Download New insights into enhancing morphine analgesia Tuomas

Document related concepts

Stimulant wikipedia , lookup

Discovery and development of angiotensin receptor blockers wikipedia , lookup

Toxicodynamics wikipedia , lookup

Cannabinoid receptor antagonist wikipedia , lookup

NMDA receptor wikipedia , lookup

Theralizumab wikipedia , lookup

Pharmacogenomics wikipedia , lookup

NK1 receptor antagonist wikipedia , lookup

Neuropharmacology wikipedia , lookup

Psychopharmacology wikipedia , lookup

Bilastine wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Transcript
New insights into enhancing
morphine analgesia
From glia to pharmacokinetics
Tuomas Lilius
Institute of Biomedicine, Pharmacology
University of Helsinki
Finland
ACADEMIC DISSERTATION
To be publicly discussed with permission of the Faculty of Medicine,
University of Helsinki, in Haartman Institute, Lecture Hall 2,
on November 28th, at 12 noon.
Helsinki 2014
Supervisors
Professor Eija Kalso, M.D., Ph.D.
Pain Clinic, Department of Anaesthesiology and Intensive Care
Medicine
Helsinki University Central Hospital
Institute of Biomedicine, Pharmacology
University of Helsinki
Finland
Docent Pekka Rauhala, M.D., Ph.D.
Institute of Biomedicine, Pharmacology
University of Helsinki
Finland
Reviewers
Professor Raimo K. Tuominen, M.D., Ph.D.
Division of Pharmacology and Pharmacotherapy
Faculty of Pharmacy
University of Helsinki
Finland
Docent Juhana J. Idänpään-Heikkilä, M.D., Ph.D.
Country Medical Director
GSK Finland
Finland
Opponent
Professor Mika Scheinin, M.D., Ph.D.
Department of Pharmacology, Drug Development and Therapeutics
University of Turku
Finland
© Tuomas Lilius 2014
ISBN 978-951-51-0488-5 (paperback)
ISBN 978-951-51-0489-2 (PDF)
ISSN 2342-3161 (print)
ISSN 2342-317X (online)
http://ethesis.helsinki.fi
Hansaprint
Helsinki, Finland 2014
“No matter what else happens, fly the airplane.”
TABLE OF CONTENTS
List of abbreviations ........................................................................................ 6
Abstract ............................................................................................................ 9
List of original publications .......................................................................... 11
1. Introduction ............................................................................................... 12
2. Review of the literature ............................................................................. 14
2.1 Physiology of nociception and pain ................................................... 14
2.1.1 Primary nociceptors .............................................................................................14
2.1.2 Nociception at the spinal cord level ..............................................................15
2.1.3 Nociception and pain in supraspinal structures ........................................15
2.1.4 Supraspinal mechanisms of pain modulation ...........................................17
2.2 Opioids .................................................................................................. 18
2.2.1 Indications and prevalence of opioid therapy in Finland ......................18
2.2.2 Opioid receptors ...................................................................................................18
2.2.3 Mechanisms of opioid-induced analgesia ...................................................20
2.2.4 Morphine .................................................................................................................21
2.2.5 Adverse effects of morphine ............................................................................22
2.3 Modulation of opioid-induced antinociception ................................ 24
2.3.1 Opioid tolerance and opioid-induced hyperalgesia ................................24
2.3.2 Role of the glutamatergic system in opioid tolerance and opioidinduced hyperalgesia ....................................................................................................28
2.3.3 Glial cells and opioid antinociception ...........................................................31
2.3.4 Opioids and the alpha-2-adrenergic receptors .........................................34
2.3.5 Access of morphine and other opioids to the central nervous system
...............................................................................................................................................43
2.4 Drugs used in the study ....................................................................... 46
2.4.1 Atipamezole, an alpha-2-adrenoceptor antagonist.................................46
2.4.2 Ibudilast, a phosphodiesterase inhibitor .....................................................48
2.4.3 Spironolactone, a mineralocorticoid receptor antagonist ....................52
2.4.4 Ketamine, an N-methyl-D-aspartate receptor antagonist .....................55
3. Aims of the study ....................................................................................... 58
4. Materials and methods .............................................................................. 59
4.1 Ethical considerations ......................................................................... 59
4.2 Materials ............................................................................................... 59
4.2.1 Animals .....................................................................................................................59
4.2.2 Drugs .........................................................................................................................59
4.3 Methods ................................................................................................ 60
4.3.1 Intrathecal cannulation ......................................................................................60
4.3.2 Morphine tolerance treatment schemes .....................................................61
4.3.3 Pain measurement in acute nociception .....................................................61
4.3.4 Assessment of motor function.........................................................................62
4.3.5 Drug concentration measurements ..............................................................62
4.3.6 Experimental design of Studies I–IV ............................................................. 63
4.4 Statistical analysis................................................................................ 66
5. Results ......................................................................................................... 67
5.1 Morphine .............................................................................................. 67
5.1.1 Effects of morphine in different models of nociception........................ 67
5.1.2 Development of morphine tolerance in different models ................... 68
5.2 Ibudilast (study I) ................................................................................. 69
5.2.1 Antinociceptive effects of ibudilast............................................................... 69
5.2.2 Effects of ibudilast cotreatment on the restoration of morphine
antinociception ............................................................................................................... 70
5.2.3 Effects of ibudilast on motor coordination ................................................ 71
5.3 Atipamezole (study II) ......................................................................... 73
5.3.1 Effects of atipamezole coadministration in acute morphine
antinociception ............................................................................................................... 73
5.3.2 Effects of spinal atipamezole coadministration in morphine tolerance
............................................................................................................................................... 74
5.3.3 Effects of systemic administration of atipamezole with morphine ... 75
5.4 Spironolactone (study III) .................................................................... 76
5.4.1 Effects of spironolactone in acute morphine antinociception and
morphine tolerance ....................................................................................................... 76
5.4.2 Effects of spironolactone on the brain morphine and morphine-3glucuronide concentrations ....................................................................................... 76
5.4.3 Antinociceptive effects of the P-glycoprotein substrate loperamide
combined with spironolactone ................................................................................. 77
5.5 Ketamine (study IV) ............................................................................. 78
5.5.1 Ketamine coadministration to morphine-tolerant rats leads to
increased brain concentrations of morphine, ketamine, and norketamine
............................................................................................................................................... 78
5.5.2 Effects of combined ketamine and morphine in acute models of
nociception ....................................................................................................................... 79
6. Discussion ................................................................................................... 81
6.1 Methodological considerations .......................................................... 81
6.2 Ibudilast in coadministration with morphine ................................... 81
6.3 Morphine combined with ultralow doses of atipamezole ................ 83
6.4 Spironolactone for the treatment of opioid tolerance ..................... 85
6.5 A novel pharmacokinetic interaction between morphine and
ketamine ..................................................................................................... 87
6.6 Future perspectives ............................................................................. 89
Conclusions .................................................................................................... 92
Acknowledgements ....................................................................................... 93
References ...................................................................................................... 95
Original publications ................................................................................... 120
LIST OF ABBREVIATIONS
5-HT
ABC
AIF1
AMPA
ANOVA
AUC
b.i.d.
BBB
Bcrp
CaMKII
cAMP
CCD
CCI
CD11b
cGMP
CGRP
CNS
CREB
CTAP
CTOP
CV
CYP450
DDD
DNA
DPDPE
DRG
EAAT
EC50
ED50
EPSC
FSH
G protein
GABA
GDNF
5-hydroxytryptamine
Adenosine triphosphate-binding cassette
Allograft inflammatory factor 1
alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid
Analysis of variance
Area under curve
Bis in die (twice daily)
Blood–brain barrier
Breast cancer resistance protein
Ca2+/calmodulin-dependent protein kinase
Cyclic adenosine monophosphate
Chronic compression of the dorsal root ganglion
Chronic constriction injury
Cluster of differentiation molecule 11b
Cyclic guanosine monophosphate
Calcitonin gene related peptide
Central nervous system
cAMP response element-binding protein
D-Phe-Cys-Tyr-D-Trp-Arg-Thr-Pen-Thr-NH2
D-Phe-Cys-Tyr-D-Trp-Orn-Thr-Pen-Thr-NH2
Coefficient of variation
Cytochrome P 450 family
Defined daily dose
Deoxyribonucleic acid
D-penicillamine(2,5)-enkephalin
Dorsal root ganglion
Excitatory amino acid transporter
Half maximal effective concentration
Half maximal effective dose
Excitatory postsynaptic current
Follicle-stimulating hormone
Guanine nucleotide-binding regulatory protein
Gamma-aminobutyric acid
Glial cell line-derived neurotrophic factor
6
GFAP
GPCR
GRK
HPLC
i.c.v.
i.m.
i.p.
i.t.
i.v.
IASP
Iba1
IFN
IL
JNK
Ki
LC-MS/MS
LD50
LE
LH
LPS
M3G
M6G
MAC
MAPK
MD2
MIF
MPE%
MPP
Mrp
MS
NGF
NMDA
NMDAR
NO
Oatp
OIH
ORL
Glial fibrillary acidic protein
G protein coupled receptor
G protein coupled receptor kinase
High pressure liquid chromatography
Intracerebroventricular
Intramuscular
Intraperitoneal
Intrathecal
Intravenous
International Association for the Study of Pain
Ionized calcium-binding adapter molecule 1
Interferon
Interleukin
c-Jun N-terminal kinase
Inhibition constant
Liquid chromatography-tandem mass spectrometry
Half maximal lethal dose
Long-Evans (rat strain)
Luteinizing hormone
Lipopolysaccharide
Morphine-3-glucuronide
Morphine-6-glucuronide
Minimum alveolar concentration
Mitogen-activated protein kinase
Lymphocyte antigen 96
Macrophage migration inhibitory factor
Percentage of the maximum possible effect
1-methyl-4-phenylpyridinium
Multidrug resistance protein
Mass spectrometry
Nerve growth factor
N-methyl-D-aspartate
N-methyl-D-aspartate receptor
Nitric oxide
Organic anion transporter polypeptide
Opioid-induced hyperalgesia
Opioid receptor-like
7
P-gp
p.o.
PAG
PB
PCA
PCP
PDE
PKA
PKC
PNS
PTX
q.d.
RVM
S/N
s.c.
SD
SD
SEM
SGC
SNL
TG
TLR
TNF
UDP
UGT
P-glycoprotein
Peroral
Periaqueductal gray
Parabrachial nucleus
Patient controlled analgesia
Phencyclidine
Phosphodiesterase
Protein kinase A
Protein kinase C
Peripheral nervous system
Pertussis toxin
Quaque die, once daily
Rostroventral medulla
Signal-to-noise ratio
Subcutaneous
Sprague-Dawley (rat strain)
Standard deviation
Standard error of the mean
Satellite glial cell
Spinal nerve ligation
Trigeminal ganglion
Toll-like receptor
Tumor necrosis factor
Uridine diphosphate
Uridine diphosphate glucuronosyl transferase
8
ABSTRACT
Opioid analgesics are effective in relieving acute and chronic pain.
However, adverse effects and the development of opioid dependence and
tolerance may restrict the use of opioids and result in inadequate pain
relief. The effects of four structurally and functionally different drugs
already on the market, ibudilast, atipamezole, spironolactone, and
ketamine, were studied in coadministration with morphine, the
prototypical mu-opioid receptor agonist. Experiments were conducted
using thermal and mechanical tests of nociception in male SpragueDawley rats. Morphine tolerance was produced during four days by
subcutaneous or intrathecal delivery of morphine. Drug and metabolite
concentrations
were
measured
using
high-pressure
liquid
chromatography-tandem mass spectrometry. The objective of the thesis
study was to search for potential drugs to augment morphine
antinociception and prevent opioid tolerance.
Ibudilast, a phosphodiesterase and macrophage inhibitory factor
inhibitor, had transient sedative effects, but it restored the antinociceptive
effect of morphine in morphine-tolerant rats after single and repeated
administration. It did not prevent the development of opioid tolerance.
Atipamezole, an alpha-2-adrenoceptor antagonist used for the
reversal of sedation in animals during anesthesia, was effective in
augmenting intrathecal morphine antinociception in both opioid-naïve
and opioid-tolerant animals. These effects were observed at doses lower
than those required for the antagonism of alpha-2-adrenoceptors. In
subcutaneous administration, low doses of atipamezole did not influence
morphine antinociception.
The mineralocorticoid receptor antagonist spironolactone dosedependently enhanced morphine antinociception. This effect was
mediated via the increased access of morphine to the central nervous
system by the inhibition of the efflux protein P-glycoprotein.
Spironolactone did not inhibit the metabolism of morphine to the
pronociceptive metabolite morphine-3-glucuronide, and it did not prevent
the development of opioid tolerance.
The effects of ketamine in augmenting opioid analgesia in tolerance
are thought to result from a beneficial pharmacodynamic interaction.
9
When acute ketamine was administered to rats under chronic morphine
treatment, the brain concentrations of morphine, ketamine and
norketamine were increased compared with the situation where either
morphine treatment or acute ketamine were administered alone. The
results indicate a potentially beneficial pharmacokinetic interaction
between the two drugs.
The results of the thesis study demonstrate that ibudilast and
atipamezole modulate nociception at systemic and spinal levels in
preclinical models of pain, and they may prove advantageous as an
adjuvant to opioid therapy. Spironolactone had a pharmacokinetic
interaction with morphine, leading to increased morphine concentrations
in the central nervous system. Ketamine, a drug used for the treatment of
opioid tolerance in cancer patients, may undergo previously unrecognized
beneficial pharmacokinetic interactions with morphine.
10
LIST OF ORIGINAL PUBLICATIONS
This thesis is based on the following original publications (studies I–IV)
and some unpublished data.
I.
Lilius, T.O., Rauhala, P.V., Kambur, O., Kalso, E.A. (2009).
Modulation of morphine-induced antinociception in acute and
chronic opioid treatment by ibudilast. Anesthesiology 111, 1356–
1364.
II.
Lilius, T.O., Rauhala, P.V., Kambur, O., Rossi, S.M., Väänänen,
A.J., Kalso, E.A. (2012). Intrathecal atipamezole augments the
antinociceptive effect of morphine in rats. Anesth Analg 114,
1353–1358.
III.
Lilius, T.O.*, Jokinen, V.*, Neuvonen, M.S., Väänänen, A.J.,
Niemi, M., Rauhala, P.V., Kalso, E.A. (2014). The
mineralocorticoid receptor antagonist spironolactone enhances
morphine antinociception. Eur J Pain 18, 386–395.
IV.
Lilius, T.O., Jokinen, V., Neuvonen, M.S., Niemi, M., Kalso,
E.A., Rauhala, P.V. (2014). Ketamine coadministration
attenuates morphine tolerance and leads to increased brain
concentrations of both drugs. In press, Br J Pharmacol.
DOI: 10.1111/bph.12974
* These authors contributed equally to the publication.
The original publications are reproduced with the permission of their
copyright holders.
11
1. INTRODUCTION
The treatment of acute and chronic severe pain remains a major challenge.
Opioids, such as morphine and oxycodone, are used as primary analgesics in
moderate to severe pain (Hoskin, 2008; Kalso et al., 2004). However, long-term
opioid treatment, in particular, has several problems, such as the development of
tolerance and dependence with withdrawal symptoms. Tolerance is manifested
as a need to increase the opioid dose to reach an equipotent analgesic effect,
whereas opioid-induced hyperalgesia is a process of sensitization in which
opioids cause hypersensitivity to painful stimuli. Dose escalation of opioids is
typically limited by adverse effects such as respiratory depression, nausea, and
constipation (McNicol et al., 2003). The adverse effects and tolerance with
hyperalgesia may necessitate the discontinuation of opioid treatment and result
in inadequate pain control.
The classic models for explaining the development of opioid tolerance and
hyperalgesia have mostly focused on cellular and network changes in neurons
themselves. The changes caused by repeated administration of opioids and
consequent opioid receptor activation may involve adaptive changes in the
neurons, such as the internalization of opioid receptors (Zuo, 2005),
upregulation of N-methyl-D-aspartate receptor function (Mao et al., 2002;
Shimoyama et al., 2005), production of nitric oxide, or down-regulation of
glutamate transporters (Pasternak, 2007). Moreover, a counter-regulatory
antiopioid system could be involved, including neuromodulators such as
cholecystokinin or dynorphin (Xie et al., 2005). The formation of heterodimeric
receptor complexes between mu-opioid and other receptors, such as the alpha-2adrenoceptor, may modulate the function of the mu-opioid–receptor coupled
effectors and the efficacy of opioids (Jordan et al., 2003).
Recent studies suggest that opioid administration may also induce
tolerance via mechanisms other than those involving neurons. The glial cells of
the central nervous system (CNS) have previously been considered as
neuroimmune cells that mainly provide support and nutrition for the neurons.
New data indicate that activated glial cells may modulate the activity of the
nociceptive neurons in the CNS (Ji et al., 2013; Grace et al., 2014) by releasing
various proinflammatory substances and actively oppose the analgesic action of
opioids on pain-transmission neurons (Wang et al., 2012b). Another line of
opioid tolerance research has focused on the pharmacokinetics of opioids,
12
especially their potentially reduced access to the CNS through the blood–brain
barrier (BBB) (Mercer and Coop, 2011). Of the known BBB drug transporters,
morphine is an acknowledged substrate of the efflux transporter P-glycoprotein
(P-gp) (Letrent et al., 1998; Schinkel et al., 1995; Xie et al., 1999). Upregulation
of this transporter by opioids could lead to decreased access of opioids to the
brain and compromised analgesia. However, limited knowledge exists
concerning other potential opioid-transporting proteins and pharmacokinetic
drug–drug interactions between analgesics.
The present investigations were conducted to acquire further information
on possible adjuvant drugs that could improve the efficacy of opioids in acute
and chronic treatment, with particular emphasis on both pharmacodynamic and
pharmacokinetic interactions.
13
2. REVIEW OF THE LITERATURE
2.1 Physiology of nociception and pain
2.1.1 Primary nociceptors
Somatosensation includes several submodalities, including touch (detection of
light mechanical stimuli), proprioception (detection of mechanical stretch of
muscles and joints), thermosensation (detection of cold and warmth), and
nociception (detection of noxious thermal, mechanical, or chemical stimuli), that
evoke pain sensations (Julius, 2013). Noxious stimuli are initially detected by
primary afferent sensory nerve fibers that innervate a peripheral target. This
information is transmitted to secondary afferent dorsal horn neurons in the
spinal cord and further to the brain via ascending neural circuits. The cell bodies
of primary afferent sensory nerves, referred to as nociceptors, lie in the dorsal
root ganglia (DRG) for the body and the trigeminal ganglion for the face. They
have a peripheral and a central axon innervating the target organ and the spinal
cord, respectively. The nociceptors only discharge action potentials when they
are stimulated to a noxious range, indicating that they are only capable of
detecting and responding to potentially injurious stimuli (Todd and Koerber,
2013).
Nociceptors are divided into two major classes: 1) medium-diameter
myelinated A-delta fibers that convey acute, fast pain and 2) small-diameter
unmyelinated C fibers that convey poorly localized, slow pain. C fibers are
usually divided into peptidergic and nonpeptidergic neurons. Peptidergic
neurons may contain substance P and calcitonin gene-related peptide (CGRP)
and express the nerve growth factor receptor TrkA. The nonpeptidergic
population expresses the c-Ret neurotrophin receptor with glial-derived
neurotrophic factor as its agonist. They usually lack substance P and CGRP, but
express purinergic P2X3 and P2Y1 receptors. The large-diameter A-beta fibers
are rapidly conducting primary afferent nerves that do not convey nociceptive
information. However, their activity may modulate pain perception by activating
inhibitory interneurons in the spinal cord (Todd and Koerber, 2013).
14
Review of the literature
2.1.2 Nociception at the spinal cord level
The dorsal horn of the spinal cord is organized into histologically and
functionally distinct laminae. They receive their input by the primary afferent
nerve fibers. A-delta fibers generally project to lamina I and also to the deep
lamina V, whereas the C fibers project to the superficial laminae I and II, with
some terminals in deeper laminae (Fig. 1A). The A-beta fibers, mediating light
touch, project to laminae III, IV and V (Basbaum et al., 2009; Todd and
Koerber, 2013). In addition to the primary afferent nerves and projection
neurons, one-third of the neuron population consists of inhibitory interneurons
using gamma-amino butyric acid (GABA) and/or glycine as neurotransmitters.
Inhibition at the spinal cord level is an important form of endogenous pain
control. For example, in the absence of sensory stimuli, the release of tonic
inhibitory neurotransmitter inhibits spontaneous discharges of the spinal dorsal
horn neurons (for review, see Todd, 2010).
2.1.3 Nociception and pain in supraspinal structures
Projection neurons located in laminae I and V are responsible for most of the
output from the dorsal horn to the supraspinal structures (Basbaum and Jessell,
2000). The spinothalamic tract, which conveys nociceptive messages to the
thalamus, is especially relevant for the sensory-discriminative aspects of pain,
whereas the spinoreticulothalamic tract conveys nociceptive information to the
brainstem and is more relevant in poorly localized nociception. There are also
connections to the amygdala via the parabrachial region of the dorsolateral pons.
This connection has been associated with information related to the
unpleasantness of the nociceptive stimulus. From these loci, information reaches
the cortical areas (Fig. 1B). Pain has been suggested to result from the activation
of numerous structures, with the somatosensory cortex being largely
responsibility for the sensory-discriminative properties of pain and others, such
as the anterior cingulate cortex and insular cortex, having responsibility for the
emotional aspects (Apkarian et al., 2005). The pain matrix concept involves
several interacting neuronal networks, of which the nociceptive matrix (mainly at
the posterior insular region), receiving input from the spinothalamic projection,
is responsible for cortical nociceptive processing. This information is further
transferred to a second-order network and to perception involving parietal,
prefrontal, and anterior insular areas (Garcia-Larrea and Peyron, 2013).
15
Review of the literature
Cingulate cortex
B
Somatocensory
cortex
Insular
cortex
Thalamus
A
Amygdala
A-delta myelinated
fibers
Peptidergic
C fibers
PAG
PB
Brainstem
R
V
M
Lamina I
Lamina II
Nonpeptidergic
C fibers
Nociceptor
Lamina V
Spinal cord
Figure 1. Schematic diagram of the spinal cord dorsal horn (A) and the supraspinal
structures (B) involved in nociception and pain. PAG, periaqueductal gray; PB,
parabrachial nucleus; RVM, rostroventral medulla. Adapted from Basbaum et al.
(2009) and Todd and Koerber (2009).
According to the International Association for the Study of Pain (IASP), pain is
described as “an unpleasant sensory and emotional experience associated with
actual or potential tissue damage, or described in terms of such damage”. This
implies that pain can only be experienced at the whole-organism level. Animals
cannot verbally report pain, but certain recurrent reactions to nociceptive stimuli
can be interpreted as nociception. The extrapolation of the analgesic response
from vastly differing pain models or the use of measures with unclear relevance is
a problem in animal research. Attempts are being made to develop new specific
pain measurement methods. For example, a facial expression pain coding system
for the mouse has recently been introduced (Matsumiya et al., 2012), a system
similar to those used for evaluating pain in nonverbal human populations.
16
Review of the literature
2.1.4 Supraspinal mechanisms of pain modulation
The relationship between the peripheral noxious stimulation and the perceived
pain experience depends on many variables, including the behavioral context and
other sensory inputs. Modification of the pain response results from the action
of CNS networks. Suppression of the nociceptive reflexes and pain sensation is
essential in the face of a threat, whereas enhanced pain in tissue inflammation or
injury may promote sheltering and recuperation. At present, it is widely
acknowledged that the continuous activation of descending pain facilitator
networks is a contributor to chronic pain states (Heinricher and Fields, 2013).
The first evidence for supraspinal sites controlling ascending sensory
pathways was published in the 1950s, and a specific pain modulatory system
theory was introduced by Melzack and Wall (1965) in the Gate control theory of
pain. Descending projections to the spinal cord control spinal nociception either
directly or through spinal interneurons (Todd, 2010). The discovery of
periaqueductal gray (PAG) and its central role in the brain mechanisms of pain
processing was a critical step in understanding the modulation of nociception.
Simple effects evoked by noxious stimulus at the spinal cord level may be
inhibited by the stimulation of PAG or its relay in the rostroventral medulla
(RVM). The major output of the RVM is to the spinal cord via neurons
including 5-hydroxytryptamine (5-HT), GABA, and several neuropeptides
(Skagerberg and Björklund, 1985). PAG receives direct projections from sites
involved in supraspinal pain processing, including the cingulate and insular
cortex, hypothalamus, and amygdala (Aggleton et al., 1980; Floyd et al., 2000;
Rizvi et al., 1996).
The PAG-RVM is crucial for the analgesic actions of opioids and
endogenous opioid peptides (see 2.2.3). Interestingly, other substances such as
cannabinoids also partly produce their antinociceptive effect by activating the
PAG-RVM system. Another important descending modulatory system is the
pontomedullary noradrenergic pathway, reciprocally linked with the PAG-RVM
system (see 2.3.4).
17
Review of the literature
2.2 Opioids
2.2.1 Indications and prevalence of opioid therapy in Finland
Opioids are the most potent analgesic agents used for pain therapy. They are the
gold standard in the treatment of acute postoperative pain and also severe to
moderate cancer pain (Caraceni et al., 2012). The usage of opioids in the
treatment of non-cancer chronic pain has significantly increased. In Finland,
nine different opioids are currently on the market, grouped into weak (codeine
and tramadol), intermediate (buprenorphine), and strong (fentanyl,
hydromorphone, methadone, morphine, oxycodone, and pethidine) categories
based on their pharmacological efficacy. In 1990, the total consumption of
opioids was three defined daily doses (DDD) per 1000 inhabitants, increasing to
a level of 17 DDD per 1000 inhabitants by 2011, indicating a six-fold increase
in opioid consumption during 20 years. This development goes hand in hand
with the global development: the use of prescribed opioids increased threefold
from 1989 to 2009, and the use of morphine worldwide increased from 6.5 tons
in 1989 to 41.8 tons in 2009 (Huxtable et al., 2011). The majority of opioids
consumed in Finland are weak opioids, mainly codeine. The consumption of
morphine has been stable for the last 10 years (0.27 DDD), whereas the
consumption of oxycodone has increased almost threefold to 0.81 DDD. In
2011, strong opioids were prescribed to 25 000 outpatients, and the total
number of inhabitants who had used opioids in 2011 was 410 000. Based on the
Finnish registries, it is not possible, however, to determine the individual
consumption for different diagnoses (Nevantaus et al., 2013; summary in
English).
2.2.2 Opioid receptors
Opioid receptors belong to the superfamily of guanine nucleotide-binding
regulatory protein (G protein) coupled receptors (GPCRs), a family that
contains several hundred members in the human genome (Lagerström and
Schiöth, 2008). The receptor contains seven transmembrane domains that are
connected by short loops. They display an extracellular N-terminal domain and
an intracellular C-terminal tail. The opioid receptors were first demonstrated in
the 1970s (Pert and Snyder, 1973), and in the 1990s the opioid receptor family,
including the classical mu, kappa, and delta receptors, as well as the closely
related opioid receptor like (ORL) 1 receptor, were cloned and hence
18
Review of the literature
molecularly identified (Taylor and Dickenson, 1998). Mu, delta and kappa
receptors are highly homologous, with greatest variability within the extracellular
loops and the N- or C-terminal tails. In 2012, the crystal structure of the kappa
receptors was determined with X-ray crystallography (Wu et al., 2012),
representing another milestone in opioid research. Pharmacological activities of
the prototypal opioid receptor agonists have been tested in mice lacking the mu-,
delta- or kappa-opioid receptor genes. The results suggest that the mu-opioid
receptor is responsible for the analgesic activities of morphine (Kieffer and
Gavériaux-Ruff, 2002) (Fig. 2).
Mu-opioid receptors are coupled to pertussis toxin-sensitive Gα proteins
Go and Gi to inhibit the cyclic adenosine monophosphate (cAMP) pathway by
inhibiting adenylyl cyclase activity (Connor and Christie, 1999; Laugwitz et al.,
1993). By initiating the release of the Gβγ dimer from Gi or Go, opioids reduce
the opening of voltage-dependent calcium channels (Saegusa et al., 2000) or
activate inwardly rectifying potassium channels (Ikeda et al., 2000), phenomena
leading to decreased neuronal excitability and analgesia. Opioids also may
activate mitogen-activated protein kinase (MAPK) cascades, affecting the
transcriptome of the cell.
Agonist
Receptor
Effect
Morphine
DPDPE
Deltorphin
Mu
Delta
Reward
Dependence
Analgesia
U50,488H
Kappa
Dysphoria
Figure 2. Opioid receptor-mediated effects. The main pharmacological activities of
the opioid receptor agonist prototypes have been assessed in mu-, delta-, and kappa
knockout mice. Modified from Kieffer and Gavériaux-Ruff (2002). DPDPE, Dpenicillamine(2,5)-enkephalin.
19
Review of the literature
2.2.3 Mechanisms of opioid-induced analgesia
The actions of opioids in mediating analgesia after systemic administration
present their actions in both the brain and spinal cord.
Within the spinal cord, opioid receptors are predominantly located in
laminae I and II of the dorsal horn, with smaller quantities in deeper layers. The
contribution of mu, delta, and kappa receptors to opioid binding in the rat spinal
cord is roughly 70%, 24% and 6%, respectively (Besse et al., 1990; Rahman et al.,
1998). Over 70% of the receptors are located on the central terminals of smalldiameter (mostly C and alpha-delta) primary afferent neurons, implying that the
main mechanism of spinal opioid analgesia lies in the activation of presynaptic
opioid receptors, decreasing the release of excitatory transmitters and nociceptive
transmission. Another confirmation of the presynaptic action is provided by the
observation that the release of primary afferent transmitters such as substance P
and CGRP, contained in small afferents, is reduced after opioid administration
(Go and Yaksh, 1987). Taken together, the capacity of spinal opioids to reduce
the release of excitatory neurotransmitters and to decrease neuron excitability in
the dorsal horn is considered to mediate the powerful effects of opioids on
nociception at the spinal cord level. In humans, there is extensive literature
indicating the clinical efficacy of spinal opioids (Yaksh, 1997).
As opioid receptors are synthesized in the DRG cell bodies, they are also
transported peripherally to the axons. The peripheral application of exogenous
opioid agonists has only been antinociceptive in inflammatory states (Machelska
et al., 1999).
Microinjections of opioids to specific brain sites have demonstrated that
opioid agonists, in a manner consistent with their activity on the mu-opioid
receptor, block nociceptive behavior. The supraspinal sites where morphine
microinjections have shown efficacy are the insular cortex, amygdala,
hypothalamus, PAG, and RVM. Interest in the PAG-RVM system was raised
when it was discovered that opioid analgesia is also partly mediated via direct
post-synaptic hyperpolarization of a neuron subpopulation in this area (Pan et
al., 1990; Yaksh et al., 1988). On the other hand, RVM may also contribute to
hyperalgesia and allodynia in inflammatory and neuropathic pain (Porreca et al.,
2002) or prolonged opioid administration (Vanderah et al., 2001). Additionally,
especially in humans, opioids also have supraspinal mood-altering effects that
modulate the experience of pain. In addition to relief of distress and increased
20
Review of the literature
tolerability of pain, patients may also experience euphoria (Yaksh and Wallace,
2011).
2.2.4 Morphine
Morphine
([5α,6α]-7,8-didehydro-4,5-epoxy-17-methylmorphinan-3,6-diol)
(Fig. 3) is the main active substance in opium. As it is difficult to synthesize, it is
still extracted from poppy straw or obtained from opium. It has become the gold
standard of opioids against which all other opioids are compared. Its main effect
is mediated by binding to and activating the mu-opioid receptors in the CNS.
Morphine is also the prototypical agonist of the mu receptor (Yaksh and
Wallace, 2011).
2
3
1
4
11
10
12
9
13
14
5
15
6
16
17
8
7
Figure 3. Structural formula of morphine.
After peroral administration, the absorption of morphine from the
gastrointestinal tract is modest with a bioavailability of approximately 25%.
Hence, its first-pass metabolism in the liver and gut wall is significant. In
parenteral intravenous, intramuscular, and subcutaneous (i.v., i.m., and s.c.,
respectively) administration, the onset of the effects is more rapid and the effect
of the same dose is larger. Morphine is also available as suppositories with a
rather good absorption through the rectal mucosa. About one-third of morphine
in plasma is bound to proteins after a therapeutic dose. The analgesic plasma
concentration range in patients suffering from cancer is very large, 22–364
ng/mL (Neumann et al., 1982).
Morphine is glucuronidated to morphine-3-glucuronide (M3G) and
morphine-6-glucuronide (M6G) by uridine diphosphate (UDP)-glucuronosyl
21
Review of the literature
transferase 2B7 (UGT2B7) in humans (Coffman et al., 1997), and mostly to
M3G by Ugt2b1 in rodents (Coffman et al., 1996; King et al., 1997; Pritchard
et al., 1994). A small amount is metabolized to normorphine via the cytochrome
P-450 (CYP) system, especially CYPs 3A4 and 2C8 (Projean et al., 2003). In
humans, after oral morphine administration, the areas under curve (AUC) of
M6G and M3G exceed that of the parent drug by factors of 9:1 and 50:1,
respectively (Osborne et al., 1990). Morphine is mainly excreted via the kidney
in the 3-glucuronide form; very little morphine is excreted unchanged. M6G is
also excreted by the kidney, and in kidney failure it may accumulate, probably
explaining the increased and prolonged effect of morphine in this patient group
(Owen et al., 1983). Some enterohepatic circulation of morphine and its
glucuronides occurs, which can be seen as traces of morphine in the feces for
some days after the end of therapy (Walsh and Levine, 1975).
In rodents, the disposition of morphine in the CNS is modulated by P-gp
(Letrent et al., 1999; Schinkel et al., 1995; Xie et al., 1999), an efflux protein at
the blood–brain barrier (BBB). Thus, P-gp inhibition could lead to increased
access of morphine to the CNS (see 2.3.5).
2.2.5 Adverse effects of morphine
Opioid therapy may be compromised by adverse effects (Table 1 and Fig. 4), of
which most are mediated through the opioid receptors. The severity of these
adverse effects on the patient may range from mild to severe: constipation is
common but it may generally be alleviated with pharmacological methods,
whereas respiratory depression is the most common cause of death associated
with opioid toxicity (Pattinson, 2008). The risk of respiratory depression in pain
treatment is lower than in opioid abuse. It has been suggested that the
respiratory center also receives nociceptive input, and pain thus acts as a
physiological antagonist for opioid-induced respiratory depression (Hanks and
Twycross, 1984). Acute potentially dangerous adverse effects of opioids may be
antagonized with naloxone, the prototypical opioid receptor antagonist, given
intravenously or subcutaneously. The bioavailability of naloxone is poor, leading
to minimal systemic concentrations after peroral administration. Interestingly,
this effect may be advantageous in the treatment of constipation, in which the
target of opioid receptor antagonism lies in the enteric nervous system of the gut.
Indeed, oral naloxone and subcutaneous methylnaltrexone are currently in use to
reverse opioid-induced constipation without compromising analgesia (Leppert,
2010).
22
Review of the literature
Table 1. Possible adverse effects of opioids in chronic pain treatment (Lawlor and
Bruera, 1998; Schug et al., 1992).
Neuropsychiatric
Delirium, sedation, mood changes, anxiety, sleep disturbance
Neurological
Myoclonus, miosis, hyperalgesia, allodynia, seizures
Gastrointestinal
Dry mouth, nausea and vomiting, constipation, biliary spasm
Respiratory
Respiratory depression, pulmonary edema, cough suppression
Urological
Micturitional disturbance
Dermatological
Pruritus, sweating
Endocrine
Hormonal changes, immune changes
Pharmacological
Tolerance, dependence
Opioids affect two major hormonal systems: the hypothalamus-pituitary-adrenal
axis and the hypothalamus-pituitary-gonadal axis. Morphine causes a strong
decrease in plasma cortisol levels in adults (Banki and Arato, 1987) and
laboratory animals (Bartolome and Kuhn, 1983; Collu et al., 1976; Rolandi et
al., 1983). In the gonadal axis, typical changes in hormonal secretion are a
decrease in luteinizing hormone (LH), follicle-stimulating hormone (FSH),
testosterone, and estrogen levels, whereas an increase in prolactin levels may be
observed (Malaivijitnond and Varavudhi, 1998). Effects of opioids on the
immune system may be mediated through their direct effects on immune cells,
as opioid-related receptors on immune cells have been discovered (Makman,
1994), or indirectly via changes in neuroendocrine system. Opioids appear to
alter both innate and adaptive immune systems (Risdahl et al., 1998; Roy and
Loh, 1996). Results from in vivo and in vitro studies suggest that prolonged
opioid exposure is likely to suppress immune function in the periphery (Roy et
al., 2006; Sacerdote et al., 1997); for a review, see (Ninković and Roy, 2013).
Interestingly, microglial cells, immune cells of the CNS, have recently gained
considerable attention in the study of pain (see 2.3.3).
23
Review of the literature
Opioid therapy
Cellular&
mechanisms&
Pharmacological&
tolerance&
Opioid7induced&
hyperalgesia&
Apparent&
tolerance&
Immune&&
changes&
Hormonal&
changes&
Dose&
escala;on&
Figure 4. Possible adverse effects specific to prolonged opioid therapy (Ballantyne
and Mao, 2003).
2.3 Modulation of opioid-induced antinociception
2.3.1 Opioid tolerance and opioid-induced hyperalgesia
Opioid tolerance is a pharmacological phenomenon that develops during chronic,
repeated use of opioids and leads to the need to increase the opioid dose to reach
an equipotent analgesic effect, i.e. the dose-response curve is shifted to the right.
Two types of tolerance, associative and non-associative tolerance, can be
distinguished and they appear to involve different neurotransmitter mechanisms
(Grisel et al., 1996; Mitchell et al., 2000). Associative tolerance involves
psychological factors, and they may be clinically connected to nonpharmacological aspects of patient care, for example admission to the hospital
ward. Non-associative tolerance occurs at the cellular or organism level and
involves numerous intracellular changes, as well as changes in the function of
wider neuronal networks (Grisel et al., 1996). A prevailing theory is that opioid
tolerance is a complex process that involves multiple regulatory mechanisms
occurring at the level of both neurons and neuronal networks. Many features of
opioid tolerance may be seen as a natural attempt of neuronal homeostasis to
24
Review of the literature
reach equilibrium, i.e. involving regulatory or opponent processes to overcome
prolonged activation of opioid receptors. Desensitization, on the other hand, is a
process referring to the acute loss of mu-opioid receptor and effector coupling
that occurs within seconds to minutes after the initiation of opioid exposure.
Both tolerance and desensitization may be used interchangeably to describe the
loss of receptor activity after chronic treatment. (Williams et al., 2013)
Tolerance, however, does not necessarily specify the cellular or molecular
mechanisms that are responsible for the phenomenon.
Multiple opioid receptor-related phenomena may contribute to cellular
tolerance. Phosphorylation of the mu-opioid receptor is a widely accepted theory
of desensitization. Many kinases, e.g. G protein receptor kinase (GRK), c-Jun
N-terminal kinase (JNK), protein kinases A and C (PKA and PKC),
Ca2+/calmodulin-dependent protein kinase (CaMKII), and MAPK, may be
involved in desensitization and receptor endocytosis (Koch and Höllt, 2008).
The endocytosis pathway does not necessarily lead to receptor degradation, but it
may also lead to dephosphorylation of the receptor and its replacement on the
surface of the cell. However, morphine is only a weak stimulator of mu-opioid
receptor phosphorylation (Zhang et al., 1998). Based on these data, it has been
speculated that morphine tolerance may not be a result of receptor
desensitization but rather the lack of desensitization, which may lead to
prolonged receptor signaling (Whistler and Zastrow, 1998; Williams et al.,
2013). Downregulation of functional mu-opioid receptors is not considered
relevant for morphine tolerance, as functional binding studies have implied that
the loss of over 80% of functional mu-opioid receptors is needed to observe a
rightward shift in the half maximal effective concentration (EC50) of morphine
(Stafford et al., 2001; Williams et al., 2013).
The mu-opioid receptor is coupled to cellular effectors such as adenylyl
cyclase and Ca2+ and K+ channels, affecting the release of neurotransmitters.
Repeated opioid administration initiates counterregulatory processes in these
systems, of which the best described is the superactivation of adenylyl cyclase
and increased intracellular cAMP concentrations. This, in turn, induces PKA
activation and an increased inward cation current through Ca2+ channels (Bailey
and Connor, 2005; Collier, 1980). Even though different opioids share the same
binding site (presumably the mu-opioid receptor) to produce their physiological
effects, there may be significant differences in the activation of different G
proteins and intracellular downstream responses. This is called functional
selectivity or biased agonism; for review, see Raehal et al. (2011).
25
Review of the literature
Opioid tolerance not only has receptor- and effector-level effects, but also
includes system-level counteradaptation. The observed tolerance may result from
not only pharmacological tolerance, but also enhanced excitability of the opioidregulated pathways. Indeed, opioids may initiate a pronociceptive process also
known as opioid-induced hyperalgesia (OIH). While tolerance at simplest is
characterized by a progressive decrease in the half maximal effective dose (ED50)
that can be overcome by increasing the drug dose, OIH is a process of
sensitization in which opioids cause hypersensitivity to pain (Fishbain et al.,
2012). The relative contributions of these two processes to the observed overall
decrease in the opioid analgesic effect has not been established, and these
processes may have mutual intercellular characteristics and contribute in parallel
to the decrease the clinical efficacy. The common mechanisms may involve
activation of bulbospinal pathways that increase the secondary neuron
excitability of the spinal dorsal horn via a counter-regulatory antiopioid system,
including neuromodulators such as cholecystokinin or dynorphin (Gardell et al.,
2002; Xie et al., 2005). Glutamatergic cellular mechanisms that lead to
upregulation of N-methyl-D-aspartate (NMDA) receptor function may also
contribute to OIH (Mao et al., 1994). OIH has been described in rodents
(Célèrier et al., 2000; Li et al., 2001a; Mao et al., 1994; Vanderah et al., 2001,
Minville et al., 2010) and pain-free human volunteers receiving opioid infusions
(Koppert et al., 2003; Luginbühl et al., 2003; Tröster et al., 2006, Vinik and
Kissin, 1998); for a meta-analysis, see Fishbain et al. (2009). However, data in
many other settings are contradictory (Fishbain et al., 2009), and there are no
clinical data on the prevalence of OIH (Lee et al., 2011). During opioid
maintenance therapy in addicts, a modality-specific increased pain sensitivity has
been observed in studies with small populations (Compton et al., 2001; 2012;
Prosser et al., 2008; Doverty et al., 2001). In postoperative studies, OIH has
mainly been studied after opioid-based anesthesia (Chia et al., 1999; Cooper et
al., 1997; Guignard et al., 2000; Katz et al., 1996; Lahtinen et al., 2008;
Schmidt et al., 2007; Schraag et al., 1999; Shin et al., 2010). In a recent metaanalysis, high intraoperative doses of remifentanil, a potent opioid with a short
elimination half-life, were associated with a small increase in postoperative pain
(Fletcher and Martinez, 2014). OIH may also be a clinical problem in chronic
pain (Hay et al., 2009; Chu et al., 2006; Hooten et al., 2010). In practice, the
discontinuation of a large dose of morphine and substitution with another opioid
(opioid rotation) has been a route to reverse opioid tolerance, implicating
26
Review of the literature
morphine or its accumulating metabolite morphine-3-glucuronide in the
phenomenon (Andersen et al., 2003; Sjøgren et al., 1994; Smith, 2000).
Pharmacokinetic factors, such as decreased absorption of oral preparations,
accelerated drug elimination in the liver and kidney, or changes in the BBB drug
transport function may lead to decreased opioid access to the CNS (see 2.3.5). It is
also important to distinguish the marked differences between the actions of M3G
and M6G, the primary metabolites of morphine in humans. Upon morphine
administration, both metabolites are observable in the brain, even though they
have limited permeability through the BBB (Bickel et al., 1996). Systemically
administered M6G is approximately twice as potent as morphine in humans
(Osborne et al., 1990) and also in animal models (Paul et al., 1989). During
prolonged oral administration in humans, the plasma concentration of M6G
generally exceeds that of morphine. In rats, however, M6G is not formed at
detectable concentrations (Coughtrie et al., 1989). M3G, the other primary
metabolite of morphine, has the opposite actions. It has very little affinity for the
opioid receptor, but may contribute to the harmful excitatory effects of morphine
and OIH (Andersen et al., 2003; Sjøgren et al., 1994; Smith, 2000).
On an individual level, there are marked differences in the variation of the
response to a given opioid. The genetics involved in opioid pharmacodynamics or
pharmacokinetics may play a role in the response to opioid drugs, but their
influence on this variation is complex. Personalized medicine is a rising field of
medicine that aims at the customization of medical treatment to personally suit
the needs of an individual patient. The majority of genetic studies have so far
focused on single genes or single nucleotide polymorphisms, and research into the
involvement of gene–gene interactions in the opioid response is only just starting;
for a review, see Branford et al. (2012). Environmental factors that cause changes
in chromosomes without altering the deoxyribonucleic acid (DNA) sequence by
altering gene transcription are called epigenetics. Chronic opioid exposure may
stimulate DNA methylation at the mu-opioid receptor coding OPRM1 gene. This
finding did not induce increased opioid dosing requirements, but was associated
with increased chronic pain. The mechanism is still unresolved, but may provide a
new interesting approach to research on OIH (Doehring et al., 2013).
In conclusion, the need for opioid dose elevation during chronic
administration in a clinical setting may be a sum of the pharmacologic opioid
tolerance, non-associative tolerance, OIH, alterations in pharmacokinetics, and
disease progression. Genetic factors also have to be considered.
27
Review of the literature
2.3.2 Role of the glutamatergic system in opioid tolerance and opioid-induced
hyperalgesia
Adaptation to opioid use can also result from the upregulation of glutamate and
other pathways in the brain, which can exert an opioid-opposing effect and
thereby reduce the effects of opioid drugs by altering downstream pathways,
regardless of mu-opioid receptor activation. Extensive research has centered on the
excitatory amino acid system, in particular the NMDA receptors. These are
ionotropic glutamate receptors that are crucial for excitatory synaptic transmission.
They are involved in numerous cognitive processes such as learning and synaptic
plasticity. In the resting state, the receptor is blocked by Mg2+. Activated NMDA
receptors are permeable to Ca2+, Na+, and K+ (Frohlich and Van Horn, 2014).
Acute pain is mainly signaled by the release of glutamate from the central
terminals of primary afferents, which after binding to alpha-amino-3-hydroxy-5methyl-4-isoxazolepropionic acid (AMPA) and kainate receptors causes transient
excitatory postsynaptic currents (EPSCs) in second-order dorsal horn neurons.
The sum of these EPSCs may lead to an action potential and transmission to
high-order neurons. During the constant barrage of action potentials from the
periphery, the increased neurotransmitter release and postsynaptic depolarization
will be long enough to activate NMDA receptors (Frohlich and Van Horn, 2014).
Functional NMDA receptors (Fig. 5) are heteromeric complexes consisting
of a NR1 subunit and modulatory subunits of the NR2 (NR2A–D) family
(Frohlich and Van Horn, 2014). NMDA receptors are involved in synaptic longterm modulation (long-term potentiation) and glutamate neurotoxicity when
overactivated. Systemic administration of the NMDA receptor antagonists MK801 or ketamine has been shown to reverse opioid-induced mechanical or thermal
hyperalgesia and opioid tolerance after acute or chronic opioid administration in
numerous studies (Li et al., 2001b; Manning et al., 1996; Mao et al., 2002;
Trujillo and Akil, 1991; 1994); for a review, see Angst and Clark (2006). In many
of the studies, a role for the intracellular effector PKC has been suggested to lie
behind the induction of OIH, as PKC may upregulate NMDA receptors, and the
development of OIH was not observed in PKC-gamma knockout mice (Célérier
et al., 2004; Zeitz et al., 2001). Interestingly, increased nociception and
hyperalgesia after opioid infusion was also observed in opioid receptor triple
knockout mice lacking all three genes encoding opioid receptors (Juni et al., 2007),
demonstrating a opioid receptor-independent route of hyperalgesia development.
Furthermore, the development of hyperalgesia could be reversed by the NMDA
28
Review of the literature
antagonist MK-801 in control (wild-type) mice, but not in opioid receptor
knockout mice, indicating that other mechanisms also contribute to the
development of hyperalgesia (see 2.3.3). In another triple knockout study using
acute fentanyl injections, the development of hyperalgesia could be reversed by an
MK-801 co-injection (Waxman 2009).
Polyamine(site(
Glutamate((
recogni/on(site(
K+(
Zn2+(site(
Glycine(site(
Extracellular side
Intracellular side
Mg2+(site(
PCP(site(
Na+(
Figure 5. Schematic
representation of
the N-methyl-Daspartate receptor
complex.
PCP, phencyclidine.
Adapted from
Frohlich and Van
Horn (2014).
Ca2+(
Glial cells may also influence the clearance of neurotransmitters, especially
glutamate, from the synaptic cleft. Chronic morphine induced downregulation
of the astrocyte glutamate transporters glial glutamate transporter 1 (GLT-1)
and glutamate aspartate transporter (GLAST) (Mao et al., 2002; Sung et al.,
2003), leading to increased glutamate and synaptic transmission, morphine
tolerance, and thermal hyperalgesia. Glia also incorporate NMDA receptors that
have a different subunit configuration compared to neurons, making them less
permeable to Ca2+. The function and significance of glial NMDA receptors has
not yet been established; for a review, see Dzamba et al. (2013).
Finally, excess glutamate may also play a role in functional changes and
the expression of drug transporters at the BBB. Glutamate was shown to
upregulate P-gp expression in microvessel endothelial cells of the rat brain (Zhu
and Liu, 2004), a phenomenon that was attenuated by the NMDA receptor
antagonist MK-801, implying an NMDA receptor-mediated mechanism behind
the finding. A simplified representation of the glumatergic system in opioid
tolerance is presented in Figure 6.
29
Review of the literature
1.
Glia%
EAATs&
NMDAR&
(?)&
MOR&
MOR&
Primary%afferent%neuron%
Blood6brain%barrier%
2.
Glu&
Glu&
NMDAR&
Postsynap2c%neuron%
Mo&
Mo&
Mo&
Mo&
Efflux&proteins&
Mo&
Glia%
NMDAR&
(?)&
EAATs&"&
MOR"&
MOR&"&
Primary%afferent%neuron%
Blood)brain%barrier%
Glu&
Glu&!&
NMDAR&!&
Glu&
Postsynap3c%neuron%
Glu&
Glu&
Opioid&tolerance&
Opioid=induced&hyperalgesia&
Increased&pain&
Mo&
Efflux&transport&!&
Mo&Mo&
Mo&
Figure 6. A simplified representation of the glutamatergic system in opioid tolerance at the
spinal cord level.
1. Morphine administration in acute pain.
2. Chronic pain, opioid tolerance and opioid-induced hyperalgesia.
EAAT, excitatory amino acid transporter; Glu, glutamate; Mo, morphine; MOR, mu-opioid
receptor; NMDAR, N-methyl-D-aspartate receptor.
é, increased concentration or activity; ê, decreased concentration or activity.
30
Review of the literature
2.3.3 Glial cells and opioid antinociception
Recent years have seen the introduction the neuroimmune interface as a new
concept of modulating CNS neurotransmission. Neuronal excitability can be
notably enhanced by the classical neuronal transmitters, but also by immune
mediators released from glial cells in the CNS as well as infiltrating cells such as
T cells and macrophages. In several painful conditions discussed below, glial
cells function in enhancing pain sensitivity via neuroglial interactions. Opioid
treatment may also cause glial activation, which may attenuate the analgesic
potency of opioids.
Glial cells of the CNS consist of three distinct types of cells: astrocytes,
microglia, and oligodendrocytes. In the peripheral nervous system (PNS), glial
cells are found as satellite glial cells (SGCs) in the dorsal root ganglia (DRGs)
and trigeminal ganglia (TGs), and as Schwann cells enclosing the peripheral
nerves (Gao and Ji, 2010; Hutchinson et al., 2011).
The most numerous glia in the CNS are called astrocytes. They have an
essential role in the regulation of the chemical content of the extracellular space.
Astrocytes form a complex network, an envelope, over the synaptic junctions in
the brain to restrict the spread of neurotransmitters. They also communicate
intercellularly via Ca2+ signaling. Astrocytes are typically immunostained with
the detection of glial fibrillary acidic protein (GFAP), although this only stains
the major branches of the cells. Astrocytes are a part of the “tripartite synapse”
theory, in which glial cells have a crucial role as active modulators of synaptic
transmission. Glia may react to neuronal activity and either enhance or suppress
neuronal activity and the synaptic strength with chemical transmitters (Gao and
Ji, 2010; Hutchinson et al., 2011; Ji et al., 2013).
Oligodendroglial cells provide layers of membrane that insulate the
neuronal axons. One oligodendroglial cell can provide myelin for several axons
(Hutchinson et al., 2011).
Microglial cells are responsible for the primary immune response in the
brain. These cells originate from bone marrow-derived monocytes that migrate
to the CNS during the prenatal period. Their distribution in the CNS is
heterogeneous, and they account for 5–12% of its cells. Under normal conditions,
microglia are considered inactivated, but they continuously sense their
environment with their ramified processes (Nimmerjahn et al., 2005). Under any
kind of brain damage or injury, microglia are activated and undergo
morphological and functional changes. The most selective markers for microglia
31
Review of the literature
are intracellular proteins CD11b (cluster of differentiation molecule 11b) and
ionized calcium-binding adapter molecule 1 (Iba1, also known as allograft
inflammatory factor 1, AIF1) (Fig. 7). In peripheral nerve injury, the spinal
microglia become activated and their morphology changes from a ramified to an
ameboid shape (Suter et al., 2007).
Figure 7. Iba1
immunohistochemical
staining of microglial cells
in the brain cortex of the
rat. 200 × magnification.
Iba1, ionized calciumbinding adapter
molecule 1. (Lilius et al.,
unpublished)
Glia exert different types of functional changes upon neuronal injury or
continuous painful stimulation. Many studies have defined glial activation as
upregulation of glia-selective markers. This is often associated with
morphological changes such as cell hypertrophy. Upregulation of these markers
has been demonstrated in numerous experimental pain states, for example nerve
injury, spinal cord injury, paw incision, joint arthritis, cancer, chemotherapy, and
diabetes models; for reviews, see Grace et al. (2014) and Ji et al. (2013). Chronic
morphine exposure has also been shown to result in considerable upregulation of
the microglial markers Iba1 and CD11b (Daulhac et al., 2006; Wen et al., 2009;
Zhou et al., 2010; Zhu et al., 2014), as well as the astrocyte marker GFAP
(Song and Zhao, 2001) in the spinal cord.
Another indicator of glial cell activation is the upregulation of intracellular
signaling cascade proteins such as the MAPK family. These pathways have an
important role in the intracellular signaling of glia. The enhanced function of
MAPKs by phosphorylation is essential for the development of chronic pain (Ji
et al., 2009), and phosphorylation of p38, a MAPK family member, has been
observed during chronic opioid treatment (Cui et al., 2006). Thirdly, many
receptors and transporters having a role in glial intracellular signaling are
regulated in painful states. At present, there is evidence for the upregulation of
32
Review of the literature
purinergic P2X4 and P2X7 ion channel receptors in spinal microglia (Horvath
and DeLeo, 2009; Zhou et al., 2010).
Toll-like receptors (TLRs) are important regulators of innate immunity
(Nicotra et al., 2012) by sensing various xenobiotic substances, for example
bacterial parts (lipopolysaccharide, LPS), as well as drugs such as opioids.
Activation of the TLR4-MD2 (lymphocyte antigen 96) receptor complex has
been shown to cause activation of both microglia and astrocytes. Morphine was
suggested to induce glial responses via the TLR4-MD2 complex (Wang et al.,
2012b; Watkins et al., 2009), and the blockade of TLR4-MD2 inhibited the
development of opioid tolerance, withdrawal and hyperalgesia. Interestingly,
glial activation via TLR4 was also observed when mu-opioid receptor inactive
(+) opioid isomers were administered (Hutchinson et al., 2010a). In two studies,
however, morphine-induced hyperalgesia was not altered in mice with Tlr4 gene
knockout (Ferrini et al., 2013), and morphine tolerance developed similarly as in
wild-type animals (Fukagawa et al., 2012).
The effects of opioids on glial cell activation have been reported in
multiple studies, with Song and Zhao being the first to report the phenomenon
(Song and Zhao, 2001). Particularly in the spinal cord, glial activation after
chronic intrathecal and/or systemic morphine administration has been
demonstrated by multiple methods, such as upregulation of the astroglial marker
GFAP (Cui et al., 2006; 2008; Huang et al., 2012; Raghavendra et al., 2002;
2004; Song and Zhao, 2001; Zhou et al., 2010) and the microglia-selective
markers Cd11b and Iba1 (Cui et al., 2006; Ferrini et al., 2013; Horvath et al.,
2010; Liu et al., 2006; Raghavendra et al., 2002; Shen et al., 2011; Tai et al.,
2009; 2006; Zhou et al., 2010). After morphine, a smaller number of studies
have shown GFAP and/or Iba1 upregulation in multiple brain areas, such as the
ventral tegmental area, nucleus accumbens, and frontal cortex (Garrido et al.,
2005), the posterior cingulate cortex and hippocampus (Song and Zhao, 2001),
PAG (Eidson and Murphy, 2013; Hutchinson et al., 2009), and midbrain
(Harada et al., 2013). In several studies, glial activation has been indirectly
observed by upregulation of glial cell membrane receptors such as P2X4
(Horvath et al., 2010), P2X7 (Zhou et al., 2010), and upregulation of the
intracellular p38 protein (Cui et al., 2006; 2008; Liu et al., 2006).
Glial cells affect neural pain processing via soluble mediators that may be
released for months after an injury or insult as a result of pathological activation
(Milligan et al., 2006). Compared with classical neurotransmitters, glia-secreted
immune modulators may modulate spinal cord neurotransmission at much lower
33
Review of the literature
concentrations (Ji et al., 2013). Many immune modulators such as tumor
necrosis factor (TNF), interleukin (IL)-1 beta, and interferon gamma (IFNgamma) are capable of directly affecting excitatory transmission by increasing
glutamate release from primary afferent neurons (Gao et al., 2009; Kawasaki et
al., 2008; Nishio et al., 2013; Yan and Weng, 2013; Zhang et al., 2010).
Another mechanism for increased synaptic transmission may be increased
surface AMPA receptor expression or phosphorylation of the NMDA receptor
(Meng et al., 2013; Stellwagen et al., 2005; Viviani et al., 2003; Zhang et al.,
2008). Increased release of proinflammatory cytokines by glia after morphine has
been documented (Shen et al., 2011; Tai et al., 2009; 2006). However, the
activation of spinal cord glia has been much more extensively studied compared
with different supraspinal areas. Furthermore, all of these studies have been
conducted with morphine as the study opioid.
The efficacy of several glia-affecting drugs has been investigated in various
models of opioid antinociception. Minocycline, a broad-spectrum tetracycline
antibiotic and glial inhibitor, enhanced opioid-induced acute antinociception
and attenuated morphine tolerance (Cui et al., 2008; Hutchinson et al., 2008a;
2008b). Interestingly, it also reduced opioid-induced respiratory depression
(Hutchinson et al., 2008b). Ibudilast (AV411), a nonspecific phosphodiesterase
(PDE) inhibitor, enhanced opioid antinociception both acutely and in models of
opioid tolerance (Hutchinson et al., 2009; Ledeboer et al., 2006), as well as
reducing spontaneous and precipitated opioid withdrawal periods (Hutchinson
et al., 2009). The xanthine derivatives propentofylline and pentoxifylline, both
inhibitors of PDE and antagonists of adenosine receptors, have shown promise
in opioid antinociception in neuropathic pain and opioid tolerance (Mika, 2008;
Mika et al., 2007; Raghavendra et al., 2004). For a more extensive review, see
Watkins et al. (2009).
2.3.4 Opioids and the alpha-2-adrenergic receptors
Catecholamine receptors are G protein-coupled receptors divided into two
categories, alpha- and beta-adrenoceptors. Alpha-adrenoceptors are further
divided to subtypes 1A, 1B, 1D, 2A, 2B, and 2C. Alpha-2-adrenoceptors
decrease intracellular adenylyl cyclase activity through Gi, or they may also
modify the activity of ion channels (Na+/H+ antiport, Ca2+ or K+ channels).
Adrenoceptors located in catecholaminergic neurons are called autoreceptors,
and those located in non-adrenergic neurons are called heteroceptors. The
autoreceptors may be located at presynaptic axon terminals, where they inhibit
34
Review of the literature
the release of the neurotransmitter, whereas in the somatodendritic area they
inhibit impulse discharge (Westfall and Westfall, 2011).
In the periphery, the main source of catecholamines is the postganglionic
sympathetic nerve fibers, mainly releasing norepinephrine, and the adrenal
medulla, mainly releasing epinephrine. In the DRG, the subtype alpha-2C is the
most common, followed by alpha-2A receptors, whereas alpha-2B is rare (Shi et
al., 2000). Whether the peripheral alpha-2-adrenoceptor system promotes pain
facilitation or suppression strongly depends on the pain model and time course. In
physiological conditions, peripheral norepinephrine has only minor effects on pain,
but in conditions of inflammation and neuropathy it may aggravate pain (Davis et
al., 1991; Drummond, 1995; Torebjörk et al., 1995). Peripheral nerve sprouting
and the formation of new adrenoceptors may follow inflammation and injury,
which may have a role in maintaining hyperalgesia (Baron, 2000). It is not exactly
known which subtypes of adrenoceptors in the periphery contribute to the
aggravation and which to the inhibition of pain.
In the CNS, the spinal cord and its dorsal horn is critical for the ascending
nociceptive pathways. It receives noradrenergic innervation from the modulatory
descending noradrenergic pathways. The source for the spinal norepinephrine is
located in axons descending from noradrenergic nuclei in the pons and brainstem
(Jones, 1991; Kwiat and Basbaum, 1992). These nuclei, in turn, receive important
projections from PAG (Bajic and Proudfit, 1999). The descending noradrenergic
fibers have sparse axon-axon contacts with central terminals of the primary
afferent fibers (Hagihira et al., 1990); however, contacts with the descending
noradrenergic axons and the spinal dorsal horn neurons are abundant (Doyle and
Maxwell, 1991). It is also possible that the noradrenergic influence on spinal
dorsal horn neurons is mainly non-synaptic, i.e. the transmitter acts by diffusion
from the release site to the site of action (Rajaofetra et al., 1992; Zoli and Agnati,
1996). The alpha-2A-adrenoceptor is localized in the whole dorsal horn (Shi et
al., 1999), especially on the primary afferent nociceptor nerve terminals (Stone et
al., 1998). The alpha-2C receptors are probably located on the axon terminals of
spinal excitatory interneurons (Olave and Maxwell, 2003), whereas the expression
of the 2B-type receptor is only minute in postnatal animals.
In supraspinal areas, alpha-2-adrenoceptors are widely distributed. The
subtype 2A is common throughout the brain, whereas subtypes 2B and 2C have
specific localizations (2B in the thalamus and 2C prominently in the striatum)
(Scheinin et al., 1994).
35
Review of the literature
The most important mechanism for alpha-2-adrenoceptor-mediated pain
regulation occurs in the spinal dorsal horn. Pain signaling is attenuated by
reducing the release of excitatory amino acids from the primary afferent nerve
fibers (Kawasaki et al., 2003; Pan et al., 2002), but also by inducing
hyperpolarization by activating K+ currents in spinal pain-relay neurons (Sonohata
et al., 2004). It is likely that the 2A subtype of the alpha-adrenoceptors mediates
both of these effects, as alpha-2-agonists failed to induce antinociception after the
knockout of alpha-2A-adrenoceptors (Lakhlani et al., 1997; Stone et al., 1997).
The activity of the spinal noradrenergic inhibitory pathways seems to be rather
inactive in animals without sustained pain, because the knockout of any of the
receptor subtypes had no effect on stimulus-evoked pain. In chronic pain, however,
the alpha-2A receptor subtype in particular seems to be involved in noradrenergic
feedback inhibition (Mansikka et al., 2004). At supraspinal sites, despite the fact
that every major structure responsible for processing pain signaling has
noradrenergic receptors, their regulatory role is not yet well established. The
effects of alpha-2-antagonists have been contradictory, depending on the pain
model and supraspinal area that has been under research.
The interplay between the alpha-2-adrenergic system and the opioidergic
system has been under extensive research. Spinal administration of morphine or
other opioids produces marked analgesia, which is mediated by opioid receptors
located both pre- and postsynaptically in the terminals of the nociceptive primary
afferent neurons and the dorsal horn neurons, respectively. During prolonged
administration of either opioids or alpha-2-adrenoceptor agonists, pharmacologic
tolerance to the drug action is observed (Milne et al., 1985; Quartilho et al., 2004;
Reddy and Yaksh, 1980; Takano and Yaksh, 1992). Numerous in vivo studies
investigating the functional interaction between opioid receptors and alpha-2adrenoceptors have shown that animals tolerant to intrathecal morphine also show
tolerance to alpha-2-agonist antinociception (Milne et al., 1985; Stevens et al.,
1988), and opioid receptor antagonists can antagonize the antinociceptive action
of spinal norepinephrine or other alpha-2-adrenoceptor agonists (Roerig et al.,
1992; Sullivan et al., 1992). On the contrary, adrenoceptor antagonists, including
the alpha-2-adrenoceptor antagonist yohimbine, have been shown to antagonize
opioid-induced antinociception (Bentley et al., 1983; Browning et al., 1982).
Furthermore, Aley and Levine (1997) showed that antagonists for mu-opioid
receptors and alpha-2-adrenoceptors antagonized the effect of agonists activating
the other receptor.
36
Review of the literature
However, some recent findings concerning the interplay between the alpha-2receptor and mu-opioid receptor render the interpretation of the supposed drug
effects more complex. Milne et al. (2008) reported that ultralow doses (an ultralow dose is defined as a dose several log units lower than the dose required to
produce functional antagonism at the respective receptor) of four structurally
different
intrathecally
administered
alpha-2-adrenoceptor
antagonists
(atipamezole, yohimbine, mirtazapine, and idazoxan) increased acute spinal
morphine antinociception, attenuated the induction of acute and chronic tolerance,
and also reversed the already established morphine tolerance in rats. The same
group has also demonstrated an augmentation in clonidine and norepinephrine
antinociception after coadministration of atipamezole (Milne et al., 2011).
Moreover, ultralow doses of systemic or intrathecal naltrexone, an opioid receptor
antagonist, administered systemically or intrathecally had a paradoxical additive
effect on the antinociceptive effect of morphine in heat nociception and also
blocked the development of tolerance (Powell et al., 2002). This has additionally
been shown in a model of neuropathic pain in rats (Largent-Milnes et al., 2008).
Similar effects were noted with ultralow doses of selective mu-, delta- and kappaopioid receptor antagonists (Abul-Husn et al., 2007; McNaull et al., 2007).
G protein-coupled receptors normally act as monomeric cell surface
receptors, but there is increasing evidence that G protein-coupled receptors may
also form homo- and heterodimeric complexes. Interestingly, mu-opioid and
alpha-2A receptors may form heterodimers at the plasma membrane of cultured
cells as well as in primary spinal cord neurons (Jordan et al., 2003). The activation
of either monomeric mu-opioid or alpha-2A-adrenoceptors normally mediates a
parallel signaling effect, leading to antinociception at the spinal level. However,
the simultaneous activation of both the mu-opioid- and alpha-2A-adrenoceptors
in heterodimer complexes changes the cellular response, which is not an additive
effect (Jordan et al., 2003). Activation of either mu-opioid or alpha-2A receptors
leads to increased signaling, whereas activation of both receptors decreases the
response. Moreover, an in vitro study (Vilardaga et al., 2008) has shown that in
the heterodimer complexes, the mu-opioid and alpha-2A receptors communicate
with each other through a cross-conformational switch that allows the direct
inhibition of one receptor by the other. Binding of morphine to the mu-opioid
receptor triggers a conformational change in the norepinephrine-occupied alpha2A-adrenoceptor that inhibits its signaling. Norepinephrine has been shown to
inhibit morphine-mediated G protein activation, suggesting that a conformational
change may also take place in the opposite direction (Jordan et al., 2003).
37
Review of the literature
Endogenous norepinephrine could inhibit the effect of morphine in heterodimers
through these mechanisms. Supporting the alpha-2A and opioid receptorinterplay related nociceptive changes, the antinociceptive effect of morphine and
especially tramadol was increased in alpha-2A-adrenoceptor knockout mice
(Ozdoğan et al., 2006).
Another theory for the mechanism behind the paradoxical augmentation of
opioid agonists by ultralow doses of opioid antagonists has been postulated. Wang
et al. (2008) showed that cotreatment with 10 ng/kg naloxone prevented the
chronic morphine-induced Gi/Go-to-Gs switch in mu-opioid receptor G protein
coupling. While opioid receptors in a normal state preferentially bind to Gi or Go
proteins to inhibit adenylyl cyclase (Laugwitz et al., 1993), chronic morphine
treatment may induce a switch to Gs coupling (Chakrabarti et al., 2005; Wang et
al., 2008), leading to cAMP accumulation. A subgroup of mu-opioid receptors
preferentially binding to Gs proteins has also been proposed, but the naloxone dose
of 10 ng/kg has been estimated to occupy only 1% of mu-opioid receptors, which
is not sufficient for antagonizing the supposed Gs-binding subpopulation of mu
receptors (Lewanowitsch and Irvine, 2003). Moreover, an ultra-low dose of
naloxone reduced Gs coupling and restored levels of coupling to native Gi or Go
proteins (Wang et al., 2005). In further studies by the same group, naloxone was
found to bind with high affinity to filamin A, a 300-kDa protein coimmunoprecipitating with mu-opioid receptors (Wang et al., 2008). Filamin A is
known as a large cytoplasmic protein responsible for regulating cell signaling by
interacting with various receptors and signaling molecules, including the muopioid receptor (Onoprishvili et al., 2003). The ultimate hypothesis is that
naloxone binding to filamin A prevents morphine tolerance and hyperalgesia by
inhibiting mu receptor coupling to excitatory Gs proteins. Interestingly, filamin A
may also inhibit the signaling cascades of the TLR4-MD2 receptor complex
(Wang et al., 2012a). Whether ultralow antagonists of alpha-2-adrenergic
receptors also have affinity for filamin A or other closely mu-opioid receptor
related proteins remains to be shown. Table 2 presents the current evidence for the
paradoxical enhancement of opioid antinociception by either opioid receptor or
alpha-2-adrenoceptor antagonists.
In clinical studies, ultralow-dose naloxone combined with morphine in
patient controlled analgesia (PCA) led to decreased nausea and pruritus, but did
not affect analgesia or opioid requirements (Cepeda et al., 2004). When the
administered naloxone concentration was ten times greater (6 µg/ml with an initial
setting of 0.5 ml bolus per demand), the combination led to poorer pain relief than
38
Review of the literature
morphine alone (Cepeda et al., 2002). Chindalore et al. (2005) reported that
Oxytrex, a preparation that combines oxycodone with ultralow-dose naltrexone,
enhanced and prolonged analgesia compared with oxycodone alone in patients
suffering from osteoarthritis. Thus, it seems that in patients the choice of the
ultralow dose may far more difficult than in laboratory animals.
Interestingly, the alpha-2-adrenoceptor is also expressed in glial cells (Hösli
et al., 1982), but its role in glial reactivity has not been thoroughly researched and
has produced some contradictory results. For example, Yan et al. (2011)
demonstrated that dexmedetomidine, an alpha-2-adrenoceptor agonist, stimulated
astrocyte activation and glial cell line-derived neurotrophic factor (GDNF)
production in vitro. In vivo, the alpha-2-adrenoceptor antagonist yohimbine (at a
dose large enough to antagonize morphine antinociception) was shown to reduce
GFAP upregulation induced by chronic morphine administration (Alonso et al.,
2007; Garrido et al., 2005).
39
(Table 2 continued)
Drug studied
POSITIVE FINDINGS
Nonselective opioid
receptor antagonist
naloxone
Pain model
Intrathecal morphine tolerance: thermal
nociception
Remifentanil infusion-induced hyperalgesia
during sevoflurane anesthesia
Intrathecal PTX-induced thermal
hyperalgesia
Nonselective opioid
receptor antagonist
naltrexone
Acute thermal and mechanical nociception
Intrathecal or systemic morphine tolerance:
acute thermal and mechanical nociception
Intrathecal morphine tolerance: acute
thermal nociception
Intrathecal or oral oxycodone
antinociception in spinal nerve ligationinduced hyperalgesia
Outcome
Route, subject
Reference
Inhibition of development of intrathecal morphine tolerance,
inhibition of Gs coupling and Gβγ signaling to adenylyl cyclase,
increasing Gi/o coupling
Inhibition of 1) development of intrathecal morphine tolerance, 2)
glutamate transporter downregulation, 3) NMDAR NR1 subunit
phosphorylation 4) PKC-gamma expression, 5) glial cell activation,
6) excitatory amino acid release
Reduction of needed minimum alveolar concentration (MAC) of
sevoflurane
Preservation of morphine antinociception, suppression of PTXinduced excitatory amino acid release, inhibition of microglial
activation, suppression of cytokine release
Prolongation of intrathecal and augmentation of systemic
morphine antinociception
Inhibition of development of intrathecal morphine tolerance;
restoration of morphine antinociceptive potency in developed
tolerance
Inhibition of development of intrathecal morphine tolerance;
restoration of morphine antinociceptive potency in developed
tolerance
Inhibition of development of i.t. morphine tolerance, attenuation
of microglial and astrocyte activation
Reduction of allodynia and hyperalgesia after SNL and oxycodone
treatment, attenuation of Gs coupling
i.t., rat
Wang et al.,
2005
i.t. infusion, rat
Lin et al.,
2010
i.v. infusion, rat
Aguado et
al., 2013
Tsai et al.,
2008
i.t., rat
i.t. and s.c., rat
Powell et
al., 2002
i.t., female and
male SD and LE
rats
i.t., rat
Terner et
al., 2006
i.t., rat
Mattioli et
al., 2010
LargentMilnes et
al., 2008
Table 2. Current evidence for the paradoxical enhancement of opioid- or alpha-2-adrenoceptor agonist antinociception by
coadministration of ultralow doses (a dose several log units lower than the dose required to produce functional antagonism at the
respective receptor) of either opioid receptor or alpha-2-adrenoceptor antagonists. CTOP, D-Phe- Cys-Tyr-D-Trp-Arg-Thr-Pen-Thr-NH2;
CTAP, D-Phe-Cys-Tyr-D-Trp-Orn-Thr-Pen-Thr-NH2; PTX, pertussis toxin; SD, Sprague-Dawley, LE, Long Evans.
Drug studied
POSITIVE FINDINGS
Alpha-2-adrenoceptor antagonists
atipamezole, yohimbine,
mirtazapine, and idazoxan
Pain model
Outcome
Route, subject
Reference
Acute thermal and mechanical nociception
Intrathecal morphine tolerance: acute
thermal and mechanical nociception
i.t., rat
Milne et al.,
2008
Alpha-2-adrenoceptor antagonist
atipamezole
Acute thermal and mechanical
antinociception
Prolongation of intrathecal morphine antinociception
Inhibition of development of intrathecal morphine
tolerance; restoration of morphine antinociceptive
potency in developed tolerance
Prolongation of intrathecal clonidine (tail-flick and hot
plate)
and
norepinephrine
(tail-flick
only)
antinociception
Inhibition of development of intrathecal norepinephrine
antinociceptive tolerance
i.t., rat
Milne et al.,
2011
Inhibition of development of intrathecal morphine
tolerance
i.t., rat
Milne et al.,
2013
Alpha-2-adrenoceptor antagonist
efaroxan, (+) and (±) isomers but
not the (-) isomer
Alpha-2-adrenoceptor antagonists
efaroxan, atipamezole, and
yohimbine
Intrathecal norepinephrine antinociceptive
tolerance: acute thermal and mechanical
antinociception
Intrathecal morphine tolerance: acute
thermal and mechanical nociception
Low-dose intrathecal morphine thermal
hyperalgesia
Inhibition of acute morphine thermal hyperalgesia
(Table 2 continued)
Drug studied
POSITIVE FINDINGS
Mu-opioid receptor antagonists
CTOP and CTAP and delta-opioid
receptor antagonist naltrindole
Mu-opioid receptor antagonist
CTAP, delta-opioid receptor
antagonist naltrindole, kappaopioid receptor antagonist norbinaltorphimine and nonselective
opioid receptor antagonist
naltrexone
Pain model
Outcome
Route, subject
Reference
Acute thermal and mechanical
nociception
Intrathecal morphine tolerance:
acute thermal and mechanical
nociception
Acute thermal and mechanical
nociception
Intrathecal morphine tolerance:
acute thermal and mechanical
nociception
Low-dose systemic and intrathecal
morphine thermal hyperalgesia
Prolongation of intrathecal morphine antinociception
i.t., rat
Abul-Husn
et al., 2007
Inhibition of development of intrathecal morphine tolerance;
restoration of morphine antinociceptive potency in developed
tolerance
Prolongation of intrathecal morphine antinociception
i.t., rat
McNaull et
al., 2007
Inhibition of development of intrathecal morphine tolerance
Inhibition of acute thermal hyperalgesia
i.t. and s.c., rat
Review of the literature
2.3.5 Access of morphine and other opioids to the central nervous system
Clinically, one interesting developing line of research is the study of opioid
pharmacokinetics and drug–drug interactions for the improvement of opioid
efficacy. The CYP450 family enzymes only have an effect on the metabolism of
opioids, whereas several drug transporter proteins may affect the absorption,
distribution, and excretion of opioid drugs (Mercer and Coop, 2011). Recent
preclinical evidence suggests that the BBB efflux transporters may contribute to
opioid efficacy and the development of tolerance. P-gp, a member of the ATPbinding cassette (ABC) superfamily, is one of the best characterized efflux
transporters to date. It has functions in the extrusion of xenobiotics and uptake
of nutrients (Hennessy and Spiers, 2007). In the brain, the expression of P-gp is
high in the frontal cortex, choroid plexus, and PAG, with lower levels in the
cerebellum and hippocampus (King et al., 2001).
A role of P-gp in the transport of opioids was first observed by Callaghan
and Riordan in an in vitro system (Callaghan and Riordan, 1993). Since then,
multiple in vivo and in vitro studies have been conducted with different
experimental approaches to confirm this finding. In Caco-2 and L-MDR1 cells,
morphine was acknowledged as a weak P-gp substrate with a efflux:influx ratio
of 1.5 (Crowe, 2002; Wandel et al., 2002), a ratio that was supported by in situ
studies (P-gp effect of 1.24) (Dagenais et al., 2004). In P-gp knockout mice,
morphine antinociception was increased and prolonged (Thompson et al., 2000).
Human evidence is sparse and it only consists of acute studies in healthy
volunteer subjects. In the presence of the P-gp inhibitor quinidine, enhanced
CNS effects were not experienced after an i.v. morphine dose (Skarke et al.,
2003), whereas increased plasma concentrations but no influence on morphine
pharmacodynamics after morphine p.o. administration were observed in another
study (Kharasch et al., 2003). Finally, pharmacokinetic modeling for morphine
has been employed, resulting in three major factors limiting the access of
morphine to the brain: limited passive diffusion, active efflux, and low-capacity
active uptake (Groenendaal et al., 2007). Interestingly, a variety of the P-gp gene
ABCB1 polymorphisms have been discovered, and variability in pain relief in
cancer patients was noted with the 3435C>T variant (Sai et al., 2006).
Much less information exists on the interaction of other opioids and P-gp.
For the commonly used oxycodone, there are conflicting results (Boström and
Simonsson, 2005; Hassan et al., 2007). Interestingly, a correlation for P-gp and
the CYP3A substrate co-specificity exists (Schuetz et al., 1996; Wacher et al.,
43
Review of the literature
1995), and oxycodone is an example of this phenomenon (Hassan et al., 2007;
Lalovic et al., 2004). Loperamide, a peripherally restricted opioid often used as
an anti-diarrheal drug, has been identified as a good P-gp substrate with an
efflux:influx ratio of 10 (Wandel et al., 2002). In P-gp knockout mice, its
concentration was increased 13-fold in the brain and doubled in plasma
(Schinkel et al., 1995), suggesting that P-gp is a key factor in constraining
loperamide to the periphery.
A number of studies have implicated the upregulation of the BBB P-gp in
morphine tolerance (Aquilante et al., 1999; Hassan et al., 2009; King et al.,
2001; Mercer and Coop, 2011; Yousif et al., 2012). Aquilante et al. (1999)
reported that the whole brain P-gp content was doubled in morphine-tolerant
rats. Oxycodone was also shown to induce P-gp up to 4-fold in the liver and 1.3fold in brain tissues (Hassan et al., 2007), but in a transcriptome study, the P-gp
(Abcb1) mRNA levels were not increased (Hassan et al., 2009). Furthermore,
King et al. (2001) reported that P-gp antisense or mismatch treatment
eliminated morphine tolerance. Interestingly, in all these studies, measurements
were conducted at the withdrawal phase, pointed out by Yousif et al. (Yousif et
al., 2012). In their study, the expression of P-gp and breast cancer resistance
protein (Bcrp) after chronic morphine administration was not directly
upregulated after the last morphine dose, but only after 9 hours of the last dose.
They hypothesized that P-gp upregulation is not induced by chronic morphine
administration itself, but by the following withdrawal period. Furthermore, the
expression of P-gp and Bcrp was clearly downregulated by the administration of
MK-801 (NMDA receptor antagonist) or meloxicam (cyclooxygenase-2
inhibitor) after the last morphine dose.
Only a few studies have reported other transporters than P-gp for
morphine. Bcrp was not found to transport morphine, even though chronic
morphine induced it (Tournier et al., 2010; Yousif et al., 2012). Like P-gp,
multidrug resistance protein (Mrp) is a member of the ABC superfamily of
transport proteins. In the brain, Mrp is expressed in the cerebellum, PAG, and
hippocampus, and at lower levels in the frontal cortex and choroid plexus (Su
and Pasternak, 2013). In the same study, morphine displayed increased
antinociception after Mrp antisense administration. Dose–response studies have
revealed a twofold-enhanced potency of morphine after Mrp downregulation.
Furthermore, in the Mrp antisense group, morphine antinociception in repeated
intracerebroventricular (i.c.v.) administration developed significantly more slowly,
indicating slower transport from the cerebrospinal fluid to the blood. Rao et al.
44
Review of the literature
(1999) utilized a choroid plexus (blood–cerebrospinal fluid barrier) model to
suggest that Mrp and P-gp are working in series, with P-gp located subapically
and Mrp situated basally. This might explain why downregulation of either
transporter resulted in a similar modulation of opioid analgesia. Finally,
Tzvetkov et al. (2013) showed that morphine uptake was fourfold increased in
OCT1-overexpressing HEK293 cells and morphine uptake in human
hepatocytes was reduced by 1-methyl-4-phenylpyridinium (MPP), an OCT
inhibitor. Furthermore, the mean morphine AUC in OCT1 loss-of-function
polymorphism carriers was significantly higher.
As noted in chapter 2.2.4, morphine is mainly metabolized in phase II
metabolism to water-soluble glucuronide conjugates M3G and M6G in the liver
by UGT2B7 in humans and Ugt2b1 in rats. Interestingly, human brain
homogenates have also been shown to metabolize morphine to its glucuronides,
and the M3G:M6G ratio produced has been associated with the morphine
concentration, so that higher morphine concentrations lead to a higher
M3G:M6G ratio (Yamada et al., 2003). M6G is an attractive analgesic, but its
BBB permeability is very poor, with a permeability that is 7.5 times less than
that of morphine and equivalent to that of sucrose (Yoshimura et al., 1973).
Bourasset et al. (2003) found that the brain uptake of M6G was not increased in
P-gp- or Mrp1-deficient mice, but suggested M6G to be a substrate of the
organic anion transporter (Oatp) 2, as the brain M6G concentration was
decreased when Oatp2 substrates digoxin and PSC833 were co-perfused. Oatp2
is known to exist on both the luminal and basolateral sides of the endothelium of
the brain vasculature. Moreover, co-perfusion of D-glucose with M6G also led
to decreased brain M6G, indicating that these compounds may use the same
transporter, glucose transporter (GLUT) 1, expressed on the both sides of brain
endothelial cells and used for facilitated diffusion of glucose to the brain.
Interestingly, the binding of the opioid peptide enkephalin with β-D-glucose by
an O-β-linkage has also been shown to enhance their brain access via the
glucose transporter GLUT1 (Polt et al., 1994). Even less information exists on
the transport of M3G, the main glucuronide metabolite formed in the SpragueDawley rat. In microdialysis studies, Xie et al. (2000) demonstrated that
probenecid, a typical Oatp inhibitor, increased the brain:blood ratio of M3G.
Current knowledge of the possible BBB transporters for morphine and its
metabolites is presented in Figure 8.
45
Review of the literature
Brain parenchyma
Mrp
BBB
endothelium
Oct1
Oatp2
MO
P-gp
M6G
Oatp2
Probenecid-sensitive
transporter*
Glut1
M3G
Glut1
Probenecid-sensitive
transporter*
Bloodstream
Figure 8. Current knowledge of the possible transporters associated with morphine,
M6G, and M3G transport at the BBB. Morphine-associated transporters in black, M6Gassociated transporters in blue, and M3G-associated transporters in red. M3G,
morphine-3-glucuronide; M6G, morphine-6-glucuronide; BBB, blood-brain barrier;
Glut, glucose transporter; Mrp, multidrug resistance protein; Oatp, organ anionic
transporter polypeptide; Oct, organic cation transporter; P-gp, P-glycoprotein.
* The exact transporter and its location (basal or apical side of endothelium) are not
known.
2.4 Drugs used in the study
2.4.1 Atipamezole, an alpha-2-adrenoceptor antagonist
Atipamezole (4-[2-ethyl-2,3-dihydro-1H-inden-2-yl]-1H-imidazole) (Fig. 9) is
an alpha-2-adrenoceptor antagonist (Scheinin et al., 1988; Virtanen, 1989). Its
binding affinity for alpha-2 receptors (Ki = 1.6 nM) is approximately 100 times
higher than for alpha-1 receptors (Haapalinna et al., 1997; Virtanen, 1989). It
has affinity for alpha-2A-, alpha-2B-, and alpha-2C-adrenoceptor subtypes, and
also for the alpha-2D human receptor variant (Haapalinna et al., 1997;
Renouard et al., 1994). It has no affinity for opioid (mu and delta) receptors.
However, in an in vivo situation, the increase in norepinephrine release in the
46
Review of the literature
CNS following the inhibition of alpha-2-adrenoceptors may result in the
indirect activation of other systems (Haapalinna et al., 2000).
Figure 9. Structural formula of atipamezole.
In rodents, atipamezole is well tolerated. It is rapidly absorbed and distributed,
and peak concentration levels in tissues are two- to threefold higher than the
peak plasma concentration. Atipamezole undergoes first-pass metabolism and its
elimination half-life is 1.3 hours. The LD50 for mice and rats is >30 mg/kg after
i.p., s.c. or i.v. administration (Haapalinna et al., 1997; 1999). In human phase I
studies, atipamezole was also well tolerated after a single dose (10–100 mg).
After the largest dose, some subjective drug effects (e.g. motor restlessness,
coldness, sweating, increased salivation) were reported. It also increased the
plasma norepinephrine concentration and both the systolic and diastolic blood
pressure. At the dose of 30 mg/subject, no subjective or cardiovascular side
effects were reported (Karhuvaara et al., 1990).
The majority of the brain noradrenergic neuron somas are based in the
brainstem and medulla, and alpha-2-antagonists have extensive effects on the
modulation of brain function. The locus coeruleus, the principal site for
noradrenergic neurons, is affiliated with attention, memory, and learning
(Berridge et al., 1993). Atipamezole, by stimulating endogenous norepinephrine
release (presynaptic modulation), has potential effects on these parameters. The
effects on cognitive performance have varied depending on atipamezole dose,
stress experienced during the test, and the age of the animal. Low doses
improved alertness, attention, planning, and recall of animals. At higher doses,
the overactivation of the noradrenergic system has led to impairment in
cognitive tasks; for a review, see Pertovaara et al. (2005). Atipamezole treatment
has also shown promise in recovery after traumatic brain injury (Pitkänen et al.,
2004) and enhanced the positive effects and reduced the adverse effects of
dopaminergic drugs used for the treatment of Parkinson’s disease (Haapalinna et
al., 2003).
47
Review of the literature
As the noradrenergic system has a role in modulating pain in both the brain and
spinal cord (see 2.3.4), atipamezole has also been studied in various pain models.
Supporting the hypothesis that the activity in the spinal descending
noradrenergic pain regulatory pathways is low in non-painful conditions,
atipamezole has not been shown to have effects in baseline nociceptive latencies
(Pertovaara, 1993). In sustained nociception (formalin and capsaicin models),
atipamezole administration antagonized the spinal descending noradrenergic
pathways and increased the nociceptive responses (Green et al., 1998; Mansikka
et al., 2004). At the brainstem level, however, microinjection of atipamezole
produced antinociception in a neurogenic inflammation model (Mansikka and
Idänpään-Heikkilä, 1996). Therefore, atipamezole may have two opposing
mechanism of action regarding pain in different areas of the CNS.
Interestingly, a recent study reported that intrathecal atipamezole (and
other chemically distinct alpha-2-adrenergic antagonists) administered in
ultralow doses increased the effect of morphine in thermal and mechanical
nociception tests. The synergistic effect of atipamezole on morphine
antinociception was seen at a dose (0.08 ng i.t.) lower than needed for blocking
the spinal clonidine antinociception (5–10 µg i.t.). Atipamezole also inhibited
the development of morphine tolerance and reversed the analgesic tolerance that
had already developed (Milne et al., 2008; 2013). Clinically, alpha-2-agonists
exert sedative and anesthetic properties, having similar effects to general
anesthetics such as isoflurane. Currently, alpha-2-adrenoceptor agonists (e.g.
medetomidine and dexmedetomidine) are in clinical use in both human and
veterinary medicine. Atipamezole is registered for clinical use in veterinary
medicine in several countries. It has proven effective in reversing the anesthesia
induced by alpha-2-adrenoceptor activation by medetomidine. The animals are
awake within minutes of administration of the antagonist, which is useful in
minor veterinary surgical operations (Ewing et al., 1993; Jalanka, 1989; Vainio
and Vähä-Vahe, 1990).
2.4.2 Ibudilast, a phosphodiesterase inhibitor
Ibudilast (3-isobutyryl-2-isopropylpyrazolo-[1,5-a]pyridine) (Fig. 10), also
named AV411, MN-166, or KC-404, is used in Asia to treat bronchial asthma
(Kawasaki et al., 1992), cerebrovascular disorders, and post-stroke dizziness
(Armstead et al., 1988; Fukuyama et al., 1993; Shinohara and Kusunoki, 2002).
These applications are based on the effect of ibudilast on tracheal smooth muscle
contractility (Souness et al., 1994), cerebral blood flow improvement (Fukuyama
48
Review of the literature
et al., 1993), and platelet aggregation (Kishi et al., 2000; Ohashi et al., 1986;
Rile et al., 2001). The effects of ibudilast for its clinical implications are
presumed to be related to PDE inhibition, which is key to the regulation of
important second messengers, i.e. cAMP and cyclic GMP (cGMP). The
inhibitory profile of ibudilast on human phosphodiesterases includes PDEs 3A,
4, 10 and 11, with a lesser inhibition on other families (Gibson et al., 2006). In
addition, ibudilast is a weak adenosine receptor antagonist, but other target
proteins have not been identified (Ledeboer et al., 2006; 2007).
Figure 10. Structural formula of ibudilast.
In addition, ibudilast has anti-inflammatory properties in the CNS. In cultured
glial cells, it suppressed LPS-induced production of inflammation markers TNF,
IL-1 and IL-6, nitric oxide (NO) and reactive oxygen species concentrationdependently, and increased the production of the anti-inflammatory cytokine
IL-10 and neurotrophic factors GDNF and nerve growth factor (NGF)
(Mizuno et al., 2004). It also decreased LPS-induced neuronal cell death in
mouse neuron-glia co-cultures and glutamate-induced cell death in hippocampal
neurons. The exact target protein for this effect of ibudilast on glial cells is not
known. Inhibition of phosphodiesterases may additionally contribute to the antiinflammatory effects of ibudilast in the CNS (Sebastiani et al., 2006; Suzumura
et al., 1999). Another potential mechanism of action is the inhibition of the
proinflammatory macrophage migration inhibitory factor (MIF) (Cho et al.,
2010b). MIF is a proinflammatory cytokine itself (Alexander et al., 2012), but it
is required for both IL-1 and TNF release and IL-1- and TNF-induced MAPK
activation (Toh et al., 2006).
Concerning opioid treatment, ibudilast was shown to attenuate morphineinduced glial proinflammatory responses in rats (Hutchinson et al., 2009).
Ibudilast also augmented the antinociceptive action of morphine and oxycodone
in the tail-flick test without having noticeable effects of its own (Hutchinson et
49
Review of the literature
al., 2009; Ledeboer et al., 2007). It attenuated chronic constriction injury (CCI)
induced allodynia and was suggested to affect the development of opioid
tolerance in CCI rats (Ledeboer et al., 2006). It has also been studied for the
treatment of opioid dependency; for a review, see Coller and Hutchinson (2012).
For a summary of findings on ibudilast in pain models, see Table 3.
In phase 1 human studies, ibudilast was shown to be well tolerated as a
single 30-mg oral dose, as well as in repeated administration (30 mg p.o. b.i.d.).
Adverse events were minor. The mean elimination half-life was 19 h (Rolan et
al., 2008). It is currently in clinical trials for the treatment of migraine,
medication overuse headache, metamphetamine dependence, multiple sclerosis,
and opioid withdrawal (http://www.clinicaltrials.gov, searched on May 1, 2014).
50
Table 3. Studies on ibudilast in different in vivo pain models. q.d., once daily; b.i.d., twice daily.
Research model
POSITIVE FINDINGS
Outcome
Dosing, route
Other notes
Reference
CCI and SNL models, rat
Alleviation of
allodynia
Ibudilast is well distributed in the CNS
Ibudilast tolerability was rather good with only transient
changes in the neurobehavioral tests
Ledeboer et al.,
2006
Paclitaxel-induced allodynia, rat
Alleviation of
allodynia
Attenuation of
morphine tolerance
Augmentation of
opioid effect
Decrease in
withdrawal behavior
Decreased
dopamine release
10 mg/kg i.p. b.i.d.;
22 mg/kg p.o. b.i.d.;
25 μg i.t. single dose
7.5 mg/kg i.p. b.i.d.
Plasma morphine levels were not altered by ibudilast
Hutchinson et al.,
2009
Ibudilast also reduced withdrawal symptoms
Bland et al., 2009
Morphine tolerance in CCI, rat
Acute morphine and oxycodone
antinociception, rat
Precipitated and spontaneous
opioid withdrawal periods, rat
Morphine-induced dopamine
release in nucleus accumbens,
rat
Spinal cord injury, rat
CCI model, rat
T13/L1 level dorsal root avulsion
7.5 mg/kg i.p. b.i.d.
7.5 mg/kg i.p.
7.5 mg/kg i.p. b.i.d. along
morphine treatment
7.5 mg/kg i.p. b.i.d. along
morphine treatment
Decreased
hyperalgesia
10 mg/kg i.p.
Decreased central
allodynia
10 mg/kg s.c. q.d.
No effect
7.5 mg/kg p.o.
Hama et al., 2012
Sciatic nerve TNF-alpha ê
Ellis et al., 2014
NEGATIVE FINDINGS
CCI model, rat
Low oral bioavailability
51
Ledeboer et al.,
2006
Review of the literature
2.4.3 Spironolactone, a mineralocorticoid receptor antagonist
Spironolactone
(7α-acetylthio-3-oxo-17α-pregn-4-ene-21,17-carbolactone)
(Fig. 11) is an old mineralocorticoid receptor antagonist with indications for the
treatment of diseases associated with primary or secondary hyperaldosteronism
(Doggrell and Brown, 2001). As it is a potassium-sparing diuretic, it can also be
used to alleviate heart failure or to reduce ascites formation. Many patients
diagnosed with acute heart failure or liver metastases are co-treated with
spironolactone and morphine. For the management of treatment-resistant
hypertension, the daily spironolactone doses are usually small, ranging from 12.5
to 50 mg/day, but patients suffering from ascites due to cancer may be prescribed
high doses of up to 400 mg/day (Ginès et al., 2004). In the rat, spironolactone
has been reported to potentiate the cataleptic effect of morphine (Chu et al.,
1978). Possible mechanisms include the inhibition of efflux transporters leading
to greater morphine disposition into the brain or the inhibition of morphine
metabolism.
Figure 11. Structural formula of spironolactone.
Spironolactone has been shown to inhibit the production of TNF, IL-1, and
several other proinflammatory cytokines in peripheral blood mononuclear cells.
Whether this action is mediated by mineralocorticoid receptors is unknown.
Spironolactone also shows promise in patients suffering from autoinflammatory
diseases such as rheumatoid arthritis (Bendtzen et al., 2003; Hansen et al., 2004;
Miura et al., 2006; Syngle et al., 2009). In an observational case series,
spironolactone improved fibromyalgia-associated pain, stiffness, fatigue,
depression, and mood (Wernze and Herdegen, 2014). Interestingly,
spironolactone reversed corticosterone- and aldosterone-induced microglial
52
Review of the literature
activation in vitro (Tanaka et al., 1997). In a cerebral stroke model (middle
cerebral artery occlusion), the knockout of mineralocorticoid receptors led to
reduced microglial activation and infarct volume (Frieler et al., 2011).
Spironolactone also showed efficacy in the CCI neuropathic pain model (Jaggi
and Singh, 2010) in peroral administration, and also when intrathecally
administered in the chronic compression of the DRG model (Sun et al., 2012).
In the same study, spironolactone decreased the expression and phosphorylation
of the NMDA receptor, an important player in opioid tolerance and
hyperalgesia. Interestingly, eplerenone, another more selective mineralocorticoid
receptor antagonist, also reduced mechanical hypersensitivity in an inflamed
DRG model (Dong et al., 2012), indicating that the mineralocorticoid receptors
may play a role in the modulation of pain, probably via the inhibition of
proinflammatory cytokine production (Chantong et al., 2012). The preclinical
studies on spironolactone in different pain models are listed in Table 4.
Spironolactone inhibits human microsomal UGT2B7 (Knights et al.,
2010), and could thus also inhibit morphine glucuronidation. Even though there
is no direct evidence for inhibition of the rat homologue Ugt2b1 by
spironolactone, the genes encoding the two enzymes have a high gene sequence
similarity (Tukey and Strassburg, 2000). In rodents, the disposition of morphine
in the CNS is regulated by P-gp (Letrent et al., 1999; Schinkel et al., 1995; Xie
et al., 1999), an efflux protein at the BBB. Inhibition of P-gp could lead to
increased access of morphine to the CNS. Spironolactone has been reported to
inhibit the transcellular transport of the P-gp substrate digoxin in P-gpoverexpressing LLC-GA5-COL150 cells (Nakamura et al., 2001) and to
increase the plasma concentration of the P-gp substrate digoxin in humans
(Hedman et al., 1992), indicating a potential interaction with P-gp substrate
drugs.
Taken together, spironolactone possesses three properties that make it
interesting for combining with morphine in the treatment of pain. First, it may
suppress the pathological activation of glial cells via an unknown mechanism;
second, it may inhibit the formation of the pronociceptive morphine metabolite
M3G; and third, it may increase the access of morphine to the brain and delay
its elimination by the inhibition of P-gp.
53
Table 4. Studies on spironolactone in different in vivo pain models. CCI, chronic constriction injury; CCD, chronic compression of the
dorsal root ganglion; DRG, dorsal root ganglion.
Pain model
Outcome
POSITIVE FINDINGS
Streptozotocin
Attenuation of hyperalgesia
diabetes model,
mouse
Dosing, route
CCI, rat
10 and 20 mg/kg
p.o. for 14 days
3 μg i.t. b.i.d.
CCD, rat
CCD, rat
Attenuated CCI-related pain
behavior
Attenuated CCD-related pain
behavior
Attenuated CCD-related pain
behavior
NEGATIVE FINDINGS
CCI, rat
No effect on hyperalgesia or
allodynia
Spinal morphine
No effect on development of
tolerance, rat
morphine tolerance
Vincristine
neuropathy, rat
No alleviation of hyperalgesia
or allodynia
Other notes
7 or 15 mg/kg p.o.
Khan et al., 2009
Sciatic nerve TNF-alpha ê
Synergistic effects with glucocorticoid receptor
antagonists
Spinal microglial reactivity ê
Concentrations of IL-1b and TNF-alpha in the spinal
cord and DRGs ê
NMDA receptor expression and phosphorylation ê
3 μg i.t. b.i.d.
Reference
Jaggi and Singh,
2010
Gu et al., 2011
Sun et al., 2012
3 μg i.t. b.i.d.
Wang et al., 2004
Co-treatment 3 μg
i.t. b.i.d.
Lim et al., 2005
10 and 20 mg/kg
p.o. for 14 days
Jaggi and Singh,
2010
54
Review of the literature
2.4.4 Ketamine, an N-methyl-D-aspartate receptor antagonist
Ketamine [2-(2-chlorophenyl)-2-(methylamino)-cyclohexanone] (Fig. 12) is a
phencyclidine (PCP) derivative available as a racemic mixture and also as an S(+)
enantiomer in Europe. It is highly lipophilic and easily crosses the BBB (White,
1988). It is a general anesthetic that produces a state of dissociative anesthesia at
high doses and analgesia at low doses (Sinner and Graf, 2008). The Senantiomer is a 2–4 times more potent anesthetic and analgesic than the Renantiomer (Arendt-Nielsen et al., 1996; Oye et al., 1992; White et al., 1980).
The actions of ketamine are mainly based on noncompetitive antagonism of the
NMDA receptor by binding to the PCP binding site (Fig. 5). Ketamine reduces
the calcium influx and during a longer period modulates long-term potentiation,
the wind-up phenomenon, and glutamate-mediated neurotoxicity. Ketamine has
other possible interactions with sodium (Frenkel and Urban, 1992) and calcium
channels (Sikand et al., 1995), the potassium/sodium hyperpolarizationactivated cyclic nucleotide-gated (HCN) 1 channel (Chen et al., 2009), as well
as the nicotinic (Scheller et al., 1996), muscarinic (Hustveit et al., 1995), and
opioid receptors (Finck and Ngai, 1982; Hirota et al., 1999; Hustveit et al.,
1995; Smith et al., 1980).
Figure 12. Structural formula of ketamine.
In humans, the elimination half-life of ketamine is two to three hours after oral
administration and approximately 80% of the drug undergoes N-demethylation
to norketamine by CYP3A4 and CYP2B6 enzymes (Hijazi, 2002; Sinner and
Graf, 2008; Yanagihara et al., 2001). The bioavailability of ketamine is 17–24%
in humans after oral administration (Chong et al., 2009; Clements et al., 1982).
Norketamine and other metabolites are mainly excreted renally, and only small
55
Review of the literature
amounts of ketamine are excreted unchanged (Geisslinger et al., 1993).
Norketamine is an active metabolite of ketamine (Ebert et al., 1997; Holtman et
al., 2008; Shimoyama et al., 1999). In a binding study, Ebert et al. (1997)
reported that S-norketamine is an approximately five times weaker NMDA
receptor antagonist than S-ketamine. Racemic norketamine was found
antinociceptive in the rat (Holtman et al., 2008; Shimoyama et al., 1999) and it
potentiated morphine antinociception in thermal, peripheral neuropathy, and
tonic inflammatory pain models (Holtman et al., 2008).
At low doses, ketamine has been proven efficacious as an analgesic
perioperatively or in chronic pain states (Bell et al., 2005; Laskowski et al., 2011;
Visser and Schug, 2006). Empirical evidence also indicates that ketamine
coadministered with morphine can significantly increase the efficacy of
morphine in cancer pain management. However, there has been little clinical
research to support this (Bell, 2012). The majority of preclinical and human
experimental studies (Arroyo-Novoa et al., 2011; Carstensen and Møller, 2010;
Honarmand et al., 2012; Suppa et al., 2012) have focused on the
pharmacodynamic properties of ketamine as an NMDA receptor antagonist
leading to decreased opioid tolerance and hyperalgesia.
Some preclinical research regarding the pharmacokinetic interaction
between opioids and ketamine has been conducted. Pretreatment with ketamine
caused cross-tolerance to morphine, but pretreatment with morphine caused an
increase in the cataleptic response to ketamine. This finding was suggested to be
caused by residual morphine in the brain after morphine treatment (Hance et al.,
1989). A similar effect was also observed with the use of sufentanil, fentanyl,
alfentanil, nalbuphine, and butorphanol in combination with ketamine
(Benthuysen et al., 1989). Alfentanil infusion increased the brain concentration
of ketamine and decreased its plasma AUC, indicating a greater distribution
volume. On the other hand, the brain concentration of alfentanil was decreased
in ketamine coadministration (Edwards et al., 2002). Interestingly, acute LPS
treatment lowered plasma ketamine AUC and increased its elimination half-life,
indicating an increase in BBB permeability and increased access of ketamine to
the brain (Veilleux-Lemieux et al., 2012). In repeated administration, morphine
may increase the permeability of the BBB (Sharma and Ali, 2006). This provides
another interesting theory for the possibility of altered ketamine
pharmacokinetics during chronic opioid administration. Regarding drug
transporters, ketamine has not been identified as a substrate of P-gp (Doan,
2002; Tournier et al., 2010) or Bcrp (Tournier et al., 2010), but an older study
56
Review of the literature
(Ullrich et al., 1993) listed ketamine as a candidate substrate for both organic
cation transporter (OCT) and organic anion transporter (OAT) families. In
HEK293 cells transfected with human or rat OCT1, OCT2, and OCT3,
ketamine was found to inhibit radiolabeled 1-methyl-4-phenylpyridinium
uptake (Amphoux et al., 2006), a finding later confirmed by Massmann et al.
(2013) regarding OCT3.
57
3. AIMS OF THE STUDY
Drugs with different mechanisms of action such as ibudilast, atipamezole,
spironolactone, and ketamine have shown promise in preclinical studies of
offering new pain treatment possibilities. This study focused on these drugs by
evaluating them in coadministration with morphine, the gold standard of
opioids. The rat was used as the study subject. Studies were performed with
special emphasis on both pharmacodynamic and pharmacokinetic properties of
the studied drugs.
The specific aims of this thesis study were:
I.
To investigate the antinociceptive and other nociceptionmodulating properties of ibudilast alone and in coadministration
with morphine;
II.
To assess the antinociceptive effects of ultralow doses of the
alpha-2-adrenergic antagonist atipamezole as an adjuvant to
morphine therapy in intrathecal and systemic administration;
III.
To determine the effects of the mineralocorticoid receptor
antagonist spironolactone in acute coadministration with
morphine and in morphine tolerance with special emphasis on
potential pharmacokinetic interactions;
IV.
To assess the potential pharmacokinetic interactions between
morphine and the NMDA receptor antagonist ketamine in acute
and chronic morphine administration, with a focus on possible
changes in brain drug and metabolite concentrations.
58
4. MATERIALS AND METHODS
4.1 Ethical considerations
The research procedures were conducted in accordance with the guidelines of
the local authorities and the IASP (Zimmermann, 1983). The provincial
government of Southern Finland had approved the study concept (Uudenmaan
lääninhallitus, Hämeenlinna, Finland). For ethical reasons, the smallest possible
number of animals and the smallest number of nociceptive tests per animal were
used. In all nociceptive behavioral models, the animals were able to terminate
the noxious stimulation. Cut-off latencies were used to avoid tissue damage.
4.2 Materials
4.2.1 Animals
Male Sprague-Dawley rats (Harlan, Horst, Netherlands, and Scanbur,
Sollentuna, Sweden) weighing 180–300 g at the beginning of experiments) were
used. Rats were housed in plastic cages in light- and temperature-controlled
rooms (artificial 12 h/12 h light–dark cycle, temperature 23 ± 2 ˚C). Water and
standard laboratory chow were available ad libitum. For the behavioral studies,
the rats were habituated to the testing environment 60–90 min/day for at least
three days. Behavioral measurements were performed in a randomized and
blinded fashion in all studies. After the experiments, the rats were euthanized by
decapitation and brain, liver and serum samples were collected.
4.2.2 Drugs
Morphine hydrochloride, racemic ketamine hydrochloride (Ketaminol vet®,
Intervet, Boxmeer, The Netherlands) and medetomidine hydrochloride
(Domitor, Orion Pharma, Espoo, Finland) were purchased from the University
Pharmacy (Helsinki, Finland). Ibudilast was purchased from APAC
Pharmaceutical, LLC (Columbia, MD, USA), and atipamezole hydrochloride
was provided by Orion Pharma (Espoo, Finland). Spironolactone and
loperamide hydrochloride were purchased from Sigma-Aldrich (St. Louis, MO,
USA).
59
Materials and methods
Morphine and atipamezole were dissolved in physiological saline. Ibudilast was
dissolved in 2% polysorbate 20 (Tween® 20; Fluka Chemika, Buchs,
Switzerland) in physiologic saline. Spironolactone was diluted with 4%
polysorbate 80 (Tween® 80, University Pharmacy, Helsinki, Finland) and
loperamide with 2% polysorbate 80 in physiological saline. Ketamine and
medetomidine were diluted to the administration concentrations with
physiological saline.
The injection volumes were 2 ml/kg subcutaneously and 10 ml/kg (Study
I) or 5 ml/kg (Study III) intraperitoneally. In the intrathecal experiments, drugs
were administered in separate phases using 50-µl Hamilton microsyringes in
injection volumes of 10 µl. Between the drug phases, 1 µl of air was used to
prevent ex vivo mixing of the solutions. Finally, 10 µl of saline was administered
to flush the catheter.
4.3 Methods
4.3.1 Intrathecal cannulation
The catheter implantation (Study II) was performed according to the modified
method of Yaksh and Rudy (Yaksh and Rudy, 1976). Each rat was anesthetized
with subcutaneous (s.c.) 0.4 mg/kg medetomidine (Domitor®, Orion Pharma,
Espoo, Finland) and 60 mg/kg ketamine hydrochloride (Ketaminol vet®,
Intervet, Boxmeer, The Netherlands). The rats were placed prone in a
stereotactic frame, and the membrane of the cisterna magna was surgically
exposed and punctured with care. A polyethylene cannula (8 cm, PE-10) was
inserted through the opening and advanced caudally to the level of the lumbar
enlargement. The cannula was fixed to the paravertebral muscles, and the skin
was closed with 3-0 sutures. The cannula was then flushed with 0.9% saline.
Rats showing any neurological symptoms such as movement of extremities or
muscle twitches indicating potential CNS trauma leading to neurological deficits
were immediately euthanized. Five days after cannulation, the placement of the
cannula was verified by administering 10 µl of 10 mg/ml lidocaine intrathecally
(Lidocain®, Orion Pharma, Espoo, Finland). Only rats with reversible
symmetrical paralysis of both hindlimbs after injection were accepted for testing.
To enable recovery from the cannulation operation and anesthesia, rats were
allowed to rest for at least 8 days before any nociceptive tests.
60
Materials and methods
4.3.2 Morphine tolerance treatment schemes
In all studies, morphine tolerance was induced with a 4-day scheme. In studies I,
II and III, the rats received 2 daily s.c. injections at 10:00 am and 8:00 pm. The
individual doses were 10 mg/kg on day 1, 15 mg/kg on day 2, 20 mg/kg on day 3,
and 30 mg/kg on day 4. In the intrathecal experiments in study II, rats received
intrathecal injections of morphine twice daily for four days. The individual dose
was 15 µg each day. In experiment IV, Alzet minipumps pumping 10 µl/h for 4
days (model 2ML1, Durect, Cupertino, CA, USA) were used to provide a
continuous morphine exposure. Under brief isoflurane anesthesia (induction in a
chamber with 4–5% isoflurane at a flow rate of 1.5 l/min of air and maintenance
with 2% at 1.5 l/min), a small incision was made between the scapulae of the rat
and a subcutaneous pocket was created. The minipumps had been previously
filled with vehicle or morphine hydrochloride to release 6 mg of morphine base
per day. The pumps were installed subcutaneously and the wound was closed
with 3-0 sutures.
4.3.3 Pain measurement in acute nociception
Hot plate tests (Woolfe and Macdonald, 1944) (Studies I–IV) were performed
with a Harvard Apparatus Ltd hotplate apparatus (Edenbridge, Kent, United
Kingdom). In the test, the rats were put onto a circular transparent plastic ring
on the hot plate (52 ± 0.2 °C). Licking or brisk shaking of the hindpaw or
jumping was considered a sign of thermal nociception. Latency to the first
reaction was measured. To avoid tissue damage, the cutoff time was set to 60
seconds.
Tail-flick latencies (D’Amour and Smith, 1941) (Studies I–IV) were
assessed with a Ugo Basile 37360 (Comerio, Italy) tail-flick apparatus. In the
test, the rats were kept in hard plastic tubes covered with a dark cloth. After
accustomization, an infrared beam (radiant heat) was directed in turn to three
different points on the middle third of the tail of the test subject. The intensity
was adjusted to produce a baseline latency of approximately 3.5 s. A flick of the
tail was considered as a sign of thermal nociception and it automatically stopped
the timer of the apparatus. The measurements were repeated three times at each
time point, and the mean of the values was used as the result. To avoid tissue
damage, the cutoff was set at 10 s. If an individual measurement reached the
cutoff time, no further tests were performed for the particular time point.
61
Materials and methods
The paw pressure test (Study II) (Randall and Selitto, 1957) was performed with
the Ugo Basile paw pressure device (Comerio, Italy). Rats were gently wrapped
in a towel during the test where the left hindpaw was placed under a pivot and
the force applied to the paw was gradually increased. A brisk withdrawal of the
hind paw terminated the measurement. A cutoff of 500 g was used.
At each time point, the tail-flick test was performed first followed by the
hot plate and/or paw pressure tests. A short waiting period was used between
each test. The predrug (baseline) latencies were measured separately for each
experiment day immediately before the administration of any drugs.
4.3.4 Assessment of motor function
A rotarod apparatus (Palmer electric recording drum; United Kingdom;
diameter, 80 mm; speed, 27 rpm (Study I) and 12 rpm (Study IV)) was used to
assess the actions of the test drugs on motor coordination. A rat was placed on
the rotating rod, and the time the rat stayed there was measured. Training on
the rod was allowed on three consecutive days. On the fourth day, animals that
remained for at least 60 s on the rotating rod before drug administration were
accepted to the test, and 60 s was also used as a cutoff time in the test proper.
Possible effects of the drugs on spontaneous locomotor activity (Study I)
were tested in a measurement box (70 × 70 × 35 cm; Kungsbacka Regler- &
Mätteknik, Kungsbacka, Sweden) isolated from sound and light. Photocells were
located at two different levels (2 and 12 cm) above the floor of the box to
automatically detect movements of the animal. A 30-min measurement period
was started 15 min after the drug injections.
4.3.5 Drug concentration measurements
The drug concentrations of morphine, M3G, M6G, normorphine (Study III),
and ketamine and norketamine (Study IV) were determined with LC-MS/MS
(liquid chromatography-tandem mass spectrometry) analyses as previously
described (Zheng et al., 1998) with some modifications. An Agilent 1100 series
high performance liquid chromatography (HPLC) system (Agilent
Technologies, Waldbronn, Germany) coupled to an API 3000 tandem mass
spectrometer (AB Sciex, Toronto, ON, Canada) operating in a positive turbo
ion spray mode was used. The chromatographic separation was achieved on an
Atlantis HILIC Silica column (3 µm particle size, 2.1 × 100 mm I.D.) (Waters,
Milford, MA, USA) using a gradient elution of the mobile phase consisting of
62
Materials and methods
acetonitrile with 10 mmol/l ammonium formate in 0.2% formic acid (v/v) (Study
III) or 20 mmol/l ammonium acetate (pH 3.0, adjusted with formic acid) (Study
IV). An aliquot (7 µl, Study III; 5 µl, Study IV) was injected at a flow rate of
250 µl/min (Study III) or 200 µl/min (Study IV) to give a total chromatographic
run time of 18 min (Study III) or 24 min (Study IV). Oxycodone served as an
internal standard for morphine and its metabolites. Deuterium-labeled internal
standards were used for ketamine and norketamine (Study IV). The following
ion transitions were monitored: morphine, m/z 286 to m/z 152; M3G and M6G,
m/z 462 to m/z 286; normorphine, m/z 272 to m/z 152; and oxycodone, m/z
316 to m/z 241. The limit of quantification was 1.0 ng/ml for morphine, M3G,
and M6G, and 0.5 ng/ml for ketamine and norketamine. A signal-to-noise ratio
(S/N) of 20:1 was used as the limit of detection for normorphine, and the
quantities were given in arbitrary units relative to the ratio of the peak area of
normorphine to that of the internal standard. The day-to-day coefficients of
variation (CV) were below 15% at relevant concentrations for all analytes.
4.3.6 Experimental design of Studies I–IV
The experimental designs of the studies are described in brief in Table 5 and in
detail in the original publications (Studies I–IV, Methods).
63
Materials and methods
Table 5. Designs of studies I-IV in brief.
Question addressed
Study I
• Antinociceptive effects of ibudilast
alone or in combination with
morphine
• Effects of ibudilast coadministration
on the development of morphine
tolerance
• Effects of ibudilast on morphine
withdrawal
• Effects of acute coadministration of
ibudilast and morphine in morphinetolerant animals
• Effects of ibudilast on opioid tolerance
in already developed tolerance
• Effects of ibudilast on motor
coordination or spontaneous
locomotor activity
Study II
• Effects of subcutaneously
administrated low-dose atipamezole
on subcutaneous morphine
antinociception
• Effects of subcutaneously
administrated low-dose atipamezole
on subcutaneous morphine
antinociception in opioid tolerance
• Effects of intrathecally administered
low-dose atipamezole on intrathecal
morphine antinociception
• Effects of intrathecally administered
low-dose atipamezole on intrathecal
morphine antinociception in opioid
tolerance
Experimental design
Methods used
Ibudilast with or without morphine
administered to drug-naïve animals
Tail-flick
Hot plate
Ibudilast co-treatment with the morphine
tolerance scheme for four days; nociceptive
tests on day 5; observations of weight
change during withdrawal
Tail-flick
Hot plate
Morphine tolerance treatment protocol,
drug coadministration and nociceptive
tests on day 5, continuation of drug
coadministration and repeated nociceptive
tests on day 9
Tail-flick
Hot plate
Ibudilast alone and in coadministration
with morphine
Rotarod
Locomotor
boxes
Morphine combined with three
atipamezole doses
Tail-flick
Hot plate
Morphine combined with three
atipamezole doses in opioid-tolerant
animals
Tail-flick
Hot plate
Morphine combined with three
atipamezole doses
Tail-flick
Paw pressure
Morphine combined with three
atipamezole doses in opioid-tolerant
animals
Tail-flick
Paw pressure
64
Materials and methods
(Table 5 continued)
Question addressed
Experimental design
Methods used
Study III
• Effects of systemical spironolactone on
morphine-induced antinociception
• Effects of acute morphine with or
Two doses of morphine
combined with spironolactone
Tail-flick
Hot plate
Four-day morphine tolerance
Tail-flick
without spironolactone in morphine-
scheme with or without
Hot plate
tolerant rats
spironolactone cotreatment
• Effects of repeated spironolactone
administration on morphine-induced
followed by drug
coadministration on day 5
tolerance
• The concentrations of morphine and its
metabolites in the brain, liver, and serum
after morphine and spironolactone
coadministrations
Morphine combined with one or
two doses of spironolactone in
opioid-naïve rats
• Effects of spironolactone on the
antinociceptive properties of
loperamide, a peripherally restricted Pglycoprotein substrate opioid
Loperamide combined with
spironolactone in opioid-naïve
rats
Study IV
• Effects of acute small dose of ketamine
on morphine antinociception and tissue
ketamine, norketamine and morphine
concentrations in opioid-tolerant
animals
• Effects of acute small dose of ketamine
on morphine antinociception and tissue
ketamine, norketamine and morphine
concentrations in opioid-tolerant
animals under withdrawal
• Effects of ketamine and/or morphine
administration on observed
antinociception and drug tissue
concentrations
Tail-flick
Hot plate
High performance
liquid chromatography
Tail-flick
Hot plate
Preceding morphine pump
treatment for four days;
administration of ketamine on
day 5
Tail-flick
Hot plate
Rotarod
High performance
liquid chromatography
Removal of morphine pumps
after day 5, withdrawal for two
days, acute doses of ketamine
and/or morphine on day 7
Tail-flick
Hot plate
Rotarod
High performance
liquid chromatography
Acute tests using small doses of
morphine and/or ketamine
Tail-flick
Hot plate
Rotarod
High performance
liquid chromatography
65
Materials and methods
4.4 Statistical analysis
The hot plate and tail-flick results are expressed as a percentage of the maximum
possible effect (MPE%), calculated as MPE% = [(postdrug latency – baseline
latency)/(cutoff time – baseline latency)] × 100%, which takes into account the
differences in baseline nociceptive latencies. In the text and figures, results are
presented as means of the groups (± SEM). The behavioral data were tested for
statistically significant differences in mean values by two-way analysis of variance
(ANOVA) followed by a Bonferroni correction for multiple comparisons (group
× dose or group × time) (Studies I and II) or followed by a Holm-Sidak
correction for multiple comparisons (group × dose or group × time) (Studies III
and IV). For the concentration data in Studies III and IV, a two-tailed t-test or
one-way analysis of variance followed by the Holm-Sidak correction was used.
For the nonparametric rotarod test data (Study IV), the Kruskal-Wallis test
followed by Dunn’s multiple comparison was used. The difference was
considered significant at P < 0.05 in both the analysis of variance and the post hoc
test. The data were analyzed using GraphPad Prism, versions 4.0c (Study I),
5.0b (Study II), 6.0a (Study III), and 6.0c (Study IV) for Macintosh (GraphPad
Software, Inc., San Diego, CA, USA).
66
5. RESULTS
5.1 Morphine
5.1.1 Effects of morphine in different models of nociception
Morphine was used as the study opioid in all studies. The peak antinociception
was measured at 30 minutes after s.c. administration, when it had an ED50 of 3.1
mg/kg in the tail-flick test and 3.7 mg/kg in the hot plate test (Study III) (Fig.
13). After i.t. administration, 1.5 µg morphine was needed to produce over 50%
MPE. The peak antinociception was measured at 30 min in both the tail-flick
and paw pressure tests (Study II, Fig. 1).
MPE%
100
Tail-flick latency
Hot plate latency
Figure 13. ED50 graphs for morphine in the tailflick and hot plate tests 30 minutes after
subcutaneous administration. n = 8.
50
0
1.25
2.5
5
Morphine dose (mg/kg s.c.)
The tissue concentrations of morphine and their corresponding antinociceptive
responses were determined (Study III). After 4 mg/kg of s.c. morphine,
significant antinociception was observed in the tail-flick test at 30 and 90
minutes after administration and in the hot plate test at 30 min after
administration (Fig. 14A and B). The brain morphine concentrations were 406
± 22 nmol/kg and 258 ± 9 nmol/kg at 30 and 90 minutes, respectively (n.s.). In
serum, the morphine concentration at 90 min was significantly lower than at 30
min (2000 ± 210 nmol/l vs. 410 ± 22 nmol/l, P < 0.001) (Fig. 14C).
67
***
Vehicle
Mo 4 mg/kg s.c.
80
60
***
40
20
0
30
90
Time after treatment (min)
B 100
Hot plate latency (MPE%)
Tail-flick latency (MPE%)
A 100
80
Vehicle
Mo 4 mg/kg s.c.
***
60
40
20
C
2500
Morphine concentration
(nmol/l or nmol/kg)
Results
2000
0
Brain (nmol/l)
Serum (nmol/kg)
###
1500
1000
500
0
30
90
Time after treatment (min)
30
90
Time after treatment (min)
Figure 14. Antinociceptive properties of subcutaneous 4 mg/kg morphine (Mo) in the
tail-flick (A) and the hot plate (B) tests and its corresponding serum and brain
concentrations (C). *** P < 0.001 compared with the vehicle group. ### P < 0.001
between the timepoints. n = 5–10.
5.1.2 Development of morphine tolerance in different models
A 4-day scheme for inducing morphine tolerance was used both in s.c. and i.t.
models. On day 5, morphine-tolerant and morphine-naïve rats received 5 mg/kg
(Studies I and II) and 4 mg/kg (Study III) s.c. morphine as the test dose, which
resulted in significant antinociception in morphine-naïve rats, whereas only
minor (n.s.) antinociception was observed in morphine-pretreated rats (Table 6).
In the i.t. model, the test dose was 1.5 µg (Study II).
In Study IV, osmotic minipumps delivering a steady dose of morphine
were used. After 24 hours of minipump implantation, significant antinociception
was observed. Tolerance to morphine antinociception developed on the fifth day
of pump treatment (Study IV, Fig. 2).
In Studies I–IV, no opioid-induced hyperalgesia (reduction of nociceptive
thresholds below baseline after morphine administration) was observed (Lilius et
al., unpublished finding).
In Study III, brain morphine concentrations were measured 90 min after
an acute 8 mg s.c. dose in morphine tolerant and morphine-naïve rats. No
difference in the mean brain morphine concentration was observed (202 ± 8 ng/g
in tolerant rats vs. 200 ± 10 ng/g in naïve rats, n = 4–8) (Lilius et al.,
unpublished data).
68
Results
Table 6. Antinociceptive effects of a test dose of morphine (Mo) administered after
the induction of morphine tolerance. B.i.d., twice daily; MPE%, percentage of the
maximum possible effect; TF, tail-flick; HP, hot plate; n/a, data not available.
* Statistically significant difference compared with the vehicle-administered group. n
= 7–8.
Study
I
II
III
II
Dosing
s.c. b.i.d.
s.c. b.i.d.
s.c. b.i.d.
i.t. b.i.d.
Test dose
5 mg/kg s.c.
5 mg/kg s.c.
4 mg/kg s.c.
1.5 μg i.t.
TF latency after test
dose (MPE%)
HP latency after test
dose (MPE%)
Mo-naïve
81*
98*
56*
53*
Mo-naïve
71*
83*
34*
n/a
Tolerant
26
34
18
22
Tolerant
7
21
0
n/a
5.2 Ibudilast (Study I)
5.2.1 Antinociceptive effects of ibudilast
Antinociception was assessed with the tail-flick (Fig. 15A) and hot plate (Fig.
15B) tests after the administration of four different doses of ibudilast with or
without morphine to drug-naïve rats. An antinociceptive effect of ibudilast
administered alone at doses of 7.5 and 22.5 mg/kg i.p. was observed. Morphine
2.5 mg/kg s.c. caused significant antinociception in both tests, and ibudilast and
morphine had an additive effect in the tests.
In a separate experiment, at 30 minutes after drug administrations,
ibudilast (22.5 mg/kg i.p.) did not modify the mean brain concentrations of
morphine (4 mg/kg s.c.) (Lilius et al., unpublished data).
69
Results
100
80
§
§
§
# §
§
60
**
*
40
20
0
0
0.83
2.5
7.5
22.5
B
Hot plate latency (MPE%)
Tail-flick latency (MPE%)
A
Ibudilast dose (mg/kg)
#
# §
100
*
80
60
40
*
§
20
0
0
0.83
2.5
7.5
22.5
Ibudilast dose (mg/kg)
Vehicle + ibudilast i.p. or vehicle
Morphine 2.5 mg/kg s.c. + ibudilast i.p. or vehicle
Figure 15. The antinociceptive dose–response curve for ibudilast in the tail-flick (A)
and hot plate (B) tests with or without 2.5 mg/kg s.c. morphine at 30 minutes after
administration. The mean of the maximum possible effect (MPE%) ± SEM
is plotted. For the ibudilast groups: *, ** Statistically significant difference (P < 0.05, P
< 0.01, respectively) compared with the group that was given the vehicle alone. For
the morphine groups: # Statistically significant difference (P < 0.05) compared with
the group that was given morphine without ibudilast. § Statistically significant
difference (P < 0.05) compared with the group that was given the same dose of
ibudilast without morphine. n = 7.
5.2.2 Effects of ibudilast cotreatment on the restoration of morphine
antinociception
Ibudilast was studied in several models of opioid tolerance. The exact treatment
schemes are presented in Study I, Table 1. When ibudilast was acutely
coadministered with morphine to opioid-tolerant rats on day 5, it restored the
antinociceptive effect of morphine in the tail-flick test (Fig. 16). However, in the
hot plate test, ibudilast had antinociceptive effects of the same magnitude in
opioid tolerant rats as in naïve animals (Study I, Fig. 5).
From this point, the cotreatment of morphine and ibudilast twice daily
was continued for four more days and the rats were retested on day 9 after drug
coadministration. Morphine alone produced no significant antinociception, but
ibudilast 7.5 mg/kg i.p. combined with morphine enhanced morphine
antinociception without having antinociceptive effects of its own (Study I, Fig.
6).
70
Results
Tail-flick latency (MPE%)
§
80
#
40
20
Morphine tolerance treatment
Morphine (mg/kg s.c.)
Ibudilast (mg/kg i.p.)
*
60
Pretreatment (days 1–4):
Acute treatment (day 5):
* #
* #
100
0
–
–
+
+
+
+
+
0
0
5
0
5
0
5
2.5
5
7.5
0
2.5
0
7.5
Figure 16. Effects of acute ibudilast treatment on morphine antinociception in
morphine-tolerant rats in the tail-flick test. Morphine tolerant rats were given two
different doses of ibudilast and/or a test dose of morphine. The mean of the
maximum possible effect (MPE%) ± SEM is shown. *, P < 0.05 compared with the
vehicle. #, P < 0.05 compared with the morphine-tolerant, morphine-administered
group (black bar). §, P < 0.05 between selected groups.
In a separate experiment, three different doses of ibudilast (0.83, 2.5 and 7.5
mg/kg) administered to rats twice daily under the four-day morphine tolerance
scheme to assess the potential of ibudilast in inhibiting the development of
morphine antinociceptive tolerance. However, in the hot plate test, pretreatment
with 7.5 mg/kg ibudilast increased the development of tolerance to an acute test
dose of 5 mg/kg morphine (Study I, Fig. 2B).
5.2.3 Effects of ibudilast on motor coordination
As ibudilast showed sedative effects in routine handling of the rats, the effects of
acutely administered ibudilast were investigated in a measurement box
measuring spontaneous locomotor activity in a novel environment. The rats were
placed inside the box 15 min after drug administration for a total of 30 minutes.
At doses of 2.5 and 7.5 mg/kg i.p., compared with vehicle, ibudilast significantly
reduced spontaneous locomotor activity during the 30-min observation period
(Fig. 17A). In a separate experiment, rats first received a 4-day pretreatment of
20 mg/kg i.p. ibudilast twice daily and locomotor experiments were conducted
71
Results
on day 5, demonstrating that no tolerance to the observed effect had developed
(Lilius et al., unpublished data, Fig. 17B).
Vehicle
Ibudilast 2.5 mg/kg i.p.
Ibudilast 7.5 mg/kg i.p.
A 600
400
300
200
100
0
***
***
***
***
0-5
***
***
**
**
***
***
Vehicle
Ibudilast 7.5 mg/kg i.p.
Ibudilast 20 mg/kg i.p. b.i.d.
for 4 days + ibudilast 7.5 mg/kg i.p.
500
Activity counts
Activity counts
500
B 600
5-10 10-15 15-20 20-25 25-30
400
300
200
100
***
0
***
***
0-5
Time (min)
***
*
**
5-10 10-15 15-20 20-25 25-30
Time (min)
Figure 17. Effects of acute ibudilast (A) and ibudilast pretreatment (B) on spontaneous
locomotor activity. The activity counts (number of photocell crossings in 5-min
periods) over time ± SEM are shown 15 min after drug administration. *, **, ***
Statistically significant difference (P < 0.05, P < 0.01, P < 0.001, respectively) compared
with the vehicle group. n = 6.
The effects of ibudilast on motor coordination were also examined using the
rotarod test. On the first treatment day, all rats (100%) that had received vehicle
achieved the cutoff time (60 s) at 30 and 120 min. Ibudilast i.p. in doses of 0.83
and 2.5 mg/kg caused no significant differences (52.3 ± 7.7 and 53.0 ± 7.0 s,
respectively) compared with vehicle. After administration of 7.5 mg/kg ibudilast,
the mean survival time was reduced to 11.2 ± 4.0 s at 30 min and 31.5 ± 12.6 s at
120 min (P < 0.05 compared with vehicle). These measurements were repeated
on day 4 after three preceding days of ibudilast treatment (0.83, 2.5, and 7.5
mg/kg i.p. twice daily), and the same significant results were achieved as on day
1, indicating that tolerance to the observed effect did not develop during four
days of treatment (Study I, Results).
72
Results
5.3 Atipamezole (study II)
5.3.1 Effects of atipamezole coadministration in acute morphine antinociception
The choice of the initial intrathecal morphine dose was based on pilot
experiments, in which 1.5 µg of morphine produced approximately 50 MPE% in
the tail-flick test. The studied ultralow i.t. doses of atipamezole did not produce
antinociception in any of the nociceptive tests on their own.
Morphine was administered in the same syringe in a separate phase with
different doses of atipamezole (0.01, 0.1 and 1 ng) or vehicle (Fig. 18).
Morphine alone produced significant antinociception at 30 (P < 0.05) and 60
min (P < 0.01) after drug administration. Morphine combined with 1 ng
atipamezole produced significantly greater antinociception than the combination
of morphine and vehicle 30 minutes after drug administration in the tail-flick
test (P < 0.05). However, a longer-lasting effect of the drug combination was not
observed. In the paw pressure test, morphine alone produced significant
antinociception from 15 to 120 minutes after administration, but atipamezole
did not augment this effect (Study II, Results). In further studies, intrathecal
morphine doses of 0.5 and 5 µg produced significant antinociception at 30 and
60 min, but the 1-ng dose of intrathecal atipamezole did not have an additive
effect on the observed antinociception (Study II, Results).
73
Results
Veh + Veh
Mo 1.5 μg + Veh
Mo 1.5 μg + At 0.01 ng
Mo 1.5 μg + At 0.1 ng
Mo 1.5 μg + At 1 ng
100
Tail-flick latency (MPE%)
*
80
60
40
20
0
0
30
60
90 120 150 180 210 240
Time after treatment (min)
Figure 18. Effects of intrathecal atipamezole on intrathecal morphine antinociception
in the tail-flick test. Rats were administered three different doses of atipamezole (At)
or vehicle (Veh) with morphine (Mo) 1.5 μg. The mean of the maximum possible effect
(MPE%) ± SEM is plotted as a function of time. * Statistically significant difference (P <
0.05) compared with the group that was administered morphine and vehicle. n = 6–7,
except n = 4 in the vehicle group.
5.3.2 Effects of spinal atipamezole coadministration in morphine tolerance
Spinal morphine tolerance was induced during four days by administering 15 µg
of intrathecal morphine twice daily. On day 5, in the tail-flick test, intrathecal
1.5 µg morphine produced significant antinociception at 30 and 60 minutes (P <
0.001 and P < 0.01, respectively) (Fig. 19), whereas the same dose in morphinepretreated animals did not cause significant antinociception compared with
vehicle. At the intrathecal dose of 10 ng but not 1 ng, atipamezole
coadministered with morphine partly restored morphine antinociception at 30
minutes (P < 0.05).
74
Results
Tail flick latency (MPE%)
100
Vehicle-pretreated Morphine-pretreated
Veh + Veh
Mo 1.5 μg + Veh
Mo 1.5 μg + Veh
Mo 1.5 μg + At 1 ng
Mo 1.5 μg + At 10 ng
80
***
60
**
*
40
20
0
0
30
60
90
120
Time after treatment (min)
Figure 19. Effects of intrathecal atipamezole on intrathecal morphine antinociception
in the tail-flick test in morphine-tolerant rats. Morphine-tolerant rats were
administered two different doses of atipamezole (At) or saline (Sal) with morphine
(Mo) 1.5 μg. As a control, morphine was also administered to previously morphinenaïve animals. The mean of the maximum possible effect (MPE%) ± SEM is plotted as a
function of time. *, **, *** Statistically significant difference (P < 0.05, P <0.01, P <
0.001 respectively) compared with the morphine-tolerant group that received acute
morphine only. n = 6–9.
5.3.3 Effects of systemic administration of atipamezole with morphine
As systemic administration is the preferred route for the ease of drug use, the
effects of systemic atipamezole on systemic acute morphine antinociception and
morphine tolerance were also evaluated. Because in the pilot experiments
atipamezole doses of 30 and 300 µg/kg s.c. induced motor restlessness, the
atipamezole doses chosen were 0.03, 0.3, and 3 µg/kg s.c. Morphine 2.5 mg/kg
s.c. produced significant antinociception in the tail-flick test 30 min after
administration (P < 0.05) and in the hot plate test at 30, 75, and 120 minutes (P
< 0.05 at all time points and P < 0.001 at 75 minutes). In acute tests, however,
none of the coadministered atipamezole doses significantly increased the effect
of morphine in either of the tests (Study II, Results).
75
Results
5.4 Spironolactone (Study III)
5.4.1 Effects of spironolactone in acute morphine antinociception and morphine
tolerance
Spironolactone 50 mg/kg i.p. had no antinociceptive effects of its own, but it
increased the antinociceptive potency of 2.5 and 5.0 mg/kg of s.c. morphine in
both the tail-flick and hot plate tests (Study III, Fig. 1). In morphine-tolerant
rats, the magnitude of increased antinociception observed was at the same level
as in naïve animals (Study III, Fig. 2).
Chronic spironolactone pretreatment alone (50 mg/kg i.p. twice daily for
four days) did not affect antinociception by acute morphine alone. To study the
effects of spironolactone treatment on the development of morphine tolerance,
one group of the morphine tolerance-treated rats was cotreated twice daily with
50 mg/kg i.p. spironolactone, and acute 4 mg/kg s.c. morphine alone was given
on day 5. Antinociception observed in cotreated animals was minimal and did
not differ from rats that had received the morphine tolerance treatment without
spironolactone (Study III, Results).
5.4.2 Effects of spironolactone on morphine and morphine-3-glucuronide
concentrations in the brain
In acute coadministration, spironolactone administered 60 min and 0 min before
morphine 4 mg/kg s.c. significantly increased morphine antinociception at 30
and 90 min (Fig. 20A). In the brain, coadministration increased the mean
morphine concentration by 69% and 290% at 30 and 90 min, respectively (Fig.
20B). At 30 minutes, serum morphine concentrations were essentially at the
same level, but the liver concentrations were increased in the group that was
coadministered two doses of spironolactone (Fig. 20C and D). Continuing to 90
minutes, morphine concentrations in all tissues were elevated in the presence of
spironolactone. Spironolactone did not, however, reduce the formation of M3G
in the liver (Study III, Fig. 3G).
76
Results
***
***
***
###
###
***
*
50
VEH + VEH
MO 4 + VEH
MO 4 + SPR 50
MO 4 + SPR 50 × 2
###
0
MO concentration (ng/g)
B
400 BRAIN
300
###
200
###
##
100
#
C
MO concentration (ng/ml)
30
90
Time after treatment (min)
D
SERUM
800
600
400
## ###
200
0
0
30
90
Time after treatment (min)
MO concentration (ng/g)
Hot plate latency (MPE%)
A100
LIVER
800
600
#
400
200
#
##
0
30
90
Time after treatment (min)
30
90
Time after treatment (min)
Figure 20. Effects of 50 mg/kg i.p. spironolactone treatment 60 min before morphine
(SPR 50) or 60 min before and coadministered with morphine (SPR 50 x 2) in the hot
plate test 30 and 90 min after morphine administration (A). The mean of the
maximum possible effect (MPE%) ± SEM is shown. The corresponding brain (B), serum
(C), and liver morphine concentrations (D) are presented (mean ± SEM). *, ***
Statistically significant difference (P < 0.05, P < 0.001, respectively) compared with the
vehicle-administered group. #, ##, ### Statistically significant difference (P < 0.05, P <
0.01, P <0.001, respectively) between the selected groups. MO, morphine. n = 10 at 30
min and n = 5 at 90 min in behavioral experiments; n = 5 per timepoint in
concentration studies.
5.4.3 Antinociceptive effects of the P-glycoprotein substrate loperamide
combined with spironolactone
To assess the effects of spironolactone on BBB transporter proteins,
spironolactone (50 mg/kg i.p.) was administered 60 minutes before and
concomitantly with a peripheral opioid, P-gp substrate loperamide (10 mg/kg
s.c.). In coadministration, loperamide showed significant antinociceptive
77
Results
properties in the hot plate test 30 and 90 minutes after administration,
indicating potential changes in the drug transporter function of the BBB
induced by spironolactone (Study III, Fig. 4).
5.5 Ketamine (Study IV)
5.5.1 Ketamine coadministration to morphine-tolerant rats leads to increased
brain concentrations of morphine, ketamine, and norketamine
The effects of ketamine in morphine tolerance have been suggested to be
mediated via a pharmacodynamic interaction. In this study, the possibility of a
pharmacokinetic interaction between the drugs was investigated. Subcutaneous
minipumps delivering morphine 6 mg/day induced tolerance (Study IV, Fig. 2),
and on day 5 rats received a small dose of ketamine (10 mg/kg s.c.). In
morphine-naïve rats, ketamine caused no antinociception, whereas in morphinetolerant rats it caused significant antinociception, peaking at 90 minutes (54
MPE% increase, Fig. 21A and B). In the rotarod test, ketamine induced motor
dysfunction at 30 min and the effect was enhanced by the chronic morphine
treatment. At peak antinociception, however, the rotarod performance of all
groups had returned to normal (Fig. 21C).
In morphine-tolerant ketamine-treated rats, the mean morphine,
ketamine and norketamine brain concentrations were 2.1-, 1.4-, and 3.4-fold
higher, respectively, compared with the rats that received the morphine tolerance
treatment or ketamine only (Fig. 19). In the liver of morphine-tolerant
ketamine-treated rats, the mean ketamine concentration was six-fold higher and
the norketamine concentration two-fold higher than in morphine-naïve rats.
Hence, morphine tolerance affects the pharmacokinetic properties of ketamine,
while on the other hand, acute ketamine affects the pharmacokinetic properties
of morphine.
In a separate experiment, the effect of morphine withdrawal on morphine
and ketamine drug concentrations was investigated. The morphine pump
treatment lasted for 5 days, after which the pumps were removed. On day 7, the
rats were administered the same low ketamine dose (10 mg/kg s.c.). As noted on
day 5, during withdrawal ketamine and norketamine concentrations were also
increased, although significantly only in the liver (Study IV, Fig. 4).
78
Results
5.5.2 Effects of combined ketamine and morphine in acute models of nociception
As ketamine administered to morphine-tolerant rats caused notable increases in
the brain concentrations of both ketamine and morphine, it was of interest
whether this interaction would also be observable in an acute experiment.
Morphine (2.5 mg/kg s.c.) was administered 15 minutes before ketamine (10
mg/kg s.c.). In behavioral experiments, the coadministration of morphine and
ketamine caused significant motor coordination impairment compared with
morphine only at 30 min, which was also reflected in the results of the hot plate
test (Study IV, Fig. 5A–C). Morphine coadministration had no effect on
ketamine or norketamine brain concentrations 90 min after ketamine
administration, but ketamine slightly increased the mean morphine
concentration in the brain and liver (Study IV, Fig. 5D–I).
Figure 21 (next page). Effects of an acute small dose of ketamine on antinociception
in morphine-tolerant rats. Rats undergoing s.c. morphine (6 mg/day, Mo) or vehicle
(Veh) pump treatment were administered an acute s.c. dose of ketamine (10 mg/kg,
Ket) or vehicle (Veh) on day 5. Antinociception was measured using tail-flick (A) and
hot plate (B) tests. The mean of the maximum possible effect (MPE%) ± SEM is plotted.
In the rotarod test (C), the mean (± SEM) survival time (seconds) is plotted. From
separate animals with the same pretreatments, whole brain, serum, and liver samples
were collected 90 min after ketamine administration. The concentrations of the
experimental drugs and their main metabolites (MO, morphine; M3G, morphine-3glucuronide; KET, ketamine; NORKET, norketamine) were quantified. The mean tissue
concentrations (± SEM) are plotted in graphs D–I. The increase or decrease in tissue
concentrations between the treatment groups is shown in percentages for each
substance. n = 7–8. **, *** Statistically significant differences (P < 0.01, P < 0.001,
respectively) compared with the vehicle control group. #, ##, ### Statistically
significant differences (P < 0.05, P < 0.01, P < 0.001, respectively) compared with the
morphine-pump-treated group that received acute vehicle. §§, §§§ Statistically
significant differences (P < 0.01, P < 0.001, respectively) compared with the vehiclepretreated group that received acute ketamine.
79
Results
Day 5 (tolerance): Behavioral data
***
###
§§§
60
***
###
§§§
40
20
0
C
80
Time on rotarod (s)
Tail-flick latency (MPE%)
Hot plate latency (MPE%)
B
A 80
60
40
***
###
§§
20
***
###
§§
60
40
**
##
20
0
0 30
90
150
Time after Ket (min)
0 30 90 150 210
Time after Ket (min)
0
30
90
Time after Ket (min)
Concentration data 90 min after ketamine administration
Serum
E 2000
110%
#
60
40
20
–4%
1500
240%
§§§
1000
45%
§§
500
0
0
MO
G2000
F 600
KET
M3G
NORKET
Concentration (ng/mL)
80
Concentration (ng/g)
Concentration (ng/g)
D100
–5%
400
10%
Concentration (ng/mL)
Brain
1500
190%
§§§
1000
200
0
MO
M3G
500
61%
§§§
0
KET
NORKET
Liver
–29%
2000
120%
1000
0
##
MO
M3G
Veh pumps
I 20000
Concentration (ng/g)
Concentration (ng/g)
H 3000
120%
§§§
15000
10000
5000
0
510%
§§§
KET
NORKET
80
Mo pumps
Veh (behavioral data)
Ket 10 mg/kg s.c.
Veh
Ket 10 mg/kg s.c.
6. DISCUSSION
6.1 Methodological considerations
The use of animals for the study of pain is necessary and justified because of the
complex pain and nociception processing systems. Studies on the complex
processing of nociception and pain require a neuronal network with the
supraspinal parts of the CNS. The struggle to alleviate pain and suffering and to
improve existing pain remedies is one of the most important and humane areas
of scientific research. Current animal models used for the study of pain have
proven invaluable when they are used carefully and with attention to supporting
data from in vitro, in silico, and other methods (Mogil et al., 2010). In the future,
when evaluating changes (i.e. synaptic plasticity and potential glial activation) in
the CNS during opioid treatment, the development of noninvasive imaging
techniques such as magnetic resonance imaging may especially help in reducing
the number of animals needed and also enable more versatile studies in humans.
All animal treatment procedures of this study were carefully conducted following
the guidelines of local authorities and the IASP (Zimmermann, 1983). The
suffering of the animals was minimized and the number of animals used was
reduced by using a washout period when applicable.
Of the thermal nociceptive tests used, the hot plate test reflects
supraspinal antinociception, whereas the tail-flick reaction is a spinal reflex. Both
of these are widely used and have been found predictive in opioid research (Le
Bars et al., 2001). For the intrathecal experiments (Study II), the method for
measuring spinal nociception was extended to also cover mechanical nociception
measured by the Randall-Sellitto paw pressure test.
6.2 Ibudilast in coadministration with morphine
Study I demonstrated that ibudilast restored the antinociceptive effect of
morphine in developed morphine tolerance after single and repeated
administration. However, when ibudilast treatment was given together with the
morphine tolerance treatment, the development of tolerance could not be
attenuated.
Previous studies (Hutchinson et al., 2009; Ledeboer et al., 2007) have
demonstrated the ability of ibudilast to augment morphine and oxycodone
81
Discussion
antinociception. However, in these studies, the administration of ibudilast alone
at a dose of 7.5 mg/kg i.p. was reported not to produce any change in hind paw
or tail-flick latencies over a 100-minute time course after administration
(Hutchinson et al., 2009). In Study I, already the 7.5-mg/kg dose of ibudilast
caused a marginal (significant in one experiment) increase in the tail-flick, but a
marked increase in hot plate latencies 30 and 120 min after acute administration.
As the hot plate test involves supraspinal nociceptive processing (Le Bars et al.,
2001), it may be more sensitive to unspecific behavioral effects such as sedation.
Indeed, ibudilast decreased locomotor activity at doses of 2.5 and 7.5 mg/kg and
impaired motor coordination at 7.5 mg/kg. Supporting this, 7.5 mg/kg ibudilast
has previously been reported to cause transient sedation and decreased reactivity
to touch (Ledeboer et al., 2007). Therefore, the observed ibudilast-induced
antinociception is most likely caused by sedation and impaired motor
coordination, especially in the hot plate test. However, the mechanisms by
which ibudilast reduces locomotor activity and impairs motor coordination are
not known. It is unlikely that these effects are mediated through actions on glial
cells, because they are thought not to be reactive in healthy animals. The routine
screening of ibudilast binding to various target proteins did not reveal any target
proteins that could mediate the decreased locomotor activity (Ledeboer et al.,
2007). In humans, however, sedation has not been reported to be a major
problem after acute (30 mg p.o.) or subchronic (30 mg p.o. b.i.d.) administration
regimen (Rolan et al., 2008).
The development of morphine dependence and tolerance has been
classically linked to the cAMP pathway (see page 25). Activation of mu-opioid
receptors leads to the inhibition of adenylate cyclase (Sharma et al., 1975; Wang
et al., 1994). However, repeated administration of morphine may lead to
upregulation of adenylate cyclase, increased intracellular cAMP concentrations,
and activation of PKA. This may contribute to the development of tolerance via
cyclic adenosine monophosphate response element-binding protein (CREB), a
transcription factor (Montminy and Bilezikjian, 1987) that regulates genes
responsible for the development of physical dependence (Lane-Ladd et al.,
1997). Ibudilast, as an inhibitor of PDE, should counteract acute opioid effects
by increasing intracellular cAMP levels. Therefore, it seems likely that the
demonstrated effect of ibudilast is not linked to PDE inhibition in opioid
receptor-containing neurons. In glial cells, however, the effect of PDE
inhibition on the development of opioid tolerance may be the opposite. In an
Alzheimer’s disease model, PDE4B, a cAMP-specific phosphodiesterase, was
82
Discussion
upregulated by the administration of amyloid beta peptide leading to increased
production of proinflammatory TNF (Sebastiani et al., 2006). After spinal cord
injury, the PDE4 B2 isoform was upregulated from 24 h to 1 week after injury,
at peak microglial activation (Ghosh et al., 2012). Thus, in microglial cells, the
inhibition of PDE may lead to beneficial anti-inflammatory effects. Supporting
this, various PDE inhibitors also show anti-inflammatory properties when
treating peripheral diseases such as asthma and chronic obstructive pulmonary
disease (theophylline and roflumilast; for a review, see Gavaldà and Roberts
(2013)). An important observation, however, is that AV1013, an amino analog
of ibudilast, shares many of the same anti-inflammatory properties as ibudilast,
but lacks PDE inhibitory capabilities (Cho et al., 2010a). This suggests that the
tolerance-suppressive effects of ibudilast may also result from alternative
mechanisms, proposedly the inhibition of MIF (Cho et al., 2010a; 2010b) or
antagonism of TLR4 receptors (Hutchinson et al., 2010b).
6.3 Morphine combined with ultralow doses of atipamezole
The main findings of Study II are that atipamezole, an alpha-2-adrenergic
antagonist, had no independent antinociceptive effect at the studied ultralow
doses, but it increased the antinociceptive effect of spinal morphine in the tailflick test in both morphine-naïve and morphine-tolerant rats. The
administration of systemic atipamezole did not alter systemic morphine
antinociception in either morphine-naïve or morphine-tolerant animals.
Intrathecal atipamezole increased the peak antinociceptive potency of
morphine in morphine-naïve animals, but no prolongation of the effect was
observed. Milne et al. (2008) demonstrated that atipamezole decreased the acute
antinociceptive effect of morphine during the first 30 minutes after
administration, but after this period, it dramatically prolonged the
antinociceptive response, resulting in an increase in the AUC of antinociception
over time during 180 minutes after administration. Moreover, in morphinetolerant rats, the same study demonstrated a much more pronounced
augmentation of the morphine antinociceptive effect in comparison with the
results of Study II. Differences in experimental design might at least partly
explain the observed differences in antinociceptive responses. In Study II, the
optimal dose of spinal morphine was 10 times lower (1.5 µg in comparison with
15 µg), while a 10 times higher dose of atipamezole (10 ng) was needed to
achieve a significant augmentation of the morphine effect. Similar variation
83
Discussion
between studies in the optimal dose of ultralow opioid antagonists has been
observed (Mattioli et al., 2010). The rats in our laboratory appeared to be more
sensitive to morphine (or more resistant to the development of acute tolerance)
because the 15-µg morphine dose used by Milne et al. caused a long-lasting
100% MPE (unpublished results).
There were also several minor differences in study design. In the
atipamezole study by Milne et al. (2008) and other reports from the same
laboratory, morphine sulfate was used in contrast to morphine hydrochloride in
Study II. Study II also used slightly longer tail-flick baseline response times and
different anesthetics (ketamine vs. halothane) for the intrathecal cannulation
operation. The intrathecal cannulation procedure itself may induce glial
activation (DeLeo et al., 1997), and as previously noted, ketamine may have
neuroprotective effects leading to decreased glial activation. This could also
explain the better efficacy of morphine in our laboratory. Previously, i.t.
morphine doses of 1 to 5 µg in 10 µl provided antinociception for several hours
(Lemberg et al., 2008), in line with the observed responses in Study II. The best
intrathecal morphine dose for evaluating the possible augmentative effects of
atipamezole was 1.5 µg, resulting in approximately a 50% MPE at 30 min after
administration.
The mechanism by which alpha-2-adrenoceptor antagonists at doses
considerably below those producing blockade of the alpha-2-adrenoceptor
increase opioid antinociception is not known. One proposal is the alpha-2adrenoceptor-mu-opioid receptor heterodimer theory (see 2.3.4). The
simultaneous activation of both the mu-opioid and alpha-2A-receptors in
heterodimer complexes may change the cellular response (Jordan et al., 2003):
activation of either mu-opioid or alpha-2A receptors leads to increased parallel
intracellular signaling, whereas activation of both receptors decreases the
response. An explanation for the paradoxical effect could be that the affinity of
heterodimer-involved alpha-2-adrenoceptors for atipamezole would be greater
than their monomeric counterparts. In the administration of ultralow doses, only
the receptors in heterodimers would be manned, leaving the monomeric
receptors unmanned to bind endogenic norepinephrine. However, this theory
remains to be investigated. Interestingly, ultralow doses of opioid antagonists
have shown similar properties to alpha-2-antagonists. For naloxone, a novel
binding site in the cytoplasmic protein filamin A has been discovered, and it
remains to be seen whether similar novel binding sites for alpha-2-adrenergic
antagonists will be discovered. However, Ozdoğan et al. (2006) demonstrated
84
Discussion
that morphine and especially tramadol analgesia were augmented in alpha-2Aadrenoceptor knockout mice, a finding defending the involvement of the alpha2A receptor binding site in the augmentation of opioid antinociception.
6.4 Spironolactone for the treatment of opioid tolerance
The results of Study III indicate that spironolactone had no antinociceptive effect
of its own in thermal models of nociception, but it dose-dependently increased the
mean brain morphine concentration and corresponding antinociception in both
tail-flick and hot plate tests. Spironolactone did not prevent the development of
morphine tolerance, but in acute coadministration with morphine it restored the
effect of morphine in tolerant rats. The peripheral µ-opioid agonist and P-gp
substrate loperamide showed significant antinociception when combined with
spironolactone. The results suggest that spironolactone increased the mean
morphine brain concentration by inhibiting an efflux transporter, most likely P-gp,
at the BBB.
During coadministration of spironolactone, the brain:serum ratio of
morphine was increased, indicating an increased distribution of morphine into the
brain. Of the known BBB transporters, morphine is a substrate of P-gp (Letrent
et al., 1999; Schinkel et al., 1995; Xie et al., 1999) and Mrp (Su and Pasternak,
2013), transporters that move their substrates from endothelial cells to the blood
(see 2.3.5). Spironolactone is a P-gp inhibitor (Nakamura et al., 2001), and the
augmentation of the mean brain morphine concentration may therefore be due to
the inhibition of P-gp. The concentrations of the substances in the spinal cord
were not measured in the study. However, parallel results were achieved in the
tail-flick and hot plate tests, indicating a probable increase in also the spinal cord
concentrations of morphine. In the elimination phase, the mean serum morphine
concentration was increased after spironolactone co-treatment, indicating a
probable decrease in the elimination of morphine. The higher serum morphine
concentration as well as the possible P-gp inhibition additively increased the brain
morphine disposition, leading to enhanced and prolonged antinociception in
spironolactone-co-treated rats. The P-gp substrate loperamide only had an
antinociceptive effect when coadministered with spironolactone, suggesting that
spironolactone is an effective P-gp inhibitor at the studied doses. In the study,
however, no direct evidence for the inhibition of P-gp was shown.
Morphine is mostly eliminated in the liver by metabolism and to a small
extent by excretion in the kidney (Andersen et al., 2003). It is metabolized to
85
Discussion
M3G by Ugt2b1 and to normorphine by CYP3A4 in humans and CYP3a1 in rats,
and is mainly excreted in its glucuronide forms (Yaksh and Wallace, 2011). In
spironolactone-cotreated animals, the mean M3G concentration in the liver was
not decreased. However, it is difficult to evaluate the inhibitor effects of
spironolactone on the Ugt2b1 enzyme, because spironolactone reduced morphine
elimination, leading to increased morphine concentrations in the liver and thus an
increased amount of substrate for the morphine glucuronidization reaction. In the
presence of spironolactone, the concentration of M3G was vastly increased.
Spironolactone had no effect by itself in acute thermal nociceptive tests. It is
interesting, however, that intrathecal spironolactone alleviated neuropathic pain in
the chronic compression of the dorsal root ganglion (CCD) model in rats (Sun et
al., 2012). The spironolactone-induced antinociceptive effect in the CCD model
was suggested to be mediated through the inhibition of activated glial cells, as the
expression of proinflammatory cytokines and the phosphorylation of NMDA
receptor subunit NR1 was decreased. In Study III, acute pain was measured in
healthy animals in which glial cells are not supposed to be pathologically activated.
Therefore, no effects of spironolactone were expected. In line with this,
spironolactone did not have any effect on its own when the contralateral (healthy)
paws of the animals were investigated (Sun et al., 2012).
We examined whether spironolactone co-treatment with morphine could
attenuate the development of tolerance. This effect could hypothetically have been
mediated through the inhibition of M3G formation or inhibition of the
pathological activation of glial cells. Importantly, 4-day spironolactone cotreatment during the morphine tolerance scheme did not affect the response of an
acute test dose of morphine. Regarding M3G, spironolactone treatment did not
lead to the desired results, as M3G concentrations even increased as a consequence
of decreased morphine elimination. The study did not directly evaluate the
possible activation of glial cells, because the pharmacokinetic interaction between
spironolactone and morphine led to vastly increased brain morphine
concentrations in co-treated animals, rendering the comparison between the
groups unfeasible. Analysis of glial activation would have required alterations in
experiment design, such as the intrathecal administration of morphine or
administration of drugs by turns in order to avoid a pharmacokinetic interaction.
Interestingly, the development of morphine tolerance was not attenuated by
intrathecal spironolactone (Lim et al., 2005), supporting the conclusion that the
augmentation of the antinociceptive effect of morphine also observed in study III
is mainly due to the increased CNS concentrations of morphine. Even though it is
86
Discussion
difficult to directly extrapolate the results of animal studies to patient treatment,
previous studies have demonstrated that inhibition of the efflux transporter P-gp
may contribute to the effects of morphine and other opioids in humans.
Spironolactone may be administered in high doses to patients with liver cirrhosis,
and these patients could be exposed to respiratory depression if given morphine to
treat pain or dyspnea. Therefore, the clinical relevance of spironolactone in cotreatment with morphine in patients should be investigated.
6.5 A novel pharmacokinetic interaction between morphine
and ketamine
Study IV demonstrated that a single dose of ketamine (10 mg/kg s.c.) did not
cause antinociception in drug-naïve animals, but in rats under chronic morphine
pump treatment it caused long-lasting antinociception up to 150 minutes after
administration. At the time point of maximum antinociception, co-treatment
caused significant increases in the brain morphine, ketamine and norketamine
concentrations, indicating that morphine and ketamine have pharmacokinetic
interactions during chronic morphine treatment. After two days of morphine
withdrawal, the synergistic antinociceptive effect of ketamine and morphine in
coadministration was still observable, but to a smaller extent. Acute co-treatment
with ketamine and morphine in drug-naïve rats did not have such effects,
suggesting that chronic morphine pump treatment may modify ketamine
pharmacokinetics.
The most prominent co-treatment-induced increases in the brain
concentrations of morphine, ketamine, and norketamine were observed during
chronic morphine administration, during which the increase in antinociception
was also marked. The mean brain norketamine concentration was most notably
elevated (3.4-fold) in morphine-tolerant rats 90 min after ketamine
administration, and the concentration exceeded that of ketamine. Norketamine
has been identified as a potent NMDA receptor antagonist (Ebert et al., 1997;
Holtman et al., 2008; Shimoyama et al., 1999; see also 2.4.4), and it can
therefore be assumed that the vastly increased brain norketamine concentration
contributed to the increased antinociception.
After ketamine administration to rats under chronic morphine treatment,
the brain concentration of morphine was doubled at maximum antinociception,
while the serum concentrations were essentially unchanged, implicating an
87
Discussion
increased brain:serum ratio. Morphine is an acknowledged substrate of P-gp
(Letrent et al., 1999; Schinkel et al., 1995; Xie et al., 1999) and possibly also Mrp
(Su and Pasternak, 2013; see also 2.3.5), both of which are efflux proteins at the
BBB. The inhibition of such efflux proteins by ketamine could underlie the
decreased efflux and increased disposition of morphine in the brain. However, it is
not known whether ketamine inhibits or is a substrate of P-gp (Doan, 2002;
Varma et al., 2005) or Mrp. Thus, the mechanism behind the ketamine-induced
increase in the brain:serum ratio of morphine remains undetermined.
Ketamine increased morphine concentrations in the liver in all
experiments. In acute coadministration, the M3G:morphine metabolic ratio was
significantly decreased in the liver, supporting the inhibition of Ugt2b1 by
ketamine. Furthermore, when an acute high dose of ketamine was combined
with morphine in pilot studies, this effect was even more pronounced (Lilius et
al., unpublished data). Indeed, ketamine has previously inhibited UGT2B7 in
vitro (Qi et al., 2010; Uchaipichat et al., 2011). However, decreased elimination
did not consistently lead to increased serum concentrations. Thus, the inhibition
of efflux proteins rather than decreased metabolism may explain the increased
brain morphine concentrations after ketamine treatment.
Chronic morphine treatment caused a significant accumulation of
ketamine and norketamine in the liver, serum, and brain compared with
morphine-naïve rats. This effect was also to a lesser extent present during
morphine withdrawal, and it was independent of an acute morphine dose.
Similar accumulation could not be seen after acute coadministration in
morphine-naïve animals, indicating long-term induced changes in the function
of metabolic enzymes, transporter proteins or plasma drug carrier proteins that
affect the pharmacokinetics of ketamine. The results did not support significant
morphine-induced changes in ketamine metabolism. Thus, morphine toleranceinduced ketamine accumulation could lie behind the downregulation of a
ketamine (or its metabolite) efflux transporter, or the upregulation of a reuptake
transporter.
More research is needed to clarify the mechanisms of morphine toleranceinduced changes in ketamine pharmacokinetics. These findings also imply that
possible changes in brain concentrations of ketamine and morphine should be
taken into account in future clinical and preclinical studies that involve the
coadministration of these drugs. This novel pharmacokinetic drug interaction
could explain the variation in results when treating opioid tolerance in cancer
pain management, where high doses of morphine or other opioids are used.
88
Discussion
6.6 Future perspectives
The studies of this thesis demonstrate that tolerance to antinociceptive effects of
systemic and intrathecal morphine develops rapidly and reliably in in vivo
models. Opioid tolerance in animals has usually been suggested to be
predominantly pharmacodynamic, time- and dose-dependent, and reversible
(Collett, 1998). The first proposed mechanism for the development of tolerance
was the upregulation of adenylyl cyclase inhibition by repeated administration of
opioids (Collier, 1980), but tolerance has subsequently been shown to be a far
more complex phenomenon, involving neuronal changes at both the cellular
level and neuronal network level. In preclinical models, recent evidence has
brought glial cells and drug transporters into the spotlight of opioid tolerance
research.
In the clinic, pain is difficult to study due to problems in the
quantification of pain and pain relief, the influence of psychological factors in its
perception, and variability in the sensitivity of individuals to pain and to opioids.
Tolerance may often be difficult to dissociate from disease progression and/or
opioid-induced hyperalgesia when assessing the reasons for inadequate analgesia
and the need for escalating the opioid dose. The continuous progress in
noninvasive imaging techniques may provide new possibilities for the study of
glial responses in humans; for reviews, see Gerhard (2013), Inglese and Petracca
(2013), and Obenaus (2013).
The results of this thesis show that when two drugs are studied in
coadministration, possible pharmacokinetic interactions must be considered,
even when the studied drugs or drug combinations also have pharmacodynamic
interactions. Increasing knowledge of drug transporters, their substrates, and
potential genetic influences may aid in personalizing drug therapy and choosing
better drug combinations while avoiding harmful ones. The pharmacokinetic
interactions described in Studies III and IV may be advantageous, but one also
has to be aware of potential adverse effects. An unidentified drug interaction is
the most dangerous interaction. The undisputed advantage of animal research
regarding pharmacokinetics is the possibility to determine the drug and
metabolite concentrations in situ, i.e. the CNS. In humans, CNS drug
concentrations can be estimated with pharmacokinetic modeling, whereas
indirect radiological measurements have so far only been performed with a few
radiolabelled drugs using fluorine magnetic resonance spectroscopy (Strauss et
al., 1998).
89
Discussion
The data of Studies III and IV and previous information indicate that
spironolactone and ketamine may enhance morphine analgesia via both
pharmacodynamic and pharmacokinetic mechanisms. NMDA antagonists have
pharmacodynamically advantageous effects that may also indirectly affect the
pharmacokinetics of morphine, e.g. by attenuating the upregulation of P-gp at
the BBB, not by directly binding to the target P-gp protein but by antagonizing
the effects of excess glutamate (Zhu and Liu, 2004), linking pharmacodynamic
and pharmacokinetic interactions inseparably together.
Ibudilast has already been proven safe in the treatment of asthma in Asian
countries. Its multiple anti-inflammatory actions (inhibition of PDE and MIF
as well as antagonism of TLR4 signaling) make it an interesting molecule for the
treatment of various CNS diseases with inflammatory properties. It would be of
great value to carefully analyze all the target proteins of ibudilast to fully
understand its mechanism of action and to eliminate possibly harmful properties
in the development of upcoming ibudilast-derived molecules.
Almost all preclinical studies regarding opioid-induced pathological
activation of glia have focused on morphine, the gold standard opioid agonist,
even though the clinical use of other opioids, especially that of oxycodone, has
increased. Thus, more attention should be paid to the basic pharmacology of
oxycodone. The contribution of M3G, the pronociceptive metabolite of
morphine, to glial activation warrants further study, despite the difficult
synthesis process of the molecule (Moreira et al., personal communication).
M3G has been suggested to be a very potent activator of the TLR4-MD2
complex (Lewis et al., 2010), but the clinically important oxycodone and its
metabolites have not yet been characterized in this respect. An interesting
molecule, PTI-609 (Wang et al., 2012a), combines opioid receptor agonism
with the inhibition of TLR4-MD2 signaling by binding to filamin A, an
intracellular cytoskeleton protein (Burns and Wang, 2010). The development of
new molecules with such properties may allow the positive analgesic actions of
mu-opioid receptor activation while avoiding pronociceptive xenobiotic-induced
central immune signaling.
Ultralow doses of various alpha-2-adrenoceptor antagonists and opioid
receptor antagonists have shown promise for the treatment of acute opioid
antinociception and also tolerance (Table 2, page 42). As a nonselective alpha-2adrenoceptor antagonist, atipamezole is not ideal for dissecting the
pharmacology behind the phenomenon, but on the other hand, its pharmacology
is rather well characterized in veterinary medicine and also in human studies.
90
Discussion
More research is needed to understand the role of dimerization of alpha-2adrenoceptors and opioid receptors in opioid analgesia, bearing in mind the
possibility of other novel binding sites for alpha-2-antagonists.
91
CONCLUSIONS
The conclusions for the specific aims set in Aims of the study are as follows:
I.
Acutely administered ibudilast decreased spontaneous locomotion and
impaired motor coordination. Ibudilast did not inhibit the
development of opioid tolerance, but it restored the antinociceptive
effect of morphine in morphine-tolerant animals after single and
repeated administration. It did not prevent the development of opioid
tolerance in a four-day tolerance model.
II.
When administered subcutaneously, ultralow doses of atipamezole
failed to augment morphine antinociception in acute administration or
in morphine tolerance. In intrathecal administration, ultralow
nanogram-scale doses of atipamezole augmented morphine
antinociception both acutely and in morphine tolerance.
III.
Spironolactone dose-dependently increased the antinociceptive effect
of morphine. The effect was mediated via increased concentrations of
morphine in the CNS. Spironolactone did not inhibit the development
of the pronociceptive morphine metabolite M3G, and it did not
prevent the development of opioid tolerance. Loperamide, the
peripheral opioid and P-gp substrate, showed antinociceptive
properties when coadministered with spironolactone.
IV.
In acute cotreatment, a small dose of ketamine increased the brain
access of morphine only marginally, whereas the same dose of
ketamine administered to rats under chronic morphine treatment led
to markedly increased brain concentrations of morphine, ketamine,
and norketamine compared to the situation in which the drugs were
administered alone. Moreover, chronic morphine treatment caused
increased accumulation of ketamine, particularly in the liver.
92
ACKNOWLEDGEMENTS
These studies were carried out at Institute of Biomedicine, Pharmacology,
University of Helsinki during years 2008–2014. A significant contribution and
support was received from the Haartman Institute, Department of Clinical
Pharmacology, University of Helsinki and the Pain Clinic, Department of
Anaesthesia and Intensive Care Medicine, Helsinki University Central Hospital.
I wish to thank Professor Esa Korpi and Professor Eero Mervaala for
providing outstanding research facilities at Institute of Biomedicine,
Pharmacology. Both Esa and Eero are fine gentlemen with a continuous drive for
new discoveries. Esa’s immortal lunch table quote, “What’s new?” is an excellent
guiding principle for a junior scientist. Eero is thanked for always having time for
a friendly and supportive discussion, no matter what the topic was.
I have had the privilege to work and develop myself under the supervision of
Docent Pekka Rauhala. During his lectures, Pekka drew my attention to
neuropharmacology as a dynamic field of science and medicine. He practically
introduced me to laboratory work and especially in vivo experimentation. At
Pharmacology, a tradition of Ostrobothnian uprightness and honesty has prevailed
and Pekka is the best example of such values. I wish to express my deepest
gratitude for always having sincere interest in my studies and ideas and for guiding
me forward in pharmacology, both in research and teaching.
Since the beginning of my studies, Professor Eija Kalso’s enthusiasm
towards science has been one of my main motivators to push my MD and PhD
theses forward. I deeply value her as a scientist, clinician, and as a person. Whether
it was about science, patient care, or cultural aspects of life, Eija’s positive presence
and resilience is beyond comparison. I am deeply thankful for your supervision and
support during these years.
All my co-authors at Pharmacology are acknowledged: Viljami Jokinen,
Oleg Kambur, Sami Rossi, and Antti Väänänen. Viljami, without your friendship,
help, support, and incisive sense of humor, I wouldn’t be writing this. Oleg, thank
you for your help, especially at the beginning of my studies. Sami, thank you for
your friendship and support during our medical studies. The research collaboration
with Professor Mikko Niemi, Jouko Laitila, and Mikko Neuvonen at Department
of Clinical Pharmacology is greatly acknowledged and valued.
I wish to thank the senior scientists of our department, especially Professor
Heikki Vapaatalo and Docents Petri Hyytiä, Esko Kankuri, and Anni-Maija
93
Linden, for creating a positive and welcoming atmosphere. Of the junior
colleagues, special thanks to Vilma Aho, Juha Lempiäinen, and Kristo Nuutila for
sharing the scientific journey with me. Lotta Stenman is acknowledged for
amusing company. Warm thanks belong to Docent Mikko Airavaara and Katrina
Albert at Viikki Campus for introducing me to immunohistochemistry. Eeva
Harju, Ritva Henriksson, Sari Laakkonen, Lahja Eurajoki, Heidi Pehkonen, and
Vuokko Pahlsten are acknowledged for administrative and technical assistance
during my studies.
The work communities of Jorvi Hospital Anesthesia Unit and Kankaanpää
Health Center are acknowledged for encouraging me to finish this timeconsuming project.
My journey into science wouldn’t have been the same without music, the
perfect counterweight for long workdays. I wish to thank all my friends from
music circles, especially the guys and gals of Las Casetas Humidas, Sininen
Huvimaja, Syncope, and Tuba Auditiva, for all those fun moments and great gigs.
At activities revolving around medical school, I have been privileged to
make good friends with exceptional and warm-hearted people, of whom I would
like to especially thank Samuli, Risto, Eeva, Johanna, Tuukka, Elisa, Henna, Anja,
Salla, Mikael, Maria, Anne, and Ilari. Antti and Sami, thank you for your
friendship, support, and those unforgettable journeys around the world. Natalia,
thank you for the unique moments and adventures we’ve had together.
Finally, the deepest gratitude I owe to my parents, Marita and Reijo. Your
continuous support, love, and encouragement to work hard and strive for my goals
have brought me here.
Biomedicum Helsinki Foundation, the Emil Aaltonen Foundation, the
Finnish Medical Foundation, the Maud Kuistila Memorial Foundation, the
Orion-Farmos research foundation, Orion Pharma, and the Sohlberg Foundation
are thanked for financial support. The Finnish Medical Society (Finska
Läkaresällskapet) is especially acknowledged for continuous financial support and
belief in my studies.
Helsinki, November 2014
Tuomas Lilius
94
REFERENCES
Abul-Husn, N.S., Sutak, M., Milne, B., Jhamandas, K. (2007). Augmentation of spinal morphine
analgesia and inhibition of tolerance by low doses of mu- and delta-opioid receptor antagonists. Br
J Pharmacol 151, 877–887.
Aggleton, J.P., Burton, M.J., Passingham, R.E. (1980). Cortical and subcortical afferents to the
amygdala of the rhesus monkey (Macaca mulatta). Brain Res 190, 347–368.
Aguado, D., Abreu, M., Benito, J., Garcia-Fernandez, J., Gómez de Segura, I.A. (2013). Effects
of naloxone on opioid-induced hyperalgesia and tolerance to remifentanil under sevoflurane
anesthesia in rats. Anesthesiology 118, 1160–1169.
Alexander, J.K., Cox, G.M., Tian, J.-B., Zha, A.M., Wei, P., Kigerl, K.A., Reddy, M.K., Dagia,
N.M., Sielecki, T., Zhu, M.X., Satoskar, A.R., McTigue, D.M., Whitacre, C.C., Popovich, P.G.
(2012). Macrophage migration inhibitory factor (MIF) is essential for inflammatory and
neuropathic pain and enhances pain in response to stress. Exp Neurol 236, 351–362.
Aley, K.O., Levine, J.D. (1997). Multiple receptors involved in peripheral alpha 2, mu, and A1
antinociception, tolerance, and withdrawal. J Neurosci 17, 735–744.
Alonso, E., Garrido, E., Díez-Fernández, C., Pérez-García, C., Herradón, G., Ezquerra, L.,
Deuel, T.F., Alguacil, L.F. (2007). Yohimbine prevents morphine-induced changes of glial
fibrillary acidic protein in brainstem and alpha2-adrenoceptor gene expression in hippocampus.
Neurosci Lett 412, 163–167.
Amphoux, A., Vialou, V., Drescher, E., Brüss, M., La Cour, C.M., Rochat, C., Millan, M.J.,
Giros, B., Bönisch, H., Gautron, S. (2006). Differential pharmacological in vitro properties of
organic cation transporters and regional distribution in rat brain. Neuropharmacology 50, 941–952.
Andersen, G., Christrup, L., Sjøgren, P. (2003). Relationships among morphine metabolism, pain
and side effects during long-term treatment: an update. J Pain Symptom Manage 25, 74–91.
Angst, M.S., Clark, J.D. (2006). Opioid-induced hyperalgesia: a qualitative systematic review.
Anesthesiology 104, 570–587.
Apkarian, A.V., Bushnell, M.C., Treede, R.-D., Zubieta, J.-K. (2005). Human brain mechanisms
of pain perception and regulation in health and disease. Eur J Pain 9, 463–484.
Aquilante, C.L., Letrent, S.P., Pollack, G.M., Brouwer, K.L.R. (1999). Increased brain Pglycoprotein in morphine tolerant rats. Life Sci 66, PL47–PL51.
Arendt-Nielsen, L., Nielsen, J., Petersen-Felix, S., Schnider, T.W., Zbinden, A.M. (1996). Effect
of racemic mixture and the (S+)-isomer of ketamine on temporal and spatial summation of pain.
Br J Anaesth 77, 625–631.
Armstead, W.M., Mirro, R., Leffler, C.W., Busija, D.W. (1988). The role of prostanoids in the
mediation of responses to KC-404, a novel cerebrovasodilator. J Pharmacol Exp Ther 244, 138–
143.
Arroyo-Novoa, C.M., Figueroa-Ramos, M.I., Miaskowski, C., Padilla, G., Paul, S.M.,
Rodríguez-Ortiz, P., Stotts, N.A., Puntillo, K.A. (2011). Efficacy of small doses of ketamine with
morphine to decrease procedural pain responses during open wound care. Clin J Pain 27, 561–566.
Bailey, C.P., Connor, M. (2005). Opioids: cellular mechanisms of tolerance and physical
dependence. Curr Opin Pharmacol 5, 60–68.
95
References
Bajic, D., Proudfit, H.K. (1999). Projections of neurons in the periaqueductal gray to pontine and
medullary catecholamine cell groups involved in the modulation of nociception. J Comp Neurol
405, 359–379.
Ballantyne, J.C., Mao, J. (2003). Opioid therapy for chronic pain. N Engl J Med 349, 1943–1953.
Banki, C.M., Arato, M. (1987). Multiple hormonal responses to morphine: relationship to
diagnosis and dexamethasone suppression. Psychoneuroendocrinology 12, 3–11.
Baron, R. (2000). Peripheral neuropathic pain: from mechanisms to symptoms. Clin J Pain 16,
S12–S20.
Bartolome, M.B., Kuhn, C.M. (1983). Endocrine effects of methadone in rats; acute effects in
adults. Eur J Pharmacol 95, 231–238.
Basbaum, A.I., Bautista, D.M., Scherrer, G., Julius, D. (2009). Cellular and molecular
mechanisms of pain. Cell 139, 267–284.
Basbaum, A.I., Jessell, T. (2000). The Perception of Pain. In Kandel's Principles of Neural Science,
E. Kandel, and J. Schwartz, eds. (McGraw-Hill Medical).
Bell, R.F. (2012). Ketamine for chronic noncancer pain: concerns regarding toxicity. Curr Opin
Support Palliat Care 6, 183–187.
Bell, R.F., Dahl, J.B., Moore, R.A., Kalso, E. (2005). Peri-operative ketamine for acute postoperative pain: a quantitative and qualitative systematic review (Cochrane review). Acta Anaesthesiol
Scand 49, 1405–1428.
Bendtzen, K., Hansen, P.R., Rieneck, K., Spironolactone/Arthritis Study Group (2003).
Spironolactone inhibits production of proinflammatory cytokines, including tumour necrosis
factor-alpha and interferon-gamma, and has potential in the treatment of arthritis. Clin Exp
Immunol 134, 151–158.
Benthuysen, J.L., Hance, A.J., Quam, D.D., Winters, W.D. (1989). Synthetic opioids compared
with morphine and ketamine: catalepsy, cross-tolerance and interactions in the rat.
Neuropharmacology 28, 1011–1015.
Bentley, G.A., Newton, S.H., Starr, J. (1983). Studies on the antinociceptive action of alphaagonist drugs and their interactions with opioid mechanisms. Br J Pharmacol 79, 125–134.
Berridge, C.W., Arnsten, A.F., Foote, S.L. (1993). Noradrenergic modulation of cognitive
function: clinical implications of anatomical, electrophysiological and behavioural studies in animal
models. Psychol Med 23, 557–564.
Besse, D., Lombard, M.C., Zajac, J.M. (1990). Pre-and postsynaptic location of mu, delta and
kappa opioid receptors in the superficial layers of the dorsal horn of the rat spinal cord. Prog Clin
Biol Res 328, 183-186.
Bickel, U., Schumacher, O.P., Kang, Y.S., Voigt, K. (1996). Poor permeability of morphine 3glucuronide and morphine 6-glucuronide through the blood-brain barrier in the rat. J Pharmacol
Exp Ther 278, 107–113.
Bland, S.T., Hutchinson, M.R., Maier, S.F., Watkins, L.R., Johnson, K.W. (2009). The glial
activation inhibitor AV411 reduces morphine-induced nucleus accumbens dopamine release. Brain
Behav Immun 23, 492–497.
Boström, E., Simonsson, U. (2005). Oxycodone pharmacokinetics and pharmacodynamics in the
rat in the presence of the P‐glycoprotein inhibitor PSC833. J Pharm Sci 94, 1060-1066.
96
References
Bourasset, F., Cisternino, S., Temsamani, J. (2003). Evidence for an active transport of
morphine‐6‐β‐d‐glucuronide but not P‐glycoprotein‐mediated at the blood–brain barrier. J
Neurochem 86, 1564-1567.
Branford, R., Droney, J., Ross, J.R. (2012). Opioid genetics: the key to personalized pain control?
Clin Genet 82, 301–310.
Browning, S., Lawrence, D., Livingston, A., Morris, B. (1982). Interactions of drugs active at
opiate receptors and drugs active at alpha 2-receptors on various test systems. Br J Pharmacol 77,
487–491.
Burns, L.H., Wang, H.-Y. (2010). PTI-609: a novel analgesic that binds filamin A to control
opioid signaling. Recent Pat CNS Drug Discov 5, 210–220.
Callaghan, R., Riordan, J.R. (1993). Synthetic and natural opiates interact with P-glycoprotein in
multidrug-resistant cells. J Biol Chem 268, 16059–16064.
Caraceni, A., Hanks, G., Kaasa, S., Bennett, M.I., Brunelli, C., Cherny, N., Dale, O., De Conno,
F., Fallon, M., Hanna, M., Haugen, D.F., Juhl, G., King, S., Klepstad, P., Laugsand, E.A.,
Maltoni, M., Mercadante, S., Nabal, M., Pigni, A., Radbruch, L., Reid, C., Sjøgren, P., Stone,
P.C., Tassinari, D., Zeppetella, G., European Palliative Care Research Collaborative (EPCRC),
European Association for Palliative Care (EAPC) (2012). Use of opioid analgesics in the
treatment of cancer pain: evidence-based recommendations from the EAPC. Lancet Oncol 13,
e58–e68.
Carstensen, M., Møller, A.M. (2010). Adding ketamine to morphine for intravenous patientcontrolled analgesia for acute postoperative pain: a qualitative review of randomized trials. Br J
Anaesth 104, 401–406.
Cepeda, M.S., Africano, J.M., Manrique, A.M., Fragoso, W., Carr, D.B. (2002). The
combination of low dose of naloxone and morphine in PCA does not decrease opioid requirements
in the postoperative period. Pain 96, 73–79.
Cepeda, M.S., Alvarez, H., Morales, O., Carr, D.B. (2004). Addition of ultralow dose naloxone
to postoperative morphine PCA: unchanged analgesia and opioid requirement but decreased
incidence of opioid side effects. Pain 107, 41–46.
Célèrier, E., Rivat, C., Jun, Y., Laulin, J.P., Larcher, A., Reynier, P., Simonnet, G. (2000). Longlasting hyperalgesia induced by fentanyl in rats: preventive effect of ketamine. Anesthesiology 92,
465–472.
Célérier, E., Simonnet, G., Maldonado, R. (2004). Prevention of fentanyl-induced delayed
pronociceptive effects in mice lacking the protein kinase Cgamma gene. Neuropharmacology 46,
264–272.
Chakrabarti, S., Regec, A., Gintzler, A.R. (2005). Biochemical demonstration of mu-opioid
receptor association with Gsα: enhancement following morphine exposure. Mol Brain Res 135,
217–224.
Chantong, B., Kratschmar, D.V., Nashev, L.G., Balazs, Z., Odermatt, A. (2012).
Mineralocorticoid and glucocorticoid receptors differentially regulate NF-kappaB activity and proinflammatory cytokine production in murine BV-2 microglial cells. J Neuroinflammation 9, 260.
Chen, X., Shu, S., Bayliss, D.A. (2009). HCN1 channel subunits are a molecular substrate for
hypnotic actions of ketamine. J Neurosci 29, 600–609.
Chia, Y.Y., Liu, K., Wang, J.J., Kuo, M.C., Ho, S.T. (1999). Intraoperative high dose fentanyl
induces postoperative fentanyl tolerance. Can J Anaesth 46, 872–877.
97
References
Chindalore, V.L., Craven, R.A., Yu, K.P., Butera, P.G., Burns, L.H., Friedmann, N. (2005).
Adding ultralow-dose naltrexone to oxycodone enhances and prolongs analgesia: a randomized,
controlled trial of Oxytrex. J Pain 6, 392–399.
Cho, Y., Crichlow, G., Vermeire, J., Leng, L. (2010a). AV411 (Ibudilast) and AV1013 are noncompetitive inhibitors of macrophage migration inhibitory factor: a novel induced-fit allosteric
inhibition mechanism. J Immunol 184, Meeting abstract supplement.
Cho, Y., Crichlow, G.V., Vermeire, J.J., Leng, L., Du, X., Hodsdon, M.E., Bucala, R., Cappello,
M., Gross, M., Gaeta, F., Johnson, K., Lolis, E.J. (2010b). Allosteric inhibition of macrophage
migration inhibitory factor revealed by ibudilast. Proc Natl Atl Sci USA 107, 11313–11318.
Chong, C., Schug, S.A., Page-Sharp, M., Jenkins, B., Ilett, K.F. (2009). Development of a
sublingual/oral formulation of ketamine for use in neuropathic pain: Preliminary findings from a
three-way randomized, crossover study. Clin Drug Investig 29, 317–324.
Chu, H., Klemp, A., Stille, G. (1978). Effects of furosemide and spironolactone on the behavior
of morphine-tolerant rats. Psychopharmacology (Berl) 59, 309–310.
Chu, L.F., Clark, D.J., Angst, M.S. (2006). Opioid tolerance and hyperalgesia in chronic pain
patients after one month of oral morphine therapy: a preliminary prospective study. J Pain 7, 43–
48.
Clements, J.A., Nimmo, W.S., Grant, I.S. (1982). Bioavailability, pharmacokinetics, and analgesic
activity of ketamine in humans. J Pharm Sci 71, 539–542.
Coffman, B.L., Rios, G.R., King, C.D., Tephly, T.R. (1997). Human UGT2B7 catalyzes
morphine glucuronidation. Drug Metab Dispos 25, 1–4.
Coffman, B.L., Rios, G.R., Tephly, T.R. (1996). Purification and properties of two rat liver
phenobarbital-inducible UDP-glucuronosyltransferases that catalyze the glucuronidation of
opioids. Drug Metab Dispos 24, 329–333.
Coller, J.K., Hutchinson, M.R. (2012). Implications of central immune signaling caused by drugs
of abuse: Mechanisms, mediators and new therapeutic approaches for prediction and treatment of
drug dependence. Pharmacol Ther 134, 219–245.
Collett, B.J. (1998). Opioid tolerance: the clinical perspective. Br J Anaesth 81, 58–68.
Collier, H.O. (1980). Cellular site of opiate dependence. Nature 283, 625–629.
Collu, R., Clermont, M.J., Ducharme, J.R. (1976). Effects of thyrotropin-releasing hormone on
prolactin, growth hormone and corticosterone secretions in adult male rats treated with
pentobarbital or morphine. Eur J Pharmacol 37, 133–140.
Compton, P., Canamar, C.P., Hillhouse, M., Ling, W. (2012). Hyperalgesia in heroin dependent
patients and the effects of opioid substitution therapy. J Pain 13, 401–409.
Compton, P., Charuvastra, V.C., Ling, W. (2001). Pain intolerance in opioid-maintained former
opiate addicts: effect of long-acting maintenance agent. Drug Alcohol Depend 63, 139–146.
Connor, M., Christie, M.D. (1999). Opioid receptor signalling mechanisms. Clin Exp Pharmacol
Physiol 26, 493–499.
Cooper, D.W., Lindsay, S.L., Ryall, D.M., Kokri, M.S., Eldabe, S.S., Lear, G.A. (1997). Does
intrathecal fentanyl produce acute cross-tolerance to i.v. morphine? Br J Anaesth 78, 311–313.
Coughtrie, M.W., Ask, B., Rane, A., Burchell, B., Hume, R. (1989). The enantioselective
glucuronidation of morphine in rats and humans. Evidence for the involvement of more than one
UDP-glucuronosyltransferase isoenzyme. Biochem Pharmacol 38, 3273–3280.
98
References
Crowe, A. (2002). The influence of P-glycoprotein on morphine transport in Caco-2 cells.
Comparison with paclitaxel. Eur J Pharmacol 440, 7–16.
Cui, Y., Chen, Y., Zhi, J.-L., Guo, R.-X., Feng, J.-Q., Chen, P.-X. (2006). Activation of p38
mitogen-activated protein kinase in spinal microglia mediates morphine antinociceptive tolerance.
Brain Res 1069, 235–243.
Cui, Y., Liao, X.-X., Liu, W., Guo, R.-X., Wu, Z.-Z., Zhao, C.-M., Chen, P.-X., Feng, J.-Q.
(2008). A novel role of minocycline: attenuating morphine antinociceptive tolerance by inhibition
of p38 MAPK in the activated spinal microglia. Brain Behav Immun 22, 114–123.
D'Amour, F.E., Smith, D.L. (1941). A method for determining loss of pain sensation. J Pharmacol
Exp Ther 72, 74–79.
Dagenais, C., Graff, C.L., Pollack, G.M. (2004). Variable modulation of opioid brain uptake by
P-glycoprotein in mice. Biochem Pharmacol 67, 269–276.
Daulhac, L., Mallet, C., Courteix, C., Etienne, M., Duroux, E., Privat, A.M., Eschalier, A.,
Fialip, J. (2006). Diabetes-Induced Mechanical Hyperalgesia Involves Spinal Mitogen-Activated
Protein Kinase Activation in Neurons and Microglia via N-Methyl-D-aspartate-Dependent
Mechanisms. Mol Pharmacol 70, 1246–1254.
Davis, K.D., Treede, R.D., Raja, S.N., Meyer, R.A., Campbell, J.N. (1991). Topical application
of clonidine relieves hyperalgesia in patients with sympathetically maintained pain. Pain 47, 309–
317.
DeLeo, J.A., Colburn, R.W., Rickman, A.J., Yeager, M.P. (1997). Intrathecal catheterization
alone induces neuroimmune activation in the rat. Eur J Pain 1, 115–122.
Doan, K.M.M. (2002). Passive Permeability and P-Glycoprotein-Mediated Efflux Differentiate
Central Nervous System (CNS) and Non-CNS Marketed Drugs. J Pharm Exp Ther 303, 1029–
1037.
Doehring, A., Oertel, B.G., Sittl, R., Lötsch, J. (2013). Chronic opioid use is associated with
increased DNA methylation correlating with increased clinical pain. Pain 154, 15–23.
Doggrell, S.A., Brown, L. (2001). The spironolactone renaissance. Expert Opin Investig Drugs 10,
943–954.
Dong, F., Xie, W., Strong, J.A., Zhang, J.-M. (2012). Mineralocorticoid receptor blocker
eplerenone reduces pain behaviors in vivo and decreases excitability in small-diameter sensory
neurons from local inflamed dorsal root ganglia in vitro. Anesthesiology 117, 1102–1112.
Doverty, M., White, J.M., Somogyi, A.A., Bochner, F., Ali, R., Ling, W. (2001). Hyperalgesic
responses in methadone maintenance patients. Pain 90, 91–96.
Doyle, C.A., Maxwell, D.J. (1991). Catecholaminergic innervation of the spinal dorsal horn: a
correlated light and electron microscopic analysis of tyrosine hydroxylase-immunoreactive fibres in
the cat. Neuroscience 45, 161–176.
Drummond, P.D. (1995). Noradrenaline increases hyperalgesia to heat in skin sensitized by
capsaicin. Pain 60, 311–315.
Dzamba, D., Honsa, P., Anderova, M. (2013). NMDA Receptors in Glial Cells: Pending
Questions. Cn 11, 250–262.
Ebert, B., Mikkelsen, S., Thorkildsen, C., Borgbjerg, F.M. (1997). Norketamine, the main
metabolite of ketamine, is a non-competitive NMDA receptor antagonist in the rat cortex and
spinal cord. Eur J Pharmacol 333, 99–104.
99
References
Edwards, S.R., Minto, C.F., Mather, L.E. (2002). Concurrent ketamine and alfentanil
administration: pharmacokinetic considerations. Br J Anaesth 88, 94–100.
Eidson, L.N., Murphy, A.Z. (2013). Blockade of Toll-Like Receptor 4 Attenuates Morphine
Tolerance and Facilitates the Pain Relieving Properties of Morphine. J Neurosci 33, 15952–15963.
Ellis, A., Wieseler, J., Favret, J., Johnson, K.W., Rice, K.C., Maier, S.F., Falci, S., Watkins, L.R.
(2014). Systemic Administration of Propentofylline, Ibudilast, and (+)-Naltrexone Each Reverses
Mechanical Allodynia in a Novel Rat Model of Central Neuropathic Pain. J Pain 15, 407–421.
Ewing, K.K., Mohammed, H.O., Scarlett, J.M. (1993). Reduction of isoflurane anesthetic
requirement by medetomidine and its restoration by atipamezole in dogs. Am J Vet Res 54, 294–
299.
Ferrini, F., Trang, T., Mattioli, T.-A.M., Laffray, S., Del'Guidice, T., Lorenzo, L.-E.,
Castonguay, A., Doyon, N., Zhang, W., Godin, A.G., Mohr, D., Beggs, S., Vandal, K., Beaulieu,
J.-M., Cahill, C.M., Salter, M.W., De Koninck, Y. (2013). Morphine hyperalgesia gated through
microglia-mediated disruption of neuronal Cl− homeostasis. Nature Neuroscience 16, 183–192.
Finck, A.D., Ngai, S.H. (1982). Opiate receptor mediation of ketamine analgesia. Anesthesiology
56, 291–297.
Fishbain, D.A., Cole, B., Lewis, J.E., Gao, J., Rosomoff, R.S. (2009). Do opioids induce
hyperalgesia in humans? An evidence-based structured review. Pain Med 10, 829–839.
Fletcher, D., Martinez, V. (2014). Opioid-induced hyperalgesia in patients after surgery: a
systematic review and a meta-analysis. Br J Anaesth 112, 991–1004.
Floyd, N.S., Price, J.L., Ferry, A.T., Keay, K.A., Bandler, R. (2000). Orbitomedial prefrontal
cortical projections to distinct longitudinal columns of the periaqueductal gray in the rat. J Comp
Neurol 422, 556–578.
Frenkel, C., Urban, B.W. (1992). Molecular actions of racemic ketamine on human CNS sodium
channels. Br J Anaesth 69, 292–297.
Frieler, R.A., Meng, H., Duan, S.Z., Berger, S., Schütz, G., He, Y., Xi, G., Wang, M.M.,
Mortensen, R.M. (2011). Myeloid-specific deletion of the mineralocorticoid receptor reduces
infarct volume and alters inflammation during cerebral ischemia. Stroke 42, 179–185.
Frohlich, J., Van Horn, J.D. (2014). Reviewing the ketamine model for schizophrenia. J
Psychopharmacol (Oxford) 28, 287–302.
Fukagawa, H., Koyama, T., Kakuyama, M., Fukuda, K. (2012). Microglial activation involved in
morphine tolerance is not mediated by toll-like receptor 4. J Anesth 27, 93–97.
Fukuyama, H., Kimura, J., Yamaguchi, S., Yamauchi, H., Ogawa, M., Doi, T., Yonekura, Y.,
Konishi, J. (1993). Pharmacological effects of ibudilast on cerebral circulation: a PET study.
Neurol Res 15, 169–173.
Gao, Y.-J., Ji, R.-R. (2010). Targeting astrocyte signaling for chronic pain. Neurotherapeutics 7,
482–493.
Gao, Y.-J., Zhang, L., Samad, O.A., Suter, M.R., Yasuhiko, K., Xu, Z.-Z., Park, J.-Y., Lind, A.L., Ma, Q., Ji, R.-R. (2009). JNK-induced MCP-1 production in spinal cord astrocytes
contributes to central sensitization and neuropathic pain. J Neurosci 29, 4096–4108.
Garcia-Larrea, L., Peyron, R. (2013). Pain matrices and neuropathic pain matrices: a review. Pain
154 Suppl 1, S29–S43.
100
References
Gardell, L.R., Wang, R., Burgess, S.E. (2002). Sustained morphine exposure induces a spinal
dynorphin-dependent enhancement of excitatory transmitter release from primary afferent fibers. J
Neurosci 22, 6747–6755.
Garrido, E., Pérez-García, C., Alguacil, L.F., Díez-Fernández, C. (2005). The alpha2adrenoceptor antagonist yohimbine reduces glial fibrillary acidic protein upregulation induced by
chronic morphine administration. Neurosci Lett 383, 141–144.
Gavaldà, A., Roberts, R.S. (2013). Phosphodiesterase-4 inhibitors: a review of current
developments (2010 - 2012). Expert Opin Ther Pat 23, 997–1016.
Geisslinger, G., Hering, W., Thomann, P., Knoll, R., Kamp, H.D., Brune, K. (1993).
Pharmacokinetics and pharmacodynamics of ketamine enantiomers in surgical patients using a
stereoselective analytical method. Br J Anaesth 70, 666–671.
Gerhard, A. (2013). Imaging of neuroinflammation in parkinsonian syndromes with positron
emission tomography. Curr Neurol Neurosci Rep 13, 405.
Ghosh, M., Garcia-Castillo, D., Aguirre, V., Golshani, R., Atkins, C.M., Bramlett, H.M.,
Dietrich, W.D., Pearse, D.D. (2012). Proinflammatory cytokine regulation of cyclic AMPphosphodiesterase 4 signaling in microglia in vitro and following CNS injury. Glia 60, 1839–1859.
Gibson, L.C.D., Hastings, S.F., McPhee, I., Clayton, R.A., Darroch, C.E., Mackenzie, A.,
MacKenzie, F.L., Nagasawa, M., Stevens, P.A., MacKenzie, S.J. (2006). The inhibitory profile of
Ibudilast against the human phosphodiesterase enzyme family. Eur J Pharmacol 538, 39–42.
Ginès, P., Cárdenas, A., Arroyo, V., Rodés, J. (2004). Management of cirrhosis and ascites. N
Engl J Med 350, 1646–1654.
Go, V.L., Yaksh, T.L. (1987). Release of substance P from the cat spinal cord. J Physiol (Lond)
391, 141–167.
Grace, P.M., Hutchinson, M.R., Maier, S.F., Watkins, L.R. (2014). Pathological pain and the
neuroimmune interface. Nat Rev Immunol 14, 217–231.
Green, G.M., Lyons, L., Dickenson, A.H. (1998). α2-Adrenoceptor antagonists enhance
responses of dorsal horn neurones to formalin induced inflammation. Eur J Pharmacol 347, 201–
204.
Grisel, J.E., Watkins, L.R., Maier, S.F. (1996). Associative and non-associative mechanisms of
morphine analgesic tolerance are neurochemically distinct in the rat spinal cord.
Psychopharmacology (Berl) 128, 248–255.
Groenendaal, D., Freijer, J., de Mik, D., Bouw, M.R., Danhof, M., de Lange, E.C.M. (2007).
Population pharmacokinetic modelling of non-linear brain distribution of morphine: influence of
active saturable influx and P-glycoprotein mediated efflux. Br J Pharmacol 151, 701–712.
Gu, X., Peng, L., Yang, D., Ma, Q., Zheng, Y., Liu, C., Zhu, B., Song, L., Sun, X., Ma, Z.
(2011). The respective and interaction effects of spinal GRs and MRs on radicular pain induced by
chronic compression of the dorsal root ganglion in the rat. Brain Res 1396, 88–95.
Guignard, B., Bossard, A.E., Coste, C., Sessler, D.I., Lebrault, C., Alfonsi, P., Fletcher, D.,
Chauvin, M. (2000). Acute opioid tolerance: intraoperative remifentanil increases postoperative
pain and morphine requirement. Anesthesiology 93, 409–417.
Haapalinna, A., Leino, T., Heinonen, E. (2003). The alpha 2-adrenoceptor antagonist
atipamezole potentiates anti-Parkinsonian effects and can reduce the adverse cardiovascular effects
of dopaminergic drugs in rats. Naunyn-Schmiedeberg's Arch Pharmacol 368, 342–351.
101
References
Haapalinna, A., MacDonald, E., Viitamaa, T., Salonen, J.S., Sirviö, J., Virtanen, R. (1999).
Comparison of the effects of acute and subchronic administration of atipamezole on reaction to
novelty and active avoidance learning in rats. Naunyn-Schmiedeberg's Arch Pharmacol 359, 194–203.
Haapalinna, A., Sirviö, J., MacDonald, E., Virtanen, R., Heinonen, E. (2000). The effects of a
specific alpha(2)-adrenoceptor antagonist, atipamezole, on cognitive performance and brain
neurochemistry in aged Fisher 344 rats. Eur J Pharmacol 387, 141–150.
Haapalinna, A., Viitamaa, T., MacDonald, E. (1997). Evaluation of the effects of a specific α2adrenoceptor antagonist, atipamezole, on α1-and α2-adrenoceptor subtype binding, brain
neurochemistry and behaviour in comparison with yohimibine. Naunyn-Schmiedeberg's Arch
Pharmacol 356, 570-582.
Hagihira, S., Senba, E., Yoshida, S., Tohyama, M., Yoshiya, I. (1990). Fine structure of
noradrenergic terminals and their synapses in the rat spinal dorsal horn: an immunohistochemical
study. Brain Res 526, 73–80.
Hama, A.T., Broadhead, A., Lorrain, D.S., Sagen, J. (2012). The antinociceptive effect of the
asthma drug ibudilast in rat models of peripheral and central neuropathic pain. J Neurotrauma 29,
600–610.
Hance, A.J., Winters, W.D., Quam, D.D., Benthuysen, J.L., Cadd, G.G. (1989). Catalepsy
induced by combinations of ketamine and morphine: potentiation, antagonism, tolerance and
cross-tolerance in the rat. Neuropharmacology 28, 109–116.
Hanks, G.W., Twycross, R.G. (1984). Pain, the physiological antagonist of opioid analgesics.
Lancet 1, 1477–1478.
Hansen, P.R., Rieneck, K., Bendtzen, K. (2004). Spironolactone inhibits production of
proinflammatory cytokines by human mononuclear cells. Immunol Lett 91, 87–91.
Harada, S., Nakamoto, K., Tokuyama, S. (2013). The involvement of midbrain astrocyte in the
development of morphine tolerance. Life Sci 93, 573–578.
Hassan, H.E., Myers, A.L., Lee, I.J., Chen, H., Coop, A., Eddington, N.D. (2009). Regulation
of Gene Expression in Brain Tissues of Rats Repeatedly Treated by the Highly Abused Opioid
Agonist, Oxycodone: Microarray Profiling and Gene Mapping Analysis. Drug Metab Dispos 38,
157–167.
Hassan, H.E., Myers, A.L., Lee, I.J., Coop, A., Eddington, N.D. (2007). Oxycodone induces
overexpression of P-glycoprotein (ABCB1) and affects paclitaxel's tissue distribution in Sprague
Dawley rats. J Pharm Sci 96, 2494–2506.
Hay, J.L., White, J.M., Bochner, F., Somogyi, A.A., Semple, T.J., Rounsefell, B. (2009).
Hyperalgesia in opioid-managed chronic pain and opioid-dependent patients. J Pain 10, 316–322.
Hedman, A., Angelin, B., Arvidsson, A., Dahlqvist, R. (1992). Digoxin-interactions in man:
spironolactone reduces renal but not biliary digoxin clearance. Eur J Clin Pharmacol 42, 481–485.
Heinricher, M.M., Fields, H.L. (2013). Central Nervous System Mechanisms of Pain
Modulation. In Wall and Melzack's Textbook of Pain, S.B. McMahon, M. Koltzenburg, I. Tracey,
and D.C. Turk, eds. (Elsevier Saunders).
Hennessy, M., Spiers, J.P. (2007). A primer on the mechanics of P-glycoprotein the multidrug
transporter. Pharmacological Research.
Hijazi, Y. (2002). Contribution of CYP3A4, CYP2B6, and CYP2C9 Isoforms to NDemethylation of Ketamine in Human Liver Microsomes. Drug Metab Dispos 30, 853–858.
102
References
Hirota, K., Okawa, H., Appadu, B.L., Grandy, D.K., Devi, L.A., Lambert, D.G. (1999).
Stereoselective interaction of ketamine with recombinant mu, kappa, and delta opioid receptors
expressed in Chinese hamster ovary cells. Anesthesiology 90, 174–182.
Holtman, J.R., Crooks, P.A., Johnson-Hardy, J.K., Hojomat, M., Kleven, M., Wala, E.P. (2008).
Effects of norketamine enantiomers in rodent models of persistent pain. Pharmacol Biochem Behav
90, 10–10.
Honarmand, A., Safavi, M., Karaky, H. (2012). Preincisional administration of intravenous or
subcutaneous infiltration of low-dose ketamine suppresses postoperative pain after appendectomy.
J Pain Res 5, 1–6.
Hooten, W.M., Mantilla, C.B., Sandroni, P., Townsend, C.O. (2010). Associations between heat
pain perception and opioid dose among patients with chronic pain undergoing opioid tapering.
Pain Med 11, 1587–1598.
Horvath, R.J., DeLeo, J.A. (2009). Morphine enhances microglial migration through modulation
of P2X4 receptor signaling. J Neurosci 29, 998–1005.
Horvath, R.J., Romero-Sandoval, E.A., De Leo, J.A. (2010). Inhibition of microglial P2X4
receptors attenuates morphine tolerance, Iba1, GFAP and mu opioid receptor protein expression
while enhancing perivascular microglial ED2. Pain 150, 401–413.
Hoskin, P.J. (2008). Opioids in context: relieving the pain of cancer. The role of comprehensive
cancer management. Palliat Med 22, 303–309.
Hösli, L., Hösli, E., Zehntner, C., Lehmann, R., Lutz, T.W. (1982). Evidence for the existence
of α-and β-adrenoceptors on cultured glial cells—an electrophysiological study. Neuroscience 7,
2867–2872.
Huang, Y.-N., Tsai, R.-Y., Lin, S.-L., Chien, C.-C., Cherng, C.-H., Wu, C.-T., Yeh, C.-C.,
Wong, C.-S. (2012). Amitriptyline attenuates astrocyte activation and morphine tolerance in rats:
role of the PSD-95/NR1/nNOS/PKCγ signaling pathway. Behav Brain Res 229, 401–411.
Hustveit, O., Maurset, A., Oye, I. (1995). Interaction of the chiral forms of ketamine with opioid,
phencyclidine, sigma and muscarinic receptors. Pharmacol Toxicol 77, 355–359.
Hutchinson, M.R., Coats, B.D., Lewis, S.S., Zhang, Y., Sprunger, D.B., Rezvani, N., Baker,
E.M., Jekich, B.M., Wieseler, J.L., Somogyi, A.A., Martin, D., Poole, S., Judd, C.M., Maier,
S.F., Watkins, L.R. (2008a). Proinflammatory cytokines oppose opioid-induced acute and chronic
analgesia. Brain Behav Immun 22, 1178–1189.
Hutchinson, M.R., Lewis, S.S., Coats, B.D., Rezvani, N., Zhang, Y., Wieseler, J.L., Somogyi,
A.A., Yin, H., Maier, S.F., Rice, K.C., Watkins, L.R. (2010a). Possible involvement of toll-like
receptor 4/myeloid differentiation factor-2 activity of opioid inactive isomers causes spinal
proinflammation and related behavioral consequences. Neuroscience 167, 880–893.
Hutchinson, M.R., Lewis, S.S., Coats, B.D., Skyba, D.A., Crysdale, N.Y., Berkelhammer, D.L.,
Brzeski, A., Northcutt, A., Vietz, C.M., Judd, C.M., Maier, S.F., Watkins, L.R., Johnson, K.W.
(2009). Reduction of opioid withdrawal and potentiation of acute opioid analgesia by systemic
AV411 (ibudilast). Brain Behav Immun 23, 240–250.
Hutchinson, M.R., Northcutt, A.L., Chao, L.W., Kearney, J.J., Zhang, Y., Berkelhammer, D.L.,
Loram, L.C., Rozeske, R.R., Bland, S.T., Maier, S.F., Gleeson, T.T., Watkins, L.R. (2008b).
Minocycline suppresses morphine-induced respiratory depression, suppresses morphine-induced
reward, and enhances systemic morphine-induced analgesia. Brain Behav Immun 22, 1248–1256.
103
References
Hutchinson, M.R., Shavit, Y., Grace, P.M., Rice, K.C., Maier, S.F., Watkins, L.R. (2011).
Exploring the neuroimmunopharmacology of opioids: an integrative review of mechanisms of
central immune signaling and their implications for opioid analgesia. Pharmacol Rev 63, 772–810.
Hutchinson, M.R., Zhang, Y., Shridhar, M., Evans, J.H., Buchanan, M.M., Zhao, T.X., Slivka,
P.F., Coats, B.D., Rezvani, N., Wieseler, J., Hughes, T.S., Landgraf, K.E., Chan, S., Fong, S.,
Phipps, S., Falke, J.J., Leinwand, L.A., Maier, S.F., Yin, H., Rice, K.C., Watkins, L.R. (2010b).
Evidence that opioids may have toll-like receptor 4 and MD-2 effects. Brain Behav Immun 24, 83–
95.
Huxtable, C.A., Roberts, L.J., Somogyi, A.A. (2011). Acute pain management in opioid-tolerant
patients: a growing challenge. Anaesth Intensive Care 39, 804–823.
Ikeda, K., Kobayashi, T., Kumanishi, T., Niki, H., Yano, R. (2000). Involvement of G-proteinactivated inwardly rectifying K (GIRK) channels in opioid-induced analgesia. Neurosci Res 38,
113–116.
Inglese, M., Petracca, M. (2013). Imaging multiple sclerosis and other neurodegenerative diseases.
Prion 7, 47–54.
Jaggi, A.S., Singh, N. (2010). Differential effect of spironolactone in chronic constriction injury
and vincristine-induced neuropathic pain in rats. Eur J Pharmacol 648, 102–109.
Jalanka, H. (1989). The use of medetomidine, medetomidine-ketamine combinations and
atipamezole at Helsinki Zoo--a review of 240 cases. Acta Vet Scand Suppl 85, 193–197.
Ji, R.-R., Berta, T., Nedergaard, M. (2013). Glia and pain: Is chronic pain a gliopathy? Pain 154
Suppl 1, S10–S28.
Ji, R.-R., Gereau, R.W., IV, Malcangio, M., Strichartz, G.R. (2009). MAP kinase and pain.
Brain Research Reviews 60, 135–148.
Jones, S.L. (1991). Descending noradrenergic influences on pain. Prog Brain Res 88, 381–394.
Jordan, B.A., Gomes, I., Rios, C., Filipovska, J., Devi, L.A. (2003). Functional interactions
between mu opioid and alpha 2A-adrenergic receptors. Mol Pharmacol 64, 1317–1324.
Julius, D. (2013). TRP channels and pain. Annu Rev Cell Dev Biol 29, 355–384.
Juni, A., Klein, G., Pintar, J.E., Kest, B. (2007). Nociception increases during opioid infusion in
opioid receptor triple knock-out mice. Neuroscience 147, 439–444.
Kalso, E., Edwards, J.E., Moore, R.A., McQuay, H.J. (2004). Opioids in chronic non-cancer
pain: systematic review of efficacy and safety. Pain 112, 372–380.
Karhuvaara, S., Kallio, A., Scheinin, M., Anttila, M., Salonen, J.S., Scheinin, H. (1990).
Pharmacological effects and pharmacokinetics of atipamezole, a novel alpha 2-adrenoceptor
antagonist--a randomized, double-blind cross-over study in healthy male volunteers. Br J Clin
Pharmacol 30, 97–106.
Katz, J., Clairoux, M., Redahan, C., Kavanagh, B.P., Carroll, S., Nierenberg, H., Jackson, M.,
Beattie, J., Taddio, A., Sandler, A.N. (1996). High dose alfentanil pre-empts pain after abdominal
hysterectomy. Pain 68, 109–118.
Kawasaki, A., Hoshino, K., Osaki, R., Mizushima, Y., Yano, S. (1992). Effect of ibudilast: a novel
antiasthmatic agent, on airway hypersensitivity in bronchial asthma. J Asthma 29, 245–252.
Kawasaki, Y., Kumamoto, E., Furue, H., Yoshimura, M. (2003). Alpha 2 adrenoceptor-mediated
presynaptic inhibition of primary afferent glutamatergic transmission in rat substantia gelatinosa
neurons. Anesthesiology 98, 682–689.
104
References
Kawasaki, Y., Zhang, L., Cheng, J.-K., Ji, R.-R. (2008). Cytokine mechanisms of central
sensitization: distinct and overlapping role of interleukin-1beta, interleukin-6, and tumor necrosis
factor-alpha in regulating synaptic and neuronal activity in the superficial spinal cord. J Neurosci
28, 5189–5194.
Khan, N., Bakshi, K.S., Jaggi, A.S., Singh, N. (2009). Ameliorative Potential of Spironolactone in
Diabetes Induced Hyperalgesia in Mice. Yakugaku Zasshi 129, 593–599.
Kharasch, E.D., Hoffer, C., Whittington, D., Sheffels, P. (2003). Role of P-glycoprotein in the
intestinal absorption and clinical effects of morphine. Clin Pharmacol Ther 74, 543–554.
Kieffer, B.L., Gavériaux-Ruff, C. (2002). Exploring the opioid system by gene knockout. Prog
Neurobiol 66, 285–306.
King, C.D., Rios, G.R., Green, M.D., MacKenzie, P.I., Tephly, T.R. (1997). Comparison of
stably expressed rat UGT1.1 and UGT2B1 in the glucuronidation of opioid compounds. Drug
Metab Dispos 25, 251–255.
King, M., Su, W., Chang, A., Zuckerman, A., Pasternak, G.W. (2001). Transport of opioids
from the brain to the periphery by P-glycoprotein: peripheral actions of central drugs. Nature
Neuroscience 4, 268–274.
Kishi, Y., Ohta, S., Kasuya, N., Tatsumi, M., Sawada, M., Sakita, S., Ashikaga, T., Numano, F.
(2000). Ibudilast modulates platelet-endothelium interaction mainly through cyclic GMPdependent mechanism. J Cardiovasc Pharmacol 36, 65–70.
Knights, K.M., Bowalgaha, K., Miners, J.O. (2010). Spironolactone and canrenone inhibit
UGT2B7-catalyzed human liver and kidney microsomal aldosterone 18beta-glucuronidation: a
potential drug interaction. Drug Metab Dispos 38, 1011–1014.
Koch, T., Höllt, V. (2008). Role of receptor internalization in opioid tolerance and dependence.
Pharmacol Ther 117, 199–206.
Koppert, W., Sittl, R., Scheuber, K., Alsheimer, M., Schmelz, M., Schüttler, J. (2003).
Differential modulation of remifentanil-induced analgesia and postinfusion hyperalgesia by Sketamine and clonidine in humans. Anesthesiology 99, 152–159.
Kwiat, G.C., Basbaum, A.I. (1992). The origin of brainstem noradrenergic and serotonergic
projections to the spinal cord dorsal horn in the rat. Somatosens Mot Res 9, 157–173.
Lagerström, M.C., Schiöth, H.B. (2008). Structural diversity of G protein-coupled receptors and
significance for drug discovery. Nat Rev Drug Discov 7, 339–357.
Lahtinen, P., Kokki, H., Hynynen, M. (2008). Remifentanil infusion does not induce opioid
tolerance after cardiac surgery. J Cardiothorac Vasc Anesth 22, 225–229.
Lakhlani, P.P., MacMillan, L.B., Guo, T.Z., McCool, B.A., Lovinger, D.M., Maze, M.,
Limbird, L.E. (1997). Substitution of a mutant alpha2a-adrenergic receptor via “hit and run” gene
targeting reveals the role of this subtype in sedative, analgesic, and anesthetic-sparing responses in
vivo. Proc Natl Acad Sci USA 94, 9950–9955.
Lalovic, B., Phillips, B., Risler, L.L., Howald, W. (2004). Quantitative contribution of CYP2D6
and CYP3A to oxycodone metabolism in human liver and intestinal microsomes. Drug Metab
Dispos. 32, 447–454.
Lane-Ladd, S.B., Pineda, J., Boundy, V.A., Pfeuffer, T., Krupinski, J., Aghajanian, G.K., Nestler,
E.J. (1997). CREB (cAMP response element-binding protein) in the locus coeruleus: biochemical,
physiological, and behavioral evidence for a role in opiate dependence. J Neurosci 17, 7890–7901.
105
References
Largent-Milnes, T.M., Guo, W., Wang, H.-Y., Burns, L.H., Vanderah, T.W. (2008).
Oxycodone plus ultra-low-dose naltrexone attenuates neuropathic pain and associated mu-opioid
receptor-Gs coupling. J Pain 9, 700–713.
Laskowski, K., Stirling, A., McKay, W.P., Lim, H.J. (2011). Revue méthodique de l’utilisation de
la kétamine intraveineuse pour l’analgésie postopératoire. Can J Anesth/J Can Anesth 58, 911–923.
Laugwitz, K.L., Offermanns, S., Spicher, K., Schultz, G. (1993). mu and delta opioid receptors
differentially couple to G protein subtypes in membranes of human neuroblastoma SH-SY5Y
cells. Neuron 10, 233–242.
Lawlor, P.G., Bruera, E. (1998). Side-effects of opioids in chronic pain treatment. Curr Opin
Anaesthesiol 11, 539–545.
Le Bars, D., Gozariu, M., Cadden, S.W. (2001). Animal models of nociception. Pharmacol Rev
53, 597–652.
Ledeboer, A., Hutchinson, M.R., Watkins, L.R., Johnson, K.W. (2007). Ibudilast (AV-411). A
new class therapeutic candidate for neuropathic pain and opioid withdrawal syndromes. Expert
Opin Investig Drugs 16, 935–950.
Ledeboer, A., Liu, T., Shumilla, J.A., Mahoney, J.H., Vijay, S., Gross, M.I., Vargas, J.A.,
Sultzbaugh, L., Claypool, M.D., Sanftner, L.M., Watkins, L.R., Johnson, K.W. (2006). The glial
modulatory drug AV411 attenuates mechanical allodynia in rat models of neuropathic pain.
Neuron Glia Biol 2, 279–291.
Lee, M., Silverman, S.M., Hansen, H., Patel, V.B., Manchikanti, L. (2011). A comprehensive
review of opioid-induced hyperalgesia. Pain Physician 14, 145–161.
Lemberg, K.K., Siiskonen, A.O., Kontinen, V.K., Yli-Kauhaluoma, J.T., Kalso, E.A. (2008).
Pharmacological characterization of noroxymorphone as a new opioid for spinal analgesia. Anesth
Analg 106, 463–470.
Leppert, W. (2010). The role of opioid receptor antagonists in the treatment of opioid-induced
constipation: a review. Adv Ther 27, 714–730.
Letrent, S.P., Pollack, G.M., Brouwer, K.R., Brouwer, K.L. (1998). Effect of GF120918, a potent
P-glycoprotein inhibitor, on morphine pharmacokinetics and pharmacodynamics in the rat. Pharm
Res 15, 599–605.
Letrent, S.P., Pollack, G.M., Brouwer, K.R., Brouwer, K.L. (1999). Effects of a potent and
specific P-glycoprotein inhibitor on the blood-brain barrier distribution and antinociceptive effect
of morphine in the rat. Drug Metab Dispos 27, 827–834.
Lewanowitsch, T., Irvine, R.J. (2003). Naloxone and its quaternary derivative, naloxone
methiodide, have differing affinities for µ, δ, and κ opioid receptors in mouse brain homogenates.
Brain Res 964, 302–305.
Lewis, S.S., Hutchinson, M.R., Rezvani, N., Loram, L.C., Zhang, Y., Maier, S.F., Rice, K.C.,
Watkins, L.R. (2010). Evidence that intrathecal morphine-3-glucuronide may cause pain
enhancement via toll-like receptor 4/MD-2 and interleukin-1beta. Neuroscience 165, 569–583.
Li, X., Angst, M.S., Clark, J.D. (2001a). A murine model of opioid-induced hyperalgesia. Brain
Res Mol Brain Res 86, 56–62.
Li, X., Angst, M.S., Clark, J.D. (2001b). Opioid-induced hyperalgesia and incisional pain. Anesth
Analg 93, 204–209.
106
References
Lim, G., Wang, S., Zeng, Q., Sung, B., Mao, J. (2005). Spinal glucocorticoid receptors contribute
to the development of morphine tolerance in rats. Anesthesiology 102, 832–837.
Lin, S.-L., Tsai, R.-Y., Shen, C.-H., Lin, F.-H., Wang, J.-J., Hsin, S.-T., Wong, C.-S. (2010).
Co-administration of ultra-low dose naloxone attenuates morphine tolerance in rats via
attenuation of NMDA receptor neurotransmission and suppression of neuroinflammation in the
spinal cords. Pharmacol Biochem Behav 96, 236–245.
Liu, W., Wang, C.-H., Cui, Y., Mo, L.-Q., Zhi, J.-L., Sun, S.-N., Wang, Y.-L., Yu, H.-M.,
Zhao, C.-M., Feng, J.-Q., Chen, P.-X. (2006). Inhibition of neuronal nitric oxide synthase
antagonizes morphine antinociceptive tolerance by decreasing activation of p38 MAPK in the
spinal microglia. Neurosci Lett 410, 174–177.
Luginbühl, M., Gerber, A., Schnider, T.W., Petersen-Felix, S., Arendt-Nielsen, L., Curatolo, M.
(2003). Modulation of remifentanil-induced analgesia, hyperalgesia, and tolerance by small-dose
ketamine in humans. Anesth Analg 96, 726–732.
Machelska, H., Pflüger, M., Weber, W., Piranvisseh-Völk, M., Daubert, J.D., Dehaven, R.,
Stein, C. (1999). Peripheral effects of the kappa-opioid agonist EMD 61753 on pain and
inflammation in rats and humans. J Pharmacol Exp Ther 290, 354–361.
Makman, M.H. (1994). Morphine receptors in immunocytes and neurons. Adv Neuroimmunol 4,
69–82.
Malaivijitnond, S., Varavudhi, P. (1998). Evidence for morphine-induced galactorrhea in male
cynomolgus monkeys. J Med Primatol 27, 1–9.
Manning, B.H., Mao, J., Frenk, H., Price, D.D., Mayer, D.J. (1996). Continuous coadministration of dextromethorphan or MK-801 with morphine: attenuation of morphine
dependence and naloxone-reversible attenuation of morphine tolerance. Pain 67, 79–88.
Mansikka, H., Idänpään-Heikkilä, J.J., Pertovaara, A. (1996). Different roles of α 2-adrenoceptors
of the medulla versus the spinal cord in modulation of mustard oil-induced central hyperalgesia in
rats. Eur J Pharmacol 297, 19–26.
Mansikka, H., Lähdesmäki, J., Scheinin, M., Pertovaara, A. (2004). Alpha(2A) adrenoceptors
contribute to feedback inhibition of capsaicin-induced hyperalgesia. Anesthesiology 101, 185–190.
Mao, J., Price, D.D., Mayer, D.J. (1994). Thermal hyperalgesia in association with the
development of morphine tolerance in rats: roles of excitatory amino acid receptors and protein
kinase C. J Neurosci 14, 2301–2312.
Mao, J., Sung, B., Ji, R.-R., Lim, G. (2002). Chronic morphine induces downregulation of spinal
glutamate transporters: implications in morphine tolerance and abnormal pain sensitivity. J
Neurosci 22, 8312–8323.
Massmann, V., Edemir, B., Schlatter, E., Al-Monajjed, R., Harrach, S., Klassen, P., Holle, S.K.,
Sindic, A., Dobrivojevic, M., Pavenstädt, H., Ciarimboli, G. (2013). The organic cation
transporter 3 (OCT3) as molecular target of psychotropic drugs: transport characteristics and acute
regulation of cloned murine OCT3. Pflugers Arch. doi:10.1007/s00404-013-1335-8.
Matsumiya, L.C., Sorge, R.E., Sotocinal, S.G., Tabaka, J.M., Wieskopf, J.S., Zaloum, A., King,
O.D., Mogil, J.S. (2012). Using the Mouse Grimace Scale to reevaluate the efficacy of
postoperative analgesics in laboratory mice. J Am Assoc Lab Anim Sci 51, 42–49.
Mattioli, T.-A.M., Milne, B., Cahill, C.M. (2010). Ultra-low dose naltrexone attenuates chronic
morphine-induced gliosis in rats. Mol Pain 6, 22.
107
References
McNaull, B., Trang, T., Sutak, M., Jhamandas, K. (2007). Inhibition of tolerance to spinal
morphine antinociception by low doses of opioid receptor antagonists. Eur J Pharmacol 560, 132–
141.
McNicol, E., Horowicz-Mehler, N., Fisk, R.A., Bennett, K., Gialeli-Goudas, M., Chew, P.W.,
Lau, J., Carr, D., Americal Pain Society (2003). Management of opioid side effects in cancerrelated and chronic noncancer pain: a systematic review. J Pain 4, 231–256.
Melzack, R., Wall, P.D. (1965). Pain mechanisms: a new theory. Science 150, 971–979.
Meng, X., Zhang, Y., Lao, L., Saito, R., Li, A., Bäckman, C.M., Berman, B.M., Ren, K., Wei,
P.-K., Zhang, R.-X. (2013). Spinal interleukin-17 promotes thermal hyperalgesia and NMDA
NR1 phosphorylation in an inflammatory pain rat model. Pain 154, 294–305.
Mercer, S.L., Coop, A. (2011). Opioid analgesics and P-glycoprotein efflux transporters: a
potential systems-level contribution to analgesic tolerance. Curr Top Med Chem 11, 1157–1164.
Mika, J. (2008). Modulation of microglia can attenuate neuropathic pain symptoms and enhance
morphine effectiveness. Pharmacol Rep. 60, 297–307.
Mika, J., Osikowicz, M., Makuch, W., Przewlocka, B. (2007). Minocycline and pentoxifylline
attenuate allodynia and hyperalgesia and potentiate the effects of morphine in rat and mouse
models of neuropathic pain. Eur J Pharmacol 560, 142–149.
Milligan, E.D., Soderquist, R.G., Malone, S.M., Mahoney, J.H., Hughes, T.S., Langer, S.J.,
Sloane, E.M., Maier, S.F., Leinwand, L.A., Watkins, L.R., Mahoney, M.J. (2006). Intrathecal
polymer-based interleukin-10 gene delivery for neuropathic pain. Neuron Glia Biol 2, 293–308.
Milne, B., Cervenko, F., Jhamandas, K., Loomis, C., Sutak, M. (1985). Analgesia and tolerance to
intrathecal morphine and norepinephrine infusion via implanted mini-osmotic pumps in the rat.
Pain 22, 165–172.
Milne, B., Jhamandas, K., Sutak, M., Grenier, P., Cahill, C.M. (2013). Stereo-selective inhibition
of spinal morphine tolerance and hyperalgesia by an ultra-low dose of the alpha-2-adrenoceptor
antagonist efaroxan. Eur J Pharmacol 702, 227–234.
Milne, B., Sutak, M., Cahill, C.M., Jhamandas, K. (2008). Low doses of alpha 2-adrenoceptor
antagonists augment spinal morphine analgesia and inhibit development of acute and chronic
tolerance. Br J Pharmacol 155, 1264–1278.
Milne, B., Sutak, M., Cahill, C.M., Jhamandas, K. (2011). Low dose alpha-2 antagonist
paradoxically enhances rat norepinephrine and clonidine analgesia. Anesth Analg 112, 1500–1503.
Minville, V., Fourcade, O., Girolami, J.-P., Tack, I. (2010). Opioid-induced hyperalgesia in a
mice model of orthopaedic pain: preventive effect of ketamine. Br J Anaesth 104, 231–238.
Mitchell, J.M., Basbaum, A.I., Fields, H.L. (2000). A locus and mechanism of action for
associative morphine tolerance. Nature Neuroscience 3, 47–53.
Miura, R., Nakamura, K., Miura, D., Miura, A., Hisamatsu, K., Kajiya, M., Nagase, S., Morita,
H., Fukushima Kusano, K., Ohe, T., Ishihara, K. (2006). Anti-inflammatory effect of
spironolactone on human peripheral blood mononuclear cells. J Pharmacol Sci 101, 256–259.
Mizuno, T., Kurotani, T., Komatsu, Y., Kawanokuchi, J., Kato, H., Mitsuma, N., Suzumura, A.
(2004). Neuroprotective role of phosphodiesterase inhibitor ibudilast on neuronal cell death
induced by activated microglia. Neuropharmacology 46, 404–411.
Mogil, J.S., Davis, K.D., Derbyshire, S.W. (2010). The necessity of animal models in pain
research. Pain 151, 12–17.
108
References
Montminy, M.R., Bilezikjian, L.M. (1987). Binding of a nuclear protein to the cyclic-AMP
response element of the somatostatin gene. Nature 328, 175–178.
Nakamura, T., Kakumoto, M., Yamashita, K., Takara, K., Tanigawara, Y., Sakaeda, T.,
Okumura, K. (2001). Factors influencing the prediction of steady state concentrations of digoxin.
Biol Pharm Bull 24, 403–408.
Neumann, P.B., Henriksen, H., Grosman, N., Christensen, C.B. (1982). Plasma morphine
concentrations during chronic oral administration in patients with cancer pain. Pain 13, 247–252.
Nevantaus, J., Simojoki, K., Hamunen, K., Heiskanen, T., Kalso, E. (2013). Opioidit
pitkäaikaisen kivun hoidossa. Suom Lääkäril 68, 3329–3335.
Nicotra, L., Loram, L.C., Watkins, L.R., Hutchinson, M.R. (2012). Toll-like receptors in chronic
pain. Exp Neurol 234, 316–329.
Nimmerjahn, A., Kirchhoff, F., Helmchen, F. (2005). Resting microglial cells are highly dynamic
surveillants of brain parenchyma in vivo. Science 308, 1314–1318.
Ninković, J., Roy, S. (2013). Role of the mu-opioid receptor in opioid modulation of immune
function. Amino Acids 45, 9–24.
Nishio, N., Taniguchi, W., Sugimura, Y.K., Takiguchi, N., Yamanaka, M., Kiyoyuki, Y.,
Yamada, H., Miyazaki, N., Yoshida, M., Nakatsuka, T. (2013). Reactive oxygen species enhance
excitatory synaptic transmission in rat spinal dorsal horn neurons by activating TRPA1 and
TRPV1 channels. Neuroscience 247, 201–212.
Obenaus, A. (2013). Neuroimaging biomarkers for epilepsy: advances and relevance to glial cells.
Neurochem Int 63, 712–718.
Ohashi, M., Ohkubo, H., Kito, J., Nishino, K. (1986). A new vasodilator 3-isobutyryl-2isopropylpyrazolo[1,5-a]pyridine (KC-404) has a dual mechanism of action on platelet
aggregation. Arch Int Pharmacodyn Ther 283, 321–334.
Olave, M.J., Maxwell, D.J. (2003). Neurokinin-1 projection cells in the rat dorsal horn receive
synaptic contacts from axons that possess alpha2C-adrenergic receptors. J Neurosci 23, 6837–6846.
Onoprishvili, I., Andria, M.L., Kramer, H.K. (2003). Interaction between the µ opioid receptor
and filamin A is involved in receptor regulation and trafficking. Mol Pharm 64, 1092–1100.
Osborne, R., Joel, S., Trew, D., Slevin, M. (1990). Morphine and metabolite behavior after
different routes of morphine administration: demonstration of the importance of the active
metabolite morphine-6-glucuronide. Clin Pharmacol Ther 47, 12–19.
Owen, J.A., Sitar, D.S., Berger, L., Brownell, L. (1983). Age-related morphine kinetics. Clin
Pharmacol Ther 34, 364–368.
Oye, I., Paulsen, O., Maurset, A. (1992). Effects of ketamine on sensory perception: evidence for a
role of N-methyl-D-aspartate receptors. J Pharmacol Exp Ther 260, 1209–1213.
Ozdoğan, U.K., Lähdesmäki, J., Scheinin, M. (2006). The analgesic efficacy of partial opioid
agonists is increased in mice with targeted inactivation of the alpha2A-adrenoceptor gene. Eur J
Pharmacol 529, 105–113.
Pan, Y.-Z., Li, D.-P., Pan, H.-L. (2002). Inhibition of glutamatergic synaptic input to spinal
lamina II(o) neurons by presynaptic alpha(2)-adrenergic receptors. J Neurophysiol 87, 1938–1947.
Pan, Z.Z., Williams, J.T., Osborne, P.B. (1990). Opioid actions on single nucleus raphe magnus
neurons from rat and guinea-pig in vitro. J Physiol (Lond) 427, 519–532.
109
References
Pasternak, G.W. (2007). When it comes to opiates, just say NO. J Clin Invest 117, 3185–3187.
Pattinson, K.T.S. (2008). Opioids and the control of respiration. Br J Anaesth 100, 747–758.
Paul, D., Standifer, K.M., Inturrisi, C.E. (1989). Pharmacological characterization of morphine-6
beta-glucuronide, a very potent morphine metabolite. J Pharm Exp Ther 251, 477–483.
Pert, C.B., Snyder, S.H. (1973). Opiate receptor: demonstration in nervous tissue. Science 179,
1011–1014.
Pertovaara, A. (1993). Antinociception induced by alpha-2-adrenoceptor agonists, with special
emphasis on medetomidine studies. Prog Neurobiol 40, 691–709.
Pertovaara, A., Haapalinna, A., Sirviö, J., Virtanen, R. (2005). Pharmacological properties, central
nervous system effects, and potential therapeutic applications of atipamezole, a selective alpha2adrenoceptor antagonist. CNS Drug Rev 11, 273–288.
Pitkänen, A., Narkilahti, S., Bezvenyuk, Z., Haapalinna, A., Nissinen, J. (2004). Atipamezole, an
alpha(2)-adrenoceptor antagonist, has disease modifying effects on epileptogenesis in rats. Epilepsy
Res 61, 119–140.
Polt, R., Porreca, F., Szabò, L.Z., Bilsky, E.J., Davis, P., Abbruscato, T.J., Davis, T.P., Harvath,
R., Yamamura, H.I., Hruby, V.J. (1994). Glycopeptide enkephalin analogues produce analgesia in
mice: evidence for penetration of the blood-brain barrier. Proc Natl Acad Sci USA 91, 7114–7118.
Porreca, F., Ossipov, M.H., Gebhart, G.F. (2002). Chronic pain and medullary descending
facilitation. Trends Neurosci 25, 319–325.
Powell, K.J., Abul-Husn, N.S., Jhamandas, A., Olmstead, M.C., Beninger, R.J., Jhamandas, K.
(2002). Paradoxical effects of the opioid antagonist naltrexone on morphine analgesia, tolerance,
and reward in rats. J Pharmacol Exp Ther 300, 588–596.
Pritchard, M., Fournel-Gigleux, S., Siest, G., Mackenzie, P., Magdalou, J. (1994). A recombinant
phenobarbital-inducible rat liver UDP-glucuronosyltransferase (UDP-glucuronosyltransferase
2B1) stably expressed in V79 cells catalyzes the glucuronidation of morphine, phenols, and
carboxylic acids. Mol Pharmacol 45, 42–50.
Projean, D., Morin, P.-E., Tu, T.M., Ducharme, J. (2003). Identification of CYP3A4 and
CYP2C8 as the major cytochrome P450 s responsible for morphine N-demethylation in human
liver microsomes. Xenobiotica 33, 841–854.
Prosser, J.M., Steinfeld, M., Cohen, L.J., Derbyshire, S., Eisenberg, D.P., Cruciani, R.A.,
Galynker, I.I. (2008). Abnormal heat and pain perception in remitted heroin dependence months
after detoxification from methadone-maintenance. Drug Alcohol Depend 95, 237–244.
Qi, X., Evans, A.M., Wang, J., Miners, J.O., Upton, R.N., Milne, R.W. (2010). Inhibition of
morphine metabolism by ketamine. Drug Metab Dispos 38, 728–731.
Quartilho, A., Mata, H.P., Ibrahim, M.M., Vanderah, T.W., Ossipov, M.H., Lai, J., Porreca, F.,
Malan, T.P. (2004). Production of paradoxical sensory hypersensitivity by alpha 2-adrenoreceptor
agonists. Anesthesiology 100, 1538–1544.
Raehal, K.M., Schmid, C.L., Groer, C.E., Bohn, L.M. (2011). Functional selectivity at the µopioid receptor: implications for understanding opioid analgesia and tolerance. Pharmacol Rev 63,
1001–1019.
Raghavendra, V., Rutkowski, M.D., DeLeo, J.A. (2002). The role of spinal neuroimmune
activation in morphine tolerance/hyperalgesia in neuropathic and sham-operated rats. J Neurosci
22, 9980–9989.
110
References
Raghavendra, V., Tanga, F.Y., DeLeo, J.A. (2004). Attenuation of morphine tolerance,
withdrawal-induced hyperalgesia, and associated spinal inflammatory immune responses by
propentofylline in rats. Neuropsychopharmacology 29, 327–334.
Rahman, W., Dashwood, M.R., Fitzgerald, M., Aynsley-Green, A., Dickenson, A.H. (1998).
Postnatal development of multiple opioid receptors in the spinal cord and development of spinal
morphine analgesia. Brain Res Dev Brain Res 108, 239–254.
Rajaofetra, N., Ridet, J.L., Poulat, P., Marlier, L., Sandillon, F., Geffard, M., Privat, A. (1992).
Immunocytochemical mapping of noradrenergic projections to the rat spinal cord with an
antiserum against noradrenaline. J Neurocytol 21, 481–494.
Randall, L.O., Selitto, J.J. (1957). A method for measurement of analgesic activity on inflamed
tissue. Arch Int Pharmacodyn Ther 111, 409–419.
Rao, V.V., Dahlheimer, J.L., Bardgett, M.E., Snyder, A.Z., Finch, R.A., Sartorelli, A.C.,
Piwnica-Worms, D. (1999). Choroid plexus epithelial expression of MDR1 P glycoprotein and
multidrug resistance-associated protein contribute to the blood-cerebrospinal-fluid drugpermeability barrier. Proceedings of the National Academy of Sciences 96, 3900–3905.
Reddy, S.V., Yaksh, T.L. (1980). Spinal noradrenergic terminal system mediates antinociception.
Brain Res 189, 391–401.
Renouard, A., Widdowson, P.S., Millan, M.J. (1994). Multiple alpha 2 adrenergic receptor
subtypes. I. Comparison of [3H]RX821002-labeled rat R alpha-2A adrenergic receptors in
cerebral cortex to human H alpha2A adrenergic receptor and other populations of alpha-2
adrenergic subtypes. J Pharmacol Exp Ther 270, 946–957.
Rile, G., Yatomi, Y., Qi, R., Satoh, K., Ozaki, Y. (2001). Potentiation of ibudilast inhibition of
platelet aggregation in the presence of endothelial cells. Thromb Res 102, 239–246.
Risdahl, J.M., Khanna, K.V., Peterson, P.K. (1998). Opiates and infection. J Neuroimmunol 83, 4–
18.
Rizvi, T.A., Murphy, A.Z., Ennis, M., Behbehani, M.M., Shipley, M.T. (1996). Medial preoptic
area afferents to periaqueductal gray medullo-output neurons: a combined Fos and tract tracing
study. J Neurosci 16, 333–344.
Roerig, S.C., Lei, S., Kitto, K., Hylden, J.K., Wilcox, G.L. (1992). Spinal interactions between
opioid and noradrenergic agonists in mice: multiplicativity involves delta and alpha-2 receptors. J
Pharmacol Exp Ther 262, 365–374.
Rolan, P., Gibbons, J.A., He, L., Chang, E., Jones, D., Gross, M.I., Davidson, J.B., Sanftner,
L.M., Johnson, K.W. (2008). Ibudilast in healthy volunteers: safety, tolerability and
pharmacokinetics with single and multiple doses. Br J Clin Pharmacol 66, 792–801.
Rolandi, E., Marabini, A., Franceschini, R., Messina, V., Bongera, P., Barreca, T. (1983).
Changes in pituitary secretion induced by an agonist-antagonist opioid drug, buprenorphine. Acta
Endocrinol 104, 257–260.
Roy, S., Loh, H.H. (1996). Effects of opioids on the immune system. Neurochem Res 21, 1375–
1386.
Roy, S., Wang, J., Kelschenbach, J., Koodie, L. (2006). Modulation of immune function by
morphine: implications for susceptibility to infection. J Neuroimmune Pharmacol 1, 77–89.
Sacerdote, P., Manfredi, B., Mantegazza, P., Panerai, A.E. (1997). Antinociceptive and
immunosuppressive effects of opiate drugs: a structure-related activity study. Br J Pharmacol 121,
834–840.
111
References
Saegusa, H., Kurihara, T., Zong, S., Minowa, O., Kazuno, A., Han, W., Matsuda, Y., Yamanaka,
H., Osanai, M., Noda, T., Tanabe, T. (2000). Altered pain responses in mice lacking alpha 1E
subunit of the voltage-dependent Ca2+ channel. Proc Natl Acad Sci USA 97, 6132–6137.
Sai, K., Itoda, M., Saito, Y., Kurose, K., Katori, N., Kaniwa, N., Komamura, K., Kotake, T.,
Morishita, H., Tomoike, H., Kamakura, S., Kitakaze, M., Tamura, T., Yamamoto, N., Kunitoh,
H., Yamada, Y., Ohe, Y., Shimada, Y., Shirao, K., Minami, H., Ohtsu, A., Yoshida, T., Saijo, N.,
Kamatani, N., Ozawa, S., Sawada, J. (2006). Genetic variations and haplotype structures of the
ABCB1 gene in a Japanese population: an expanded haplotype block covering the distal promoter
region, and associated ethnic differences. Ann Hum Genet 70, 605–622.
Scheinin, H., MacDonald, E., Scheinin, M. (1988). Behavioural and neurochemical effects of
antipamezole, a novel alpha 2-adrenoceptor antagonist. Eur J Pharmacol 151, 35–42.
Scheinin, M., Lomasney, J.W., Hayden-Hixson, D.M., Schambra, U.B., Caron, M.G.,
Lefkowitz, R.J., Fremeau, R.T. (1994). Distribution of alpha 2-adrenergic receptor subtype gene
expression in rat brain. Brain Res Mol Brain Res 21, 133–149.
Scheller, M., Bufler, J., Hertle, I., Schneck, H.J., Franke, C., Kochs, E. (1996). Ketamine blocks
currents through mammalian nicotinic acetylcholine receptor channels by interaction with both
the open and the closed state. Anesth Analg 83, 830–836.
Schinkel, A.H., Wagenaar, E., van Deemter, L., Mol, C.A., Borst, P. (1995). Absence of the
mdr1a P-Glycoprotein in mice affects tissue distribution and pharmacokinetics of dexamethasone,
digoxin, and cyclosporin A. J Clin Invest 96, 1698–1705.
Schmidt, S., Bethge, C., Förster, M.H., Schäfer, M. (2007). Enhanced postoperative sensitivity to
painful pressure stimulation after intraoperative high dose remifentanil in patients without
significant surgical site pain. Clin J Pain 23, 605–611.
Schraag, S., Checketts, M.R., Kenny, G.N. (1999). Lack of rapid development of opioid tolerance
during alfentanil and remifentanil infusions for postoperative pain. Anesth Analg 89, 753–757.
Schuetz, E.G., Beck, W.T., Schuetz, J.D. (1996). Modulators and substrates of P-glycoprotein
and cytochrome P4503A coordinately up-regulate these proteins in human colon carcinoma cells.
Mol Pharmacol 49, 311–318.
Schug, S.A., Zech, D., Grond, S. (1992). Adverse effects of systemic opioid analgesics. Drug Saf
7, 200–213.
Sebastiani, G., Morissette, C., Lagacé, C., Boulé, M., Ouellette, M.-J., McLaughlin, R.W.,
Lacombe, D., Gervais, F., Tremblay, P. (2006). The cAMP-specific phosphodiesterase 4B
mediates Abeta-induced microglial activation. Neurobiol Aging 27, 691–701.
Sharma, H.S., Ali, S.F. (2006). Alterations in Blood-Brain Barrier Function by Morphine and
Methamphetamine. Ann. N. Y. Acad. Sci. 1074, 198–224.
Sharma, S.K., Klee, W.A., Nirenberg, M. (1975). Dual regulation of adenylate cyclase accounts
for narcotic dependence and tolerance. Proc Natl Acad Sci USA 72, 3092–3096.
Shen, C.-H., Tsai, R.-Y., Shih, M.-S., Lin, S.-L., Tai, Y.-H., Chien, C.-C., Wong, C.-S.
(2011). Etanercept restores the antinociceptive effect of morphine and suppresses spinal
neuroinflammation in morphine-tolerant rats. Anesth Analg 112, 454–459.
Shi, T.J., Winzer-Serhan, U., Leslie, F., Hökfelt, T. (1999). Distribution of alpha2-adrenoceptor
mRNAs in the rat lumbar spinal cord in normal and axotomized rats. Neuroreport 10, 2835–2839.
Shi, T.S., Winzer-Serhan, U., Leslie, F., Hökfelt, T. (2000). Distribution and regulation of
alpha(2)-adrenoceptors in rat dorsal root ganglia. Pain 84, 319–330.
112
References
Shimoyama, M., Shimoyama, N., Gorman, A.L., Elliott, K.J., Inturrisi, C.E. (1999). Oral
ketamine is antinociceptive in the rat formalin test: role of the metabolite, norketamine. Pain 81,
85–93.
Shimoyama, N., Shimoyama, M., Davis, A.M., Monaghan, D.T., Inturrisi, C.E. (2005). An
antisense oligonucleotide to the N-methyl-D-aspartate (NMDA) subunit NMDAR1 attenuates
NMDA-induced nociception, hyperalgesia, and morphine tolerance. J Pharmacol Exp Ther 312,
834–840.
Shin, S.-W., Cho, A.-R., Lee, H.-J., Kim, H.-J., Byeon, G.J., Yoon, J.-W., Kim, K.-H., Kwon,
J.-Y. (2010). Maintenance anaesthetics during remifentanil-based anaesthesia might affect
postoperative pain control after breast cancer surgery. Br J Anaesth 105, 661–667.
Shinohara, Y., Kusunoki, T. (2002). Clinical Efficacy of Ibudilast on Vertigo and Dizziness
Observed in Patients with Cerebral Infarction in the Chronic Stage. A Double-blind,
Randomized, Placebo-controlled Trial with a Run-in Period. Shinnkeitiryougaku (Neurol Ther) 19,
177–187.
Sikand, K.S., Smith, G., Lambert, D.G. (1995). Ketamine inhibits K+ evoked [3H] noradrenaline
release from SH-SY5Y cells by reducing calcium influx. Biochem Soc Trans 23, 417S.
Sinner, B., Graf, B.M. (2008). Ketamine. In Handb Exp Pharmacol, J. Schuttler and H. Schwilden,
eds. (Springer).
Sjøgren, P., Jensen, N.H., Jensen, T.S. (1994). Disappearance of morphine-induced hyperalgesia
after discontinuing or substituting morphine with other opioid agonists. Pain 59, 313–316.
Skagerberg, G., Björklund, A. (1985). Topographic principles in the spinal projections of
serotonergic and non-serotonergic brainstem neurons in the rat. Neuroscience 15, 445–480.
Skarke, C., Jarrar, M., Erb, K., Schmidt, H., Geisslinger, G., Lötsch, J. (2003). Respiratory and
miotic effects of morphine in healthy volunteers when P-glycoprotein is blocked by quinidine. Clin
Pharmacol Ther 74, 303–311.
Smith, D.J., Pekoe, G.M., Martin, L.L., Coalgate, B. (1980). The interaction of ketamine with
the opiate receptor. Life Sci 26, 789–795.
Smith, M.T. (2000). Neuroexcitatory effects of morphine and hydromorphone: evidence
implicating the 3‐glucuronide metabolites. Clin Exp Pharmacol Physiol 27, 524–528.
Song, P., Zhao, Z.Q. (2001). The involvement of glial cells in the development of morphine
tolerance. Neurosci Res 39, 281–286.
Sonohata, M., Furue, H., Katafuchi, T., Yasaka, T., Doi, A., Kumamoto, E., Yoshimura, M.
(2004). Actions of noradrenaline on substantia gelatinosa neurones in the rat spinal cord revealed
by in vivo patch recording. J Physiol (Lond) 555, 515–526.
Souness, J.E., Villamil, M.E., Scott, L.C., Tomkinson, A., Giembycz, M.A., Raeburn, D. (1994).
Possible role of cyclic AMP phosphodiesterases in the actions of ibudilast on eosinophil
thromboxane generation and airways smooth muscle tone. Br J Pharmacol 111, 1081–1088.
Stafford, K., Gomes, A.B., Shen, J., Yoburn, B.C. (2001). mu-Opioid receptor downregulation
contributes to opioid tolerance in vivo. Pharmacol Biochem Behav 69, 233–237.
Stellwagen, D., Beattie, E.C., Seo, J.Y., Malenka, R.C. (2005). Differential regulation of AMPA
receptor and GABA receptor trafficking by tumor necrosis factor-alpha. J Neurosci 25, 3219–3228.
Stevens, C.W., Monasky, M.S., Yaksh, T.L. (1988). Spinal infusion of opiate and alpha-2
agonists in rats: tolerance and cross-tolerance studies. J Pharmacol Exp Ther 244, 63–70.
113
References
Stone, L.S., Broberger, C., Vulchanova, L., Wilcox, G.L., Hökfelt, T., Riedl, M.S., Elde, R.
(1998). Differential distribution of alpha2A and alpha2C adrenergic receptor immunoreactivity in
the rat spinal cord. J Neurosci 18, 5928–5937.
Stone, L.S., MacMillan, L.B., Kitto, K.F., Limbird, L.E., Wilcox, G.L. (1997). The alpha2a
adrenergic receptor subtype mediates spinal analgesia evoked by alpha2 agonists and is necessary
for spinal adrenergic-opioid synergy. J Neurosci 17, 7157–7165.
Strauss, W.L., Layton, M.E., Dager, S.R. (1998). Brain elimination half-life of fluvoxamine
measured by 19F magnetic resonance spectroscopy. Am J Psychiatry 155, 380–384.
Su, W., Pasternak, G.W. (2013). The role of multidrug resistance-associated protein in the bloodbrain barrier and opioid analgesia. Synapse 67, 609–619.
Sullivan, A.F., Kalso, E.A., McQuay, H.J., Dickenson, A.H. (1992). The antinociceptive actions
of dexmedetomidine on dorsal horn neuronal responses in the anaesthetized rat. Eur J Pharmacol
215, 127–133.
Sun, Y.-E., Peng, L., Sun, X., Bo, J., Yang, D., Zheng, Y., Liu, C., Zhu, B., Ma, Z., Gu, X.
(2012). Intrathecal injection of spironolactone attenuates radicular pain by inhibition of spinal
microglia activation in a rat model. PLoS ONE 7, e39897.
Sung, B., Lim, G., Mao, J. (2003). Altered expression and uptake activity of spinal glutamate
transporters after nerve injury contribute to the pathogenesis of neuropathic pain in rats. J Neurosci
23, 2899–2910.
Suppa, E., Valente, A., Catarci, S., Zanfini, B., Draisci, G. (2012). A study of low-dose SKetamine infusion as “preventive” pain treatment for cesarean section with spinal anesthesia:
benefits and side effects. Minerva Anestesiol 78, 774–81.
Suter, M.R., Wen, Y.-R., Decosterd, I., Ji, R.-R. (2007). Do glial cells control pain? Neuron Glia
Biol 3, 255–268.
Suzumura, A., Ito, A., Yoshikawa, M., Sawada, M. (1999). Ibudilast suppresses TNFα production
by glial cells functioning mainly as type III phosphodiesterase inhibitor in the CNS. Brain Res
837, 203–212.
Syngle, A., Vohra, K., Kaur, L., Sharma, S. (2009). Effect of spironolactone on endothelial
dysfunction in rheumatoid arthritis. Scand J Rheumatol 38, 15–22.
Tai, Y.-H., Wang, Y.-H., Wang, J.-J., Tao, P.-L., Tung, C.-S., Wong, C.-S. (2006).
Amitriptyline suppresses neuroinflammation and up-regulates glutamate transporters in
morphine-tolerant rats. Pain 124, 77–86.
Tai, Y.H., Tsai, R.Y., Lin, S.L., Yeh, C.C., Wang, J.J. (2009). Amitriptyline suppresses
neuroinflammation-dependent interleukin-10-p38 mitogen-activated protein kinase-heme
oxygenase-1 signaling pathway in chronic morphine-infused rats. Anesthesiology 110, 1379–1389.
Takano, Y., Yaksh, T.L. (1992). Characterization of the pharmacology of intrathecally
administered alpha-2 agonists and antagonists in rats. J Pharmacol Exp Ther 261, 764–772.
Tanaka, J., Fujita, H., Matsuda, S., Toku, K., Sakanaka, M., Maeda, N. (1997). Glucocorticoidand mineralocorticoid receptors in microglial cells: the two receptors mediate differential effects of
corticosteroids. Glia 20, 23–37.
Taylor, F., Dickenson, A. (1998). Nociceptin/orphanin FQ. A new opioid, a new analgesic?
Neuroreport 9, R65–R70.
114
References
Terner, J.M., Barrett, A.C., Lomas, L.M., Negus, S.S., Picker, M.J. (2006). Influence of low
doses of naltrexone on morphine antinociception and morphine tolerance in male and female rats
of four strains. Pain 122, 90–101.
Thompson, S.J., Koszdin, K., Bernards, C.M. (2000). Opiate-induced Analgesia Is Increased and
Prolonged in Mice Lacking P-glycoprotein. Anesthesiology 92, 1392.
Todd, A.J. (2010). Neuronal circuitry for pain processing in the dorsal horn. Nat Rev Neurosci 11,
823–836.
Todd, A.J., Koerber, R.H. (2013). Neuroanatomical Substrates of Spinal Nociception. In Wall and
Melzack's Textbook of Pain, S.B. McMahon, M. Koltzenburg, I. Tracey, and D.C. Turk, eds.
(Elsevier Saunders).
Toh, M.-L., Aeberli, D., Lacey, D., Yang, Y., Santos, L.L., Clarkson, M., Sharma, L., Clyne, C.,
Morand, E.F. (2006). Regulation of IL-1 and TNF receptor expression and function by
endogenous macrophage migration inhibitory factor. J Immunol 177, 4818–4825.
Torebjörk, E., Wahren, L., Wallin, G., Hallin, R., Koltzenburg, M. (1995). Noradrenalineevoked pain in neuralgia. Pain 63, 11–20.
Tournier, N., Chevillard, L., Megarbane, B., Pirnay, S., Scherrmann, J.-M., Declèves, X. (2010).
Interaction of drugs of abuse and maintenance treatments with human P-glycoprotein (ABCB1)
and breast cancer resistance protein (ABCG2). Int J Neuropsychopharmacol 13, 905–915.
Tröster, A., Sittl, R., Singler, B., Schmelz, M., Schüttler, J., Koppert, W. (2006). Modulation of
remifentanil-induced analgesia and postinfusion hyperalgesia by parecoxib in humans.
Anesthesiology 105, 1016–1023.
Trujillo, K.A., Akil, H. (1991). Inhibition of morphine tolerance and dependence by the NMDA
receptor antagonist MK-801. Science 251, 85–87.
Trujillo, K.A., Akil, H. (1994). Inhibition of opiate tolerance by non-competitive N-d-aspartate
receptor antagonists. Brain Res 633, 178–188.
Tsai, R.-Y., Jang, F.-L., Tai, Y.-H., Lin, S.-L., Shen, C.-H., Wong, C.-S. (2008). Ultra-LowDose Naloxone Restores the Antinociceptive Effect of Morphine and Suppresses Spinal
Neuroinflammation in PTX-Treated Rats. Neuropsychopharmacology 33, 2772–2782.
Tukey, R.H., Strassburg, C.P. (2000). Human UDP-glucuronosyltransferases: metabolism,
expression, and disease. Annu Rev Pharmacol Toxicol 40, 581–616.
Tzvetkov, M.V., Santos Pereira, dos, J.N., Meineke, I., Saadatmand, A.R., Stingl, J.C.,
Brockmöller, J. (2013). Morphine is a substrate of the organic cation transporter OCT1 and
polymorphisms in OCT1 gene affect morphine pharmacokinetics after codeine administration.
Biochem Pharmacol 86, 666–678.
Uchaipichat, V., Raungrut, P., Chau, N., Janchawee, B., Evans, A.M., Miners, J.O. (2011).
Effects of ketamine on human UDP-glucuronosyltransferases in vitro predict potential drug-drug
interactions arising from ketamine inhibition of codeine and morphine glucuronidation. Drug
Metab Dispos 39, 1324–1328.
Ullrich, K.J., Rumrich, G., David, C., Fritzsch, G. (1993). Bisubstrates: substances that interact
with renal contraluminal organic anion and organic cation transport systems. I. Amines,
piperidines, piperazines, azepines, pyridines, quinolines, imidazoles, thiazoles, guanidines and
hydrazines. Pflugers Arch 425, 280–299.
Vainio, O., Vähä-Vahe, T. (1990). Reversal of medetomidine sedation by atipamezole in dogs. J
Vet Pharmacol Ther 13, 15–22.
115
References
Vanderah, T.W., Suenaga, N.M., Ossipov, M.H., Malan, T.P., Lai, J., Porreca, F. (2001). Tonic
descending facilitation from the rostral ventromedial medulla mediates opioid-induced abnormal
pain and antinociceptive tolerance. J Neurosci 21, 279–286.
Varma, M.V.S., Sateesh, K., Panchagnula, R. (2005). Functional Role of P-Glycoprotein in
Limiting Intestinal Absorption of Drugs: Contribution of Passive Permeability to P-Glycoprotein
Mediated Efflux Transport. Mol Pharmaceutics 2, 12–21.
Veilleux-Lemieux, D., Beaudry, F., Hélie, P., Vachon, P. (2012). Effects of endotoxemia on the
pharmacodynamics and pharmacokinetics of ketamine and xylazine anesthesia in Sprague–Dawley
rats. Vet Med Res Rep 3, 99–109.
Vilardaga, J.-P., Nikolaev, V.O., Lorenz, K., Ferrandon, S., Zhuang, Z., Lohse, M.J. (2008).
Conformational cross-talk between alpha2A-adrenergic and mu-opioid receptors controls cell
signaling. Nat Chem Biol 4, 126–131.
Vinik, H.R., Kissin, I. (1998). Rapid development of tolerance to analgesia during remifentanil
infusion in humans. Anesth Analg 86, 1307–1311.
Virtanen, R. (1989). Pharmacological profiles of medetomidine and its antagonist, atipamezole.
Acta Vet Scand Suppl 85, 29–37.
Visser, E., Schug, S.A. (2006). The role of ketamine in pain management. Biomed Pharmacother
60, 341–348.
Viviani, B., Bartesaghi, S., Gardoni, F., Vezzani, A., Behrens, M.M., Bartfai, T., Binaglia, M.,
Corsini, E., Di Luca, M., Galli, C.L., Marinovich, M. (2003). Interleukin-1beta enhances
NMDA receptor-mediated intracellular calcium increase through activation of the Src family of
kinases. J Neurosci 23, 8692–8700.
Wacher, V.J., Wu, C.Y., Benet, L.Z. (1995). Overlapping substrate specificities and tissue
distribution of cytochrome P450 3A and P-glycoprotein: implications for drug delivery and
activity in cancer chemotherapy. Mol Carcinog 13, 129–134.
Walsh, C.T., Levine, R.R. (1975). Studies of the enterohepatic circulation of morphine in the rat.
J Pharmacol Exp Ther 195, 303–310.
Wandel, C., Kim, R., Wood, M., Wood, A. (2002). Interaction of morphine, fentanyl, sufentanil,
alfentanil, and loperamide with the efflux drug transporter P-glycoprotein. Anesthesiology 96, 913–
920.
Wang, H.-Y., Friedman, E., Olmstead, M.C., Burns, L.H. (2005). Ultra-low-dose naloxone
suppresses opioid tolerance, dependence and associated changes in mu opioid receptor-G protein
coupling and Gbetagamma signaling. Neuroscience 135, 247–261.
Wang, H.-Y., Frankfurt, M., Burns, L.H. (2008). High-affinity naloxone binding to filamin a
prevents mu opioid receptor-Gs coupling underlying opioid tolerance and dependence. PLoS ONE
3, e1554.
Wang, H.-Y., Bakshi, K., Frankfurt, M., Stucky, A., Goberdhan, M., Shah, S.M., Burns, L.H.
(2012a). Reducing amyloid-related Alzheimer's disease pathogenesis by a small molecule targeting
filamin A. J Neurosci 32, 9773–9784.
Wang, S., Lim, G., Zeng, Q., Sung, B., Ai, Y., Guo, G., Yang, L., Mao, J. (2004). Expression of
central glucocorticoid receptors after peripheral nerve injury contributes to neuropathic pain
behaviors in rats. J Neurosci 24, 8595–8605.
Wang, X., Loram, L.C., Ramos, K., de Jesus, A.J., Thomas, J., Cheng, K., Reddy, A., Somogyi,
A.A., Hutchinson, M.R., Watkins, L.R., Yin, H. (2012b). Morphine activates neuroinflammation
116
References
in a manner parallel to endotoxin. Proc Natl Acad Sci USA 109, 6325–6330.
Wang, Z., Bilsky, E.J., Porreca, F., Sadée, W. (1994). Constitutive mu opioid receptor activation as a
regulatory mechanism underlying narcotic tolerance and dependence. Life Sci 54, PL339–PL350.
Watkins, L.R., Hutchinson, M.R., Rice, K.C., Maier, S.F. (2009). The “toll” of opioid-induced
glial activation: improving the clinical efficacy of opioids by targeting glia. Trends Pharmacol Sci 30,
581–591.
Wen, Y.-R., Suter, M.R., Ji, R.-R., Yeh, G.-C., Wu, Y.-S., Wang, K.-C., Kohno, T., Sun, W.Z., Wang, C.-C. (2009). Activation of p38 mitogen-activated protein kinase in spinal microglia
contributes to incision-induced mechanical allodynia. Anesthesiology 110, 155–165.
Wernze, H., Herdegen, T. (2014). Long-term efficacy of spironolactone on pain, mood, and
quality of life in women with fibromyalgia: An observational case series. Scandinavian Journal of
Pain 5, 63–71.
Westfall, T.C., Westfall, D.P. (2011). Neurotransmission: The Autonomic and Somatic Motor
Nervous Systems. In Goodman & Gilman's the Pharmacological Basis of Therapeutics, L. Brunton, B.
Chabner, and B. Knollman, eds. (McGraw-Hill Medical).
Whistler, J.L., Zastrow, von, M. (1998). Morphine-activated opioid receptors elude
desensitization by beta-arrestin. Proc Natl Acad Sci USA 95, 9914–9919.
White, P.F., Ham, J., Way, W.L., Trevor, A.J. (1980). Pharmacology of ketamine isomers in
surgical patients. Anesthesiology 52, 231–239.
White, P.F. (1988). Clinical pharmacology of intravenous induction drugs. Int Anesthesiol Clin 26,
98–104.
Williams, J.T., Ingram, S.L., Henderson, G., Chavkin, C., Zastrow, von, M., Schulz, S., Koch,
T., Evans, C.J., Christie, M.J. (2013). Regulation of µ-opioid receptors: desensitization,
phosphorylation, internalization, and tolerance. Pharmacol Rev 65, 223–254.
Woolfe, G., Macdonald, A.D. (1944). The evaluation of the analgesic action of pethidine
hydrochloride (Demerol). J Pharmacol Exp Ther 80, 300–307.
Wu, H., Wacker, D., Mileni, M., Katritch, V., Han, G.W., Vardy, E., Liu, W., Thompson,
A.A., Huang, X.-P., Carroll, F.I., Mascarella, S.W., Westkaemper, R.B., Mosier, P.D., Roth,
B.L., Cherezov, V., Stevens, R.C. (2012). Structure of the human κ-opioid receptor in complex
with JDTic. Nature 485, 327–332.
Xie, J.Y., Herman, D.S., Stiller, C.-O., Gardell, L.R., Ossipov, M.H., Lai, J., Porreca, F.,
Vanderah, T.W. (2005). Cholecystokinin in the rostral ventromedial medulla mediates opioidinduced hyperalgesia and antinociceptive tolerance. J Neurosci 25, 409–416.
Xie, R., Hammarlund-Udenaes, M., de Boer, A.G., de Lange, E.C. (1999). The role of Pglycoprotein in blood-brain barrier transport of morphine: transcortical microdialysis studies in
mdr1a (-/-) and mdr1a (+/+) mice. Br J Pharmacol 128, 563–568.
Xie, R., Bouw, M.R., Hammarlund-Udenaes, M. (2000). Modelling of the blood-brain barrier
transport of morphine-3-glucuronide studied using microdialysis in the rat: involvement of
probenecid-sensitive transport. Br J Pharmacol 131, 1784–1792.
Yaksh, T.L., Rudy, T.A. (1976). Chronic catheterization of the spinal subarachnoid space. Physiol
Behav 17, 1031–1036.
Yaksh, T.L., Al-Rodhan, N.R., Jensen, T.S. (1988). Sites of action of opiates in production of
analgesia. Prog Brain Res 77, 371–394.
117
References
Yaksh, T.L. (1997). Pharmacology and mechanisms of opioid analgesic activity. Acta Anaesthesiol
Scand 41, 94–111.
Yaksh, T.L., Wallace, M.S. (2011). Opioids, Analgesia, and Pain Management. In Goodman &
Gilman's the Pharmacological Basis of Therapeutics, L. Brunton, B. Chabner, and B. Knollman, eds.
(McGraw-Hill Medical).
Yamada, H., Ishii, K., Ishii, Y., Ieiri, I., Nishio, S., Morioka, T., Oguri, K. (2003). Formation of
highly analgesic morphine-6-glucuronide following physiologic concentration of morphine in
human brain. J Toxicol Sci 28, 395–401.
Yan, M., Dai, H., Ding, T., Dai, A., Zhang, F., Yu, L., Chen, G., Chen, Z. (2011). Effects of
dexmedetomidine on the release of glial cell line-derived neurotrophic factor from rat astrocyte
cells. Neurochem Int 58, 549–557.
Yan, X., Weng, H.-R. (2013). Endogenous interleukin-1β in neuropathic rats enhances glutamate
release from the primary afferents in the spinal dorsal horn through coupling with presynaptic Nmethyl-D-aspartic acid receptors. J Biol Chem 288, 30544–30557.
Yanagihara, Y., Kariya, S., Ohtani, M., Uchino, K., Aoyama, T., Yamamura, Y., Iga, T. (2001).
Involvement of CYP2B6 in n-demethylation of ketamine in human liver microsomes. Drug Metab
Dispos 29, 887–890.
Yoshimura, H., Ida, S., Oguri, K., Tsukamoto, H. (1973). Biochemical basis for analgesic activity
of morphine-6-glucuronide. I. Penetration of morphine-6-glucuronide in the brain of rats.
Biochem Pharmacol 22, 1423–1430.
Yousif, S., Chaves, C., Potin, S., Margaill, I., Scherrmann, J.-M., Declèves, X. (2012). Induction
of P-glycoprotein and Bcrp at the rat blood-brain barrier following a subchronic morphine
treatment is mediated through NMDA/COX-2 activation. J Neurochem 123, 491–503.
Zeitz, K.P., Malmberg, A.B., Gilbert, H., Basbaum, A.I. (2001). Reduced development of
tolerance to the analgesic effects of morphine and clonidine in PKC gamma mutant mice. Pain 94,
245–253.
Zhang, H., Nei, H., Dougherty, P.M. (2010). A p38 mitogen-activated protein kinase-dependent
mechanism of disinhibition in spinal synaptic transmission induced by tumor necrosis factor-alpha.
J Neurosci 30, 12844–12855.
Zhang, J., Ferguson, S.S., Barak, L.S., Bodduluri, S.R., Laporte, S.A., Law, P.Y., Caron, M.G.
(1998). Role for G protein-coupled receptor kinase in agonist-specific regulation of mu-opioid
receptor responsiveness. Proc Natl Acad Sci USA 95, 7157–7162.
Zhang, R.-X., Li, A., Liu, B., Wang, L., Ren, K., Zhang, H., Berman, B.M., Lao, L. (2008). IL1ra alleviates inflammatory hyperalgesia through preventing phosphorylation of NMDA receptor
NR-1 subunit in rats. Pain 135, 232–239.
Zheng, M., McErlane, K.M., Ong, M.C. (1998). High-performance liquid chromatography-mass
spectrometry-mass spectrometry analysis of morphine and morphine metabolites and its
application to a pharmacokinetic study in male Sprague-Dawley rats. J Pharm Biomed Anal 16,
971–980.
Zhou, D., Chen, M.L., Zhang, Y.Q., Zhao, Z.Q. (2010). Involvement of spinal microglial P2X7
receptor in generation of tolerance to morphine analgesia in rats. J Neurosci 30, 8042–8047.
Zhu, H.-J., Liu, G.-Q. (2004). Glutamate up-regulates P-glycoprotein expression in rat brain
microvessel endothelial cells by an NMDA receptor-mediated mechanism. Life Sci 75, 1313–1322.
118
References
Zhu, X., Cao, S., Zhu, M.D., Liu, J.Q., Chen, J.J., Gao, Y.J. (2014). Contribution of chemokine
CCL2/CCR2 signaling in the dorsal root ganglion and spinal cord to the maintenance of
neuropathic pain in a rat model of lumbar disc herniation. J Pain 15, 516–526.
Zimmermann, M. (1983). Ethical guidelines for investigations of experimental pain in conscious
animals. Pain 16, 109–110.
Zoli, M., Agnati, L.F. (1996). Wiring and volume transmission in the central nervous system: the
concept of closed and open synapses. Prog Neurobiol 49, 363–380.
Zuo, Z. (2005). The role of opioid receptor internalization and beta-arrestins in the development
of opioid tolerance. Anesth Analg 101, 728–734.
119
ORIGINAL PUBLICATIONS
120