotoxic amyloid plaques (Sfera et al., 2016). Both dehydration and sleeplessness are likely to be marked by symptoms of fatigue, confusion, "brain fog," and memory lapses. ...
Article
Advances in science come both from discovering new facts and from thinking in new ways about known facts. For example, Albert Einstein’s Special Theory of Relativity didn’t add new data, but it better connected the dots about experimental facts that were at odds with Newtonian physics. Now, medical science requires new thinking to connect the dots as scientists search for the causes of Alzheimer's Disease (AD). Using a systems-thinking lens offers significant advantages in hypothesizing the potential role of dehydration in causing AD. Most critically, it enables us to consider (a) how multiple elements function and interact within the various systems in the body and (b) how dehydration can operate in combination with other AD risk factors -- whether sequentially, cumulatively, synergistically, or by way of a destructive feedback loop -- to damage the brain and cause cognitive dysfunction. Numerous dehydrating episodes, created by chronic use of dehydrating medicines over many years, can inflict an inciting injury that creates a self-reinforcing loop that causes AD. Rapid neurodegeneration late in life because of AD is explained by the operation of the destructive reinforcing loop. Women are more likely than men to become diagnosed as having AD because they are more likely to experience during their lifetimes chronic use of dehydrating medicines. Clearly, it would be a breakthrough in medical science and practice if significantly lower rates for women developing AD could be achieved by simply developing medicine-taking regimens that would be less dehydrating.
... but with no effect of occupancy and no interaction between activity in the COVID-19 pandemic and occupancy (see Figure 4a) (2). Next, we modelled each pandemic period separately (see Figure 4b) (3), with the main effect of COVID-19 being a significant increase in kitchen activity from P1 onwards (F(6,126) = 10.77***) but with no significant effect of occupancy or interaction between activity across the pandemic periods and household occupancy (4). ...
Preprint
Full-text available
Malnutrition and dehydration are strongly associated with increased cognitive and functional decline in people living with dementia (PLWD), as well as an increased rate of hospitalisations in comparison to their healthy counterparts. Extreme changes in eating and drinking behaviours can often lead to mal- nutrition and dehydration, accelerating the progression of cognitive and functional decline and resulting in a marked reduction in quality of life. Unfortunately, there are currently no established methods by which to objectively detect such changes. Here, we present the findings of a quantitative analysis conducted on in-home monitoring data collected from 73 households of PLWD. The Coronavirus 2019 (COVID-19) pandemic has previously been shown to have dramatically altered the behavioural habits, particularly the eating and drinking habits, of PLWD. Using the COVID-19 pandemic as a natural experiment, we show that there are significant changes in eating and drinking habits at the group level within a subset of 21 households of PLWD that were continuously monitored for 499 days, with an overall increase in day-time activities and a decrease in night-time activity observed in both single and multiple occupancy households. We further present preliminary results suggesting it is possible to proactively detect episodic and gradual changes in behaviours. Together, these results pave the way to introduce improvements into the monitoring of PLWD in naturalistic settings and for shifting from reactive to proactive care.
Article
A number of experimental studies are described that challenge the significance of synaptic plasticity and prove the role of transposable elements in memory consolidation. This is due to the cis-regulatory influence of activated transposable elements on gene expression, as well as insertions into new genomic loci near the genes involved in brain functioning. RNAs and proteins of endogenous retroviruses are transported to dendritic synapses and transmit information to change gene expression in neighboring cells through the formation of virus-like particles in vesicles. Due to this, the relationship between synaptic plasticity and nuclear coding is ensured, since transposable elements are also drivers of epigenetic regulation due to relationship with the non-coding RNAs derived from them. Our analysis of the scientific literature allowed us to identify the role of 17 microRNAs derived from transposable elements in normal memory formation. In neurodegenerative diseases with memory impairment, we identified impaired expression of 44 microRNAs derived from transposable elements. This demonstrates the potential for targeting pathological transposon activation in neurodegenerative diseases for memory restoration using microRNAs as tools.
Preprint
Malnutrition and dehydration are strongly associated with increased cognitive and functional decline in people living with dementia (PLWD), as well as an increased rate of hospitalisations in comparison to their healthy counterparts. Extreme changes in eating and drinking behaviours can often lead to malnutrition and dehydration, accelerating the progression of cognitive and functional decline and resulting in a marked reduction in quality of life. Unfortunately, there are currently no established methods by which to objectively detect such changes. Here, we present the findings of an extensive quantitative analysis conducted on in-home monitoring data collected from 73 households of PLWD using Internet of Things technologies. The Coronavirus 2019 (COVID-19) pandemic has previously been shown to have dramatically altered the behavioural habits, particularly the eating and drinking habits, of PLWD. Using the COVID-19 pandemic as a natural experiment, we conducted linear mixed-effects modelling to examine changes in mean kitchen activity within a subset of 21 households of PLWD that were continuously monitored for 499 days. We report an observable increase in day-time kitchen activity and a significant decrease in night-time kitchen activity (t(147) = -2.90, p < 0.001). We further propose a novel analytical approach to detecting changes in behaviours of PLWD using Markov modelling applied to remote monitoring data as a proxy for behaviours that cannot be directly measured. Together, these results pave the way to introduce improvements into the monitoring of PLWD in naturalistic settings and for shifting from reactive to proactive care.
Article
Malnutrition and dehydration are strongly associated with increased cognitive and functional decline in people living with dementia (PLWD), as well as an increased rate of hospitalisations in comparison to their healthy counterparts. Extreme changes in eating and drinking behaviors can often lead to malnutrition and dehydration, accelerating the progression of cognitive and functional decline and resulting in a marked reduction in quality of life. Unfortunately, there are currently no established methods by which to objectively detect such changes. Here, we present the findings of an extensive quantitative analysis conducted on in-home monitoring data collected from 73 households of PLWD using Internet of Things technologies. The Coronavirus 2019 (COVID-19) pandemic has previously been shown to have dramatically altered the behavioral habits, particularly the eating and drinking habits, of PLWD. Using the COVID-19 pandemic as a natural experiment, we conducted linear mixed-effects modeling to examine changes in mean kitchen activity within a subset of 21 households of PLWD that were continuously monitored for 499 days. We report an observable increase in day-time kitchen activity and a significant decrease in night-time kitchen activity ( t (147) =2.90=\,\,-2.90 , p{p} < 0.001). We further propose a novel analytical approach to detecting changes in behaviors of PLWD using Markov modeling applied to remote monitoring data as a proxy for behaviors that cannot be directly measured. Together, these results pave the way to introduce improvements into the monitoring of PLWD in naturalistic settings and for shifting from reactive to proactive care.
Article
Objectives: Delirium is highly prevalent in hospitalised older adults, under-diagnosed and associated with poor outcomes. We aim to determine (i) association of frailty measured using Hospital Frailty Risk Score (HFRS) with delirium, (ii) impact of delirium on mortality, 30-days readmission, extended length of stay (eLOS) and cost (eCOST). Methods: Retrospective cohort study was conducted on 902 older adults ≥75 years discharged from an academic tertiary hospital between March and September 2021. Data was obtained from hospital administrative database. Results: Delirium was prevalent in 39.1%, 58.1% were female with mean age 85.3 ± 6.2 years. Patients with delirium were significantly older, had higher HFRS, pneumonia, urinary tract infection (UTI), E.coli and Klebsiella infection, constipation, dehydration, stroke and intracranial bleed, with comorbidities including dementia, diabetes, hypertension, hyperlipidaemia and chronic kidney disease. In-hospital mortality, 30-days mortality, 30-days readmission, median LOS and cost was significantly higher. Delirium was significantly associated with at least intermediate frailty (OR = 3.52; CI = 2.48-4.98), dementia (OR = 2.39; CI = 1.61-3.54), UTI (OR = 1.95; CI = 1.29-2.95), constipation (OR = 2.49; CI = 1.43-4.33), Klebsiella infection (OR = 3.06; CI = 1.28-7.30), dehydration (OR = 2.01; CI = 1.40 - 2.88), 30-day mortality (OR = 2.52; CI = 1.42-4.47), 30-day readmission (OR = 2.18; CI = 1.36-3.48), eLOS (OR = 1.80; CI = 1.30-2.49) and eCOST (OR = 1.67; CI = 1.20-2.35). Conclusions: Delirium was highly prevalent in older inpatients, and associated with dementia, frailty, increased cost, LOS, 30-day readmissions and mortality. Hospital Frailty Risk Score had robust association with delirium and can be auto-populated from electronic medical records. Prospective studies are needed on multicomponent delirium preventive measures in high-risk groups identified by HFRS in acute care settings.
Article
Objectives Older adults have an elevated risk of dehydration, a state with proven detrimental cognitive and physical effects. Furthermore, the use of diuretics by hypertensive patients further compounds this risk. This prospective study investigated the diagnostic accuracy of point-of-care (POC) salivary osmolarity (SOSM) measurement for the detection of dehydration in hypertensive adults with and without diuretic pharmacotherapy. Design Prospective diagnostic accuracy study. Setting Home visits to patients recruited from 4 community health centers in West Sulawesi, Indonesia. Participants A total of 148 hypertensive older adults (57 men, 91 women). The mean ages of male and female patients were 69.4 ± 11.4 and 68.1 ± 7.8 years, respectively. Methods Hypertensive adults were divided into 2 groups based on the presence of diuretics in their pharmacotherapeutic regimen. First-morning mid-stream urine samples were used to perform urine specific gravity (USG) testing. Same-day SOSM measurements were obtained using a POC saliva testing system. Results Both USG (P = .0002) and SOSM (P < .0001) were significantly elevated in hypertensive patients with diuretic pharmacotherapy. At a USG threshold of ≥1.030, 86% of diuretic users were classified as dehydrated compared with 55% of non-using participants. A strong correlation was observed between USG and SOSM measurements (r = 0.78, P < .0001). Using a USG threshold of ≥1.030 as a hydration classifier, an SOSM threshold of ≥93 mOsm had a sensitivity of 78.6% and a specificity of 91.1% for detecting dehydration. Conclusions and Implications Hypertensive patients on diuretics have significantly higher first-morning USG and SOSM values, indicating a higher likelihood of dehydration relative to those on other classes of antihypertensive medication. POC SOSM assessment correlates strongly with first-morning USG assessment, and represents a rapid and noninvasive alternative to urinary hydration assessment that may be applicable for routine use in populations with elevated risk of dehydration.
Article
Full-text available
Parkinson's disease (PD) is primarily due to the progressive, selective and irreversible loss of dopaminergic (DA) neurons in the substantia nigra (SN). Interestingly, DA neurons in the ventral and lateral SN are much more susceptible than adjacent dopamine neurons in the ventral tegmental area (VTA) not only in human PD but in many PD model systems. However, the molecular causes of regional vulnerability in PD remain unknown. In our previous studies, we established acute PD animal models by administration of MPTP (1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine), and found that AQP4 knockout mice were significantly more prone to MPTP-induced neurotoxicity. Here, we further observe that AQP4 deficiency resulted in the same susceptible to MPTP between SN DA neuron and VTA neurons both in acute and chronic PD model. Moreover, we show that AQP4 deficiency increased the numbers of reactive astrocytes and microglias not only in the SN and but also in the VTA under basal and MPTP-induced situations. Meanwhile, AQP4 deficiency disrupted the balance of the pro-inflammatory cytokine/neurotrophin in midbrain. Taken together, these results demonstrate that glial AQP4 is involved in the susceptibility differences of DA neurons between SN and VTA, although the precise mechanism of AQP4 remains to be explored. Moreover, these findings also suggest that these susceptibility differences are not only due to intrinsic neuronal factors, but also attribute to differences in astrocytes of these regions.
Article
Full-text available
Aquaporin 4 (AQP4) is the major water channel expressed in the central nervous system (CNS), and it is primarily expressed in astrocytes. It has been studied in various brain pathological conditions. However, the potential for AQP4 to influence Alzheimer's disease (AD) is still unclear. Research regarding AQP4 functions related to AD can be traced back several years and has gradually progressed toward a better understanding of the potential mechanisms. Currently, it has been suggested that AQP4 influences synaptic plasticity, and AQP4 deficiency may impair learning and memory, in part, through glutamate transporter-1 (GLT-1). AQP4 may mediate the clearance of amyloid beta peptides (Aβ). In addition, AQP4 may influence potassium (K(+)) and calcium (Ca(2+)) ion transport, which could play decisive roles in the pathogenesis of AD. Furthermore, AQP4 knockout is involved in neuroinflammation and interferes with AD. To date, no specific therapeutic agents have been developed to inhibit or enhance AQP4. However, experimental results strongly emphasize the importance of this topic for future investigations.
Article
Full-text available
Until recently it was held that the neurocomputations conducted by the brain involved only whole neurons as the operating units. This may however represent only a part of the mechanism. This theoretical and academic position article reviews the considerable evidence that allosteric interactions between proteins (as extensively described by Fuxe et al., 2014), and in particular protein vibrations in neurons, form small scale codes that are involved as parts of the complex information processing systems of the brain. The argument is then developed to suggest that the protein allosteric and vibration codes (that operate at the molecular level) are nested within a medium scale coding system whose computational units are organelles (such as microtubules). This medium scale code is nested in turn inside a large scale coding system, whose computational units are individual neurons. The hypothesis suggests that these three levels interact vertically in both directions thus materially increasing the computational capacity of the brain. The whole hierarchy is thus similar to three nested Russian dolls. This theoretical development may be of use in the design of experiments to test it.
Article
Full-text available
Aquaporins (AQPs) are the water-channels that play important roles in brain water homeostasis and in cerebral edema induced by brain injury. In this study we investigated the relationship between AQPs and a neuroprotective agent curcumin that was effective in the treatment of brain edema in mice with intracerebral hemorrhage (ICH). ICH was induced in mice by autologous blood infusion. The mice immediately received curcumin (75, 150, 300 mg/kg, ip). The Rotarod test scores, brain water content and brain expression of AQPs were measured post ICH. Cultured primary mouse astrocytes were used for in vitro experiments. The expression of AQP1, AQP4 and AQP9 and NF-κB p65 were detected using Western blotting or immunochemistry staining. Curcumin administration dose-dependently reduced the cerebral edema at day 3 post ICH, and significantly attenuated the neurological deficits at day 5 post ICH. Furthermore, curcumin dose-dependently decreased the gene and protein expression of AQP4 and AQP9, but not AQP1 post ICH. Treatment of the cultured astrocytes with Fe(2+) (10-100 μmol/L) dose-dependently increased the expression and nuclear translocation of NF-κB p65 and the expression of AQP4 and AQP9, which were partly blocked by co-treatment with curcumin (20 μmol/L) or the NF-κB inhibitor PDTC (10 μmol/L). Curcumin effectively attenuates brain edema in mice with ICH through inhibition of the NF-κB pathway and subsequently the expression of AQP4 and AQP9. Curcumin may serve as a potential therapeutic agent for ICH.
Article
Full-text available
Alzheimer’s disease (AD), an age-related neurodegenerative disorder with progressive cognition deficit, is characterized by extracellular senile plaques (SP) of aggregated β-amyloid (Aβ) and intracellular neurofibrillary tangles, mainly containing the hyperphosphorylated microtubule-associated protein tau. Multiple factors contribute to the etiology of AD in terms of initiation and progression. Melatonin is an endogenously produced hormone in the brain and decreases during aging and in patients with AD. Data from clinical trials indicate that melatonin supplementation improves sleep, ameliorates sundowning and slows down the progression of cognitive impairment in AD patients. Melatonin efficiently protects neuronal cells from Aβ-mediated toxicity via antioxidant and anti-amyloid properties. It not only inhibits Aβ generation, but also arrests the formation of amyloid fibrils by a structure-dependent interaction with Aβ. Our studies have demonstrated that melatonin efficiently attenuates Alzheimer-like tau hyperphosphorylation. Although the exact mechanism is still not fully understood, a direct regulatory influence of melatonin on the activities of protein kinases and protein phosphatases is proposed. Additionally, melatonin also plays a role in protecting the cholinergic system and in anti-inflammation. The aim of this review is to stimulate interest in melatonin as a potentially useful agent in the prevention and treatment of AD.
Article
Full-text available
Anorexia nervosa is an eating disorder associated with severe weight loss as a consequence of voluntary food intake avoidance. Animal models such as dehydration-induced anorexia (DIA) mimic core features of the disorder, including voluntary reduction in food intake, which compromises the supply of energy to the brain. Glial cells, the major population of nerve cells in the central nervous system, play a crucial role in supplying energy to the neurons. The corpus callosum (CC) is the largest white matter tract in mammals, and more than 99% of the cell somata correspond to glial cells in rodents. Whether glial cell density is altered in anorexia is unknown. Thus, the aim of this study was to estimate glial cell density in the three main regions of the CC (genu, body, and splenium) in a murine model of DIA. The astrocyte density was significantly reduced (~34%) for the DIA group in the body of the CC, whereas in the genu and the splenium no significant changes were observed. DIA and forced food restriction (FFR) also reduced the ratio of astrocytes to glial cells by 57.5% and 22%, respectively, in the body of CC. Thus, we conclude that DIA reduces astrocyte density only in the body of the rat CC.
Article
Full-text available
Dehydration is the most common fluid and electrolyte problem among elderly patients. It is reported to be widely prevalent and costly to individuals and to the health care system. The purpose of this review is to summarize the literature on the economic burden of dehydration in the elderly. A comprehensive search of several databases from database inception to November 2013, only in English language, was conducted. The databases included Pubmed and ISI Web of Science. The search terms «dehydration» / "hyponaremia" / "hypernatremia" AND «cost» AND «elderly» were used to search for comparative studies of the economic burden of dehydration. A total of 15 papers were identified. Dehydration in the elderly is an independent factor of higher health care expenditures. It is directly associated with an increase in hospital mortality, as well as with an increase in the utilization of ICU, short and long term care facilities, readmission rates and hospital resources, especially among those with moderate to severe hyponatremia. Dehydration represents a potential target for intervention to reduce healthcare expenditures and improve patients' quality of life.
Article
Delirium is a common problem associated with considerable morbidity and mortality for older hospitalized patients. Recognition of this condition may be difficult but can be improved by cognitive assessment and the use of a simplified diagnostic tool. The etiology is multifactorial; thus, solving one factor is unlikely to resolve the delirium. However, many cases may be preventable through a targeted risk factor approach. Nonpharmacologic approaches to management - such as the use of massage, relaxation music, and reorientation strategies - are recommended. Medications with psychoactive effects (including sedative-hypnotics, narcotics, and anticholinergic drugs) are among the most frequent contributors to delirium. Pharmacologic management with haloperidol should be reserved for patients whose severe agitation will result in interruption of essential medical therapies or who pose a danger to themselves or staff members.
Article
The extracellular aggregation of proteins into proteotoxic oligomers and amyloid fibrils is implicated in the onset and pathology of numerous diseases referred to as amyloid diseases. All of the proteins that aggregate extracellularly in association with amyloid disease pathogenesis originate in the endoplasmic reticulum (ER) and are secreted through the secretory pathway. Disruptions in ER protein homeostasis or proteostasis (i.e., ER stress) can facilitate the aberrant secretion of misfolded protein conformations to the extracellular space and exacerbate pathologic protein aggregation into proteotoxic species. Activation of an ER stress-responsive signaling pathway, the Unfolded Protein Response (UPR), restores ER proteostasis through the transcriptional regulation of ER proteostasis pathways. In contrast, the functional role for the UPR in regulating extracellular proteostasis during ER stress is poorly defined. We recently identified ERdj3 as a UPR-regulated secreted chaperone that increases extracellular proteostasis capacity in response to ER stress, revealing a previously-unanticipated direct mechanism by which the UPR impacts extracellular proteostasis. Here, we discuss the functional implications of ERdj3 secretion on extracellular proteostasis maintenance and define the mechanisms by which ERdj3 secretion coordinates intra- and extracellular proteostasis environments during ER stress.
Article
Protein aggregation in aqueous cellular environments is linked to diverse human diseases. Protein aggregation proceeds through a multistep process initiated by conformational transitions, called protein misfolding, of monomer species toward aggregation-prone structures. Various forms of aggregate species are generated through the association of misfolded monomers including soluble oligomers and amyloid fibrils. Elucidating the molecular mechanisms and driving forces involved in the misfolding and subsequent association has been a central issue for understanding and preventing protein aggregation diseases such as Alzheimer's, Parkinson's, and type II diabetes. In this Account, we provide a thermodynamic perspective of the misfolding and aggregation of the amyloid-beta (Aβ) protein implicated in Alzheimer's disease through the application of fluctuating thermodynamics. This approach "dissects" the conventional thermodynamic characterization of the end states into the one of the fluctuating processes connecting them, and enables one to analyze variations in the thermodynamic functions that occur during the course of protein conformational changes. The central quantity in this approach is the solvent-averaged effective energy, f = Eu + Gsolv, comprising the protein potential energy (Eu) and the solvation free energy (Gsolv), whose time variation reflects the protein dynamics on the free energy landscape. Protein configurational entropy is quantified by the magnitude of fluctuations in f. We find that misfolding of the Aβ monomer when released from a membrane environment to an aqueous phase is driven by favorable changes in protein potential energy and configurational entropy, but it is also accompanied by an unfavorable increase in solvation free energy. The subsequent dimerization of the misfolded Aβ monomers occurs in two steps. The first step, where two widely separated monomers come into contact distance, is driven by water-mediated attraction, that is, by a decrease in solvation free energy, harnessing the monomer solvation free energy earned during the misfolding. The second step, where a compact dimer structure is formed, is driven by direct protein-protein interactions, but again it is accompanied by an increase in solvation free energy. The increased solvation free energy of the dimer will function as the driving force to recruit another Aβ protein in the approach stage of subsequent oligomerizations. The fluctuating thermodynamics analysis of the misfolding and dimerization of the Aβ protein indicates that the interaction of the protein with surrounding water plays a critical role in protein aggregation. Such a water-centric perspective is further corroborated by demonstrating that, for a large number of Aβ mutants and mutants of other protein systems, the change in the experimental aggregation propensity upon mutation has a significant correlation with the protein solvation free energy change. We also find striking discrimination between the positively and negatively charged residues on the protein surface by surrounding water molecules, which is shown to play a crucial role in determining the protein aggregation propensity. We argue that the protein total charge dictates such striking behavior of the surrounding water molecules. Our results provide new insights for understanding and predicting the protein aggregation propensity, thereby offering novel design principles for producing aggregation-resistant proteins for biotherapeutics.