Вы находитесь на странице: 1из 88

Digital Comprehensive Summaries of Uppsala Dissertations

from the Faculty of Medicine 1089

Hyperglycemia in Experimental
Cerebral Ischemia

MARIA MOLNAR

ACTA
UNIVERSITATIS
UPSALIENSIS ISSN 1651-6206
ISBN 978-91-554-9216-8
UPPSALA urn:nbn:se:uu:diva-247763
2015
Dissertation presented at Uppsala University to be publicly examined in Enghoffsalen,
Entrance 50, Akademiska Sjukhuset, Uppsala, Thursday, 28 May 2015 at 13:00 for the
degree of Doctor of Philosophy (Faculty of Medicine). The examination will be conducted
in Swedish. Faculty examiner: Professor emeritus Sten Lindahl (Department of Physiology
and Pharmacology, Karolinska Institute).

Abstract
Molnar, M. 2015. Hyperglycemia in Experimental Cerebral Ischemia. Digital Comprehensive
Summaries of Uppsala Dissertations from the Faculty of Medicine 1089. 86 pp. Uppsala: Acta
Universitatis Upsaliensis. ISBN 978-91-554-9216-8.

Cerebral ischemia is a life-threatening condition associated with a substantial morbidity and


mortality. Hyperglycemia, a common coexisting phenomenon in both stroke and cardiac arrest
(CA), may further aggravate ischemic brain injury. To date, the therapeutic possibilities are lim-
ited and the search for new treatment modalities is warranted. One aspect of such a research
could be to better understand the cerebral pathogenesis induced by hyperglycemic ischemia-
reperfusion.
We investigated the combination of ischemia and hyperglycemia in two experimental models
of stroke and CA. The aims were to test the neuroprotective potential of the sulfonated nitrone
2-sulfophenyl-N-tert-butylnitrone (S-PBN) in focal hyperglycemic cerebral ischemia (1), to
outline the short-terms effects of hyperglycemia in prolonged (2) and short CA (3) and to
performed a global transcriptome analysis of brain from hyperglycemic and normoglycemic
CA (4).
In a stroke model rats were made hyperglycemic prior to transient middle cerebral artery oc-
clusion and randomized to S-PBN or saline. We found that S-PBN may ameliorate hyperglyce-
mic-ischemic brain damage by improving the neurological performance after 1 day of survival,
but did not reduce the infarct size.
To study the cerebral oxidative state and perfusion after CA, pigs were randomized and
clamped at blood glucose levels of 8.5 ̶ 10.0 mmol/L (high) and 4.0 ̶ 5.5 mmol/L (normal), sub-
jected to 12 ̶ min of CA, followed by 8 min of cardiopulmonary resuscitation (CPR), and ob-
served for 180 min.
Increased oxygenation was found at higher glucose levels measured by near-infrared light
spec-troscopy after CA. Tendencies toward increased protein S100β and 15-keto-dihydro-
prostaglandin F2α were observed in the hyperglycemic group.
We hypothesized that in combination with a brief period of CA, the preischemic
hyperglycemia would worsen the cerebral injury compared with normoglycemia. We used a
glycemic protocol similar to that in Paper II, whereby pigs were subjected to 5 ̶ min of CA,
followed by 8 min of CPR, and observed for 180 mins. An increased level of the cerebral marker
S100β was found in hyperglycemic pigs compared with normoglycemic pigs after CA.
Global transcriptome analysis using microarray analysis revealed a different early metabolic
gene expression in hyperglycemic CA compared with normoglycemic CA.

Keywords: brain ischemia, cardiac arrest, cytokines, gene expression, glucose, hyperglycemia,
microarray, oxidative stress, pigs, rats, reperfusion, resuscitation S100β

Maria Molnar, Department of Surgical Sciences, Anaesthesiology and Intensive Care,


Akademiska sjukhuset, Uppsala University, SE-75185 Uppsala, Sweden.

© Maria Molnar 2015

ISSN 1651-6206
ISBN 978-91-554-9216-8
urn:nbn:se:uu:diva-247763 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-247763)
In memory of my beloved mother and to my family
List of Papers

This thesis is based on the following papers, which are referred to in the text
by their Roman numerals.

I Molnar M, Lennmyr L. Neuroprotection by S-PBN in


hyperglycemic ischemic injury in rats. Upsala Journal of
Medical Sciences 2010;115:163-168

II Lennmyr F; Molnar M, Basu S, Wiklund L. Cerebral effects of


hyperglycemia in experimental cardiac arrest. Crit Care Med
2010;38:1726-32

III Molnar M, Bergquist M, Larsson A, Wiklund L, Lennmyr F.


Hyperglycemia increases S100β after short experimental
cardiac arrest. Acta Anaesthesiol Scand 2014;58: 106-113

IV Molnar M*, Lindblom R*, Israelsson Ch, Fridman B, Wiklund


L, Lennmyr F. Differential regulation of cerebral metabolic
genes after hyperglycemic and normoglycemic cardiac arrest.
Manuscript
*Equal contribution

Paper I, II and III are printed with permission from the respective publishers.
Contents

Introduction ................................................................................................... 11 


Cerebral ischemia ..................................................................................... 11 
Background .......................................................................................... 11 
Pathogenesis of cerebral ischemia ....................................................... 12 
Cerebral ischemia and cell signal pathways.................................... 14 
Cerebral ischemia and free radicals ................................................ 14 
Hyperglycemia as a risk factor in focal cerebral ischemia .................. 16 
Pathogenesis of transient hyperglycemic cerebral ischemia ................ 17 
Acidosis and additional mechanisms .............................................. 17 
Oxidative stress and hyperglycemia .................................................... 18 
Hyperglycemia as a prognostic factor after cardiac arrest ................... 19 
Neuroprotection with nitrones ............................................................. 21 
Aims .............................................................................................................. 22 
Material and Methods ................................................................................... 23 
Animals ................................................................................................ 23 
Experimental models of focal and global cerebral ischemia ............... 23 
Models of focal cerebral ischemia .................................................. 23 
Model of global cerebral ischemia .................................................. 24 
Middle cerebral artery occlusion (MCAO) (Paper I)........................... 24 
Neurological testing (Paper I) .............................................................. 25 
Triphenyltetrazolium (TTC) staining (Paper I) ................................... 25 
Cardiac arrest (Paper II and Paper III) ................................................. 26 
Cerebral oxygenation (Paper II and Paper III)..................................... 27 
Near-infrared spectroscopy (NIRS) ................................................ 27 
Laser Doppler flowmetry (LDF) ..................................................... 28 
Jugular venous oxygen saturation (SjvO2) ...................................... 28 
Isoprostanes and prostaglandins (Paper II) .......................................... 29 
Cerebral and myocardial injury marker (Paper II and Paper III) ......... 29 
Protein S100β .................................................................................. 29 
Cytokines IL-6 and TNF (Paper III) .................................................... 30 
Biochemical analysis (Paper II and Paper III) ..................................... 30 
RNA preparation (Paper IV) ................................................................ 31 
Microarray expression analysis (Paper IV) ......................................... 31 
Statistical analysis................................................................................ 32 
Paper I .................................................................................................. 33 
Paper II and Paper III ........................................................................... 33 
Paper IV ............................................................................................... 33 
Results and discussion .................................................................................. 34 
Paper I ...................................................................................................... 34 
General findings .................................................................................. 34 
The main finding ................................................................................. 35 
Glycemic levels ................................................................................... 35 
Glucose metabolism and cerebral blood flow under ischemic
conditions............................................................................................. 36 
Role of RONS in mismatch of glucose metabolism and cerebral
blood flow ............................................................................................ 36 
Nitrone-based trapping agents ............................................................. 36 
Therapeutic time window of S-PBN.................................................... 38 
Failure of translation of neuroprotection in animal studies into
clinical practice .................................................................................... 38 
Paper II ..................................................................................................... 40 
Resuscitability, blood glucose levels and hemodynamics ................... 40 
rSO2, SjvO2 and HbO2 saturation in the sagittal sinus (SssvO2) .......... 40 
Markers of injury ................................................................................. 40 
Cerebral oxygenation (rSO2, SjvO2, Sssvo2) ....................................... 41 
Oxidative stress and inflammation ...................................................... 43 
Protein S100β ...................................................................................... 45 
Time window and duration of ischemia .............................................. 45 
Paper III .................................................................................................... 46 
Resuscitability, blood glucose and hemodynamics ............................. 46 
SjvO2 and SssvO2 ................................................................................. 46 
Cerebral cortical flow .......................................................................... 46 
IL-6 and TNF ....................................................................................... 46 
S100β ................................................................................................... 46 
The main finding ................................................................................. 47 
Cytokines IL-6 and TNF ...................................................................... 48 
Glycemic levels ................................................................................... 50 
Paper IV ................................................................................................... 50 
The main finding ................................................................................. 51 
SLC5A11 .............................................................................................. 55 
BDH1 ................................................................................................... 55 
SLCO1A2 ............................................................................................. 56 
CNTN3 ................................................................................................. 57 
MicroRNA ........................................................................................... 57 
Conclusions ................................................................................................... 59 
Conclusive remarks and future perspective .................................................. 60 
Acknowledgements ....................................................................................... 61 
Funding ......................................................................................................... 63 
References ..................................................................................................... 64 
Abbreviations

AEG advanced glycation endproducts


ATP adenosine triphosphate
BBB blood-brain barrier
BDH1 3-hydroxybutyrate dehydrogenase type 1
BP blood pressure
CA cardiac arrest
CBF cerebral blood flow
CNTN3 contactin3
CO cardiac output
COX cyclooxygenase
CPP cerebral perfusion pressure
CPR cardiopulmonary resuscitation
DNA deoxyribonucleic acid
ERK extracellular signal-regulated kinase
FFA free fatty acids
IHCA in-hospital cardiac arrest
IL interleukin
IIT intensive insulin therapy
IPC ischemic preconditioning
I/R ischemia-reperfusion
LDF laser Doppler flowmetry
MAP mean arterial pressure
MAPK mitogen activated protein kinase
MCAO middle cerebral artery occlusion
MMP metalloproteinase
NADPH nicotinamide adenine dinucleotide phosphate
NO nitric oxide
NIRS near-infrared light spectroscopy
NMDA N-methyl-D-aspartate
NOS nitric oxide synthase
NXY-059 disodium 2,4 disulfophenyl-N-tert-butylnitrone
OOHCA out-of-hospital cardiac arrest
PBP α-phenyl N-tert-butylnitrone
PGF prostaglandin F
PKC protein kinase C
PMN polymorphonuclear
PUFA polyunsaturated fatty acid
ROS reactive oxygen species
RONS reactive oxygen and nitrogen species
ROSC restoration of spontaneous circulation
RNA ribonucleic acid
SAINT Stroke Acute Ischemic NXY-059 Study
SLCO1A2 Solute carrier organic anion transporter family,
member 1A2
SLC5A11 Solute carrier family 5, member 11
S-PBN sodium 2-sulfophenyl-N-tert-butylnitrone
TBI traumatic brain injury
TCA tricarboxylic acid
Trop I Troponin I
tPA tissue plasminogen activator
TTC Tri-phenyl-tetrazolium-chloride
XO xanthine oxidase
VF ventricular fibrillation
Introduction

Cerebral ischemia

Background
Ischemic stroke and CA are common primary causes of cerebral ischemia in
humans.
Ischemic stroke is a life-threatening condition in which a specific vascular
territory of brain is affected by either permanent or transient cessation of blood
flow. Stroke is the third leading cause of mortality in the industrialized world1.
Furthermore, it leads to short- and long-term morbidity, psychosocial and so-
cioeconomic consequences.
While primary and secondary preventive efforts have been successful, the
treatment possibilities for the acute phase of ischemic stroke are restricted to
serine protease tissue plasminogen activator (tPA) as the only effective treat-
ment2. Primarily targeting the cerebral vasculature, this neuroprotective mo-
dality can be described as extrinsic or indirect 3.
During focal cerebral ischemia the blood flow is depleted in the center of
the brain region called the ischemic core. In the surrounding region, defined
as the penumbra by Astrup and coworkers4, the circulation is partially pre-
served by residual collateral flow, thereby creating an endogen or direct po-
tential neuroprotective target3.
During CA, the cerebral blood flow (CBF) is affected in the entire brain.
Although reperfusion can restore the CBF, it can also lead to secondary brain
injury, which is the leading cause of death in initial survivors of CA. The ther-
apeutic possibilities are limited and currently rely on induced mild hypother-
mia5.
Hyperglycemia is a commonly coexisting phenomenon both in focal and
global transient ischemia6, 7, associated with increased ischemic brain dam-
age8, 9, particularly during the reperfusion period10. Experimental11, 12 and clin-
ical studies13, 14 have demonstrated the worsening effect of hyperglycemia on
the outcome after cerebral ischemia but its prognostic significance is still con-
troversial14, 15.
This thesis will focus on hyperglycemia and transient experimental cerebral
ischemia, with particular emphasis on neuroprotection in hyperglycemic
stroke and glycemic control in CA.

11
Pathogenesis of cerebral ischemia
High energy demands and the use almost exclusively glucose as a metabolic
fuel are unique characteristics of the brain. Since the endogenous level of this
principal substrate is low in the brain, it is essential that the blood continuously
supply it together with oxygen in order to maintain normal brain metabolism.
Cessation or occlusion of circulation disturbs this equilibrium and initiates a
cascade of molecular events, such as progressive relapse of high energy me-
tabolism, anaerobic glycolysis, lactate acidosis, disruption of membrane elec-
trochemical gradients and cellular influx of calcium/sodium and efflux of po-
tassium leading to increased intracellular water and cytotoxic edema. Together
these events lead to increased intracellular calcium resulting in release of ex-
citatory amine acids, mitochondrial failure, free radical production, accumu-
lation of free fatty acids (FFA), inflammatory processes and activation of gene
expression resulting in cerebral tissue damage1. Furthermore, accumulation of
precursors of oxidative phosphorylation initiates conformational alterations of
enzymes such as xanthine oxidase (XO). These processes in turn play an im-
portant role in increased reactive oxygen species (ROS) production when re-
introduction of oxygen occurs16.
If the ischemia is transient, the reestablished blood flow precipitates further
potentially pathologic events - termed as reperfusion injury by Dietrich17. The
oxygen has a paradox role in this pathogenesis, since both the absence of ox-
ygen during ischemia and the reinstitution of oxygen during reperfusion can
lead to tissue damage. In the latter phase injury occurs mainly as a result of
the production of ROS16.

12
The schematic ischemia/reperfusion (I/R) is summarized in fig.1. (Published
with permission from the publisher.)

Fig. 1. The cascade of biochemical events leading to apoptosis or necrosis following


cerebral ischaemia. Vascular occlusion in a blood vessel initiates a complex signaling
cascade that leads to neuronal cell death. The reduction in blood flow produces ionic
pump failure and anoxic depolarization leading to enhanced glutamate release and a
sudden increase in intracellular calcium. This rise in calcium triggers mitochondrial
collapse, free radical production, cytotoxic oedema, and increased NO generation.
Reperfusion also produces injury by augmenting BBB breakdown, inflammation, and
free radical production leading to apoptosis. Red borders signify important events in
the cascade. The blue border indicates reperfusion. This figure has been adapted
from18. AA, arachidonic acid; AMPA, a-amino-3-hydroxy-5-methyl-4-isoxa-
zolepropionic acid; BBB, blood-brain barrier; iNOS, inducible nitric oxide synthase;
NMDA, N-methyl-D-aspartate; nNOS, neuronal nitric oxide synthase; NO, nitric ox-
ide; PLA2, phospholipase A2.

13
Cerebral ischemia and cell signal pathways
Cerebral ischemia affects cell signaling, a phenomenon previously studied in
our laboratory11, 19.
During ischemia, exitotoxicity of glutamate - the main excitatory neuro-
transmitter - occurs via NMDA receptors leading to increased Ca2+permeabil-
ity and causing activation of Ca2+ ̶ dependent nitric oxide synthases (NOS)
and production of nitric oxide (NO)20. NO is a water- and lipid-soluble free
radical compound and plays an important role in the reactive oxygen and ni-
trogen species (RONS) mediated injury.
The interaction between the NMDA receptor mediated exitotoxicity and
Src family kinases (SFKs) has been studied in cerebral ischemia as one possi-
ble implication of SFK mediated injury mechanisms21. SFKs may participate
in exemplifying inflammatory and free radical induced damage and mediating
vascular endothelial growth factor (VEGF) induced vascular permeability
during ischemic conditions22.
SFKs can be engaged by other pathways to modulate NMDA-receptors,
such as a route involving cell-adhesion kinase-β (CAKβ) providing a link to
the mitogen activated protein kinases (MAPKs) pathway23, which involves a
group of signal transduction proteins regulating cell proliferation, differentia-
tion and survival/apoptosis during physiological conditions. Cerebral ische-
mia provides a presumptive stimulus for activation of the MAPK pathway19.
The principal MAPK pathways include three subgroups, such as extracel-
lular signal-regulated kinase (ERK), c-jun N-terminal kinase (JNK) and p38,
all of which were to be found differentially activated by cerebral ischemia19.
ERK is activated by protein tyrosine phosphorylation through the Ras path-
way in response to reactive oxygen species (ROS) and NO24 and has been
suggested to mediate brain ischemic injury25. Phospho-JNK is implicated in
apoptosis26. Activation of JNK has been reported to induce endothelial prolif-
eration by interacting with the VEGF ̶ induced ERK pathway, and it may also
participate in the angiogenic response in cerebral ischemia27. Activation of
p38 leads to the promotion of expression of inflammatory cytokines28.
Experimental evidence has demonstrated that Src kinases may act as an
intermediate link between MAPK and ROS29 .

Cerebral ischemia and free radicals


Free radicals are comprised of one or more unpaired valence electrons without
intramolecular bonding making these compounds highly reactive. Free radi-
cals play a role as signal transducing molecules in response to extracellular
stimuli under physiological conditions and are involved in various signaling
processes, such as regulation of vascular tone, sensing of oxygen tension, am-
plification of signal transduction from numerous membrane receptors and ox-
idative stress responses (see review30). The major types of free radicals are
ROS and reactive nitrogen species (RNS), together termed RONS. They are

14
counterbalanced by antioxidative enzymes such as superoxide dismutase
(SOD), glutathione peroxidase (GSPx) and catalase, as well as non ̶ enzymatic
compounds. When the production of pro ̶ oxidants exceeds that of anti ̶ oxi-
dants it can lead to harmful consequences termed oxidative stress, which has
been implicated as an underlying mechanism in various brain injury models31.
The brain is highly vulnerable to ROS ̶ mediated injury due to its high uti-
lization rate of oxygen combined with low antioxidant capacity, abundant
amount of polyunsaturated lipids and high content of iron and cooper32.
During the immediate reperfusion period and the ischemic penumbra phase
enhanced production of ROS, such as hydrogen peroxide (H2O2), superoxide
(O2-), hydroxyl radical (OH) occurs induced by a glutamate ̶ mediated in-
crease of intracellular calcium33. As a substrate for these reactive compounds,
NO plays a central role in this process34. NO is produced by the NOS family,
which catalyze the conversion of L-arginine to L-citrulline and NO. This pro-
cess is coupled to the donation of two electrons by nicotinamide adenine di-
nucleotide phosphate (NADPH)33. The chemical biological reactions of NO
can either be classified as direct effects (occurring between NO and specific
biological molecules), or indirect effects mediated by RONS35. NO has dual
role in ischemic neuropathology - beneficial as a potent vasodilator and dele-
terious as a cytotoxic agent34.
The three main types of NOS exhibit increased activity during ischemia33.
The endothelial NOS has a protective effect via NO ̶ mediated vasodilatation
and enhanced penumbral perfusion36, while the overall effects of neuronal and
inducible NOS are deleterious33.
During reduced accessibility of L-arginine or other cofactors, the NOS ac-
tivity is uncoupled from electron donation by NADPH, which in addition to
the accompanied synthesis of NO leads to increased production of O2- and
H2O233. NO and O2- rapidly unite forming peroxynitrite37, which is more toxic
to DNA than either of its precursors38.
Peroxynitrite generates a single-strand breakdown of DNA which activates
the DNA repair protein poly ADP-ribose polymerase (PARP). Over-activation
of PARP is reported as a major contributive factor to brain damage after tran-
sient focal cerebral ischemia33.
During I/R in addition to NOS, deleteriously high levels of ROS may be
derived from cyclooxygenases (COXs), XO, NADPH oxidases and mitochon-
dria, the latter which forms ROS via the electron transport chain (ETC)30, 33.
The elevated level of mitochondrial ROS reduces ETC activity, resulting in
diminished adenosine triphosphate (ATP) leading to further failure of mito-
chondrial function39.
COXs promote the ischemic injury by enhanced ROS generation, since the
oxidative transformation of arachidonic acid results in concomitant O2- pro-
duction31, 40.
XO has also been demonstrated to contribute to I/R injury41. During ische-
mia hypoxanthine accumulates as a result of increased breakdown of adenine

15
nucleotides and inhibition of XO activity42. During reperfusion hypoxanthine
acts as a substrate for .O2- production43,which occurs through the conversation
of hypoxanthine to xanthine by XO. The end product of this process is uric
acid, which has been identified as an important endogenous radical scaven-
ger44, but its production is also an indicator of XO activity43. ROS production
by XO is the major source of oxidative stress during ischemia and reperfu-
sion30.
The alteration in the levels of hypoxanthine, uric acid and allantoin (a prod-
uct of uric acid oxidation) can be used as endogenous markers of free radical
formation in hypoxia/ischemia by using high performance liquid chromatog-
raphy (HPLC), considering that direct measurements of ROS in vivo is diffi-
cult45.
NADPH oxidases are the key enzymes of interest since they represent the
only enzyme family whose exclusive function is to produce ROS46. Almost all
cells possess NADPH oxidase activity, but neutrophils in particular show high
expression of NADPH oxidases. NADPH oxidases has a dual role in I/R in-
jury16. During ischemia, NADPH oxidases are up ̶ regulated by the activation
of hypoxia-induced factor (HIF)-1α47, suggesting an adaptation mechanisms
of the tissue to the decreased substrate availability. During reintroduction of
oxygen, NADPH oxidases act as key mediators in reperfusion brain injury by
being triggers of mitochondrial ROS production which leads to secondary mi-
tochondrial dysfunction16.
Mitochondrial ROS enhance inflammatory response occurring during I/R
and this issue will be discussed later.
ROS can also have direct cytotoxic effects48 and together with inflamma-
tion these events can result in apoptosis or necrosis.

Hyperglycemia as a risk factor in focal cerebral ischemia


The main sources of acute high levels of glucose are diabetes and stress ̶ in-
duced hyperglycemia, e.g. the absence of pre ̶ existing diagnosis of diabetes.
The risk of developing ischemic stroke is reported to be 4-12 times higher
in people with diabetes than the general population49 and diabetic patients ex-
periencing stroke had worse outcomes than those without diabetes50.
Stress ̶ induced admission hyperglycemia is common in acute stroke pa-
tients7, 51 and is an independent predictor of mortality52 and poor functional
recovery10.
Regardless of the causes of high glucose level, the majority of experimental
studies on focal transient ischemia demonstrated that acute hyperglycemia ag-
gravates ischemic brain damage11, 53-55. Furthermore, hyperglycemia acceler-
ates neuronal injury caused by transient cerebral ischemia and so possibly
shortens the therapeutic window56, 57.
The therapeutic window for revascularization to attenuate neurological
damage following ischemic stroke was limited to 3h2 and now can be extended

16
to 4.5 ̶ 6 h58 in selected patients, but might hypothetically be prolonged as long
as 24 h post ̶ stroke for by using direct neuroprotective agents49.

Pathogenesis of transient hyperglycemic cerebral ischemia


Acidosis and additional mechanisms
In addition to accumulation of intracellular calcium and lactate leading to in-
tracellular acidosis, free radical formation, impaired mitochondrial function
and lipid peroxidation have also been emerged as main factors introducing
hyperglycemic ischemic brain injury56.
The effects of these various factors are particularly important when hyper-
glycemia is combined with I/R49.
A cornerstone in the pathophysiology of cerebral ischemia is metabolic ac-
idosis. Under conditions of complete ischemia, lactate formation correlates
with pre ̶ existing tissue stores of glycogen and glucose59. The relationship
between the lactate concentration and the intracellular/extracellular pH
(pHi/pHe) is almost linear60 during ischemia, and under normoglycemic is-
chemic conditions these pH values are 6.4 and 6.7, respectively56.
In hyperglycemic animals a further decrease in intra-ischemic pH was seen
as the correspondent pHi and pHe was more acid during ischemia and short
time after the reintroduction of circulation61. The pHi returned than to normal
values and even shifted above controls, caused possible by the upregulated
activity of the Na+/H+ exchanger62. The pHe also returned to normal levels
after 15 ̶ 30 min of reperfusion. Therefore it is unlikely that hyperglycemia-
aggravated injury can be caused by lactic acidosis61 alone.

Additional mechanisms other than exaggerated acidosis may be involved in


enhanced hyperglycemic cerebral ischemia during reperfusion, for example
prolonged translocation of protein kinase Cγ (PKCγ) to cell membranes and
increased tyrosine- phosphorylation of proteins p120 and p180. Tyrosine
phosphorylation of p180 (known as NMDA receptor subunit NR2A/2B) in-
creases the intracellular calcium levels, which in turn leads to free radical ac-
tivation of Src, and to Ras activation of the ERK pathway in the cytosol63.
During I/R hyperglycemia/acidosis ̶ induced alteration of signal transduc-
tion pathway MAPKs/ERK64 has been identified as a deleterious factor due to
its role in causing increased phosphorylation of ERK 1/2 in the cingulated
cortex, frontoparietal cortex, hippocampal CA3 and dentate gyrus, areas
where neurons are destined to die in hyperglycemia. Phosphorylation of ERK
in dentate gyrus was more pronounced at a later phase of reperfusion, indicat-
ing a biphasic effect of hyperglycemia on cell death by initially facilitating
mitochondrial respiration and conserving mitochondrial energetic status dur-
ing early reperfusion65, but accelerating cell death in the late phase via unclas-
sified mechanisms66.

17
Differential activation of MAPK family members ERK and JNK was found
in focal hyperglycemic cerebral ischemia, where ERK was hyper ̶ activated
on the lesion side and JNK was activated on the contralateral hemisphere in
early phases of reperfusion. One possible mechanism may be ROS accumula-
tion, which in turn can be triggered by accumulation of extracellular glutamate
thereby mediating ERK ̶ dependent neuronal oxidative damage11.

Oxidative stress and hyperglycemia


Acute hyperglycemia leads to increased oxidative stress especially when com-
bined with cerebral I/R.
Superoxide, NO and peroxynitrite are the major sources of oxidative stress
in hyperglycemic cerebral ischemia49 because the concentration of these three
compounds increases during hyperglycemic I/R.
Hyperglycemia leads to increased mitochondrial production of O2-, which
in turn has been implicated as a common element in five main mechanisms
that lead to hyperglycemic ̶ ischemic damage, namely protein kinase C (PKC)
activation, advanced glycation endproducts (AGE), increased expression of
the receptor for AGEs and its activating ligands, increased aldose reductase
and hexosamine pathway fluxes67, 68. Increased glycolysis leads to excessive
NADH processing, which in turn enhances the electron transfer via the ETC
in the mitochondria, causing accumulation of intermediate substrates, such as
ubisemiquinone. The latter promotes formation of O2-, which in turn partially
inhibits glyceraldehyde-3-phosphate dehydrogenase (GAPDH) leading to in-
creased production of glycolytic metabolites for these pathways and resulting
in depletion of antioxidants, thereby causing secondary mitochondrial dys-
function67.
Increased production of O2- can also result in imbalance between NO and
O , which in turn causes endothelial dysfunction69 because hyperglycemia-
2-

induced depletion of endothelium and vascular muscle nicotinamide adenine


dinucleotide phosphate (NADPH) leads to reduced intracellular antioxidant
capacity70, which might render the tissue vulnerable to RONS damage during
I/R.
Much of the deleterious effect of excess NO and O2- production is attributed
to peroxynitrite formation in focal transient hyperglycemic ischemic cerebral
damage71.
The lipid ̶ rich environmental of the brain allows increased lipid peroxida-
tion under certain conditions, such as hyperglycemic ischemia and F2-isopros-
tanes can be used as a marker of oxidative stress72.
Inflammation is another source of oxidative stress in hyperglycemic cere-
bral I/R and its role in this context will be discussed later.
Thus, hyperglycemic-ischemic damage has been associated with increased ox-
idative stress, which occurs especially through the augmented production of
RONS. These reactions in turn are involved in increased enrollment of various

18
blood cells that promotes inflammation. Scavenging of free radicals may
therefore play an important role in strategies to decrease hyperglycemic cere-
bral injury.

Hyperglycemia as a prognostic factor after cardiac arrest


Cerebral dysfunction, ranging from temporary impairment to brain death, is a
well-recognized complication in survivors after CA73. While the two most im-
portant determinants of neurological outcome are duration of CA and that of
cardiopulmonary resuscitation (CPR)74, 75, several investigators have also
pointed out the importance of hyperglycemia in the context of CA and resus-
citation76, 77.
Hyperglycemia is commonly recognized in both diabetic and non ̶ diabetic
patients during the immediate post ̶ restoration of spontaneous circulation
(ROSC) following in-hospital CA (IHCA)15 and out-of-hospital CA
(OOHCA)78.
Studies on IHCA and OOHCA have revealed that derangements of blood
glucose immediately following ROSC6, 79 and during the first few days (12,
24 and 72 h) was found to be independently correlated with patient survival80
and was associated with increased hospital mortality14.
Beiser et al14 demonstrated that decreased survival odds for diabetes pa-
tients could only be detected with extreme hyperglycemia (>13.3 mmol/L).
Non ̶ diabetic patients appeared more sensitive to derangement outside of the
minimum (3.9 ̶ 9.4 mmol/L) and maximum (6.2 ̶ 13.3 mmol/L) glucose values,
assuming that glycemic control may be more important in non ̶ diabetic pa-
tients than diabetic patients.
Others, however, found no association between post CPR glucose levels
and neurologic outcome, although such an effect could not be excluded15.
In another study, no improvement was demonstrated after resuscitation
from ventricular fibrillation (VF) using tight glycemic control81.

Acute hyperglycemia (>6.1 mmol/L) is seen in 90 % of all intensive care unit


(ICU) patients82, probably because of the hypermetabolic state83, which is
most likely a secondary reaction to the activation of counter ̶ regulatory hor-
mone and cytokine responses84. These reactions result in central (increased
hepatic glucose production)85 and peripheral insulin resistance leading to hy-
perglycemia86. A high blood glucose level is augmented further by the admin-
istration of a glucose containing infusion84, sympathomimetic drugs and the
use of corticosteroid87. Stress ̶ induced hyperglycemia is therefore an indicator
of illness severity and was significantly related to increased ICU mortality88.
Aggressive treatment of hyperglycemia was suggested by several clinical
trials - including the Leuven studies - revealing improved mortality and mor-
bidity in critically ill patients89-92. Intensive insulin therapy (IIT) - maintaining
blood glucose levels below 6.1 mmol/L - reduced both mortality from multiple

19
organ failure with sepsis as well as morbidity among critically ill patients in
the cardiac surgical intensive care unit, regardless of whether they had a his-
tory of diabetes90. However, in a subsequent publication the authors proposed
better management of blood glucose rather than increased administration of
insulin93. As a result of the Leuven trials more intensive blood glucose control
was advocated in order to improve the outcome of critically ill patients94.
In contrast, others could not demonstrate reduced mortality and morbidity
after applying intensive intraoperative insulin therapy as a part of strict post-
operative glucose control after cardiac surgery, and the possibility of harm to
patients receiving IIT could not be excluded95.
The concerns raised against IIT as a standard treatment include potential
harm from the increased risks of hypoglycemia, the administration of glucose-
containing infusion and the frequent early use of parenteral nutrition96. Fur-
thermore, hypoglycemia both mild (<4.5 mmol/L) and severe (<2.2 mmol/L)
was independently associated with higher risk of death97.
Therefore, randomized controlled trials were performed to compare IIT
with conventional insulin therapy. Accordingly, the Normoglycemia in Inten-
sive Care Evaluation – Survival Using Glucose Algorithm Regulation (NICE-
SUGAR) study revealed that the intensive glucose control targeted at 4.5 ̶ 6.0
mmol/L resulted in increased mortality among adults in the ICU, while blood
glucose targeted of 10.0 mmol/L or less resulted in lower mortality96. Severe
hypoglycemia (<2.2 mmol/L) was seen significantly more frequent with in-
tensive glucose control96. However, recent meta-analysis revealed that tight
glucose control decreased the rate of septicemia98. In the NICE-SUGAR study
the patients received predominately enteral nutrition - in accordance with the
evidence-based feeding guidelines99 - compared with a substantial proportion
of the patients in the meta-analysis89, 98 who received predominantly parenteral
nutrition. One explanation of discrepancies between the Leuven trials and the
subsequent NICE-SUGAR study could be found in the high rate of use of total
parenteral nutrition82.
There is also a significant association between increased blood glucose var-
iability and worse outcome, suggesting that blood glucose variability is even
more important than hyperglycemia itself at the biological and clinical lev-
els100.
Thus, hyperglycemia is often seen in patients undergoing CA and ICU
treatment, but it is still unclear how and/or whether hyperglycemia can act as
a marker of severity of illness, and how the pathogenic influence of hypergly-
cemia on damage can actually understood. Different strategies exist to coun-
terbalance the effect of hyperglycemia in the context of CA but the impact of
such regimens on the pathogenesis of CA remains unclear101.

20
Neuroprotection with nitrones
Cerebral ischemia triggers a cascade of biochemical events, and within this pro-
cess many molecular targets can be modulated to achieve neuroprotection. A
widely investigated neuroprotective strategy both in experimental and clinical
studies has been the administering of free radical scavengers, such as α-phenyl
N-tert-butylnitrone (PBN), disodium 2,4-disulfophenyl-N-tert-butylnitrone
(NXY-059) and sodium 2-sulfophenyl-N-tert-butylnitrone (S-PBN).
The neuroprotective effect of PBN was first demonstrated in a gerbil model
of global ischemia102, 103. PBN has been shown to be neuroprotective in an ex-
perimental permanent middle cerebral artery occlusion (MCAO) model even if
given 12 h after ischemia104. Pretreatment with PBN reduced the infarct size to
< 50% of control in transient ischemia and this beneficial effect was obtained
even when the drug was given 1 and 3 h following recirculation after 2 h of
MCAO105.
PBN has also been proven to be neuroprotective by reducing the infarct vol-
ume by approximately 70% in a hyperglycemic transient MCAO model54.
NXY-059 (which has similar physicochemical properties to S-PBN) was
found more efficient than PBN in transient MCAO model by reducing infarct
volume and improving neurological deficits. It also had a wide therapeutic win-
dow given 5 ̶ 8 h after the onset of ischemia, which was 3 ̶ 6 h after the start of
reperfusion106.
NXY-059 was found effective in a transient ischemia model in a rodent re-
sulting in dose-dependent neurological improvement, however the correlation
with neurological improvement and decrease in infarct size was not strict107. In
the same study the permanent ischemia group required a higher dose of NXY-
059 for effective neuroprotection than that required to attenuate the damage in
the transient ischemia model, but this dose was well tolerated in Phase II clinical
study of NXY-059108. The therapeutic window of NXY-059 was also found to
be significantly effective given 4 h after the start of permanent MCAO, reveal-
ing a relevant time window used also by clinical practice.
NXY-059 was also neuroprotective given either immediately or 4 h after per-
manent MCAO in a primate model, where aside from the attenuated motor def-
icit and spatial neglect, there was a substantial decrease in the infarct size109, 110.
S-PBN has been found to be neuroprotective in a malonate injection
model111, in experimental transient global ischemia 112, 113, experimental focal
embolic cerebral ischemia114 and in experimental traumatic brain injury
(TBI)115-117.
Thus, these free radical scavengers, called also as spin trapping agents, have
demonstrated neuroprotective ability in various experimental ischemia model
with normoglycemia. However, their efficacy in hyperglycemic-ischemic brain
damage has not been widely explored.

21
Aims

Paper I To investigate the neuroprotective effect of S-PBN on hyper-


glycemic focal ischemia reperfusion.

Paper II To investigate the effects of CA and resuscitation on cerebral


perfusion and oxidative stress during hyperglycemia and
normoglycemia.

Paper III To investigate the effects on cerebral perfusion of 5 min of


CA and resuscitation during hyperglycemia and normoglyce-
mia.

Paper IV To perform a global transcriptome analysis of cerebral cortex


after hyperglycemic and normoglycemic CA, respectively.

22
Material and Methods

Animals
Male-Sprague-Dawley rats were used in Paper I. Triple-breed pigs obtained
from a single provider were utilized in Paper II and Paper III. After termi-
nation, pieces of frontoparietal cortex taken from pigs included in Paper III
were used in Paper IV.

Experimental models of focal and global cerebral ischemia


There are several animal models of focal and global cerebral ischemia, each
with advantages and disadvantages (see review118).

Models of focal cerebral ischemia


The use of rodent models of focal cerebral ischemia is widespread and has
several advantages over larger animals, such as lower supply and maintenance
costs and the relative homogeneity of the breeding strains. Also, from the an-
imal rights viewpoints the use of rodents for scientific research is more desir-
able. Furthermore, the similarity of the cerebrovascular anatomy, physiology
and neuroglial cellular biology to humans119 makes these animals suitable for
this purpose .
Most focal cerebral ischemia models include occlusion of one major cere-
bral artery, mainly the middle cerebral artery (MCA). The pathophysiology of
cerebral ischemia restricted to the distribution of the MCA has been studied
widely because of its similarity to the clinical condition of thromboembolic
stroke120, 121. The anatomic distribution of the anterior communicating artery
(ACA), posterior cerebral artery (PCA) and MCA in rats resembles that in
humans122. The blood supply to the rat thalamus and basal ganglia is also anal-
ogous to that in humans123. Furthermore, rats have a preserved circle of Willis,
which exists in almost all mammals124.
MCAO of the rats can be achieved by either surgical or vascular access to
the MCA, exhibiting ischemia predominantly to the cortex and the middle and
posterior portions of the caudoputamen region, including the internal capsule
and anterior thalamus.
The surgical approach of MCAO has been used since 1975125. Since then,
this technique has been further developed and a direct microsurgical technique

23
for permanently occluding the MCA was introduced by Tamura et al126, re-
quiring a subtemporal craniectomy. Additional modification of this technique
allows transient or permanent MCAO by using clips127 or reversible snare lig-
ature128.
A relatively simple extracranial non ̶ invasive vascular model of reversible
MCAO was first described by Koizumi et al. in 1986129 and then by Zea-Longa
in 1989. Here, a 4 ̶ 0 bunted nylon suture was inserted into the right internal
carotid artery (ICA), advancing it cranially to block the MCA and the collat-
eral blood flow from the ACA and the PCA130.
This technique was used in Paper I.

Model of global cerebral ischemia


VF is commonly used method to mimic the clinical situation of CA in large
animals (most frequently dogs) and as defibrillation is usually successful in
this context, CPR has been added to follow CA131. Pigs have also been utilized
for the model described above132.
This CA/CPR model means that complete ischemia is followed by incom-
plete ischemia as the cerebral perfusion pressure (CPP) is low during CPR,
and the level of cerebral blood flow (CBF) produced is considerably lower
than the control level131.
Much effort has been made to improve neurological outcome by increasing
the systemic blood flow and thereby the CBF during CPR133. Higher CBF and
cardiac output (CO) have been achieved by using a mechanical device (LU-
CAS) in experimental CRP134.
The CA/CPR technique was used in Paper II and Paper III.

MCAO (Paper I)
Male Sprague-Dawley rats were subjected to focal cerebral ischemia using the
filament technique of MCAO maintained for 90 min followed by 1 d of reper-
fusion.
Anesthesia was induced and maintained by injection of fentanyl/fluanisone
and midazolam.
Pre ̶ ischemic hyperglycemia was achieved by intraperitoneal (IP) glucose
bolus injection (2 g/kg)55.
The tail artery was catheterized for measurement of blood pressure and
blood gas sampling, and a catheter was introduced into the left femoral vein
for intravenous access. The temperature was kept at 37.5±0.5 ˚C by using a
heating lamp.
The surgical procedure included exposure of the left carotid bifurcation,
after which the small branches of the external carotid artery (ECA) were di-
vided and the pterygopalatine artery was ligated. Then, a monofilament (Ethi-
lon 3/0) was inserted via the ECA and placed at the origin of the MCA.

24
Electroencephalogram (EEG) was recorded over the left MCA region in
order to verify ischemia135.
The rats were randomized into a therapeutic regime of S-PBN (156
mg/kg54) and saline (0.9% NaCl) given as a 60 ̶ min intravenous starting 60
min after MCAO.
The randomization was kept blinded to the operator until the experiment
was terminated.

Neurological testing (Paper I)


After 1 d of reperfusion the muscular strength and proprioception were exam-
ined using the inclined plane test136.
The device used in this study consists of a box (25x50 cm) with a rubber-
coated floor. One end of the box can be adjusted to achieve inclined plane.
The angle of the plane is measured by a digital water level. The maximal angle
of inclined plane that the rats could manage before gliding down was docu-
mented and the average of three attempts was assigned as the individual`s rec-
ord.

The neurological performance was also scored using the four ̶ level Bederson
scoring scale137.
Bederson et al (1986) produced a uniform site and size of areas of infarction
by transcranially diathermic occlusion of the MCA at various sites (divided
into five groups) and developed a method of 100 % specificity and 88% sen-
sitivity for assessment of the neurologic status of rats after MCAO137.
The neurological status was evaluated 24 ̶ h after surgery using a grading
scale of 0 ̶ 3. Rats that extended both forelimbs toward the floor and had no
other neurological deficit ranked as 0. Rats with variable forelimb flexion
were graded 1. Decreased resistance to lateral push and forelimb flexion with-
out circling was graded 2. Grade 3 had the same behavior as grade 2 but with
circling.
The developed neurologic examination correlated with the severity of neu-
rological deficit, and the size and location of lesion areas.

Triphenyltetrazolium (TTC) staining (Paper I)


The rats were perfused using 2% tri-phenyl-tetrazolium-chloride red (TTC)
and phosphate-buffered formaldehyde (4%, pH 7.4) in deep anesthesia. The
brains were sectioned into 2 ̶ mm slices and photographed. The lesions were
delineated by drawing free hand regions of interest around the pale areas on
TTC sections.
TTC reacts with succinate dehydrogenase enzyme in mitochondria and is
reduced to insoluble red formazan by electron acceptance138. This reaction is

25
lost in injured mitochondria and the resulting absence of staining in the is-
chemic area demarcates it from normal brain tissue. Park at al139 combined
TTC staining with perfusion-fixation using formalin. This method has the ad-
vantage to allow the histological verification of the presence or absence of
tissue damage in areas lacking TTC staining.

Cardiac arrest (Paper II and Paper III)


Thirty three pigs (weight: 22 ̶ 27 kg) were openly randomized into groups of
either hyperglycemic with blood glucose levels of 8.5–10 mmol/L (H) or
normoglycemic with blood glucose levels of 4 ̶ 5.5 mmol/L (N) groups, re-
sulting in 17 and 16 pigs, respectively (Paper II).
In the next experiment, twenty pigs (weight: 22 ̶ 29 kg) were openly random-
ized into groups of either hyperglycemic with blood glucose levels of 8.5–10
mmol/L (H) or normoglycemic with blood glucose levels of 4 ̶ 5.5 mmol/L
(N), resulting in 10 and 10 pigs, respectively (Paper III).

Anesthesia was induced by intramuscular injection of tiletamine ̶ zolazepam,


xylazine and atropine. An additional dose of morphine was given before tra-
cheotomy, thereafter anesthesia was maintained with an infusion of pentobar-
bital, morphine and pancuronium bromide.
The animals were ventilated mechanically via tracheostomy with
normoventilation (PaO2 5 ̶ 5.5 kPa).
Perspiration was compensated for by a fluid of Ringer’s solution (30 mL
kg-1) during the first hour, thereafter, bolus doses of 100 mL were given on
demand to treat hypovolemia, defined as hypotension (mean arterial pressure
(MAP) < 70 mmHg) along with tachycardia and increased respiratory varia-
tions in blood pressure.
An arterial catheter was inserted through a right external carotid artery
branch for blood pressure monitoring and blood sampling. A central venous
catheter and a pulmonary artery catheter were inserted via the right external
jugular vein, and a venous catheter was inserted cranially into the left internal
jugular vein and advanced as far as possible with intact backflow.
For cerebral oxygen saturation monitoring a single combined emitter ̶ de-
tector device was placed subcutaneously over the parietal skull and secured to
obtain a stable baseline regional hemoglobin oxygen saturation (rSO2) reading
from the connected near ̶ infrared light spectroscopy (NIRS) instrument (Pa-
per II). In the animals with sagittal sinus sampling, a catheter was inserted
into the sagittal sinus for sampling and via burr hole a laser Doppler flow
(LDF) probe was placed intrathecally through a minimal dural incision for
recording (Paper II and Paper III).
After surgical preparation, a glucose infusion (20 mL/h) was started in both
groups and short ̶ acting insulin (0 ̶ 2 units/h) was given to the normoglycemic
group. The blood glucose level was measured at 5 ̶ to 15 mins intervals to

26
achieve desired blood glucose targets and avoid hypoglycemia. Hemodynamic
variables including electrocardiographic recordings, heart rate, systemic arte-
rial blood pressure, central venous pressure and pulmonary artery pressure
were monitored continuously and recorded intermittently at baseline 0A (tar-
get glucose established), followed by 30 mins stabilizing interval 0B (time to
CA), and at 5, 15, 30, 60,120 and 180 mins after ROSC. Hemodynamic pa-
rameters such as the mean pressures were derived from the monitor, CO was
measured by thermodilution and used together with pulmonary capillary
wedge pressure recordings for hemodynamic calculations. LDF (Paper II and
Paper III) and rSO2 (Paper II) was also monitored continuously. HbO2 satu-
ration in the internal jugular vein and sagittal sinus were measured at each
time point indicated above (Paper II and Paper III).
CA was induced by using alternating current resulting in VF and hemody-
namic collapse. The VF was allowed to continue for 12 min (Paper II) or 5
min (Paper III) before starting closed ̶ chest compressions using pneumatic
device LUCAS at a rate of 100 cpm. After 1 min of CPR, Arg ̶ vasopressin
was given (0.4 units/kg intravenously). After 8 min of CPR, the VF was direct-
current defibrillated for ROSC.
Norepinephrine infusion was used to MAP of ˃ 70 mmHg. During the ini-
tial 5 mins after ROSC, FiO2 was 100%, thereafter 30% and the ventilatory
setting was adjusted to obtain normoventilation.
The blood samples included arterial, jugular venous, and sagittal sinus venous
samples. The plasma obtained after centrifugation was stored at ̶ 70°C until ana-
lyzed for F2-isoprostanes, protein S100β and troponin I (Paper II) and for protein
S100 β and proinflammatory cytokines IL-6 and TNF (Paper III).

Cerebral oxygenation (Paper II and Paper III)


The cerebral oxygenation was measured by three methods in Paper II: NIRS
index, saturation in the sagittal sinus and the jugular vein.

Near-infrared spectroscopy (NIRS)


The physical basic of NIRS involves several main principles, such as that human
tissues possess the property of comparatively good transparency to light in the
near-infrared (NIR) spectral window (650 ̶ 1000 nm), and that this light is either
absorbed by chromophores, such as hemoglobin (HbO2), deoxyhemoglobin (Hb)
or scattered in tissues140, 141. The relatively high attenuation of NIR light is caused
by the hemoglobin situated in small vessels (˂ 1 mm) of microcirculation, such
as capillary, arteriolar and venular vessels. Since the approximate arterial fraction
of cerebral blood volume is only 30 % in the human brain142, the NIRS technique
provides information predominantly regarding the changes in oxygenation taking
place in the venous compartment.

27
The absorption spectrum of hemoglobin is determined by its level of oxy-
genation, which in turns relies on the inspired oxygen and the pulmonary gas
exchange.
NIRS uses laser diode and/or light emitting diode (LED) within the optical
window of 650 ̶ 1000 nm, by using flexible fiber optics to convey the NIR
light to (source) and from (detector) tissues.
Several types of NIRS equipment are available based on a specific type of
illumination. The most commonly used NIRS instruments are based on the
continuous wave (CW) technique, as was the single combined emitter ̶ detec-
tor INVOS Somanetics used in Paper II.
The CW ̶ based modality measures the changes in Hb and HbO2 using the
modified Lambert ̶ Beer law. The Lambert ̶ Beer law states that for an absorb-
ing compound dissolved in a non ̶ absorbing medium, the attenuation of an
incident light is proportional to the concentration of the compound in the so-
lution and the optical pathlength (i.e. the length of the NIR light through the
tissue). For a highly scattered medium this law incorporates an additive term
G, describing the scattering attenuation and a multiplying factor to account for
the increased optical path ̶ length due to scattering.
The Somanetics INVOS Cerebral Oximeter measures rSO2 of the brain, (the
absolute ratio of HbO2 to the total Hb content)143 using two wavelengths (730
and 810 nm). The spatially resolved spectroscopy (SRS) method is established
to improve the resolution using a sensor with two source ̶ detector distances,
such as a ´near´ (shallow) and a ´far´ (deep). Subtraction of the near signal from
the far signal should provide a signal originating mainly in the brain cortex.
Practical advances of NIRS methods include its easy use for routine con-
tinuous or intermittent non ̶ invasive bedside measurements of changes in cer-
ebral hemodynamics.

Laser Doppler flowmetry (LDF)


LDF has the same basic principle as ultrasound Doppler. The sensor emits a
monochromatic laser light and by measuring the volume or concentration of
red blood cells and their velocity and multiplying these two parameters, a flow
signal is created.
LDF provides a continuous and qualitative estimate of regional CBF
changes presented in arbitrary units using small sample volume144.
This method was used in Paper II and Paper III.

Jugular venous oxygen saturation (SjvO2)


Since jugular venous blood carries unexploited oxygen away from the brain,
SjvO2 provides an approximate measurements of the global balance between
cerebral oxygen delivery and utilization.
The monitoring can be carried out either continuously by using fiber-optic
catheters or intermittently by analyzing the blood sample using a co-oximeter.
The latter method was used in Paper II and Paper III.

28
Isoprostanes and prostaglandins (Paper II)
Isoprostanes, a family of prostaglandin ̶ like compounds, are synthetized non ̶
enzymatically in vivo from esterified polyunsaturated fatty acid (PUFA)
through a free radical mediated process (see review145).
Isoprostanes operate as full or partial agonists through thromboxane re-
ceptors as well as through some independent mechanism, and are considered
reliable biomarkers of lipid peroxidation and pathologic mediators of oxida-
tive stress due to their vasoconstrictive and inflammatory effects. Isopros-
tanes, distinct to primary prostaglandins, do not request COX for their bio-
transformation. Based on the process of formation, there are four F-ring iso-
prostanes regioisomers shaped: -5,-12,-8- and -15-series. Since F-ring com-
pounds are structurally isomeric to the COX ̶ derived primary PGF2α, these
compounds are entitled as F2-isoprostanes and can be found in high quantity
in the tissues.
F2-isoprostanes, mainly 8-iso-PGF2α, are produced in their esterified form
in tissue and converted by hydrolytic enzyme reaction to their free acid form,
which secretes to the peripheral circulation. Further metabolism takes place
by reductive and oxidative enzymes and the end products excrete via urine.
15-keto-13,14-dihydro-PGF2α (15-keto-dihydro-PGF2α) is a major metab-
olite of PGF2α and can be used as an indicator of lipid peroxidation through
the COX pathway. Furthermore, it could indicate inflammation as its level
rapidly increased following infusion of intravenous endotoxin in the porcine
model146, 147.
A specific radioimmunoassay (RIA) was developed by raising specific an-
tibodies in rabbits to 8-iso-PGF2α and 15-keto-dihydro-PGF2α in order to
measure free and total levels of 8-iso-PGF2α148 and 15-keto-dihydro-PGF2α149,
respectively.

Cerebral and myocardial injury marker (Paper II and Paper III)


Cerebral damage was measured as the neuroglial marker protein S100β and
for myocardial damage as serum mass concentration of troponin I.

Protein S100β
S100β, a small dimeric calcium binding ̶ protein, belongs to a large family of
over 20 different calcium ̶ modulated proteins, and was first described by
Moore in 1965150. Moore isolated it from the bovine brain as a partially puri-
fied mixture of two S100 proteins (S100β and S100A1) and called it´S100´
because of its solubility in 100% saturated ammonium sulfate. S100β is pro-
duced mainly by astrocytes, but can be localized in oligodendrocytes, neurons
and the choroid plexus, metabolized and excreted by the kidneys, has a plasma
half ̶ life of about 30 min151 and is found in healthy individuals at low levels152.

29
In low concentration, S100β has beneficial activity, such as inducing neu-
rite overgrowth153, stimulating proliferation154, phosphorylation of the
ERK1/2 through MAPK pathway155 in astrocytes and it exhibits neuroprotec-
tive activity after various injury paradigms153, 156, 157.
Excessive production of S100β from astrocytes has detrimental conse-
quences, producing pro ̶ inflammatory cytokines, such as IL-1β, TNF and IL-
6 in glia158 and IL-6 in neurons159. IL-1β enhances S100β ̶ induced activation
of inducible NOS via generating NO160 and by creating complex response cas-
cades which can result ultimately in neuronal death160. Furthermore, micromo-
lar concentration of S100β can elicit direct neurotoxic effects161.
Elevated levels of S100β in serum or cerebrospinal fluid (CSF) can be used
as a biomarker for evaluating the presence of brain damage and predicting the
outcome in a variety of pathologic conditions, such as stroke162 and CA163 . Its
prognostic role in these contexts will further be discussed later.

Cytokines IL-6 and TNF (Paper III)


IL-6 is the major cytokine of the brain and is produced by neurons, but also
by diverse populations of microglia and macrophages. Not only is IL-6 impli-
cated in the immune response and inflammation, but plays a major role in
many physiological processes in the central nervous system (CNS) as well as
in diverse neuropathology. As a pleiotropic cytokine, IL-6 exerts both proin-
flammatory and anti ̶ inflammatory effects164.
TNF are expressed by neurons, microglia and some astrocytes, along with
the endothelial cells of the cerebral vasculature, but also by the peripheral im-
mune system. TNF has also a pleiotropic role in the ischemic brain, most
likely by acting via two different pathways. At least two TNF receptors have
been recognized, such as TNF receptor 1 (TNFR1) and TNF receptor 2
(TNFR2) expressing together on neuronal and glial membrane surfaces and it
has been suggested that they have an antagonistic function. Most TNF ̶ in-
duced signals are mediated by TNFR1, which includes a death domain (DD)
and may represent a dividing route for signaling cell death or cell survival.
Activation of TNFR1 leads to apoptosis by involving cysteine protease caspa-
ses. However, the connection of TNF associated factor-2 (TRAF-2) protein
with DD of TNFR1 can result in cell survival by activating NF-κB gene tran-
scription and by up ̶ regulation of neuroprotective mediators, such as calbindin
and free radical scavenger MnSOD. This pathway is also coupled to TNFR2 ̶
mediated cascade mechanism. The overall effect of TNF is highly multifac-
eted and may be determined by the ratio of TNFR1/TNFR2 expression165.

Biochemical analysis (Paper II and Paper III)


The plasma concentration of 8-iso-PGF2α and 15-keto-dihydro-PGF2α indicat-
ing oxidative stress and inflammation, respectively were measured according

30
to methods previously described148, 149 with detection limits of 23 pmol/L for
8-iso-PGF2α and 45 pmol/L for 15-keto-dihydro-PGF2α, respectively. Unex-
tracted plasma from jugular vein samples was analyzed by using the RIA
method (Paper II).
Myocardial damage was measured in jugular vein samples as serum mass
concentration of Troponin I using monoclonal/polyclonal mass immunoassay
(ADVIA Centaur, Bayer) with normal level of ˂ 0.12 µg/L (Paper II).
Cerebral damage was assessed by measuring the neuroglial marker S100β
in blood. The analysis was carried out at the department of Clinical Chemistry
(LIA-mat, Sangtec 100 in Paper II and Cobas 6000 analyzer, e601 module,
Roche Diagnostics in Paper III). The blood samples were drawn from the
jugular vein in Paper II and the sagittal sinus in Paper III.
For quantification of proinflammatory markers IL-6 and TNF, plasma sam-
ples from the sagittal sinus were analyzed in duplicate using commercially
available Enzyme Linked Immunosorbent Assays (ELISA) according to the
manufacturer’s instructions. The optical density was measured at 450 nm and
correction wavelength 540 nm, using a Tecan Sunrise instrument running the
Magellan software (Paper III).

RNA preparation (Paper IV)


Isolation of total RNA from twelve pigs (H = 6 and N = 6) was performed by
quickly dissecting approximately 100 mg of the frontoparietal cortex and in-
stantly putting the brain tissue in homogenizing buffer containing β-mercap-
toethanol according to the manufacturer’s protocol (Qiagen Inc., USA). The
tissue was then immediately homogenized using a Polytron homogenizer and
total RNA isolated by RNeasy Mini kit (Qiagen) with absorbance determined
at 260 and 280 nm.

Microarray expression analysis (Paper IV)


Microarray analysis by assaying gene expression - a simultaneous analysis of
large amounts of genes - can be describes as a power method for biomedical
research.
DNA microarray comprises tens of thousands of immobilized single ̶
stranded DNA (ssDNA) sequences, known also as probes attached to a solid
surface in an orderly arrangement, which are utilized to hybridize a mobile
labeled complementary DNA (cDNA) or complementary RNA (cRNA) sam-
ple, also called as a target.
The first phase of microarray analysis is the isolation of total RNA followed
by the control of its quality and integrity.
The first-stranded cDNA is synthetized from total RNA as a primer by us-
ing reverse transcriptase. This single ̶ stranded cDNA is (ss ̶ cDNA) than trans-
lated to double–stranded cDNA, which in turn plays as a template to produce

31
antisense RNA or cRNA via in vitro transcription (IVT). The next step is to
purify the cRNA in order to create the second cycle ss ̶ cDNA. The final step
of this process is to remove the template cRNA followed by fragmentation of
ss ̶ cDNA. The target ss ̶ cDNA can then be labeled by several different agents,
such as fluorescence dyes, a hapten group (biotin, aminoallyl) or by radioac-
tive compounds.
The main principle of microarray is the hybridization between the labeled
mobile targets and the complementary immobile sequences through the for-
mation of base pairs.
After hybridization the arrays are moved to fluidic station. By washing off
the non ̶ specific bonding, the remaining strongly attached base pairs stay hy-
bridized thus creating a signal. The total strength of the signal is determined
by the quantity of bindings between the targets and the probes on a specific
spot.
The hybridized arrays are then stained with a fluorescent streptavidin ̶ phy-
coerythrin conjugate, which attaches to the biotins and creates fluorescent
tags. Here, the microarrays are scanned in order to create a digital image of
the array through the use of different scanning techniques, such as autoradi-
ography, chemiluminescence or fluorescence scanning. This process makes
use of a special microscope with a laser, which excites the fluorophores and
creates fluorescence intensities. This process is followed by identification of
the differentially expressed genes, by applying either a two ̶ channel or one ̶
channel detection system.
The data analysis is complex and comprises several steps, such as annota-
tion of all spots, data processing, statistical analysis and clustering of the genes
exhibiting similar patterns of gene expression166.

RNA quality was evaluated using the Agilent 2100 Bioanalyzer system. 250
nanograms of total RNA from each sample were used to generate amplified
and biotinylated sense ̶ strand cDNA from the entire expressed genome ac-
cording to the GeneChip® WT PLUS Reagent Kit User Manual (Affymetrix
Inc., CA). GeneChip® ST Arrays (GeneChip® Porcine Gene 1.0 ST Array)
were hybridized for 16 hours in a 45°C incubator and rotated at 60 rpm. In
accordance with the GeneChip® Expression Wash, Stain and Scan Manual,
the arrays were then washed and stained using the Fluidics Station 450 and
finally scanned using the GeneChip® Scanner 3000 7G.

Statistical analysis
An α level of p˂0.05 was considered statistically significance in Papers I-III,
while an α level of p˂0.001 was chosen as statistically significant in Paper IV.

32
Paper I
Parametric data were expressed as mean ± SD and non-parametric data as
mean (range). Differences between groups were tested for statistical signifi-
cance using unpaired t ̶ test for parametric and Mann-Whitney U test for non-
parametric data, respectively.

Paper II and Paper III


Data from serial hemodynamic and oxygenation parameters were expressed
as mean ± SEM. Biochemical markers were expressed as medians and inter-
quartile range.
Kolmogorov-Smirnov test (Paper II) or D’Agostino & Pearson test (Paper
III) was performed for normal distribution of the baseline values. All variables
had normal distribution or log ̶ normal (biochemical markers) distribution.
The serial measurements were analyzed by using a two ̶ way analysis of vari-
ance (ANOVA) for repeated measurements, followed by the Bonferroni post
hoc test.
A summary statistic - defined as the peak ̶ baseline ratio - was calculated
for each variable and each pig. The group difference for this statistic was ex-
pressed as 95% confidence interval (CI) and tested by using an unpaired Stu-
dent`s t test.

Paper IV
The raw data was normalized in the free software Expression Console using
the robust multi ̶ array average (RMA) method. Principal component analysis
(PCA) was applied in order to visualize the data using, more specifically, the
“princomp” function in R. One sample was highly aberrant and therefore ex-
cluded from the analysis. Normalization was performed again. Subsequent
analysis of the gene expression data from the remaining 11 samples was car-
ried out in the freely available statistical computing language R, using pack-
ages available from the Bioconductor project. In order to search for the differ-
entially expressed genes between the hyperglycemic and the normoglycemic
groups, an empirical Bayes moderated t ̶ test was applied, employing the
‘limma’ package with the robust method. To address the problem with multi-
ple testing, the p ̶ values were adjusted, corresponding to the false discovery
rate (FDR). Heat maps were generated in Genesis 1.0 using the normalized
data for each sample. The data was adjusted by using the command “mean
center genes” and hierarchical clustering was performed.

33
Results and discussion

Paper I

General findings
A total of 42 rats were successfully subjected to MCAO, of which 13 were
included for analysis of infarct size and neurological tests.
All included rats showed contralateral motor deficit at recovery from anes-
thesia indicating successful MCAO.
After 1 d of survival the group treated with S-PBN showed significantly
better neurological scores than the controls, according to the scoring de-
scribed by Bederson137
The performance on the inclined plane was similar at baseline, but signifi-
cantly better in the S-PBN ̶ treated group than controls after 1 d of survival136.
All included rats showed pallor anatomically corresponding to the MCA
territory, leaving normal tissue red after TTC staining139.
The volumetric assessment integrating areas of 2 ̶ mm slices revealed sim-
ilar total brain and infarct volumes in the S-PBN ̶ treated and control rats.
The summary of results are shown in Table 1.

Table.1.
Groups S-PBN Control P
     
Bederson1(score) 1 (1 ̶ 3) 3 (2 ̶ 3) ˂0.05
Inclined plane2 (%)
Absolut 74.5±4.6 66±8.3 ˂0.05 
Relative ̶ 1.7±2.8 ̶ 9±3.3 ˂0.01 
Volumetric           
analysis3(mm3)
Total 1054±61 1102±86 NS
Infarct 206±18 213±57 NS
     
1 137
Neurological scoring according to Bederson et al
2
Indicating muscular strength and proprioception136.
3
Volumetric analysis with TTC (tri-phenyl-tetrazolium-chloride)139.

34
The main finding 
The present results indicate that S-PBN may ameliorate hyperglycemic ̶ is-
chemic brain damage in terms of neurological performance.
In two separate studies, non ̶ sulfonated PBN decreased the infarct size by
approximately 30%167 in normoglycemic compared with 70% in hyperglyce-
mic ischemic injury54.
Our finding that S-PBN did not reduce the infarct size but improved the
neurological performance, contrasts with the more comprehensive data re-
ported for PBN105. It is possible that structural differences between the two
compounds affecting, for instance, cerebral glucose metabolism - as demon-
strated in TBI115 - might explain this discrepancy.

Glycemic levels
In the present study, the induced hyperglycemia was well above the levels
associated with poor stroke outcome13.
Admission hyperglycemia presenting in 20 ̶ 50% of acute stroke patients is
a significant risk factor of poor outcome following stroke7, 10, 51, 52, 168, and also
an independent prognosticator for long ̶ term morbidity and mortality with this
condition168.
Capes et al found that even a mildly high admission glucose level (> 6.1 ̶
7.0 mmol/L) increased the risk of in ̶ hospital or 30 ̶ day mortality in non-
diabetic patients13.
Admission hyperglycemia was significantly more common in patients with
more severe stroke, especially in subtype total anterior circulation syndrome
(TACS)7 suggesting a relationship between high glucose level and severity of
stroke.
In models of brief transient cerebral ischemia hyperglycemia was consist-
ently associated with increased infarct size169, but it had slight or no effect in
models of permanent ischemia without reperfusion170, a finding compatible
with human studies where higher admission blood glucose level was associ-
ated with worse outcome in non ̶ lacunar stroke10.
Furthermore, the detrimental effect of hyperglycemia on stroke outcome
has been associated with early rather than late stages of reperfusion. This sug-
gests that the load of excessive glucose at the time of reperfusion may hasten
the recruitment of penumbral tissue into infarction171.

In line with this observation, the mortality not attributed to subarachnoid hem-
orrhage (SAH) was high among rats with successful MCAO. A post ̶ hoc anal-
ysis of survivors and non ̶ survivors revealed that the blood glucose level at
the time of MCAO and at 120 min was the only discrepant parameter between
these two groups. The values in the non-survivor group were 2.3 and 3.6
mmol/L higher, respectively, at these two time points.

35
Glucose metabolism and cerebral blood flow under ischemic
conditions
During ischemic conditions, classically defined as a blood flow of ˂ 18
mL/100g/min4, action potential cannot be generated and consequently the
electrical activity of the brain ceases172. Furthermore, the relationship between
CBF and glucose metabolism is disturbed173. This effect is more pronounced
when hyperglycemia precedes the ischemic insult174.
In the model of focal MCAO, a markedly reduced glucose metabolic rate
in the core of the infarct region has been observed and extreme degrees of
metabolisms ̶ greater than blood flow uncoupling were detected in the early
ischemic penumbra175.
A similar marked mismatch of local glucose metabolism and blood flow
was found in TBI together with astrocytic activation, suggesting the preferen-
tial role of stimulated glycolytic metabolism in astrocytes in both injury mech-
anisms175, 176.
During experimental MCAO, a positive correlation has been revealed be-
tween glucose levels, rCBF and CMRO2, while during the early reperfusion
phase the glucose levels were determined by the presence of hyperperfusion
and the occurrence of flow ̶ metabolism mismatch177.

Role of RONS in mismatch of glucose metabolism and cerebral


blood flow
Oxidative stress, mainly in terms of RONS, has been implicated in uncoupling
the rCBF and glucose metabolism in TBI, and both spin ̶ trapping agents PBN
and S-PBN were found effective in attenuating this effect115. The possible con-
tribution of this mechanism in attenuating the hyperglycemic ischemic cere-
bral damage by S-PBN will be discussed later.
Clinical interest has been directed towards sulfonated nitrones, such as S-
PBN and dual sulfonated NXY-059, as the latter was found to be neuropro-
tective by the Stroke Ischemic NXY-059 Treatment Study (SAINT) I trial178.

Nitrone-based trapping agents


PBN, S-PBN and NXY-059 differ structurally in the number of sulfo groups
attached to the phenyl ring (Fig.2).

36
Fig.2.Structures of PBN, S-PBN and NXY-059 (Published with permission from the
publisher.)

All three of these nitrones trap carbon ̶ and oxygen ̶ centered radicals to form
radical adducts according to the electron spin resonance (ESR) studies. The
strongest ESR adduct intensity is shown by NXY-059. Each nitrone can coun-
teract in vivo salicylate oxidation, PBN being the most efficient179. NXY-059
has been proven more effective than PBN in a transient MCAO model in rat106
although it is not a more effective antioxidant compared with PBN179.
In addition to scavenging free radicals and reducing NO production180,
PBN influences inflammatory mediators181, transmitter systems182 , blocks
Ca2+ channels183 and improves mitochondrial function184. PBN may act by pro-
tecting against the breakdown of phospholipids and attenuating lactate for-
mation after TBI176.
PBN prominently improved the energy state of mitochondria by normaliz-
ing the ATP and lactate contents in both the ischemic core and penumbra sug-
gesting a recovered microvascular function in a focal MCAO model. One pos-
sible explanation of this effect could be that PBN blocks the activation of ad-
hesion molecules for PMNs or the inflammatory reactions induced by PMN
and platelets adhesion185.
Both PBN and S-PBN have been found effective in trapping ROS in TBI116
although they have different abilities for penetrate the BBB. PBN readily pen-
etrates the BBB and has a half-life in plasma of 3h186, while S-PBN has a half-
life in plasma of 9 min and penetrates the BBB poorly116. In spite of measur-
able plasma concentrations of S-PBN, there was no detectable concentration
in the brain tissue measured 30 and 60 min after experimental cerebral injury,
however, the reduction of ROS production by both agents was similarly effec-
tive116. These findings suggest that PBN may act both by scavenging intracel-
lular sources of ROS, and like S-PBN, may perform ROS scavenging at the
blood–endothelial interface116. The latter mechanism has been advocated to
explain the early improvement in rCBF and glucose metabolism observed in
PBN and S-PBN treated animals after experimental TBI, suggesting that ROS
are implicated in the uncoupling of CBF and glucose metabolism under these
circumstances115.

37
S-PBN has been found to improve both morphological and functional out-
comes in TBI without any detectable drug level116. Based on the neurovascular
unit concept (an coordinated interaction between neuron, glia, endothelial cell
and immune system) it has been hypothesized that S-PBN - as a non ̶ pene-
trating nitrone - interacts with cross ̶ signaling between the injured brain and
the peripheral immune system at the microvascular endothelial level, thereby
modulating immune cell trafficking into the brain after TBI117. Clausen et al117
found that pretreatment with S-PBN significantly reduced the TcR+ T lym-
phocytes and attenuated the neutrophil infiltration into the injured brain. These
findings propose that T lymphocyte trafficking at the microvascular endothe-
lial level involves ROS, which when targeted by S-PBN may offer a possible
new approach of neuroprotection.
S-PBN has also been found effective in lowering the cerebral cortical blood
flow (CCBF), - which mirrors the global cerebral perfusion - in transient
global cerebral ischemia probably by redirecting the blood flow to areas where
blood circulation was deficient112. As S-PBN does not enter the brain via BBB,
Preston et al. hypothesized that it may travel through pore-like openings in the
BBB after forebrain ischemia in rats187.
S-PBN has also been found to attenuate the excitotoxic cerebral lesion in
rats111.

Therapeutic time window of S-PBN


A short period of reduced ROS formation after vessel occlusion, followed by
a burst ̶ like pattern of enhanced ROS production (approximately fourfold)
during early reperfusion, was demonstrated in an experimental transient
MCAO model70.
In the present study, S-PBN was given primarily to cover this time window
of anticipated RONS provocation.

Failure of translation of neuroprotection in animal studies into


clinical practice
The SAINT I trial demonstrated that NXY-059 improved the outcome after
acute ischemic stroke178. NXY-059 given within 6 h of the onset of acute is-
chemic stroke reduced the cerebral injury and significantly improved the pri-
mary outcome (reduced disability to 90 d), however, re-examination of the
summary findings in the SAINT I trial led to the conclusion that both statisti-
cal and clinical significance of the efficacy of NXY-059 was at best equivo-
cal188.
Furthermore, the subsequent SAINT II trial could not confirm these re-
sults189 and several shortcomings of these studies have been pointed out, such

38
as low methodological quality of the studies, lack of randomization190 and
publication bias191.
Methodological concerns regarding nitrone decompositions might be of
relevance in interpreting these findings given that PBN can decompose during
storage to benzaldehyde and N-tert-butyl-hydroxylamine (NtBHA), the latter
has been proposed to act as antioxidant192. A corresponding decomposition of
S-PBN or NXY-059 is not available but there is no evidence of any in vivo
decomposition of S-PBN or NXY-059.

In pre ̶ clinical studies, over 1000 potential neuroprotective therapies have


been trialed193, but only a few clinical attempts at neuroprotection for ischemic
stroke have been able to show a beneficial effect (Stroke Trial Registry Home
Page).
Despite all the drawbacks regarding the clinical effectiveness of the neuro-
protective approach, the concept of targeting RONS should not be rejected as
a feasible therapy for cerebral ischemia. Genetic and pharmacologic blocking
of NADPH oxidase NOX seems to be a promising treatment option for acute
stroke194. Hypothermia has also been proved efficient in experimental models
of cerebral ischemia195 and one possible mechanism of the neuroprotective
effect involves inhibiting the formation of RONS196.

39
Paper II

Resuscitability, blood glucose levels and hemodynamics


Thirty ̶ three pigs were openly randomized into a hyperglycemic (H) and a
normoglycemic (N) groups, resulting in 17 and 16 pigs, respectively. Six pigs
were excluded because of failure to obtain ROSC and premature mortality 5 ̶
10 min after ROSC. Of the included pigs, 17 were monitored with rSO2 (H=9,
N=8) and 10 (H=6, N=4) with a sagittal sinus catheter and LDF probe.
The blood glucose levels were kept within the target range during the 30
min before CA, then increased in both groups at ROSC, and returned to base-
line approximately 120 min after ROSC. There were no significant differences
between the two groups in MAP, CO, systemic vascular resistance or demand
of vasopressor support.

rSO2, SjvO2 and HbO2 saturation in the sagittal sinus (SssvO2)


No difference was seen in rSO2 immediately after placement of the sensor and
there was no significant difference between the groups during preparation or
CPR.
The rSO2 responded promptly to the alterations in perfusion associated with
the CA and was significantly higher after ROSC in the H group compared with
the N group (p<0.05). This difference was already seen at the glucose clamp-
ing immediately before CA (but not from the first obtained values before the
clamp) and a similar pattern was observed in the jugular but not in the sagittal
sinus HbO2 measurements.
The laser Doppler measurements of the cortical blood flow in a subset of
experiments did not reveal any differences compatible with the rSO2 changes.
The peak in HbO2 saturation was seen 5 min after ROSC for both parame-
ters, although there was no significant difference in HbO2 saturation between
the two groups at any time.
The tendency toward higher rSO2 in the H group appeared to be correlated
with the SvjO2, rather than the SssvO2 measurements.

Markers of injury
The concentrations of 8-iso-PGF2α and 15-keto-dihydro-PGF2α were measured
in jugular venous samples.
Peak concentrations of 8-iso-PGF2α were seen 5 min after ROSC in both
groups, but there were no differences between the groups throughout the ob-
servations.
The peak levels of 15-keto-dihydro-PGF2α were seen throughout the ROSC
period in both groups, but no significant differences were detected. However,

40
tendencies toward increased 15-keto-dihydro-PGF2α were observed in the H
group.
Protein S100β peaked 5 min after ROSC in both groups. There was a higher
but non ̶ significant mean value of protein S100β in the H group during the
initial time points after ROSC.
Troponin I increased markedly after the insult in a similar manner in both
groups throughout the experiment.
Thus, the present results showed no significant differences in oxidative
stress or markers for cerebral or myocardial damage between the hyperglyce-
mic and normoglycemic pigs.

Cerebral oxygenation (rSO2, SjvO2, Sssvo2)


The cerebral oxygenation by means of regional NIRS was significantly higher
in the H than the N group after ROSC (p˂ 0.05) (fig.3).

Fig.3.Cerebral oxygenation. Time points A and B represent the 30-min interval of


blood glucose stabilization. Cardiac arrest (12 mins, black), cardiopulmonary resus-
citation (CPR, 8 mins, dark gray) and restoration of spontaneous circulation (ROSC,
first 5 mins, light gray). A Near-infrared light spectroscopy index of regional oxygen
saturation (rSO2). No difference was present between groups at baseline; however,
higher levels developed after time point A in the H group compared with the N group.
(*p ˂.05).

The rSO2 represents a tissue oxygenation extraction, with functions as a sum-


mation of the HbO2 saturation of all vascular beds (but mainly of the venous
compartment) and therefore reflects the perfusion state.
The sagittal sinus measurements are considered to represent the cortical
perfusion because of the cortical drainage into the sinus, while the jugular ve-
nous blood contains a mixture of variable proportions of cerebral and non-
cerebral drainage. Dual measurements of isoprostanes in the sagittal sinus and
in the jugular vein - using the porcine CA model - suggest that the average
admixture from extracerebral tissue is approximately 50 %113

41
Rivers et al used continuous measurements of central venous saturation
(ScvO2) during and after CPR to evaluate the clinical significance of mixed
venous hyperoxia197.
Mixed venous hyperoxia (˃ 75%) has been associated with poor prognosis
after cardiogenic shock, indicating the damaged ability of the cell to utilize
oxygen in the post ̶ resuscitation phase197. Non ̶ survivors after CA presented
an impairment of oxygen extraction and a more supply ̶ independent oxygen
consumption (Vo2) pattern197. The pathogenesis of low Vo2 may depend on
several factors, such as microcirculatory failure, high doses of vasopressor
(epinephrine) therapy by redirecting the blood flow from the peripheral to cen-
tral compartment and longer duration of CA197.
In this study, no included animal received epinephrine for resuscitation pur-
poses, but it cannot be ruled out that the increased rSO2 in response to hyper-
glycemia may represent an aggravated microcirculatory dysfunction.
The inflammatory responses - triggered by I/R injury - can also precipitate
microcirculatory dysfunction198. The reestablished circulation with oxygen-
ated blood after CA results in overproduction of ROS which can be additive
by simultaneous generation of ROS in ischemic neurons and immune cells.
ROS production in turn promotes secretion of inflammatory cytokines and
chemokines by ischemic neurons, astrocytes and activated microglia199. These
responses mediate upregulation of adhesion molecules in the cerebral vascu-
lature, which in turn stimulate the enrollment of peripheral leucocytes into the
brain parenchyma200. After entering into the brain, the activated leukocytes
and microglia generate numerous inflammatory mediators, such as TNF, IL-
1β, and NO via iNOS, which act through common key regulators, the tran-
scription factor NF-κB being the most important198.
In experimental I/R, the role of endothelial P-selectin and ICAM-1 has been
highlighted in rolling and firm adhesion of leukocytes creating a surface onto
platelets can be attached201. This process is especially pronounced in the post-
capillary segment of cerebral microcirculation201 and can lead to prothrom-
botic state of microvessels200. Prolonged time of ischemia and a longer period
of reperfusion tend to intensify this process201.
Proinflammatory cytokines, such as IL-6 and TNF contribute to this path-
ogenesis by producing metalloproteinases (MMPs), especially MMP-9. Inter-
action between cytokines and leukocytes results in adhesion of leukocytes to
vascular wall. MMPs then alleviates the migration of leukocytes through the
endothelium, which leads to breakdown of the BBB, oedema building, hem-
orrhage and ultimately cell death 202.
Thus, prolonged CA may cause cerebral microcirculatory failure under
reperfusion period that might be aggravated further by the coexisting hyper-
glycemia.

Untreated CA of 10 ̶ 12.5 min is associated with a dynamic rather than fixed


no ̶ flow hypoperfusion after ROSC203.

42
Under normoglycemic conditions, regional differences in CBF and cerebral
metabolic rate have been demonstrated after CA in a porcine model204. During
the first 30 min of ROSC, a transient cerebral hyperemia followed by a pro-
tracted global and multifocal hypoperfusion were seen in animal models of
CA and resuscitation205. Mörtberg et al found a bilateral cortical hypoperfu-
sion especially in the parietal and temporal cortex and in the subcortical re-
gions at 60 min after ROSC of 10 min of CA204. Furthermore, the same study
demonstrated an increasing hyperperfusion in close proximity of the hippo-
campus - a region most vulnerable to hypoxia206 - in the hours after ROSC,
but no significant difference was revealed in oxygen extraction, reflecting cer-
ebral metabolic rate.
The intact brain can adapt the CBF to the cerebral metabolic rate, while the
impact of hyperglycemic CA on regional changes of CBF during reperfusion
remains to be shown.

Oxidative stress and inflammation


Oxidative stress and inflammation have been implicated in CPR after CA207-209.
In this study, oxidative stress was measured as the level of 8-iso-PGF2α,
and was not affected by hyperglycemia.
Levels of 8-iso-PGF2α, were measured both in mixed venous and the jugu-
lar bulb - that primarily drains the brain - after induced CA and ROSC in por-
cine model and were found quickly elevated indicating oxidative injury207.
Since higher concentrations have been revealed in jugular venous samples, the
role of oxidative stress in I/R brain injury113 has been implicated. Furthermore,
increase of CA time led to an additional increase in 8-iso-PGF2α. Elevated
levels of 8-iso-PGF2α were also found to be linked to neurological deficits113.
By using the same model, the administration of various free radical scaven-
gers, such as S-PBN and PBN, resulted in diminished oxidativ stress - indi-
cated by the decreased level of 8-iso-PGF2α - and improved neurological out-
come, respectively112, 209.

The concentration of 15-keto-dihydro-PGF2α, indicating inflammation, tended


to be non ̶ significantly higher in the H group.
The role of inflammation in the pathogenesis of hyperglycemic ischemic
brain injury has been highlighted by several investigators8, 210. Hyperglycemia
accelerates the inflammatory reaction to I/R by various mechanisms, for ex-
ample, increased cycloxygenase-2 (COX-2) and IL-1β protein expression has
been revealed in a rat model of focal cerebral ischemia57. COX-2 is considered
to contribute to ischemic damage by producing O2- radicals211.
IL-6 is released from monocytes under hyperglycemic condition, suggest-
ing its potential role in hyperglycemic cerebral I/R damage212.
Hyperglycemia enhances the adhesion of polymorphonuclear (PMN) leu-
kocytes adhesion in blood vessels of the ischemic territory213. Adhesion of

43
PMN leukocytes induces a burst of RONS production, thus contributing to
secondary injury by activating a group of lipases, proteinases and DNAases
leading to disruption of BBB and microvessel occlusion214.
Furthermore, hyperglycemia increases O ̶ linked glycosylation of cerebral
protein as a contributing factor to I/R injury210 by activating the hexosamine
pathway67. O ̶ glycosylation can produce modification of Ser/Thr residues (O ̶
GlcNAcylation) on a substantial number of signaling molecules resulting in
transiently blocking residue phosphorylation215. O ̶ GlcNAcylation inhibits
Akt activation in response to insulin signaling and also worsens cardiomyocite
function in hyperglycemic conditions216.
In addition to its pro ̶ oxidant effect, glucose has been shown to have a pro ̶
inflammatory action217. Intake of glucose (75 g) in humans induced production of
free radicals by PMN leukocytes and mononuclear cells and the expression of
p47phox, an essential constituent of the enzyme NADPH oxidase. The augmented
formation of ROS in turn activates an amount of redox ̶ sensitive pro-inflamma-
tory transcription factors, such as NF-κB, activator protein-1, early growth re-
sponse -1 and HIF-1α217, 218. Furthermore, hyperglycemia increased the expres-
sion of a substantial number of genes at the messenger RNA (mRNA) level, such
TNF, IL-6 and monocyte chemoattractant protein-1219, 220.

The anti-inflammatory and neuroprotective effect of insulin have been high-


lighted in cardiovascular context217, 221.
Insulin plays a key function as a metabolic hormone with an intense effect
on glucose and lipid metabolism. Insulin possesses however several other bi-
ological actions, such as induces vasodilatation by prompt release of NO by
the endothelium222 and by the expression of endothelial NOS223.
Insulin has been demonstrated an anti ̶ inflammatory effect by suppressing
three important inflammatory mediators: intercellular cell-adhesion molecule-
1, monocyte chemoattractant protein-1 expression and NF-κB binding in hu-
man aortic endothelial cells in vitro224. It was also shown that insulin has ROS ̶
suppressive effects225.
Furthermore, insulin can elicit an antiplatelet and antithrombotic effect by
inhibition of platelets aggregation caused by the activation of platelet NOS
and subsequent NO production226.
During acute focal and global ischemia insulin reduces ischemic brain dam-
age221, 227, either by reducing of blood glucose level or eliciting direct neuro-
protective effect221. In experimental studies with transient global ischemia in-
sulin has been proven to be neuroprotective and even improved neurobehav-
ioral outcome228. In contrast to global ischemia, the neuroprotective effect of
insulin in experimental focal ischemia was found to be due to its reduction of
glucose levels, as the simultaneous administration of glucose infusion abol-
ished most of the mitigating effect of insulin229.

44
Protein S100β
During the initial time points after ROSC the neuroglia marker protein S100β
showed a numerically, but non ̶ significantly higher peak in the H group com-
pared with the N group, a discrepancy that was attenuated cross the observa-
tion time.
The role of protein S100β as cerebral injury biomarker in the context of
hyperglycemic CA is discussed in Results and Discussion Paper III.

Time window and duration of ischemia


The time window of 3 hours after CA covered in Paper II and Paper III may
represent the early phase of post ̶ CA syndrome described in humans230. Pre-
vious data from our laboratory113 also revealed that ischemic brain damage is
evident in early after CA, which underlines the importance of focusing on this
phase of event.
The ischemic brain damage triggered by CA and resuscitation is likely to
reach full magnitude over the hours and days after ROSC230. In addition to
cerebral microcirculatory failure, hyperemic macroscopic reperfusion is seen
in the first few minutes after CA, further aggravating the reperfusion injury231.
The intensity and quality of reperfusion are important factors regarding
neuronal damage after CA and ROSC. Therefore, in a porcine model of CPR,
maximization of CBF during CPR and after ROSC has been attempted in or-
der to ameliorate the brain damage by using several methods, such as occlu-
sion of the abdominal aorta232, aortic administration of vasopressin205 and ex-
pansion of the plasma volume233. Using the same model, PBN has been found
to reduce the neurological damage even if administered after the incident,
while maximization of CBF during CPR and after ROSC did not improve neu-
rological dysfunction209.
Apart from the reperfusion injury, the ischemic duration per se is likely to
influence the neurological outcome. Development of autonomic dysfunction,
such as spontaneous hyperthermia was found to be related to the duration of
ischemia (≥120 min of temporary MCAO) and associated with ischemic
damage in the hypothalamus. Furthermore, this spontaneous hyperthermia
closely resembled the ischemic injury in a permanent MCAO rat model234.
Experience from recanalized focal ischemic stroke has pointed out that the
role of hyperglycemia as an aggravating factor was inversely related to the
duration of the ischemia171. The validity of this mechanism in the context of
brief hyperglycemic CA will be discussed later.
Thus, it is likely that when the CA is prolonged, the concomitant factors
such as hyperglycemia may play a relatively less important role.

45
Paper III

Resuscitability, blood glucose and hemodynamics


Twenty pigs were openly randomized into the hyperglycemic (H) and the
normoglycemic (N) groups, resulting in 10 and 10 pigs, respectively. No pigs
were excluded.
The blood glucose levels were kept within the target range for 30 min be-
fore CA, then increased in both groups after ROSC with a preserved discrep-
ancy (p˂0.001) between groups and returned to baseline approximately 120
min after ROSC.
The hemodynamic responses were similar in the two groups. No difference
in MAP was seen between groups. A decrease in CO was observed in both
groups immediately after ROSC, returning to baseline at the end of experi-
ment.
There was a non ̶ significant tendency towards increased cumulative de-
mand of norepinephrine in the N group compared with the H group. The vol-
ume of crystalloid fluid boluses was similar in the two groups.

SjvO2 and SssvO2


The HbO2 saturation peaked at 5 min after ROSC at both sites in both groups,
but no significant difference was seen between the groups at any time point.

Cerebral cortical flow


The cortical cerebral blood flow was measured by LDF. There was a tendency
towards higher levels in the H group after ROSC, but not statistically signifi-
cant.

IL-6 and TNF


The proinflammatory cytokines IL-6 and TNF were quantified in sinus sagittal
venous samples at baseline and 180 min after ROSC. IL-6 and TNF increased
significantly compared with baseline in both the N (IL-6:p˂0.05; TNF:
p˂0.05) and H (IL-6:p˂0.01; TNF:p˂0.001). There was no statistical differ-
ence between the groups at any point.

S100β
The baseline concentrations of S100β did not differ between the two groups,
and no distinct peak was observed after ROSC in either group. The concen-
tration of S100β tended to increase over time in the H group, but not in the N

46
group. At the end of the experiment, S100β was increased by +20±10% in the
H group, while in the N group the change was ̶ 9±7% (N vs. H, p˂0.05).

The main finding


There was a subtle, but significant rise in the levels of S100β after 5 ̶ min CA
in the H group compared with the N group (p˂0.05) (fig 4.). The observed
difference in S100β responses developed across the experiment. This finding
contrast to the observation from Paper II, where there was an early peak in
S100β in the post ̶ ischemic phase and a tendency towards increased levels in
H group compared with the N group.

Fig.4. S100β. The baseline concentration of protein S100β did not differ between
groups (a). The concentration of protein S100β tended to increase over time in the H
group, but not in the N group (b). At the end of the experiment, there was a statisti-
cally significant differences in the development of protein S100β (c) (p˂0.05).

S100β has been widely investigated as one of many aspects of acute stroke235-
238
and CA.
The significance of protein S100β as a brain tissue injury biomarker has
been debated239, 240 and its role as a surrogate predictor for cerebral damage in
acute ischemic stroke239 and in clinical CA241 has been discussed.
Dassan et al239 reviewed the literature on S100β in the context of acute is-
chemic stroke. The collected data revealed that S100β levels increase gradu-
ally starting 8 ̶ 10 h after onset of symptoms, peaking at 72 h and decreasing
at 96 h. The use of S100β as a biomarker for diagnosing acute ischemic stroke
thereby is not suitable because of its delayed pharmacokinetics. There is how-
ever substantial evidence from a number of studies that S100β measured be-
yond 24 h after stroke onset significantly correlates with the degree of neuro-
logical deficit and the final infarct volume236, 242, 243. S100β has also a surrogate
role in the research setting of interventional stroke trials and in predicting
functional outcome after ischemic stroke239.
In patients with isolated trauma to extremities, increased levels of S100β
were found (higher serum levels of S100β were seen in the larger bone frac-
tures) and therefore the extra ̶ cerebral source of S100β (from the bone tissue

47
or adipocytes) should be taken into account when interpreting S100β as a pre-
dictor of cerebral brain damage240.
S100β and neurone ̶ specific enolase (NSE) - a neurone ̶ derived enzyme -
are both released after CA244. Serum S100β levels at 24 ̶ 48 h after CA have
been found elevated and correlated significantly with cognitive impairment at
discharge after OOHCA. Furthermore, S100β predicted in ̶ hospital death after
CA and was found to be more useful than NSE in this clinical setting because
of its greater specificity, although the sensitivity of this test was relatively
low241.
However, several issues have to be solved in order to improve the clinical
usefulness of measuring S100β as a biomarker. It must be taken into account
that the variety of techniques used for measuring the S100β protein differ in
sensitivity and specificity of detection. The time points of blood sampling
should also follow the release pattern of S100β, since the change over time of
S100β in the early phase of brain injury more reliable reflects the ongoing
pathological process rather than its magnitude per se163.

Hyperglycemia has been shown to accelerate ischemic brain injury in both


experimental focal cerebral ischemia11 and in humans with non ̶ lacunar
strokes in which reperfusion occurs but not in lacunar strokes in which very
little or no reperfusion takes place10. Furthermore, hyperglycemia before
reperfusion has been revealed to partly counterbalance the favorable effect of
treatment with tPA245.
Alvarez-Sabin et al also showed that the detrimental effect of admission
hyperglycemia was significant only after early recanalization (<3 h). This phe-
nomenon may be a result of the high glucose load at the time of reperfusion
accelerating the enrollment of penumbral tissue into infarction, an effect that
may progressively diminish as the area of penumbra decreases over time of
arterial occlusion171.
This phenomenon has been described in focal stroke, but not after CA. Our
finding would be consistent with the tissue being salvageable early in the
course of the ischemia, and therefore efforts to achieve normoglycemia may
be ineffective once the damage is already established.
Thus, the impact of pre ̶ ischemic hyperglycemia may be more significant
when the CA is short enough.

Cytokines IL-6 and TNF


We measured IL-6 and TNF in sagittal sinus to trace the inflammatory re-
sponse based on our previous results described in Paper II.

48
The most investigated cytokines associated with inflammation in stroke are
TNF, IL-1, IL-6, interleukin 10 (IL-10) and transforming growth factor-β
(TGF-β). While TNF and IL-1 are mainly connected with aggravated brain
injury, IL-6, IL-10 and TGF-β may exert a neuroprotective effect246.
Although IL-6 is proposed mainly as pro ̶ inflammatory cytokine, express-
ing at high levels in the cerebrospinal fluid (CSF) and plasma in patients with
ischemic stroke, the significance of IL-6 in these settings is far from clear.
Clinical studies with stroke patients demonstrated a positive correlation be-
tween serum concentration of IL-6 and in ̶ hospital mortality247. The admission
concentration of IL-6 has been found to be positively correlated with the early
clinical deterioration and the strength of the acute phase ̶ response, the latter
being a susceptible predictor of poor short ̶ term survival248. Furthermore, in a
double ̶ blind clinical trial on patients with acute ischemic stroke, the treatment
with rhIL-1a - a neuroprotective drug - resulted in decreased level of IL-6 and
the patients demonstrated better neurological outcome249.
However, others have found inverse correlation between level of IL-6 with
both infarct volume and final neurological outcome in patients with acute is-
chemic stroke suggesting the neuroprotective role of this cytokine250. The up-
regulation of the neuronal adenosine A1 receptor induced by IL-6 has been
demonstrated essential for neuroprotective effect of IL-6251.
Regarding the relationship between the peripheral levels of TNF and stroke
severity, the data are inconsistent reporting either unchanged or elevated lev-
els of this cytokine251.
An early, but transient activation of TNF together with IL-1β is seen in
response to I/R injury, which in turn initiates a more persisting secondary in-
flammatory reaction mediated by IL-6 and IL-8.
Experimental data with stroke model has shown that the systemic cytokine
response corresponds to that of in the brain252. Increased IL-6 levels in plasma
could be detected as early as 3 h after onset of experimental stroke and from
24 h in the brain253, while increased TNF levels were observed in serum 6 h
and 24 h after MCAO252.

The role of cytokines in response to CA in a porcine model has also been


studied by several investigators254, 255. Niemann et al found a sharp increase in
TNF and IL-1β levels within 30 min under the reperfusion period after CA
and these levels continued elevated for the first 3 h of reperfusion255.
The role of TNF and IL-6 in post ̶ CA syndrome has been highlighted by
several studies256, 257. High plasma levels of TNF and IL-6 have been found to
be positively correlated with mortality257. This finding contrasted with the re-
sults demonstrated by Niemann et al, which revealed elevated TNF but not IL-
6 in response to CA255, but it is in line with our results showing significantly
higher levels of TNF and IL-6 at 3 h after CA.

49
Furthermore, there were no group differences in these cytokines and there-
fore the effect of hyperglycemia on production of inflammatory response in
this model is yet to be established.

Glycemic levels
In Paper II and Paper III the glycemic levels were achieved by glucose-
clamping rather than occurring spontaneously and therefore may not entirely
represent the situation seen in stress ̶ induced hyperglycemic response after
CA.
Experimental models using only glucose administration to achieve hyper-
glycemia have shown that hyperglycemia enhances the generation of ROS
even without ischemia12. This suggests that, at least in part, hyperglycemia
may aggravate cerebral damage by enhancing basal ̶ and ischemia ̶ triggered
production of ROS.
The pathogenesis of hyperglycemia ̶ induced cerebral damage may relate
to the glucose levels, particularly when the glucose concentration exceeds 15-
20 mmol/L, and then secondary lactate acidosis is present56 .
In the present model such high levels were only reached temporarily after
ROSC. Previous data from our laboratory11 and from other experiments in ro-
dents64 have suggested that even lower levels of glucose concentration can
induce pathogenic mechanisms such as oxidative stress, possibly involving
MARK signaling.
The present design was aimed to describe clinically relevant levels using
both tight and more liberal glucose regimens to mimic CA at different pre-
arrest glycemic levels. The ranges we have chosen are debated and varies be-
tween studies. The N range (4 ̶ 5.5 mM) can correspond to the IIT in the Leu-
ven and NICE-SUGAR studies, respectively; while the H range (8.5 ̶ 10 mM)
can represent the liberal targets of these studies90, 96. The efficacy and effec-
tiveness of the Leuven trial protocols have been discussed earlier.
The route and timing of nutrition may have significant impact on the out-
come258. New technologies that can determine accurately and near ̶ continu-
ously the blood glucose together with the computerized decision support sys-
tem would be essential to optimize further the glycemic control82.
Thus, our result of investigating the influence of pre ̶ ischemic hyperglyce-
mia allows few conclusions, although the slight increase of S100β in the H
group compared with the N group is compatible with a subtle cerebral benefit
from the N compared with the H regimen.

50
Paper IV
Global expression profiling of the cerebral cortex from 5 hyperglycemic (H)
and 6 normoglycemic (N) animals was performed. A piece of frontoparietal
cortex was taken post-mortem for subsequent analysis of the transcriptome
using microarrays. The expression of approximately 19000 transcripts were
studied. Analysis revealed significant differences in expression of 102 tran-
scripts at p<0.001.
To correct for multiple testing the Benjamini and Hochberg method was
used259.

The main finding


Among the most strongly differentially regulated genes we identified several
that are involved in glucose transport and metabolism pathways. Furthermore,
a gene coding for solute carrier organic anion polypeptide, a gene coding for
a neuronal membrane associated with synapses and two microRNAs have
been found significantly different regulated.
Solute carrier family 5, member 11 (SLC5A11), a sodium/glucose co-trans-
porter was more highly expressed in N animals (p=4.3x10-7) (fig. 5A).
3-hydroxybutyrate dehydrogenase type 1 (BDH1) levels more highly ex-
pressed in H animals (p=1.5x10-5) (fig.5.B).
Solute carrier organic anion transporter family, member 1A2, (SLCO1A2)
was up-regulated in the N group compared with the H group (p=8.33x10-6)
(fig. 5C).
Contactin 3 (CNTN3) was the transcript that differed most strongly between
the groups (p= 3.2x10-8) and was more highly expressed in H animals (fig.
5D).
There were two microRNAs among the top genes, mir-664 with higher ex-
pression in N animals (p= 1.5x10-5) and mir9-2 with higher expression in H
animals (p= 3.3x10-5) (fig. 5E and F).

51
Fig.5.Solute carrier family 5, member 11 (SLC5A11) was more highly expressed in N
pigs (A) whereas 3-hydroxybutyrate dehydrogenase type 1 (BDH1) levels were higher
in H pigs (B). Solute carrier organic anion transporter family, member 1A2,
(SLCO1A2) was up ̶ regulated in N pigs (C). Contactin 3 (CNTN3) was more highly
expressed in H pigs (D). Mir-664 expressed more highly in N pigs (E), while mir9-2
was more highly expressed in H pigs (F).

52
In order to assess whether the identified genes could be linked together in re-
lated biologic processes, the transcripts were analyzed using the DAVID func-
tional annotation tool260.
This revealed significant enrichment for several biological pathways, for
instance membrane processes (p=2.1x10-6) and ion transport (p=0.0012) as al-
ready suggested by the presence of several solute carrier genes among the
most differentially regulated genes.
Also, glycosylation site (p=0.0015) and glycoproteins (p=0.0026) were en-
riched processes amongst the identified genes verifying that glucose related
pathways were differentially activated between the conditions.
Regulation of actin cytoskeleton organization (p=0.0021) and actin fila-
ment ̶ based process (p=0.0023), were enriched, suggesting that intracellu-
lar/structural processes were also differentially activated.

To further visualize the data hierarchical clustering was used to create heat
maps using the normalized gene expression data for each sample (fig. 6).

53
Fig.6. The pattern of expression for the 102 genes with p<0.001.
A heat map illustrating the pattern of expression for the 102 genes with p<0.001 was
constructed, with hierarchical clustering depicting the internal relationships between
the genes. Red color indicates up ̶ regulation and green color down ̶ regulation. The
more intense the color the stronger the difference. Transcript ID and gene symbols
are in the right column.

54
SLC5A11
The glucose transport into the cells are mediated by two transport mecha-
nisms, such as Na+ independent facilitated diffusion system accomplished by
a family of glucose transporters (GLUTs)261 and the other is an ATP-depend-
ent Na+-monosaccharide co ̶ transport system (SGLT).
In the first system, the transporters facilitate the travel of glucose down
from a high concentration to a lower one without requiring energy. In the latter
system, also called as a sodium/glucose co ̶ transport family (SLC5), the
movement of glucose is combined with the concentration gradient of Na+ and
requires energy since this transport is executed against the concentration gra-
dient262. This mechanism could play an important role in the brain for neuron
survival, when the concentration of glucose is low (e.g. hypoxia or ischemia)
by facilitating the prompt membrane insertion of SGLTs from the intracellular
stores and may thereby contribute to the endogenous neuroprotection263.
The distribution of functional SGLTs in rat brain has revealed their high
abundance in neurons sensitive to hypoxia, such as in the cerebellum, hippo-
campus and frontoparietal cortex263.
The SLC5 protein encoding gene family includes more than 220 members
found in animal and bacteria cells. The human SLC5 gene family incorporates
12 human genes and are expressed throughout the body and also in the CNS.
The encoded proteins are implicated in processes such as Na+ cotransporter
for sugars, myo-inositol, iodide, short-chain fatty acids and choline.
SLC5A11, member 11 of SLC5, is also known as human sodium inositol
cotransporter 2 (SMIT2). The encoded protein (SMIT2) is responsible for
transport of myo-Inositol (MI) and its epimer D-chiro-Inositol (DCI) (derivate
from glucose 6-phosphatase), compounds that are involved in mediating insu-
lin bioactivity. SMIT2 expression and function measured by DCI transport
may be influenced by insulin and glucose or both264.
SLC5A11 may provoke apoptosis via the TNF-α, programmed cell death 1
(PDCD1) pathway, although this has yet to be confirmed265. It has also been
suggested that SLC5A11 may act as an autoimmune modifier gene265 in hu-
mans.
The current finding that SLC5A11 gene was down-regulated in H group
may in this context lend support to our previous results that suggest aggravated
brain injury in the H group.

BDH1
The BDH1 gene is a protein ̶ coding gene. The encoded protein D-3-hydroxy-
butyrate dehydrogenase (HBDH) - a member of short ̶ chain dehydrogenase/re-
ductase (SDR) family - is a catalytic mitochondrial enzyme that reversible oxi-
dizes β-hydroxybutyrate (β-OHB) to acetoacetate (AcAc) using NAD+ as a co-
enzyme. β-OHB, AcAc and acetone are termed together as ketone bodies.

55
Ketone bodies are produced mainly in the liver and can be absorbed by
brain tissue because they are water ̶ soluble. Once in the brain tissue, they are
decomposed to acetyl-CoA, which - by oxidizing via the tricarboxylic acid
(TCA) cycle - constitutes energy266.
The brain has the capability to adjust its metabolism to ketone bodies which
may reveal a mechanism of cerebral metabolic adaptation267. During pro-
longed period of starving, the increased synthesis of ketone bodies via abun-
dant acetyl CoA leads to a decrease need for glucose as energy supply268. The
brain may be more susceptible to this ketone adaptation during acute neuro-
pathologic conditions such as ischemia and it may provide neuroprotection by
inducing rapid metabolic changes. These adaptive mechanisms are the
promptly increased levels of monocarboxylate tranporters269 - which convey
ketones, lactate and pyruvate – and the up ̶ regulated enzymatic activity of
HBDH together with the high availability of endogenously or exogenously
supplied ketone bodies may drive the brain to rely on ketones rather than glu-
cose as the metabolic position of glucose as primary fuel is no longer guaran-
teed266.
Experimental studies using intravenous infusion of β-OHB both pre ̶ and
postinjury showed decreased infarct volume and lipid peroxidation270, de-
creased oedema formation and attenuated ATP reduction271.
Our result showing up ̶ regulated BDH1 in hyperglycemic CA would sug-
gest that this mechanism could theoretically ameliorate the ischemic injury.

SLCO1A2
SLCO1A2 was significantly down ̶ regulated in H compared with N animals.
Organic anion transporting polypeptides (OATPs) encoded by the SCLO
(solute carrier organic anion transporter family) gene family are expressed in
epithelial cells originating from various organs including the brain. Substrates
of OATPs incorporate bile salts, steroid hormones and their conjugates, thy-
roid hormones, inflammatory mediators and numerous drugs272. Since many
members of OATPs are expressed in cells - whose functions include regula-
tion of solute exchange between compartments, such as BBB and choroid
plexus epithelium - OATPs play an important role in preservation of tissue
bioavailability and distribution of different drugs273.
OATP1A2 encoded by the gene SLCO1A2 is found abundantly in the hu-
man brain, especially in the frontal cortex and hippocampus, but also in capil-
lary endothelial cells, and it has a role in transport activity and reuptake for
neuroactive peptides, such as substance P and vasoactive intestinal peptide273.
OATP1A2 has also demonstrated an excessive transport activity for dehydroe-
piandrosterone sulfate (DHEAS), which in turn inhibits the GABAA recep-
tors274. This activity furthermore could indirectly lead to neuroprotection.

56
Referring to this mechanism, down ̶ regulation of SLCO1A2 gene in hyper-
glycemic compared with normoglycemic animals would theoretically contrib-
ute to the ischemic damage.

CNTN3
CNTN3, contactin-3 was the most significantly differing gene between the
groups, an immunoglobulin superfamily member expressed exclusively in the
nervous system. The contactin family includes six cell adhesion molecules and
is involved in developmental processes, such as neural cell migration, neurite
overgrowth and axonal guidance275.
CNTN3 is expressed uniquely in subsets of neurons, for instance in the neu-
rons of superficial layers of the cerebral cortex276 and is essential to sustain
the normal function of Golgi apparatus277. Members of the contactin family
have been linked to autism spectrum disease278, but otherwise little is known
about the role of CNTN3 in human disease.

MicroRNA
MicroRNAs (miRNAs) represent a novel subgroup of untranslated regulatory
RNAs also known as non ̶ coding RNAs, and are mainly situated as mi-
croRNA nuclear genes either in introns or in intergenic regions of genome.
After going through several steps of maturation, the large primary microRNAs
convert to the final transcript of miRNAs containing 18 ̶ 25 nucleotides. MiR-
NAs control gene expression either transcriptionally by regulating transcrip-
tion factors, or post ̶ transcriptionally by down ̶ regulation of mRNA via trans-
lation/ transcription inhibition279.
Approximately 30 % of all genes are controlled by miRNAs and 70% of
known miRNAs are expressed in the brain. Seven brain specific miRNAs
(miR-9, miR-124a, miR-124b, miR-135, miR-153, miR-183 and miR-219)
have been demonstrated to be involved in regulating neuronal differentiation,
maturation and cell survival in the mammalian brain280.
The role of miR-9 has been revealed in proper neuronal maturation by tar-
geting a transcription protein connected with speech and language develop-
ment, in neurogenesis of pallium and subpallium by regulating the expression
of genes involving in progenitor proliferation and differentiation, and in brain
development by regulating the dynamics of neuronal stem cells281.
The role of miR-9 in cerebral pathogenesis has also been thoroughly inves-
tigated. Expression of miR-9 has been found expressed particularly in the hip-
pocampus and in medial frontal gyrus in patients with Alzheimer disease282.
Mir-9 was found to be the second most highly expressed miRNA in hippo-
campus after experimental global ischemia followed by 30 min of reperfusion.
Considering that global ischemia negatively affects the translational recovery283
and that the predicted targets of miR-9 are the translation initiation factors

57
EIF4E and EIF5 and also the transcription factors cAMP response element bind-
ing protein 5 (CREB5) and CREB2F, the observed upregulation of miR-9 could
be partly responsible for the inhibition of gene expression in that study284.
In our study, the miR-9 was significantly up-regulated in hyperglycemic
compared with normoglycemic CA. Whether this represented a difference in
translational recovery between the groups was not investigated.
MicroRNA profiling on rats subjected to transient MAO revealed a distinct
microRNA expression pattern in I/R injury demonstrating the dynamic nature
of expression during ischemic injury. Up ̶ regulation of the gene MMP9 with
down ̶ regulation of the corresponding miR-132 and miR-664 has been re-
vealed both in the early and late reperfusion phase285.
MiR-664 was found down ̶ regulated at 4 of 5 reperfusion time points star-
ing 3 h after reestablished perfusion286. Furthermore, bioinformatics analysis
of that study suggested the relationship between altered expression of miRNA
and their mRNA targets, many of which are recognized in inflammation, neu-
roprotection, transcription and ion homeostasis
Our results indicate that the down-regulation of miR-664 after cerebral is-
chemia can be further enhanced by hyperglycemia after CA, but the im-
portance of such a mechanism remains to be elucidated.
Furthermore, the nature of the miRNA-mRNA interaction is multifaceted,
since a small number of miRNA can affect a substantial number of mRNA, which
in turn can influence the specific miRNA expression in a feed ̶ forward manner
and thereby make it difficult to predict the overall effect of a specific miRNA.
It is also noteworthy to point out that the altered regulation of miRNAs in
the early phase of reperfusion demonstrated by earlier studies286 is compatible
with the framework of the present study.

Data from miRNA profiling studies performed on experimental cerebral is-


chemia models may also open new therapeutic possibilities. For example, the
antagonization of some specific miRNAs showed neuroprotective character-
istics that could be studied in future experiments in order to apply this possible
therapeutic approach in prevention of postischemic pathophysiological conse-
quences or stimulation of tissue regeneration286.
Furthermore, the discovery of ischemic preconditioning (ICP) - since this
process necessitates new protein synthesis - has revealed the role of miRNAs
as regulators of these mechanisms287. According a recent experimental study,
even a single period of ischemia has been proved neuroprotective288. Although
our aim with this study was not to induce ICP, this relatively short period of
lack of oxygen might act as a preconditioning stimulus. Revealing specific
patterns of the up ̶ and down ̶ regulation of the miRNAs can contribute to
better understanding the brain`s own endogen repair mechanisms which can
lead to more tailor ̶ made therapeutic approaches for cerebral ischemia289.

58
Conclusions

 S-PBN may improve short-term neurological performance after is-


chemic hyperglycemic stroke in rats.
 Hyperglycemia elevated cerebral oxygenation after 12–min of CA
compared with normoglycemia.
 Hyperglycemia elevated S100β after 5–min of CA compared with
normoglycemia.
 Hyperglycemia lead to differential early gene expression after 5 ̶
min of CA compared with normoglycemia.
 

59
Conclusive remarks and future perspective

We found that S-PBN may improve short-term neurological performance in


experimental hyperglycemic focal cerebral ischemia.
Hyperglycemia is common in CA but its role as a marker of severity of
illness and pathogenic influence on hyperglycemic ischemic cerebral injury is
unclear.
We observed higher cerebral tissue oxygenation as well as tendencies to-
ward higher concentrations of S100β and 15-keto-dihydro PGF2α. These ob-
servations as well as the biological significance of these findings will need to
be addressed in future studies.
The cerebral injury from prolonged CA is likely to be severe and therefore
the contribution of concomitant hyperglycemia may be of minor importance
in comparison with the ischemic damage. Experience from recanalized focal
ischemic stroke indicates that the aggravating role of hyperglycemia is in-
versely related to the duration of the ischemia. In line with this, we observed
a subtle, but significant rise in the levels of S100β after reducing the duration
of the CA from 12 to 5 min in the hyperglycemic group compared with the
normoglycemia group.
The biology of hyperglycemia in CA is relatively unexplored and we also
find the prospect of descriptive studies interesting. To that end we performed
a global transcriptome analysis of brains from hyperglycemic and normogly-
cemic CA and found that hyperglycemia during CA leads to different early
gene expression compared with normoglycemia.
To date, the therapeutic options for treating the acute phase of ischemic
stroke is restricted to tPA. However, the usefulness of tPA treatment is limited
by several aspects, such as its narrow therapeutic window, the risk of hemor-
rhagic transformation and its aggravating effect on alteration of integrity of
the neurovascular unit290. Likewise, the applicability of mild induced hypo-
thermia or targeted temperature control in treating survivors of CA also still
has several unanswered questions291.
Therefore these transcriptional changes found by transcriptome analysis
constitute targets for further investigation in future experiments that explore
whether and/or how the significance of glycemic control and its relation to the
time point and duration of cerebral ischemia play a role in the pathophysiology
of cerebral ischemic damage. Ultimately, such research can contribute toward
to develop our strategy for preventing and treating neurological complication
in survivors of ischemic stroke and CA.

60
Acknowledgements

I would like to express my sincere gratitude to everyone who has contributed


to my research, but especially to following persons:

Fredrik Lennmyr, my main supervisor for his invaluable help with my re-
search and his imperturbable support in all conceivable situations. Thank you
so much for EVERYTHING Fredrik!
Lars Wiklund, my co ̶ supervisor for co ̶ authorships, for providing time for
research, for all kind of help in the laboratory, and for stimulating scientific
discussions.
Sten Rubertsson for giving me time for research and for comforting me with
wise advice on how to teach medical students.
Torsten Gordh for offering me a position and for always showing great hu-
manity.
Niklas Marklund for critically reviewing Paper I and sharing his expertise.
Samar Basu, Maria Bergquist and Anders Larsson for co ̶ authorships, for
performing biochemical analysis and for their insightful review of respective
manuscript.
Ulla Svensson, Anders Nordgren, Monica Hall and Elisabeth Pettersson
for their invaluable help at the animal laboratory and for their patient in teach-
ing me the technical basics of animal experimentation.
Belinda Fridman for performing microarray analysis and for her helpfulness
in teaching me the very basic of bioinformatics analysis.
Hanna Göransson Kultima for carefully reviewing the microarray data anal-
ysis of manuscript IV.
Malin Olsson for technical implementation of microarray analysis.
Charlotte Israelsson for performing the RNA preparation.
Rickard Lindblom, my co ̶ author and mentor in biogenetic. Thank you for
helping me with both technical and theoretical knowledge and for introducing
me to the “real field” of research!
Katja Andersson and Marianne Tufveson for all their help with logistics
and administration.

61
All my colleagues working at the Department of Anesthesiology and Inten-
sive Care, Uppsala University Hospital, Sweden for working hard while I
was off duty and busy with my research.

62
Funding

Grants were obtained from Erik, Karin and Gösta Selander’s Foundation (Pa-
per I, Paper III and Paper IV) and from the Laerdal Foundation for Acute
Medicine (Paper II and Paper III).

63
References

1. Sutherland BA, Minnerup J, Balami JS, Arba F, Buchan AM,


Kleinschnitz C. Neuroprotection for ischaemic stroke: translation
from the bench to the bedside. Int J Stroke 2012; 7: 407-18.
2. Tissue plasminogen activator for acute ischemic stroke. The National
Institute of Neurological Disorders and Stroke rt-PA Stroke Study
Group. N Engl J Med 1995; 333: 1581-7.
3. Minnerup J, Sutherland BA, Buchan AM, Kleinschnitz C.
Neuroprotection for stroke: current status and future perspectives. Int
J Mol Sci 2012; 13: 11753-72.
4. Astrup J, Siesjo BK, Symon L. Thresholds in cerebral ischemia - the
ischemic penumbra. Stroke 1981; 12: 723-5.
5. Nolan JP, Morley PT, Hoek TL, Hickey RW, Advancement Life
support Task Force of the International Liaison committee on R.
Therapeutic hypothermia after cardiac arrest. An advisory statement
by the Advancement Life support Task Force of the International
Liaison committee on Resuscitation. Resuscitation 2003; 57: 231-5.
6. Calle PA, Buylaert WA, Vanhaute OA. Glycemia in the post-
resuscitation period. The Cerebral Resuscitation Study Group.
Resuscitation 1989; 17 Suppl: S181-8; discussion S99-206.
7. Scott JF, Robinson GM, French JM, O'Connell JE, Alberti KG, Gray
CS. Prevalence of admission hyperglycaemia across clinical subtypes
of acute stroke. Lancet 1999; 353: 376-7.
8. Kagansky N, Levy S, Knobler H. The role of hyperglycemia in acute
stroke. Arch Neurol 2001; 58: 1209-12.
9. Lanier WL, Stangland KJ, Scheithauer BW, Milde JH, Michenfelder
JD. The effects of dextrose infusion and head position on neurologic
outcome after complete cerebral ischemia in primates: examination of
a model. Anesthesiology 1987; 66: 39-48.
10. Bruno A, Biller J, Adams HP, Jr., Clarke WR, Woolson RF, Williams
LS, Hansen MD. Acute blood glucose level and outcome from
ischemic stroke. Trial of ORG 10172 in Acute Stroke Treatment
(TOAST) Investigators. Neurology 1999; 52: 280-4.
11. Farrokhnia N, Roos MW, Terent A, Lennmyr F. Differential early
mitogen-activated protein kinase activation in hyperglycemic
ischemic brain injury in the rat. Eur J Clin Invest 2005; 35: 457-63.

64
12. Li PA, Liu GJ, He QP, Floyd RA, Siesjo BK. Production of hydroxyl
free radical by brain tissues in hyperglycemic rats subjected to
transient forebrain ischemia. Free Radic Biol Med 1999; 27: 1033-40.
13. Capes SE, Hunt D, Malmberg K, Pathak P, Gerstein HC. Stress
hyperglycemia and prognosis of stroke in nondiabetic and diabetic
patients: a systematic overview. Stroke 2001; 32: 2426-32.
14. Beiser DG, Carr GE, Edelson DP, Peberdy MA, Hoek TL.
Derangements in blood glucose following initial resuscitation from in-
hospital cardiac arrest: a report from the national registry of
cardiopulmonary resuscitation. Resuscitation 2009; 80: 624-30.
15. Steingrub JS, Mundt DJ. Blood glucose and neurologic outcome with
global brain ischemia. Crit Care Med 1996; 24: 802-6.
16. Kleikers PW, Wingler K, Hermans JJ, Diebold I, Altenhofer S,
Radermacher KA, Janssen B, Gorlach A, Schmidt HH. NADPH
oxidases as a source of oxidative stress and molecular target in
ischemia/reperfusion injury. J Mol Med (Berl) 2012; 90: 1391-406.
17. Dietrich WD. Morphological manifestations of reperfusion injury in
brain. Ann N Y Acad Sci 1994; 723: 15-24.
18. Durukan A, Tatlisumak T. Acute ischemic stroke: overview of major
experimental rodent models, pathophysiology, and therapy of focal
cerebral ischemia. Pharmacol Biochem Behav 2007; 87: 179-97.
19. Lennmyr F, Karlsson S, Gerwins P, Ata KA, Terent A. Activation of
mitogen-activated protein kinases in experimental cerebral ischemia.
Acta Neurol Scand 2002; 106: 333-40.
20. Dawson VL, Dawson TM, London ED, Bredt DS, Snyder SH. Nitric
oxide mediates glutamate neurotoxicity in primary cortical cultures.
Proc Natl Acad Sci U S A 1991; 88: 6368-71.
21. Pei L, Li Y, Zhang GY, Cui ZC, Zhu ZM. Mechanisms of regulation
of tyrosine phosphorylation of NMDA receptor subunit 2B after
cerebral ischemia/reperfusion. Acta Pharmacol Sin 2000; 21: 695-
700.
22. Paul R, Zhang ZG, Eliceiri BP, Jiang Q, Boccia AD, Zhang RL,
Chopp M, Cheresh DA. Src deficiency or blockade of Src activity in
mice provides cerebral protection following stroke. Nat Med 2001; 7:
222-7.
23. Dikic I, Tokiwa G, Lev S, Courtneidge SA, Schlessinger J. A role for
Pyk2 and Src in linking G-protein-coupled receptors with MAP kinase
activation. Nature 1996; 383: 547-50.
24. Kanterewicz BI, Knapp LT, Klann E. Stimulation of p42 and p44
mitogen-activated protein kinases by reactive oxygen species and
nitric oxide in hippocampus. J Neurochem 1998; 70: 1009-16.
25. Alessandrini A, Namura S, Moskowitz MA, Bonventre JV. MEK1
protein kinase inhibition protects against damage resulting from focal
cerebral ischemia. Proc Natl Acad Sci U S A 1999; 96: 12866-9.

65
26. Herdegen T, Claret FX, Kallunki T, Martin-Villalba A, Winter C,
Hunter T, Karin M. Lasting N-terminal phosphorylation of c-Jun and
activation of c-Jun N-terminal kinases after neuronal injury. J
Neurosci 1998; 18: 5124-35.
27. Pedram A, Razandi M, Levin ER. Extracellular signal-regulated
protein kinase/Jun kinase cross-talk underlies vascular endothelial cell
growth factor-induced endothelial cell proliferation. J Biol Chem
1998; 273: 26722-8.
28. Raingeaud J, Gupta S, Rogers JS, Dickens M, Han J, Ulevitch RJ,
Davis RJ. Pro-inflammatory cytokines and environmental stress cause
p38 mitogen-activated protein kinase activation by dual
phosphorylation on tyrosine and threonine. J Biol Chem 1995; 270:
7420-6.
29. Yoshizumi M, Abe J, Haendeler J, Huang Q, Berk BC. Src and Cas
mediate JNK activation but not ERK1/2 and p38 kinases by reactive
oxygen species. J Biol Chem 2000; 275: 11706-12.
30. Droge W. Free radicals in the physiological control of cell function.
Physiol Rev 2002; 82: 47-95.
31. Lewen A, Matz P, Chan PH. Free radical pathways in CNS injury. J
Neurotrauma 2000; 17: 871-90.
32. Shohami E, Beit-Yannai E, Horowitz M, Kohen R. Oxidative stress in
closed-head injury: brain antioxidant capacity as an indicator of
functional outcome. J Cereb Blood Flow Metab 1997; 17: 1007-19.
33. Love S. Oxidative stress in brain ischemia. Brain Pathol 1999; 9: 119-
31.
34. Kader A, Frazzini VI, Solomon RA, Trifiletti RR. Nitric oxide
production during focal cerebral ischemia in rats. Stroke 1993; 24:
1709-16.
35. Thomas DD, Ridnour LA, Isenberg JS, Flores-Santana W, Switzer
CH, Donzelli S, Hussain P, Vecoli C, Paolocci N, Ambs S, Colton CA,
Harris CC, Roberts DD, Wink DA. The chemical biology of nitric
oxide: implications in cellular signaling. Free Radic Biol Med 2008;
45: 18-31.
36. Iadecola C. Bright and dark sides of nitric oxide in ischemic brain
injury. Trends Neurosci 1997; 20: 132-9.
37. Koppenol WH. The basic chemistry of nitrogen monoxide and
peroxynitrite. Free Radic Biol Med 1998; 25: 385-91.
38. Beckman JS, Koppenol WH. Nitric oxide, superoxide, and
peroxynitrite: the good, the bad, and ugly. Am J Physiol 1996; 271:
C1424-37.
39. Pundik S, Xu K, Sundararajan S. Reperfusion brain injury: focus on
cellular bioenergetics. Neurology 2012; 79: S44-51.
40. Fraser PA. The role of free radical generation in increasing
cerebrovascular permeability. Free Radic Biol Med 2011; 51: 967-77.

66
41. Nishino T, Tamura I. The mechanism of conversion of xanthine
dehydrogenase to oxidase and the role of the enzyme in reperfusion
injury. Adv Exp Med Biol 1991; 309A: 327-33.
42. Siesjo BK, Wieloch T. Cerebral metabolism in ischaemia:
neurochemical basis for therapy. Br J Anaesth 1985; 57: 47-62.
43. McCord JM. Oxygen-derived free radicals in postischemic tissue
injury. N Engl J Med 1985; 312: 159-63.
44. Ames BN, Cathcart R, Schwiers E, Hochstein P. Uric acid provides an
antioxidant defense in humans against oxidant- and radical-caused
aging and cancer: a hypothesis. Proc Natl Acad Sci U S A 1981; 78:
6858-62.
45. Marklund N, Ostman B, Nalmo L, Persson L, Hillered L.
Hypoxanthine, uric acid and allantoin as indicators of in vivo free
radical reactions. Description of a HPLC method and human brain
microdialysis data. Acta Neurochir (Wien) 2000; 142: 1135-41;
discussion 41-2.
46. Cave AC, Brewer AC, Narayanapanicker A, Ray R, Grieve DJ,
Walker S, Shah AM. NADPH oxidases in cardiovascular health and
disease. Antioxid Redox Signal 2006; 8: 691-728.
47. Diebold I, Flugel D, Becht S, Belaiba RS, Bonello S, Hess J,
Kietzmann T, Gorlach A. The hypoxia-inducible factor-2alpha is
stabilized by oxidative stress involving NOX4. Antioxid Redox Signal
2010; 13: 425-36.
48. Elahi MM, Kong YX, Matata BM. Oxidative stress as a mediator of
cardiovascular disease. Oxid Med Cell Longev 2009; 2: 259-69.
49. Bemeur C, Ste-Marie L, Montgomery J. Increased oxidative stress
during hyperglycemic cerebral ischemia. Neurochem Int 2007; 50:
890-904.
50. Pulsinelli WA, Levy DE, Sigsbee B, Scherer P, Plum F. Increased
damage after ischemic stroke in patients with hyperglycemia with or
without established diabetes mellitus. Am J Med 1983; 74: 540-4.
51. Melamed E. Reactive hyperglycaemia in patients with acute stroke. J
Neurol Sci 1976; 29: 267-75.
52. Moulin T, Tatu L, Crepin-Leblond T, Chavot D, Berges S, Rumbach T.
The Besancon Stroke Registry: an acute stroke registry of 2,500
consecutive patients. Eur Neurol 1997; 38: 10-20.
53. Nedergaard M. Transient focal ischemia in hyperglycemic rats is
associated with increased cerebral infarction. Brain Res 1987; 408:
79-85.
54. Farrokhnia N, Roos MW, Terent A, Lennmyr F. Experimental
treatment for focal hyperglycemic ischemic brain injury in the rat. Exp
Brain Res 2005; 167: 310-4.
55. Yip PK, He YY, Hsu CY, Garg N, Marangos P, Hogan EL. Effect of
plasma glucose on infarct size in focal cerebral ischemia-reperfusion.
Neurology 1991; 41: 899-905.

67
56. Siesjo BK, Siesjo P. Mechanisms of secondary brain injury. Eur J
Anaesthesiol 1996; 13: 247-68.
57. Bemeur C, Ste-Marie L, Desjardins P, Vachon L, Butterworth RF,
Hazell AS, Montgomery J. Dehydroascorbic acid normalizes several
markers of oxidative stress and inflammation in acute hyperglycemic
focal cerebral ischemia in the rat. Neurochem Int 2005; 46: 399-407.
58. Balami JS, Hadley G, Sutherland BA, Karbalai H, Buchan AM. The
exact science of stroke thrombolysis and the quiet art of patient
selection. Brain 2013; 136: 3528-53.
59. Ljunggren B, Norberg K, Siesjo BK. Influence of tissue acidosis upon
restitution of brain energy metabolism following total ischemia. Brain
Res 1974; 77: 173-86.
60. Katsura K, Asplund B, Ekholm A, Siesjo BK. Extra- and Intracellular
pH in the Brain During Ischaemia, Related to Tissue Lactate Content
in Normo- and Hypercapnic rats. Eur J Neurosci 1992; 4: 166-76.
61. Smith ML, von Hanwehr R, Siesjo BK. Changes in extra- and
intracellular pH in the brain during and following ischemia in
hyperglycemic and in moderately hypoglycemic rats. J Cereb Blood
Flow Metab 1986; 6: 574-83.
62. Siesjo BK, Katsura K, Kristian T. Acidosis-related damage. Adv
Neurol 1996; 71: 209-33; discussion 34-6.
63. Kurihara J, Katsura K, Siesjo BK, Wieloch T. Hyperglycemia and
hypercapnia differently affect post-ischemic changes in protein
kinases and protein phosphorylation in the rat cingulate cortex. Brain
Res 2004; 995: 218-25.
64. Li PA, He QP, Yi-Bing O, Hu BR, Siesjo BK. Phosphorylation of
extracellular signal-regulated kinase after transient cerebral ischemia
in hyperglycemic rats. Neurobiol Dis 2001; 8: 127-35.
65. Hillered L, Smith ML, Siesjo BK. Lactic acidosis and recovery of
mitochondrial function following forebrain ischemia in the rat. J
Cereb Blood Flow Metab 1985; 5: 259-66.
66. Li PA, He QP, Miyashita H, Howllet W, Siesjo BK, Shuaib A.
Hypothermia ameliorates ischemic brain damage and suppresses the
release of extracellular amino acids in both normo- and hyperglycemic
subjects. Exp Neurol 1999; 158: 242-53.
67. Brownlee M. Biochemistry and molecular cell biology of diabetic
complications. Nature 2001; 414: 813-20.
68. Giacco F, Brownlee M. Oxidative stress and diabetic complications.
Circ Res 2010; 107: 1058-70.
69. Hoshiyama M, Li B, Yao J, Harada T, Morioka T, Oite T. Effect of
high glucose on nitric oxide production and endothelial nitric oxide
synthase protein expression in human glomerular endothelial cells.
Nephron Exp Nephrol 2003; 95: e62-8.

68
70. Peters O, Back T, Lindauer U, Busch C, Megow D, Dreier J, Dirnagl
U. Increased formation of reactive oxygen species after permanent
and reversible middle cerebral artery occlusion in the rat. J Cereb
Blood Flow Metab 1998; 18: 196-205.
71. Wei J, Quast MJ. Effect of nitric oxide synthase inhibitor on a
hyperglycemic rat model of reversible focal ischemia: detection of
excitatory amino acids release and hydroxyl radical formation. Brain
Res 1998; 791: 146-56.
72. Cherubini A, Ruggiero C, Polidori MC, Mecocci P. Potential markers
of oxidative stress in stroke. Free Radic Biol Med 2005; 39: 841-52.
73. Laver S, Farrow C, Turner D, Nolan J. Mode of death after admission
to an intensive care unit following cardiac arrest. Intensive Care Med
2004; 30: 2126-8.
74. Abramson NS, Safar P, Detre KM, Kelsey SF, Monroe J, Reinmuth O,
Snyder JV. Neurologic recovery after cardiac arrest: effect of duration
of ischemia. Brain Resuscitation Clinical Trial I Study Group. Crit
Care Med 1985; 13: 930-1.
75. Hallstrom AP, Cobb LA, Swain M, Mensinger K. Predictors of
hospital mortality after out-of-hospital cardiopulmonary resuscitation.
Crit Care Med 1985; 13: 927-9.
76. Pulsinelli WA, Waldman S, Rawlinson D, Plum F. Moderate
hyperglycemia augments ischemic brain damage: a neuropathologic
study in the rat. Neurology 1982; 32: 1239-46.
77. Welsh FA, Ginsberg MD, Rieder W, Budd WW. Deleterious effect of
glucose pretreatment on recovery from diffuse cerebral ischemia in
the cat. II. Regional metabolite levels. Stroke 1980; 11: 355-63.
78. Longstreth WT, Jr., Diehr P, Cobb LA, Hanson RW, Blair AD.
Neurologic outcome and blood glucose levels during out-of-hospital
cardiopulmonary resuscitation. Neurology 1986; 36: 1186-91.
79. Langhelle A, Tyvold SS, Lexow K, Hapnes SA, Sunde K, Steen PA.
In-hospital factors associated with improved outcome after out-of-
hospital cardiac arrest. A comparison between four regions in Norway.
Resuscitation 2003; 56: 247-63.
80. Skrifvars MB, Pettila V, Rosenberg PH, Castren M. A multiple logistic
regression analysis of in-hospital factors related to survival at six
months in patients resuscitated from out-of-hospital ventricular
fibrillation. Resuscitation 2003; 59: 319-28.
81. Oksanen T, Skrifvars MB, Varpula T, Kuitunen A, Pettila V, Nurmi J,
Castren M. Strict versus moderate glucose control after resuscitation
from ventricular fibrillation. Intensive Care Med 2007; 33: 2093-100.
82. Egi M, Finfer S, Bellomo R. Glycemic control in the ICU. Chest
2011; 140: 212-20.
83. Mizock BA. Alterations in carbohydrate metabolism during stress: a
review of the literature. Am J Med 1995; 98: 75-84.

69
84. McCowen KC, Malhotra A, Bistrian BR. Stress-induced
hyperglycemia. Crit Care Clin 2001; 17: 107-24.
85. Khaodhiar L, McCowen K, Bistrian B. Perioperative hyperglycemia,
infection or risk? Curr Opin Clin Nutr Metab Care 1999; 2: 79-82.
86. Robinson LE, van Soeren MH. Insulin resistance and hyperglycemia
in critical illness: role of insulin in glycemic control. AACN Clin
Issues 2004; 15: 45-62.
87. Gearhart MM, Parbhoo SK. Hyperglycemia in the critically ill patient.
AACN Clin Issues 2006; 17: 50-5.
88. Christiansen C, Toft P, Jorgensen HS, Andersen SK, Tonnesen E.
Hyperglycaemia and mortality in critically ill patients. A prospective
study. Intensive Care Med 2004; 30: 1685-8.
89. Van den Berghe G, Wilmer A, Milants I, Wouters PJ, Bouckaert B,
Bruyninckx F, Bouillon R, Schetz M. Intensive insulin therapy in
mixed medical/surgical intensive care units: benefit versus harm.
Diabetes 2006; 55: 3151-9.
90. van den Berghe G, Wouters P, Weekers F, Verwaest C, Bruyninckx F,
Schetz M, Vlasselaers D, Ferdinande P, Lauwers P, Bouillon R.
Intensive insulin therapy in critically ill patients. N Engl J Med 2001;
345: 1359-67.
91. Krinsley JS. Effect of an intensive glucose management protocol on
the mortality of critically ill adult patients. Mayo Clin Proc 2004; 79:
992-1000.
92. Van den Berghe G, Wilmer A, Hermans G, Meersseman W, Wouters
PJ, Milants I, Van Wijngaerden E, Bobbaers H, Bouillon R. Intensive
insulin therapy in the medical ICU. N Engl J Med 2006; 354: 449-61.
93. Van den Berghe G, Wouters PJ, Bouillon R, Weekers F, Verwaest C,
Schetz M, Vlasselaers D, Ferdinande P, Lauwers P. Outcome benefit
of intensive insulin therapy in the critically ill: Insulin dose versus
glycemic control. Crit Care Med 2003; 31: 359-66.
94. Dellinger RP, Levy MM, Carlet JM, Bion J, Parker MM, Jaeschke R,
Reinhart K, Angus DC, Brun-Buisson C, Beale R, Calandra T,
Dhainaut JF, Gerlach H, Harvey M, Marini JJ, Marshall J, Ranieri M,
Ramsay G, Sevransky J, Thompson BT, Townsend S, Vender JS,
Zimmerman JL, Vincent JL, International Surviving Sepsis Campaign
Guidelines C, American Association of Critical-Care N, American
College of Chest P, American College of Emergency P, Canadian
Critical Care S, European Society of Clinical M, Infectious D,
European Society of Intensive Care M, European Respiratory S,
International Sepsis F, Japanese Association for Acute M, Japanese
Society of Intensive Care M, Society of Critical Care M, Society of
Hospital M, Surgical Infection S, World Federation of Societies of I,
Critical Care M. Surviving Sepsis Campaign: international guidelines
for management of severe sepsis and septic shock: 2008. Crit Care
Med 2008; 36: 296-327.

70
95. Gandhi GY, Nuttall GA, Abel MD, Mullany CJ, Schaff HV, O'Brien
PC, Johnson MG, Williams AR, Cutshall SM, Mundy LM, Rizza RA,
McMahon MM. Intensive intraoperative insulin therapy versus
conventional glucose management during cardiac surgery: a
randomized trial. Ann Intern Med 2007; 146: 233-43.
96. Investigators N-SS, Finfer S, Chittock DR, Su SY, Blair D, Foster D,
Dhingra V, Bellomo R, Cook D, Dodek P, Henderson WR, Hebert PC,
Heritier S, Heyland DK, McArthur C, McDonald E, Mitchell I,
Myburgh JA, Norton R, Potter J, Robinson BG, Ronco JJ. Intensive
versus conventional glucose control in critically ill patients. N Engl J
Med 2009; 360: 1283-97.
97. Egi M, Bellomo R, Stachowski E, French CJ, Hart GK, Taori G,
Hegarty C, Bailey M. Hypoglycemia and outcome in critically ill
patients. Mayo Clin Proc 2010; 85: 217-24.
98. Wiener RS, Wiener DC, Larson RJ. Benefits and risks of tight glucose
control in critically ill adults: a meta-analysis. JAMA 2008; 300: 933-
44.
99. Jeejeebhoy KN. Enteral nutrition versus parenteral nutrition--the risks
and benefits. Nat Clin Pract Gastroenterol Hepatol 2007; 4: 260-5.
100. Egi M, Bellomo R, Stachowski E, French CJ, Hart G. Variability of
blood glucose concentration and short-term mortality in critically ill
patients. Anesthesiology 2006; 105: 244-52.
101. Nolan JP, Neumar RW, Adrie C, Aibiki M, Berg RA, Bbttiger BW,
Callaway C, Clark RS, Geocadin RG, Jauch EC, Kern KB, Laurent I,
Longstreth WT, Merchant RM, Morley P, Morrison LJ, Nadkarni V,
Peberdy MA, Rivers EP, Rodriguez-Nunez A, Sellke FW, Spaulding
C, Sunde K, Vanden Hoek T, International Liaison Committee on R,
Emergency Cardiovascular Care Committee AHA, Council on
Cardiovascular S, Anesthesia, Council on Cardiopulmonary P, Critical
C, Council on Clinical C, Council on S. Post-cardiac arrest syndrome:
epidemiology, pathophysiology, treatment, and prognostication: a
scientific statement from the International Liaison Committee on
Resuscitation; the American Heart Association Emergency
Cardiovascular Care Committee; the Council on Cardiovascular
Surgery and Anesthesia; the Council on Cardiopulmonary,
Perioperative, and Critical Care; the Council on Clinical Cardiology;
the Council on Stroke (Part II). Int Emerg Nurs 2010; 18: 8-28.
102. Oliver CN, Starke-Reed PE, Stadtman ER, Liu GJ, Carney JM, Floyd
RA. Oxidative damage to brain proteins, loss of glutamine synthetase
activity, and production of free radicals during ischemia/reperfusion-
induced injury to gerbil brain. Proc Natl Acad Sci U S A 1990; 87:
5144-7.
103. Clough-Helfman C, Phillis JW. The free radical trapping agent N-tert.-
butyl-alpha-phenylnitrone (PBN) attenuates cerebral ischaemic injury
in gerbils. Free Radic Res Commun 1991; 15: 177-86.

71
104. Cao X, Phillis JW. alpha-Phenyl-tert-butyl-nitrone reduces cortical
infarct and edema in rats subjected to focal ischemia. Brain Res 1994;
644: 267-72.
105. Zhao Q, Pahlmark K, Smith ML, Siesjo BK. Delayed treatment with
the spin trap alpha-phenyl-N-tert-butyl nitrone (PBN) reduces infarct
size following transient middle cerebral artery occlusion in rats. Acta
Physiol Scand 1994; 152: 349-50.
106. Kuroda S, Tsuchidate R, Smith ML, Maples KR, Siesjo BK.
Neuroprotective effects of a novel nitrone, NXY-059, after transient
focal cerebral ischemia in the rat. J Cereb Blood Flow Metab 1999;
19: 778-87.
107. Sydserff SG, Borelli AR, Green AR, Cross AJ. Effect of NXY-059 on
infarct volume after transient or permanent middle cerebral artery
occlusion in the rat; studies on dose, plasma concentration and
therapeutic time window. Br J Pharmacol 2002; 135: 103-12.
108. Lees KR, Sharma AK, Barer D, Ford GA, Kostulas V, Cheng YF,
Odergren T. Tolerability and pharmacokinetics of the nitrone NXY-
059 in patients with acute stroke. Stroke 2001; 32: 675-80.
109. Marshall JW, Duffin KJ, Green AR, Ridley RM. NXY-059, a free
radical--trapping agent, substantially lessens the functional disability
resulting from cerebral ischemia in a primate species. Stroke 2001; 32:
190-8.
110. Marshall JW, Cummings RM, Bowes LJ, Ridley RM, Green AR.
Functional and histological evidence for the protective effect of NXY-
059 in a primate model of stroke when given 4 hours after occlusion.
Stroke 2003; 34: 2228-33.
111. Schulz JB, Matthews RT, Jenkins BG, Brar P, Beal MF. Improved
therapeutic window for treatment of histotoxic hypoxia with a free
radical spin trap. J Cereb Blood Flow Metab 1995; 15: 948-52.
112. Liu XL, Wiklund L, Nozari A, Rubertsson S, Basu S. Differences in
cerebral reperfusion and oxidative injury after cardiac arrest in pigs.
Acta Anaesthesiol Scand 2003; 47: 958-67.
113. Wiklund L, Sharma HS, Basu S. Circulatory arrest as a model for
studies of global ischemic injury and neuroprotection. Ann N Y Acad
Sci 2005; 1053: 205-19.
114. Yang Y, Li Q, Shuaib A. Neuroprotection by 2-h postischemia
administration of two free radical scavengers, alpha-phenyl-n-tert-
butyl-nitrone (PBN) and N-tert-butyl-(2-sulfophenyl)-nitrone (S-
PBN), in rats subjected to focal embolic cerebral ischemia. Exp
Neurol 2000; 163: 39-45.
115. Marklund N, Sihver S, Langstrom B, Bergstrom M, Hillered L. Effect
of traumatic brain injury and nitrone radical scavengers on relative
changes in regional cerebral blood flow and glucose uptake in rats. J
Neurotrauma 2002; 19: 1139-53.

72
116. Marklund N, Lewander T, Clausen F, Hillered L. Effects of the nitrone
radical scavengers PBN and S-PBN on in vivo trapping of reactive
oxygen species after traumatic brain injury in rats. J Cereb Blood
Flow Metab 2001; 21: 1259-67.
117. Clausen F, Lorant T, Lewen A, Hillered L. T lymphocyte trafficking: a
novel target for neuroprotection in traumatic brain injury. J
Neurotrauma 2007; 24: 1295-307.
118. Traystman RJ. Animal models of focal and global cerebral ischemia.
ILAR J 2003; 44: 85-95.
119. Graham SM, McCullough LD, Murphy SJ. Animal models of
ischemic stroke: balancing experimental aims and animal care. Comp
Med 2004; 54: 486-96.
120. Garcia JH. Experimental ischemic stroke: a review. Stroke 1984; 15:
5-14.
121. Erdo F, Hossmann KA. [Animal models of cerebral ischemia--testing
therapeutic strategies in vivo]. Ideggyogy Sz 2007; 60: 356-69.
122. Coyle P. Arterial patterns of the rat rhinencephalon and related
structures. Exp Neurol 1975; 49: 671-90.
123. Rieke GK, Bowers DE, Jr., Penn P. Vascular supply pattern to rat
caudatoputamen and globus pallidus: scanning electronmicroscopic
study of vascular endocasts of stroke-prone vessels. Stroke 1981; 12:
840-7.
124. Mitchell WK, Himwich WA. Hemodynamic studies in the circle of
Willis in the rat. Experientia 1966; 22: 673-4.
125. Robinson RG, Shoemaker WJ, Schlumpf M, Valk T, Bloom FE. Effect
of experimental cerebral infarction in rat brain on catecholamines and
behaviour. Nature 1975; 255: 332-4.
126. Tamura A, Graham DI, McCulloch J, Teasdale GM. Focal cerebral
ischaemia in the rat: 1. Description of technique and early
neuropathological consequences following middle cerebral artery
occlusion. J Cereb Blood Flow Metab 1981; 1: 53-60.
127. Morikawa E, Ginsberg MD, Dietrich WD, Duncan RC, Kraydieh S,
Globus MY, Busto R. The significance of brain temperature in focal
cerebral ischemia: histopathological consequences of middle cerebral
artery occlusion in the rat. J Cereb Blood Flow Metab 1992; 12: 380-
9.
128. Shigeno T, Teasdale GM, McCulloch J, Graham DI. Recirculation
model following MCA occlusion in rats. Cerebral blood flow,
cerebrovascular permeability, and brain edema. J Neurosurg 1985; 63:
272-7.
129. Koizumi J YY, Nakazawa T, Ooneda G. "Experimental studies of
ischemic brain edema,I: a new experimental model of cerebral
emolism in rats in which recirculation can be introduced in the
ischemic area." Jpn J Stroke 1986; 8: 1-8.

73
130. Longa EZ, Weinstein PR, Carlson S, Cummins R. Reversible middle
cerebral artery occlusion without craniectomy in rats. Stroke 1989; 20:
84-91.
131. Koehler RC, Chandra N, Guerci AD, Tsitlik J, Traystman RJ, Rogers
MC, Weisfeldt ML. Augmentation of cerebral perfusion by
simultaneous chest compression and lung inflation with abdominal
binding after cardiac arrest in dogs. Circulation 1983; 67: 266-75.
132. Berkowitz ID, Gervais H, Schleien CL, Koehler RC, Dean JM,
Traystman RJ. Epinephrine dosage effects on cerebral and myocardial
blood flow in an infant swine model of cardiopulmonary resuscitation.
Anesthesiology 1991; 75: 1041-50.
133. Hossmann KA. The hypoxic brain. Insights from ischemia research.
Adv Exp Med Biol 1999; 474: 155-69.
134. Rubertsson S, Karlsten R. Increased cortical cerebral blood flow with
LUCAS; a new device for mechanical chest compressions compared
to standard external compressions during experimental
cardiopulmonary resuscitation. Resuscitation 2005; 65: 357-63.
135. Williams AJ, Lu XM, Slusher B, Tortella FC. Electroencephalogram
analysis and neuroprotective profile of the N-acetylated-alpha-linked
acidic dipeptidase inhibitor, GPI5232, in normal and brain-injured
rats. J Pharmacol Exp Ther 2001; 299: 48-57.
136. Roos MW, Ericsson A, Berg M, Sperber GO, Sjoquist M, Meyerson
BJ. Functional evaluation of cerebral microembolization in the rat.
Brain Res 2003; 961: 15-21.
137. Bederson JB, Pitts LH, Tsuji M, Nishimura MC, Davis RL,
Bartkowski H. Rat middle cerebral artery occlusion: evaluation of the
model and development of a neurologic examination. Stroke 1986;
17: 472-6.
138. Nachlas MM, Tsou KC, De Souza E, Cheng CS, Seligman AM.
Cytochemical demonstration of succinic dehydrogenase by the use of
a new p-nitrophenyl substituted ditetrazole. J Histochem Cytochem
1957; 5: 420-36.
139. Park CK, Mendelow AD, Graham DI, McCulloch J, Teasdale GM.
Correlation of triphenyltetrazolium chloride perfusion staining with
conventional neurohistology in the detection of early brain ischaemia.
Neuropathol Appl Neurobiol 1988; 14: 289-98.
140. Jobsis FF. Noninvasive, infrared monitoring of cerebral and
myocardial oxygen sufficiency and circulatory parameters. Science
1977; 198: 1264-7.
141. Ferrari M, Quaresima V. A brief review on the history of human
functional near-infrared spectroscopy (fNIRS) development and fields
of application. Neuroimage 2012; 63: 921-35.
142. Ito H, Kanno I, Fukuda H. Human cerebral circulation: positron
emission tomography studies. Ann Nucl Med 2005; 19: 65-74.

74
143. Sakatani K, Katayama Y, Yamamoto T, Suzuki S. Changes in cerebral
blood oxygenation of the frontal lobe induced by direct electrical
stimulation of thalamus and globus pallidus: a near infrared
spectroscopy study. J Neurol Neurosurg Psychiatry 1999; 67: 769-73.
144. Bolognese P, Miller JI, Heger IM, Milhorat TH. Laser-Doppler
flowmetry in neurosurgery. J Neurosurg Anesthesiol 1993; 5: 151-8.
145. Basu S. F2-isoprostanes in human health and diseases: from molecular
mechanisms to clinical implications. Antioxid Redox Signal 2008; 10:
1405-34.
146. Basu S. Metabolism of 8-iso-prostaglandin F2alpha. FEBS Lett 1998;
428: 32-6.
147. Lipcsey M, Larsson A, Eriksson MB, Sjolin J. Effect of the
administration rate on the biological responses to a fixed dose of
endotoxin in the anesthetized pig. Shock 2008; 29: 173-80.
148. Basu S. Radioimmunoassay of 8-iso-prostaglandin F2alpha: an index
for oxidative injury via free radical catalysed lipid peroxidation.
Prostaglandins Leukot Essent Fatty Acids 1998; 58: 319-25.
149. Basu S. Radioimmunoassay of 15-keto-13,14-dihydro-prostaglandin
F2alpha: an index for inflammation via cyclooxygenase catalysed
lipid peroxidation. Prostaglandins Leukot Essent Fatty Acids 1998;
58: 347-52.
150. Moore BW. A soluble protein characteristic of the nervous system.
Biochem Biophys Res Commun 1965; 19: 739-44.
151. Jonsson H, Johnsson P, Hoglund P, Alling C, Blomquist S.
Elimination of S100B and renal function after cardiac surgery. J
Cardiothorac Vasc Anesth 2000; 14: 698-701.
152. Wiesmann M, Missler U, Gottmann D, Gehring S. Plasma S-100b
protein concentration in healthy adults is age- and sex-independent.
Clin Chem 1998; 44: 1056-8.
153. Haglid KG, Yang Q, Hamberger A, Bergman S, Widerberg A,
Danielsen N. S-100beta stimulates neurite outgrowth in the rat sciatic
nerve grafted with acellular muscle transplants. Brain Res 1997; 753:
196-201.
154. Selinfreund RH, Barger SW, Pledger WJ, Van Eldik LJ. Neurotrophic
protein S100 beta stimulates glial cell proliferation. Proc Natl Acad
Sci U S A 1991; 88: 3554-8.
155. Goncalves DS, Lenz G, Karl J, Goncalves CA, Rodnight R.
Extracellular S100B protein modulates ERK in astrocyte cultures.
Neuroreport 2000; 11: 807-9.
156. Ahlemeyer B, Beier H, Semkova I, Schaper C, Krieglstein J. S-
100beta protects cultured neurons against glutamate- and
staurosporine-induced damage and is involved in the antiapoptotic
action of the 5 HT(1A)-receptor agonist, Bay x 3702. Brain Res 2000;
858: 121-8.

75
157. Barger SW, Van Eldik LJ, Mattson MP. S100 beta protects
hippocampal neurons from damage induced by glucose deprivation.
Brain Res 1995; 677: 167-70.
158. Hu J, Van Eldik LJ. Glial-derived proteins activate cultured astrocytes
and enhance beta amyloid-induced glial activation. Brain Res 1999;
842: 46-54.
159. Li Y, Barger SW, Liu L, Mrak RE, Griffin WS. S100beta induction of
the proinflammatory cytokine interleukin-6 in neurons. J Neurochem
2000; 74: 143-50.
160. Van Eldik LJ, Wainwright MS. The Janus face of glial-derived S100B:
beneficial and detrimental functions in the brain. Restor Neurol
Neurosci 2003; 21: 97-108.
161. Mariggio MA, Fulle S, Calissano P, Nicoletti I, Fano G. The brain
protein S-100ab induces apoptosis in PC12 cells. Neuroscience 1994;
60: 29-35.
162. Jauch EC, Lindsell C, Broderick J, Fagan SC, Tilley BC, Levine SR,
Group Nr-PSS. Association of serial biochemical markers with acute
ischemic stroke: the National Institute of Neurological Disorders and
Stroke recombinant tissue plasminogen activator Stroke Study. Stroke
2006; 37: 2508-13.
163. Shinozaki K, Oda S, Sadahiro T, Nakamura M, Abe R, Nakada TA,
Nomura F, Nakanishi K, Kitamura N, Hirasawa H. Serum S-100B is
superior to neuron-specific enolase as an early prognostic biomarker
for neurological outcome following cardiopulmonary resuscitation.
Resuscitation 2009; 80: 870-5.
164. Scheller J, Chalaris A, Schmidt-Arras D, Rose-John S. The pro- and
anti-inflammatory properties of the cytokine interleukin-6. Biochim
Biophys Acta 2011; 1813: 878-88.
165. Watters O, Pickering M, O'Connor JJ. Preconditioning effects of
tumor necrosis factor-alpha and glutamate on calcium dynamics in rat
organotypic hippocampal cultures. J Neuroimmunol 2011; 234: 27-
39.
166. Dalma-Weiszhausz DD, Warrington J, Tanimoto EY, Miyada CG. The
affymetrix GeneChip platform: an overview. Methods Enzymol 2006;
410: 3-28.
167. Zausinger S, Hungerhuber E, Baethmann A, Reulen H, Schmid-
Elsaesser R. Neurological impairment in rats after transient middle
cerebral artery occlusion: a comparative study under various treatment
paradigms. Brain Res 2000; 863: 94-105.
168. Weir CJ, Murray GD, Dyker AG, Lees KR. Is hyperglycaemia an
independent predictor of poor outcome after acute stroke? Results of a
long-term follow up study. BMJ 1997; 314: 1303-6.

76
169. Prado R, Ginsberg MD, Dietrich WD, Watson BD, Busto R.
Hyperglycemia increases infarct size in collaterally perfused but not
end-arterial vascular territories. J Cereb Blood Flow Metab 1988; 8:
186-92.
170. Kraft SA, Larson CP, Jr., Shuer LM, Steinberg GK, Benson GV, Pearl
RG. Effect of hyperglycemia on neuronal changes in a rabbit model of
focal cerebral ischemia. Stroke 1990; 21: 447-50.
171. Alvarez-Sabin J, Molina CA, Ribo M, Arenillas JF, Montaner J,
Huertas R, Santamarina E, Rubiera M. Impact of admission
hyperglycemia on stroke outcome after thrombolysis: risk
stratification in relation to time to reperfusion. Stroke 2004; 35: 2493-
8.
172. Branston NM, Symon L, Crockard HA, Pasztor E. Relationship
between the cortical evoked potential and local cortical blood flow
following acute middle cerebral artery occlusion in the baboon. Exp
Neurol 1974; 45: 195-208.
173. Pulsinelli WA, Levy DE, Duffy TE. Regional cerebral blood flow and
glucose metabolism following transient forebrain ischemia. Ann
Neurol 1982; 11: 499-502.
174. Kozuka M, Smith ML, Siesjo BK. Preischemic hyperglycemia
enhances postischemic depression of cerebral metabolic rate. J Cereb
Blood Flow Metab 1989; 9: 478-90.
175. Ginsberg MD, Zhao W, Alonso OF, Loor-Estades JY, Dietrich WD,
Busto R. Uncoupling of local cerebral glucose metabolism and blood
flow after acute fluid-percussion injury in rats. Am J Physiol 1997;
272: H2859-68.
176. Lewen A, Hillered L. Involvement of reactive oxygen species in
membrane phospholipid breakdown and energy perturbation after
traumatic brain injury in the rat. J Neurotrauma 1998; 15: 521-30.
177. Frykholm P, Hillered L, Langstrom B, Persson L, Valtysson J, Enblad
P. Relationship between cerebral blood flow and oxygen metabolism,
and extracellular glucose and lactate concentrations during middle
cerebral artery occlusion and reperfusion: a microdialysis and positron
emission tomography study in nonhuman primates. J Neurosurg 2005;
102: 1076-84.
178. Lees KR, Zivin JA, Ashwood T, Davalos A, Davis SM, Diener HC,
Grotta J, Lyden P, Shuaib A, Hardemark HG, Wasiewski WW, Stroke-
Acute Ischemic NXYTTI. NXY-059 for acute ischemic stroke. N
Engl J Med 2006; 354: 588-600.
179. Maples KR, Ma F, Zhang YK. Comparison of the radical trapping
ability of PBN, S-PPBN and NXY-059. Free Radic Res 2001; 34:
417-26.

77
180. Miyajima T, Kotake Y. Spin trapping agent, phenyl N-tert-butyl
nitrone, inhibits induction of nitric oxide synthase in endotoxin-
induced shock in mice. Biochem Biophys Res Commun 1995; 215:
114-21.
181. Floyd RA, Hensley K, Jaffery F, Maidt L, Robinson K, Pye Q,
Stewart C. Increased oxidative stress brought on by pro-inflammatory
cytokines in neurodegenerative processes and the protective role of
nitrone-based free radical traps. Life Sci 1999; 65: 1893-9.
182. Joseph JA, Cao G, Cutler RC. In vivo or in vitro administration of the
nitrone spin-trapping compound, n-tert-butyl-alpha-phenylnitrone,
(PBN) reduces age-related deficits in striatal muscarinic receptor
sensitivity. Brain Res 1995; 671: 73-7.
183. Anderson DE, Yuan XJ, Tseng CM, Rubin LJ, Rosen GM, Tod ML.
Nitrone spin-traps block calcium channels and induce pulmonary
artery relaxation independent of free radicals. Biochem Biophys Res
Commun 1993; 193: 878-85.
184. Kuroda S, Katsura K, Hillered L, Bates TE, Siesjo BK. Delayed
treatment with alpha-phenyl-N-tert-butyl nitrone (PBN) attenuates
secondary mitochondrial dysfunction after transient focal cerebral
ischemia in the rat. Neurobiol Dis 1996; 3: 149-57.
185. Folbergrova J, Zhao Q, Katsura K, Siesjo BK. N-tert-butyl-alpha-
phenylnitrone improves recovery of brain energy state in rats
following transient focal ischemia. Proc Natl Acad Sci U S A 1995;
92: 5057-61.
186. Chen GM, Bray TM, Janzen EG, McCay PB. Excretion, metabolism
and tissue distribution of a spin trapping agent, alpha-phenyl-N-tert-
butyl-nitrone (PBN) in rats. Free Radic Res Commun 1990; 9: 317-
23.
187. Preston E, Foster DO. Evidence for pore-like opening of the blood-
brain barrier following forebrain ischemia in rats. Brain Res 1997;
761: 4-10.
188. Koziol JA, Feng AC. On the analysis and interpretation of outcome
measures in stroke clinical trials: lessons from the SAINT I study of
NXY-059 for acute ischemic stroke. Stroke 2006; 37: 2644-7.
189. Shuaib A, Lees KR, Lyden P, Grotta J, Davalos A, Davis SM, Diener
HC, Ashwood T, Wasiewski WW, Emeribe U, Investigators SIT.
NXY-059 for the treatment of acute ischemic stroke. N Engl J Med
2007; 357: 562-71.
190. Macleod MR, van der Worp HB, Sena ES, Howells DW, Dirnagl U,
Donnan GA. Evidence for the efficacy of NXY-059 in experimental
focal cerebral ischaemia is confounded by study quality. Stroke 2008;
39: 2824-9.

78
191. Bath PM, Gray LJ, Bath AJ, Buchan A, Miyata T, Green AR,
Investigators NXYEM-aiIAwS. Effects of NXY-059 in experimental
stroke: an individual animal meta-analysis. Br J Pharmacol 2009; 157:
1157-71.
192. Clausen F. Delayed cell death after traumatic brain injury:role of
reactive oxygen species. . In: Comprehensive Summaries of Uppsala
Dissertations from the Faculty of Medicine. Uppsala: Uppsala
University, Faculty of Medicine; 2004;
193. O'Collins VE, Macleod MR, Donnan GA, Horky LL, van der Worp
BH, Howells DW. 1,026 experimental treatments in acute stroke. Ann
Neurol 2006; 59: 467-77.
194. Chen H, Song YS, Chan PH. Inhibition of NADPH oxidase is
neuroprotective after ischemia-reperfusion. J Cereb Blood Flow
Metab 2009; 29: 1262-72.
195. van der Worp HB, Sena ES, Donnan GA, Howells DW, Macleod MR.
Hypothermia in animal models of acute ischaemic stroke: a systematic
review and meta-analysis. Brain 2007; 130: 3063-74.
196. Yenari M, Kitagawa K, Lyden P, Perez-Pinzon M. Metabolic
downregulation: a key to successful neuroprotection? Stroke 2008; 39:
2910-7.
197. Rivers EP, Rady MY, Martin GB, Fenn NM, Smithline HA, Alexander
ME, Nowak RM. Venous hyperoxia after cardiac arrest.
Characterization of a defect in systemic oxygen utilization. Chest
1992; 102: 1787-93.
198. Ridder DA, Schwaninger M. NF-kappaB signaling in cerebral
ischemia. Neuroscience 2009; 158: 995-1006.
199. Minami M, Katayama T, Satoh M. Brain cytokines and chemokines:
roles in ischemic injury and pain. J Pharmacol Sci 2006; 100: 461-70.
200. DeGraba TJ. The role of inflammation after acute stroke: utility of
pursuing anti-adhesion molecule therapy. Neurology 1998; 51: S62-8.
201. Ishikawa M, Cooper D, Arumugam TV, Zhang JH, Nanda A, Granger
DN. Platelet-leukocyte-endothelial cell interactions after middle
cerebral artery occlusion and reperfusion. J Cereb Blood Flow Metab
2004; 24: 907-15.
202. Rodriguez-Yanez M, Castillo J. Role of inflammatory markers in
brain ischemia. Curr Opin Neurol 2008; 21: 353-7.
203. Fischer M, Bottiger BW, Popov-Cenic S, Hossmann KA.
Thrombolysis using plasminogen activator and heparin reduces
cerebral no-reflow after resuscitation from cardiac arrest: an
experimental study in the cat. Intensive Care Med 1996; 22: 1214-23.
204. Mortberg E, Cumming P, Wiklund L, Rubertsson S. Cerebral
metabolic rate of oxygen (CMRO2) in pig brain determined by PET
after resuscitation from cardiac arrest. Resuscitation 2009; 80: 701-6.

79
205. Gedeborg R, Silander HC, Rubertsson S, Wiklund L. Cerebral
ischaemia in experimental cardiopulmonary resuscitation--comparison
of epinephrine and aortic occlusion. Resuscitation 2001; 50: 319-29.
206. Kirino T, Sano K. Selective vulnerability in the gerbil hippocampus
following transient ischemia. Acta Neuropathol 1984; 62: 201-8.
207. Basu S, Nozari A, Liu XL, Rubertsson S, Wiklund L. Development of
a novel biomarker of free radical damage in reperfusion injury after
cardiac arrest. FEBS Lett 2000; 470: 1-6.
208. Basu S. Isoprostanes: novel bioactive products of lipid peroxidation.
Free Radic Res 2004; 38: 105-22.
209. Liu XL, Nozari A, Basu S, Ronquist G, Rubertsson S, Wiklund L.
Neurological outcome after experimental cardiopulmonary
resuscitation: a result of delayed and potentially treatable neuronal
injury? Acta Anaesthesiol Scand 2002; 46: 537-46.
210. Martin A, Rojas S, Chamorro A, Falcon C, Bargallo N, Planas AM.
Why does acute hyperglycemia worsen the outcome of transient focal
cerebral ischemia? Role of corticosteroids, inflammation, and protein
O-glycosylation. Stroke 2006; 37: 1288-95.
211. Cheung RT, Pei Z, Feng ZH, Zou LY. Cyclooxygenase-1 gene
knockout does not alter middle cerebral artery occlusion in a mouse
stroke model. Neurosci Lett 2002; 330: 57-60.
212. Devaraj S, Venugopal SK, Singh U, Jialal I. Hyperglycemia induces
monocytic release of interleukin-6 via induction of protein kinase c-
{alpha} and -{beta}. Diabetes 2005; 54: 85-91.
213. Ste-Marie L, Hazell AS, Bemeur C, Butterworth R, Montgomery J.
Immunohistochemical detection of inducible nitric oxide synthase,
nitrotyrosine and manganese superoxide dismutase following
hyperglycemic focal cerebral ischemia. Brain Res 2001; 918: 10-9.
214. Clemens JA. Cerebral ischemia: gene activation, neuronal injury, and
the protective role of antioxidants. Free Radic Biol Med 2000; 28:
1526-31.
215. Wells L, Vosseller K, Hart GW. Glycosylation of nucleocytoplasmic
proteins: signal transduction and O-GlcNAc. Science 2001; 291:
2376-8.
216. Federici M, Menghini R, Mauriello A, Hribal ML, Ferrelli F, Lauro D,
Sbraccia P, Spagnoli LG, Sesti G, Lauro R. Insulin-dependent
activation of endothelial nitric oxide synthase is impaired by O-linked
glycosylation modification of signaling proteins in human coronary
endothelial cells. Circulation 2002; 106: 466-72.
217. Dandona P, Chaudhuri A, Ghanim H, Mohanty P. Proinflammatory
effects of glucose and anti-inflammatory effect of insulin: relevance to
cardiovascular disease. Am J Cardiol 2007; 99: 15B-26B.
218. Mohanty P, Hamouda W, Garg R, Aljada A, Ghanim H, Dandona P.
Glucose challenge stimulates reactive oxygen species (ROS)
generation by leucocytes. J Clin Endocrinol Metab 2000; 85: 2970-3.

80
219. Aljada A, Ghanim H, Mohanty P, Syed T, Bandyopadhyay A,
Dandona P. Glucose intake induces an increase in activator protein 1
and early growth response 1 binding activities, in the expression of
tissue factor and matrix metalloproteinase in mononuclear cells, and
in plasma tissue factor and matrix metalloproteinase concentrations.
Am J Clin Nutr 2004; 80: 51-7.
220. Esposito K, Nappo F, Marfella R, Giugliano G, Giugliano F, Ciotola
M, Quagliaro L, Ceriello A, Giugliano D. Inflammatory cytokine
concentrations are acutely increased by hyperglycemia in humans:
role of oxidative stress. Circulation 2002; 106: 2067-72.
221. Auer RN. Insulin, blood glucose levels, and ischemic brain damage.
Neurology 1998; 51: S39-43.
222. Zeng G, Quon MJ. Insulin-stimulated production of nitric oxide is
inhibited by wortmannin. Direct measurement in vascular endothelial
cells. J Clin Invest 1996; 98: 894-8.
223. Aljada A, Dandona P. Effect of insulin on human aortic endothelial
nitric oxide synthase. Metabolism 2000; 49: 147-50.
224. Aljada A, Ghanim H, Saadeh R, Dandona P. Insulin inhibits
NFkappaB and MCP-1 expression in human aortic endothelial cells. J
Clin Endocrinol Metab 2001; 86: 450-3.
225. Dandona P, Mohanty P, Chaudhuri A, Garg R, Aljada A. Insulin
infusion in acute illness. J Clin Invest 2005; 115: 2069-72.
226. Trovati M, Anfossi G, Massucco P, Mattiello L, Costamagna C,
Piretto V, Mularoni E, Cavalot F, Bosia A, Ghigo D. Insulin stimulates
nitric oxide synthesis in human platelets and, through nitric oxide,
increases platelet concentrations of both guanosine-3', 5'-cyclic
monophosphate and adenosine-3', 5'-cyclic monophosphate. Diabetes
1997; 46: 742-9.
227. Voll CL, Auer RN. Insulin attenuates ischemic brain damage
independent of its hypoglycemic effect. J Cereb Blood Flow Metab
1991; 11: 1006-14.
228. Wass CT, Scheithauer BW, Bronk JT, Wilson RM, Lanier WL. Insulin
treatment of corticosteroid-associated hyperglycemia and its effect on
outcome after forebrain ischemia in rats. Anesthesiology 1996; 84:
644-51.
229. Hamilton MG, Tranmer BI, Auer RN. Insulin reduction of cerebral
infarction due to transient focal ischemia. J Neurosurg 1995; 82: 262-
8.
230. Nolan JP, Neumar RW, Adrie C, Aibiki M, Berg RA, Bbttiger BW,
Callaway C, Clark RS, Geocadin RG, Jauch EC, Kern KB, Laurent I,
Longstreth WT, Merchant RM, Morley P, Morrison LJ, Nadkarni V,
Peberdy MA, Rivers EP, Rodriguez-Nunez A, Sellke FW, Spaulding
C, Sunde K, Hoek TV. Post-cardiac arrest syndrome: Epidemiology,
pathophysiology, treatment, and prognostication: A scientific
statement from the International Liaison Committee on Resuscitation;

81
the American Heart Association Emergency Cardiovascular Care
Committee; the Council on Cardiovascular Surgery and Anesthesia;
the Council on Cardiopulmonary, Perioperative, and Critical Care; the
Council on Clinical Cardiology; the Council on Stroke (Part 1). Int
Emerg Nurs 2009; 17: 203-25.
231. Sundgreen C, Larsen FS, Herzog TM, Knudsen GM, Boesgaard S,
Aldershvile J. Autoregulation of cerebral blood flow in patients
resuscitated from cardiac arrest. Stroke 2001; 32: 128-32.
232. Gedeborg R, Rubertsson S, Wiklund L. Improved haemodynamics
and restoration of spontaneous circulation with constant aortic
occlusion during experimental cardiopulmonary resuscitation.
Resuscitation 1999; 40: 171-80.
233. Fischer M, Dahmen A, Standop J, Hagendorff A, Hoeft A, Krep H.
Effects of hypertonic saline on myocardial blood flow in a porcine
model of prolonged cardiac arrest. Resuscitation 2002; 54: 269-80.
234. Li F, Omae T, Fisher M. Spontaneous hyperthermia and its
mechanism in the intraluminal suture middle cerebral artery occlusion
model of rats. Stroke 1999; 30: 2464-70; discussion 70-1.
235. Persson L, Hardemark HG, Gustafsson J, Rundstrom G, Mendel-
Hartvig I, Esscher T, Pahlman S. S-100 protein and neuron-specific
enolase in cerebrospinal fluid and serum: markers of cell damage in
human central nervous system. Stroke 1987; 18: 911-8.
236. Fassbender K, Schmidt R, Schreiner A, Fatar M, Muhlhauser F,
Daffertshofer M, Hennerici M. Leakage of brain-originated proteins in
peripheral blood: temporal profile and diagnostic value in early
ischemic stroke. J Neurol Sci 1997; 148: 101-5.
237. Buttner T, Weyers S, Postert T, Sprengelmeyer R, Kuhn W. S-100
protein: serum marker of focal brain damage after ischemic territorial
MCA infarction. Stroke 1997; 28: 1961-5.
238. Herrmann M, Vos P, Wunderlich MT, de Bruijn CH, Lamers KJ.
Release of glial tissue-specific proteins after acute stroke: A
comparative analysis of serum concentrations of protein S-100B and
glial fibrillary acidic protein. Stroke 2000; 31: 2670-7.
239. Dassan P, Keir G, Brown MM. Criteria for a clinically informative
serum biomarker in acute ischaemic stroke: a review of S100B.
Cerebrovasc Dis 2009; 27: 295-302.
240. Unden J, Bellner J, Eneroth M, Alling C, Ingebrigtsen T, Romner B.
Raised serum S100B levels after acute bone fractures without cerebral
injury. J Trauma 2005; 58: 59-61.
241. Grubb NR, Simpson C, Sherwood RA, Abraha HD, Cobbe SM,
O'Carroll RE, Deary I, Fox KA. Prediction of cognitive dysfunction
after resuscitation from out-of-hospital cardiac arrest using serum
neuron-specific enolase and protein S-100. Heart 2007; 93: 1268-73.

82
242. Missler U, Wiesmann M, Friedrich C, Kaps M. S-100 protein and
neuron-specific enolase concentrations in blood as indicators of
infarction volume and prognosis in acute ischemic stroke. Stroke
1997; 28: 1956-60.
243. Wunderlich MT, Wallesch CW, Goertler M. Release of
neurobiochemical markers of brain damage is related to the
neurovascular status on admission and the site of arterial occlusion in
acute ischemic stroke. J Neurol Sci 2004; 227: 49-53.
244. Martens P, Raabe A, Johnsson P. Serum S-100 and neuron-specific
enolase for prediction of regaining consciousness after global cerebral
ischemia. Stroke 1998; 29: 2363-6.
245. Alvarez-Sabin J, Molina CA, Montaner J, Arenillas JF, Huertas R,
Ribo M, Codina A, Quintana M. Effects of admission hyperglycemia
on stroke outcome in reperfused tissue plasminogen activator--treated
patients. Stroke 2003; 34: 1235-41.
246. Wang Q, Tang XN, Yenari MA. The inflammatory response in stroke.
J Neuroimmunol 2007; 184: 53-68.
247. Rallidis LS, Vikelis M, Panagiotakos DB, Rizos I, Zolindaki MG,
Kaliva K, Kremastinos DT. Inflammatory markers and in-hospital
mortality in acute ischaemic stroke. Atherosclerosis 2006; 189: 193-7.
248. Vila N, Castillo J, Davalos A, Chamorro A. Proinflammatory
cytokines and early neurological worsening in ischemic stroke. Stroke
2000; 31: 2325-9.
249. Emsley HC, Smith CJ, Georgiou RF, Vail A, Hopkins SJ, Rothwell
NJ, Tyrrell PJ, Acute Stroke I. A randomised phase II study of
interleukin-1 receptor antagonist in acute stroke patients. J Neurol
Neurosurg Psychiatry 2005; 76: 1366-72.
250. Sotgiu S, Zanda B, Marchetti B, Fois ML, Arru G, Pes GM, Salaris
FS, Arru A, Pirisi A, Rosati G. Inflammatory biomarkers in blood of
patients with acute brain ischemia. Eur J Neurol 2006; 13: 505-13.
251. Lambertsen KL, Biber K, Finsen B. Inflammatory cytokines in
experimental and human stroke. J Cereb Blood Flow Metab 2012; 32:
1677-98.
252. Chang L, Chen Y, Li J, Liu Z, Wang Z, Chen J, Cao W, Xu Y.
Cocaine-and amphetamine-regulated transcript modulates peripheral
immunity and protects against brain injury in experimental stroke.
Brain Behav Immun 2011; 25: 260-9.
253. Chapman KZ, Dale VQ, Denes A, Bennett G, Rothwell NJ, Allan SM,
McColl BW. A rapid and transient peripheral inflammatory response
precedes brain inflammation after experimental stroke. J Cereb Blood
Flow Metab 2009; 29: 1764-8.
254. Niemann JT, Rosborough JP, Youngquist S. Is the tumour necrosis
factor-alpha response following resuscitation gender dependent in the
swine model? Resuscitation 2008; 77: 258-63.

83
255. Niemann JT, Rosborough JP, Youngquist S, Shah AP, Lewis RJ, Phan
QT, Filler SG. Cardiac function and the proinflammatory cytokine
response after recovery from cardiac arrest in swine. J Interferon
Cytokine Res 2009; 29: 749-58.
256. Ito T, Saitoh D, Fukuzuka K, Kiyozumi T, Kawakami M, Sakamoto T,
Okada Y. Significance of elevated serum interleukin-8 in patients
resuscitated after cardiopulmonary arrest. Resuscitation 2001; 51: 47-
53.
257. Adrie C, Adib-Conquy M, Laurent I, Monchi M, Vinsonneau C,
Fitting C, Fraisse F, Dinh-Xuan AT, Carli P, Spaulding C, Dhainaut JF,
Cavaillon JM. Successful cardiopulmonary resuscitation after cardiac
arrest as a "sepsis-like" syndrome. Circulation 2002; 106: 562-8.
258. Casaer MP, Mesotten D, Hermans G, Wouters PJ, Schetz M,
Meyfroidt G, Van Cromphaut S, Ingels C, Meersseman P, Muller J,
Vlasselaers D, Debaveye Y, Desmet L, Dubois J, Van Assche A,
Vanderheyden S, Wilmer A, Van den Berghe G. Early versus late
parenteral nutrition in critically ill adults. N Engl J Med 2011; 365:
506-17.
259. Benjamini Y, Drai D, Elmer G, Kafkafi N, Golani I. Controlling the
false discovery rate in behavior genetics research. Behav Brain Res
2001; 125: 279-84.
260. Huang da W, Sherman BT, Lempicki RA. Systematic and integrative
analysis of large gene lists using DAVID bioinformatics resources.
Nat Protoc 2009; 4: 44-57.
261. Dwyer DS, Vannucci SJ, Simpson IA. Expression, regulation, and
functional role of glucose transporters (GLUTs) in brain. Int Rev
Neurobiol 2002; 51: 159-88.
262. Wright EM. Glucose transport families SLC5 and SLC50. Mol
Aspects Med 2013; 34: 183-96.
263. Yu AS, Hirayama BA, Timbol G, Liu J, Diez-Sampedro A, Kepe V,
Satyamurthy N, Huang SC, Wright EM, Barrio JR. Regional
distribution of SGLT activity in rat brain in vivo. Am J Physiol Cell
Physiol 2013; 304: C240-7.
264. Lin X, Ma L, Fitzgerald RL, Ostlund RE, Jr. Human sodium/inositol
cotransporter 2 (SMIT2) transports inositols but not glucose in L6
cells. Arch Biochem Biophys 2009; 481: 197-201.
265. Tsai LJ, Hsiao SH, Tsai LM, Lin CY, Tsai JJ, Liou DM, Lan JL. The
sodium-dependent glucose cotransporter SLC5A11 as an autoimmune
modifier gene in SLE. Tissue Antigens 2008; 71: 114-26.
266. Prins ML. Cerebral metabolic adaptation and ketone metabolism after
brain injury. J Cereb Blood Flow Metab 2008; 28: 1-16.
267. McIlwain H. Metabolic adaptation in the brain. Nature 1970; 226:
803-6.
268. Sass JO. Inborn errors of ketogenesis and ketone body utilization. J
Inherit Metab Dis 2012; 35: 23-8.

84
269. Zhang F, Vannucci SJ, Philp NJ, Simpson IA. Monocarboxylate
transporter expression in the spontaneous hypertensive rat: effect of
stroke. J Neurosci Res 2005; 79: 139-45.
270. Suzuki M, Suzuki M, Sato K, Dohi S, Sato T, Matsuura A, Hiraide A.
Effect of beta-hydroxybutyrate, a cerebral function improving agent,
on cerebral hypoxia, anoxia and ischemia in mice and rats. Jpn J
Pharmacol 2001; 87: 143-50.
271. Suzuki M, Suzuki M, Kitamura Y, Mori S, Sato K, Dohi S, Sato T,
Matsuura A, Hiraide A. Beta-hydroxybutyrate, a cerebral function
improving agent, protects rat brain against ischemic damage caused
by permanent and transient focal cerebral ischemia. Jpn J Pharmacol
2002; 89: 36-43.
272. Hagenbuch B, Stieger B. The SLCO (former SLC21) superfamily of
transporters. Mol Aspects Med 2013; 34: 396-412.
273. Gao B, Vavricka SR, Meier PJ, Stieger B. Differential cellular
expression of organic anion transporting peptides OATP1A2 and
OATP2B1 in the human retina and brain: implications for carrier-
mediated transport of neuropeptides and neurosteriods in the CNS.
Pflugers Arch 2014;
274. Maninger N, Wolkowitz OM, Reus VI, Epel ES, Mellon SH.
Neurobiological and neuropsychiatric effects of
dehydroepiandrosterone (DHEA) and DHEA sulfate (DHEAS). Front
Neuroendocrinol 2009; 30: 65-91.
275. Mohebiany AN, Harroch S, Bouyain S. New insights into the roles of
the contactin cell adhesion molecules in neural development. Adv
Neurobiol 2014; 8: 165-94.
276. Yoshihara Y, Kawasaki M, Tani A, Tamada A, Nagata S, Kagamiyama
H, Mori K. BIG-1: a new TAG-1/F3-related member of the
immunoglobulin superfamily with neurite outgrowth-promoting
activity. Neuron 1994; 13: 415-26.
277. Boal F, Stephens DJ. Specific functions of BIG1 and BIG2 in
endomembrane organization. PLoS One 2010; 5: e9898.
278. Chen J, Yu S, Fu Y, Li X. Synaptic proteins and receptors defects in
autism spectrum disorders. Frontiers in cellular neuroscience 2014; 8:
276.
279. Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and
function. Cell 2004; 116: 281-97.
280. Sempere LF, Freemantle S, Pitha-Rowe I, Moss E, Dmitrovsky E,
Ambros V. Expression profiling of mammalian microRNAs uncovers
a subset of brain-expressed microRNAs with possible roles in murine
and human neuronal differentiation. Genome Biol 2004; 5: R13.
281. Adlakha YK, Saini N. Brain microRNAs and insights into biological
functions and therapeutic potential of brain enriched miRNA-128.
Mol Cancer 2014; 13: 33.

85
282. Cogswell JP, Ward J, Taylor IA, Waters M, Shi Y, Cannon B, Kelnar
K, Kemppainen J, Brown D, Chen C, Prinjha RK, Richardson JC,
Saunders AM, Roses AD, Richards CA. Identification of miRNA
changes in Alzheimer's disease brain and CSF yields putative
biomarkers and insights into disease pathways. J Alzheimers Dis
2008; 14: 27-41.
283. Martin de la Vega C, Burda J, Nemethova M, Quevedo C, Alcazar A,
Martin ME, Danielisova V, Fando JL, Salinas M. Possible
mechanisms involved in the down-regulation of translation during
transient global ischaemia in the rat brain. Biochem J 2001; 357: 819-
26.
284. Yuan Y, Wang JY, Xu LY, Cai R, Chen Z, Luo BY. MicroRNA
expression changes in the hippocampi of rats subjected to global
ischemia. J Clin Neurosci 2010; 17: 774-8.
285. Jeyaseelan K, Lim KY, Armugam A. MicroRNA expression in the
blood and brain of rats subjected to transient focal ischemia by middle
cerebral artery occlusion. Stroke 2008; 39: 959-66.
286. Dharap A, Bowen K, Place R, Li LC, Vemuganti R. Transient focal
ischemia induces extensive temporal changes in rat cerebral
microRNAome. J Cereb Blood Flow Metab 2009; 29: 675-87.
287. Saugstad JA. MicroRNAs as effectors of brain function with roles in
ischemia and injury, neuroprotection, and neurodegeneration. J Cereb
Blood Flow Metab 2010; 30: 1564-76.
288. Xu H, Lu A, Sharp FR. Regional genome transcriptional response of
adult mouse brain to hypoxia. BMC Genomics 2011; 12: 499.
289. Lee ST, Chu K, Jung KH, Yoon HJ, Jeon D, Kang KM, Park KH, Bae
EK, Kim M, Lee SK, Roh JK. MicroRNAs induced during ischemic
preconditioning. Stroke 2010; 41: 1646-51.
290. Roussel BD, Mysiorek C, Rouhiainen A, Jullienne A, Parcq J,
Hommet Y, Culot M, Berezowski V, Cecchelli R, Rauvala H, Vivien
D, Ali C. HMGB-1 promotes fibrinolysis and reduces neurotoxicity
mediated by tissue plasminogen activator. J Cell Sci 2011; 124: 2070-
6.
291. Picchi A, Valente S, Gensini G. Therapeutic hypothermia in the
intensive cardiac care unit. J Cardiovasc Med (Hagerstown) 2014;
 

86
Acta Universitatis Upsaliensis
Digital Comprehensive Summaries of Uppsala Dissertations
from the Faculty of Medicine 1089
Editor: The Dean of the Faculty of Medicine

A doctoral dissertation from the Faculty of Medicine, Uppsala


University, is usually a summary of a number of papers. A few
copies of the complete dissertation are kept at major Swedish
research libraries, while the summary alone is distributed
internationally through the series Digital Comprehensive
Summaries of Uppsala Dissertations from the Faculty of
Medicine. (Prior to January, 2005, the series was published
under the title “Comprehensive Summaries of Uppsala
Dissertations from the Faculty of Medicine”.)

ACTA
UNIVERSITATIS
UPSALIENSIS
Distribution: publications.uu.se UPPSALA
urn:nbn:se:uu:diva-247763 2015

Вам также может понравиться