ArticlePDF AvailableLiterature Review

Post-Operative Cognitive Dysfunction: An exploration of the inflammatory hypothesis and novel therapies

Authors:

Abstract and Figures

Post-Operative Cognitive Dysfunction (POCD) is a highly prevalent condition with significant clinical, social and financial impacts for patients and their communities. The underlying pathophysiology is becoming increasingly understood, with the role of neuroinflammation and oxidative stress secondary to surgery and anaesthesia strongly implicated. This review aims to describe the putative mechanisms by which surgery-induced inflammation produces cognitive sequelae, with a focus on identifying potential novel therapies based upon their ability to modify these pathways.
No caption available
… 
Content may be subject to copyright.
Post-Operative Cognitive Dysfunction: An exploration of the inflammatory hypothesis
and novel therapies.
*David R. Skvarc a, c, dskvarc@deakin.edu.au
Michael Berk c, d, e, mikebe@barwonhealth.org.au
Linda K. Byrne a, linda.byrne@deakin.edu.au
Olivia M. Dean c, d, e, olivia.dean@barwonhealth.org.au
Seetal Dodd c, d, seetald@barwonhealth.org.au
Matthew Lewis a, f, m.lewis@cgmc.org.au
Andrew Marriott b, c, d, e, amarri@barwonhealth.org.au
Eileen M. Moore b, c, eileenmo@barwonhealth.org.au
Gerwyn Morris g, activatedmicroglia@gmail.com
Richard S. Paged, h, Richard.page@deakin.edu.au
& Laura Gray d, l.gray@deakin.edu.au
*Corresponding author: dskvarc@deakin.edu.au
A. School of Psychology, Deakin University, Melbourne, Australia. B. Department of
Anaesthesia, Perioperative Medicine & Pain Management, Barwon Health, Geelong,
Australia. C. Deakin University, Innovations in Mental and Physical Health and Clinical
Treatment (IMPACT) Strategic Research Centre, Barwon Health, Geelong, Australia. D.
Deakin University, School of Medicine, Geelong, Australia. E. Orygen, The National Centre of
Excellence in Youth Mental Health, the Department of Psychiatry and the Florey Institute of
Neuroscience and Mental Health, the University of Melbourne, Parkville, Australia. F. Aged
Psychiatry Service, Caulfield Hospital, Alfred Health, Caulfield, Australia. G. Tir Na Nog. H.
Department of Orthopaedics, Barwon Health, Geelong, Australia.
Post-Operative Cognitive Dysfunction: An exploration of the inflammatory hypothesis
and novel therapies.
Abstract
Post-Operative Cognitive Dysfunction (POCD) is a highly prevalent condition with significant
clinical, social and financial impacts for patients and their communities. The underlying
pathophysiology is becoming increasingly understood, with the role of neuroinflammation
and oxidative stress secondary to surgery and anaesthesia strongly implicated. This review
aims to describe the putative mechanisms by which surgery-induced inflammation produces
cognitive sequelae, with a focus on identifying potential novel therapies based upon their
ability to modify these pathways.
Abbreviations
AD, Alzheimer’s disease; AMPA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid;
ApoE, apolipoprotein E; BBB, Blood brain barrier; BDNF, brain-derived neurotrophic factor;
CNS; Central nervous system; CRP, C-reactive protein; GABA, gamma-aminobutyric acid;
HMGB1; High-mobility group 1 box; IL, Interleukin; NAC, N-acetylcysteine; NADPH,
nicotinamide adenine dinucleotide phosphate; NMDA, N-Methyl-D-aspartate; POCD, Post-
Operative Cognitive Dysfunction; TNF-α; Tumour necrosis factor alpha.
Keywords
Cognition, Post-operative cognitive dysfunction, biomarkers, neuroinflammation, oxidative
stress
1. Introduction
Cognitive dysfunction following a surgical procedure can have a significant impact upon
patients’ health, function and wellbeing. Post-Operative Cognitive Dysfunction or Decline
(POCD) is a controversial condition that is the subject of much debate, but is purported to
encompass acute or persistent deficits in attention, concentration, learning and memory
following surgery that are not attributable to an overt complication or insult arising from the
procedure (Niccolò Terrando et al., 2011; Tsai, Sands, & Leung, 2010). POCD is distinct from
delirium (Tsai, Sands, & Leung, 2010), which describes an acute fluctuating and transient
disturbance to perception and cognition that is not attributable to a pre-existing
neurocognitive disorder (European Delirium Association and American Delirium Society,
2014). The definition of POCD varies markedly across studies but is typically inferred from a
comparison of pre-operative to post-operative cognitive function. The investigation of
POCD typically has focussed on an acute phase, lasting up to the time of discharge, and a
more prolonged cognitive deficit that can occur months to years following the surgical event
(Tsai, Sands, & Leung). In the absence of a DSM or ICD classification or biological marker, the
condition has been defined by consensus as “a spectrum of postoperative central nervous
dysfunction both acute and persistent” (Tsai, Sands, & Leung).
POCD can occur in between 25% to 40% of elderly patients at the point of discharge, and
the long-term impact of this is marked by a significantly higher mortality rate as compared
to age and sex matched controls without POCD (Moller et al., 1998; Monk et al., 2008).
These changes persist for some time, with 25% of patients over 65 years of age displaying
evidence of POCD at 7 days following hip replacement surgery (M.-H. Ji et al., 2012; LI, XI,
AN, DONG, & ZHOU, 2012) and 25% of patients displaying significant impairments in
executive functions and/or memory functions 3 months following surgical intervention
(Price, Garvan, & Monk, 2008). The consequences for such patients can be significant; a
longitudinal follow up of 700 patients diagnosed with POCD 3 months following surgery
revealed a significantly higher rate of death and disability leading to loss of employment and
a greater reliance on social security (Steinmetz, Christensen, Lund, Lohse, & Rasmussen,
2009).
Bedford (1955), first described post-operative symptoms consistent with POCD. Since then
there has been a concerted effort to determine aetiological factors that underpin the
cognitive dysfunction. The majority of the early work has focused on factors specific to
cardiac surgery but no aetiological factors conclusively emerged. Subsequent work reported
that in the medium term similar numbers of people undergoing non-cardiac surgery develop
POCD (Evered, Scott, Silbert, & Maruff, 2011) which has necessitated a broader examination
of possible underlying causes. Increased attention has been focused on inflammatory
markers. Preclinical studies suggest that neuroinflammation and oxidative events are key
mechanisms underpinning the development of POCD (Cibelli et al., 2010; N. Terrando et al.,
2010). This has also been shown in replicated clinical studies, with patients also displaying
elevated levels of pro-inflammatory cytokines in both the systemic circulatory and central
nervous systems following surgery associated with the degree of cognitive decline
(Beloosesky et al., 2007; Buvanendran et al., 2006). However, despite the strengthening
evidence supporting this aetiological model, there has been minimal translation into
therapeutic development, with no established preventive or ameliorating therapies. This
review will explore the pathophysiology of POCD focussing on the aetiological role of
neuroinflammatory and oxidative processes. Such mechanisms suggest potential therapies
acting on such targets, including N-acetyl cysteine, whose capacity to act as an anti-oxidant,
anti-inflammatory and modifier of glutamate signalling, hypothetically could prove effective
in ameliorating the pathways involved.
2. Aetiology of POCD
2.1. Surgery induced inflammation
Current theories regarding the mechanisms underlying cognitive dysfunction following
surgery highlight the role of inflammation and immune activation. Surgery itself induces
inflammatory processes; both localised and systemic expression of pro-inflammatory
cytokines and related signalling molecules (Kohl & Deutschman, 2006; Reikerås, 2010).
Tissue damage following surgery provokes the release of IL-1 and TNF-α from endothelial
cells and phagocytes, triggering a cascade (Mannick, Rodrick, & Lederer, 2001; Menger &
Vollmar, 2004) of downstream signalling events (E. Lin, Calvano, & Lowry, 2000). Elevated
levels of IL-1 and TNF α induce the later production of IL-6, whose levels correlate positively
with the extent of tissue trauma(Mannick, Rodrick, & Lederer, 2001; Menger & Vollmar,
2004).
Copious pre-clinical data from animal models supports the association of immune system
activation following surgery with the development of cognitive decline (Cibelli et al., 2010;
N. Terrando et al., 2010; Niccolò Terrando et al., 2010, 2011). Multiple, although relatively
small-scale, clinical studies have reported correlations between circulating peripheral
inflammatory markers and cognitive dysfunction after surgery. Li et al observed higher
serum levels of IL-6 in patients developing POCD (LI et al., 2012), and similarly, Hudetz et al
documented impaired memory in patients with high IL-6 or high C-reactive protein (CRP), a
marker of the acute-phase inflammatory response (Hudetz, Gandhi, Iqbal, Patterson, &
Pagel, 2010). The latter study also observed persistent decrements of cognitive function in
the high-IL-6 group 1 month after surgery. A meta-analysis of clinical data showed that
elevated peripheral inflammatory markers were indeed associated with POCD, and that this
relationship was strongest for IL-6 (L. Peng, Xu, & Ouyang, 2013). Further support for the
role of Il-β, TNF-α and other pro-inflammatory cytokines in the genesis of cognitive
dysfunction following surgery is provided by a number of animal experiments. These
preclinical studies revealed that surgical trauma is associated with prolonged activation of
macrophages with the capacity of maintaining elevated levels of cytokines and that those
levels correlate with the severity of cognitive dysfunction as measured in several domains,
especially in memory and learning (Cibelli et al., 2010; Rosczyk, Sparkman, & Johnson, 2008;
Niccolò Terrando et al., 2011).
Insights into the interaction between inflammation and cognitive function can also be drawn
from other inflammatory states. Associations between elevated circulating inflammatory
markers and cognitive dysfunction have been documented in numerous clinical populations,
such as in viral and bacterial infection (Cvejic, Lemon, Hickie, Lloyd, & Vollmer-Conna, 2014),
HIV infection (Yuan et al., 2015), following chemotherapy (Ganz et al., 2013), in sickle-cell
disease (Andreotti, King, Macy, Compas, & DeBaun, 2014), and in non-surgery-related
delirium (Adamis et al., 2014). The mechanism behind this association has been explored
further in experimental mice models of induced inflammation, which show cognitive deficits
in response to immune activation (Ming, Sawicki, & Bekar, 2015), or transgenic animal
models which overexpress cytokines, including IL-6, in the brain (Heyser, Masliah, Samimi,
Campbell, & Gold, 1997). IL-6, for example, increases with aging, in animal models and in
human plasma (Ye & Johnson, 1999), and has been associated with the cognitive decline
associated with aging (Godbout & Johnson, 2004). However, mice model studies of the
effects of IL-6 on learning and memory have shown that the effects of this cytokine are
complex and dependent on age, with both very low levels and excessive levels impairing
cognition (Braida et al., 2004; Heyser et al., 1997). These findings are reflective of the
requirement that inflammatory signalling be tightly controlled; baseline levels of signalling
are required for normal function, but dysregulated inflammation has pathological outcomes.
The signalling pathways initiating, controlling and fine-tuning inflammation are complex.
However certain key players have been identified which sit at nodes or intersections of
these pathways and have particularly strong associations with the cognitive dysfunction
associated with surgery. The high-mobility group box-1 chromatin protein (HMGB1) has a
pivotal role in inflammatory pathways. The post-surgical environment of elevated cytokines
and oxidative stress drives both the release and activation of HMGB1 in humans (Manganelli
et al., 2010;) and in mice (N. Terrando et al., 2010; Niccolò Terrando et al., 2010). HMGB1
then drives further inflammation, in a feed-forward mechanism which promotes further
activation of inflammatory signalling pathways and the production of damaging oxidative
molecules by inflammatory cells as demonstrated in both animal models (Lee et al., 2014);
and in humans (Tang et al., 2011). HGMB1 can also promote long-term inflammation via
promotion of immune cell proliferation and maturation (X.-M. Zhu et al., 2009).
The pivotal role of this cytokine as a driver of systemic inflammation is highlighted by clinical
evidence that levels of HMGB1 after surgery are predictive of the degree of post- surgery
inflammation and morbidity over the short and long term (Kornblit et al., 2008; Suda et al.,
2006; Takahata et al., 2011). Clinical and animal studies have demonstrated that elevations
of HGMB1 post-surgery are associated with poor performance on cognitive tasks in both
mice (R.-L. Li et al., 2013); and in humans (LIN, WANG, CHEN, HU, & OUYANG, 2014).
Critically, levels of this inflammatory molecule are mechanistically linked with the
development of POCD. In a landmark experiment, a single peripheral administration of
HMGB1 induced cognitive deficits in a mice model (Vacas, Degos, Tracey, & Maze, 2014).
Furthermore, administration of an antibody neutralising HMGB1 prevented the cognitive
decline associated with surgery and reduced the circulating inflammatory cytokine burden
(Vacas et al.). These findings emphasise the central role of this factor in regulating post-
surgical inflammatory responses and consequential cognitive deficits. Notably, in other
animal studies, elevated expression of HMGB1 was observed not only in the periphery, but
in the hippocampus, an area of the brain associated with memory functions (Li et al., Lin et
al.) Both surgery and anaesthesia induce elevations in HMGB1 expression in the
hippocampus in conjunction with the development of cognitive deficits (He et al., 2012).
This raises the important question of the mechanisms by which alterations in peripheral
inflammatory signals can impact on CNS function and therefore cognition.
2.2 CNS inflammation
The central nervous system was classically thought to be impervious to peripheral
inflammatory factors, due to the existence of the blood brain barrier (BBB) and paucity of
lymphatic drainage, although the recent discovery of a brain lymphatic system has
challenged this assumption (Iliff, Goldman, & Nedergaard, 2015; Louveau et al., 2015). The
BBB does tightly regulate the transit of inflammatory factors, cells and other substances into
the CNS, and also monitors the peripheral environment. However, peripheral inflammatory
states can compromise the integrity of the BBB, allowing for increased entry of
inflammatory factors. Surgery and anaesthesia both induce blood-brain barrier dysfunction
which may be driven by peripheral inflammatory factors such as HMGB1 but may also
encourage further CNS expression of inflammatory cytokines (He et al., 2012).
Considerable evidence from preclinical models demonstrates that elevated inflammatory
cytokines in the periphery are a well-documented contributor to neuroinflammation via
their capacity to disrupt blood brain barrier integrity (Cibelli et al., 2010; Terrando 2010a,
2010b, Poon, Ho, Chiu, Wong, & Chang, 2015). The activation and dysfunctional activity of
endothelial cells in the BBB in response to peripheral inflammatory stimuli is now
considered to be the initial event in the development of neuroinflammation (X. Zheng et al.,
2014). Pro-inflammatory cytokines can transverse the BBB directly using specific receptors
and transporters on the surface of BBB endothelial cells (Pan et al., 2011). These
inflammatory mediators can also gain access to the CNS via the circumventricular regions
where the BBB is either absent or discontinuous. Markers of BBB dysfunction are elevated
post-surgery, and are associated with both peripheral and central inflammation in animals
(He et al., 2012); and in humans (Bromander et al., 2012).
Increased levels of inflammatory markers in the CSF following surgery, including IL-6, TNFα
and HMGB1, have been repeatedly observed in both cardiac and non-cardiac surgery
(MacLullich et al., 2011; Reinsfelt et al., 2012; Tang et al., 2011; Woiciechowsky et al., 1997).
Importantly, such elevations have been directly associated with cognitive dysfunction in
humans (Barrientos, Frank, Watkins, & Maier, 2012;) and mouse models Cibelli et al., 2010;;
K. Ji, Akgul, Wollmuth, & Tsirka, 2013; K. Ji, Miyauchi, & Tsirka, 2013), suggestive of a link
between peripheral and central inflammation driving neurocognitive dysfunction. In the
CNS, inflammation is mediated by the action of glial cells, to which we turn.
2.3 Glial activation
Microglia are the resident immune cells of the CNS and have a lineage similar to peripheral
macrophages. Microglia have the capacity to modulate rearrangement of surface molecules
mediating cell-cell and cell-matrix interactions, as well as releasing multiple factors and
compounds with proinflammatory, immunoregulatory and oxidative effects (Kettenmann,
Hanisch, Noda, & Verkhratsky, 2011). Microglia express specialised receptors for
inflammatory signals such as bacterial endotoxin or endogenous ligands such as HGMB1, in
particular the toll-like receptor 4 (TLR-4) (I.-D. Kim & Lee, 2013; J.-B. Kim et al., 2006). The
elevation of peripheral inflammatory factors induced by surgery prompts microglial
morphology changes and upregulation of TLR-4 (F. Wang, 2014). Activated microglia
themselves produce cytokines, resulting in de novo production of a range of inflammatory
factors in areas of the CNS, in particular but not restricted to the hippocampus. These
factors include TNFα, IL-1, IL-6, oxidative species and other signalling molecules (Norden &
Godbout, 2013).
Astrocytes are numerically the dominant CNS glial cell. The astrocytic protein S100β (S100
calcium binding protein β) is highly abundant in the brain and serves key roles in multiple
homeostatic and damage/infection associated processes, through both autocrine and
paracrine signalling (Donato et al., 2013). Although this protein is expressed by some
selected cells in the periphery, peripheral circulating levels of S100β are a key potential
biomarker of CNS stress (Michetti et al., 2012; Sun & Feng, 2013).
At low levels, S100β acts as a neurotrophic or supportive factor (Arcuri, Bianchi, Brozzi, &
Donato, 2004). However, elevations of this protein can drive neurotoxic pathways, via a
number of mechanisms. S100β may induce direct neuronal damage (Fano’ et al., 1993; Sorci
et al., 2010), or activate both microglia and astrocytes (Petrova, Hu, & Van Eldik, 2000;
Bianchi et al., 2007; Hu et al., 1997) and induce microglial oxidative species production
(Adami, Bianchi, Pula, & Donato, 2004). When viewed as a whole there is considerable in
vitro evidence that high levels of S100B have a causative role in the development and
maintenance of neuroinflammation (Bianchi, Adami, Giambanco, & Donato, 2006; Bianchi,
Kastrisianaki, Giambanco, & Donato, 2011) operating through similar receptors to HMGB1
(Han, Kim, & Mook-Jung, 2011; R.-L. Li et al., 2013; Ryckman, Vandal, Rouleau, Talbot, &
Tessier, 2003).
Peripheral inflammation, microglial activation, and microglial cytokine production are
temporally and spatially correlated with the development of cognitive dysfunction in animal
models (Wang et al., 2012; and Wang, 2014). This finding has been corroborated in several
animal models of surgery-induced cognitive changes (Hovens et al., 2014, 2015).
Interestingly these cognitive deficits, hippocampal microgliosis and cytokine expression
associated with peripheral surgery could be ameliorated by administration of the non-
selective inflammatory inhibitor ibuprofen, as suggested by mice models (Xu et al., 2014).
Cognitive deficits after surgery could also be ameliorated by blockade of TNFα, a key player
in inflammatory cascades (Terrando et al., 2010a). This was associated with suppressed
microgliosis in the hippocampus and reduced peripheral and central inflammation in mice.
Similarly, experimentally-induced peripheral inflammation induced hippocampal microglial
activation was correlated with HMGB1 production and cognitive dysfunction, and this could
be prevented by treatment with inhibitors of IL-1 signalling (Terrando, 2010b).
Peripheral surgery is known to induce elevations of both CSF and serum S100β, and these
have been correlated with neuropathological processes in humans (Reinsfelt et al., 2012).
Multiple clinical studies have documented an association between serum S100β and
impaired cognitive function after various surgery types (Bayram et al., 2013; Wang, 2014)
and studies using narrower definitions of POCD have replicated this finding (LI et al., 2012;
Linstedt et al., 2002). Preclinical studies show that S100β, HGMB1 expression and astrocyte
morphological changes occur in concert with cognitive dysfunction after experimental
surgical procedures (Li et al., 2013). A recent meta-analysis examined the association
between S100β levels in the periphery and the development of POCD and confirmed that
elevated levels of S100β post-operatively were strongly associated with POCD in humans
(Peng, Xu, & Ouyang, 2013). These data further reinforce both the hypothesis that POCD
involves disordered CNS inflammatory signalling but also that peripheral levels of factors
such as S100β can be used as potential biomarkers of cognitive dysfunction associated with
inflammatory mechanisms.
2.4 Oxidative stress in the CNS
Production of reactive oxygen species is intrinsic to a variety of processes occurring during
surgery (e.g. tissue trauma and wound healing) in addition to being a consequence of the
inflammatory reaction peri- and post-surgery (Fu et al., 2014). Elevated levels of reactive
oxygen species and markers of oxidative damage, along with depleted antioxidants have
repeatedly been reported post-surgery in animal (Ko, Isoda, & Mobbs, 2015; Kotzampassi et
al., 2009); and human studies as summarised in the review byTsai et al., (2010).
Peripheral oxidative stress may induce or exacerbate neuroinflammation via dysregulation
of BBB integrity (Abdul-Muneer, Chandra, & Haorah, 2014). Damage to the BBB may also
induce release of cytokines and oxidative species into the brain. The activation of the
endothelium may also provoke microcirculatory dysfunction ultimately leading to
compromised cerebral perfusion (Taccone et al., 2010). However, inflammation in the CNS is
a major driver of the local production of oxidative species, primarily through microglial
activation. Activated microglia release reactive oxygen species and nitric oxide along with
pro-inflammatory cytokines, together providing a toxic milieu in which neuronal function is
compromised (Heneka, Kummer, & Latz, 2014). Activated microglia and neurones may
become entwined in a degenerative positive feedback loop, in which release of HMGB1
from activated microglia and damaged neurones promotes release of oxidative species from
microglia, further precipitating injury (H.-M. Gao et al., 2011). Redox homeostasis may also
modulate microglial function in an autocrine manner, potentiating feed-forward activation
of microglia in an environment of oxidative stress (Rojo et al., 2014).
The neuroinflammatory effect of surgery is emphasised by the observation that surgical
stress raises the levels of brain nicotinamide adenine dinucleotide phosphate (NADPH)
oxidase, a key player in oxidative stress regulation, and that this was correlated with
behavioural changes in mice (Zhang et al., 2015). Levels of oxidative species were also
positively correlated with the severity of cognitive impairment in patients with major
depressive disorder (Talarowska et al., 2011, 2012). Experimental models of POCD show
elevated markers of oxidative damage to lipids and proteins in the hippocampus (An et al.,
2012). Critically, the level of the key oxidative species nitric oxide following surgery is
predictive of the development of cognitive dysfunction in humans(Harmon et al., 2005;
Iohom et al., 2004).
2.5 Impact of inflammation and oxidative stress on neuronal integrity & signalling
Although multiple brain regions are involved in cognitive processes, the role of the
hippocampus in many of the processes underpinning learning and memory formation is
particularly well established (Epp, Chow, & Galea, 2013; K. Z. Tanaka et al., 2014).
This region of the brain also arguably contains the largest number of cytokine receptors
(Gemma, Fister, Hudson, & Bickford, 2005; Parnet et al., 1994). In particular, the
hippocampus contains the highest density of receptors for IL-1, and whilst physiological
levels of this cytokine are essential for optimal memory and learning processes, excessive
levels have been associated with diminished cognitive function in animal models (J. Chen et
al., 2008; Rachal Pugh, Fleshner, Watkins, Maier, & Rudy, 2001). The hippocampus appears
especially vulnerable to the detrimental effects of systemically elevated inflammatory
mediators such as HGMB1 due to the high density of the TNF-α receptor and numerous
other receptors on the surface of endothelial cells in this region of the brain (Rothwell,
Luheshi, & Toulmond, 1996). The mechanism underpinning impaired learning and
memory resulting from excessive levels of IL-1 and other inflammatory mediators in the
hippocampus involves detrimental effects on long term potentiation and neurogenesis in
this region of the brain, as has been demonstrated in various animal models (A. J.
Cunningham, Murray, O’Neill, Lynch, & O’Connor, 1996; Kelly et al., 2001; Vereker et al.,
2001).
Glutamate signalling is particularly important for the process of long term potentiation
thought to underpin aspects of memory formation. Experimentally-induced peripheral
inflammation fundamentally changes hippocampal neuronal activity, long-term potentiation
and synaptic plasticity via modulation of glutamate receptors in rats (Riazi et al., 2015).
Expression of the GluR2 subtype of glutamate receptor was decreased, leading to enhanced
AMPA and NMDA mediated signalling. Mice genetically susceptible to inflammatory
challenge through deficiency of a negative regulator of cytokine signalling show deficits in
multiple cognitive domains and have impaired hippocampal long-term potentiation through
pathways involving Il-1β, TNFα and HGMB1 (Costello et al., 2011). HMGB1 plays a critical
role in modulating glutamatergic neuronal activity underlying cognitive function. HMGB1
potentiates glutamate signalling through the NMDA receptor, and exacerbates glutamate-
induced toxicity associated with degenerative processes (Balosso, Liu, Bianchi, & Vezzani,
2014; Pedrazzi et al., 2012).
Suppression of microglial activity abolished the effect of peripheral inflammation on
hippocampal function, highlighting the key role of microglial-derived factors on the
modulation of glutamate signalling and neuronal function. These findings were echoed in a
similar study which demonstrated that induction of peripheral inflammation promotes Il-1β
production in the hippocampus and results in parallel decreases in glutamate release and
long-term potentiation in mice (Di Filippo et al., 2013). Further evidence for the key role of
glial inflammatory factors in the maintenance and modulation of hippocampal circuitry
comes from a mice model study focussing on TNF-α (Stellwagen & Malenka, 2006). The level
of this glial-derived pro-inflammatory cytokine was found to modulate the level of the
AMPA form of glutamate receptor and thus fine-tunes the sensitivity of neurones to
glutamate signalling. Inhibitory neurotransmission in the hippocampus is also compromised
by neuroinflammation. Acute application of TNFα induces a persistent decrease of gamma-
aminobutyric acid (GABA) receptors and a pervasive dysregulation of this important
component of synaptic plasticity regulation (Pribiag & Stellwagen, 2013).
Astrocytes, and to some extent microglia, also have the capacity to modulate availability of
the key neurotransmitter glutamate. This facilitates the balance required between
glutamate availability for signalling processes and prevention of the excessive levels
associated with excitotoxicity. The expression of specific glutamate transporters on the
astrocyte surface are regulated by cytokine stimulation, with the pro-inflammatory cytokine
TNFα upregulating glutamate transporters on cultured astrocytes (Dumont, Goursaud,
Desmet, & Hermans, 2014). Infiltration of activated T cells into the CNS provokes glutamate
release from microglia via a different subtype of glutamate transporter (Evonuk et al., 2015).
Together this suggests that altered glutamate availability, in addition to glutamate receptor
changes, may contribute to the changes in neuronal function and neuroplasticity observed
in response to inflammation.
Microglia and astrocytes also actively regulate synaptic function via modulation of synaptic
morphology and synaptogenesis ( Ji, Akgul, et al., 2013; K. Ji, Miyauchi, et al., 2013). Key to
this is the release of neurotrophic factors such as brain-derived neurotrophic factor (BDNF).
BDNF is now well established as a key player in multiple aspects of neural plasticity, learning
and memory in various animal models (Heldt, Stanek, Chhatwal, & Ressler, 2007; Lu,
Nagappan, & Lu, 2014). Inflammation and cytokine expression levels modulate the
production of BDNF (Derecki et al., 2010; Yirmiya & Goshen, 2011). Stimulation of the
immune response via administration of bacterial lipopolysaccharide induces hippocampal
cytokine production, depletion of BDNF and learning and memory deficits (S. Tanaka et al.,
2006). Intrahippocampal administration of Il-1β was shown to block the upregulation of
BDNF required for learning (Barrientos et al., 2004). Importantly, inflammatory processes
associated with surgery also result in decrements in BDNF. In an animal model of surgery,
rats showed cognitive deficits which were spatially and temporally correlated with
decreased BDNF, increased inflammatory cytokines in the periphery and brain, and
increased microglial activity (Hovens et al., 2014).
A key facet of microglial activity is their capacity to induce production and release of
reactive oxygen and nitrogen species, which are essential for the immune response to
infection but in the context of prolonged activation, have deleterious effects on brain
cytoarchitecture and neuronal function. Despite its high dependence on oxidative
metabolism, the brain has relatively low levels of antioxidants and is vulnerable to oxidative
stress. Microglia express 90% of brain glutathione, the main antioxidant species, which
highlights both the centrality of these cells to oxidative processes in the CNS and the relative
deficit of antioxidants in neurones (Rojo et al., 2014). In fact, circulating levels of glutathione
have been proposed to be a useful predictive biomarker of cognitive decline in patients with
neurodegenerative disease (Revel et al., 2015).
The role of inflammation-driven oxidative stress in neurodegenerative disease, and the
centrality of this oxidative stress to pathophysiology has been the subject of much
speculation. (Di Filippo, Chiasserini, Tozzi, Picconi, & Calabresi, 2010; Fischer & Maier, 2015).
For example, mice models have demonstrated that levels of the inflammatory cytokine TNF-
α, which approximate those in the post-stroke brain, induce striking loss of neuronal
integrity and function which was driven by dysfunction of mitochondria, the major source of
oxidant species (Doll, Rellick, Barr, Ren, & Simpkins, 2015). However, more subtle
dysregulation of inflammatory drive may also modulate the oxidative milieu of the CNS, and
in turn alter neuronal function without inducing cell death. Elevation of the pro-
inflammatory cytokine Il-1β in the CNS induces elevations of reactive oxygen species which
are coupled to disruption of long-term potentiation, an element of neuroplasticity which is
intrinsic to learning, memory and other aspects of cognition (Vereker et al., 2001). Oxidative
stress may also be considered as a significant contributor Oxidative stress is now seen as a
key contributor to the cognitive decline associated with aging, and supplementation with
anti-oxidants might ameliorate this process (Haxaire et al., 2012; Robillard, Gordon, Choi,
Christie, & MacVicar, 2011).
3. Potential contributions of anaesthetic agents to POCD
In the surgical setting, the question of whether anaesthetic agents are contributors to
cognitive dysfunction and underlying pathological processes is of particular importance.
These agents penetrate the CNS and have direct activity on neuronal signalling and activity.
It is therefore a question of some interest as to whether the changes in inflammatory and
oxidative stress parameters associated with cognitive dysfunction can be attributed to
surgical procedures and their systemic consequences as outlined above, anaesthetic use or,
perhaps most likely, a combination of the two.
Anaesthetics used in surgery are a heterogeneous group of agents with differing
pharmacological properties and mechanisms of action. In addition, studies of the effects of
anaesthesia in the clinical setting are impeded by the confounding effects of the surgery
itself. However, emerging evidence from pre-clinical studies suggests that anaesthetic use
is, in isolation, a risk for cognitive dysfunction. Studies in aged mice suggest that isoflurane
anaesthesia alone induced cognitive dysfunction, which was associated with microglial
activation and CNS inflammation (F. Wang, 2014). These results are echoed by studies
examining exposure to another volatile anaesthetic agent, sevoflurane, in which working
memory impairments and alterations in proteins associated with synaptic plasticity and
neuronal signalling were observed in rats (Ling, Ma, Yu, Zhang, & Liang, 2015). A recent
study comparing epidural anaesthesia and general anaesthesia in elderly patients
undergoing hip replacement demonstrated a higher risk of POCD in patients receiving
general anaesthesia, which was associated with elevations of circulating amyloidogenic
proteins (Shi et al., 2015). Similarly, poorer cognitive outcomes were observed following
general anaesthesia as compared to epidural anaesthesia in an earlier study of hip and knee
surgery (Pal et al., 2011). Anaesthesia depth may also impact upon cognitive outcomes
(Radtke et al., 2013; Short et al., 2015), with research suggesting lower bispectral index
scores are associated with greater incidence of post-operative delirium, but not cognitive
dysfunction in elderly surgical patients (Radtke et al).
Both clinical and pre-clinical evidence suggests that the elevations in inflammatory and
oxidative stress parameters observed following surgery may in part be attributed to the
effects of anaesthetic agents. When administered to aged animals, isoflurane induces
significant upregulation of a number of pro-inflammatory cytokines and decrements in
spatial memory skills (Li et al., 2013; Li et al., 2013). Conversely, very young mice were
shown to be more susceptible to inflammation, microglial activation and cognitive
dysfunction induced by general anaesthesia than adult (not aged) animals (Shen et al.,
2013). Use of the anti-inflammatory agent parecoxib mitigated some of the cognitive deficit
induced by general anaesthesia.
There is contrary evidence regarding the aetiological role of general anaesthesia, with
preclinical studies showing changes in S100B, HGMB1 expression and astrocyte morphology
in concert with cognitive dysfunction after experimental surgical procedures, but not in an
isoflurane anaesthesia control group (Li et al., 2013). The discrepancy between these studies
may relate to the pathways being investigated, or methodological differences in isoflurane
administration (i.e. duration of anaesthesia).
Evidence that the type of general anaesthesia impacts on outcomes post-surgery is
potentially suggestive of an aetiological role, though the nature and extent of this impact is
far from clearElderly patients undergoing major surgery were found to experience a higher
incidence of POCD following inhalational anaesthesia with sevoflurane than those who
received intravenous propofol (Qiao et al., 2015). Importantly, this was associated with
increased circulating levels of S100β and the proinflammatory cytokines TNFα and IL-6,
which could be ameliorated by adjunctive treatment with the anti-inflammatory
methylprednisolone. An aetiological role for anaesthesia in POCD was also evidenced in a
study where patients (N=921) were randomly assigned to receive either bispectral index
guided anaesthesia, associated with reduced anaesthetic exposure, or routine care. Guided
anaesthesia was associated with fewer patients with delirium (15.6% vs. 24.1%, P=0.01) and
a lower rate of POCD at 3 months (10.2% vs. 14.7%, p=0.025) (Chan, Cheng, Lee, & Gin,
2013).
Evidence from rat models also indicates that exposure to anaesthetics may also impair
mitochondrial function and potentiate oxidative damage to neurones (Sanchez et al., 2011).
Anaesthetic agents, particularly volatile anaesthetic agents targeting inhibitory GABA
receptors, directly modulate neuronal signalling. Differences in GABA receptor affinity may
underpin some of the variation between anaesthetic agents. Isoflurane, for example,
depresses long-term potentiation and shifts the ability of neuronal synapses to respond to
stimulation, which may contribute to cognitive impairment (Simon, Hapfelmeier, Kochs,
Zieglgänsberger, & Rammes, 2001). Prolonged elevation of GABA receptor activity induced
by general anaesthesia is associated with sustained impairments of cognitive function in
mouse models (Zurek et al., 2014). Importantly, blockade of the oxidative damage induced
by general anaesthetic agents by EUK-134 (a synthetic oxidative species scavenger) or
pramipexole appears to prevent the development of cognitive impairment in the rat brain
(Boscolo et al., 2012). Further, systemic inflammation can also modulate GABA receptor
function and contribute to cognitive deficits, as observed in mouse models (Wang et al.,
2012). This suggests a mechanism by which both surgery and anaesthesia-induced
inflammation may intersect with anaesthetic modulation of synaptic plasticity, a confluence
of factors contributing to cognitive dysfunction. In addition, findings that surgical insults
alone can induce significant inflammation and cognitive dysfunction in mice (Xu et al., 2014)
highlight the likelihood that surgery and anaesthesia interact in a complex and synergistic
manner to influence postoperative cognitive function.
4 Pre-existing risk factors for POCD
POCD is far from a universal consequence of surgery, while neuroinflammation following
even minor surgery appears to be a ubiquitous event. Hence while neuroinflammation is
likely an essential element underpinning the development of POCD it is clearly not the sole
driver. Other factors appear to be essential in the transition from ‘expected’ events post-
surgery and enhanced neuropathology seen in patients with POCD. Pre-existing conditions
or demographics may also act to decrease cognitive reserve; such as the presence of
physical illness, frailty, lower education levels, history of alcohol or opiate use, and the
presence of anticholinergic medication to risk post-operative cognitive change (Silbert et al.,
2007; 2015; Wang et al., 2014b) (B. Silbert et al., 2015; B. S. Silbert, Scott, Evered, Lewis, &
Maruff, 2007; F. Wang, 2014). Nadelson et al. argue that the preoperative cognitive
trajectory, the success of the surgery, and events in the perioperative period including
ongoing pain have been argued to be the most important determinant of POCD (2014) and
also make the point that cognitive decline will be observed in surgical patients if cognitive
function is measured pre- and post-surgery if the patient is already on a trajectory of
cognitive decline.
One confounding factor that may be a significant contributor to the development of POCD is
the presence of an abnormally potent and or prolonged neuroinflammatory response in a
patient with pre-existing inflammatory pathology. For example, post-stroke patients have a
significantly increased risk of developing POCD even without extensive neurological damage,
due to the underlying inflammation pre-existing as part of the stroke event (Monk et al.,
2008). Similarly, pre-existing metabolic syndrome is a recently confirmed risk factor in the
development of POCD in both cardiac and non-cardiac patients (Hudetz et al., 2010; Hudetz,
Patterson, Amole, Riley, & Pagel, 2011; Hudetz, Patterson, Iqbal, Gandhi, & Pagel, 2011a,
2011b) which is of interest considering the role of systemic inflammation and
neuroinflammation in the development and pathological consequences of the illness (Cai &
Liu, 2012; Purkayastha & Cai, 2013). Diabetes is also associated with an elevated
inflammatory burden. In a study of patients with type-2 diabetes undergoing carotid
endarterectomy, elevated pre-operative levels of C-reactive protein and monocytes were
predictive of a higher risk of cognitive dysfunction post-surgery (E. J. Heyer, Mergeche,
Bruce, & Connolly, 2013). These pathways have been further explored in animal models of
metabolic syndrome, in which mice selectively bred for poor metabolic function were more
susceptible to post-operative cognitive dysfunction (Su et al., 2012). This was associated
with an elevated inflammatory milieu as measured by several different parameters, which
appeared to be suggestive of a diminished capacity for resolving inflammation.
Advancing age is another well documented risk factor which is in itself noteworthy as levels
of circulating inflammatory markers generally increase with age. Neuroinflammatory
responses to peripheral stimuli also change with age, and this may be important for
cognition and memory (Barrientos et al., 2012; Dilger & Johnson, 2008). There is now
evidence that microglia become highly reactive or ‘primed’ with age, in a similar manner to
that observed in mild cognitive impairment or Alzheimer's disease ( Cunningham, 2012;
Norden & Godbout, 2013; Perry & Teeling, 2013). This primed state sets the scene for an
exaggerated neuroinflammatory response following activation, with increased intensity and
duration of microglial activation and cytokine production in the brain (Corona, Fenn, &
Godbout, 2011; Godbout, 2005; Ritzel et al., 2015; Wynne, Henry, & Godbout, 2009). The
link between these elevations in neuroinflammation and cognitive dysfunction has been
illustrated in several experiments in which either peripheral experimentally induced
inflammation or surgical procedures precipitate much more florid microgliosis and cytokine
expression, along with more pronounced cognitive changes, in old versus young mice (J.
Chen et al., 2008; Xu et al., 2014). Importantly, these effects were restricted to aged mice
receiving surgery under anaesthesia, and not to those receiving anaesthesia alone (Li et al.,
2013).
It is also of interest that there is a body of evidence suggesting that surgery accelerates the
development of Alzheimer’s disease (AD) or worsens its severity in some patients (Bohnen,
Warner, Kokmen, & Kurland, 1994; Fong et al., 2009; T. A. Lee, Wolozin, Weiss, & Bednar,
2005), although the latter finding is not universally replicated (Avidan et al., 2009) and
alternate biomarkers such as the ratio of amyloid-β to tau in cerebrospinal fluid have been
proposed (Xie et al., 2013). Fluctuations of perioperative amyloid-β to tau ratios are
predictive of verbal and visual memory dysfunction in AD (Xie et al., Ji et al., 2013) but this
hypothesis cannot be applied to POCD without additional high quality research. Meta-
analysis appears to show no association between exposure to general anaesthetics and the
risk of Alzheimer’s disease in humans, although there has been a paucity of high-quality
studies (Seitz, Shah, Herrmann, Beyene, & Siddiqui, 2011). Nonetheless, the presence of
mild cognitive impairment has been observed to be a possible risk factor for the
development of post-operative cognitive dysfunction, which in some studies has been
correlated with higher levels of inflammatory factors (Kazmierski et al., 2014; Oldham et al.,
2015; Veliz-Reissmüller, Torres, van der Linden, Lindblom, & Jönhagen, 2007). Interestingly
the link between POCD and dementia and mild cognitive impairment may be under-
reported as inclusion criteria for most studies are likely to exclude pre-existing diagnoses of
neurocognitive impairment, and many cohorts could be considered healthier than a
representative population of elderly surgical patients.
The presence of the ApoE allele has also been proposed as a risk factor for the development
of POCD (Ely et al., 2007; E. J. Heyer et al., 2005; Leung et al., 2007), which may relate to the
capacity for ApoE to regulate the intensity of microglial activation (Y. Liu et al., 2015)and
pathways governing the magnitude of inflammation and its resolution (Laskowitz et al.,
2001). However, the association between ApoE genotype and postoperative cognitive
changes has not been supported by all studies (Abildstrom, Christiansen, Siersma, &
Rasmussen, 2004; B. S. Silbert, Evered, Scott, & Cowie, 2008; van Munster, Korevaar, de
Rooij, Levi, & Zwinderman, 2007; Vasunilashorn et al., 2015), which is perhaps indicative of
the complexity of genetic risk factors for both dementia and cognitive decline as a whole.
Together these findings highlight the multilayered interactions between brain inflammatory
signalling, glial activity, glutamate synaptic morphology and activity, long term potentiation
and processes involved in cognition. Even in the non-injured or inflamed state, cytokines
and neurotrophic factors have the capacity to modulate these processes, (Di Filippo et al.,
2008; Yirmiya & Goshen, 2011) and therefore in states where the inflammatory drive is
heightened, including the perioperative setting, there is increased potential for disruption of
neuronal function and therefore cognitive dysfunction.
Figure 1 - Flowchart of the neuroinflammatory hypothesis of POCD
[INSERT FIGURE 1 ABOUT HERE]
The neuroinflammatory hypothesis of POCD provides a framework for potential intervention
agents. The flowchart in figure 1 provides a descriptive mechanism for how peripheral
trauma can feed forward into a cascade resulting in cognitive dysfunction, and each step is
an opportunity to mitigate this potentiality. To date, neither preventative nor intervening
agent has been established though a number of therapies have been investigated. A suitable
effective treatment could also ameliorate the longitudinal sequelae of POCD.
Novel therapies for POCD
5.1 Parecoxib/COX-II Inhibitors
Cyclooxygenase (COX) inhibitors, such as parecoxib and celecoxib, are a subtype of non-
steroidal anti-inflammatory drugs (NSAIDs) routinely used as an intervention for
inflammatory pain, including that experienced post-surgery (Derry & Moore, 2013). NSAIDs
inhibit the production of the enzyme COX that acts as a precursor to vasodilator
prostaglandins. Selective COX-II inhibitors are a class of NSAIDs that specifically bind to the
inflammatory COX-II isoform, and by interrupting the enzymatic conversion cascade
secondary to COX-II production, ameliorate inflammatory pain (Mattia & Coluzzi, 2005). For
this reason, a recent Cochrane review of clinical trials has suggested that COX-II inhibitors
have demonstrated efficacy in the treatment and management of inflammatory arthritis
and associated conditions (Kroon et al., 2015).
The mechanism of COX inhibitors also reduces the activation of microglia and subsequent
neuroinflammation (Choi, Aid, & Bosetti, 2009), and COX inhibitors are therefore routinely
the subject of translational research into conditions associated with neuroinflammation,
albeit with mixed results. A Cochrane review and meta-analysis for the use of COX-II
inhibitors for Alzheimer’s disease found no evidence to suggest any significant positive
effect, statistically or clinically (Jaturapatporn, Isaac, McCleery, & Tabet, 2012) a finding
which has recently been replicated in a large scale RCT (ADAPT-FS research group, 2015).
Similarly, evidence for the use COX-II inhibitors as an adjunct or monotherapy for
depression is extremely heterogeneous – studies showing significant improvement are
roughly approximate in weight to studies showing no effect (Eyre, Air, Proctor, Rositano, &
Baune, 2015). Meta-analysis showed that the use of celecoxib was efficacious in reducing
the severity of positive and negative symptoms of schizophrenia, but only for first-episode
schizophrenia participants as opposed to chronic participants (W. Zheng et al., 2017).
However, beyond symptomology associated with psychiatric disorders, cognitive function
appears to be a rarely examined outcome of COX-II inhibitor use. Some preclinical evidence
suggests a cognitive protection effect in rats after traumatic brain injury, but controversy
remains regarding the superiority of selective interventions for COX-I (Cernak, O’Connor, &
Vink, 2002) or II (Girgis et al., 2013) over placebo. In sum, there appears to be little support
for the use of COX-II inhibitors to treat chronic conditions, but research into acute
inflammatory states may hold some promise.
In humans, COX-II inhibitors are commonly used to treat post-operative pain, but the effect
for post-operative cognition is rarely directed examined in research. As such, even though
the existent evidence appears quite positive for the cognitive protection effect of COX-II
inhibitors, the amount of evidence is quite small. For example, in a large (n = 1062) double-
blind randomized control trial, Langford et al., (2009) demonstrated that the use of
perioperative parecoxib and valdecoxib was associated with significantly lower rates of
opioid-related confusion and concentration difficulties after major non-cardiac surgery, but
this appeared to equalize at approximately four days post-surgery. In addition, confusion
and concentration difficulties were approximated using patient self-report rating scales for
descriptions of severity and duration rather than objective measures, which complicate
comparisons with other studies.
By comparison, Y.-Z. Zhu, Yao, Zhang, Xu, and Wang, (2016) compared post-operative
cognitive performance on 122 elderly patients undergoing total knee arthroplasty at 7 days,
and 3 months post-surgery, using a battery of neuropsychological tests across a range of
cognitive domains. Participants were randomised to receive additional parecoxib or saline
after administration of general anaesthesia, and while cognitive performance at POD7 was
significantly better in the parecoxib group compared with controls, though there was no
difference at 3 months. Crucially, this convergence in performance appears to be driven by
an improvement in performance among the control group at 3 months, rather than a
delayed decline for the parecoxib group. To date, only one new study appears to be
examining the cognitive effect of NSAIDs after surgery (NCT02689024), and appears to be
focussed mainly on delirium incidence.
5.2 Statins
Statins are used to treat high cholesterol in humans. As the product of an enzymatic
reaction cascade, cholesterol production is in part rate-limited by the production of
mevalonic acid, which is itself rate-limited by conversion of the enzyme 3-hydroxy-3-
methylglutaryl-CoA reductase (HMGCR). Statins work by inhibiting conversion of HMGCR
into mevalonic acid through competitive binding of the active site of the enzyme (Friesen &
Rodwell, 2004). It is by intervention into this melavonic pathway that statins are also
purported to moderate inflammation. NADPH reduces HMGCR into mevalonic acid as part
of cholesterol production, and is associated with the production of phagocytic ROS. Statins
inhibit the HMGCR conversion process resulting in downregulation of NADPH production,
and the subsequent production of oxidative species (Margaritis, Channon, & Antoniades,
2014).
Statins have been widely researched as potential interventions for myriad of conditions,
particularly those with cognitive sequelae. In 2012 the FDA warned of potential cognitive
dysfunction after use of statins, though more recent safety reviews have found little
evidence to support this, and conversely suggest a potential cognitive protective effect in
some conditions (Mospan, 2016; Ott et al., 2015). However, the evidence for the use of
statins for cognitive change associated with neurological conditions is promising but
ultimately controversial. For example, a recent Cochrane review reported that statins do not
appear to prevent or treat dementia (McGuinness, Craig, Bullock, & Passmore, 2016).
However, McGuiness and colleagues limited included articles to only RCTs with a sample
aged over 65 years where a statin was administered for at least 12 months, which limited
the available evidence to two (albeit large) RCTs. By contrast, an earlier systematic review
and meta-analysis examining the use of statins for both short and long term cognitive
function reported a 29% reduction in long-term dementia incidence compared with controls
(Swiger, Manalac, Blumenthal, Blaha, & Martin, 2013). However, the review by Swiger and
colleagues allowed the inclusion of less controlled studies and included considerable
heterogeneity across administration regimen, dosage, and duration, as well study design
and sample characteristics. The clinical efficacy for the use of statins to treat neurological
conditions appears to remain theoretical in the absence of more definitive evidence
(Malfitano et al., 2014).
Researching examining the use of statins to moderate acute neuroinflammatory states is
promising but appears relatively undeveloped as a field. For example, in a systematic review
of preclinical data, W. Peng et al., (2014) suggested that statins appeared to be particularly
effective in protecting memory function after closed-head TBI in rodent models, though
little cognitive data is available from clinical studies and so the translational potential has
not yet been realised (Wible & Laskowitz, 2010). Similarly, the post-operative cognitive
impact of pre and perioperative statins remains largely unexplored, with only a handful of
non-randomised trials published. The evidence that is available is inconsistent – two
retrospective analyses and one smaller non-randomised prospective study reported
conflicting findings. For example, Mathew et al., (2005) observed no apparent impact of
perioperative statins on cognitive function 6 weeks post-operation with 440 elderly
cardiopulmonary bypass patients, while a post-hoc analysis of 585 carotid endarterectomy
patients revealed that the use of perioperative statins ameliorated morbidity associated
with post-operative cognitive dysfunction to the same degree as patients who did not
experience post-operative cognitive dysfunction (Heyer, Mergeche, Wang, Gaudet, &
Connolly, 2015). However, due to the retrospective nature of these studies the impact of
confounding variables cannot be estimated. By comparison, in a smaller non-randomised
control trial Das, Nanda, Bisoi, and Wadhawan (2016) compared the cognitive protection
effect of statins after off-pump CABG. Both the statin and control group demonstrated
significant memory dysfunction at 6 days post-operation in comparison with pre-surgical
performance, but the severity and breadth of dysfunction was significantly greater for the
control group. However, the absence of longitudinal follow-ups does not allow for direct
comparison with the earlier retrospective analyses.
5.3 Pregabalin
Initially developed as an anticonvulsant, pregabalin is being used in the treatment of
neuropathic pain and fibromyalgia (Toth, 2013). The antiepileptic mechanism for pregabalin
is thought to involve binding to voltage-gated calcium channels on the presynaptic
membrane, specifically the α2δ subunit (Schulze-Bonhage, 2012). This effectively limits the
rate of proconvulsant neurotransmitter release, and subsequently restricts seizures.
Pregabalin has demonstrated efficacy in reversing induced inflammatory hyperalgesia (Patel
et al., 2001) and suppresses the edema, neuronal damage, bleeding, and inflammation
resultant from traumatic brain injury in rats (Calikoglu et al., 2015). Clinical studies have
shown efficacy in modulating pain associated with multiple sclerosis; post-herpetic
neuralgia, peripheral nerve injury, radiculopathy, spinal cord injury, neuropathy, and post-
stroke pain (Dunn, Durieux, & Nemergut, 2016, p. 201; Moore, Wiffen, Derry, & McQuay,
2009).
Modulation of voltage-gated channels interferes with the activation of cytokine receptors,
and may subsequently contribute to neurotoxicity and neuron hyper excitability (Pexton,
Moeller-Bertram, Schilling, & Wallace, 2011; Vezzani & Viviani, 2015). The α2δ subunit
voltage-gated calcium channel is prominently expressed in the hippocampus (Cole et al.,
2005). By binding to the α2δ subunit, it has been proposed that pregabalin alters the release
of neurotransmitters in the hippocampus, moderating microglia activation (Kawano et al.,
2016), and downregulating subsequent cytokine production. Kawano and colleagues
demonstrated that pre-operative pregabalin reduced hippocampal production of IL-1β and
TNF-α and was protective of cognition in a rat model, while post-operative administration
was ineffective on these outcomes. In a model of diabetic mice, 10mg/kg pregabalin
significantly reduced levels of brain COX-II, glucose-transport type 4, cytosolic prostaglandin
E synthase, and nuclear factor-κB expression compared with controls, impair but did not
improve measures of learning (Sałat et al., 2016).
In a rat model of cardiopulmonary bypass with deep hypothermic circulatory arrest (Shim,
Ma, Zhang, Podgoreanu, & Mackensen, 2014), acute pre-operative administration of
pregabalin (30mg/kg) preserved some indices of neurologic function compared with saline,
though this did not translate to significant between-groups differences on the water-maze
test of long-term spatial memory. Percentages of necrotic neurones were significantly
reduced within the cerebral cortex for the pregabalin group, though no significant effect
was detected within the hippocampus. In sum, pre-clinical evidence suggests a
neuroprotective dose-response effect for pregabalin in inflammatory models, but the
impact upon cognition is less clear.
In humans, evidence for cognitive preservation is less well established. On the contrary,
acute cognitive change is an established risk of pregabalin use (Zaccara, Gangemi, Perucca,
& Specchio, 2011). In their meta-analysis, Zaccara and colleagues identified a dose-
dependent relationship between pregabalin administration in clinical populations and acute
confusional state and disturbance of attention, with positive correlation between dose and
risk. However, despite the large number of clinical studies examining psychiatric and
cognitive symptoms associated with pregabalin administration, comparatively few studies
directly examine and measure change within specific cognitive domains.
Evaluation of the cognitive impact of pregabalin is complicated due to the heterogeneity of
dose regimen across studies, though it is possible to observe a dose-response effect for
cognition. For example, Hindmarch, Trick, and Ridout (2005) found mixed evidence of
cognitive change after administration of 150mg TID for three days in a small sample of
healthy adults. Small but significant dysfunction in attention, CNS arousal and subjective
arousal was observed, while no significant changes to reaction speed, visual processing, and
memory were detected. Comparatively, significant dysfunction to executive function and
visual processing in healthy adults was associated with a larger dose of 600mg per day for 8
weeks (Salinsky, Storzbach, & Munoz, 2010). By contrast, Myhre, Diep, and Stubhaug (2016)
found no evidence of cognitive change in healthy adults after two 150mg doses of
pregabalin, at 13 hours and 1 hour prior to testing respectively. However, the combination
of remifentanil and pregabalin adversely impacted performance on cognitive tests beyond
that of either drug in isolation, suggesting an interaction effect of pregabalin with other
medications.
Despite some promising pre-clinical data, the cognitive impairment witnessed in the clinical
setting suggests it would not be a prime candidate for targeted exploration as an agent to
protect against POCD. Given its increasing use in the perioperative setting for the treatment
of acute neuropathic pain and in the prevention of chronic post-surgical pain, it would be
useful to analyse any influence it does have on early and late cognitive function within
studies examining POCD.
5.4 Dexmedetomidine
Dexmedetomidine is a α2 receptor agonist, used as a sedative agent with analgesic
properties (Dunn et al., 2016). In comparison with other α2 receptor agonists,
dexmedetomidine is highly selective for the locus coeruleus in the brainstem, suppressing
release of noradrenaline and adrenaline resulting in sedation purported to be akin to
natural sleep (Giovannitti, Thoms, & Crawford, 2015; Nelson et al., 2003). The modulation of
spinal noradrenaline and adrenaline is purported to be responsible for the analgesic
properties of dexmedetomidine (Coursin, Coursin, & Maccioli, 2001; Funai et al., 2014;
Yaksh, 1985). In vitro evidence suggests that dexmedetomidine mediates levels of
inflammation in human whole blood (ex vivo) through the moderation of inflammatory
cytokines induced by administration of lipopolysaccharide (Kawasaki et al., 2013), while
administration of dexmedetomidine in rats was associated with increased vagal nerve
activity and inhibited cytokine production (S.-T. Li & Wu, 2015; Xiang, Hu, Li, & Li, 2014).
Pre-clinical evidence suggests that dexmedetomidine is effective in protecting against
anaesthetic-induced cognitive impairment in a dose-dependent manner, though this is not a
universal finding. In a model of neonatal rats exposed to isoflurane 0.75% for 6 hours,
higher doses of dexmedetomidine (1, 10, and 25µg/kg respectively) were associated with
greater levels of neuroprotection compared with controls, with a specific reduction in
caspase-3 induced apoptosis in the hippocampus, thalamus, and cortex (Sanders et al.,
2009). High dose dexmedetomidine (25µg/kg) also preserved learning at 40 days post
isoflurane exposure controlsrelative to isofluorane alone, and reversed neurotoxicity to pre-
isoflurane exposure baseline levels in the hippocampus and thalamus, but not in the cortex.
In a similar investigation, Qian et al., (2015) reported that the pre-operative administration
of 25 µg/kg but not 15 µg/kg of dexmedetomidine was sufficient to protect against cognitive
dysfunction induced by splenectomy accompanied by 1.5-2.0% isoflurane and 100mg/kg
pre-operative ketamine in a rat model. By comparison, both 15 and 25 µg/kg
dexmedetomidine significantly reduced Il-β and TNF-α mRNA and protein expression
relative to controls. Qian and colleagues were also able to demonstrate that
dexmedetomidine administration reduced hippocampal levels of the pro-apoptotic genes
caspase-3 and BAX to attenuate neuronal apoptosis induced by anaesthesia and surgery
trauma. Hippocampal protection of 10 µg/kg dexmedetomidine after laparotomy in rats, in
comparison to the control group has also been observed in conjunction with preserved
learning and memory function (Wan, Xu, & Bo, 2014).
However, the dose-response effect of dexmedetomidine and neurotoxicity secondary to
inflammation may not adhere to a normal distribution. Sato, Kimura, Nishikawa, Tobe, and
Masaki (2010) demonstrated that 100µg/kg of intraperitoneal dexmedetomidine alone was
insufficient to protect against neurologic performance impairment 3 days after cerebral
ischemia in rats, despite a statistically significant protective effect upon neuronal survival in
the hippocampus compared with controls. However, the combination of dexmedetomidine
and induced hypothermia after ischemia (body temperature cooled to a mean of 35°C) not
only preserved hippocampal neuron survival, but also preserved neurological function
scores 3 days after ischemia. Dexmedetomidine alone was also ineffective for reducing
caudate nucleus apoptosis, and only when combined with hypothermic treatment was
caudate nucleus survival preserved.
Meta-analyses by B. Li, Wang, Wu, and Gao, (2015) and Man, Guo, Cao, and Mi, (2015) show
that dexmedetomidine was superior to controls and comparative anaesthesia groups for
reducing the risk of post-operative delirium in ICU patients, across a broad range of research
methodologies and dose regimens. Additionally, the majority of included studies only
included assessments of global cognition such as the Mini-Mental State Exam or the ICU-
Confusion Assessment Method, potentially obscuring cognitive change in specific domains.
For example, clinical evidence has shown that intra-operative administration of
dexmedetomidine has shown promise in significantly reducing acute episodic delirium after
cholecystectomy (Y. Li, He, Chen, & Qu, 2015) and carotid endarterectomy (Ge et al., 2016),
with approximately comparable outcomes despite dexmedetomidine dose disparity; 1mg/kg
for ten minutes followed by 0.4mg/kg/hour for the duration of surgery in Li and colleagues,
and 0.3 µg/kg for ten minutes before anaesthesia followed by a continuous 0.3 µg/kg/hour
until 30 minutes before surgery conclusion in Ge and colleagues. Both studies demonstrated
marked reduction in inflammatory biomarkers; IL-1β, IL-6 and CRP (Li et al.,) and in S100B
and the concentration of malondialdehyde but not neuron-specific enolase (Ge et al.).
Interestingly, in both studies the dexmedetomidine group demonstrated significantly
reduced rates of cognitive impairment at around 6 to 24 hours after surgery compared with
the control groups, in which cognition tended to recover and approximate the
dexmedetomidine groups at approximately days 3 to 4. In sum, the available evidence
suggests that dexmedetomidine is a potentially effective protective agent for cognition in
the early post-operative period, though the longitudinal impact of this protection is not yet
widely examined.
5.5 Lidocaine
Lidocaine works by targeting specific voltage-gated sodium channels associated with
neuropathic and inflammatory pain (Lauretti, 2008; Oliveira, Issy, & Sakata, 2010). Lidocaine
allosterically couples with the channel to stabilise it in an inactive state (Amir et al., 2006;
Sheets & Hanck, 2003), effectively halting the transmission of action potentials from the
administration site.
Due to capacity of lidocaine to suppress pain, it has been proposed that this agent may
likewise reduce inflammatory responses. The majority of available research supports this
hypothesis, with in vivo and in vitro evidence demonstrating a significant reduction in
biological correlates of inflammation such as IL-6, & IL-8 in rabbits; TNF-α and IL-1β in mice
and rats, and levels of IL-8 and CRP in humans after surgery (Caracas, Maciel, Martins, de
Souza, & Maia, 2009; van der Wal et al., 2015). Reduction in biological correlates of
inflammation in preclinical models has also translated to cognitive protection, including the
attenuation of memory impairment after intraoperative lidocaine administration (Tan, Cao,
Zhang, & Zuo, 2014) increased neuronal survival and cognitive protection after transient
global ischaemia (Popp et al., 2011); and the attenuation of memory and learning sequelae
2-4 weeks after isoflurane exposure (D. Lin et al., 2012); all in rat models, and all with
comparable lidocaine administration (1.5-2 mg/kg during the perioperative period).
Clinical evidence for post-operative cognitive protection due to lidocaine is mixed (Bilotta et
al., 2013). The majority of studies examine elderly cardiac patients, and typically use 1-1.5
mg/kg loading lidocaine doses immediately prior to or at the commencement of surgery,
followed by maintenance until surgical conclusion (Bilotta et al.). Lidocaine administration in
this fashion appears to provide neuroprotection early in the post-operative period (Mitchell
et al., 2009; D. Wang et al., 2002), but does not appear to be associated with longer-term
outcomes (Mitchell et al.; Mathew et al., 2009). For example, cardiac artery bypass surgical
patients were administered 1.5 mg/kg lidocaine immediately after surgical incision, and
maintained at 4mg/min for the duration of the procedure and post-operative cognitive
outcomes were compared with a control group that was given saline. Both groups
experienced significant levels of cognitive dysfunction at 9 days after surgery with significant
decreases to executive function, processing speed, short-term memory, and visuospatial
motor function compared with pre-surgical performance; while working memory and verbal
associative learning were preserved in the lidocaine group (D. Wang et al.).
Lidocaine for neuroprotection in non-cardiac patients is less extensively examined, and may
be similarly controversial. In an examination of urologic and orthopaedic patients, Heidari,
Rahavi, and Hashemi, (2013) found no difference in post-operative cognition as measured
by the MMSE in the first 24 hours between a 1.5 mg/kg lidocaine bolus-and-maintenance
treated group compared with saline controls; while K. Chen, Wei, Zheng, Zhou, and Li,
(2015) found that intravenous 1 mg/kg bolus followed by 1.5 mg/kg maintenance lidocaine
was neuroprotective in elderly spinal surgery patients. At 3 days after surgery, the lidocaine
group had significantly lower levels of S100B, malonaldehyde, and NSE compared with
controls, while the levels of TNFα were unchanged in either group. Post-operative cognition
at day 3 was showed a small but significant decline in the control group compared with
baseline function, whereas cognition in the lidocaine group was preserved. More recently,
Y. Peng et al., (2016) demonstrated that perioperative lidocaine did not confer greater
cognitive protection from neurological compromise in patients following supratentorial
tumour removal at any stage up to 6 months after surgery, compared with saline controls,
though brain oxygen consumption and metabolism was significantly lower throughout
surgery in the lidocaine group and this is a neurologically compromised group.
5.6 Ketamine
Ketamine is a common anaesthetic agent with hypnotic, analgesic, and amnesiac properties,
shown to be an NMDA receptor antagonist (M. Gao, Rejaei, & Liu, 2016). At sub-anaesthetic
doses, ketamine can be used to manage chronic neuropathic pain (Zgaia et al., 2015).
Ketamine has been shown to moderate inflammatory macrophage activation and
production of cytokines including IL-1β, IL-6 and TNF-α in both preclinical and clinical models
(De Kock, Loix, & Lavand’homme, 2013; F.-L. Liu, Chen, & Chen, 2012). As a glutamate
antagonist, ketamine has also been purported to potentially ameliorate depressive
symptoms associated with excessive glutamate levels, as observed in a number of clinical
trials (DeWilde, Levitch, Murrough, Mathew, & Iosifescu, 2015). Subsequent upregulation of
BDNF and mammalian target of rapamycin (mTOR) in the prefrontal cortex after ketamine
administration may promote synaptogenesis in preclinical models, along with cognitive
protective effects (Y. Lee et al., 2016). X. Zhu et al., (2015) demonstrated that administration
of 100mg/kg ketamine reversed upregulation of IL-β, TNF-α and amyloid- β after
electroshock therapy in a rat model, and attenuated associated memory dysfunction.
Reversing the upregulation of amyloids suggests that ketamine may show therapeutic
promise for management of neurodegenerative disorders. A study by Murman et al., (1997),
was able to demonstrate in a small pilot that patients with Huntington’s disease did not
appear to be more susceptible to cognitive change due to ketamine infusion compared with
controls, but this area is not well explored.
Few human studies have directly examined perioperative ketamine administration for
protection of post-operative cognition. Mortero et al., (2001) demonstrated that a
combination of propofol and ketamine (mean dose: 3.7 µg/kg) significantly improved post-
operative cognition in a non-cardiac sample in comparison to propofol alone. Hudetz et al.,
(2009) and Nagels et al., (2004) both examined elderly cardiac surgery patients, but with
differing ketamine regimens and inconsistent cognitive effects. Nagel and colleagues
administered a 2.5mg/kg bolus prior to surgical commencement and maintained this
infusion, whereas participants in Hudetz and colleagues were administered a single
0.5mg/kg bolus at induction. Hudetz and colleagues were able to demonstrate a significant
protective cognitive effect at one-week post-surgery, while Nagels and colleagues found no
effect at either one or ten weeks with the exception of improved performance on a single
outcome for the ketamine group.
5.7 Minocycline
Minocycline is a tetracycline antibiotic typically used in the treatment of acne, and has
shown anti-inflammatory and microglial inhibitory properties (Dean, Data-Franco,
Giorlando, & Berk, 2012; Garner et al., 2012). Minocycline is highly fat soluble and is able to
easily cross the blood-brain barrier (Garrido-Mesa, Zarzuelo, & Gálvez, 2013). Substantial
evidence from animal models demonstrates the potential of minocycline as an adjunct
therapy in managing neurodegenerative disease (Garrido-Messa et al.). In vitro evidence
suggests that the therapeutic mechanism of minocycline relates to its anti-inflammatory and
microglial inhibitory effects (Karachitos, Solis Garcia del Pozo, W.J. de Groot, Kmita, &
Jordan, 2012; H.-S. Kim & Suh, 2009). Specifically, it has been proposed that minocycline to
reduce the excitoxicity resultant from increased production of glutamate, mediating the
subsequent production of mitochondrial ROS and associated neurodegeneration (Karachitos
et al.). The potential for pro-cognitive effects of minocycline is further evidenced by the
capacity of this drug to regulate long-term potentiation and glutamatergic function via
modulation of microglial inflammatory cytokine production (S.-Y. Li et al., 2013; Riazi et al.,
2015).
Clinical evidence for minocycline as a pro-cognitive agent is promising but mixed, with few
placebo-controlled trials to date. Grieco et al., (2014) demonstrated a significant cognitive
improvement in children with Angelman syndrome through open-label administration of
minocycline, with medium-to-large effects upon communication and language
comprehension, at doses of 3mg/kg BID for 16 weeks; positive effects of comparable size
were observed in an open-label trial of minocycline for Fragile X syndrome (Paribello et al.,
2010). Open-label adjuvant administration of 100mg of minocycline each day for 2 years
appeared to significantly delay the progression of cognitive and motor decline in patients
with Huntington’s disease (Bonelli, et al., 2004). Evidence from placebo-controlled trials is
less encouraging; Nakasujja et al., (2012) and Sacktor et al., (2011) both found that the
administration of 100mg of minocycline orally every 12 hours for 24 weeks did not
attenuate cognitive dysfunction associated with HIV in humans when compared with
placebo, and later examination of the biomarkers of oxidative stress revealed limited
positive change in minocycline patients in comparison with placebos (Sacktor et al., 2014).
Interestingly, minocycline may actually promote CNS apoptosis in the brain of adolescents
and children via exacerbation of synaptic pruning (Inta, Lang, Borgwardt, Meyer-Lindenberg,
& Gass, 2016; Na, Jung, & Kim, 2014); suggesting that the effects of minocycline in the
developing brain may differ from those observed in the adult.
Given the putative neuroinflammatory hypothesis of POCD and similarity with other
conditions associated with neuroinflammation, it is possible that minocycline may have
potential for the prevention and treatment of POCD. Pre-clinical evidence suggests that
minocycline is effective in reducing the severity of post-operative cognitive dysfunction,
purportedly via anti-inflammatory mechanisms. For example, H.-L. Wang, Liu, Xue, Liao, and
Fang, (2016) examined minocycline for reducing post-operative cognitive dysfunction in
aged mice, with mice separated into control, minocycline, isoflurane, minocycline +
isoflurane, surgery, or minocycline + surgery. All minocycline was administered at 45 mg/kg
12 hours before surgery or isoflurane exposure, with isoflurane at 1.4% concentration in
50% oxygen for 2 hours. Mice who received isoflurane or surgery without minocycline
demonstrated significantly increased levels of IL-β, TNF-α, and TNF-γ as well as significantly
impaired processing speed and learning compared with controls. Pre-treatment with
minocycline attenuated these differences substantially. The findings of Wang and
colleagues support earlier findings documenting the capacity of minocycline to alleviate
surgery and anaesthesia-induced cognitive deficits via anti-inflammatory mechanisms in
rodent models (Jin et al., 2013; Kong et al., 2013; S.-Y. Li et al., 2013; Tian, Guo, Wu, Ma, &
Zhao, 2015). Similarly, in transgenic models of Alzheimer’s disease (Biscaro, Lindvall, Tesco,
Ekdahl, & Nitsch, 2012) and animal models of amyloid angiopathy-related cognitive
dysfunction (Fan et al., 2007), minocycline demonstrated significant pro-cognitive effects
which were correlated with anti-inflammatory rather than anti-amyloid effects. Minocycline
has also been shown to alleviate age-related cognitive deficits via modulation of microglial
activation in aged and adult mice (Kohman, Bhattacharya, Kilby, Bucko, & Rhodes, 2013).
The protective effects of minocycline appear to be dose-dependent in mice; Z. Li et al.,
(2016) found significant post-operative spinal microglia and macrophage inhibition with
0.05mg/kg minocycline for 14 days, but not within the prefrontal cortex.
5.8 N-acetylcysteine
N-acetylcysteine (NAC) is a widely available supplement, and is FDA approved for a number
of indications, including paracetamol overdose. Additionally, NAC is used to assist in the
management of bronchitis and COPD. NAC is a bioavailable precursor for glutathione, and
acts to reduce systemic markers of inflammation and suppress mitochondrial dysfunction.
NAC thus appears to target many of the cellular processes implicated in cognitive
dysfunction. Substantive evidence from multiple pre-clinical pathological models indicates
that NAC exerts protective effects via anti-inflammatory activity. In vitro evidence has
shown that NAC has the capacity to downregulate pro-inflammatory cytokine production
and downstream signalling pathways, including critical pro-inflammatory cytokines such as
IL-6 and HGMB-1 (Gabryel, Bielecka, Bernacki, Łabuzek, & Herma, 2011) in multiple tissue
types (Y.-H. Lee et al., 2014b; Usatyuk et al., 2003; C. Wang et al., 2015). NAC also
suppresses the transcription and activity of TLR-4 and similar key receptors in the
inflammatory system (Hou et al., 2012). Interestingly, evidence from clinical studies suggests
that NAC also increases the production of anti-inflammatory cytokines, notably IL-10, IL-4
and IL-1 (Santiago et al., 2008). Strong preclinical evidence from rat models also
demonstrates that NAC reduces the reactivity of microglia (Amar Karalija, Novikova,
Kingham, Wiberg, & Novikov, 2012), their level of activation (Berman et al., 2010) and the
levels of cytokines and reactive oxygen species released following the activation of these
glial cells (B. Wang, Navath, Romero, Kannan, & Kannan, 2009). This in turn reduces
damage to neuronal cells due to inflammatory processes (A. Karalija, Novikova, Kingham,
Wiberg, & Novikov, 2014).
A series of studies have reported improvements in cognitive function in animals with
induced traumatic brain injury following NAC supplementation (Eakin et al., 2014; Günther
et al., 2015). These studies indicate that suppression of inflammation and microglial
activation may be key to the therapeutic response to NAC (G. Chen, Shi, Hu, & Hang, 2008;
Nazıroğlu, Şenol, Ghazizadeh, & Yürüker, 2014a; Şenol, Nazıroğlu, & Yürüker, 2014). NAC
acting as a glutathione precursor can reduce the evoked oxidative stress and reverse the
behavioural deficits in mice resultant from glutathione depletion (Choy, Dean, Berk, Bush, &
van den Buuse, 2010; Dean et al., 2011), and mice engineered to exhibit mitochondrial
dysfunction and elevated reactive oxygen species show deficits of learning and memory on
complex tests which can be reversed by NAC treatment (Otte et al., 2011). Mitochondrial
dysfunction and oxidative stress is also a feature of the neurodegenerative Huntington’s
disease. NAC is able to ameliorate the cognitive deficits observed in mice expressing
elevated levels of the amyloidogenic proteins implicated in Alzheimer’s disease (Hsiao, Kuo,
Chen, & Gean, 2012; Parachikova, Green, Hendrix, & LaFerla, 2010), and Huntingtons’s
disease (Sandhir, Sood, Mehrotra, & Kamboj, 2012) which appears to be associated with
modulation of oxidative stress (Huang et al., 2010; Tchantchou, Graves, Rogers, Ortiz, &
Shea, 2005). The improvement in cognitive function with NAC treatment following brain
injury following resolution of oxidative stress has also been observed in preclinical models
(G. Chen et al., 2008; Nazıroğlu, Şenol, Ghazizadeh, & Yürüker, 2014b; Şenol et al., 2014;
Xiong, Peterson, & Lee, 1999). Promisingly, support for this mechanism and subsequent
cognitive protection has also been demonstrated in preclinical rat models of post-operative
cognitive change. (Dhanda, Kaur, & Sandhir, 2013; Prakash, Kalra, & Kumar, 2015).
There is accumulating clinical evidence for the potential efficacy of NAC in various
psychiatric and neurological conditions (Berk, Malhi, Gray, & Dean, 2013) and a recent
systematic review suggested that supplementation with NAC as either an adjunct or
monotherapy was associated with significantly pro-cognitive effects for a variety of
conditions, including psychiatric disorders, Alzheimer’s disease, traumatic brain injury, and
healthy populations (Skvarc et al., 2017). Importantly for translation to acute induction of
cognitive dysfunction, a recent study demonstrated efficacy of NAC in the setting of mild
traumatic brain injury (Hoffer, Balaban, Slade, Tsao, & Hoffer, 2013). Individuals exposed to
blast injury in the course of active military service received NAC (4 grammes daily) for seven
days following the event, and a series of symptoms including headaches, memory loss, sleep
disturbance and neurocognitive dysfunction were assessed. Those receiving NAC showed
improved performance in multiple cognitive domains relative to the placebo group, and also
showed milder ancillary symptoms of brain injury. Although limited and preliminary, the
available clinical studies provide intriguing evidence for the pro-cognitive effects of NAC.
Perhaps more importantly, these studies highlight the potential efficacy of NAC in settings
where inflammation and oxidative stress play a role in the pathological process. As far as
could be ascertained, only one trial is currently examining the efficacy of NAC for POCD,
though no data are available yet (Skvarc et al., 2016).
5.9 Preoperative cognitive exercise
Cognitive reserve is an established protective factor for POCD (Silbert et al., 2007; F. Wang,
2014) and increased levels of cognitive reserve are strongly protective for incidence of
dementia; including higher educational attainment, increased occupational complexity,
premorbid IQ, and preference for mentally stimulating leisure activities (Valenzuela &
Sachdev, 2006). Additionally, the possibility of enhancing this reserve in humans through
cognitive exercise (“brain-training”) has been demonstrated, albeit with low quality data
(Valenzuela & Sachdev, 2009).
Evidence from preclinical models suggests that the provision of a cognitively stimulating
environment actively promotes neurogenesis. Fares et al., (2013) and Hannan, (2014)
reported significantly larger proportions of hippocampal surface area and cortical thickness
in rats housed in enriched environments compared to conventionally housed animals.
Strikingly, enrichment of the preoperative environment has demonstrated significant
neuroprotection in rat models of POCD, with significant reduction in TNF-α and IL1-β after
surgery in aged rats (Kawano et al., 2015). In cognitively healthy humans, physical but not
cognitive exercise improved global cognition and was associated with increased grey matter
volume, while cognitive but not physical exercise was associated with reduced memory loss
(Suo et al., 2016). Two clinical trials are currently planned to examine the efficacy of pre-
operative cognitive exercise for reducing cognitive dysfunction in cardiac patients
(NCT02053207) and delirium in non-cardiac patients (NCT02230605), but no data was
available at the time of writing.
6.0 Conclusions
There is speculative yet promising evidence to suggest that immuno-inflammatory and
redox processes contribute to the disruptions in synaptic plasticity and glutamate signalling
which underpins cognitive dysfunction in a number of settings, including POCD. POCD may
have long-term consequences for quality of life following surgery and may have a significant
impact on the experience of the patient, their family and the healthcare burden in the
community. Emerging evidence supports the potential efficacy of neuroprotective agents in
ameliorating POCD, though clinical research remains relatively embryonic and can be
considered mixed at best without considerably more investigation. The majority of potential
therapies showed great pre-clinical promise for the mitigation of POCD, but translation into
clinical efficacy has proven to be problematic so far for a number of reasons. For example,
evidence for the efficacy of lidocaine or ketamine for POCD in humans is scant and
conflicting, with the few papers available rarely in agreement. Research into the use of
dexmedetomidine or COX-II inhibitors is extremely encouraging but requires larger and
more longitudinal randomised control trials. Conversely, perioperative use of statins or
pregabalin may be associated with increased severity of POCD, though the reliance upon
retrospective studies does not allow for potential confounds to be adequately controlled.
Further, despite good evidence for the preclinical efficacy of minocycline, NAC, or
preoperative cognitive exercise for mitigation of POCD, clinical studies remain in the
embryonic stage. To date, most research has focused upon exploration of analgesic
interventions, though anti-inflammatory interventions are a sizeable minority.
Given the lack of conclusive findings, the prevalence and impact of POCD, exploration into
potential therapies should be encouraged to the point of priority. Given the inherent
vulnerability of surgical patients, researchers and clinicians should be especially vigilant of
potential risks and complications resultant from such interventions. In particular,
dexmedetomidine, COX-II inhibitors, minocycline, and NAC appear to show the greatest
promise for potential translation via research and should be thoroughly investigated
through randomised control trials. Presently, a large-scale trial specifically exploring NAC for
POCD is currently underway (ANZCTR12614000411640). Trials such as this allow the
evaluation of suitable therapies, and once more therapies (both pharmacological and
behavioural) have been similarly evaluated, the clinical, health, and economic impact of a
multi-pronged evidence-based intervention could then be evaluated for overall efficacy.
7.0 Competing interests and role of funding
DRS is supported by the Sydney Parker Smith bequest and Centre for Research Cooperation
in Mental Health scholarships. MB is supported by a NHMRC Senior Principal Research
Fellowship 1059660. OMD has received grant support from the Brain and Behavior
Foundation, Marion and EH Flack Trust, Simons Autism Foundation, Australian Rotary
Health, Stanley Medical Research Institute, Deakin University, Brazilian Society Mobility
Program, Lilly, NHMRC, Australasian Society for Bipolar and Depressive Disorders and
Servier. She has also received in kind support from BioMedica Nutracuticals, NutritionCare
and Bioceuticals. All other authors are supported by their affiliated institutions.
The authors are currently conducting several intervention studies examining the efficacy of
NAC and Minocycline.
References
Abdul-Muneer, P. M., Chandra, N., & Haorah, J. (2014). Interactions of Oxidative Stress and
Neurovascular Inflammation in the Pathogenesis of Traumatic Brain Injury. Molecular
Neurobiology, 51(3), 966–979. https://doi.org/10.1007/s12035-014-8752-3
Abildstrom, H., Christiansen, M., Siersma, V. D., & Rasmussen, L. S. (2004). Apolipoprotein E
Genotype and Cognitive Dysfunction after Noncardiac Surgery. Anesthesiology, 101(4), 855–
861. https://doi.org/10.1097/00000542-200410000-00009
Adami, C., Bianchi, R., Pula, G., & Donato, R. (2004). S100B-stimulated NO production by BV-2
microglia is independent of RAGE transducing activity but dependent on RAGE extracellular
domain. Biochimica et Biophysica Acta (BBA) - Molecular Cell Research, 1742(1–3), 169–177.
https://doi.org/10.1016/j.bbamcr.2004.09.008
Adamis, D., Meagher, D., Treloar, A., Dunne, C., Larvin, M., Martin, F. C., & Macdonald, A. J. D.
(2014). Phenomenological and biological correlates of improved cognitive function in
hospitalized elderly medical inpatients. Archives of Gerontology and Geriatrics, 59(3), 593–
598. https://doi.org/10.1016/j.archger.2014.08.007
ADAPT-FS research group. (2015). Follow-up evaluation of cognitive function in the randomized
Alzheimer’s Disease Anti-inflammatory Prevention Trial and its Follow-up Study. Alzheimer’s
& Dementia, 11(2), 216–225.
Amir, R., Argoff, C. E., Bennett, G. J., Cummins, T. R., Durieux, M. E., Gerner, P., … Strichartz, G. R.
(2006). The Role of Sodium Channels in Chronic Inflammatory and Neuropathic Pain. The
Journal of Pain, 7(5), S1–S29. https://doi.org/10.1016/j.jpain.2006.01.444
An, L.-N., Yue, Y., Guo, W.-Z., Miao, Y.-L., Mi, W.-D., Zhang, H., … Dong, L. (2012). Surgical Trauma
Induces Iron Accumulation and Oxidative Stress in a Rodent Model of Postoperative
Cognitive Dysfunction. Biological Trace Element Research, 151(2), 277–283.
https://doi.org/10.1007/s12011-012-9564-9
Andreotti, C., King, A. A., Macy, E., Compas, B. E., & DeBaun, M. R. (2014). The Association of
Cytokine Levels With Cognitive Function in Children With Sickle Cell Disease and Normal MRI
Studies of the Brain. Journal of Child Neurology, 30(10), 1349–1353.
https://doi.org/10.1177/0883073814563140
Arcuri, C., Bianchi, R., Brozzi, F., & Donato, R. (2004). S100B Increases Proliferation in PC12 Neuronal
Cells and Reduces Their Responsiveness to Nerve Growth Factor via Akt Activation. Journal
of Biological Chemistry, 280(6), 4402–4414. https://doi.org/10.1074/jbc.m406440200
Avidan, M. S., Searleman, A. C., Storandt, M., Barnett, K., Vannucci, A., Saager, L., … Evers, A. S.
(2009). Long-term Cognitive Decline in Older Subjects Was Not Attributable to Noncardiac
Surgery or Major Illness. Anesthesiology, 111(5), 964–970.
https://doi.org/10.1097/aln.0b013e3181bc9719
Balosso, S., Liu, J., Bianchi, M. E., & Vezzani, A. (2014). Disulfide-Containing High Mobility Group Box-
1 PromotesN-Methyl-d-Aspartate Receptor Function and Excitotoxicity by Activating Toll-
Like Receptor 4-Dependent Signaling in Hippocampal Neurones. Antioxidants & Redox
Signaling, 21(12), 1726–1740. https://doi.org/10.1089/ars.2013.5349
Barrientos, R. M., Frank, M. G., Watkins, L. R., & Maier, S. F. (2012). Aging-related changes in
neuroimmune-endocrine function: Implications for hippocampal-dependent cognition.
Hormones and Behavior, 62(3), 219–227. https://doi.org/10.1016/j.yhbeh.2012.02.010
Barrientos, R. M., Sprunger, D. B., Campeau, S., Watkins, L. R., Rudy, J. W., & Maier, S. F. (2004).
BDNF mRNA expression in rat hippocampus following contextual learning is blocked by
intrahippocampal IL-1β administration. Journal of Neuroimmunology, 155(1–2), 119–126.
https://doi.org/10.1016/j.jneuroim.2004.06.009
Bayram, H., Hidiroglu, M., Cetin, L., Kucuker, A., Iriz, E., Uguz, E., … Sener, E. (2013). Comparing S-100
beta protein levels and neurocognitive functions between patients undergoing on-pump and
off-pump coronary artery bypass grafting. Journal of Surgical Research, 182(2), 198–202.
https://doi.org/10.1016/j.jss.2012.10.047
Bedford, P. . (1955). ADVERSE CEREBRAL EFFECTS OF ANÆSTHESIA ON OLD PEOPLE. The Lancet,
266(6884), 259–264. https://doi.org/10.1016/s0140-6736(55)92689-1
Beloosesky, Y., Hendel, D., Weiss, A., Hershkovitz, A., Grinblat, J., Pirotsky, A., & Barak, V. (2007).
Cytokines and C-Reactive Protein Production in Hip-Fracture-Operated Elderly Patients. The
Journals of Gerontology Series A: Biological Sciences and Medical Sciences, 62(4), 420–426.
https://doi.org/10.1093/gerona/62.4.420
Berk, M., Malhi, G. S., Gray, L. J., & Dean, O. M. (2013). The promise of N-acetylcysteine in
neuropsychiatry. Trends in Pharmacological Sciences, 34(3), 167–177.
https://doi.org/10.1016/j.tips.2013.01.001
Berman, A. E., Chan, W. Y., Brennan, A. M., Reyes, R. C., Adler, B. L., Suh, S. W., … Swanson, R. A.
(2010). N-acetylcysteine prevents loss of dopaminergic neurones in theEAAC1−/−mouse.
Annals of Neurology, 69(3), 509–520. https://doi.org/10.1002/ana.22162
Bianchi, R., Adami, C., Giambanco, I., & Donato, R. (2006). S100B binding to RAGE in microglia
stimulates COX-2 expression. Journal of Leukocyte Biology, 81(1), 108–118.
https://doi.org/10.1189/jlb.0306198
Bianchi, Roberta, Kastrisianaki, E., Giambanco, I., & Donato, R. (2011). S100B Protein Stimulates
Microglia Migration via RAGE-dependent Up-regulation of Chemokine Expression and
Release. Journal of Biological Chemistry, 286(9), 7214–7226.
https://doi.org/10.1074/jbc.m110.169342
Bilotta, F., Gelb, A. W., Stazi, E., Titi, L., Paoloni, F. P., & Rosa, G. (2013). Pharmacological
perioperative brain neuroprotection: a qualitative review of randomized clinical trials. British
Journal of Anaesthesia, 110(suppl 1), i113–i120. https://doi.org/10.1093/bja/aet059
Biscaro, B., Lindvall, O., Tesco, G., Ekdahl, C. T., & Nitsch, R. M. (2012). Inhibition of Microglial
Activation Protects Hippocampal Neurogenesis and Improves Cognitive Deficits in a
Transgenic Mouse Model for Alzheimer’s Disease. Neurodegenerative Diseases, 9(4), 187–
198. https://doi.org/10.1159/000330363
Bohnen, N., Warner, M. A., Kokmen, E., & Kurland, L. T. (1994). Early and Midlife Exposure to
Anesthesia and Age of Onset of Alzheimer’s Disease. International Journal of Neuroscience,
77(3–4), 181–185. https://doi.org/10.3109/00207459408986029
Bonelli, R. M., H??dl, A. K., Hofmann, P., & Kapfhammer, H.-P. (2004). Neuroprotection in
Huntington??s disease: a 2-year study on minocycline. International Clinical
Psychopharmacology, 19(6), 337–342. https://doi.org/10.1097/00004850-200411000-00004
Boscolo, A., Starr, J. A., Sanchez, V., Lunardi, N., DiGruccio, M. R., Ori, C., … Jevtovic-Todorovic, V.
(2012). The abolishment of anesthesia-induced cognitive impairment by timely protection of
mitochondria in the developing rat brain: The importance of free oxygen radicals and
mitochondrial integrity. Neurobiology of Disease, 45(3), 1031–1041.
https://doi.org/10.1016/j.nbd.2011.12.022
Braida, D., Sacerdote, P., Panerai, A. E., Bianchi, M., Aloisi, A. M., Iosuè, S., & Sala, M. (2004).
Cognitive function in young and adult IL (interleukin)-6 deficient mice. Behavioural Brain
Research, 153(2), 423–429. https://doi.org/10.1016/j.bbr.2003.12.018
Bromander, S., Anckarsäter, R., Kristiansson, M., Blennow, K., Zetterberg, H., Anckarsäter, H., &
Wass, C. E. (2012). Changes in serum and cerebrospinal fluid cytokines in response to non-
neurological surgery: an observational study. Journal of Neuroinflammation, 9(1).
https://doi.org/10.1186/1742-2094-9-242
Buvanendran, A., Kroin, J. S., Berger, R. A., Hallab, N. J., Saha, C., Negrescu, C., … Tuman, K. J. (2006).
Upregulation of Prostaglandin E2 and Interleukins in the Central Nervous System and
Peripheral Tissue during and after Surgery in Humans. Anesthesiology, 104(3), 403–410.
https://doi.org/10.1097/00000542-200603000-00005
Cai, D., & Liu, T. (2012). Inflammatory cause of metabolic syndrome via brain stress and NF-κB.
Aging, 4(2), 98–115. https://doi.org/10.18632/aging.100431
Calikoglu, C., Aytekin, H., Akgül, O., Akgül, M. H., Gezen, A. F., Akyuz, F., & Cakir, M. (2015). Effect of
Pregabalin in Preventing Secondary Damage in Traumatic Brain Injury: An Experimental
Study. Medical Science Monitor : International Medical Journal of Experimental and Clinical
Research, 21, 813–820. https://doi.org/10.12659/MSM.893887
Caracas, H. C. P. M., Maciel, J. V. B., Martins, P. M. R. e S., de Souza, M. M. G., & Maia, L. C. (2009).
The use of lidocaine as an anti-inflammatory substance: A systematic review. Journal of
Dentistry, 37(2), 93–97. https://doi.org/10.1016/j.jdent.2008.10.005
Cernak, I., O’Connor, C., & Vink, R. (2002). Inhibition of cyclooxygenase 2 by nimesulide improves
cognitive outcome more than motor outcome following diffuse traumatic brain injury in rats.
Experimental Brain Research, 147(2), 193–199. https://doi.org/10.1007/s00221-002-1245-z
Chan, M. T. V., Cheng, B. C. P., Lee, T. M. C., & Gin, T. (2013). BIS-guided Anesthesia Decreases
Postoperative Delirium and Cognitive Decline. Journal of Neurosurgical Anesthesiology,
25(1), 33–42. https://doi.org/10.1097/ana.0b013e3182712fba
Chen, G., Shi, J., Hu, Z., & Hang, C. (2008). Inhibitory Effect on Cerebral Inflammatory Response
following Traumatic Brain Injury in Rats: A Potential Neuroprotective Mechanism of N-
Acetylcysteine. Mediators of Inflammation, 2008, 1–8.
https://doi.org/10.1155/2008/716458
Chen, J., Buchanan, J. B., Sparkman, N. L., Godbout, J. P., Freund, G. G., & Johnson, R. W. (2008).
Neuroinflammation and disruption in working memory in aged mice after acute stimulation
of the peripheral innate immune system. Brain, Behavior, and Immunity, 22(3), 301–311.
https://doi.org/10.1016/j.bbi.2007.08.014
Chen, K., Wei, P., Zheng, Q., Zhou, J., & Li, J. (2015). Neuroprotective effects of intravenous lidocaine
on early postoperative cognitive dysfunction in elderly patients following spine surgery.
Medical Science Monitor : International Medical Journal of Experimental and Clinical
Research, 21, 1402–1407. https://doi.org/10.12659/MSM.894384
Choi, S.-H., Aid, S., & Bosetti, F. (2009). The distinct roles of cyclooxygenase-1 and -2 in
neuroinflammation: implications for translational research. Trends in Pharmacological
Sciences, 30(4), 174–181. https://doi.org/10.1016/j.tips.2009.01.002
Choy, K. H. C., Dean, O., Berk, M., Bush, A. I., & van den Buuse, M. (2010). Effects of N-acetyl-
cysteine treatment on glutathione depletion and a short-term spatial memory deficit in 2-
cyclohexene-1-one-treated rats. European Journal of Pharmacology, 649(1–3), 224–228.
https://doi.org/10.1016/j.ejphar.2010.09.035
Cibelli, M., Fidalgo, A. R., Terrando, N., Ma, D., Monaco, C., Feldmann, M., … Maze, M. (2010). Role
of interleukin-1β in postoperative cognitive dysfunction. Annals of Neurology, 68(3), 360–
368. https://doi.org/10.1002/ana.22082
Cole, R. L., Lechner, S. M., Williams, M. E., Prodanovich, P., Bleicher, L., Varney, M. A., & Gu, G.
(2005). Differential distribution of voltage-gated calcium channel alpha-2 delta (α2δ) subunit
mRNA-containing cells in the rat central nervous system and the dorsal root ganglia. The
Journal of Comparative Neurology, 491(3), 246–269. https://doi.org/10.1002/cne.20693
Corona, A. W., Fenn, A. M., & Godbout, J. P. (2011). Cognitive and Behavioral Consequences of
Impaired Immunoregulation in Aging. Journal of Neuroimmune Pharmacology, 7(1), 7–23.
https://doi.org/10.1007/s11481-011-9313-4
Costello, D. A., Watson, M. B., Cowley, T. R., Murphy, N., Murphy Royal, C., Garlanda, C., & Lynch, M.
A. (2011). Interleukin-1 and HMGB1 Mediate Hippocampal Dysfunction in SIGIRR-Deficient
Mice. Journal of Neuroscience, 31(10), 3871–3879. https://doi.org/10.1523/jneurosci.6676-
10.2011
Coursin, D. B., Coursin, D. B., & Maccioli, G. A. (2001). Dexmedetomidine. Current Opinion in Critical
Care, 7(4), 221–226. https://doi.org/10.1097/00075198-200108000-00002
Cunningham, A. J., Murray, C. A., O’Neill, L. A. J., Lynch, M. A., & O’Connor, J. J. (1996). Interleukin-1β
(IL-1β) and tumour necrosis factor (TNF) inhibit long-term potentiation in the rat dentate
gyrus in vitro. Neuroscience Letters, 203(1), 17–20. https://doi.org/10.1016/0304-
3940(95)12252-4
Cunningham, C. (2012). Microglia and neurodegeneration: The role of systemic inflammation. Glia,
61(1), 71–90. https://doi.org/10.1002/glia.22350
Cvejic, E., Lemon, J., Hickie, I. B., Lloyd, A. R., & Vollmer-Conna, U. (2014). Neurocognitive
disturbances associated with acute infectious mononucleosis, Ross River fever and Q fever:
A preliminary investigation of inflammatory and genetic correlates. Brain, Behavior, and
Immunity, 36, 207–214. https://doi.org/10.1016/j.bbi.2013.11.002
Das, S., Nanda, S., Bisoi, A., & Wadhawan, A. (2016). Effect of preoperative statin therapy on early
postoperative memory impairment after off-pump coronary artery bypass surgery. Annals of
Cardiac Anaesthesia, 19(1), 38. https://doi.org/10.4103/0971-9784.173018
De Kock, M., Loix, S., & Lavand’homme, P. (2013). Ketamine and Peripheral Inflammation. CNS
Neuroscience & Therapeutics, 19(6), 403–410. https://doi.org/10.1111/cns.12104
Dean, O. M., Data-Franco, J., Giorlando, F., & Berk, M. (2012). Minocycline. CNS Drugs, 26(5), 391–
401. https://doi.org/10.2165/11632000-000000000-00000
Dean, O. M., van den Buuse, M., Berk, M., Copolov, D. L., Mavros, C., & Bush, A. I. (2011). N-acetyl
cysteine restores brain glutathione loss in combined 2-cyclohexene-1-one and d-
amphetamine-treated rats: Relevance to schizophrenia and bipolar disorder. Neuroscience
Letters, 499(3), 149–153. https://doi.org/10.1016/j.neulet.2011.05.027
Derecki, N. C., Cardani, A. N., Yang, C. H., Quinnies, K. M., Crihfield, A., Lynch, K. R., & Kipnis, J.
(2010). Regulation of learning and memory by meningeal immunity: a key role for IL-4. The
Journal of Experimental Medicine, 207(5), 1067–1080.
https://doi.org/10.1084/jem.20091419
Derry, S., & Moore, R. A. (2013). Single dose oral celecoxib for acute postoperative pain in adults. In
Cochrane Database of Systematic Reviews. John Wiley & Sons, Ltd.
https://doi.org/10.1002/14651858.cd004233.pub4
DeWilde, K. E., Levitch, C. F., Murrough, J. W., Mathew, S. J., & Iosifescu, D. V. (2015). The promise of
ketamine for treatment-resistant depression: current evidence and future directions. Annals
of the New York Academy of Sciences, 1345(1), 47–58. https://doi.org/10.1111/nyas.12646
Dhanda, S., Kaur, S., & Sandhir, R. (2013). Preventive effect of N-acetyl-L-cysteine on oxidative stress
and cognitive impairment in hepatic encephalopathy following bile ductligation. Free Radical
Biology and Medicine, 56, 204–215. https://doi.org/10.1016/j.freeradbiomed.2012.09.017
Di Filippo, M., Chiasserini, D., Gardoni, F., Viviani, B., Tozzi, A., Giampà, C., … Calabresi, P. (2013).
Effects of central and peripheral inflammation on hippocampal synaptic plasticity.
Neurobiology of Disease, 52, 229–236. https://doi.org/10.1016/j.nbd.2012.12.009
Di Filippo, M., Chiasserini, D., Tozzi, A., Picconi, B., & Calabresi, P. (2010). Mitochondria and the Link
Between Neuroinflammation and Neurodegeneration. Journal of Alzheimer’s Disease,
20(s2), S369–S379. https://doi.org/10.3233/jad-2010-100543
Dilger, R. N., & Johnson, R. W. (2008). Aging, microglial cell priming, and the discordant central
inflammatory response to signals from the peripheral immune system. Journal of Leukocyte
Biology, 84(4), 932–939. https://doi.org/10.1189/jlb.0208108
Doll, D. N., Rellick, S. L., Barr, T. L., Ren, X., & Simpkins, J. W. (2015). Rapid mitochondrial dysfunction
mediates TNF-alpha-induced neurotoxicity. Journal of Neurochemistry, 132(4), 443–451.
https://doi.org/10.1111/jnc.13008
Donato, R., R. Cannon, B., Sorci, G., Riuzzi, F., Hsu, K., J. Weber, D., & L. Geczy, C. (2013). Functions of
S100 Proteins. Current Molecular Medicine, 13(1), 24–57.
https://doi.org/10.2174/156652413804486214
Dumont, A. O., Goursaud, S., Desmet, N., & Hermans, E. (2014). Differential Regulation of Glutamate
Transporter Subtypes by Pro-Inflammatory Cytokine TNF-α in Cortical Astrocytes from a Rat
Model of Amyotrophic Lateral Sclerosis. PLoS ONE, 9(5), e97649.
https://doi.org/10.1371/journal.pone.0097649
Dunn, L. K., Durieux, M. E., & Nemergut, E. C. (2016). Non-opioid analgesics: Novel approaches to
perioperative analgesia for major spine surgery. Best Practice & Research Clinical
Anaesthesiology, 30(1), 79–89. https://doi.org/10.1016/j.bpa.2015.11.002
Eakin, K., Baratz-Goldstein, R., Pick, C. G., Zindel, O., Balaban, C. D., Hoffer, M. E., … Hoffer, B. J.
(2014). Efficacy of N-Acetyl Cysteine in Traumatic Brain Injury. PLoS ONE, 9(4), e90617.
https://doi.org/10.1371/journal.pone.0090617
Ely, E. W., Girard, T. D., Shintani, A. K., Jackson, J. C., Gordon, S. M., Thomason, J. W. W., …
Laskowitz, D. T. (2007). Apolipoprotein E4 polymorphism as a genetic predisposition to
delirium in critically ill patients*. Critical Care Medicine, 35(1), 112–117.
https://doi.org/10.1097/01.ccm.0000251925.18961.ca
Epp, J. R., Chow, C., & Galea, L. A. M. (2013). Hippocampus-dependent learning influences
hippocampal neurogenesis. Frontiers in Neuroscience, 7.
https://doi.org/10.3389/fnins.2013.00057
Evered, L., Scott, D. A., Silbert, B., & Maruff, P. (2011). Postoperative cognitive dysfunction is
independent of type of surgery and anesthetic. Anesthesia & Analgesia, 112(5), 1179–1185.
Evonuk, K. S., Baker, B. J., Doyle, R. E., Moseley, C. E., Sestero, C. M., Johnston, B. P., … DeSilva, T. M.
(2015). Inhibition of SystemXc−Transporter Attenuates Autoimmune Inflammatory
Demyelination. The Journal of Immunology, 195(2), 450–463.
https://doi.org/10.4049/jimmunol.1401108
Eyre, H. A., Air, T., Proctor, S., Rositano, S., & Baune, B. T. (2015). A critical review of the efficacy of
non-steroidal anti-inflammatory drugs in depression. Progress in Neuro-
Psychopharmacology and Biological Psychiatry, 57, 11–16.
https://doi.org/10.1016/j.pnpbp.2014.10.003
Fan, R., Xu, F., Previti, M. L., Davis, J., Grande, A. M., Robinson, J. K., & Van Nostrand, W. E. (2007).
Minocycline Reduces Microglial Activation and Improves Behavioral Deficits in a Transgenic
Model of Cerebral Microvascular Amyloid. Journal of Neuroscience, 27(12), 3057–3063.
https://doi.org/10.1523/jneurosci.4371-06.2007
Fano’, G., Mariggio’, M. A., Angelella, P., Nicoletti, I., Antonica, A., Fulle, S., & Calissano, P. (1993).
The S-100 protein causes an increase of intracellular calcium and death of PC12 cells.
Neuroscience, 53(4), 919–925. https://doi.org/10.1016/0306-4522(93)90477-w
Fares, R. P., Belmeguenai, A., Sanchez, P. E., Kouchi, H. Y., Bodennec, J., Morales, A., … Bezin, L.
(2013). Standardized Environmental Enrichment Supports Enhanced Brain Plasticity in
Healthy Rats and Prevents Cognitive Impairment in Epileptic Rats. PLoS ONE, 8(1), e53888.
https://doi.org/10.1371/journal.pone.0053888
Fidalgo, A. R. (2013). Experimental insights into age-exacerbated cognitive dysfunction after
peripheral surgery. Aging Cell, 12(3), 523–524. https://doi.org/10.1111/acel.12066
Fischer, R., & Maier, O. (2015). Interrelation of Oxidative Stress and Inflammation in
Neurodegenerative Disease: Role of TNF. Oxidative Medicine and Cellular Longevity, 2015,
1–18. https://doi.org/10.1155/2015/610813
Fong, T. G., Jones, R. N., Shi, P., Marcantonio, E. R., Yap, L., Rudolph, J. L., … Inouye, S. K. (2009).
Delirium accelerates cognitive decline in Alzheimer disease. Neurology, 72(18), 1570–1575.
https://doi.org/10.1212/wnl.0b013e3181a4129a
Friesen, J. A., & Rodwell, V. W. (2004). Genome Biology, 5(11), 248. https://doi.org/10.1186/gb-
2004-5-11-248
Fu, X.-J., Peng, Y.-B., Hu, Y.-P., Shi, Y.-Z., Yao, M., & Zhang, X. (2014). NADPH Oxidase 1 and Its
Derived Reactive Oxygen Species Mediated Tissue Injury and Repair. Oxidative Medicine and
Cellular Longevity, 2014, 1–10. https://doi.org/10.1155/2014/282854
Funai, Y., Pickering, A. E., Uta, D., Nishikawa, K., Mori, T., Asada, A., … Furue, H. (2014). Systemic
dexmedetomidine augments inhibitory synaptic transmission in the superficial dorsal horn
through activation of descending noradrenergic control: An in vivo patch-clamp analysis of
analgesic mechanisms. Pain, 155(3), 617–628. https://doi.org/10.1016/j.pain.2013.12.018
Gabryel, B., Bielecka, A., Bernacki, J., Łabuzek, K., & Herma, Z. S. (2011). Immunosuppressant
cytoprotection correlates with HMGB1 suppression in primary astrocyte cultures exposed to
combined oxygen-glucose deprivation. Pharmacological Reports, 63(2), 392–402.
https://doi.org/10.1016/s1734-1140(11)70505-9
Ganz, P. A., Bower, J. E., Kwan, L., Castellon, S. A., Silverman, D. H. S., Geist, C., … Cole, S. W. (2013).
Does tumor necrosis factor-alpha (TNF-α) play a role in post-chemotherapy cerebral
dysfunction? Brain, Behavior, and Immunity, 30, S99–S108.
https://doi.org/10.1016/j.bbi.2012.07.015
Gao, H.-M., Zhou, H., Zhang, F., Wilson, B. C., Kam, W., & Hong, J.-S. (2011). HMGB1 Acts on
Microglia Mac1 to Mediate Chronic Neuroinflammation That Drives Progressive
Neurodegeneration. Journal of Neuroscience, 31(3), 1081–1092.
https://doi.org/10.1523/jneurosci.3732-10.2011
Gao, M., Rejaei, D., & Liu, H. (2016). Ketamine use in current clinical practice. Acta Pharmacologica
Sinica, 37(7), 865–872. https://doi.org/10.1038/aps.2016.5
Garner, S. E., Eady, A., Bennett, C., Newton, J. N., Thomas, K., & Popescu, C. M. (2012). Minocycline
for acne vulgaris: efficacy and safety. In Cochrane Database of Systematic Reviews. John
Wiley & Sons, Ltd. https://doi.org/10.1002/14651858.cd002086.pub2
Garrido-Mesa, N., Zarzuelo, A., & Gálvez, J. (2013). Minocycline: far beyond an antibiotic. British
Journal of Pharmacology, 169(2), 337–352. https://doi.org/10.1111/bph.12139
Ge, Y.-L., Li, X., Gao, J. U., Zhang, X., Fang, X., Zhou, L., … Lin, S. (2016). Beneficial effects of
intravenous dexmedetomidine on cognitive function and cerebral injury following a carotid
endarterectomy. Experimental and Therapeutic Medicine, 11(3), 1128–1134.
https://doi.org/10.3892/etm.2016.2978
Gemma, C., Fister, M., Hudson, C., & Bickford, P. C. (2005). Improvement of memory for context by
inhibition of caspase-1 in aged rats. European Journal of Neuroscience, 22(7), 1751–1756.
https://doi.org/10.1111/j.1460-9568.2005.04334.x
Giovannitti, J. A., Thoms, S. M., & Crawford, J. J. (2015). Alpha-2 Adrenergic Receptor Agonists: A
Review of Current Clinical Applications. Anesthesia Progress, 62(1), 31–38.
https://doi.org/10.2344/0003-3006-62.1.31
Girgis, H., Palmier, B., Croci, N., Soustrat, M., Plotkine, M., & Marchand-Leroux, C. (2013). Effects of
selective and non-selective cyclooxygenase inhibition against neurological deficit and brain
oedema following closed head injury in mice. Brain Research, 1491, 78–87.
https://doi.org/10.1016/j.brainres.2012.10.049
Godbout, J. P. (2005). Exaggerated neuroinflammation and sickness behavior in aged mice after
activation of the peripheral innate immune system. The FASEB Journal.
https://doi.org/10.1096/fj.05-3776fje
Godbout, J. P., & Johnson, R. W. (2004). Interleukin-6 in the aging brain. Journal of
Neuroimmunology, 147(1–2), 141–144. https://doi.org/10.1016/j.jneuroim.2003.10.031
Grieco, J. C., Ciarlone, S. L., Gieron-Korthals, M., Schoenberg, M. R., Smith, A. G., Philpot, R. M., …
Weeber, E. J. (2014). An open-label pilot trial of minocycline in children as a treatment for
Angelman syndrome. BMC Neurology, 14(1). https://doi.org/10.1186/s12883-014-0232-x
Günther, M., Davidsson, J., Plantman, S., Norgren, S., Mathiesen, T., & Risling, M. (2015).
Neuroprotective effects of N-acetylcysteine amide on experimental focal penetrating brain
injury in rats. Journal of Clinical Neuroscience, 22(9), 1477–1483.
https://doi.org/10.1016/j.jocn.2015.03.025
Han, S.-H., Kim, Y. H., & Mook-Jung, I. (2011). RAGE: The beneficial and deleterious effects by diverse
mechanisms of actions. Molecules and Cells, 31(2), 91–97. https://doi.org/10.1007/s10059-
011-0030-x
Hannan, A. J. (2014). Review: Environmental enrichment and brain repair: harnessing the
therapeutic effects of cognitive stimulation and physical activity to enhance experience-
dependent plasticity. Neuropathology and Applied Neurobiology, 40(1), 13–25.
https://doi.org/10.1111/nan.12102
Harmon, D., Eustace, N., Ghori, K., Butler, M., OʼCallaghan, S., OʼDonnell, A., … Shorten, G. (2005).
Plasma concentrations of nitric oxide products and cognitive dysfunction following coronary
artery bypass surgery. European Journal of Anaesthesiology, 22(4), 269–276.
https://doi.org/10.1017/s0265021505000451
Haxaire, C., Turpin, F. R., Potier, B., Kervern, M., Sinet, P.-M., Barbanel, G., … Billard, J.-M. (2012).
Reversal of age-related oxidative stress prevents hippocampal synaptic plasticity deficits by
protecting d-serine-dependent NMDA receptor activation. Aging Cell, 11(2), 336–344.
https://doi.org/10.1111/j.1474-9726.2012.00792.x
He, H.-J., Wang, Y., Le, Y., Duan, K.-M., Yan, X.-B., Liao, Q., … Ouyang, W. (2012). Surgery Upregulates
High Mobility Group Box-1 and Disrupts the Blood-Brain Barrier causing Cognitive
Dysfunction in Aged Rats. CNS Neuroscience & Therapeutics, 18(12), 994–1002.
https://doi.org/10.1111/cns.12018
Heidari, S., Rahavi, A., & Hashemi, S. (2013). Lidocaine administration before tracheal extubation
cannot reduce post-operative cognition disorders in elderly patients. Advanced Biomedical
Research, 2(1), 81. https://doi.org/10.4103/2277-9175.120869
Heldt, S. A., Stanek, L., Chhatwal, J. P., & Ressler, K. J. (2007). Hippocampus-specific deletion of BDNF
in adult mice impairs spatial memory and extinction of aversive memories. Molecular
Psychiatry, 12(7), 656–670. https://doi.org/10.1038/sj.mp.4001957
Heneka, M. T., Kummer, M. P., & Latz, E. (2014). Innate immune activation in neurodegenerative
disease. Nature Reviews Immunology, 14(7), 463–477. https://doi.org/10.1038/nri3705
Heyer, E. J., Mergeche, J. L., Bruce, S. S., & Connolly, E. S. (2013). Inflammation and Cognitive
Dysfunction in Type 2 Diabetic Carotid Endarterectomy Patients. Diabetes Care, 36(10),
3283–3286. https://doi.org/10.2337/dc12-2507
Heyer, E. J., Wilson, D. A., Sahlein, D. H., Mocco, J., Williams, S. C., Sciacca, R., … Connolly, E. S.
(2005). APOE- 4 predisposes to cognitive dysfunction following uncomplicated carotid
endarterectomy. Neurology, 65(11), 1759–1763.
https://doi.org/10.1212/01.wnl.0000184579.23624.6b
Heyer, Eric J., Mergeche, J. L., Wang, S., Gaudet, J. G., & Connolly, E. S. (2015). Impact of Cognitive
Dysfunction on Survival in Patients With and Without Statin Use Following Carotid
Endarterectomy. Neurosurgery, 77(6), 880–887.
https://doi.org/10.1227/neu.0000000000000904
Heyser, C. J., Masliah, E., Samimi, A., Campbell, I. L., & Gold, L. H. (1997). Progressive decline in
avoidance learning paralleled by inflammatory neurodegeneration in transgenic mice
expressing interleukin 6 in the brain. Proceedings of the National Academy of Sciences, 94(4),
1500–1505. https://doi.org/10.1073/pnas.94.4.1500
Hindmarch, I., Trick, L., & Ridout, F. (2005). A double-blind, placebo- and positive-internal-controlled
(alprazolam) investigation of the cognitive and psychomotor profile of pregabalin in healthy
volunteers. Psychopharmacology, 183(2), 133–143. https://doi.org/10.1007/s00213-005-
0172-7
Hoffer, M. E., Balaban, C., Slade, M. D., Tsao, J. W., & Hoffer, B. (2013). Amelioration of Acute
Sequelae of Blast Induced Mild Traumatic Brain Injury by N-Acetyl Cysteine: A Double-Blind,
Placebo Controlled Study. PLoS ONE, 8(1), e54163.
https://doi.org/10.1371/journal.pone.0054163
Hou, Y., Wang, L., Yi, D., Ding, B., Yang, Z., Li, J., … Wu, G. (2012). N-acetylcysteine reduces
inflammation in the small intestine by regulating redox, EGF and TLR4 signaling. Amino Acids,
45(3), 513–522. https://doi.org/10.1007/s00726-012-1295-x
Hovens, I. B., Schoemaker, R. G., van der Zee, E. A., Absalom, A. R., Heineman, E., & van Leeuwen, B.
L. (2014). Postoperative cognitive dysfunction: Involvement of neuroinflammation and
neuronal functioning. Brain, Behavior, and Immunity, 38, 202–210.
https://doi.org/10.1016/j.bbi.2014.02.002
Hovens, I. B., van Leeuwen, B. L., Nyakas, C., Heineman, E., van der Zee, E. A., & Schoemaker, R. G.
(2015). Postoperative cognitive dysfunction and microglial activation in associated brain
regions in old rats. Neurobiology of Learning and Memory, 118, 74–79.
https://doi.org/10.1016/j.nlm.2014.11.009
Hsiao, Y.-H., Kuo, J.-R., Chen, S.-H., & Gean, P.-W. (2012). Amelioration of social isolation-triggered
onset of early Alzheimer’s disease-related cognitive deficit by N-acetylcysteine in a
transgenic mouse model. Neurobiology of Disease, 45(3), 1111–1120.
https://doi.org/10.1016/j.nbd.2011.12.031
Huang, Q., Aluise, C. D., Joshi, G., Sultana, R., St. Clair, D. K., Markesbery, W. R., & Butterfield, D. A.
(2010). Potential in vivo amelioration by N-acetyl-L-cysteine of oxidative stress in brain in
human double mutant APP/PS-1 knock-in mice: Toward therapeutic modulation of mild
cognitive impairment. Journal of Neuroscience Research, n/a-n/a.
https://doi.org/10.1002/jnr.22422
Hudetz, J. A., Gandhi, S. D., Iqbal, Z., Patterson, K. M., & Pagel, P. S. (2010). Elevated postoperative
inflammatory biomarkers are associated with short- and medium-term cognitive dysfunction
after coronary artery surgery. Journal of Anesthesia, 25(1), 1–9.
https://doi.org/10.1007/s00540-010-1042-y
HUDETZ, J. A., IQBAL, Z., GANDHI, S. D., PATTERSON, K. M., BYRNE, A. J., HUDETZ, A. G., … WARLTIER,
D. C. (2009). Ketamine attenuates post-operative cognitive dysfunction after cardiac surgery.
Acta Anaesthesiologica Scandinavica, 53(7), 864–872. https://doi.org/10.1111/j.1399-
6576.2009.01978.x
Hudetz, J. A., Patterson, K. M., Amole, O., Riley, A. V., & Pagel, P. S. (2011). Postoperative cognitive
dysfunction after noncardiac surgery: effects of metabolic syndrome. Journal of Anesthesia,
25(3), 337–344. https://doi.org/10.1007/s00540-011-1137-0
Hudetz, J. A., Patterson, K. M., Iqbal, Z., Gandhi, S. D., & Pagel, P. S. (2011a). Metabolic Syndrome
Exacerbates Short-term Postoperative Cognitive Dysfunction in Patients Undergoing Cardiac
Surgery: Results of a Pilot Study. Journal of Cardiothoracic and Vascular Anesthesia, 25(2),
282–287. https://doi.org/10.1053/j.jvca.2010.06.008
Hudetz, J. A., Patterson, K. M., Iqbal, Z., Gandhi, S. D., & Pagel, P. S. (2011b). Metabolic Syndrome
Exacerbates Short-term Postoperative Cognitive Dysfunction in Patients Undergoing Cardiac
Surgery: Results of a Pilot Study. Journal of Cardiothoracic and Vascular Anesthesia, 25(2),
282–287. https://doi.org/10.1053/j.jvca.2010.06.008
Iliff, J. J., Goldman, S. A., & Nedergaard, M. (2015). Implications of the discovery of brain lymphatic
pathways. The Lancet Neurology, 14(10), 977–979. https://doi.org/10.1016/s1474-
4422(15)00221-5
Inta, D., Lang, U. E., Borgwardt, S., Meyer-Lindenberg, A., & Gass, P. (2016). Microglia Activation and
Schizophrenia: Lessons From the Effects of Minocycline on Postnatal Neurogenesis,
Neuronal Survival and Synaptic Pruning. Schizophrenia Bulletin, sbw088.
https://doi.org/10.1093/schbul/sbw088
Iohom, G., Szarvas, S., Larney, V., O’Brien, J., Buckley, E., Butler, M., & Shorten, G. (2004).
Perioperative Plasma Concentrations of Stable Nitric Oxide Products Are Predictive of
Cognitive Dysfunction After Laparoscopic Cholecystectomy. Anesthesia & Analgesia, 99(4),
1245–1252. https://doi.org/10.1213/01.ane.0000132971.00206.4a
Jaturapatporn, D., Isaac, M. G. E. K. N., McCleery, J., & Tabet, N. (2012). Aspirin, steroidal and non-
steroidal anti-inflammatory drugs for the treatment of Alzheimer’s disease. In Cochrane
Database of Systematic Reviews. John Wiley & Sons, Ltd.
https://doi.org/10.1002/14651858.cd006378.pub2
Ji, K., Akgul, G., Wollmuth, L. P., & Tsirka, S. E. (2013). Microglia Actively Regulate the Number of
Functional Synapses. PLoS ONE, 8(2), e56293.
https://doi.org/10.1371/journal.pone.0056293
Ji, K., Miyauchi, J., & Tsirka, S. E. (2013). Microglia: An Active Player in the Regulation of Synaptic
Activity. Neural Plasticity, 2013, 1–9. https://doi.org/10.1155/2013/627325
Ji, M.-H., Yuan, H.-M., Zhang, G.-F., Li, X.-M., Dong, L., Li, W.-Y., … Yang, J.-J. (2012). Changes in
plasma and cerebrospinal fluid biomarkers in aged patients with early postoperative
cognitive dysfunction following total hip-replacement surgery. Journal of Anesthesia, 27(2),
236–242. https://doi.org/10.1007/s00540-012-1506-3
Jin, W.-J., Feng, S.-W., Feng, Z., Lu, S.-M., Qi, T., & Qian, Y.-N. (2013). Minocycline improves
postoperative cognitive impairment in aged mice by inhibiting astrocytic activation.
NeuroReport, 1. https://doi.org/10.1097/wnr.0b013e3283629195
Karachitos, A., Solis Garcia del Pozo, J., W.J. de Groot, P., Kmita, H., & Jordan, J. (2012). Minocycline
Mediated Mitochondrial Cytoprotection: Premises for Therapy of Cerebrovascular and
Neurodegenerative Diseases. Current Drug Targets, 14(1), 47–55.
https://doi.org/10.2174/1389450111314010006
Karalija, A., Novikova, L. N., Kingham, P. J., Wiberg, M., & Novikov, L. N. (2014). The effects of N-
acetyl-cysteine and acetyl-l-carnitine on neural survival, neuroinflammation and
regeneration following spinal cord injury. Neuroscience, 269, 143–151.
https://doi.org/10.1016/j.neuroscience.2014.03.042
Karalija, Amar, Novikova, L. N., Kingham, P. J., Wiberg, M., & Novikov, L. N. (2012). Neuroprotective
Effects of N-Acetyl-Cysteine and Acetyl-L-Carnitine after Spinal Cord Injury in Adult Rats.
PLoS ONE, 7(7), e41086. https://doi.org/10.1371/journal.pone.0041086
Kawano, T., Eguchi, S., Iwata, H., Tamura, T., Kumagai, N., & Yokoyama, M. (2015). Impact of
Preoperative Environmental Enrichment on Prevention of Development of Cognitive
Impairment following Abdominal Surgery in a Rat Model. Anesthesiology, 123(1), 160–170.
https://doi.org/10.1097/aln.0000000000000697
Kawano, T., Eguchi, S., Iwata, H., Yamanaka, D., Tateiwa, H., Locatelli, F. M., & Yokoyama, M. (2016).
Pregabalin can prevent, but not treat, cognitive dysfunction following abdominal surgery in
aged rats. Life Sciences, 148, 211–219. https://doi.org/10.1016/j.lfs.2016.02.021
Kawasaki, T., Kawasaki, C., Ueki, M., Hamada, K., Habe, K., & Sata, T. (2013). Dexmedetomidine
suppresses proinflammatory mediator production in human whole blood in vitro. Journal of
Trauma and Acute Care Surgery, 74(5), 1370–1375.
https://doi.org/10.1097/ta.0b013e31828db978
Kazmierski, J., Banys, A., Latek, J., Bourke, J., Jaszewski, R., Sobow, T., & Kloszewska, I. (2014). Mild
Cognitive Impairment with Associated Inflammatory and Cortisol Alterations as Independent
Risk Factor for Postoperative Delirium. Dementia and Geriatric Cognitive Disorders, 38(1–2),
65–78. https://doi.org/10.1159/000357454
Kelly, Á., Lynch, A., Vereker, E., Nolan, Y., Queenan, P., Whittaker, E., … Lynch, M. A. (2001). The Anti-
inflammatory Cytokine, Interleukin (IL)-10, Blocks the Inhibitory Effect of IL-1β on Long Term
Potentiation. Journal of Biological Chemistry, 276(49), 45564–45572.
https://doi.org/10.1074/jbc.m108757200
Kettenmann, H., Hanisch, U.-K., Noda, M., & Verkhratsky, A. (2011). Physiology of Microglia.
Physiological Reviews, 91(2), 461–553. https://doi.org/10.1152/physrev.00011.2010
Kim, H.-S., & Suh, Y.-H. (2009). Minocycline and neurodegenerative diseases. Behavioural Brain
Research, 196(2), 168–179. https://doi.org/10.1016/j.bbr.2008.09.040
Kim, I.-D., & Lee, J.-K. (2013). HMGB1-Binding Heptamer Confers Anti-Inflammatory Effects in
Primary Microglia Culture. Experimental Neurobiology, 22(4), 301.
https://doi.org/10.5607/en.2013.22.4.301
Kim, J.-B., Sig Choi, J., Yu, Y.-M., Nam, K., Piao, C.-S., Kim, S.-W., … Lee, J.-K. (2006). HMGB1, a Novel
Cytokine-Like Mediator Linking Acute Neuronal Death and Delayed Neuroinflammation in
the Postischemic Brain. Journal of Neuroscience, 26(24), 6413–6421.
https://doi.org/10.1523/jneurosci.3815-05.2006
Ko, F., Isoda, F., & Mobbs, C. (2015). Laparotomy in Mice Induces Blood Cell Expression of
Inflammatory and Stress Genes. Journal of Interferon & Cytokine Research, 35(4), 302–312.
https://doi.org/10.1089/jir.2014.0031
Kohl, B. A., & Deutschman, C. S. (2006). The inflammatory response to surgery and trauma. Current
Opinion in Critical Care, 12(4), 325–332.
https://doi.org/10.1097/01.ccx.0000235210.85073.fc
Kohman, R. A., Bhattacharya, T. K., Kilby, C., Bucko, P., & Rhodes, J. S. (2013). Effects of minocycline
on spatial learning, hippocampal neurogenesis and microglia in aged and adult mice.
Behavioural Brain Research, 242, 17–24. https://doi.org/10.1016/j.bbr.2012.12.032
Kong, F., Chen, S., Cheng, Y., Ma, L., Lu, H., Zhang, H., & Hu, W. (2013). Minocycline Attenuates
Cognitive Impairment Induced by Isoflurane Anesthesia in Aged Rats. PLoS ONE, 8(4),
e61385. https://doi.org/10.1371/journal.pone.0061385
Kornblit, B., Munthe-Fog, L., Madsen, H. O., Strom, J., Vindelov, L., & Garred, P. (2008). Association
of HMGB1 polymorphisms with outcome in patients with systemic inflammatory response
syndrome. Critical Care, 12(3), R83. https://doi.org/10.1186/cc6935
Kotzampassi, K., Kolios, G., Manousou, P., Kazamias, P., Paramythiotis, D., Papavramidis, T. S., …
Eleftheriadis, E. (2009). Oxidative stress due to anesthesia and surgical trauma: Importance
of early enteral nutrition. Molecular Nutrition & Food Research, 53(6), 770–779.
https://doi.org/10.1002/mnfr.200800166
Kroon, F. P., van der Burg, L. R., Ramiro, S., Landewé, R. B., Buchbinder, R., Falzon, L., & van der
Heijde, D. (2015). Non-steroidal anti-inflammatory drugs (NSAIDs) for axial spondyloarthritis
(ankylosing spondylitis and non-radiographic axial spondyloarthritis). In Cochrane Database
of Systematic Reviews. John Wiley & Sons, Ltd.
https://doi.org/10.1002/14651858.cd010952.pub2
Langford, R. M., Joshi, G. P., Gan, T. J., Mattera, M. S., Chen, W.-H., Revicki, D. A., … Zlateva, G.
(2009). Reduction in Opioid-Related Adverse Events and Improvement in Function with
Parecoxib followed by Valdecoxib Treatment after Non-Cardiac Surgery. Clinical Drug
Investigation, 29(9), 577–590. https://doi.org/10.2165/11317570-000000000-00000
Laskowitz, D. T., Thekdi, A. D., Thekdi, S. D., Han, S. K. D., Myers, J. K., Pizzo, S. V., & Bennett, E. R.
(2001). Downregulation of Microglial Activation by Apolipoprotein E and ApoE-Mimetic
Peptides. Experimental Neurology, 167(1), 74–85. https://doi.org/10.1006/exnr.2001.7541
Lauretti, G. R. (2008). Mecanismos envolvidos na analgesia da lidocaína por via venosa. Revista
Brasileira de Anestesiologia, 58(3), 280–286. https://doi.org/10.1590/s0034-
70942008000300011
Lee, T. A., Wolozin, B., Weiss, K. B., & Bednar, M. M. (2005). Assessment of the emergence of
Alzheimer’s disease following coronary artery bypass graft surgery or percutaneous
transluminal coronary angioplasty1. Journal of Alzheimer’s Disease, 7(4), 319–324.
https://doi.org/10.3233/jad-2005-7408
Lee, Y., Syeda, K., Maruschak, N. A., Cha, D. S., Mansur, R. B., Wium-Andersen, I. K., … McIntyre, R. S.
(2016). A New Perspective on the Anti-Suicide Effects With Ketamine Treatment. Journal of
Clinical Psychopharmacology, 36(1), 50–56. https://doi.org/10.1097/jcp.0000000000000441
Lee, Y.-H., Su, S.-B., Huang, C.-C., Sheu, H.-M., Tsai, J.-C., Lin, C.-H., … Wang, B.-J. (2014a). N-
Acetylcysteine Attenuates Hexavalent Chromium-Induced Hypersensitivity through
Inhibition of Cell Death, ROS-Related Signaling and Cytokine Expression. PLoS ONE, 9(9),
e108317. https://doi.org/10.1371/journal.pone.0108317
Lee, Y.-H., Su, S.-B., Huang, C.-C., Sheu, H.-M., Tsai, J.-C., Lin, C.-H., … Wang, B.-J. (2014b). N-
Acetylcysteine Attenuates Hexavalent Chromium-Induced Hypersensitivity through
Inhibition of Cell Death, ROS-Related Signaling and Cytokine Expression. PLoS ONE, 9(9),
e108317. https://doi.org/10.1371/journal.pone.0108317
Leung, J. M., Sands, L. P., Wang, Y., Poon, A., Kwok, P., Kane, J. P., & Pullinger, C. R. (2007).
Apolipoprotein E e4 Allele Increases the Risk of Early Postoperative Delirium in Older
Patients Undergoing Noncardiac Surgery. Anesthesiology, 107(3), 406–411.
https://doi.org/10.1097/01.anes.0000278905.07899.df
Li, B., Wang, H., Wu, H., & Gao, C. (2015). Neurocognitive Dysfunction Risk Alleviation With the Use
of Dexmedetomidine in Perioperative Conditions or as ICU Sedation. Medicine, 94(14), e597.
https://doi.org/10.1097/md.0000000000000597
Li, R.-L., Zhang, Z.-Z., Peng, M., Wu, Y., Zhang, J.-J., Wang, C.-Y., & Wang, Y.-L. (2013). Postoperative
impairment of cognitive function in old mice: a possible role for neuroinflammation
mediated by HMGB1, S100B, and RAGE. Journal of Surgical Research, 185(2), 815–824.
https://doi.org/10.1016/j.jss.2013.06.043
Li, S., Chen, X., Chen, Y., Tan, L., Zhao, Y., Wang, J., … Luo, A. (2013). Role of GSK-3β in isoflurane-
induced neuroinflammation and cognitive dysfunction in aged rats. Journal of Huazhong
University of Science and Technology [Medical Sciences], 33(4), 530–535.
https://doi.org/10.1007/s11596-013-1154-3
Li, S.-T., & Wu, J. (2015). Dexmedetomidine May Produce Extra Protective Effects on Sepsis-induced
Diaphragm Injury. Chinese Medical Journal, 128(10), 1407. https://doi.org/10.4103/0366-
6999.156808
Li, S.-Y., Xia, L.-X., Zhao, Y.-L., Yang, L., Chen, Y.-L., Wang, J.-T., & Luo, A.-L. (2013). Minocycline
mitigates isoflurane-induced cognitive impairment in aged rats. Brain Research, 1496, 84–
93. https://doi.org/10.1016/j.brainres.2012.12.025
Li, Y., He, R., Chen, S., & Qu, Y. (2015). Effect of dexmedetomidine on early postoperative cognitive
dysfunction and peri-operative inflammation in elderly patients undergoing laparoscopic
cholecystectomy. Experimental and Therapeutic Medicine.
https://doi.org/10.3892/etm.2015.2726
LI, Y.-C., XI, C.-H., AN, Y.-F., DONG, W.-H., & ZHOU, M. (2012). Perioperative inflammatory response
and protein S-100β concentrations - relationship with post-operative cognitive dysfunction
in elderly patients. Acta Anaesthesiologica Scandinavica, 56(5), 595–600.
https://doi.org/10.1111/j.1399-6576.2011.02616.x
Li, Z., Wei, H., Piirainen, S., Chen, Z., Kalso, E., Pertovaara, A., & Tian, L. (2016). Spinal versus brain
microglial and macrophage activation traits determine the differential neuroinflammatory
responses and analgesic effect of minocycline in chronic neuropathic pain. Brain, Behavior,
and Immunity, 58, 107–117. https://doi.org/10.1016/j.bbi.2016.05.021
Lin, D., Cao, L., Wang, Z., Li, J., Washington, J. M., & Zuo, Z. (2012). Lidocaine attenuates cognitive
impairment after isoflurane anesthesia in old rats. Behavioural Brain Research, 228(2), 319–
327. https://doi.org/10.1016/j.bbr.2011.12.010
Lin, E., Calvano, S. E., & Lowry, S. F. (2000). Inflammatory cytokines and cell response in surgery.
Surgery, 127(2), 117–126. https://doi.org/10.1067/msy.2000.101584
LIN, G.-X., WANG, T., CHEN, M.-H., HU, Z.-H., & OUYANG, W. (2014). Serum high-mobility group box
1 protein correlates with cognitive decline after gastrointestinal surgery. Acta
Anaesthesiologica Scandinavica, 58(6), 668–674. https://doi.org/10.1111/aas.12320
Ling, Y., Ma, W., Yu, L., Zhang, Y., & Liang, Q. (2015). Decreased PSD95 expression in medial
prefrontal cortex (mPFC) was associated with cognitive impairment induced by sevoflurane
anesthesia. Journal of Zhejiang University SCIENCE B, 16(9), 763–771.
https://doi.org/10.1631/jzus.b1500006
Linstedt, U., Meyer, O., Kropp, P., Berkau, A., Tapp, E., & Zenz, M. (2002). Serum concentration of S-
100 protein in assessment of cognitive dysfunction after general anesthesia in different
types of surgery. Acta Anaesthesiologica Scandinavica, 46(4), 384–389.
https://doi.org/10.1034/j.1399-6576.2002.460409.x
Liu, F.-L., Chen, T.-L., & Chen, R.-M. (2012). Mechanisms of ketamine-induced immunosuppression.
Acta Anaesthesiologica Taiwanica, 50(4), 172–177.
https://doi.org/10.1016/j.aat.2012.12.001
Liu, Y., Xu, X., Dou, H., Hua, Y., Xu, J., & Hui, X. (2015). Apolipoprotein E knockout induced
inflammatory responses related to microglia in neonatal mice brain via astrocytes.
International Journal of Clinical and Experimental Medicine, 8(1), 737–743.
Louveau, A., Smirnov, I., Keyes, T. J., Eccles, J. D., Rouhani, S. J., Peske, J. D., … Kipnis, J. (2015).
Structural and functional features of central nervous system lymphatic vessels. Nature,
523(7560), 337–341. https://doi.org/10.1038/nature14432
Lu, B., Nagappan, G., & Lu, Y. (2014). BDNF and Synaptic Plasticity, Cognitive Function, and
Dysfunction. Neurotrophic Factors, 223–250. https://doi.org/10.1007/978-3-642-45106-5_9
MacLullich, A. M. J., Edelshain, B. T., Hall, R. J., de Vries, A., Howie, S. E. M., Pearson, A., … van
Munster, B. C. (2011). CEREBROSPINAL FLUID INTERLEUKIN-8 LEVELS ARE HIGHER IN PEOPLE
WITH HIP FRACTURE WITH PERIOPERATIVE DELIRIUM THAN IN CONTROLS. Journal of the
American Geriatrics Society, 59(6), 1151–1153. https://doi.org/10.1111/j.1532-
5415.2011.03428.x
Malfitano, A. M., Marasco, G., Proto, M. C., Laezza, C., Gazzerro, P., & Bifulco, M. (2014). Statins in
neurological disorders: An overview and update. Pharmacological Research, 88, 74–83.
https://doi.org/10.1016/j.phrs.2014.06.007
Man, Y., Guo, Z., Cao, J., & Mi, W. (2015). Efficacy of perioperative dexmedetomidine in
postoperative neurocognitive function: a meta-analysis. Clinical and Experimental
Pharmacology and Physiology, 42(8), 837–842. https://doi.org/10.1111/1440-1681.12432
Manganelli, V., Signore, M., Pacini, I., Misasi, R., Tellan, G., Garofalo, T., … Delogu, G. (2010).
Increased HMGB1 expression and release by mononuclear cells following surgical/anesthesia
trauma. Critical Care, 14(6), R197. https://doi.org/10.1186/cc9316
Mannick, J. A., Rodrick, M. L., & Lederer, J. A. (2001). The immunologic response to injury. Journal of
the American College of Surgeons, 193(3), 237–244. https://doi.org/10.1016/s1072-
7515(01)01011-0
Margaritis, M., Channon, K. M., & Antoniades, C. (2014). Statins as Regulators of Redox State in the
Vascular Endothelium: Beyond Lipid Lowering. Antioxidants & Redox Signaling, 20(8), 1198–
1215. https://doi.org/10.1089/ars.2013.5430
Mathew, J. P., Grocott, H. P., McCurdy, J. R., Ti, L. K., Davis, R. D., Laskowitz, D. T., … Newman, M. F.
(2005). Preoperative Statin Therapy Does Not Reduce Cognitive Dysfunction After
Cardiopulmonary Bypass. Journal of Cardiothoracic and Vascular Anesthesia, 19(3), 294–299.
https://doi.org/10.1053/j.jvca.2005.03.004
Mattia, C., & Coluzzi, F. (2005). COX-2 inhibitors: pharmacological data and adverse effects. Minerva
Anestesiologica, 71(7–8), 461–470.
McGuinness, B., Craig, D., Bullock, R., & Passmore, P. (2016). Statins for the prevention of dementia.
In Cochrane Database of Systematic Reviews. John Wiley & Sons, Ltd.
https://doi.org/10.1002/14651858.cd003160.pub3
Menger, M. D., & Vollmar, B. (2004). Surgical trauma: hyperinflammation versus
immunosuppression? Langenbeck’s Archives of Surgery, 389(6), 475–484.
https://doi.org/10.1007/s00423-004-0472-0
Michetti, F., Corvino, V., Geloso, M. C., Lattanzi, W., Bernardini, C., Serpero, L., & Gazzolo, D. (2012).
The S100B protein in biological fluids: more than a lifelong biomarker of brain distress.
Journal of Neurochemistry, 120(5), 644–659. https://doi.org/10.1111/j.1471-
4159.2011.07612.x
Ming, Z., Sawicki, G., & Bekar, L. K. (2015). Acute systemic LPS-mediated inflammation induces
lasting changes in mouse cortical neuromodulation and behavior. Neuroscience Letters, 590,
96–100. https://doi.org/10.1016/j.neulet.2015.01.081
Mitchell, S. J., Merry, A. F., Frampton, C., Davies, E., Grieve, D., Mills, B. P., … Gorman, D. F. (2009).
Cerebral Protection by Lidocaine During Cardiac Operations: A Follow-Up Study. The Annals
of Thoracic Surgery, 87(3), 820–825. https://doi.org/10.1016/j.athoracsur.2008.12.042
Moller, J., Cluitmans, P., Rasmussen, L., Houx, P., Rasmussen, H., Canet, J., … Gravenstein, J. (1998).
Long-term postoperative cognitive dysfunction in the elderly: ISPOCD1 study. The Lancet,
351(9106), 857–861. https://doi.org/10.1016/s0140-6736(97)07382-0
Monk, T. G., Weldon, B. C., Garvan, C. W., Dede, D. E., Van Der Aa, M. T., Heilman, K. M., &
Gravenstein, J. S. (2008). Predictors of cognitive dysfunction after major noncardiac surgery.
The Journal of the American Society of Anesthesiologists, 108(1), 18–30.
Moore, R. A., Wiffen, P. J., Derry, S., & McQuay, H. J. (2009). Gabapentin for chronic neuropathic
pain in adults. In Cochrane Database of Systematic Reviews. John Wiley & Sons, Ltd.
https://doi.org/10.1002/14651858.cd007938
Mortero, R. F., Clark, L. D., Tolan, M. M., Metz, R. J., Tsueda, K., & Sheppard, R. A. (2001). The Effects
of Small-Dose Ketamine on Propofol Sedation: Respiration, Postoperative Mood, Perception,
Cognition, and Pain. Anesthesia and Analgesia, 1465–1469.
https://doi.org/10.1097/00000539-200106000-00022
Mospan, C. M. (2016). Are statins protective or harmful to cognitive function? Journal of the
American Academy of Physician Assistants, 29(1), 11–12.
https://doi.org/10.1097/01.jaa.0000475471.02134.37
Murman, D. L., Giordani, B., Mellow, A. M., Johanns, J. R., Little, R. J. A., Hariharan, M., & Foster, N. L.
(1997). Cognitive, behavioral, and motor effects of the NMDA antagonist ketamine in
Huntington’s disease. Neurology, 49(1), 153–161. https://doi.org/10.1212/wnl.49.1.153
Myhre, M., Diep, L. M., & Stubhaug, A. (2016). Pregabalin Has Analgesic, Ventilatory, and Cognitive
Effects in Combination with Remifentanil. Anesthesiology, 124(1), 141–149.
https://doi.org/10.1097/aln.0000000000000913
Na, K.-S., Jung, H.-Y., & Kim, Y.-K. (2014). The role of pro-inflammatory cytokines in the
neuroinflammation and neurogenesis of schizophrenia. Progress in Neuro-
Psychopharmacology and Biological Psychiatry, 48, 277–286.
https://doi.org/10.1016/j.pnpbp.2012.10.022
Nadelson, M., Sanders, R., & Avidan, M. (2014). Perioperative cognitive trajectory in adults. British
Journal of Anaesthesia, aet420.
Nagels, W., Demeyere, R., Van Hemelrijck, J., Vandenbussche, E., Gijbels, K., & Vandermeersch, E.
(2004). Evaluation of the Neuroprotective Effects of S(+)-Ketamine During Open-Heart
Surgery. Anesthesia & Analgesia, 1595–1603.
https://doi.org/10.1213/01.ane.0000117227.00820.0c
Nakasujja, N., Miyahara, S., Evans, S., Lee, A., Musisi, S., Katabira, E., … Sacktor, N. (2012).
Randomized trial of minocycline in the treatment of HIV-associated cognitive impairment.
Neurology, 80(2), 196–202. https://doi.org/10.1212/wnl.0b013e31827b9121
Nazıroğlu, M., Şenol, N., Ghazizadeh, V., & Yürüker, V. (2014a). Neuroprotection Induced by N-
acetylcysteine and Selenium Against Traumatic Brain Injury-Induced Apoptosis and Calcium
Entry in Hippocampus of Rat. Cellular and Molecular Neurobiology, 34(6), 895–903.
https://doi.org/10.1007/s10571-014-0069-2
Nazıroğlu, M., Şenol, N., Ghazizadeh, V., & Yürüker, V. (2014b). Neuroprotection Induced by N-
acetylcysteine and Selenium Against Traumatic Brain Injury-Induced Apoptosis and Calcium
Entry in Hippocampus of Rat. Cellular and Molecular Neurobiology, 34(6), 895–903.
https://doi.org/10.1007/s10571-014-0069-2
Nelson, L. E., Lu, J., Guo, T., Saper, C. B., Franks, N. P., & Maze, M. (2003). The α2-Adrenoceptor
Agonist Dexmedetomidine Converges on an Endogenous Sleep-promoting Pathway to Exert
Its Sedative Effects. Anesthesiology, 98(2), 428–436. https://doi.org/10.1097/00000542-
200302000-00024
Norden, D. M., & Godbout, J. P. (2013). Review: Microglia of the aged brain: primed to be activated
and resistant to regulation. Neuropathology and Applied Neurobiology, 39(1), 19–34.
https://doi.org/10.1111/j.1365-2990.2012.01306.x
Oldham, M. A., Hawkins, K. A., Yuh, D. D., Dewar, M. L., Darr, U. M., Lysyy, T., & Lee, H. B. (2015).
Cognitive and functional status predictors of delirium and delirium severity after coronary
artery bypass graft surgery: an interim analysis of the Neuropsychiatric Outcomes After
Heart Surgery study. International Psychogeriatrics, 27(12), 1929–1938.
https://doi.org/10.1017/s1041610215001477
Oliveira, C. M. B. de, Issy, A. M., & Sakata, R. K. (2010). Lidocaína por via venosa intraoperatória.
Revista Brasileira de Anestesiologia, 60(3), 325–332. https://doi.org/10.1590/s0034-
70942010000300012
Ott, B. R., Daiello, L. A., Dahabreh, I. J., Springate, B. A., Bixby, K., Murali, M., & Trikalinos, T. A.
(2015). Do Statins Impair Cognition? A Systematic Review and Meta-Analysis of Randomized
Controlled Trials. Journal of General Internal Medicine, 30(3), 348–358.
https://doi.org/10.1007/s11606-014-3115-3
Otte, D.-M., Sommersberg, B., Kudin, A., Guerrero, C., Albayram, Ö., Filiou, M. D., … Zimmer, A.
(2011). N-acetyl Cysteine Treatment Rescues Cognitive Deficits Induced by Mitochondrial
Dysfunction in G72/G30 Transgenic Mice. Neuropsychopharmacology, 36(11), 2233–2243.
https://doi.org/10.1038/npp.2011.109
Pal, R., Kar, S., Kundu, K., Sarkar, U., Gupta, S., Mandal, S., … Sarbapalli, D. (2011). Impact of general
versus epidural anesthesia on early post-operative cognitive dysfunction following hip and
knee surgery. Journal of Emergencies, Trauma, and Shock, 4(1), 23.
https://doi.org/10.4103/0974-2700.76829
Pan, W., P. Stone, K., Hsuchou, H., K. Manda, V., Zhang, Y., & J. Kastin, A. (2011). Cytokine Signaling
Modulates Blood-Brain Barrier Function. Current Pharmaceutical Design, 17(33), 3729–3740.
https://doi.org/10.2174/138161211798220918
Parachikova, A., Green, K. N., Hendrix, C., & LaFerla, F. M. (2010). Formulation of a Medical Food
Cocktail for Alzheimer’s Disease: Beneficial Effects on Cognition and Neuropathology in a
Mouse Model of the Disease. PLoS ONE, 5(11), e14015.
https://doi.org/10.1371/journal.pone.0014015
Paribello, C., Tao, L., Folino, A., Berry-Kravis, E., Tranfaglia, M., Ethell, I. M., & Ethell, D. W. (2010).
Open-label add-on treatment trial of minocycline in fragile X syndrome. BMC Neurology,
10(1). https://doi.org/10.1186/1471-2377-10-91
Parnet, P., Amindari, S., Wu, C., Brunke-Reese, D., Goujon, E., Weyhenmeyer, J. A., … Kelley, K. W.
(1994). Expression of type I and type II interleukin-1 receptors in mouse brain. Molecular
Brain Research, 27(1), 63–70. https://doi.org/10.1016/0169-328x(94)90185-6
Patel, S., Naeem, S., Kesingland, A., Froestl, W., Capogna, M., Urban, L., & Fox, A. (2001). The effects
of GABAB agonists and gabapentin on mechanical hyperalgesia in models of neuropathic and
inflammatory pain in the rat. Pain, 90(3), 217–226. https://doi.org/10.1016/s0304-
3959(00)00404-8
Pedrazzi, M., Averna, M., Sparatore, B., Patrone, M., Salamino, F., Marcoli, M., … Melloni, E. (2012).
Potentiation of NMDA Receptor-Dependent Cell Responses by Extracellular High Mobility
Group Box 1 Protein. PLoS ONE, 7(8), e44518.
https://doi.org/10.1371/journal.pone.0044518
Peng, L., Xu, L., & Ouyang, W. (2013). Role of Peripheral Inflammatory Markers in Postoperative
Cognitive Dysfunction (POCD): A Meta-Analysis. PLoS ONE, 8(11), e79624.
https://doi.org/10.1371/journal.pone.0079624
Peng, W., Yang, J., Yang, B., Wang, L., Xiong, X., & Liang, Q. (2014). Impact of Statins on Cognitive
Deficits in Adult Male Rodents after Traumatic Brain Injury: A Systematic Review. BioMed
Research International, 2014, 1–13. https://doi.org/10.1155/2014/261409
Peng, Y., Zhang, W., Zhou, X., Ji, Y., Kass, I. S., & Han, R. (2016). Lidocaine Did Not Reduce
Neuropsychological-Cognitive Decline in Patients 6 Months After Supratentorial Tumor
Surgery. Journal of Neurosurgical Anesthesiology, 28(1), 6–13.
https://doi.org/10.1097/ana.0000000000000171
Perry, V. H., & Teeling, J. (2013). Microglia and macrophages of the central nervous system: the
contribution of microglia priming and systemic inflammation to chronic neurodegeneration.
Seminars in Immunopathology, 35(5), 601–612. https://doi.org/10.1007/s00281-013-0382-8
Petrova, T. V., Hu, J., & Van Eldik, L. J. (2000). Modulation of glial activation by astrocyte-derived
protein S100B: differential responses of astrocyte and microglial cultures. Brain Research,
853(1), 74–80. https://doi.org/10.1016/s0006-8993(99)02251-9
Pexton, T., Moeller-Bertram, T., Schilling, J. M., & Wallace, M. S. (2011). Targeting voltage-gated
calcium channels for the treatment of neuropathic pain: a review of drug development.
Expert Opinion on Investigational Drugs, 20(9), 1277–1284.
https://doi.org/10.1517/13543784.2011.600686
Poon, D. C.-H., Ho, Y.-S., Chiu, K., Wong, H.-L., & Chang, R. C.-C. (2015). Sickness: From the focus on
cytokines, prostaglandins, and complement factors to the perspectives of neurones.
Neuroscience & Biobehavioral Reviews, 57, 30–45.
https://doi.org/10.1016/j.neubiorev.2015.07.015
Popp, S. S., Lei, B., Kelemen, E., Fenton, A. A., Cottrell, J. E., & Kass, I. S. (2011). Intravenous
antiarrhythmic doses of lidocaine increase the survival rate of CA1 neurones and improve
cognitive outcome after transient global cerebral ischemia in rats. Neuroscience, 192, 537–
549. https://doi.org/10.1016/j.neuroscience.2011.06.086
Prakash, A., Kalra, J. K., & Kumar, A. (2015). Neuroprotective effect of N-acetyl cysteine against
streptozotocin-induced memory dysfunction and oxidative damage in rats. Journal of Basic
and Clinical Physiology and Pharmacology, 26(1). https://doi.org/10.1515/jbcpp-2013-0150
Pribiag, H., & Stellwagen, D. (2013). TNF- Downregulates Inhibitory Neurotransmission through
Protein Phosphatase 1-Dependent Trafficking of GABAA Receptors. Journal of Neuroscience,
33(40), 15879–15893. https://doi.org/10.1523/jneurosci.0530-13.2013
Price, C. C., Garvan, C. W., & Monk, T. G. (2008). Type and severity of cognitive decline in older
adults after noncardiac surgery. Anesthesiology: The Journal of the American Society of
Anesthesiologists, 108(1), 8–17.
Purkayastha, S., & Cai, D. (2013). Neuroinflammatory basis of metabolic syndrome. Molecular
Metabolism, 2(4), 356–363. https://doi.org/10.1016/j.molmet.2013.09.005
Qian, X.-L., Zhang, W., Liu, M.-Z., Zhou, Y.-B., Zhang, J.-M., Han, L., … Wang, Q.-D. (2015).
Dexmedetomidine improves early postoperative cognitive dysfunction in aged mice.
European Journal of Pharmacology, 746, 206–212.
https://doi.org/10.1016/j.ejphar.2014.11.017
Qiao, Y., Feng, H., Zhao, T., Yan, H., Zhang, H., & Zhao, X. (2015). Postoperative cognitive dysfunction
after inhalational anesthesia in elderly patients undergoing major surgery: the influence of
anesthetic technique, cerebral injury and systemic inflammation. BMC Anesthesiology, 15(1).
https://doi.org/10.1186/s12871-015-0130-9
Rachal Pugh, C., Fleshner, M., Watkins, L. R., Maier, S. F., & Rudy, J. W. (2001). The immune system
and memory consolidation: a role for the cytokine IL-1β. Neuroscience & Biobehavioral
Reviews, 25(1), 29–41. https://doi.org/10.1016/s0149-7634(00)00048-8
Radtke, F. M., Franck, M., Lendner, J., Kruger, S., Wernecke, K. D., & Spies, C. D. (2013). Monitoring
depth of anaesthesia in a randomized trial decreases the rate of postoperative delirium but
not postoperative cognitive dysfunction. British Journal of Anaesthesia, 110(suppl 1), i98–
i105. https://doi.org/10.1093/bja/aet055
Reikerås, O. (2010). Immune depression in musculoskeletal trauma. Inflammation Research, 59(6),
409–414. https://doi.org/10.1007/s00011-010-0167-7
Reinsfelt, B., Ricksten, S.-E., Zetterberg, H., Blennow, K., Fredén-Lindqvist, J., & Westerlind, A. (2012).
Cerebrospinal Fluid Markers of Brain Injury, Inflammation, and Blood-Brain Barrier
Dysfunction in Cardiac Surgery. The Annals of Thoracic Surgery, 94(2), 549–555.
https://doi.org/10.1016/j.athoracsur.2012.04.044
Revel, F., Gilbert, T., Roche, S., Drai, J., Blond, E., Ecochard, R., & Bonnefoy, M. (2015). Influence of
oxidative stress biomarkers on cognitive decline. Journal of Alzheimer’s Disease, 45(2), 553–
560.
Riazi, K., Galic, M. A., Kentner, A. C., Reid, A. Y., Sharkey, K. A., & Pittman, Q. J. (2015). Microglia-
Dependent Alteration of Glutamatergic Synaptic Transmission and Plasticity in the
Hippocampus during Peripheral Inflammation. Journal of Neuroscience, 35(12), 4942–4952.
https://doi.org/10.1523/jneurosci.4485-14.2015
Ritzel, R. M., Patel, A. R., Pan, S., Crapser, J., Hammond, M., Jellison, E., & McCullough, L. D. (2015).
Age- and location-related changes in microglial function. Neurobiology of Aging, 36(6),
2153–2163. https://doi.org/10.1016/j.neurobiolaging.2015.02.016
Robillard, J. M., Gordon, G. R., Choi, H. B., Christie, B. R., & MacVicar, B. A. (2011). Glutathione
Restores the Mechanism of Synaptic Plasticity in Aged Mice to That of the Adult. PLoS ONE,
6(5), e20676. https://doi.org/10.1371/journal.pone.0020676
Rojo, A. I., McBean, G., Cindric, M., Egea, J., López, M. G., Rada, P., … Cuadrado, A. (2014). Redox
Control of Microglial Function: Molecular Mechanisms and Functional Significance.
Antioxidants & Redox Signaling, 21(12), 1766–1801. https://doi.org/10.1089/ars.2013.5745
Rosczyk, H. A., Sparkman, N. L., & Johnson, R. W. (2008). Neuroinflammation and cognitive function
in aged mice following minor surgery. Experimental Gerontology, 43(9), 840–846.
https://doi.org/10.1016/j.exger.2008.06.004
Rothwell, N. J., Luheshi, G., & Toulmond, S. (1996). Cytokines and their receptors in the central
nervous system: Physiology, pharmacology, and pathology. Pharmacology & Therapeutics,
69(2), 85–95. https://doi.org/10.1016/0163-7258(95)02033-0
Ryckman, C., Vandal, K., Rouleau, P., Talbot, M., & Tessier, P. A. (2003). Proinflammatory Activities of
S100: Proteins S100A8, S100A9, and S100A8/A9 Induce Neutrophil Chemotaxis and
Adhesion. The Journal of Immunology, 170(6), 3233–3242.
https://doi.org/10.4049/jimmunol.170.6.3233
Sacktor, N., Miyahara, S., Deng, L., Evans, S., Schifitto, G., Cohen, B. A., … Clifford, D. B. (2011).
Minocycline treatment for HIV-associated cognitive impairment: Results from a randomized
trial. Neurology, 77(12), 1135–1142. https://doi.org/10.1212/wnl.0b013e31822f0412
Sacktor, Ned, Miyahara, S., Evans, S., Schifitto, G., Cohen, B., Haughey, N., … Clifford, D. (2014).
Impact of minocycline on cerebrospinal fluid markers of oxidative stress, neuronal injury,
and inflammation in HIV-seropositive individuals with cognitive impairment. Journal of
NeuroVirology, 20(6), 620–626. https://doi.org/10.1007/s13365-014-0292-0
Sałat, K., Gdula-Argasińska, J., Malikowska, N., Podkowa, A., Lipkowska, A., & Librowski, T. (2016).
Effect of pregabalin on contextual memory deficits and inflammatory state-related protein
expression in streptozotocin-induced diabetic mice. Naunyn-Schmiedeberg’s Archives of
Pharmacology, 389(6), 613–623. https://doi.org/10.1007/s00210-016-1230-x
Salinsky, M., Storzbach, D., & Munoz, S. (2010). Cognitive effects of pregabalin in healthy volunteers:
A double-blind, placebo-controlled trial. Neurology, 74(9), 755–761.
https://doi.org/10.1212/wnl.0b013e3181d25b34
Sanchez, V., Feinstein, S. D., Lunardi, N., Joksovic, P. M., Boscolo, A., Todorovic, S. M., & Jevtovic-
Todorovic, V. (2011). General Anesthesia Causes Long-term Impairment of Mitochondrial
Morphogenesis and Synaptic Transmission in Developing Rat Brain. Anesthesiology, 115(5),
992–1002. https://doi.org/10.1097/aln.0b013e3182303a63
Sanders, R. D., Xu, J., Shu, Y., Januszewski, A., Halder, S., Fidalgo, A., … Maze, M. (2009).
Dexmedetomidine Attenuates Isoflurane-induced Neurocognitive Impairment in Neonatal
Rats. Anesthesiology, 110(5), 1077–1085. https://doi.org/10.1097/aln.0b013e31819daedd
Sandhir, R., Sood, A., Mehrotra, A., & Kamboj, S. S. (2012). N-Acetylcysteine Reverses Mitochondrial
Dysfunctions and Behavioral Abnormalities in 3-Nitropropionic Acid-Induced Huntington’s
Disease. Neurodegenerative Diseases, 9(3), 145–157. https://doi.org/10.1159/000334273
Santiago, F. M., Bueno, P., Olmedo, C., Muffak-Granero, K., Comino, A., Serradilla, M., … Ferrón, J. A.
(2008). Effect of N-Acetylcysteine Administration on Intraoperative Plasma Levels of
Interleukin-4 and Interleukin-10 in Liver Transplant Recipients. Transplantation Proceedings,
40(9), 2978–2980. https://doi.org/10.1016/j.transproceed.2008.08.103
SATO, K., KIMURA, T., NISHIKAWA, T., TOBE, Y., & MASAKI, Y. (2010). Neuroprotective effects of a
combination of dexmedetomidine and hypothermia after incomplete cerebral ischemia in
rats. Acta Anaesthesiologica Scandinavica, 54(3), 377–382. https://doi.org/10.1111/j.1399-
6576.2009.02139.x
Schulze-Bonhage, A. (2012). Pharmacokinetic and pharmacodynamic profile of pregabalin and its
role in the treatment of epilepsy. Expert Opinion on Drug Metabolism & Toxicology, 9(1),
105–115. https://doi.org/10.1517/17425255.2013.749239
Seitz, D. P., Shah, P. S., Herrmann, N., Beyene, J., & Siddiqui, N. (2011). Exposure to general
anesthesia and risk of Alzheimer’s disease: a systematic review and meta-analysis. BMC
Geriatrics, 11(1). https://doi.org/10.1186/1471-2318-11-83
Şenol, N., Nazıroğlu, M., & Yürüker, V. (2014). N-Acetylcysteine and Selenium Modulate Oxidative
Stress, Antioxidant Vitamin and Cytokine Values in Traumatic Brain Injury-Induced Rats.
Neurochemical Research, 39(4), 685–692. https://doi.org/10.1007/s11064-014-1255-9
Sheets, M. F., & Hanck, D. A. (2003). Molecular Action of Lidocaine on the Voltage Sensors of Sodium
Channels. The Journal of General Physiology, 121(2), 163–175.
https://doi.org/10.1085/jgp.20028651
Shen, X., Dong, Y., Xu, Z., Wang, H., Miao, C., Soriano, S. G., … Xie, Z. (2013). Selective Anesthesia-
induced Neuroinflammation in Developing Mouse Brain and Cognitive Impairment.
Anesthesiology, 118(3), 502–515. https://doi.org/10.1097/aln.0b013e3182834d77
Shi, H.-J., Xue, X.-H., Wang, Y.-L., Zhang, W.-S., Wang, Z.-S., & Yu, A.-L. (2015). Effects of different
anesthesia methods on cognitive dysfunction after hip replacement operation in elder
patients. International Journal of Clinical and Experimental Medicine, 8(3), 3883–3888.
Shim, J.-K., Ma, Q., Zhang, Z., Podgoreanu, M. V., & Mackensen, G. B. (2014). Effect of pregabalin on
cerebral outcome after cardiopulmonary bypass with deep hypothermic circulatory arrest in
rats. The Journal of Thoracic and Cardiovascular Surgery, 148(1), 298–303.
https://doi.org/10.1016/j.jtcvs.2014.02.076
Short, T. G., Leslie, K., Chan, M. T. V., Campbell, D., Frampton, C., & Myles, P. (2015). Rationale and
Design of the Balanced Anesthesia Study. Anesthesia & Analgesia, 121(2), 357–365.
https://doi.org/10.1213/ane.0000000000000797
Silbert, B., Evered, L., Scott, D. A., McMahon, S., Choong, P., Ames, D., … Jamrozik, K. (2015).
Preexisting Cognitive Impairment Is Associated with Postoperative Cognitive Dysfunction
after Hip Joint Replacement Surgery. Anesthesiology, 122(6), 1224–1234.
https://doi.org/10.1097/aln.0000000000000671
Silbert, B. S., Evered, L. A., Scott, D. A., & Cowie, T. F. (2008). The Apolipoprotein E ϵ4 Allele is not
Associated With Cognitive Dysfunction in Cardiac Surgery. The Annals of Thoracic Surgery,
86(3), 841–847. https://doi.org/10.1016/j.athoracsur.2008.04.085
Silbert, B. S., Scott, D. A., Evered, L. A., Lewis, M. S., & Maruff, P. T. (2007). Preexisting Cognitive
Impairment in Patients Scheduled for Elective Coronary Artery Bypass Graft Surgery.
Anesthesia & Analgesia, 104(5), 1023–1028.
https://doi.org/10.1213/01.ane.0000263285.03361.3a
Simon, W., Hapfelmeier, G., Kochs, E., Zieglgänsberger, W., & Rammes, G. (2001). Isoflurane Blocks
Synaptic Plasticity in the Mouse Hippocampus. Anesthesiology, 94(6), 1058–1065.
https://doi.org/10.1097/00000542-200106000-00021
Skvarc, D. R., Dean, O. M., Byrne, L. K., Gray, L. J., Ives, K., Lane, S. E., … Marriott, A. J. (2016). The
Post-Anaesthesia N-acetylcysteine Cognitive Evaluation (PANACEA) trial: study protocol for a
randomised controlled trial. Trials, 17(1). https://doi.org/10.1186/s13063-016-1529-4
Skvarc, D. R., Dean, O. M., Byrne, L. K., Gray, L., Lane, S., Lewis, M., … Marriott, A. (2017). The effect
of N-acetylcysteine (NAC) on human cognition – A systematic review. Neuroscience &
Biobehavioral Reviews, 78, 44–56. https://doi.org/10.1016/j.neubiorev.2017.04.013
Sorci, G., Bianchi, R., Riuzzi, F., Tubaro, C., Arcuri, C., Giambanco, I., & Donato, R. (2010). S100B
Protein, a Damage-Associated Molecular Pattern Protein in the Brain and Heart, and Beyond.
Cardiovascular Psychiatry and Neurology, 2010, 1–13. https://doi.org/10.1155/2010/656481
Steinmetz, J., Christensen, K. B., Lund, T., Lohse, N., & Rasmussen, L. S. (2009). Long-term
consequences of postoperative cognitive dysfunction. The Journal of the American Society of
Anesthesiologists, 110(3), 548–555.
Stellwagen, D., & Malenka, R. C. (2006). Synaptic scaling mediated by glial TNF-α. Nature, 440(7087),
1054–1059. https://doi.org/10.1038/nature04671
Su, X., Feng, X., Terrando, N., Yan, Y., Chawla, A., Koch, L. G., … Maze, M. (2012). Dysfunction of
Inflammation-Resolving Pathways Is Associated with Exaggerated Postoperative Cognitive
Decline in a Rat Model of the Metabolic Syndrome. Molecular Medicine, 18(1), 1481–1490.
https://doi.org/10.2119/molmed.2012.00351
Suda, K., Kitagawa, Y., Ozawa, S., Saikawa, Y., Ueda, M., Abraham, E., … Ishizaka, A. (2006). Serum
concentrations of high-mobility group box chromosomal protein 1 before and after exposure
to the surgical stress of thoracic esophagectomy: a predictor of clinical course after surgery?
Diseases of the Esophagus, 19(1), 5–9. https://doi.org/10.1111/j.1442-2050.2006.00529.x
Sun, Z.-L., & Feng, D.-F. (2013). Biomarkers of cognitive dysfunction in traumatic brain injury. Journal
of Neural Transmission, 121(1), 79–90. https://doi.org/10.1007/s00702-013-1078-x
Suo, C., Singh, M. F., Gates, N., Wen, W., Sachdev, P., Brodaty, H., … Valenzuela, M. J. (2016).
Therapeutically relevant structural and functional mechanisms triggered by physical and
cognitive exercise. Molecular Psychiatry, 21(11), 1645–1645.
https://doi.org/10.1038/mp.2016.57
Swiger, K. J., Manalac, R. J., Blumenthal, R. S., Blaha, M. J., & Martin, S. S. (2013). Statins and
Cognition: A Systematic Review and Meta-analysis of Short- and Long-term Cognitive Effects.
Mayo Clinic Proceedings, 88(11), 1213–1221. https://doi.org/10.1016/j.mayocp.2013.07.013
Taccone, F. S., Su, F., Pierrakos, C., He, X., James, S., Dewitte, O., … De Backer, D. (2010). Cerebral
microcirculation is impaired during sepsis: an experimental study. Critical Care, 14(4), R140.
https://doi.org/10.1186/cc9205
Takahata, R., Ono, S., Tsujimoto, H., Hiraki, S., Kimura, A., Kinoshita, M., … Hase, K. (2011).
Postoperative Serum Concentrations of High Mobility Group Box Chromosomal Protein-1
Correlates to the Duration of SIRS and Pulmonary Dysfunction Following Gastrointestinal
Surgery. Journal of Surgical Research, 170(1), e135–e140.
https://doi.org/10.1016/j.jss.2011.04.040
Talarowska, M., Gałecki, P., Maes, M., Gardner, A., Chamielec, M., Orzechowska, A., … Kowalczyk, E.
(2011). Malondialdehyde plasma concentration correlates with declarative and working
memory in patients with recurrent depressive disorder. Molecular Biology Reports, 39(5),
5359–5366. https://doi.org/10.1007/s11033-011-1335-8
Talarowska, M., Gałecki, P., Maes, M., Orzechowska, A., Chamielec, M., Bartosz, G., & Kowalczyk, E.
(2012). Nitric oxide plasma concentration associated with cognitive impairment in patients
with recurrent depressive disorder. Neuroscience Letters, 510(2), 127–131.
https://doi.org/10.1016/j.neulet.2012.01.018
Tan, H., Cao, J., Zhang, J., & Zuo, Z. (2014). Critical role of inflammatory cytokines in impairing
biochemical processes for learning and memory after surgery in rats. Journal of
Neuroinflammation, 11(1), 93. https://doi.org/10.1186/1742-2094-11-93
Tanaka, K. Z., Pevzner, A., Hamidi, A. B., Nakazawa, Y., Graham, J., & Wiltgen, B. J. (2014). Cortical
Representations Are Reinstated by the Hippocampus during Memory Retrieval. Neuron,
84(2), 347–354. https://doi.org/10.1016/j.neuron.2014.09.037
Tanaka, S., Ide, M., Shibutani, T., Ohtaki, H., Numazawa, S., Shioda, S., & Yoshida, T. (2006).
Lipopolysaccharide-induced microglial activation induces learning and memory deficits
without neuronal cell deathin rats. Journal of Neuroscience Research, 83(4), 557–566.
https://doi.org/10.1002/jnr.20752
Tang, J. X., Baranov, D., Hammond, M., Shaw, L. M., Eckenhoff, M. F., & Eckenhoff, R. G. (2011).
Human Alzheimer and Inflammation Biomarkers after Anesthesia and Surgery.
Anesthesiology, 115(4), 727–732. https://doi.org/10.1097/aln.0b013e31822e9306
Tchantchou, F., Graves, M., Rogers, E., Ortiz, D., & Shea, T. B. (2005). N-acteyl cysteine alleviates
oxidative damage to central nervous system of ApoE-deficient mice following folate and
vitamin E-deficiency. Journal of Alzheimer’s Disease, 7(2), 135–138.
https://doi.org/10.3233/jad-2005-7206
Terrando, N., Monaco, C., Ma, D., Foxwell, B. M. J., Feldmann, M., & Maze, M. (2010). Tumor
necrosis factor- triggers a cytokine cascade yielding postoperative cognitive decline.
Proceedings of the National Academy of Sciences, 107(47), 20518–20522.
https://doi.org/10.1073/pnas.1014557107
Terrando, Niccolò, Eriksson, L. I., Kyu Ryu, J., Yang, T., Monaco, C., Feldmann, M., … Maze, M. (2011).
Resolving postoperative neuroinflammation and cognitive decline. Annals of Neurology,
70(6), 986–995. https://doi.org/10.1002/ana.22664
Terrando, Niccolò, Rei Fidalgo, A., Vizcaychipi, M., Cibelli, M., Ma, D., Monaco, C., … Maze, M.
(2010). The impact of IL-1 modulation on the development of lipopolysaccharide-induced
cognitive dysfunction. Critical Care, 14(3), R88. https://doi.org/10.1186/cc9019
The DSM-5 criteria, level of arousal and delirium diagnosis: inclusiveness is safer. (2014). BMC
Medicine, 12(1). https://doi.org/10.1186/s12916-014-0141-2
Tian, Y., Guo, S., Wu, X., Ma, L., & Zhao, X. (2015). Minocycline Alleviates Sevoflurane-Induced
Cognitive Impairment in Aged Rats. Cellular and Molecular Neurobiology, 35(4), 585–594.
https://doi.org/10.1007/s10571-014-0154-6
Toth, C. (2013). Pregabalin: latest safety evidence and clinical implications for the management of
neuropathic pain. Therapeutic Advances in Drug Safety, 5(1), 38–56.
https://doi.org/10.1177/2042098613505614
Tsai, T. L., Sands, L. P., & Leung, J. M. (2010). An Update on Postoperative Cognitive Dysfunction.
Advances in Anesthesia, 28(1), 269–284. https://doi.org/10.1016/j.aan.2010.09.003
Usatyuk, P. V., Vepa, S., Watkins, T., He, D., Parinandi, N. L., & Natarajan, V. (2003). Redox Regulation
of Reactive Oxygen Species-Induced p38 MAP Kinase Activation and Barrier Dysfunction in
Lung Microvascular Endothelial Cells. Antioxidants & Redox Signaling, 5(6), 723–730.
https://doi.org/10.1089/152308603770380025
Vacas, S., Degos, V., Tracey, K. J., & Maze, M. (2014). High-mobility Group Box 1 Protein Initiates
Postoperative Cognitive Decline by Engaging Bone Marrow–derived Macrophages.
Anesthesiology, 120(5), 1160–1167. https://doi.org/10.1097/aln.0000000000000045
Valenzuela, M. J., & Sachdev, P. (2006). Brain reserve and dementia: a systematic review.
Psychological Medicine, 36(4), 441–454. https://doi.org/10.1017/S0033291705006264
Valenzuela, M., & Sachdev, P. (2009). Can Cognitive Exercise Prevent the Onset of Dementia?
Systematic Review of Randomized Clinical Trials with Longitudinal Follow-up. The American
Journal of Geriatric Psychiatry, 17(3), 179–187.
https://doi.org/10.1097/jgp.0b013e3181953b57
van der Wal, S. E. I., van den Heuvel, S. A. S., Radema, S. A., van Berkum, B. F. M., Vaneker, M.,
Steegers, M. A. H., … Vissers, K. C. P. (2015). Thein vitromechanisms andin vivoefficacy of
intravenous lidocaine on the neuroinflammatory response in acute and chronic pain.
European Journal of Pain, 20(5), 655–674. https://doi.org/10.1002/ejp.794
van Munster, B. C., Korevaar, J. C., de Rooij, S. E., Levi, M., & Zwinderman, A. H. (2007). The
association between delirium and the apolipoprotein E ϵ4 allele in the elderly. Psychiatric
Genetics, 17(5), 261–266. https://doi.org/10.1097/ypg.0b013e3280c8efd4
Vasunilashorn, S. M., Ngo, L., Kosar, C. M., Fong, T. G., Jones, R. N., Inouye, S. K., & Marcantonio, E.
R. (2015). Does Apolipoprotein E Genotype Increase Risk of Postoperative Delirium? The
American Journal of Geriatric Psychiatry, 23(10), 1029–1037.
https://doi.org/10.1016/j.jagp.2014.12.192
Veliz-Reissmüller, G., Torres, H. A., van der Linden, J., Lindblom, D., & Jönhagen, M. E. (2007). Pre-
operative mild cognitive dysfunction predicts risk for post-operative delirium after elective
cardiac surgery. Aging Clinical and Experimental Research, 19(3), 172–177.
https://doi.org/10.1007/bf03324686
Vereker, E., O’Donnell, E., Lynch, A., Kelly, Á., Nolan, Y., & Lynch, M. A. (2001). Evidence that
interleukin-1β and reactive oxygen species production play a pivotal role in stress-induced
impairment of LTP in the rat dentate gyrus. European Journal of Neuroscience, 14(11), 1809–
1819. https://doi.org/10.1046/j.0953-816x.2001.01809.x
Vezzani, A., & Viviani, B. (2015). Neuromodulatory properties of inflammatory cytokines and their
impact on neuronal excitability. Neuropharmacology, 96, 70–82.
https://doi.org/10.1016/j.neuropharm.2014.10.027
Wan, Q., Xu, L., & Bo, Y. (2014). Effects of Dexmedetomidine combined with Dezocine on cognition
function and hippocampal microglia activation of rats. International Journal of Clinical and
Experimental Medicine, 7(9), 2787–2792.
Wang, B., Navath, R. S., Romero, R., Kannan, S., & Kannan, R. (2009). Anti-inflammatory and anti-
oxidant activity of anionic dendrimer–N-acetyl cysteine conjugates in activated microglial
cells. International Journal of Pharmaceutics, 377(1–2), 159–168.
https://doi.org/10.1016/j.ijpharm.2009.04.050
Wang, C., Xia, Y., Zheng, Y., Dai, W., Wang, F., Chen, K., … Guo, C. (2015). Protective Effects of N-
Acetylcysteine in Concanavalin A-Induced Hepatitis in Mice. Mediators of Inflammation,
2015, 1–17. https://doi.org/10.1155/2015/189785
Wang, D., Wu, X., Li, J., Xiao, F., Liu, X., & Meng, M. (2002). The Effect of Lidocaine on Early
Postoperative Cognitive Dysfunction After Coronary Artery Bypass Surgery. Anesthesia &
Analgesia, 95(5), 1134–1141. https://doi.org/10.1097/00000539-200211000-00002
Wang, D.-S., Zurek, A. A., Lecker, I., Yu, J., Abramian, A. M., Avramescu, S., … Orser, B. A. (2012).
Memory Deficits Induced by Inflammation Are Regulated by α5-Subunit-Containing GABAA
Receptors. Cell Reports, 2(3), 488–496. https://doi.org/10.1016/j.celrep.2012.08.022
Wang, F. (2014). Postoperative Cognitive Dysfunction: Current Developments in Mechanism and
Prevention. Medical Science Monitor, 20, 1908–1912.
https://doi.org/10.12659/msm.892485
Wang, H.-L., Liu, H., Xue, Z.-G., Liao, Q.-W., & Fang, H. (2016). Minocycline attenuates post-operative
cognitive impairment in aged mice by inhibiting microglia activation. Journal of Cellular and
Molecular Medicine, 20(9), 1632–1639. https://doi.org/10.1111/jcmm.12854
Wible, E. F., & Laskowitz, D. T. (2010). Statins in traumatic brain injury. Neurotherapeutics, 7(1), 62–
73. https://doi.org/10.1016/j.nurt.2009.11.003
Woiciechowsky, C., Asadullah, K., Nestler, D., Glckner, F., Robinson, P. N., Volk, H.-D., … Lanksch,
W. R. (1997). Different release of cytokines into the cerebrospinal fluid following surgery for
intra- and extra-axial brain tumours. Acta Neurochirurgica, 139(7), 619–624.
https://doi.org/10.1007/bf01411996
Wynne, A. M., Henry, C. J., & Godbout, J. P. (2009). Immune and behavioral consequences of
microglial reactivity in the aged brain. Integrative and Comparative Biology, 49(3), 254–266.
https://doi.org/10.1093/icb/icp009
Xiang, H., Hu, B., Li, Z., & Li, J. (2014). Dexmedetomidine Controls Systemic Cytokine Levels through
the Cholinergic Anti-inflammatory Pathway. Inflammation, 37(5), 1763–1770.
https://doi.org/10.1007/s10753-014-9906-1
Xie, Z., McAuliffe, S., Swain, C. A., Ward, S. A. P., Crosby, C. A., Zheng, H., … Marcantonio, E. R.
(2013). Cerebrospinal Fluid to Tau Ratio and Postoperative Cognitive Change. Annals of
Surgery, 258(2), 364–369. https://doi.org/10.1097/sla.0b013e318298b077
XIONG, Y., PETERSON, P. L., & LEE, C. P. (1999). Effect ofN-Acetylcysteine on Mitochondrial Function
Following Traumatic Brain Injury in Rats. Journal of Neurotrauma, 16(11), 1067–1082.
https://doi.org/10.1089/neu.1999.16.1067
Xu, Z., Dong, Y., Wang, H., Culley, D. J., Marcantonio, E. R., Crosby, G., … Xie, Z. (2014). Peripheral
Surgical Wounding and Age-Dependent Neuroinflammation in Mice. PLoS ONE, 9(5), e96752.
https://doi.org/10.1371/journal.pone.0096752
Yaksh, T. L. (1985). Pharmacology of spinal adrenergic systems which modulate spinal nociceptive
processing. Pharmacology Biochemistry and Behavior, 22(5), 845–858.
https://doi.org/10.1016/0091-3057(85)90537-4
Ye, S.-M., & Johnson, R. W. (1999). Increased interleukin-6 expression by microglia from brain of
aged mice. Journal of Neuroimmunology, 93(1–2), 139–148. https://doi.org/10.1016/s0165-
5728(98)00217-3
Yirmiya, R., & Goshen, I. (2011). Immune modulation of learning, memory, neural plasticity and
neurogenesis. Brain, Behavior, and Immunity, 25(2), 181–213.
https://doi.org/10.1016/j.bbi.2010.10.015
Yuan, L., Liu, A., Qiao, L., Sheng, B., Xu, M., Li, W., & Chen, D. (2015). The Relationship of CSF and
Plasma Cytokine Levels in HIV Infected Patients with Neurocognitive Impairment. BioMed
Research International, 2015, 1–6. https://doi.org/10.1155/2015/506872
Zaccara, G., Gangemi, P., Perucca, P., & Specchio, L. (2011). The adverse event profile of pregabalin:
A systematic review and meta-analysis of randomized controlled trials. Epilepsia, 52(4), 826–
836. https://doi.org/10.1111/j.1528-1167.2010.02966.x
Zgaia, A. O., Irimie, A., Sandesc, D., Vlad, C., Lisencu, C., Rogobete, A., & Achimas-Cadariu, P. (2015).
THE ROLE OF KETAMINE IN THE TREATMENT OF CHRONIC CANCER PAIN. Clujul Medical,
88(4), 457. https://doi.org/10.15386/cjmed-500
Zhang, T., Tian, X., Wang, Q., Tong, Y., Wang, H., Li, Z., … Chui, D. (2015). Surgical stress induced
depressive and anxiety like behavior are improved by dapsone via modulating NADPH
oxidase level. Neuroscience Letters, 585, 103–108.
https://doi.org/10.1016/j.neulet.2014.11.045
Zheng, W., Cai, D.-B., Yang, X.-H., Ungvari, G. S., Ng, C. H., Müller, N., … Xiang, Y.-T. (2017).
Adjunctive celecoxib for schizophrenia: A meta-analysis of randomized, double-blind,
placebo-controlled trials. Journal of Psychiatric Research, 92, 139–146.
https://doi.org/10.1016/j.jpsychires.2017.04.004
Zheng, X., Liang, Y., Kang, A., Ma, S.-J., Xing, L., Zhou, Y.-Y., … Hao, H.-P. (2014). Peripheral
immunomodulation with ginsenoside Rg1 ameliorates neuroinflammation-induced
behavioral deficits in rats. Neuroscience, 256, 210–222.
https://doi.org/10.1016/j.neuroscience.2013.10.023
Zhu, X., Li, P., Hao, X., Wei, K., Min, S., Luo, J., … Jin, J. (2015). Ketamine-mediated alleviation of
electroconvulsive shock-induced memory impairment is associated with the regulation of
neuroinflammation and soluble amyloid-beta peptide in depressive-like rats. Neuroscience
Letters, 599, 32–37. https://doi.org/10.1016/j.neulet.2015.05.022
Zhu, X.-M., Yao, Y.-M., Liang, H.-P., Liu, F., Dong, N., Yu, Y., & Sheng, Z.-Y. (2009). Effect of high
mobility group box-1 protein on apoptosis of peritoneal macrophages. Archives of
Biochemistry and Biophysics, 492(1), 54–61. https://doi.org/10.1016/j.abb.2009.09.016
Zhu, Y.-Z., Yao, R., Zhang, Z., Xu, H., & Wang, L.-W. (2016). Parecoxib prevents early postoperative
cognitive dysfunction in elderly patients undergoing total knee arthroplasty. Medicine,
95(28), e4082. https://doi.org/10.1097/md.0000000000004082
Zurek, A. A., Yu, J., Wang, D.-S., Haffey, S. C., Bridgwater, E. M., Penna, A., … Orser, B. A. (2014).
Sustained increase in α5GABAA receptor function impairs memory after anesthesia. Journal
of Clinical Investigation, 124(12), 5437–5441. https://doi.org/10.1172/jci76669
Figure 1 - Flowchart of POCD
Positive
feedback
loop
... Although the pathophysiology of POD is not fully understood until now, there is increasing evidence which indicated that the blood-brain barrier damage, CSF biomarkers, neuroinflammation and so on play a key part in POD occurrence [14][15][16][17]. Among these mechanisms, the relationships between POD and CSF biomarkers, such as β-amyloid42 (Aβ42), total tau (T-tau), and phosphorylated tau (P-tau), have been extensively studied [18,19]. ...
Article
Full-text available
Background Postoperative delirium (POD) often occurs in elderly patients after surgery. We conducted two clinical studies to determine whether COVID-19 vaccination has a protective effect on POD and to explore the role of CSF biomarkers in this process. Methods We conducted two clinical studies, Perioperative Neurocognitive Disorder Risk Factor and Prognosis (PNDRFAP) and Perioperative Neurocognitive Disorder and Biomarker Lifestyle (PNDABLE), in which patients more than or equal to 65 years old who have had elective non-cardiac surgery were enrolled. The preoperative cognitive status of patients were evaluated by Mini-Mental State Examination (MMSE) one day preoperatively. Confusion Assessment Method (CAM) was used to diagnose POD. We used the mediation model to analyze the relationship between CSF biomarkers, COVID-19 vaccination and POD, as well as Dynamic Nomogram to calculate the incidence of Non-Postoperative Delirium (NPOD). The main outcome of these studies was the incidence of POD during seven days postoperatively or before discharge, which was assessed by CAM. Results In the final, 705 participants were enrolled in the PNDRFAP study, and 638 patients in the PNDABLE. In both studies, we found that the occurrence of POD was lower in patients who had injected COVID-19 vaccination before surgery compared with those without vaccination (PNDRFAP: 10.20 % [21/205] vs 25.80 % [129/500], P < 0.001; PNDABLE: 2.40 % [4/164] vs 34.60 % [164/474], P < 0.001). Mediation analysis showed that the protective effect of preoperative COVID-19 vaccine on POD was significantly mediated by CSF Aβ42 (proportion = 17.56 %), T-tau (proportion = 19.64 %), Aβ42/T-tau (proportion = 29.67 %), and Aβ42/P-tau (proportion = 12.26 %). Conclusions COVID-19 vaccine is a protective factor for POD in old patients, which is associated with CSF biomarkers.
... The incidence of PND ranged from 41 % to 75 % on POD7 and from 18 % to 45 % on the 3rd month after operation. PND is associated with suboptimal functional recovery and increased mortality following major surgery [2,3]. Surgical interventions can trigger acute systemic inflammation, subsequently leading to neuroinflammation and synaptic dysfunction, thereby contributing to hippocampal-dependent cognitive impairment [4][5][6]. ...
Article
Full-text available
Background Elderly patients undergoing surgery are prone to cognitive decline known as perioperative neurocognitive disorders (PND). Several studies have shown that the microglial activation and the decrease of short-chain fatty acids (SCFAs) in gut induced by surgery may be related to the pathogenesis of PND. The purpose of this study was to determine whether microglia and short-chain fatty acids were involved in cognitive dysfunction in aged rats. Methods Male wild-type Wistar rats aged 11–12 months were randomly divided into control group (Ctrl: Veh group), propionic acid group (Ctrl: PA group), exploratory laparotomy group (LP: Veh group) and propionic acid + exploratory laparotomy group (LP: PA group) according to whether exploratory laparotomy (LP) or PA pretreatment for 21 days was performed. The motor ability of the rats was evaluated by open field test on postoperative day 3 (POD3), and then the cognitive function was evaluated by Y-maze test and fear conditioning test. The expression of IL-1β, IL-6, RORγt and IL-17A mRNA in hippocampus was detected by RT-qPCR, the expression of IL-17A and IL-17RA in hippocampus was detected by Western blot, and the activation of microglia was detected by immunofluorescence. Results The PND rat model was successfully established by laparotomy. Compared with Ctrl: Veh group, the body weight of LP: Veh group decreased, the percentage of spontaneous alternations in Y maze decreased (P < 0.001), and the percentage of freezing time in contextual fear test decreased (P < 0.001). Surgery triggers neuroinflammation, manifested as the elevated levels of the inflammatory cytokines IL-1β (P < 0.001) and IL-6 (P < 0.001), the increased expression of the transcription factor RORγt (P = 0.0181, POD1; P = 0.0073, POD5)and major inflammatory cytokines IL-17A (P = 0.0215, POD1; P = 0.0071, POD5), and the increased average fluorescence intensity of Iba1 (P < 0.001, POD1; P < 0.001, POD5). After PA preconditioning, the recovery of rats in LP: PA group was faster than that in LP: Veh group as the body weight lost on POD1 (P = 0.0148) was close to the baseline level on POD5 (P = 0.1846), and they performed better in behavioral tests. The levels of IL-1β (P < 0.001) and IL-6 (P = 0.0035) inflammatory factors in hippocampus decreased on POD1 and the average fluorescence intensity of Iba1 decreased (P = 0.0024, POD1; P < 0.001, POD5), representing the neuroinflammation was significantly improved. Besides, the levels of RORγt mRNA (P = 0.0231, POD1; P = 0.0251, POD5) and IL-17A mRNA (P = 0.0208, POD1; P = 0.0071, POD5) in hippocampus as well as the expression of IL-17A (P = 0.0057, POD1; P < 0.001, POD5) and IL-17RA (P = 0.0388) decreased. Conclusion PA pretreatment results in reduced postoperative neuroinflammation and improved cognitive function, potentially attributed to the regulatory effects of PA on Th17-mediated immune responses.
... Individualized comprehensive prevention and treatment based on controllable risk factors are the main strategies for preventing and treating POCD in clinical practice [30,32]. These strategies include adjusting the patient's medications before surgery, optimizing the organ function status of elderly patients, and improving metabolic syndrome [33]. It also involves selecting the most appropriate anesthesia and surgical approach and controlling the depth of anesthesia guided by the bispectral index during general anesthesia to avoid burst suppression, maintain hemodynamic stability, and avoid significant fluctuations in circulation [30]. ...
Article
Full-text available
Background Postoperative cognitive dysfunction (POCD) is a common neurological complication of anesthesia and surgery in aging individuals. Neuroinflammation has been identified as a hallmark of POCD. However, safe and effective treatments of POCD are still lacking. Itaconate is an immunoregulatory metabolite derived from the tricarboxylic acid cycle that exerts anti-inflammatory effects by activating the nuclear factor erythroid 2-related factor 2 (Nrf2) pathway. In this study, we investigated the effects and underlying mechanism of 4-octyl itaconate (OI), a cell-permeable itaconate derivative, on POCD in aged mice. Methods A POCD animal model was established by performing aseptic laparotomy in 18-month-old male C57BL/6 mice under isoflurane anesthesia while maintaining spontaneous ventilation. OI was intraperitoneally injected into the mice after surgery. Primary microglia and neurons were isolated and treated to lipopolysaccharide (LPS), isoflurane, and OI. Cognitive function, neuroinflammatory responses, as well as levels of gut microbiota and their metabolites were evaluated. To determine the mechanisms underlying the therapeutic effects of OI in POCD, ML385, an antagonist of Nrf2, was administered intraperitoneally. Cognitive function, neuroinflammatory responses, endogenous neurogenesis, neuronal apoptosis, and Nrf2/extracellular signal-related kinases (ERK) signaling pathway were evaluated. Results Our findings revealed that OI treatment significantly alleviated anesthesia/surgery-induced cognitive impairment, concomitant with reduced levels of the neuroinflammatory cytokines IL-1β and IL-6, as well as suppressed activation of microglia and astrocytes in the hippocampus. Similarly, OI treatment inhibited the expression of IL-1β and IL-6 in LPS and isoflurane-induced primary microglia in vitro. Intraperitoneal administration of OI led to alterations in the gut microbiota and promoted the production of microbiota-derived metabolites associated with neurogenesis. We further confirmed that OI promoted endogenous neurogenesis and inhibited neuronal apoptosis in the hippocampal dentate gyrus of aged mice. Mechanistically, we observed a decrease in Nrf2 expression in hippocampal neurons both in vitro and in vivo, which was reversed by OI treatment. We found that Nrf2 was required for OI treatment to inhibit neuroinflammation in POCD. The enhanced POCD recovery and promotion of neurogenesis triggered by OI exposure were, at least partially, mediated by the activation of the Nrf2/ERK signaling pathway. Conclusions Our findings demonstrate that OI can attenuate anesthesia/surgery-induced cognitive impairment by stabilizing the gut microbiota and activating Nrf2 signaling to restrict neuroinflammation and promote neurogenesis. Boosting endogenous itaconate or supplementation with exogenous itaconate derivatives may represent novel strategies for the treatment of POCD.
Article
Full-text available
Objectives We hope to offer a comprehensive understanding of the advancements and patterns in research on PND. Methods: We performed a thorough search on the Web of Science Core Collection to locate relevant studies published from 1969 to 2022 and utilized four distinct tools, namely VOSviewer (J Data Inf Sci, 2017, 2, 1; J Am Soc Inf Sci, 1973, 24, 265; Amer Doc, 1963, 14, 10 and Scientometrics, 2010, 82, 581), CiteSpace (Scientometrics, 2010, 84, 523), Scimago Graphica, and R‐bibliometrix which allowed us to examine various aspects. Results: We included a total of 6787 articles and reviews for analysis which described PND research, the sources, and the subfields; highlighted the significant developments in this field; identified three main directions in PND.Conclusion: This study highlights the rapid growth of research on PND in recent years and provided an overview of previous studies in the field of PND, thereby establishing the overall landscape of PND research and identifying potential avenues for future investigations. Methods We performed a thorough search on the Web of Science Core Collection to locate relevant studies published from 1969 to 2022. To perform bibliometric analysis and network visualization, we utilized four distinct tools, namely VOSviewer (J Data Inf Sci, 2017, 2, 1; J Am Soc Inf Sci, 1973, 24, 265; Amer Doc, 1963, 14, 10 and Scientometrics, 2010, 82, 581), CiteSpace (Scientometrics, 2010, 84, 523), Scimago Graphica, and R‐bibliometrix. These tools allowed us to examine various aspects, including the yearly publication output, the contribution of different countries or regions, the involvement of active journals, co‐citation analysis, publication status, keywords, and terms, as well as scientific categories. We hope to offer a comprehensive understanding of the advancements and patterns in research on PND. The insights gained from this study can assist researchers and clinicians in enhancing the management and implementation of their work in this field. Results In this study, we included a total of 6787 articles and reviews for analysis. First, publication trends and contribution by country analysis described PND research. Second, a historical analysis described PND research, the sources, and the subfields. Third, an analysis of keywords highlighted the significant developments in this field. Fourth, an analysis of research themes identified three main directions in PND. Conclusion In summary, the research volume exhibits exponential growth over time. Furthermore, the majority of contributions originate from Western countries and China. The interdisciplinary nature of the field is evident, with its roots in biology and medicine and further branching into psychology and social sciences. POCD, delirium‐predominant associated clinical management were major research themes about PND.
Article
Purpose The Boston naming test (BNT), as a simple, fast, and easily administered neuropsychological test, was demonstrated to be useful in detecting language function. In this study, BNT was investigated whether it could be a screening tool for early postoperative cognitive dysfunction (POCD). Methods This prospective observational cohort study included 132 major noncardiac surgery patients and 81 nonsurgical controls. All participants underwent a mini-mental state examination (MMSE) and BNT 1 day before and 7 days after surgery. Early POCD was assessed by reliable change index and control group results. Results Seven days after surgery, among 132 patients, POCD was detected in 30 (22.7%) patients (95% CI, 15.5%-30.0%) based on MMSE, and 45 (34.1%) patients (95% CI, 26.3%-41.9%) were found with postoperative language function decline based on BNT and MMSE. Agreement between the BNT spontaneous naming and MMSE total scoring was moderate (Kappa .523), and the sensitivity of BNT spontaneous naming for detecting early POCD was .767. Further analysis showed that areas under receiver operating characteristics curves (AUC) did not show statistically significant differences when BNT spontaneous naming (AUC .862) was compared with MMSE language functional subtests (AUC .889), or non-language functional subtests (AUC .933). Conclusion This study indicates the feasibility of implementing the BNT spontaneous naming test to screen early POCD in elderly patients after major noncardiac surgery.
Article
Full-text available
Postoperative cognitive dysfunction (POCD) occurs after surgery and severely impairs patients’ quality of life. Finding POCD-associated variables can aid in its diagnosis and prognostication. POCD is associated with noncoding RNAs, such as microRNAs (miRNAs), involved in metabolic function, immune response alteration, and cognitive ability impairment; however, the underlying mechanisms remain unclear. The aim of this study was to investigate hub miRNAs (i.e., miRNAs that have an important regulatory role in diseases) regulating postoperative cognitive function and the associated mechanisms. Hub miRNAs were identified by bioinformatics, and their expression in mouse hippocampus tissues was determined using real-time quantitative polymerase chain reaction. Hub miRNAs were overexpressed or knocked down in cell and animal models to test their effects on neuroinflammation and postoperative cognitive function. Six differentially expressed hub miRNAs were identified. miR-206-3p was the only broadly conserved miRNA, and it was used in follow-up studies and animal experiments. Its inhibitors reduced the release of proinflammatory cytokines in BV-2 microglia by regulating its target gene, brain-derived neurotrophic factor (BDNF), and the downstream signaling pathways. miR-206-3p inhibition suppressed microglial activation in the hippocampi of mice and improved learning and cognitive decline. Therefore, miR-206-3p significantly affects POCD, implying its potential as a therapeutic target.
Article
Full-text available
With the aging of the population, the health of the elderly has become increasingly important. Postoperative cognitive dysfunction (POCD) is a common neurological complication in elderly patients following general anesthesia or surgery. It is characterized by cognitive decline that may persist for weeks, months, or even longer. Electroacupuncture (EA), a novel therapy that combines physical nerve stimulation with acupuncture treatment from traditional Chinese medicine, holds potential as a therapeutic intervention for preventing and treating POCD, particularly in elderly patients. Although the beneficial effects of EA on POCD have been explored in preclinical and clinical studies, the reliability of EA is limited by methodological shortcomings, and the underlying mechanisms remain largely unexplored. Therefore, we have synthesized existing evidence and proposed potential biological mechanisms underlying the effects of EA on neuroinflammation, oxidative stress, autophagy, the microbiota-gut-brain axis, and epigenetic modification. This review summarizes recent advances in EA and POCD, provides a theoretical foundation, explores potential molecular mechanisms for the prevention and treatment of POCD, and offers a basis for conducting relevant clinical trials.
Article
Full-text available
Background: Some degree of cognitive decline after surgery occurs in as many as one quarter of elderly surgical patients, and this decline is associated with increased morbidity and mortality. Cognition may be affected across a range of domains, including memory, psychomotor skills, and executive function. Whilst the exact mechanisms of cognitive change after surgery are not precisely known, oxidative stress and subsequent neuroinflammation have been implicated. N-acetylcysteine (NAC) acts via multiple interrelated mechanisms to influence oxidative homeostasis, neuronal transmission, and inflammation. NAC has been shown to reduce oxidative stress and inflammation in both human and animal models. There is clinical evidence to suggest that NAC may be beneficial in preventing the cognitive decline associated with both acute physiological insults and dementia-related disorders. To date, no trials have examined perioperative NAC as a potential moderator of postoperative cognitive changes in the noncardiac surgery setting. Methods and design: This is a single-centre, randomised, double-blind, placebo-controlled clinical trial, with a between-group, repeated-measures, longitudinal design. The study will recruit 370 noncardiac surgical patients at the University Hospital Geelong, aged 60 years or older. Participants are randomly assigned to receive either NAC or placebo (1:1 ratio), and groups are stratified by age and surgery type. Participants undergo a series of neuropsychological tests prior to surgery, 7 days, 3 months, and 12 months post surgery. It is hypothesised that the perioperative administration of NAC will reduce the degree of postoperative cognitive changes at early and long-term follow-up, as measured by changes on individual measures of the neurocognitive battery, when compared with placebo. Serum samples are taken on the day of surgery and on day 2 post surgery to quantitate any changes in levels of biomarkers of inflammation and oxidative stress. Discussion: The PANACEA trial aims to examine the potential efficacy of perioperative NAC to reduce the severity of postoperative cognitive dysfunction in an elderly, noncardiac surgery population. This is an entirely novel approach to the prevention of postoperative cognitive dysfunction and will have high impact and translatable outcomes if NAC is found to be beneficial. Trial registration: The PANACEA trial has been registered with the Therapeutic Goods Administration, and the Australian New Zealand Clinical Trials Registry: ACTRN12614000411640 ; registered on 15 April 2014.
Article
Full-text available
Background: Trial design neuroinflammation and postoperative pain after surgery are increasingly reported in association with postoperative cognitive dysfunction (POCD). Parecoxib, a selective cyclooxygenase (COX)-2 inhibitor, is used for postoperative analgesia for its potent anti-inflammatory and analgesic effects. This study aimed to evaluate parecoxib's effects on POCD in elderly patients undergoing total knee arthroplasty. Methods: Around 134 elderly patients undergoing total knee arthroplasty were randomly divided into parecoxib (group P) and control (group C) groups, and treated with parecoxib sodium and saline, respectively, shortly after induction of general anesthesia and 12-h postsurgery, respectively. Perioperative plasma IL-1β, IL-6, TNF-α, and C-reactive protein (CRP) 1evels were measured. Postoperative pain was assessed following surgery. Neuropsychological tests were performed before surgery, and 1 week and 3 months postoperation. Results: POCD incidence in group P was significantly lower compared with that of group C at 1 week after surgery (16.7% vs 33.9%; P < 0.05); no significant difference was found between groups C and P at 3-month follow-up (9.7% vs 6.7%). Compared with group C values, visual analog pain scale (VAS) scores at 3, 6, and 12 hours after surgery were significantly lower in group P(P < 0.05). Plasma IL-1β, IL-6, and TNF-α levels were lower in group P than in group C after the operation (P < 0.05). No significant difference in the plasma CRP level was found between groups P and C. Conclusions: Parecoxib sodium decreases POCD incidence after total knee arthroplasty in elderly patients and may explain how this drug suppresses inflammation and acute postoperative pain caused by surgical trauma.
Article
This is the protocol for a review and there is no abstract. The objectives are as follows: To assess the benefit and harm of NSAIDs in controlling disease activity, symptoms and radiographic progression in patients with axial SpA.
Article
Neuroinflammation has been implicated in the neurobiological pathways of schizophrenia. This meta-analysis evaluated the efficacy and tolerability of adjunctive celecoxib as a noncompetitive anti-inflammation drug in treating schizophrenia. Data were searched, extracted, checked and entered into the RevMan (version 5.3) software by two independent investigators. Standardized/weighted mean differences (SMDs/WMDs), risk ratio (RR) and their 95% confidence intervals (CIs) were calculated, as appropriate. Included were 8 randomized controlled trials (RCTs) with a total of 626 patients with schizophrenia including 316 (50.5%) treated on celecoxib (400 mg/day) and 310 (49.5%) on placebo over 8.3 ± 2.3 weeks of treatment. Adjunctive celecoxib outperformed placebo with respect to total psychopathology [3 RCTs, n = 180; SMD: −0.47; 95%CI: −0.81, −0.14; P = 0.005; I² = 18%; 'moderate quality'], symptoms positive [3 RCTs, n = 180; SMD: −0.50; 95%CI: −0.79, −0.20; P = 0.001; I² = 0%; 'moderate quality'], negative symptoms [3 RCTs, n = 180; SMD: −0.32; 95%CI: −0.66, 0.02; P = 0.06; I² = 22%; 'moderate quality'], and general psychopathology scores [3 RCTs, n = 180; SMD: −0.35; 95%CI: −0.65, −0.06; P = 0.02; I² = 0%; 'moderate quality'] in first-episode, but not chronic patients. Additionally, superiority of celecoxib for the Positive and Negative Syndrome Scale (PANSS) total scores was moderated by higher PANSS positive scores and lower PANSS negative scores at baseline. All-cause discontinuation [RR: 1.02; (95%CI: 0.56, 1.86); P = 0.94; I² = 0%] and adverse drug reactions were similar between the two groups. Adjunctive celecoxib appears to be an efficacious and safe treatment in improving psychotic symptoms, particularly in first-episode schizophrenia. Review registration CRD42016054233 (http://www.crd.york.ac.uk/prospero/).
Article
Interleukin-1 beta (IL-1beta), known to play a role in orchestrating the physiological and behavioral adjustments that occur during sickness, has also been shown to significantly influence memory consolidation. To support this assertion we present neurobiological evidence that the substrates for IL-1beta to influence memory processing and neural plasticity exist. We then present behavioral evidence that central IL-1beta administration and agents that induce central IL-1beta activity impair the consolidation of memories that depend on the hippocampal formation but have no effect on the consolidation of hippocampal-independent memories. Further, we demonstrate that the impairments in hippocampal-dependent memory consolidation produced by agents that induce IL-1beta activity are blocked by antagonizing the actions of IL-1beta. Finally, we discuss these data in terms of their implications for a physiological role of IL-1beta in memory consolidation processes and a potential role of IL-1beta in producing memory impairments associated with stress, aging, Alzheimer's disease, and AIDS related dementia complex.
Article
The implication of neuroinflammation in schizophrenia, sustained by recent genetic evidence, represents one of the most exciting topics in schizophrenia research. Drugs which inhibit microglia activation, especially the classical tetracycline antibiotic minocycline are currently under investigation as alternative antipsychotics. However, recent studies demonstrated that microglia activation is not only a hallmark of neuroinflammation, but plays important roles during brain development. Inhibition of microglia activation by minocycline was shown to induce extensive neuronal cell death and to impair subventricular zone (SVZ) neurogenesis and synaptic pruning in the early postnatal and adolescent rodent brain, respectively. These deleterious effects contrast with the neuroprotective actions of minocycline at adult stages. They are of potential importance for schizophrenia, since minocycline triggers similar pro-apoptotic effects in the developing brain as NMDA receptor (NMDAR) antagonists, known to induce long-term schizophrenia-like abnormalities. Moreover, altered postnatal neurogenesis, recently described in the human striatum, was proposed to induce striatal dopamine dysregulation associated with schizophrenia. Finally, the effect of minocycline on synapse remodeling is of interest considering the recently reported strong genetic association of the pruning-regulating complement factor gene C4A with schizophrenia. This raises the exciting possibility that in conditions of hyperactive synaptic pruning, as supposed in schizophrenia, the inhibitory action of minocycline turns into a beneficial effect, with relevance for early therapeutic interventions. Altogether, these data support a differential view on microglia activation and its inhibition. Further studies are needed to clarify the relevance of these results for the pathogenesis of schizophrenia and the use of minocycline as antipsychotic drug.
Article
Substantial evidence indicates involvement of microglia/macrophages in chronic neuropathic pain. However, the temporal-spatial features of microglial/macrophage activation and their pain-bound roles remain elusive. Here, we evaluated microglia/macrophages and the subtypes in the lumbar spinal cord (SC) and prefrontal cortex (PFC), and analgesic-anxiolytic effect of minocycline at different stages following spared nerve injury (SNI) in rats. While SNI enhanced the number of spinal microglia/macrophages since post-operative day (POD)3, pro-inflammatory MHCII+ spinal microglia/macrophages were unexpectedly less abundant in SNI rats than shams on POD21. By contrast, less abundant anti-inflammatory CD172a (SIRPα)+ microglia/macrophages were found in the PFC of SNI rats. Interestingly in naïve rats, microglial/macrophage expression of CD11b/c, MHCII and MHCII+/CD172a+ ratio were higher in the SC than the cortex. Consistently, multiple immune genes involved in anti-inflammation, phagocytosis, complement activation and M2 microglial/macrophage polarization were upregulated in the spinal dorsal horn and dorsal root ganglia but downregulated in the PFC of SNI rats. Furthermore, daily intrathecal minocycline treatment starting from POD0 for two weeks alleviated mechanical allodynia most robustly before POD3 and attenuated anxiety on POD9. Although minocycline dampened spinal MHCII+ microglia/macrophages until POD13, it failed to do so on cortical microglia/macrophages, indicating that dampening only spinal inflammation may not be enough to alleviate centralized pain at chronic stage. Taken together, our data provide the first evidence that basal microglial/macrophage traits underlie differential region-specific responses to SNI and minocycline treatment, and suggest that drug treatment efficiently targeting not only spinal but also brain inflammation may be more effective in treating chronic neuropathic pain.
Article
1-aminopropyl-3-methylimidazolium Bromide ([APMim]Br) exhibits a good selectivity for carbon dioxide (CO2) absorption and its aqueous solutions can have a great prospect for CO2 capture. In order to reveal the Effects of water concentration on the CO2 solubility in the aqueous solutions, the experiments based on an isochoric technique by using a high-pressure quartz glass cell were performed to test the solutions of 0.5590, 0.6578, 0.7680 and 0.8576 water mass fractions. The equilibrium data for CO2 solubility in these aqueous solutions were obtained in range of 5~75℃ and 0.1~7.6 MPa. The results indicated that the chemical absorption mechanism is predominant at lower pressure while the CO2 solubility can be doubled comparing to the chemical one when the water concentration increased and the physical absorption ability prevailed. In 0.65~0.85 water mass fraction, the solutions have an excellent CO2 absorption and release ability in considerable large temperature and pressure ranges and may have a promised future for engineering application.