You are on page 1of 19

doi: 10.1111/j.1472-8206.2008.00608.

x
REVI EW
ART I CL E
Drug-induced liver injury through
mitochondrial dysfunction: mechanisms and
detection during preclinical safety studies
Gilles Labbe
a
, Dominique Pessayre
b
, Bernard Fromenty
b
*
a
Sano-aventis recherche & developpement, Drug Safety Evaluation, Alfortville, France
b
INSERM, U773, Centre de Recherche Biomedicale Bichat Beaujon CRB3, Paris, France
I NTRODUCTI ON
Drug-induced liver injury (DILI) is a major issue for
pharmaceutical companies, because DILI is a frequent
cause for the failure of a drug to get approved, or for
the withdrawal of an already marketed medicine. The
nancial, legal and image-tarnishing consequences are
at their worst when toxicity is only discovered after
marketing, requiring the recall of an unsafe medicine,
which has caused lethal liver injury or has required liver
transplantation in many patients.
Although drugs can cause hepatotoxicity through
different ways [1], mitochondrial dysfunction is one
major mechanisms underlying DILI [2]. By hampering
mitochondrial energy production and/or releasing mito-
chondrial pro-apoptotic proteins into the cytoplasm,
drugs can trigger the necrosis or apoptosis of hepato-
cytes, thus causing cytolytic hepatitis, which can evolve
towards liver failure. Drug-induced mitochondrial dys-
function can also lead to steatosis and steatohepatitis,
which can sometimes progress towards cirrhosis [24].
In addition, drug-induced mitochondrial dysfunction
can also trigger diverse extrahepatic or general mani-
festations, such as hyperlactataemia, lactic acidosis,
myopathy, rhabdomyolysis, pancreatitis, peripheral
neuropathy or lipoatrophy [58]. Thus, because of
its potential severity for patients and its dire conse-
quences for the pharmaceutical industry, drug-induced
Keywords
cell death,
drug,
hepatotoxicity,
mitochondria,
steatohepatitis,
steatosis
Received 14 January 2008;
Revised 11 March 2008;
Accepted 27 March 2008
*Correspondence and reprints:
bernard.fromenty@inserm.fr
This review is the summary of
an oral communication made by
Dr B. Fromenty at the 20th
Annual ADPC Meeting, held in
Paris on 24 October 2007, and
devoted to Predictive Toxicol-
ogy and Drug Safety.
ABSTRACT
Mitochondrial dysfunction is a major mechanism whereby drugs can induce liver
injury and other serious side effects such as lactic acidosis and rhabdomyolysis in
some patients. By severely altering mitochondrial function in the liver, drugs can
induce microvesicular steatosis, a potentially severe lesion that can be associated
with profound hypoglycaemia and encephalopathy. They can also trigger hepatic
necrosis and/or apoptosis, causing cytolytic hepatitis, which can evolve into liver
failure. Milder mitochondrial dysfunction, sometimes combined with an inhibition of
triglyceride egress from the liver, can induce macrovacuolar steatosis, a benign lesion
in the short term. However, in the long term this lesion can evolve in some
individuals towards steatohepatitis, which itself can progress to extensive brosis and
cirrhosis. As liver injury caused by mitochondrial dysfunction can induce the
premature end of clinical trials, or drug withdrawal after marketing, it should
be detected during the preclinical safety studies. Several in vitro and in vivo
investigations can be performed to determine if newly developed drugs disturb
mitochondrial fatty acid oxidation (FAO) and the oxidative phosphorylation
(OXPHOS) process, deplete hepatic mitochondrial DNA (mtDNA), or trigger the
opening of the mitochondrial permeability transition (MPT) pore. As drugs can be
deleterious for hepatic mitochondria in some individuals but not in others, it may also
be important to use novel animal models with underlying mitochondrial and/or
metabolic abnormalities. This could help us to better predict idiosyncratic liver injury
caused by drug-induced mitochondrial dysfunction.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353 335
mitochondrial dysfunction should be detected during the
course of preclinical safety studies.
In the present review, we rst briey recall some key
features of mitochondria and the main molecular mech-
anisms whereby drugs can impair mitochondrial func-
tion and lead to liver injury. We then cite a few drugs,
the propensity of which to cause mitochondrial dysfunc-
tion and DILI has led to the premature termination of
clinical trials or to the withdrawal of the drug from the
market. We also indicate which in vitro and in vivo
investigations can be performed during preclinical safety
studies to detect DILI linked to mitochondrial dysfunc-
tion, or later in drug development to explain hepatotoxic
events in some individuals. Finally, we discuss the
possible means to improve the prediction of idiosyncratic
liver injury in the future, such as the use of animal
models with underlying mitochondrial and/or metabolic
abnormalities. It is noteworthy that several excellent
reviews are available for those who wish to gain insight
into preclinical testing strategies independently of the
mechanism of DILI [911].
MI TOCHONDRI AL FUNCTI ONS AND
GENOME
The mitochondria: powerhouses of the cell and
crossroads of metabolic pathways
Mitochondria oxidizes many metabolites including fatty
acids, pyruvate (generated from glycolysis) and several
amino acids such as valine, leucine and methionine
336 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
[3,12,13]. During the fasting state, fatty acid oxidation
(FAO) in the liver is incomplete; the acetyl-CoA that is
generated is not oxidized in the liver, but condenses into
ketone bodies (mainly acetoacetate and b-hydroxybuty-
rate) (Figure 1), which are released into the plasma for
subsequent oxidation in extrahepatic tissues. In kidney,
heart and brain, the ketone bodies are then cleaved into
acetyl-CoA, which is oxidized through the tricarboxylic
acid cycle [3,12].
The intra-mitochondrial oxidation of fatty acids,
pyruvate and amino acids generates reducing equiva-
lents (FADH
2
, NADH), which must be reoxidized by
the mitochondrial respiratory chain (MRC) to regenerate
the FAD and NAD
+
that are needed for the oxidative
processes such as FAO, pyruvate dehydrogenation and
the activity of the tricarboxylic acid cycle (Figure 1). In
addition to regenerating FAD and NAD
+
, the MRC also
creates a large electrochemical gradient across the inner
mitochondrial membrane, the so-called mitochondrial
transmembrane potential or Dw
m
(Figure 1). Indeed,
the transfer of electrons along the different components
of the MRC (for instance, complexes I, III and IV when
electrons come from NADH) is associated with the
ejection of protons from the mitochondrial matrix to
the intermembrane space, thus building up the Dw
m
(Figure 1). The Dw
m
serves as a reservoir of potential
energy. When cells need energy, the rise in ADP levels
within the mitochondria induces the re-entry of protons
into the mitochondrial matrix through the membrane-
spanning F
0
portion of the ATP synthase complex (also
referred to as the complex V of the MRC). This re-entry
of protons partially dissipates the Dw
m
, thus liberating
energy, which is used by the F
1
portion of ATP synthase
to phosphorylate ADP into ATP (Figure 1). Concom-
itantly, the transient decrease in Dw
m
promotes electron
transfer in the MRC and the coupled ejection of
protons towards the intermembrane space, which
restores the Dw
m
. Hence, there is a tight coupling within
the mitochondria between electron transfer within the
MRC (i.e. NADH and FADH
2
oxidation) and ATP
synthesis (i.e. ADP phosphorylation) [3,6,13]. A severe
dysfunction of this oxidative phosphorylation (OXPHOS)
Figure 1 Metabolism and energy production in liver mitochondria, and main drug targets. The entry of long-chain (C
14
C
18
) fatty acids
(LCFAs) within mitochondria requires a specic shuttle system involving four steps: (a) LCFAs are activated into LCFA-coenzyme A (acyl-
CoA) thioesters by long-chain acyl-CoA synthetases (ACS) located in the outer mitochondrial membrane, microsomes and peroxisomes.
(b) The long-chain acyl-CoA is converted into an acyl-carnitine by carnitine palmitoyltransferase-I (CPT-I) in the outer mitochondrial
membrane. (c) The acyl-carnitine is translocated across the inner mitochondrial membrane into the mitochondrial matrix by carnitine-
acylcarnitine translocase. (d) Finally, carnitine palmitoyltransferase-II (CPT-II), located on the inner side of the inner mitochondrial
membrane, transfers the acyl moiety from carnitine back to coenzyme A. LCFA-CoA thioesters are then oxidized into acetyl-CoA moieties via
the b-oxidation process. Acetyl-CoA moieties can then generate ketone bodies (mainly acetoacetate and b-hydroxybutyrate) via ketogenesis,
which is stimulated during fasting. Ketone bodies are liberated into the plasma to be used by extrahepatic tissues for energy production. After
a meal, however, the acetyl-CoA moieties are mainly generated in the liver by the glycolysis of glucose into pyruvate and the oxidative
decarboxylation of pyruvate. The formed acetyl-CoA enters the tricarboxylic acid (TCA) cycle (not shown), to be fully oxidized or to form
citrate, which can leave mitochondria and regenerate acetyl-CoA in the cytosol for subsequent de novo lipogenesis. Mitochondrial fatty acid
b-oxidation (FAO) and the TCA cycle generate NADH and FADH
2
, which transfer their electrons (e
)
) to the mitochondrial respiratory chain
(MRC), regenerating NAD
+
and FAD used for other b-oxidation (or TCA) cycles. Within the MRC, electrons are sequentially transferred to
different polypeptide complexes (numbered from I to IV) embedded within the inner membrane. The nal transfer of the electrons to oxygen
takes place at the level of complex IV, also referred to as cytochrome c oxidase. The mitochondrial DNA (mtDNA) encodes 13 polypeptides,
which are embedded within complexes I, III, IV and V. The ow of electrons within the MRC is coupled with the extrusion of protons (H
+
)
from the mitochondrial matrix to the intermembrane space, which creates the mitochondrial transmembrane potential, Dw
m
. When energy is
needed (i.e. when ATP levels are low), these protons re-enter the matrix through ATP synthase (also referred to as complex V), thus liberating
energy that is used to phosphorylate ADP into ATP. The adenine nucleotide translocator (ANT) then extrudes ATP from mitochondria, in
exchange for cytosolic ADP. Importantly, ANT is one of the components of the mitochondrial permeability transition (MPT) pore, which also
includes the voltage-dependent anion channel (VDAC) in the outer membrane and cyclophilin D in the matrix. The opening of the MPT pore
by exogenous or endogenous compounds leads to the outer membrane rupture and the release into the cytosol of cytochrome c (Cyt. c), a
component of the MRC loosely associated with the inner membrane. Cytochrome c release induces caspase activation and cell death via
apoptosis or necrosis. Drugs can impair mitochondrial function through different mechanisms. They can act (i) by inducing MPT, (ii) by
causing mtDNA depletion, which can reduce the synthesis of mtDNA-encoded MRC polypeptides, impair MRC activity and increase the
generation of the superoxide anion and other reactive oxygen species (ROS), (iii) by directly impairing MRC activity, which can secondarily
inhibit mitochondrial FAO and enhance ROS formation (not shown), (iv) by dissipating Dw
m
, (v) by directly altering FAO via the inhibition of
b-oxidation enzyme(s), and/or (vi) by indirectly impairing FAO through the sequestration of the FAO cofactors, L-carnitine and coenzyme A.
NAPQI and NRTIs signify N-acetyl-p-benzoquinone imine and nucleoside reverse transcriptase inhibitors, respectively.
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 337
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
process can jeopardize cell metabolism and cell life in the
liver and elsewhere [2,14].
Double genetic origin of the OXPHOS proteins
One unique feature of mitochondria is the double
genetic origin of the proteins (approximately 100) that
play a role in the OXPHOS process. Indeed, whereas
the majority of these polypeptides are encoded by the
nuclear genome and subsequently imported within the
mitochondria, 13 MRC polypeptides are instead
encoded by the mitochondrial genome, a small piece
of circular doubled-stranded DNA (approximately
16.6 kb in mammals) located within the mitochondrial
matrix (Figure 1). Importantly, there are several hun-
dreds (or thousands) of copies of mitochondrial DNA
(mtDNA) in a single cell, and the replication of these
mtDNA molecules occurs continuously, even in cells
that do not divide [6]. Permanent mtDNA replication by
the DNA polymerase c thus allows mtDNA levels to
remain constant in the cell, despite the continuous
degradation of some mitochondria, particularly dam-
aged and dysfunctional mitochondria [15,16]. It is
important to bear in mind that most cells (including
hepatocytes) have a surplus of mtDNA copies, and
can therefore tolerate a substantial depletion of mtDNA.
Classically, it is considered that the number of normal
mtDNA copies must fall below 2040% of basal
levels to induce mitochondrial dysfunction and adverse
events [6]. Below this threshold, hepatic mtDNA
depletion can impair the MRC. This impairment can
hamper the regeneration of NAD
+
and FAD, thus
leading to a secondary inhibition of the oxidation of
fatty acids and pyruvate. Yet another consequence
of the MRC impairment is to increase the production
of reactive oxygen species (ROS), which can damage
diverse mitochondrial constituents, including mtDNA
[6,1719]. Indeed, a key feature of the mitochondrial
genome is its high sensitivity to oxidative damage,
because of its proximity to the inner membrane (the
main cellular source of ROS, as discussed in the next
section), the absence of protective histone and an
incomplete repertoire of DNA repair enzymes within the
mitochondria [6,20,21].
The respiratory chain as a source of ROS
As pointed out previously, most of the electrons provided
by FADH
2
or NADH to the MRC migrate all the way
along the different MRC components, to nally reach
cytochrome c oxidase (COX, also referred to as complex
IV), where they safely combine with oxygen and protons
to form water (Figure 1). This leads to oxygen consump-
tion, which can be measured by a Clark-type electrode or
oxygen-sensitive probes, as mentioned later on. How-
ever, at several upstream sites of the MRC (most probably
complexes I and III), a small fraction of these electrons
can directly react with oxygen, to generate the super-
oxide anion radical. This radical is then dismutated
by the mitochondrial manganese superoxide dismutase
(MnSOD) into hydrogen peroxide (H
2
O
2
), which is
detoxied into water by the mitochondrial glutathione
peroxidase. Hence, in the normal (non-diseased) state,
most of the ROS generated by the MRC is detoxied by
the mitochondrial antioxidant defences.
However, this detoxication process can be over-
whelmed in some circumstances. One such circumstance
is the depletion of mitochondrial, reduced glutathione.
Normally, glutathione peroxidase plays a key role in
H
2
O
2
detoxication, as liver mitochondria do not
have catalase. However, glutathione peroxidase needs
adequate amounts of reduced glutathione in the mito-
chondrial matrix to detoxify H
2
O
2
. The depletion of
mitochondrial glutathione below a critical threshold can
impair mitochondrial H
2
O
2
detoxication, which can
trigger mitochondrial dysfunction, mitochondrial perme-
ability transition (MPT) and cell death [22].
Mitochondrial antioxidant enzymes can also be
overwhelmed when electron transfer within the MRC is
chronically hindered. A partial block in the ow of
electrons within the MRC leads to their accumulation
within MRC complexes. At some sites of complex I and
complex III, the accumulated electrons directly react
with oxygen to form the superoxide anion radical [23].
When mitochondrial dysfunction is prolonged, the
enhanced formation of the superoxide anion can over-
whelm the mitochondrial antioxidant defences. High
steady-state levels of ROS then damage OXPHOS pro-
teins, cardiolipin and mtDNA. This oxidative damage
aggravates mitochondrial dysfunction to further aug-
ment electron leakage and ROS formation, thus leading
to a vicious circle [2,6].
The MPT pore
The MPT pore is located at contact sites between
the outer and inner mitochondrial membranes. The pore
may be composed of several proteins, including the
voltage-dependent anion channel (VDAC) located in the
outer membrane, the adenine nucleotide translocator
(ANT) (also called ADP/ATP translocase) in the inner
mitochondrial membrane and cyclophilin D in the
mitochondrial matrix (Figure 1).
338 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
In the normal state, MPT pores are kept closed, but in
some circumstances they rapidly open, thus increasing
the permeability of the mitochondrial membranes to
compounds, the molecular weight of which is <1.5 kDa.
As a result, the transmembrane potential (Dw
m
) collapses
because of massive proton re-entry, and the mitochon-
drial matrix expands because of water accumulation,
leading to the rupture of the outer membrane and the
release of pro-apoptotic proteins from the intermembrane
space into the cytosol. The release of apoptosis-inducing
factor (AIF) triggers the fragmentation of large-
sized nuclear DNA. The release of a protein called
second mitochondria-derived activator of caspase/
direct inhibitor of apoptosis-binding protein with low pI
(Smac/DIABLO) inactivates the inhibitor of apoptosis
proteins (IAPs). Finally, the released cytochrome c binds
in the cytosol to Apaf-1 in an ATP-dependent manner to
activate caspase-9. Importantly, cytochrome c, which is
loosely bound to the outer part of the inner membrane, is
a key component of the MRC (Figure 1). Therefore, its
leakage from the mitochondria eventually impairs
electron transfer within the MRC and enhances ROS
generation.
Mitochondrial permeability transition can cause either
apoptosis or necrosis. When MPT occurs in some
mitochondria only, the unaffected organelles synthesize
ATP, thus avoiding necrosis, while the permeabilized
mitochondria release pro-apoptotic substances that trig-
ger apoptosis. In contrast, when MPT occurs in most
mitochondria in a single cell, ATP generation is com-
promised because of the collapse of the Dw
m
, and the
resulting ATP depletion triggers cell necrosis [2,24].
Mitochondrial permeability transition pore opening,
which is calcium-dependent, can be triggered by a
variety of endogenous compounds such as iron, ROS,
nitric oxide, free fatty acids (and acyl-CoA derivatives),
ceramide and bile salts [25,26]. MPT can also be
triggered by extracellular cytokines such as tumour
necrosis factor-a (TNF-a) and Fas ligand (Fas-L) acting
through their plasma membrane receptors [27,28].
Finally, MPT can be triggered by drugs, as discussed in
the next section.
MAI N MECHANI SMS AND
CONSEQUENCES OF DRUG- I NDUCED
HEPATI C MI TOCHONDRI AL
DYSFUNCTI ON
Drugs (or their metabolites) can induce mitochondrial
dysfunction in hepatocytes through several mechanisms.
Induction of MPT and cytochrome c release
Several drugs can trigger MPT in liver mitochondria,
leading to apoptosis and/or necrosis. These drugs cause
cytolytic hepatitis, leading occasionally to fulminant
hepatic failure. Some drugs (or metabolites) listed in
Table I can directly trigger MPT in isolated mitochondria
[2,2931]. The molecular mechanisms whereby these
drugs induce MPT are often poorly understood.
Other drugs (or metabolites) can induce the release
of pro-apoptotic proteins from mitochondria through
indirect mechanisms. For instance, troglitazone and
acetaminophen activate c-Jun N-terminal protein kinase
(JNK), thus inducing the cleavage of Bid, the transloca-
tion of this pro-apoptotic protein to the outer mitochon-
drial membranes and the release of cytochrome c from
mitochondria [32,33]. Another mechanism is seen
during the extensive formation of reactive metabolites,
which increases cytosolic calcium that then enters into
the mitochondria to trigger MPT and apoptosis [34].
FAO impairment
Drugs can also impair the mitochondrial oxidation of
fatty acids, leading to their accretion as triglycerides in
the cytoplasm of the hepatocytes. Although free fatty
acids can also accumulate, this seems to happen
preferentially when FAO is severely inhibited [3]. FAO
can be impaired in different ways. Some drugs directly
impair mitochondrial FAO by inhibiting FAO enzymes
and/or by sequestering the FAO cofactors, L-carnitine
and coenzyme A (Figure 1) [24,35]. Other drugs rst
inhibit MRC activity either directly or as a consequence
of detrimental effects on mtDNA replication and/
or integrity (Figure 1) [24]. Severe MRC inhibition
decreases the regeneration of the oxidized cofactors
(NAD
+
and FAD), which are required for b-oxidation.
Table I Examples of drugs able to directly induce the opening of the
MPT pore.
Drug
a
Therapeutic class
Acetaminophen-derived NAPQI
b
Antalgic drug
Alpidem Anxiolytic drug
Diclofenac Nonsteroidal anti-inammatory
drug (NSAID)
Nimesulide NSAID
Salicylic acid NSAID
Troglitazone Antidiabetic drug (PPARc agonist)
Valproic acid Antiepileptic drug
a
Drugs are listed by alphabetic order.
b
N-Acetyl-p-benzoquinone imine.
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 339
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
Thus, severe MRC inhibition secondarily inhibits mito-
chondrial b-oxidation.
Severe inhibition of mitochondrial b-oxidation can
lead to microvesicular steatosis characterized by the
presence of many tiny lipid droplets within the cytoplasm
of hepatocytes. This potentially severe liver lesion can be
associated with liver failure, profound hypoglycaemia
and encephalopathy [24]. Microvesicular steatosis can
also lead to the death of the patient. Examples of drugs
capable of inducing microvesicular steatosis are given in
Table II [24].
However, a less severe (and more prolonged) inhibi-
tion of mitochondrial b-oxidation can cause macrovac-
uolar steatosis, which is characterized by the presence
of a single, large lipid vacuole within the cytoplasm of
the hepatocytes. Although macrovacuolar steatosis is a
benign liver lesion in the short term, it can evolve in
some patients towards steatohepatitis after several years.
In this more severe condition, steatosis is associated with
necrosis, apoptosis, inammation and brosis, which can
lead to cirrhosis in some patients [2,4]. Examples of
drugs capable of inducing macrovacuolar steatosis and
steatohepatitis are listed in Table III [2,4,3639]. Inter-
estingly, drugs inducing steatohepatitis are usually com-
pounds that impair the MRC (Figure 1), thus causing not
only FAO impairment and steatosis, but also enhanced
ROS production. ROS overproduction may be a key event
in the pathogenesis of drug-induced steatohepatitis by
inducing lipid peroxidation (favoured by lipid accumu-
lation), and possibly by triggering the generation of pro-
apoptotic (TNF-a) and pro-brotic (TGF-b) cytokines by
Kupffer cells and other inammatory cells [2,4,40].
Regarding drug-induced steatosis and steatohepatitis,
several important remarks must be made.
1 Some drugs can acutely induce microvesicular steato-
sis in some patients, and can also cause subacute
or chronic macrovacuolar steatosis or steatohepatitis
in other patients (Tables II and III). It is tempting to
speculate that such drugs may be able to severely impair
mitochondrial FAO in a few patients (possibly those who
may have an underlying mitochondrial dysfunction,
as discussed later on), thus leading to microvesicular
steatosis. In other patients, however, a less severe and
more prolonged impairment of FAO may instead cause
macrovacuolar steatosis.
2 Some drugs can inhibit mitochondrial respiration
through both direct and indirect mechanisms. For
instance, tamoxifen directly inhibits the transfer of
electrons within the MRC, and also progressively depletes
hepatic mtDNA through a mechanism that may involve
the inhibition of mitochondrial topoisomerases and the
impairment of mtDNA replication by this DNA-inter-
calating drug [41,42].
3 By severely inhibiting FAO (either directly or indirectly
through an impairment of the MRC), drugs can deplete
ATP stores within the hepatocytes and lead them to cell
death.
4 Although drug-induced FAO impairment is a key
mechanism leading to steatosis, some drugs (e.g. amio-
darone, amineptine, pirprofen) also inhibit microsomal
triglyceride transfer protein (MTP), an enzyme playing a
major role in the formation of very-low-density-lipopro-
tein (VLDL) particles [43]. Such drugs therefore inhibit
the two main pathways removing fat from the liver,
namely FAO and hepatic VLDL secretion.
Table III Examples of drugs capable of inducing macrovacuolar
steatosis and steatohepatitis.
Drug
a
Therapeutic class
Amiodarone Anti-anginal, anti-arrhythmic drug
Irinotecan Antineoplastic drug (colorectal cancer)
Methotrexate Antipsoriatic, anti-rheumatoid drug
NRTIs
b
(AZT, ddI, d4T) Antiretroviral (anti-HIV) drug
Perhexiline Anti-anginal drug
Tamoxifen Antineoplastic drug (breast cancer)
Toremifene Antineoplastic drug (breast cancer)
a
Drugs are listed by alphabetic order.
b
Nucleoside reverse transcriptase inhibitors.
Table II Examples of drugs capable of inducing microvesicular
steatosis.
Drug
a
Therapeutic class
Amineptine Antidepressant drug
Amiodarone Anti-anginal, anti-arrhythmic drug
2-Arylpropionic acids
(pirprofen, ibuprofen)
Nonsteroidal anti-inammatory
drug (NSAID)
Aspirin (and salicylic acid) NSAID
Fialuridine (FIAU) Antiviral (anti-HBV) drug
NRTIs
b
(AZT, ddI, d4T) Antiretroviral (ant-HIV) drug
Panadiplon Anxiolytic drug
Perhexiline Anti-anginal drug
Tetracycline and its derivatives
(high doses)
Antibiotics
Tianeptine Antidepressant drug
Valproic acid Antiepileptic drug
a
Drugs are listed by alphabetic order.
b
Nucleoside reverse transcriptase inhibitors.
340 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
OXPHOS uncoupling
Cationic amphiphilic drugs such as amiodarone, per-
hexiline, diethylaminoethoxyhexestrol (DEAEH) and
tamoxifen can be rst protonated in the intermembrane
space of mitochondria, and then electrophoretically
transported into the mitochondrial matrix thanks to
the Dw
m
[3,4,40,42], as discussed in greater detail
below. This transfer across the inner mitochondrial
membrane dissipates Dw
m
(Figure 1) and increases the
basal (state 4) mitochondrial respiration. This phe-
nomenon, which is called OXPHOS uncoupling, can
also be observed in the case of some mitochondrial
poisons such as the protonophores 2,4-dinitrophenol
(DNP) and carbonylcyanide p-triuoromethoxy-
phenylhydrazone (FCCP). Importantly, the long-lasting
increase in mitochondrial respiration induced by ami-
odarone, perhexiline or tamoxifen can be observed in
isolated liver mitochondria incubated with relatively
low concentrations (20100 lM) of these drugs
[3,40,42]. However, for higher concentrations, their
intramitochondrial accumulation induces a rapid inhi-
bition of electron transfer within the MRC, so that the
transient increase in state 4 is followed by a progres-
sive decrease in mitochondrial respiration [3,40,42].
Finally, it must be pointed out that another important
consequence of drug-induced OXPHOS uncoupling is
the reduction of ATP synthesis, as the re-entry of
protons within the mitochondrial matrix bypasses ATP
synthase [3,40].
Multiple mechanisms
Several drugs impair mitochondrial function via several
of the above-mentioned pathways. For instance,
valproic acid inhibits mitochondrial FAO through the
sequestration of coenzyme A (a cofactor mandatory for
FAO), and possibly also through the inactivation
of b-oxidation enzymes by an electrophilic metabolite
[2,3]. In addition, this antiepileptic drug can induce
MPT (Figure 1) [29], which may explain why valproate-
induced microvesicular steatosis can be associated with
liver cell death [2].
DRUG WI THDRAWAL BECAUSE OF
MI TOCHONDRI AL DYSFUNCTI ON AND
LI VER I NJ URY
In the past 20 years, the development of several drugs
was halted during clinical trials or they were withdrawn
denitively or temporarily from the market because of
mitochondrial dysfunction-associated DILI (Table IV). For
example, aluridine (FIAU) induced unmanageable lactic
acidosis, microvesicular steatosis and severe liver failure
requiring liver transplantation or leading to death, in
several patients receiving this anti-hepatitis B virus
(HBV) nucleoside analogue during clinical trials [3,44].
After interruption of these trials, new in vivo and in vitro
investigations showed that aluridine strongly inhibits
DNA polymerase c and triggers mtDNA depletion by an
unusual mechanism [44,45]. Unlike nucleoside reverse
transcriptase inhibitors (NRTIs) such as stavudine (d4T),
zidovudine (AZT) or didanosine (ddI), the incorporation
of which into a growing chain of mtDNA interrupts
mtDNA replication, aluridine can be incorporated into
mtDNA without immediately terminating mtDNA repli-
cation. Indeed, this drug carries a hydroxyl group on the
3 carbon of the sugar moiety, allowing the subsequent
incorporation of other nucleotides [46]. However, when
several adjacent molecules of aluridine are successively
incorporated into a growing chain of mtDNA, DNA
polymerase c is strongly inhibited, blocking further
mtDNA replication (Figure 1) [45].
Panadiplon is a non-benzodiazepine anxiolytic drug,
the development of which was interrupted during phase I
clinical trials because of increased serum transaminase
activity [2,47]. Subsequent studies showed that pana-
diplon can be converted into cyclopropane carboxylic
acid, which sequesters coenzyme A and L-carnitine, two
cofactors required for mitochondrial FAO (Figure 1).
Hence, panadiplon inhibits mitochondrial FAO and
causes microvesicular steatosis (as well as necrosis) in
Dutch-belted rabbits [47]. As discussed later on, pre-
clinical safety investigations in rats, dogs and monkeys
treated with panadiplon revealed good tolerance,
although microvesicular steatosis was observed in mon-
keys [47].
Table IV Examples of drugs, the potential of which to cause
mitochondrial dysfunction and DILI has led to the interruption of
clinical trials, or their withdrawal after marketing.
Drug
a
Therapeutic class
Drug withdrawn during clinical trials
Fialuridine Antiviral (anti-HBV) drug
Panadiplon Anxiolytic drug
Drug withdrawn after marketing
Alpidem Anxiolytic drug
Perhexiline Anti-anginal drug
Pirprofen Nonsteroidal anti-inammatory drug (NSAID)
Troglitazone Antidiabetic drug (PPARc agonist)
a
Drugs are listed by alphabetic order within each subgroup.
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 341
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
The anxiolytic drug, alpidem, has been removed from
the market after it led to several cases of severe,
sometimes fatal, hepatitis [48]. Investigations sub-
sequently showed that this peripheral benzodiazepine
receptor (PBR) ligand can accelerate MPT in rat liver
mitochondria and increase the toxicity of TNF-a in rat
hepatocytes [2,49].
Perhexiline is an anti-anginal drug, the administration
of which could cause steatohepatitis that was often
severe, leading to cirrhosis and death of some patients
[3,4,37]. Consequently, perhexiline has been withdrawn
from the market, although it has recently been suggested
that therapeutic plasma monitoring could perhaps
rehabilitate the drug [50]. Similar to amiodarone,
perhexiline accumulates in mitochondria, where it
inhibits both FAO (thus causing steatosis) and mito-
chondrial respiration (thus increasing the mitochondrial
formation of ROS and favouring lipid peroxidation)
[40,51]. Perhexiline and amiodarone accumulate within
mitochondria thanks to Dw
m
[3,52]. Indeed, these
amphiphilic drugs are protonated within the acidic
intermembrane space and pulled into the mitochondrial
matrix, which is basic and negatively charged. In the
more alkaline matrix, the protonated perhexiline or
amiodarone molecule frees its proton, so that the net
result of the import of one drug molecule into the matrix
is to transfer one proton from the intermembrane space
into the matrix. This abnormal entry of protons bypasses
the F
0
portion of ATP synthase, thus uncoupling
OXPHOS and decreasing ATP production. However, as
perhexiline and amiodarone progressively accumulate
within the mitochondria, they soon reach high concen-
trations, which inhibit FAO enzymes and also electron
transfer within the MRC [3,5154]. Hence, after a
transient stimulation caused by uncoupling, mitochon-
drial respiration is progressively inhibited in mito-
chondria incubated with perhexiline or amiodarone
[3,51,52].
Pirprofen, a nonsteroidal anti-inammatory drug
(NSAID), was removed from the market after it had
caused several instances of liver lesions combining both
severe cytolytic hepatitis and microvesicular steatosis
[55]. This aryl-propionate drug was later shown to
inhibit mitochondrial b-oxidation [56] and MTP activity
[43], decreasing both the oxidation of fat within
hepatocytes and the secretion of fat from the liver,
respectively.
Troglitazone, a peroxisome proliferator-associated
receptor-c agonist, has been removed from the market
after the occurrence of several cases of acute liver failure,
which were fatal or required liver transplantation in
some patients [2,57]. In vitro investigations sub-
sequently showed that this antidiabetic drug induces
MPT in mouse liver mitochondria and inhibits ADP-
driven respiration (with glutamate and malate as sub-
strates) in rat liver mitochondria [31,58]. In addition,
troglitazone is a potent inhibitor of complex IV and V
activities in bovine heart mitochondria [58]. As men-
tioned previously, investigations also showed that trog-
litazone-induced liver cell death could also involve JNK
activation [32].
Several other drugs can induce liver toxicity
through mitochondrial dysfunction, but still have a
favourable benetrisk ratio, at least in some circum-
stances. Although these drugs have not been withdrawn
from the market, their use has been restricted by diverse
Black Box warnings added at the request of drug
agencies (Table V).
PREDI SPOSI TI ONS TO DRUG- I NDUCED
MI TOCHONDRI AL DYSFUNCTI ON AND
I DI OSYNCRATI C DI LI
Drugs are only considered for marketing when they do
not cause DILI, or rarely cause DILI. This selection
ensures that DILI only affects a minority of the treated
patients. Otherwise, the compound would not have been
marketed. At the moment, we still poorly understand
which susceptibility factor(s) may explain the higher
susceptibility of these few subjects. However, it is already
clear that genetic, metabolic and environmental factors
that impair mitochondrial function can add their effects
to those of drugs capable of causing mitochondrial
dysfunction. The added effects of the pre-existing con-
dition and the mitochondria-targeting drug can then
Table V Examples of marketed drugs capable of inducing hepato-
toxicity caused by mitochondrial dysfunction, which have received
Black Box warnings from drug agencies.
Drug
a
Therapeutic class
Amiodarone Anti-anginal, anti-arrhythmic drug
Benzbromarone Uricosuric drug
Buprenorphine Opioid agonist/antagonist
NRTIs
b
(AZT, d4T, ddI) Antiretroviral (anti-HIV) drug
Tamoxifen Antineoplastic drug (breast cancer)
Tolcapone Anti-Parkinson drug
Valproic acid Antiepileptic drug
a
Drugs are listed by alphabetic order.
b
Nucleoside reverse transcriptase inhibitors.
342 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
alter mitochondrial function to reach the low threshold,
under which hepatic manifestations start to occur.
Genetic predisposition
Several studies reported fatal valproate hepatotoxicity in
patients suffering from an underlying genetic mitochon-
drial disease, such as cytochrome c oxidase deciency
or medium-chain acyl-CoA dehydrogenase (MCAD)
deciency [5961]. MCAD is a key FAD-dependent
dehydrogenase involved in the oxidation of medium-
chain (C
6
C
12
) fatty acids within the mitochondria [3].
MCAD deciency is a frequent metabolic disorder of
genetic origin among Caucasians. Indeed, the carrier
frequency of the A985G mutation (which is the most
frequent mutation leading to MCAD deciency) ranges
from approximately one in 55 in northern Europe to one
in 140 in southern Europe [6264]. Because of this high
frequency, it may be worth determining whether the
MCAD carrier state is sufcient or not to increase the risk
of DILI in patients receiving drugs that impair mito-
chondrial function.
Although the report did not mention liver injury, a
mutation in the catalytic subunit of DNA polymerase c
was found in a patient with stavudine-induced hyper-
lactataemia [65]. Interestingly, the mutation markedly
decreased DNA polymerase activity, and rendered lym-
phoblastoid cells more sensitive to stavudine-induced
mtDNA depletion [65]. Thus, it would be interesting to
look for this mutation in patients with NRTI-induced
liver injury.
It must be also pointed out that there is a large
variation of mtDNA copy number in liver or muscle
among different individuals [66,67]. Although the origin
of this variation is currently unknown, low basal mtDNA
levels could perhaps act as a predisposing factor to the
toxicity of NRTIs or other mtDNA-altering drugs.
A recent study showed that individuals with the
C allele (T/C or C/C genotype) of MnSOD had a higher
risk of experiencing DILI with anti-tuberculosis drugs
[68]. Carrying the C/C genotype (i.e. being homo-
zygous for alanine in the mitochondrial targeting
sequence of MnSOD) was also found to greatly enhance
the risk of microvesicular steatosis or cirrhosis in French
alcoholics [69]. Furthermore, the alanine-encoding
allele also increased the risk of hepatocellular carci-
noma development in alcoholic cirrhotic patients [70].
The Ala-MnSOD precursor variant is more efciently
imported into the mitochondria than the Val-MnSOD
precursor variant, resulting in a higher mitochondrial
activity of the MnSOD homotetramer derived from the
Ala-MnSOD precursor [71,72]. MnSOD transforms the
superoxide anion into H
2
O
2
, while alcohol intoxication
can decrease hepatic mitochondrial glutathione, which
is required for the detoxication of H
2
O
2
by glutathione
peroxidase. Therefore, in alcoholics with the Ala-
MnSOD precursor variant, a high MnSOD activity
combined with a low glutathione peroxidase activity
could increase the cellular steady-state levels of
H
2
O
2
, leading to an increased formation of the highly
reactive hydroxyl radical and to liver injury and cancer
[73].
Environmental factors
Environmental factors such as alcohol, microsomal
enzyme inducers or viral infections can enhance the
risk of drug-induced mitochondrial dysfunction and
DILI. Ethanol overconsumption causes numerous dele-
terious effects on liver mitochondria, in particular on
the mitochondrial genome [3,74,75]. Alcohol abuse
may therefore induce an underlying mitochondrial
dysfunction, which may render the liver more vulner-
able to drug-induced mitochondrial dysfunction and
DILI. Surprisingly, this susceptibility factor has been
insufciently addressed in the literature, due in part to
the difculty of disentangling alcohol- and drug-induced
injury in some circumstances. Nevertheless, it is known
that the hepatotoxicity of methotrexate and that of
paracetamol is increased in alcoholics [76]. In addition
to causing mitochondrial alterations, ethanol intoxica-
tion also increases cytochrome P450 2E1 (CYP2E1)
levels, not only in the endoplasmic reticulum but
also in mitochondria [77]. The ectopic localization of
CYP2E1 may enhance the mitochondrial generation of
N-acetyl-p-benzoquinone imine (NAPQI), the toxic
metabolite of acetaminophen. As the reactive NAPQI
has a direct toxicity effect on mitochondria, in partic-
ular by triggering MPT [78,79], the generation of
NAPQI directly within the mitochondria may favour
acetaminophen-induced mitochondrial dysfunction and
liver injury.
Mitochondria contain not only CYP2E1 but also other
inducible CYPs, such as CYP1A1 and CYP2B1 [80,81].
Thus, the concomitant administration of a CYP-inducer
(such as phenobarbital or phenytoin) may increase the
mitochondrial CYP-mediated generation of reactive
metabolites in subjects co-treated with some drugs, such
as valproic acid [76,82].
Viral infections may also be a predisposing factor
to drug-induced mitochondrial dysfunction in the liver
[76]. Viral infections can release viral proteins, NO,
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 343
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
cytokines and interferon, which can impair mitochon-
drial function and lipid homeostasis, and thus could have
additive effects with drugs inducing mitochondrial dys-
function [5,6,76]. An example is the combined role of
aspirin and viral infections in the occurrence of Reyes
syndrome, a severe post-infectious metabolic disease
characterized by extensive microvesicular steatosis of the
liver, severe hypoglycaemia and a non-inammatory
encephalopathy [3,83]. Although therapeutic doses of
aspirin are well tolerated in uninfected children, they can
trigger Reyes syndrome in a few children sensitized by
varicella or inuenza.
Another example is the hepatic toxicity of some NRTIs
used to prevent the replication of the human immuno-
deciency virus (HIV) (Tables II and III). The hepatic
toxicity of NRTIs is increased in patients co-infected with
either the hepatitis B virus (HBV) or the hepatitis C virus
(HCV) [6,84]. Some HBV or HCV proteins such as the
HCV core protein or the HBV X protein (HBx) can disturb
mitochondrial function [8587]. Together with cyto-
kines, these viral proteins could impair mitochondrial
function and make co-infected subjects more susceptible
to NRTI-induced liver toxicity.
Obesity and NAFLD
A recent surge in the prevalence of obesity threatens the
health and life of millions of individuals, as obesity
increases the risk of type 2 diabetes, hyperlipaemia,
coronary heart disease, stroke, osteoarthritis and cancer.
Furthermore, obesity can cause non-alcoholic fatty liver
disease (NAFLD) [13,88], which can make the liver more
vulnerable to some forms of DILI, and to alcohol-induced
liver injury [38,8993]. This increased vulnerability
could involve several mechanisms, including an under-
lying mitochondrial dysfunction, an oxidative stress, and
an overproduction of cytotoxic cytokines such as tumour
necrosis factor (TNF)-a in NAFLD [13,88,9496]. As the
number of treated overweight or obese patients is
increasing, clinical investigations are urgently needed
to establish which drugs have an increased hepatotoxic
risk in patients with NAFLD.
METHODS TO I NVESTI GATE
DRUG- I NDUCED MI TOCHONDRI AL
DYSFUNCTI ON I N LI VER
Among the battery of in vitro and in vivo investigations
that are presented in this section, some of them can be
performed during the preclinical safety studies to detect
drug-induced hepatotoxicity linked to mitochondrial
dysfunction. These methods allow us to assess the
capacity of a drug to inhibit or uncouple mitochondrial
respiration, to impair FAO, to trigger MPT or to deplete
mtDNA. Importantly, some of these investigations can
also be carried out before preclinical studies in order to
predict mitochondrial dysfunction, thus allowing the
selection of safer compounds for subsequent develop-
ment. Finally, investigational studies can be also per-
formed after clinical trials, or marketing, in order to
explain the occurrence of hepatotoxic events in some
individuals.
The in vitro assays
Several in vitro assays can be used to detect drug-
induced mitochondrial dysfunction either in hepatic cells
or in isolated liver mitochondria (Table VI). Although
technical details are not given in the present review,
some important issues are addressed.
1 When the drug is highly bound to plasma proteins, it is
convenient to perform these in vitro investigations in the
presence of a physiological concentration of albumin
[97]. This allows a direct comparison of the concentra-
tions that impair mitochondrial function in vitro with
the plasma levels in treated patients [97].
2 A very convenient method is to assess the effects of
drugs on FAO, respiration and MPT rst in isolated rat or
mouse liver mitochondria, and then to check the effects
Table VI Examples of in vitro investigations which can be
performed to detect drug-induced mitochondrial dysfunction.
Investigations on isolated liver mitochondria
Assessment of fatty acid oxidation (FAO) with radiolabelled fatty acids
Measurement of oxygen consumption with a Clark-type electrode
Measurement of mitochondrial respiratory chain (MRC) complexes activities
by spectrophotometry
Determination of Dw
m
(mitochondrial transmembrane potential) with the
tetraphenylphosphonium chloride selective electrode, or by ow cytometry
with a uorescent probe (e.g. TMRM
a
), or by spectrouorometry with a
uorescent probe (e.g. safranine)
Assessment of mitochondrial permeability transition (MPT) pore opening by
spectrophotometry (OD recording at 540 nm)
Investigations on hepatic cells
Coloration with oil red O (for the detection of neutral lipids)
Measurement of lactic acid in the incubation medium
Assessment of FAO with radiolabelled fatty acids
Measurement of oxygen consumption with a Clark-type electrode
Determination of Dw
m
(mitochondrial transmembrane potential) by ow
cytometry with a uorescent probe (e.g. MitoTracker Red CMXRos, DiOC6
b
)
Assessment of mtDNA levels by qPCR
a
Tetramethylrhodamine methyl ester.
b
3,3-Dihexyloxacarbocyanine iodide.
344 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
of the drug in human hepatoma cell line [97] or in
human lymphocytes. Human lymphocytes can be indeed
used as a useful model to assess the ability of a drug to
inhibit FAO or deplete ATP stores in human cells [98].
3 When using hepatoma cells, it must be remembered
that most hepatic cell lines express very low (or nil) levels
of enzymes involved in drug metabolism such as CYPs
or glutathione S-transferases (GSTs), possibly due to
low expression of the nuclear receptors regulating the
expression of these enzymes such as the constitutive
androstane receptor (CAR) and pregnane X receptor
(PXR) [99,100]. This major limitation of most hepatoma
cells should be taken into consideration during preclin-
ical safety studies, as mitochondrial dysfunction can be
induced by reactive metabolites and not by the parent
drugs [101]. Hence, negative results regarding mito-
chondrial dysfunction in HepG2 cells should be inter-
preted only as indicating a lack of toxicity of the parent
compound. In contrast, the human hepatoma cell line
HepaRG expresses substantial levels of different CYPs,
GSTs and nuclear receptors [99,100], and could there-
fore be preferred for toxicological studies.
4 Hepatoma cell lines cultivated in high-glucose media
(i.e. 25 mM) generate ATP mostly from glycolysis rather
than from mitochondrial OXPHOS. The reliance of
HepG2 cells on glycolysis decreases their vulnerability
to drugs altering mitochondrial function or to OXPHOS
poisons, such as rotenone, antimycin, oligomycin or
FCCP [102]. However, when HepG2 cells are cultivated
with galactose (10 mM) and glutamine (2 mM) for more
than 10 passages, their susceptibility to mitochondria-
targeting compounds increases [102]. Because the
oxidation of galactose to pyruvate does not cause any
net ATP generation, cultivating HepG2 cells with this
substrate progressively forces cells to rely more on
mitochondrial ATP generation [102].
5 Hepatoma cell lines (such as HepG2, H4IIE and Fao)
also present a very poor capacity to oxidize long-chain
fatty acids (LCFAs), such as palmitic acid and oleic acid,
possibly because of high levels of malonyl-CoA, an
endogenous inhibitor of carnitine palmitoyltransferase I
(CPT-I), which catalyses the limiting step for LCFA
oxidation within the mitochondria (Figure 1) [103,104].
We recently conrmed these results on HepG2 cells and
showed that HuH7 and Chang Liver cells also have a
poor capacity to oxidize LCFAs [105]. Thus, these
hepatoma cell lines are not suitable to explore the effect
of drugs on mitochondrial LCFA oxidation. It would be
interesting to determine whether the mitochondrial
oxidation of LCFAs can be improved if these cells are
maintained in a galactose medium. In contrast, the
mitochondrial oxidation of medium-chain fatty acids,
such as octanoic acid, seems active in hepatoma cells
[104].
6 Caution is needed when mitochondrial function is
investigated in cryopreserved/thawed hepatocytes.
Indeed, a recent study showed profound mitochondrial
dysfunction in cryopreserved/thawed human and mouse
hepatocytes [106]. In particular, ATP levels and mito-
chondrial respiration with glutamate/malate, which give
electrons to complex I, were severely decreased. More-
over, electron microscopy showed swollen mitochondria
and partial loss of cristae [106]. Uncontrolled damage to
mitochondria and other organelles in cryopreserved/
thawed hepatocytes is likely to alter the vulnerability
of cells to drugs. Hence, further studies are needed to
improve the conditions of cryopreservation and thawing,
and determine whether cryopreserved/thawed hepato-
cytes can be valuable tools to assess the effects of drugs
on mitochondrial function.
7 Besides isolated mitochondria, it is also possible to
investigate electron transfer along the MRC thanks to
submitochondrial particules (SMPs). SMPs are prepared
from isolated mitochondria by sonication, and consist of
inverted, vesicular portions of the inner mitochondrial
membrane [107]. With SMPs, NADH can be used as an
electron donor to complex I instead of the glutamate/
malate substrates usually employed. Hence, one can take
advantage of this feature if one suspects that a drug can
inhibit the substrate carriers of the inner mitochondrial
membrane, or enzymes involved in substrate processing
and NADH generation. For instance, studies have shown
that valproic acid inhibits pyruvate and succinate
transport as well as glutamate dehydrogenase activity
[3].
8 Finally, it is important to mention that investigations
are currently being performed in order to set up and
validate high-throughput assays for mitochondrial
toxicity screening. Yvonne Will and colleagues have
developed high-throughput-applicable screens based on
luminescent, oxygen-sensitive probes and immunocap-
ture-based MRC complex assays, which allow the assess-
ment of oxygen consumption in isolated mitochondria,
or the measurement of the activities of different MRC
complexes, respectively [58,108,109].
Plasma metabolites and enzymes
In vivo investigations aimed at detecting the toxicity of a
drug on liver mitochondria may include the measure-
ment of several metabolites and enzymes in plasma
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 345
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
(Table VII). Regarding plasma chemistry, the European
Medicines Agency (EMEA) recommends the determina-
tion of lactate levels, as well as glutamate dehydrogenase
(GLDH) and ornithine carbamoyltransferase (OCT) activ-
ities [110]. GLDH and OCT are mitochondrial enzymes.
An increased plasma activity of these enzymes reects
structural damage to the mitochondria and cell mem-
brane, leading to the leakage of these enzymes into the
plasma [111,112]. OCT is particularly expressed in the
liver [111]. Therefore, a high plasma OCT activity can
occur when mitochondrial damage has specically
caused liver injury.
Elevated lactate (and/or pyruvate) levels indicate an
insufcient oxidation of pyruvate by mitochondria. This
may not only be due to drug-induced mitochondrial
dysfunction, but also to many other causes, including
anoxia and/or shock. It must be also pointed out that
hyperlactatemia is not strictly specic to mitochondrial
dysfunction in liver. We therefore also suggest measur-
ing the two plasma ketone bodies, b-hydroxybutyrate
and acetoacetate. The b-hydroxybutyrate/acetoacetate
ratio mostly reects the NADH/NAD
+
ratio present in
hepatic mitochondria [3,113], and therefore provides
indirect information on the activity of the hepatic MRC.
When the hepatic MRC is inhibited, the insufcient re-
oxidation of NADH into NAD
+
increases the NADH/
NAD
+
ratio in the liver. The increase in NADH, in turn,
favours the reduction of acetoacetate into b-hydroxybu-
tyrate by b-hydroxybutyrate dehydrogenase, a mito-
chondrial enzyme mainly expressed in the liver
[114,115]. The b-hydroxybutyrate/acetoacetate ratio
therefore increases when the MRC is inhibited in the liver
not only at the level of complex I but also downstream
to this site [116].
In addition to the b-hydroxybutyrate/acetoacetate
ratio, one should also consider the sum of b-hydroxybu-
tyrate and acetoacetate levels in plasma. FAO-inhibiting
drugs can diversely affect the plasma levels of ketone
bodies. Some FAO inhibitors such as valproic acid,
ibuprofen and pirprofen decreased, as expected, plasma
ketone bodies in rodents [3,56,117]. However, other
hepatic FAO inhibitors, such as aspirin, amineptine,
tianeptine, tetracycline or amiodarone instead aug-
mented plasma ketone bodies in rodents [3,118120].
Although the mechanism behind this paradoxical
increase has not been determined, it has been suggested
that the latter drugs may inhibit the activity of the
tricarboxylic acid cycle in extra-hepatic tissues, which
massively use ketone bodies as an energy source during
fasting [3].
Acyl-carnitine and acyl-glycine derivatives can be also
assessed in blood or urine, as their augmentation can
reect an inhibition of mitochondrial FAO within the
liver [121]. For instance, rats or mice treated with
valproic acid had increased levels of acyl-carnitine
derivatives in plasma (or serum) and urine [122,123].
However, it must be kept in mind that abnormal acyl-
carnitine proles can be also the secondary consequence
of an impaired MRC activity [124].
Liver histology and electron microscopy
Liver histology can also give precious information
regarding the mitochondrial origin of DILI (Table VII).
For instance, the presence of microvesicular steatosis is
highly suggestive of a strong inhibition of mitochondrial
FAO [3,4,56,117120,123]. It is noteworthy, however,
that pure microvesicular steatosis is rare, and that this
lesion is often accompanied by macrovacuolar steatosis.
The presence of steatohepatitis can also suggest drug-
induced mitochondrial dysfunction. However, investiga-
tions should be performed to test the possibility that this
liver lesion could be due also (or instead) to an increased
hepatic de novo lipogenesis, such as that triggered by
insulin resistance in a context of body fatness [13,88]. It
is important to mention that drugs can indirectly impair
mitochondrial FAO in liver and cause steatosis (or
steatohepatitis) by inducing lipoatrophy (i.e. a reduction
in body fat mass) [125]. Fatty liver (or steatohepatitis)
secondary to lipoatrophy is, at least in part, the con-
sequence of a lower production of leptin by the white
Table VII Examples of in vivo or ex vivo investigations which can
be performed to detect drug-induced mitochondrial dysfunction.
Plasma biochemistry: lactate (and pyruvate), ketone bodies
(b-hydroxybutyrate and acetoacetate), GLDH
a
and OCT
b
activities,
acyl-carnitine derivatives
Urine biochemistry: acyl-carnitine and acyl-glycine derivatives
Histopathology (with oil red O and haematoxylin-eosin staining) and electron
microscopy
Assessment of whole-body fatty acid oxidation (FAO) after administration
of
14
C-labelled fatty acids
Investigations on liver mitochondria isolated from treated animal:
measurement of mitochondrial FAO with
14
C-labelled fatty acids and/or
oxygen consumption with a Clark electrode, assessment of mitochondrial
permeability transition (MPT) pore opening by spectrophotometry
Investigations on liver homogenates prepared from treated animals:
assessment of mtDNA levels and/or activities of different mitochondrial
respiratory chain (MRC) complexes, immunoblot analysis of selected MRC
or FAO polypeptides
a
Glutamate dehydrogenase.
b
Ornithine carbamoyltransferase.
346 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
adipose tissue. Leptin insufciency in turn reduces
hepatic FAO and ketogenesis and favours de novo
lipogenesis [125,126].
The presence of apoptosis and necrosis in liver does
not specically imply direct toxicity of the drug on
mitochondria. Indeed, both kinds of cell death can be
induced by other mechanisms such as oxidative stress,
reticulum endoplasmic stress, nitric oxide overproduc-
tion, activation of kinases such as JNK, or alteration of
calcium homeostasis [127130].
Electron microscopy can also provide valuable pieces
of information (Table VII) as it can reveal not only
ultrastructural changes in liver mitochondria (enlarge-
ment or swelling, disruption of cristae), but also the
presence of small lipid droplets within the cytoplasm
suggesting microvesicular steatosis [21,131133]. How-
ever, ultrastructural alterations of mitochondria are not
necessarily a direct consequence of drug toxicity. Indeed,
these alterations can represent collateral effects linked
to the generation of free radicals (and/or lipid peroxida-
tion products) in extra-mitochondrial compartments
[134,135].
Other in vivo or ex vivo investigations
Other investigations can be performed in vivo or ex vivo
(Table VII). For instance, whole-body FAO can be
assessed thanks to the measurement of [
14
C]CO
2
exha-
lation after the administration of
14
C-labelled fatty acids
[3,42,119,120,123]. Interestingly,
14
C-labelled fatty
acids of different lengths can be used to determine
whether FAO inhibition affects the whole oxidative
process or only some chain length-specic processes
[117,119].
Alternatively, mitochondria can be isolated from the
liver of treated animals and immediately used to measure
mitochondrial FAO ex vivo, thanks to
14
C-labelled fatty
acids [3,117,119,120]. Finally, a fraction of the liver can
be frozen and subsequently used to assess mtDNA levels,
the activities of different MRC complexes, and the
expression of selected MRC and FAO proteins by immuno-
blotting [42,125,132,133,136].
Uncommon animal species
Rats and dogs (and sometimes monkeys) are the regu-
lation species commonly used by pharmaceutical com-
panies in preclinical studies. Thus, investigations in any
other animals are not required by drug agencies. Despite
this, uncommon animal species could be valuable tools
for the screenings carried out prior to the preclinical
studies, as they could help select safer drugs. Alterna-
tively, these animal species can help understand a
posteriori the complex ways whereby a marketed drug
induced hepatotoxicity in some patients, in particular,
through mitochondrial dysfunction.
Uncommon animal species such as marmots and
rabbits have been used to test aluridine and panadiplon
(Table IV) [11]. Indeed, aluridine did not induce hepa-
totoxicity in rats, dogs and nonhuman primates, but
liver toxicity similar to that in humans was revealed
when this antiviral drug was administered to wood-
chucks (Marmota monax) [44,132]. Panadiplon was
found to be safe to rats and dogs during preclinical
safety studies. However, this anxiolytic drug induced
microvesicular steatosis and centrilobular necrosis when
Dutch-belted rabbits were used for toxicological investi-
gations performed after the interruption of the clinical
trials [47,137]. Interestingly, preclinical safety investi-
gations in monkeys showed microvesicular steatosis, but
as this lesion was not accompanied with overt signs of
hepatotoxicity the clinical development of panadiplon
was continued [47].
Unfortunately, it is unknown why aluridine and
panadiplon are particularly toxic to woodchucks
and Dutch-belted rabbits, respectively. Owing to their
mechanisms of mitochondrial toxicity (Figure 1), aluri-
dine could be incorporated at a very high rate within
woodchuck hepatic mtDNA and panadiplon could
severely deplete coenzyme A and/or L-carnitine stores
in Dutch-belted rabbit liver. Obviously, these uncommon
animal species cannot be used on a regular basis in
toxicological studies, although they could be used, for
example, for investigational compounds that share
structural or pharmacological analogies with aluridine
or panadiplon. Finally, as discussed in the next section,
the use of genetically modied mice could also improve
our understanding of the complex ways whereby
a marketed drug can induce hepatotoxicity in some
patients, in particular when DILI is mediated by
mitochondrial dysfunction [101,123,138,139]. More
generally, signicant efforts should be devoted to assess-
ing the potential of different animal models in detecting
mitochondrial toxicity-mediated DILI.
FUTURE DI RECTI ONS
A major challenge for pharmaceutical companies in the
future will be to develop drugs, which do not induce
mitochondrial dysfunction in liver and in other tissues
too. Improvement of in vitro screenings by using high-
throughput technologies during lead selection would
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 347
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
certainly help select safer compounds for subsequent
in vivo preclinical safety investigations [7]. Recent
investigations also suggest that better predictions could
be possible thanks to the use of novel animal models
such as the heterozygous MnSOD
+/)
(sod2
+/)
) knockout
mouse [138]. Although sod2
+/)
mice are apparently
normal, they present underlying mitochondrial altera-
tions in liver and other tissues, such as decreased Dw
m
,
reduced MRC activity, and an enhanced propensity to
undergo MPT [140,141]. Interestingly, sod2
+/)
mice
treated with troglitazone developed increased serum
alanine aminotransferase activity and hepatic necrosis
whereas wild-type mice did not [136]. Using the sod2
+/)
mice model can also unmask the mitochondrial toxicity
of nimesulide and stavudine [138,142].
Another murine model with underlying mitochon-
drial/metabolic dysfunction is the juvenile visceral
steatosis (jvs) mouse. Heterozygous jvs
+/)
mice have
decreased L-carnitine stores because of impaired renal re-
absorption of this mitochondrial FAO cofactor, and thus
should be expected to be more susceptible to FAO-
impairing drugs [123]. It would be also helpful to use
MCAD
+/)
mice [143], as partial MCAD deciency is one
of the most frequent metabolic abnormalities of genetic
origin among Caucasians, as previously mentioned.
Finally, because of the increasing prevalence of obesity
and NAFLD, and the propensity of NAFLD, and partic-
ularly non-alcoholic steatohepatitis (NASH), to alter
mitochondrial function [13,88], it could be also valuable
to use rodent models of obesity and pre-existing NAFLD
such as ob/ob mice (which are leptin decient), or db/db
mice and fa/fa rats (which both are decient in leptin
receptor) [9496,144]. However, further studies are
needed to assess the interest if any of these different
animal models.
CONCLUSI ONS
Together with the metabolic activation of drugs into
reactive metabolites, mitochondrial dysfunction is yet
another frequent cause of DILI. Severe side effects such
as microvesicular steatosis, lactic acidosis or rhabdo-
myolysis can threaten the life of patients, and can lead to
the premature interruption of clinical trials, or the recall
of an unsafe medicine. The enormous investments made
for drug development are not recovered, which can
jeopardize the nancial viability and the image of a drug
company. A major challenge for the pharmaceutical
industry is therefore to be able to detect drug-induced
mitochondrial dysfunction during preclinical studies. We
therefore strongly recommend performing mechanism-
based investigations during preclinical safety studies in
order to assess whether a selected drug is able to impair
mitochondrial function, in particular in the liver. The
development and improvement of high-throughput
in vitro screening techniques should also allow pharma-
ceutical companies to assess mitochondrial effects during
lead selection, before subsequent preclinical safety inves-
tigations [7]. Finally, further investigations should be
carried out in mice with pre-existing mitochondrial and/
or metabolic abnormalities, such as sod
+/)
, jvs
+/)
or
ob/ob mice. Indeed, data collected from such investiga-
tions could help to determine if these animal models can
prove useful in the future to detect mitochondrial
dysfunction-mediated DILI, which could selectively occur
in patients with pre-existing mitochondrial and/or met-
abolic abnormalities.
REFERENCES
1 Lee W.M. Drug-induced hepatotoxicity. N. Engl. J. Med.
(2003) 349 474485.
2 Pessayre D., Fromenty B., Mansouri A., Berson A. Hepato-
toxicity due to mitochondrial injury, in: Kaplowitz N., DeLeve
L.D. (Eds), Drug-induced liver disease, 2nd edn. Informa
Healthcare, New York, 2007, pp. 4984.
3 Fromenty B., Pessayre D. Inhibition of mitochondrial beta-
oxidation as a mechanism of hepatotoxicity. Pharmacol. Ther.
(1995) 67 101154.
4 Pessayre D., Mansouri A., Haouzi D., Fromenty B. Hepato-
toxicity due to mitochondrial dysfunction. Cell Biol. Toxicol.
(1999) 15 367373.
5 Gougeon M.L., Penicaud L., Fromenty B., Leclercq P., Viard
J.P., Capeau J. Adipocytes targets and actors in the patho-
genesis of HIV-associated lipodystrophy and metabolic alter-
ations. Antivir. Ther. (2004) 9 161177.
6 Igoudjil A., Begriche K., Pessayre D., Fromenty B. Mitochon-
drial, metabolic and genotoxic effects of antiretroviral nucle-
oside reverse-transcriptase inhibitors. Curr. Med. Chem.
Anti-Infective Agents (2006) 5 273292.
7 Dykens J.A., Will Y. The signicance of mitochondrial toxicity
testing in drug development. Drug Discov. Today (2007) 12
777785.
8 Scatena R., Bottoni P., Botta G., Martorana G.E., Giardina B.
The role of mitochondria in pharmacotoxicology: a reevalu-
ation of an old, newly emerging topic. Am. J. Physiol. Cell
Physiol. (2007) 293 C12C21.
9 Ballet F. Hepatotoxicity in drug development: detection,
signicance and solutions. J. Hepatol. (1997) 26(Suppl 2) 26
36.
10 FDA Working Group. Nonclinical assessment of potential
hepatotoxicity in man; available at: http://www.fda.gov/cder/
livertox/preclinical.pdf (2000).
348 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
11 Peters T.S. Do preclinical testing strategies help predict human
hepatotoxic potentials? Toxicol. Pathol. (2005) 33 146154.
12 Salway J.G. Metabolism at a glance, 1st edn. Blackwell
Scientic Publications, London, 1994.
13 Pessayre D., Fromenty B. NASH: a mitochondrial disease.
J. Hepatol. (2005) 42 928940.
14 Navarro A., Boveris A. The mitochondrial energy transduction
system and the aging process. Am. J. Physiol. Cell Physiol.
(2007) 292 C670C686.
15 Moraes C.T. What regulates mitochondrial DNA copy number
in animal cells? Trends Genet. (2001) 17 199205.
16 Kim I., Rodriguez-Enriquez S., Lemasters J.J. Selective degra-
dation of mitochondria by autophagia. Arch. Biochem.
Biophys. (2007) 462 245253.
17 Maaswinkel-Mooij P.D., Van den Bogert C., Scholte H.R.,
Onkenhout W., Brederoo P., Poorthuis B.J. Depletion of
mitochondrial DNA in the liver of a patient with lactic
academia and hypoketotic hypoglycaemia. J. Pediatr. (1996)
128 679683.
18 Park S.Y., Chang I., Kim J.Y. et al. Resistance of mitochondrial
DNA-depleted cells against cell death. Role of mitochondrial
superoxide dismutase. J. Biol. Chem. (2004) 279 75127520.
19 Velsor LW, Kovacevic M, Goldstein M, Leitner HM, Lewis W,
Day BJ. Mitochondrial oxidative stress in human hepatoma
cells exposed to stavudine. Toxicol. Appl. Pharmacol. (2004)
199 1019.
20 Dianov G.L., Souza-Pinto N., Nyaga S.G., Thybo T., Stevnsner
T., Bohr V.A. Base excision repair in nuclear and mitochondrial
DNA. Prog. Nucleic Acid Res. Mol. Biol. (2001) 68 285297.
21 Demeilliers C., Maisonneuve C., Grodet A. et al. Impaired
adaptive resynthesis and prolonged depletion of hepatic
mitochondrial DNA after repeated alcohol binges in mice.
Gastroenterology (2002) 123 12781290.
22 Fernandez-Checa J.C., Kaplowitz N. Hepatic mitochondrial
glutathione: transport and role in disease and toxicity. Toxicol.
Appl. Pharmacol. (2005) 204 263273.
23 Brookes P.S. Mitochondrial H
+
leak and ROS generation:
an odd couple. Free Radic. Biol. Med. (2005) 38 1223.
24 Chiarugi A. Simple but not simpler: toward an unied
picture of energy requirements in cell death. FASEB J. (2005)
19 17831788.
25 Zoratti M., Szabo` I. The mitochondrial permeability transition.
Biochem. Biophys. Acta (1995) 1241 139176.
26 Yerushalmi B., Dahl R., Devereaux M.W., Gumpricht E., Sokol
R.J. Bile acid-induced rat hepatocyte apoptosis is inhibited by
antioxidants and blockers of the mitochondrial permeability
transition. Hepatology (2001) 33 616626.
27 Feldmann G., Haouzi D., Moreau A. et al. Opening of the
mitochondrial permeability transition pore causes matrix
expansion and outer membrane rupture in Fas-mediated
hepatic apoptosis in mice. Hepatology (2000) 31 674683.
28 Chen X., Ding W.X., Ni H.M. et al. Bid-independent mito-
chondrial activation in tumor necrosis alpha-induced apop-
tosis and liver injury. Mol. Cell. Biol. (2007) 27 541553.
29 Trost L.C., Lemasters J.J. The mitochondrial permeability
transition: a new pathophysiological mechanism for Reyes
syndrome and toxic liver injury. J. Pharmacol. Exp. Ther.
(1996) 278 10001005.
30 Masubuchi Y., Nakayama S., Horie T. Role of mitochondrial
permeability transition in diclofenac-induced hepatocyte
injury in rats. Hepatology (2002) 35 544551.
31 Masubuchi Y., Kano S., Horie T. Mitochondrial permeability
transition as a potential determinant of hepatotoxicity of
antidiabetic thialozidinediones. Toxicology (2006) 222 233
239.
32 Bae M.A., Song B.J. Critical role of c-Jun N-terminal protein
kinase in troglitazone-induced apoptosis of human HepG2
hepatoma cells. Mol. Pharmacol. (2003) 63 401408.
33 Gunawan B.K., Liu Z.X., Han D., Hanawa N., Gaarde W.A.,
Kaplowitz N. c-Jun N-terminal kinase plays a major role in
murine acetaminophen hepatotoxicity. Gastroenterology
(2006) 131 165178.
34 Haouzi D., Lekehal M., Moreau A. et al. Cytochrome P450-
generated reactive metabolites cause mitochondrial per-
meability transition, caspase activation and apoptosis in rat
hepatocytes. Hepatology (2000) 32 303311.
35 Deschamps D., Fisch C., Fromenty B., Berson A., Degott C.,
Pessayre D. Inhibition by salicylic acid of the activation and
thus oxidation of long chain fatty acids. Possible role in the
development of Reyes syndrome. J. Pharmacol. Exp. Ther.
(1991) 259 894904.
36 Hamada N., Ogawa Y., Saibara T. et al. Toremifene-induced
fatty liver and NASH in breast cancer patients with
breast-conservation treatment. Int. J. Oncol. (2000) 17 1119
1123.
37 Farrell G.C. Drugs and steatohepatitis. Semin. Liver Dis.
(2002) 22 185194.
38 Fernandez F.G., Ritter J., Goodwin J.W, Linehan D.C., Hawkins
W.G., Strasberg S.M. Effect of steatohepatitis associated with
irinotecan and oxaliplatin pretreatment on respectability of
hepatic colorectal metastases. J. Am. Coll. Surg. (2005) 200
845853.
39 Vauthey J.N., Pawlik T.M., Ribero D. et al. Chemotherapy
regimen predicts steatohepatitis and an increase in 90-day
mortality after surgery for hepatic colorectal metastases.
J. Clin. Oncol. (2006) 24 20652072.
40 Berson A., De Beco V., Letteron P. et al. Steatohepatitis-
inducing drugs cause mitochondrial dysfunction and lipid
peroxidation in rat hepatocytes. Gastroenterology (1998) 114
764774.
41 Moreira P.I., Custodio J., Moreno A., Oliveira C.R., Santos M.S.
Tamoxifen and estradiol interact with the avin mono-
nucleotide site of complex I leading to mitochondrial failure.
J. Biol. Chem. (2006) 281 1014310152.
42 Larosche I., Letteron P., Fromenty B. et al. Tamoxifen inhibits
topoisomerases, depletes mitochondrial DNA and triggers
steatosis in mouse liver. J. Pharmacol. Exp. Ther. (2007) 321
526535.
43 Letteron P., Sutton A., Mansouri A., Fromenty B., Pessayre D.
Inhibition of microsomal triglyceride transfer protein: another
mechanism for drug-induced steatosis in mice. Hepatology
(2003) 38 133140.
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 349
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
44 Tennant B.C., Baldwin B.H., Graham L.A. et al. Antiviral
activity and toxicity of aluridine in the woodchuck model of
hepatitis B virus infection. Hepatology (1998) 28 179191.
45 Lewis W., Levine E.S., Griniuviene B. et al. Fialuridine and its
metabolites inhibit DNA polymerase c at sites of multiple
adjacent analog incorporation, decrease mtDNA abundance,
and cause mitochondrial structural defects in cultured hepato-
blasts. Proc. Natl Acad. Sci. USA (1996) 93 35923597.
46 Richardson F.C., Engelhardt J.A., Bowsher R.R. Fialuridine
accumulates in DNA of dogs, monkeys, and rats following
long-term oral administration. Proc. Natl Acad. Sci. USA
(1994) 91 1200312007.
47 Ulrich R.G., Bacon J.A., Brass E.P., Cramer C.T., Petrella D.K.,
Sun E.L. Metabolic, idiosyncratic toxicity of drugs: overview of
the hepatic toxicity induced by the anxiolytic, panadiplon.
Chem. Biol. Interact. (2001) 134 251270.
48 Baty V., Denis B., Goudot C. et al. Hepatitis induced by
alpidem (Ananxyl). Four cases, one of them fatal. Gastroen-
terol. Clin. Biol. (1994) 18 11291131.
49 Berson A., Descatoire V., Sutton A. et al. Toxicity of alpidem,
a peripheral benzodiazepine receptor ligand, but not zolpidem,
in rat hepatocytes: role of mitochondrial permeability transi-
tion and metabolic activation. J. Pharmacol. Exp. Ther. (2001)
299 793800.
50 Ashraan H., Horowitz J.D., Frenneaux M.P. Perhexiline.
Cardiovasc. Drug Rev. (2007) 25 7697.
51 Deschamps D., DeBeco V., Fisch C., Fromenty B., Guillouzo A.,
Pessayre D. Inhibition by perhexiline of oxidative phosphor-
ylation and the b-oxidation of fatty acids: possible role in
pseudoalcoholic liver lesions. Hepatology (1994) 19 948961.
52 Fromenty B., Fisch C., Berson A, Letteron P., Larrey D.,
Pessayre D. Dual effect of amiodarone on mitochondrial
respiration. Initial protonophoric uncoupling effect followed
by inhibition of the respiratory chain at the levels of complex I
and complex II. J. Pharmacol. Exp. Ther. (1990) 255 1377
1384.
53 Kennedy J.A., Unger S.A., Horowitz J.D. Inhibition of
carnitine palmitoyltransferase-1 in rat heart and liver by
perhexiline and amiodarone. Biochem. Pharmacol. (1996)
52 273280.
54 Spaniol M., Bracher R., Ha H.R., Follath F., Krahenbu hl S.
Toxicity of amiodarone and amiodarone analogues on isolated
rat liver mitochondria. J. Hepatol. (2001) 35 628636.
55 Danan G., Trunet P., Bernuau J. et al. Pirprofen-induced
fulminant hepatitis. Gastroenterology (1985) 89 210213.
56 Gene`ve J., Hayat-Bonan B., Labbe G. et al. Inhibition of
mitochondrial -oxidation of fatty acids by pirprofen. Role
in microvesicular steatosis due to this nonsteroidal anti-
inammatory drug. J. Pharmacol. Exp. Ther. (1987) 242
11331137.
57 Chitturi S., George J. Hepatotoxicity of commonly used drugs:
nonsteroidal anti-inammatory drugs, antihypertensives,
antidiabetic agents, anticonvulsants, lipid-lowering agents,
psychotropic drugs. Semin. Liver Dis. (2002) 22 169183.
58 Nadanaciva S., Dykens J.A., Bernal A., Capaldi R.A., Will Y.
Mitochondrial impairment by PPAR agonists and statins
identied via immunocaptured OXPHOS complex activities
and respiration. Toxicol. Appl. Pharmacol. (2007) 223 277
287.
59 Chabrol B., Mancini J., Chretien D., Rustin P., Munnich A.,
Pinsard N. Valproate-induced hepatic failure in a case of
cytochrome c oxidase deciency. Eur. J. Pediatr. (1994) 153
133135.
60 Njolstad P.R., Skjeldal O.H., Agsteribbe E. et al. Medium chain
acyl-CoA dehydrogenase deciency and fatal valproate toxic-
ity. Pediatr. Neurol. (1997) 16 160162.
61 Krahenbu hl S., Brandner S., Kleinle S., Liechti S., Straumann
D. Mitochondrial diseases represent a risk factor for
valproate-induced fulminant liver failure. Liver (2000) 20
346348.
62 de Vries H.G., Niezen-Koning K., Kliphuis J.W., Smit G.P.,
Scheffer H., the Kare L.P. Prevalence of carriers of the most
common medium-chain acyl-CoA dehydrogenase (MCAD)
deciency mutation (G985A) in The Netherlands. Hum.
Genet. (1996) 98 12.
63 Fromenty B., Mansouri A., Bonnefont J.P. et al. Most cases of
medium-chain acyl-CoA dehydrogenase deciency escape
detection in France. Hum. Genet. (1996) 97 367368.
64 Tanaka K., Gregersen N., Ribes A. et al. A survey of the
newborn populations in Belgium, Germany, Poland, Czech
Republic, Hungary, Bulgaria, Spain, Turkey, and Japan for the
G985 variant allele with haplotype analysis at the medium
chain acyl-CoA dehydrogenase gene locus: clinical and
evolutionary consideration. Pediatr. Res. (1997) 41 201209.
65 Yamanaka H., Gatanaga H., Kosalaraksa P. et al. Novel
mutation of human DNA polymerase c associated with
mitochondrial toxicity induced by anti-HIV treatment.
J. Infect. Dis. (2007) 195 14191425.
66 Vittecoq D., Jardel C., Barthelemy C. et al. Mitochondrial
damage associated with long-term antiretroviral treatment:
associated alteration or causal disorder? J. Acquir. Immune
Dec. Syndr. (2002) 31 299308.
67 Walker U.A., Bauerle J., Laguno M. et al. Depletion of
mitochondrial DNA in liver under antiretroviral therapy with
didanosine, stavudine, or zalcitabine. Hepatology (2004) 39
311317.
68 Huang Y.S., Su W.J., Huang Y.H. Genetic polymorphisms of
manganese superoxide dismutase, NAD(P)H:quinone oxido-
reductase, glutathione S-transferase M1 and T1, and the
susceptibility to drug-induced liver injury. J. Hepatol. (2007)
47 128134.
69 Degoul F., Sutton A., Mansouri A. et al. Homozygosity for
alanine in the mitochondrial targeting sequence of superoxide
dismutase and risk for severe alcoholic liver disease. Gastro-
enterology (2001) 120 14681474.
70 Nahon P., Sutton A., Pessayre D. et al. Genetic susceptibility
to alcoholic cirrhosis, hepatocellular carcinoma and death.
Clin. Gastroenterol. Hepatol. (2005) 3 292298.
71 Sutton A., Khoury H., Prip-Buus C. et al. The Ala16Val
genetic dimorphism modulates the import of human manga-
nese superoxide dismutase into rat liver mitochondria.
Pharmacogenetics (2003) 13 145157.
350 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
72 Sutton A., Imbert A., Igoudjil A. et al. The manganese
superoxide dismutase Ala16Val dimorphism modulates both
mitochondrial import and mRNA stability. Pharmacogenet.
Genomics (2005) 15 311319.
73 Sutton A., Nahon P., Pessayre D. et al. Genetic polymorphisms
in antioxidant enzymes modulate hepatic iron accumulation
and hepatocellular carcinoma development in patients with
alcohol-induced cirrhosis. Cancer Res. (2006) 66 28442852.
74 Fromenty B., Grimbert S., Mansouri A. et al. Hepatic mito-
chondrial DNA deletion in alcoholics: association with
microvesicular steatosis. Gastroenterology (1995) 108 193
200.
75 Cahill A., Cunningham C.C., Adachi M. et al. Effects of alcohol
and oxidative stress on liver pathology: the role of the
mitochondrion. Alcohol. Clin. Exp. Res. (2002) 26 907915.
76 Pessayre D., Larrey D. Drug-induced liver injury, in: Rodes J.,
Benhamou J.P., Blei A.T., Reichen J., Rizzetto M. (Eds),
Textbook of hepatology: from basic science to clinical practice,
Blackwell Publishing, Oxford, 2007, pp. 12111268.
77 Robin M.A., Sauvage I., Grandperret T., Descatoire V.,
Pessayre D., Fromenty B. Ethanol increases mitochondrial
cytochrome P450 2E1 in mouse liver and rat hepatocytes.
FEBS Lett. (2005) 579 68956902.
78 Weis M., Kass G.E., Orrenius S., Moldeus P. N-acetyl-
p-benzoquinone imine induces Ca
2+
release from mitochon-
dria by stimulating pyridine nucleotide hydrolysis. J. Biol.
Chem. (1992) 267 804809.
79 Masubuchi Y., Suda C., Horie T. Involvement of mitochondrial
permeability transition in acetaminophen-induced liver injury
in mice. J. Hepatol. (2005) 42 110116.
80 Anandatheerthavarada H.K., Addya S., Dwivedi R.S., Biswas
G., Mullick J., Avadhani N.G. Localization of multiple forms
of inducible cytochromes P450 in rat liver mitochondria:
immunological characteristics and patterns of xenobiotic
substrate metabolism. Arch. Biochem. Biophys. (1997) 339
136150.
81 Sepuri N.B., Yadav S., Anandatheerthavarada H.K., Avadhani
N.G. Mitochondrial targeting of intact CYP2B1 and CYP2E1
and N-terminal truncated CYP1A1 proteins in Saccharomyces
cerevisiae: role of protein kinase A in the mitochondrial
targeting of CYP2E1. FEBS J. (2007) 274 46154630.
82 Neuman M.G., Shear N.H., Jacobson-Brown P.M. et al.
CYP2E1-mediated modulation of valproic acid-induced
hepatotoxicity. Clin. Biochem. (2001) 34 211218.
83 Belay E.D., Bresee J.S., Holman R.C., Khan A.S., Shahriari A.,
Schonberger L.B. Reyes syndrome in the United States from
1981 through 1997. N. Engl. J. Med. (1999) 340 1377
1382.
84 Novak D., Lewis J.H. Drug-induced liver disease. Curr. Opin.
Gastroenterol. (2003) 19 203215.
85 Rahmani Z., Huh K.W., Lasher R., Siddiqui A. Hepatitis B
virus X protein colocalizes to mitochondria with a human
voltage-dependent anion channel, HVDAC3, and alters its
transmembrane potential. J. Virol. (2000) 74 28402846.
86 Korenaga M., Wang T., Li Y. et al. Hepatitis C virus core
protein inhibits mitochondrial electron transport and
increases reactive oxygen species (ROS) production. J. Biol.
Chem. (2005) 280 3748137488.
87 Piccoli C., Scrima R., DAprile A. et al. Mitochondrial dys-
function in hepatitis C virus infection. Biochim. Biophys. Acta
(2006) 1757 14291437.
88 Begriche K., Igoudjil A., Pessayre D., Fromenty B. Mitochon-
drial dysfunction in NASH: causes, consequences and possible
means to prevent it. Mitochondrion (2006) 6 128.
89 Bellentani S., Saccoccio G., Masutti F. et al. Prevalence of and
risk factors for hepatic steatosis in Northern Italy. Ann. Intern.
Med. (2000) 132 112117.
90 Langman G., Hall P.M., Todd G. Role of non-alcoholic
steatohepatitis in methotrexate-induced liver injury.
J. Gastroenterol. Hepatol. (2001) 16 13951401.
91 Callery M.P. Chemotherapy-related steatohepatitis: proceed
with caution. Gastroenterology (2005) 129 2114215.
92 Bruno S., Maisonneuve P., Castellana P. et al. Incidence and
risk factors for non-alcoholic steatohepatitis: prospective study
of 5408 women enrolled in Italian tamoxifen chemopreven-
tion trial. BMJ (2005) 330 932935.
93 Tarantino G., Conca P., Basile V. et al. A prospective study
of acute drug-induced liver injury in patients suffering from
non-alcoholic fatty liver disease. Hepatol. Res. (2007) 37
410415.
94 Robin M.A., Demeilliers C., Sutton A. et al. Alcohol increases
tumor necrosis factor-a and decreases nuclear factor-jB to
activate hepatic apoptosis in genetically obese mice. Hepato-
logy (2005) 42 12801290.
95 Carmiel-Haggai M., Cederbaum A.I., Nieto N. Binge ethanol
exposure increases liver injury in obese rats. Gastroenterology
(2003) 125 18181833.
96 Tomaru M., Takano H., Inoue K. et al. Pulmonary exposure to
diesel exhaust particles enhances fatty change of the liver in
obese diabetic mice. Int. J. Mol. Med. (2007) 19 1722.
97 Berson A., Cazanave S., Descatoire V. et al. The anti-inam-
matory drug, nimesulide (4-nitro-2-phenoxymethane-sulfo-
anilide), uncouples mitochondria and induces mitochondrial
permeability transition in human hepatoma cells: protection
by albumin. J. Pharmacol. Exp. Ther. (2006) 318 444454.
98 Fromenty B., Letteron P., Fisch C., Berson A., Deschamps D.
Evaluation of human blood lymphocytes as a model to study
the effects of drugs on human mitochondria. Effects of low
concentrations of amiodarone on fatty acid oxidation, ATP
levels and cell survival. Biochem. Pharmacol. (1993) 46 421
432.
99 Aninat C., Piton A., Glaise D. et al. Expression of cytochromes
P450, conjugating enzymes and nuclear receptors in human
hepatoma HepaRGcells. Drug Metab. Dispos. (2006) 347583.
100 Guillouzo A., Corlu A., Aninat C., Glaise D., Morel F., Guguen-
Guillouzo C. The human hepatoma HepaRG cells: a highly
differentiated model for studies of liver metabolism and toxicity
of xenobiotics. Chem. Biol. Interact. (2007) 168 6673.
101 James L.P., Mayeux P.R., Hinson J.A. Acetaminophen-induced
hepatotoxicity. Drug Metab. Dispos. (2003) 31 14991506.
102 Marroquin L.D., Hynes J., Dykens J.A., Jamieson J.D., Will Y.
Circumventing the Crabtree effect: replacing media glucose
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 351
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
with galactose increases susceptibility of HepG2 cells to
mitochondrial toxicants. Toxicol. Sci. (2007) 97 539547.
103 Blackard W., Clore J.N., Powers L.P. A stimulatory effect of
FFA on glycolysis unmasked in cells with impaired oxidative
capacity. Am. J. Physiol. (1990) 259 E451E456.
104 Prip-Buus C., Bouthillier-Voisin A.C., Kohl C., Demaugre F.,
Girard J., Pegorier J.P. Evidence for an impaired long-chain
fatty acid oxidation and ketogenesis in Fao hepatoma cells.
Eur. J. Biochem. (1992) 209 291298.
105 Igoudjil A., Massart J., Begriche K., Descatoire V., Robin M.A.,
Fromenty B. High concentrations of stavudine impair fatty acid
oxidation without depleting mitochondrial DNA in cultured
rat hepatocytes. Toxicol. In Vitro (2008) 22 887898.
106 Stephenne X., Najimi M., Ngoc D.K. et al. Cryopreservation of
human hepatocytes alters the mitochondrial respiratory chain
complex 1. Cell Transplant. (2007) 16 409419.
107 Oakes D.J., Pollak J.K. Effects of a herbicide formulation,
Tordon 75D, and its individual components on the oxidative
functions of mitochondria. Toxicology (1999) 136 4152.
108 Nadanaciva S., Bernal A., Aggeler R., Capaldi R.A., Will Y.
Target identication of drug induced mitochondrial toxicity
using immunocapture based OXPHOS activity assays. Toxicol.
In Vitro (2007) 21 902911.
109 Hynes J., Marroquin L.D., Ogurtsov V.I. et al. Investigation
of drug-induced mitochondrial toxicity using uorescence-
based oxygen-sensitive probes. Toxicol. Sci. (2006) 92 186
200.
110 Committee for Medicinal Products for Human Use (CHMP) of
the European Medicines Agency (EMEA). Guideline on detec-
tion of early signals of drug-induced hepatotoxicity in non-
clinical studies; available at: http://www.emea.europa.eu/
pdfs/human/swp/15011506en.pdf (2006).
111 Murayama H., Ikemoto M., Fukuda Y., Tsunekawa S., Nagata
A. Advantage of serum type-I arginase and ornithine carba-
moyltransferase in the evaluation of acute and chronic liver
damage induced by thioacetamide in rats. Clin. Chim. Acta
(2007) 375 6368.
112 Rodriguez L.C., Araujo C.R., Posleman S.E., Rey Mdel R.
Hepatotoxic effect of cyclosporin A in the mitochondrial
respiratory chain. J. Appl. Toxicol. (2007) 27 310317.
113 Ozawa K., Chance B., Tanaka A., Iwata S., Kitai T., Ikai I.
Linear correlation between acetoacetate/b-hydroxybutyrate in
arterial blood and oxidized avoprotein/reduced pyridine
nucleotide in freeze-trapped human liver tissue. Biochim.
Biophys. Acta (1992) 1138 350352.
114 Lehninger A.L., Sudduth H.C., Wise J.B. D-b-Hydroxybutyric
dehydrogenase of mitochondria. J. Biol. Chem. (1960) 235
24502455.
115 Coquard C., Adami P., Cherkaoui-Malki M., Fellmann D.,
Latruffe N. Immunological study of the tissue expression of
D-b-hydroxybutyrate dehydrogenase, a ketone body-convert-
ing enzyme. Biol. Cell (1987) 59 137143.
116 Treem W.R., Sokol R.J. Disorders of the mitochondria. Semin.
Liver Dis. (1998) 18 237253.
117 Freneaux E., Fromenty B., Berson A. et al. Stereoselective and
nonstereoselective effects of ibuprofen enantiomers on mito-
chondrial b-oxidation of fatty acids. J. Pharmacol. Exp. Ther.
(1990) 255 529535.
118 Le Dinh T., Freneaux E., Labbe G. et al. Amineptine, a
tricyclic antidepressant, inhibits the mitochondrial oxidation
of fatty acids and produces microvesicular steatosis of the
liver in mice. J. Pharmacol. Exp. Ther. (1988) 247 745
750.
119 Fromenty B., Freneaux E., Labbe G. et al. Tianeptine, a new
tricyclic antidepressant metabolized by b-oxidation of its
heptanoic chain, inhibits the mitochondrial oxidation of
medium and short chain fatty acids in mice. Biochem.
Pharmacol. (1989) 38 37433751.
120 Fromenty B., Fisch C., Labbe G. et al. Amiodarone inhibits
the mitochondrial b-oxidation of fatty acids and produces
microvesicular steatosis of the liver in mice. J. Pharmacol. Exp.
Ther. (1990) 255 13711376.
121 Rinaldo P. Fatty acid transport and mitochondrial oxidation
disorders. Semin. Liver Dis. (2001) 21 489500.
122 Nishida N., Sugimoto T., Araki A., Woo M., Sakane Y.,
Kobayashi Y. Carnitine metabolism in valproate-treated rats:
the effect of L-carnitine supplementation. Pediatr. Res. (1987)
22 500503.
123 Knapp A.C., Todesco L., Beier K. et al. Toxicity of valproic acid
in mice with decreased plasma and tissue carnitine stores.
J. Pharmacol. Exp. Ther. (2008) 324 568575.
124 Sim K.G., Carpenter K., Hammond J., Christodoulou J.,
Wilcken B. Acylcarnitine proles in broblasts from patients
with respiratory chain defects can resemble those from
patients with mitochondrial fatty acid b-oxidation disorders.
Metabolism (2002) 51 366371.
125 Igoudjil A., Abbey-Toby A., Begriche K. et al. High doses of
stavudine induce fat wasting and mild liver damage without
impairing mitochondrial respiration in mice. Antivir. Ther.
(2007) 12 389400.
126 Beltrand J., Beregszaszi M., Chevenne D. et al. Metabolic
correction induced by leptin replacement treatment in young
children with Berardinelli-Seip congenital lipoatrophy. Pedi-
atrics (2007) 120 e291296.
127 Nicotera P., Bellomo G., Orrenius S. Calcium-mediated mech-
anisms in chemically induced cell death. Annu. Rev. Phar-
macol. Toxicol. (1992) 32 449470.
128 Fau D., Lekehal M., Farrell G. et al. Diterpenoids from
germander, an herbal medicine, induce apoptosis in
isolated rat hepatocytes. Gastroenterology (1997) 113
13341346.
129 Gardner C.R., Heck D.E., Yang C.S. et al. Role of nitric oxide in
acetaminophen-induced hepatotoxicity in the rat. Hepatology
(1998) 26 748754.
130 Kaplowitz N. Biochemical and cellular mechanisms of toxic
liver injury. Semin. Liver Dis. (2002) 22 137144.
131 Sugimoto T., Araki A., Nishida N. et al. Hepatotoxicity in rat
following administration of valproic acid: effect of L-carnitine
supplementation. Epilepsia (1987) 28 373377.
132 Lewis W., Griniuviene B., Tankersley K.O. et al. Depletion of
mitochondrial DNA, destruction of mitochondria, and
accumulation of lipid droplets result from aluridine treat-
352 G. Labbe et al.
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353
ment in woodchucks (Marmota monax). Lab. Invest. (1997)
76 7787.
133 Lebrecht D., Vargas-Infante Y.A., Setzer B., Kirschner J.,
Walker U.A. Uridine supplementation antagonizes zalcitabine-
induced microvesicular steatohepatitis in mice. Hepatology
(2007) 45 7279.
134 Matsuhashi T., Liu X., Nishizawa Y., Usukura J., Wozniak M.,
Wakabayashi T. Mechanism of the formation of megamito-
chondria in the mouse liver induced by chloramphenicol.
Toxicol. Lett. (1996) 86 4754.
135 Matsuhashi T., Liu X., Karbowski M., Wozniak M., An-
tosiewicz J., Wakabayashi T. Role of free radicals in the
mechanism of the hydrazine-induced formation of mega-
mitochondria. Free Radic. Biol. Med. (1997) 23 285293.
136 Ong M.M., Latchoumycandane C., Boelsterli U.A. Troglitaz-
one-induced hepatic necrosis in an animal model of silent
genetic mitochondrial abnormalities. Toxicol. Sci. (2007) 97
205213.
137 Ulrich R.G., Bacon J.A., Branstetter D.G. et al. Induction of a
hepatic toxic syndrome in the Dutch-belted rabbit by a
quinoxalinone anxiolytic. Toxicology (1995) 98 187198.
138 Dixit R., Boelsterli U.A. Healthy animals and animal models of
human disease(s) in safety assessment of human pharma-
ceuticals, including therapeutic antibodies. Drug Discov.
Today (2007) 12 336342.
139 McCarthy T.C., Pollack P.T., Hanniman E.A., Sinal C.J.
Disruption of hepatic lipid homeostasis in mice after amioda-
rone treatment is associated with peroxisome proliferator-
activated receptor-a target gene activation. J. Pharmacol. Exp.
Ther. (2004) 311 864873.
140 Williams M.D., Van Remmen H., Conrad C.C., Huang T.T.,
Epstein C.J., Richardson A. Increased oxidative damage is
correlated to altered mitochondrial function in heterozygous
manganese superoxide dismutase knockout mice. J. Biol.
Chem. (1998) 273 2851028515.
141 Mansouri A., Muller F.L., Liu Y. Alterations in mitochondrial
function, hydrogen peroxide release and oxidative damage in
mouse hind-limb skeletal muscle during aging. Mech. Ageing
Dev. (2006) 127 298306.
142 Ong M.M., Wang A.S., Leow K.Y., Khoo Y.M., Boelsterli U.A.
Nimesulide-induced hepatic mitochondrial injury in hetero-
zygous Sod2
+/)
mice. Free Radic. Biol. Med. (2006) 40 420
429.
143 Talwani R.J., Hamm D.A., Tian L. et al. Medium-chain acyl-
CoA dehydrogenase deciency in gene-targeted mice. PLoS
Genet. (2005) 1 e23.
144 Tomita K., Azuma T., Kitamura N. et al. Leptin deciency
enhances sensitivity of rats to alcoholic steatohepatitis
through suppression of metallothionein. Am. J. Physiol.
Gastrointest. Liver Physiol. (2004) 287 G1078G1085.
Drug-induced mitochondrial dysfunction in liver: mechanisms and detection 353
2008 The Authors Journal compilation 2008 Blackwell Publishing Ltd. Fundamental & Clinical Pharmacology 22 (2008) 335353

You might also like