Вы находитесь на странице: 1из 9

Biochimica et Biophysica Acta 1777 (2008) 763–771

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / b b a b i o

Review

Cellular energetic metabolism in sepsis: The need for a systems approach


Jane E. Carré ⁎, Mervyn Singer
Wolfson Institute for Biomedical Research, University College London, Gower Street, London WC1E 6BT, UK

A R T I C L E I N F O A B S T R A C T

Article history: Sepsis is a complex pathophysiological disorder arising from a systemic inflammatory response to infection.
Received 4 February 2008 Patients are clinically classified according to the presence of signs of inflammation alone, multiple organ
Received in revised form 4 April 2008 failure (MOF), or organ failure plus hypotension (septic shock). The organ damage that occurs in MOF is not a
Accepted 5 April 2008
direct effect of the pathogen itself, but rather of the dysregulated inflammatory response of the patient.
Available online 23 April 2008
Although mechanisms underlying MOF are not completely understood, a disruption in cellular energetic
Keywords:
metabolism is increasingly implicated. In this review, we describe how various factors affecting cellular ATP
Mitochondria supply and demand appear to be altered in sepsis, and how these vary through the timecourse. We will
Sepsis emphasise the need for an integrated systems approach to determine the relative importance of these factors
Inflammation in both the failure and recovery of different organs. A modular framework is proposed that can be used to
Multiple organ failure assess the control hierarchy of cellular energetics in this complex pathophysiological condition.
Mitochondrial dysfunction © 2008 Elsevier B.V. All rights reserved.
Top–down metabolic control
Cellular energetic metabolism

1. Introduction dying, compared to 90% of those with four or more failures [6,7].
Patients who survive their septic episode often have long-term
Sepsis describes a combination of clinical manifestations of sys- morbidities, including weakness, fatigue, neuropathies and psycholo-
temic inflammation specifically related to an infectious insult from gical problems [2]. To what extent these problems are due to the
agents such as bacteria, pathogenic fungi, yeasts and viruses. Other disease process itself and how much to sequelae and complications
non-infectious insults such as trauma, burn injury or haemorrhage can related to the treatment and support given in intensive care is cur-
cause a similar response, suggesting a common final pathway. Clinical rently a major topic of investigation.
presentations include pyrexia, variable changes in cardiac output (the
product of heart rate and ejected ventricular stroke volume), 1.1. Pathophysiology of sepsis
hypotension that is non-responsive to vasopressor agents, immuno-
suppression, and evidence of failure of one or more organ systems Several comprehensive reviews of systemic immune [8,9], endo-
including haemodynamic, respiratory, neurological, renal and hepatic crine [10,11] and metabolic [12–14] responses in sepsis are available,
systems. The severity of illness depends on the extent to which these and only a general overview will be given here (Fig. 1). The
disturbances occur and has led to the classification of sepsis into sepsis inflammatory response to infection is initiated by the binding of
syndrome, severe sepsis and septic shock ([1], Table 1). microbial products such as the bacterial cell wall components
The course of sepsis is variable and often unpredictable, and lipopolysaccharide (Gram-negative) and peptidoglycan (Gram-posi-
depends on factors such as age and pre-existing morbidities. Genetic tive) to Toll-like and other recognition receptors on immune and
polymorphisms also play a part in determining susceptibility to sepsis endothelial cells. Subsequent activation of nuclear transcription factor
and the severity of the inflammatory response. Mortality rates are κB (NF-κB) by its translocation to the nucleus leads to expression of
decreasing due largely to improved supportive care, however, the significant circulating levels of pro-inflammatory cytokines including
incidence of sepsis is rising [2,3]. Between 30–50% of critically ill tumor necrosis factor-α (TNF-α) and interleukin 1-β (IL-1β). Together
patients in intensive care units (ICUs) develop sepsis [2,4], with an with activation of complement pathways, these proinflammatory
average early mortality risk rising from 4% after 24 h to approximately cytokines stimulate a pro-coagulatory state. Cytokine stimulation
30% by day 28 [5]. Mortality increases with the number of failed organs, of circulating and resident immune cells, endothelial and some
with approximately one-third of patients with single organ failure epithelial cells results in increased production of reactive oxygen
intermediates (ROS) such as superoxide, and nitric oxide (NO).
⁎ Corresponding author. Excessive NO production is considered a major cause of the vaso-
E-mail address: j.carre@ucl.ac.uk (J.E. Carré). relaxation and decreased responsiveness to vasopressor agents

0005-2728/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.bbabio.2008.04.024
764 J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771

Table 1 “stress” hormones including cortisol, catecholamines, vasopressin, and


Definitions of systemic inflammation and sepsis both insulin and glucagon [10,11]. Together, these hormones act to
Condition Definition and clinical description maintain oxygen delivery to the tissues (q.v.) and also result in
Systemic inflammatory The systemic inflammatory response to various clinical mobilisation of carbohydrate, fat and protein stores in muscle and liver
response syndrome (SIRS) insults (infection, trauma, haemorhage). to fuel the synthesis of acute phase proteins. Resting energy ex-
Requires at least two of: penditure (REE), as measured by calorimetry, is increased during this
• core temperature b 36 °C or N 38 °C
period, although the more severely affected patients do not increase
• heart rate N 90 beats min− 1
• respiratory rate 20 breaths min− 1or PaCO2 4.3 kPa REE to the same extent [17]. Indeed, septic shock patients have REE
• white blood cell count b4000 mm− 3 or N12 000 mm− 3 or values similar to healthy people, often despite marked pyrexia. The
immature neutrophil content N 10% acute phase endocrine response is predictive of outcome with a
Sepsis The systemic response to infection. suboptimal or impaired response portending a poor prognosis [18].
• SIRS criteria PLUS
• proved or suspected infection
During prolonged sepsis, circulating hormone levels return to
Severe sepsis Severe sepsis associated with: normal, or even sub-normal levels. The “low T3” syndrome is often
• organ dysfunction manifest with decreases in thyroxine (T4) and the more active tri-
• hypoperfusion or hypotension iodothyronine (T3), and an increase in the inactive reverse T3 (rT3).
Septic shock Sepsis associated with:
REE shows a significant rebound during the recovery phase [17].
• hypotension that is unresponsive to administered
vasopressors (despite adequate fluid resuscitation)
• perfusion abnormalities 1.2. Current treatments for sepsis
Summarised from [1].
While many immunomodulatory therapies have been trialled,
none has equivocally demonstrated an overall reduction in mortality.
Even the use of activated protein C, the only immunomodulatory agent
(vascular hyporeactivity) underlying septic shock (reviewed in [15, 16]. with a current licensed indication for sepsis, is now being formally
Expression of adhesion molecules results in increased infiltration of readdressed. Most therapy is supportive rather than curative, propping
immune cells into tissues both local to, and distant from, the cause of up the failing organs until they start functioning. These include anti-
infection. In line with the pro-inflammatory response, a compensatory biotics to remove the initiating trigger, fluid resuscitation to maintain
anti-inflammatory response is also initiated to quench the flames of organ perfusion, vasopressor administration to maintain an adequate
inflammation. After an indeterminate and variable period of hours to blood pressure, and mechanical support of failing lungs (mechanical
days, this may even shift the balance of the inflammatory milieu to an ventilation) and kidneys (dialysis-like techniques). Despite these
immunosuppressed state characterised by high levels of anti-inflam- technical advances, sepsis still remains a major killer worldwide.
matory cytokines such as IL-10, T-cell anergy, accelerated apoptosis of
circulating B-lymphocytes, delayed apoptosis of neutrophils, and 1.3. Energetic failure in sepsis?
decreased phagocytic and bactericidal activity [9].
Besides the alterations in cytokine profiles, the acute phase The concept of cellular energetic dysfunction in septic organ failure
response to infection is accompanied by a massive release of various can be traced back some 40 years, with observations of swollen and

Fig. 1. Systemic interactions in the pathophysiology of sepsis. NF-κB, nuclear factor κB. Reproduced from [8] with permission.
J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771 765

distorted mitochondria in livers of mice injected with endotoxin [19]. concept of mitochondrial dysfunction in sepsis impacts upon cellular
Mitochondrial respiration, respiratory complex enzyme activities and energetics.
nucleotide levels have been studied in tissue homogenates, cells or
mitochondria isolated from a variety of animal or cell models to which 2. Control of cellular energetics: ATP supply and demand
endotoxin or bacterial insults had been administered. Highly variable
changes in mitochondrial function are reported in models lasting up to 2.1. Top–down metabolic control analysis and the “modular kinetic” approach
16 h, but a much more consistent finding of mitochondrial dysfunction
and ultrastructural damage is seen thereafter (reviewed in detail in Our understanding of the control of respiration (see [36]) has
[20]). Several factors are likely to be responsible for this variability, benefited significantly from the development of metabolic control
including the type and degree of insult, the animal, organ or cell analysis [37,38] and the subsequent extension of top–down or
species selected, and the adequacy or otherwise of resuscitation. modular control analysis [39–41]. These approaches have led to the
Differences in tissue ATP content have been reported in muscle appreciation that control of cellular respiration is shared between
biopsies from patients [21,22] and in different tissues from animal processes that provide ATP and those that consume it and that the
models [23–25]. We reported depressed ATP levels in muscle biopsies strength of control that one process has over respiration varies under
taken from critically ill patients who went on to die, while eventual different conditions [42].
survivors demonstrated maintained/elevated ATP [21]. In combina- Compared to the bottom-up approach, which establishes the
tion with several reports of elevated, rather than depressed, tissue control exerted by individual enzymes over flux through a pathway
oxygen tensions (tPO2), these findings of mitochondrial abnormalities [38,42], the top–down approach divides systems into blocks of
and changes in ATP led to the proposal that septic organ failure reactions, consisting of either individual enzymes or, usefully, groups
represents a state of “tissue dysoxia” or “cytopathic hypoxia”, i.e. a of enzyme and intermediate interactions (i.e. pathways). In a top–
cellular inability to use oxygen rather than a lack of its availability [26]. down (or “modular-kinetic”) analysis, the responses of the steady-
Studies in septic patients widely report depletion of glutathione state flux of consumer and producer pathways to changes in con-
and ascorbate in both tissues and blood [27,28]. We also found a centration of their linking intermediate(s) are measured. From these
correlation between the severity of sepsis and muscle glutathione responses (i.e. elasticities), the influence of a change in activity of these
levels in critically ill ICU patients [21]. Apart from the afore mentioned producers or consumers on both overall steady state flux (i.e. flux
link between ATP levels and subsequent death, we also reported other control) and intermediate concentration (i.e. concentration control)
prognostic features including depleted muscle levels of reduced can be calculated. A pathway having large elasticity towards an
glutathione, impaired complex I activity, elevated nitrate/nitrite intermediate has, in general, low flux control, and vice versa. The
(indicator of RNS) [21], increased EPR-detected radical species, and overall control structure of a process can thus be described quan-
decreased complex I iron–sulphur centres [29]. Plasma levels of lipid titatively. This will predict whether and, if so, how a system changes in
peroxide products were elevated in patients with multi-organ failure response to acute changes in intermediate or enzyme activity (for
(MOF), particularly in those with renal failure [30]. Plasma from septic example, resulting from hypoxia or allosteric modulations such as
patients induced a severity-linked oxidative stress when applied to phosphorylation). This modular-kinetic approach may be exploited
cells in vitro [31]]. Functional implications of such oxidative stress has further to assess (1) where and to what extent a system is affected by
been inferred by the presence of mitochondrial damage and depletion administration of effector molecules (e.g. hormones, kinases, NO) and
in patient samples of both leg muscle [21,22,28]and diaphragm [22]. (2) how the control between two or more systems (e.g. cell types,
pathophysiological states) differs.
1.4. Energetic failure or adaptive metabolic control? The need for a systems
approach 2.1.1. A modular kinetic approach to cellular energetics in sepsis
In Fig. 2, we outline a simple modular framework that can be used
Together, such data as described above have led to the formulation to determine how sepsis can affect the control of cellular energetic
of the hypothesis that septic organ failure results from a depression metabolism. The system is centred around ATP as an intermediate
in mitochondrial function, with key effectors being cytokine-induced with a block each for supply and demand. By modulating the activity
oxidant and nitrosative species, and changes in hormone levels of one block and determining the effect on the activity of the other, a
[20,26,32]. To some, this represents a damage-induced energetic description of the responses of ATP-producing and ATP-consuming
failure of mitochondria. Based on the relatively rapid recovery of pathway fluxes to a change in ATP is obtained [39–41]. From this
organs and the lack of evidence of significant cell death in failed organs kinetic description, the elasticities of supply and demand blocks
[33], an alternative hypothesis proposes that the condition manifest as towards ATP can be calculated. Subsequently, these elasticities can be
“organ failure” actually represents an adaptive and functional meta- used to calculate the control of each block over the overall steady-state
bolic suppression, similar to aestivation or hibernation, that enables flux of ATP supply and demand, and also the control exerted by each
long-term survival if the victim survives the initial insult [34]. block over the concentration of ATP. This initial analysis reveals
It is inevitably problematic to obtain tissue samples from patients broadly whether or not supply and/or demand pathways have been
with MOF to perform mechanistic studies, especially from deep “vital” affected by the pathophysiological insult. Once this is ascertained, the
organs such as liver and kidney. It is thus not surprising that human relevant block can be further subdivided, and a more detailed modular
literature is limited to snapshots of cellular events. However, these kinetic analysis (or, if more appropriate, a bottom–up analysis) can
steady-state measurements reveal simply that a process has been then performed to identify more precisely the site(s) of modification.
affected by an insult, and reveal little about whether affecting that The value of understanding how and where changes in the dis-
process at that time will have any influence on cellular behaviour. For tribution of control of a pathophysiological system have occurred
example, nitrosative damage to respiratory complex I is likely to be becomes apparent when deciding where to target therapeutic efforts.
detrimental to cellular energetics only at times where supply of ATP is For example, if at a particular time-point, control of ATP supply
unable to meet cellular ATP demand (see [35], discussed below). and demand is found to have shifted towards ATP-producing pathways
Therefore, determining the cause and effect of energetic disturbances (e.g. oxidative phosphorylation), therapies aimed towards increasing
requires a systems approach, whereby the impact of sepsis on ATP production (e.g. substrate availability, mitochondrial functional
energetic supply and demand is determined. We anticipate that by capacity) are more likely to be successful than during periods where
employing the top–down metabolic control approach described control lies with ATP-consuming pathways. In this respect, further
below, we will be able to determine to what extent the long-held analysis of the effects of external modulators including hormones
766 J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771

drial functional capacity. However, consideration should also be given


to the processes directly affecting Δψ, Ca2 and ROS production, as
these factors play significant roles both in energy metabolism and in
mediating the mitochondrial permeability transition and cell death.

3.1. ATP production and consumption

3.1.1. Control of ATP supply and demand


The control distribution for oxidative phosphorylation has been
found to be similar for isolated liver mitochondria [52] and resting
hepatocytes [53], with approximately 50% of control over respiratory
rate lying with phosphorylating reactions (equivalent to ATP turn-
over), 30% with the electron transfer chain and 20% with proton
cycling (leak). Most (~ 80%) of the control over phosphorylation lay
with the phosphorylating system itself, indicating that cellular
energetics in hepatocytes are largely demand-driven. On the other
hand, control over phosphorylation in brain mitochondria was found
Fig. 2. Modular separation of energetic metabolism around ATP for analysis of cellular to lie largely with substrate oxidation reactions (~ 70%) [54].
energetics in sepsis. Initial analysis of Supply/Demand flux and [ATP] at different The sensitivity of supply and demand pathways to small changes
modulations of flux reveal whether Supply block and/or Demand block have been in [ATP] (or ATP/ADP, energy charge or the phosphorylation potential
altered by the septic insult. Subsequently, the relevant blocks can be sub-divided and
appropriate intermediates (e.g. O2, substrate, Δψ, Ca2+) and fluxes (arrows: e.g.,
ΔGp) will reflect the elasticities of all the components in a given
glycolysis, substrate oxidation, phosphorylation, proton leak, individual demand block. For the ATP-supply pathways, this includes the sensitivities of
pathways) can be analysed to pin-point where control distribution of the system has both oxidative phosporylation and glycolysis. Several studies have
been affected. A comparison of control distribution for different time points in sepsis can shown an increase in the rate of lactate production upon stimulation
be performed for acute, established and recovery phases, in different tissues, or at
of the Na+/K+-ATPase, or a decrease in lactate production by inhibition
different severities of insult. The site of action of various effectors (hormones, NO/ROS,
mediators of mitochondrial biogenesis, temperature) can be determined. See [38–40] for of this transporter [55], combined with a relative insensitivity of
details of the modular kinetic approach to top–down metabolic control analysis. oxygen consumption (e.g. [56]). This behaviour suggests that these
changes in flux are due to a higher elasticity of glycolysis, compared to
that of oxidative phosphorylation, towards ATP.
[43,44] or even drugs [45] used in patient management (e.g. catecho- The ATP-demand block comprises those processes involved in
lamines) can be determined for the initially characterised system. maintenance of ion gradients (Na+/K+-ATPase, Ca2+-ATPase), and
anabolic functions including macromolecule synthesis (DNA, RNA,
2.2. Cell models of sepsis for systems approaches to energetic metabolism protein) and tissue-specific ATP-dependent processes (gluconeogen-
esis, ureagenesis). Atkinson [57] first proposed the concept of hierarchy
The modular kinetic approach outlined would be suitable to study among different ATP-demanding processes where less essential
ex vivo cells that have been freshly isolated from established animal anabolic functions are more sensitive to changes in energetic status
models of sepsis. In this respect, polymicrobial models of peritoneal compared to essential activities such as ion homeostasis. This concept
sepsis lasting several days have been developed for rodent and large has been demonstrated in several experimental systems including
animal studies. Methods used for the induction of sepsis in these immune-activated thymocytes [58] and responses of hepatocytes to
models include caecal ligation and puncture (reviewed in [46]), hypoxia [59,60]. Whether the difference in elasticities of reaction blocks
injection of faecal slurry (e.g. [23] or implantation of bacterially- to ATP is a direct allosteric effect of nucleotides on enzymes controlling
infected fibrin clots (e.g. [47]). The provision of adequate fluid these processes or an indirect effect via allosteric regulation by, for
resuscitation is considered to result in more clinically-relevant long- example, AMP-activated protein kinase (AMPK) [61] is, at present,
term models. Cells can be isolated from different organs at different unclear. In some systems (e.g. muscle), changes in [ATP] are buffered by
stages of disease progression and recovery, and analysed ex vivo. activity of the creatine kinase system. This system is also involved in
Alternatively, in vitro models can be used where cultured cells are effectively shuttling ATP between compartments and within the cytosol
subjected to a septic insult of varying dose or duration, followed by a [62].
washout period to follow recovery. In this respect, incubation with
either LPS alone or in combination with cytokines has been used to 3.1.2. ATP and supply-demand interactions in sepsis
investigate cellular energetic aspects of sepsis (e.g. [48,49]). Co-culture As discussed earlier, a low muscle ATP content at study entry was
and Transwell culture of hepatocytes and Kupffer cells has been used to associated with mortality in a critically ill group of patients [21]. In a
investigate the mechanism of sepsis-induced down-regulation of follow-up study [unpublished data], there was also a significant
peptides involved in vasodilation [50]. A further in vitro approach depression of phosphocreatine (PCr) content and, in survivors, a lower
investigated the effect of plasma taken from control patients or those PCr/ATP ratio. Similar findings were reported in a separate study of leg
with sepsis on the oxygen consumption rates of peripheral blood muscle, although neither PCr nor ATP content was affected in inter-
mononuclear cells (PBMCs) isolated from septic or control patients [51]. costal (respiratory) muscle [22]. Measurement of leg muscle lactate by
The choice of system for a modular kinetic analysis will depend on microdialysis in a group of patients in early septic shock revealed a high
a balance between (1) the clinical relevance, but inherently variable lactate content that decreased over the following 24 h. Parallel
response, of animal models, and (2) the more easily-controlled, but microdialysis with an infusion of ouabain indicated that muscle lactate
potentially less-directly translatable, in vitro cell model. production rate was sensitive to this Na+K+-ATPase inhibitor [63].
Although a comparison with control patients is unavailable, these data
3. Cellular energetics in sepsis suggest an enhanced glycolytic rate in muscle in early sepsis linked to
an increase in an ATP demanding process.
The discussion below concentrates on aspects of cellular energetic Together, these data suggest a stimulation of ATP turnover in leg
metabolism that, to date, have received most attention in sepsis: ATP muscle during the early phase of sepsis. In an investigation of septic
production and consumption, oxygen availability, and also mitochon- muscle energetics over time, 31P NMR analysis of PCr, Pi and ATP in
J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771 767

septic rats indicated that mitochondrial activity was enhanced early proportion of control of the phosphorylating system over respiration,
in sepsis, but diminished later, although two separate models were could significantly affect cellular energetics [75].
used in this comparison [25]. Time-dependent studies in patients
are lacking, however our preliminary (unpublished) data suggest that 3.1.5. Substrate-dependent effects in sepsis
PCr/ATP in surviving patients increases over time and, after two weeks In sepsis, changes in substrate-dependent effects on ATP
of illness, is equal to, or indeed greater than, control values. production rate and efficiency may be reflected at a number of the
levels discussed above. Firstly, circulating cytokines and catechola-
3.1.3. Efficiency of ATP production in sepsis: proton leak mines cause a shift in metabolism towards the stress state: acti-
One aspect that remains to be resolved in sepsis is the extent to vation of glycogenolysis and hepatic gluconeogenesis, together with
which REE represents metabolism that is uncoupled from ATP insulin resistance lead to increased circulating levels of glucose;
production. Under normal physiological conditions, it is estimated hepatic lipolysis increases plasma free fatty acids, while greatly
that approximately 20% of the standard metabolic rate of oxygen increased muscle protein catabolism and lactate release also occur
consumption is not coupled to ATP production, but is used to drive the (see [12,14]). While the levels of these circulating substrates are
proton leak across the mitochondrial membrane [64]. The extent to increased, the extent to which the substrates are oxidised or taken
which this changes in sepsis is uncertain, but it has been argued that up by cells and mitochondria of different tissues may be variable. For
metabolic inefficiency is a major contributor to pyrexia (see [64]). One example, insulin-independent glucose uptake may be inhibited in
key hormone that affects metabolic efficiency is thyroid hormone, sepsis in some cells by down-regulation of the glucose transporter,
which both increases the rate of respiration and the proton leak rate GLUT-4, whilst expression of insulin-dependent glucose transpor-
[44,65]. The modulations in thyroid status at different stages of critical ters is increased in the presence of cytokines. The disruptions in
illness may have a major impact on the efficiency of ATP production. circulating levels of Ca2+-modulating hormones discussed above
Thyroid hormone is significantly elevated during the acute phase of could also have significant effects on the activation state of
sepsis, whilst sub-normal levels persist during established illness mitochondrial dehydrogenases, thus affecting the availability of
(the low T3 syndrome described earlier). respiratory substrates. Other mechanisms by which substrate
Proton leak across the inner mitochondrial membrane to the availability may be affected in sepsis are by depletion of NAD+ via
matrix can be modulated by the regulated engagement of protein- peroxynitrite-induced activation of PARP (poly-(ADP ribose) poly-
dependent pathways, such as activation of uncoupling proteins (UCPs) merase), a DNA-repair enzyme [76].
(see [66,67]), or by increased basal leak via the adenine nucleotide
translocase [68]. Although the physiological role and function of 3.2. Oxygen availability
the more novel members of the UCP family remains to be firmly
established, UCP2 activity of immune cells seems to be involved 3.2.1. Cell responses to tissue hypoxia
inflammation [69,70]. The extent to which tissue hypoxia causes ATP depletion and
In addition to proton leak, any changes in the stoichiometry of associated cell damage in a particular tissue will depend on (1) the
respiratory proton pumps (redox slip) could contribute to decreased metabolic response of the system to short-term changes in O2 tension/
efficiency of ATP production (see [71,72]), although it is uncertain as to concentration [O2] and (2) the sensitivity of the system to any changes
what extent this process occurs (patho)physiologically. in [ATP], both in terms of short-term changes in control of metabolic
flux and longer-term adaptive responses to these changes (see [77]).
3.1.4. Substrate-dependent effects on ATP production efficiency and rate
Depending on the individual tissue and on nutritional status, the 3.2.2. Short-term effects of hypoxia
main respiratory substrates used by the body are glucose and fatty Whereas some cells (e.g. kidney) are highly sensitive to changes in
acids, although significant amounts of lactate are also oxidised by [O2] and reduce their consumption as [O2] decreases [78], the oxygen
heart, liver and brain. Each substrate has a different efficiency of ATP uptake rate of many other cells is relatively insensitive until O2 levels
production; for example, oxidation of palmitate yields approximately 3 as low as 3–8 µM are reached [79–81]. The apparent KmO2 in
times the amount of ATP per mole as does complete oxidation of hepatocytes is reported to be approximately 2 µM [82]. Changes in
glucose, resulting in maximum amounts of ATP synthesised per oxygen intermediate concentrations (e.g. NADH/NAD+, Δψ, reduced cyto-
atom consumed (P/O) that differ by some 14% (2.1 vs 2.4, respectively, chrome c, phosphorylated metabolites) can occur without a change in
assuming an H+/ATP ratio of 13/3, see [73]). The effective P/O of respiratory rate or [ATP] ([82], discussed in [36]). Below these oxygen
substrate oxidation varies with the extent of proton leak, increasing tensions, there are changes in both respiratory rate and levels of
towards state 3 as the proton leak rate declines [74]. The oxidation of a intermediates. Gluconeogenesis and urea synthesis – pathways in
particular substrate will affect NADH/NAD+ redox state, membrane hepatocytes that both consume and produce ATP – exhibit a similar
potential (Δψ), and respiration (including proton leak) rate to different sensitivity to changes in O2 (apparent KmO2 5 µM; [83]). As respiratory
extents. As discussed in [74], hormones that increase Ca2+, stimulate oxygen consumption decreases in low O2, glycolysis increases in rate
respiration and/or increase proton leak may also influence the effective (the Pasteur effect).
P/O and the rate of ATP production. The relative contribution of each Factors that influence the availability of O2 include (1) the effect of
the factors above to the in vivo rate of ATP production is unclear. vasoactive hormones on oxygen delivery by heart, lungs and
However, according to the relative control of each of these parameters circulation to the tissues (potentially in addition to direct endocrine
over respiration under the prevailing conditions, the effective P/O and effects on the energetics of the tissue of interest) and (2) nitric oxide,
the rate of ATP production may indeed vary (see [74]). NO, the effects of which could include (a) cGMP-dependent vaso-
As control of respiration shifts towards substrate oxidation with relaxation, affecting oxygen delivery and (b) competitive inhibition of
increasing demand for ATP, substrates that increase the rate of cytochrome c oxidase that, besides inhibiting the respiratory rate, may
respiration when demand is high are likely to increase ATP production effectively increase the local [O2] within the cell. Such an effect of NO
rate. Equally, under similar conditions of increasing demand for ATP, may be relevant to the availability of O2 to enable the seemingly
substrate deprivation could have a detrimental effect on cellular paradoxical increase in mitochondrial ROS production during hypoxia.
energetics, particularly under conditions of low oxygen. However, the quantities of NO produced in inflammation are likely to
Other substrate-dependent effects on cellular energetics include, be detrimental to the cell, especially in combination with hypoxia. For
for example, direct inhibition of the adenine nucleotide translocase example, hypoxia has been shown to sensitize mitochondria to
(ANT) by long-chain acyl-CoA esters which, given the significant inhibition by NO in isolated inflamed aorta [84], neuronal [85,86]
768 J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771

and activated macrophage cell lines [49], resulting in increased 3.3. Functional capacity
necrotic cell death. Inhibition by NO was reversible if NO was removed
from the medium, but irreversible following prolonged exposure to The concept of “reserve capacity” in mitochondria is well
NO, probably reflecting a reversible competitive inhibition by NO at recognised, in that the respiratory rate can be stimulated above
complex IV followed by more stable inhibition of complex I by state 3 in isolated mitochondria, or over the endogenous rate in cells
peroxynitrite [49,84]. upon dissipation of Δψ by the addition of a chemical uncoupler. Such
a reserve might be important in some tissues when the demand for
3.2.3. Longer-term responses to hypoxia cellular ATP acutely rises and the proportion of control over
Cellular responses to hypoxia depend on the tissue itself, and the respiration exerted by the respiratory chain increases ([52], dis-
duration and extent of hypoxia. For example, hepatocytes exhibit cussed in [35]). In this respect, under conditions of elevated ATP
progressive decreases in [ATP] and oxygen consumption rate as demand, systems which are unable to maintain functional mito-
hypoxia continued over several hours [59,87]. These decreases were chondria might be expected to compromise their function and even
fully reversible upon reoxygenation, normalising more rapidly than viability. Depending on the system, short-term adjustment in the rate
for cells exposed to a shorter period of hypoxia [59]. Prolonged of “non-essential” (e.g. macromolecule synthesis) or facultative
moderate hypoxia was associated with maintained Na+/K+- ATPase (metabolic) ATP demand pathways, may initially occur to allow the
activity while other, non-essential ATPase activities were suppressed cell to prioritise essential processes such as maintenance of ion
[59]. Such a “hibernation” response to hypoxia is also seen in gradients. In the longer term, failure to maintain mitochondrial
ischaemic myocardium. The mechanism of such adaptive responses functional capacity may result in ATP depletion and this will trigger
is unclear [77,88]. Possible candidates for molecular regulation include cell death pathways. These two responses — a decrease in metabolic
hypoxia-inducible transcription factor-1 (HIF-1a) [89] or AMP- function and increased cell death— could conceivably be manifested
activated protein kinase [61,90]. clinically as organ failure.
Mitochondrial functional capacity in the cell is determined by
3.2.4. Oxygen availability in sepsis several factors including: (1) the extent to which existing mitochondrial
A major point of contention in the sepsis field is the extent to enzymes retain functionality, including (a) the extent of oxidative/
which tissue oxygen delivery compromises the mitochondrial func- nitrosative damage and (b) the ability of enzymes to respond to
tion. Early studies interpreted a high blood lactate level (usually allosteric activators; (2) the rate of synthesis and assembly of proteins
associated with a metabolic acidosis) as indicative of tissue hypoxia, in the mitochondria; and (3) the rate of proteolytic degradation. The
representing anaerobic glycolytic metabolism [91]. As metabolic extent to which changes in functional capacity are important in
acidosis was positively correlated with mortality in sepsis [92], this defining overall energetic metabolism in sepsis will depend on the
finding was perceived as deleterious. Despite early suggestions that hierarchy of control over the system in the septic condition.
hyperlactataemia could occur in the absence of oxygen delivery
defects, the tissue hypoxia theory prevailed through the 1990s and 3.3.1. Impaired mitochondrial function in sepsis
prompted two large clinical trials aiming to optimise tissue perfusion Reports of altered respiratory capacity in isolated mitochondria in
of septic patients in established multi-organ failure. Alas, these sepsis date back nearly 40 years, with conflicting findings [reviewed in
showed either no benefit [93] or harm [94]. [20] depending on the duration and severity of the model and the
Low cardiac output states such as cardiogenic, hypoxic, or tissue studied. The majority of reports using isolated mitochondria or
haemorrhagic shock are characterised by decreases in tissue oxygen permeabilised cells and muscle fibres point towards an early increase
delivery and consequent tissue hypoxia [95]. In contrast, sepsis is in capacity (4–18 h) followed by a later depression (24–48 h). A
generally manifest as a high cardiac output state — despite hypoten- decrease in complex I-linked respiration in isolated muscle fibres was
sion, global oxygen supply is often reasonably well maintained. There associated with the severity of sepsis in a long-term rat model, while
is, however, lactic acidaemia and evidence of local impairment of succinate-dependent oxidation rates were similar between control
microvascular blood flow, predominantly through vasoconstriction. It and septic groups [102]. Differences in respiration between intact
has thus been argued that the resulting localised tissue hypoxia will hepatocytes (decreased endogenous rate) and mitochondria
lead to organ failure [96]. However, several animal and human studies (increased state 3 rate) isolated from livers of endotoxic rats have
suggest that, despite indications of microcirculatory abnormality, been explained by the existence of variably affected mitochondrial
peripheral organs do not become hypoxic during sepsis after initial populations [24]. Electron micrographs indicate a population of
resuscitation has been performed. Elevated tissue oxygen tensions swollen mitochondria amidst more normal looking, electron-dense
(PO2) have been reported for bladder [95,97,98] and gut [99] in mitochondria. Potentially, these swollen organelles may be more
volume-resuscitated septic models, as well as in the muscle of septic susceptible to rupture and removal during the mitochondrial isolation
patients [100,101]. Recently, using a short-term (3 h), fluid-resusci- procedure.
tated endotoxaemic rat model, there was an elevation in bladder PO2
and relatively maintained muscle and renal PO2 despite a significant 3.3.2. Mitochondrial biogenesis
reduction in cardiac output [95]. On the other hand, liver PO2 fell both As described earlier, limited patient studies report decreases in
early and profoundly, paralleling changes seen with severe haemor- mitochondrial enzyme activities [21,22] that are associated with
rhage and hypoxaemia. Similar findings have been seen after 6 h in a oxidative/nitrosative damage [27]. Similar findings have been made in
faecal peritonitis septic rat model yet the tissue PO2 values normalized animal models [23–25,103], and may be related to a functional
by 24 h, despite severe clinical and biochemical manifestations of inhibition and/or damage/loss of active respiratory complex protein,
organ dysfunction becoming apparent (Dyson and Singer, unpub- particularly complex I. The recovery of mitochondrial function
lished data). Thus, the organ tissue PO2 response to sepsis – which requires the process of organelle repair/regeneration. This process is
reflects the local O2 supply-demand balance – varies, even in the face called mitochondrial biogenesis.
of decreased oxygen delivery and potential micorovascular abnorm- Mitochondrial biogenesis describes the co-ordination of nuclear-
alities. A reconciling view is that both macro- and microcirculatory and mitochondrial gene expression, protein import and assembly into
abnormalities occur early in sepsis and, if not corrected sufficiently mitochondrial membranes and the changes in mitochondrial mass and
quickly, can amplify the inflammatory response with a possibly morphology. This process can be regulated, in theory, at each level,
greater impact on later mitochondrial dysfunction. However, this although the relative importance remains to be established. Both ROS
needs to be confirmed and the debate continues. and NO have been implicated as signalling molecules for regulation of
J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771 769

biogenesis pathways [104–106]. A key regulatory molecule in By establishing the control distribution of a system, the effect of
mitochondrial biogenesis is peroxisome proliferator-activated recep- modulations at different points in the system can be predicted. This
tor gamma (PPARγ) co-activator-1α (PGC-1α), which is highly aspect is fundamental to the success of therapeutic interventions, as
expressed, in oxidative tissues including heart, kidney, skeletal muscle targeting a part of the system that has little control at a particular time
and brain [107–109]. This nuclear transcriptional co-activator speci- will be unlikely to be of any benefit. We anticipate that the application
fically interacts with, and enhances, activity of transcription factors of a modular kinetic analysis [39–41] will enable the role of different
that act on promoters of genes involved in energy metabolism. Specific candidate effectors in the causality of organ dysfunction and recovery
targets of PGC-1α common to different tissues include the nuclear to be determined for different tissues and immune cells throughout
respiratory factors (NRF)-1 and NRF-2; these are co-activated to bind to the clinical course of the disease. Such an insight may prove useful
promoter regions of nuclear-encoded mitochondrial genes and to in determining the application of therapies aimed at stimulating
those of the gene encoding mitochondrially-targeted transcription energetic metabolism at one time point, whilst suppressing it at
factor A (Tfam, or mTFA). Another target of PGC-1α is the (o)estrogen- others.
related receptor-α (ERRα). The net effect of PGC-1α co-activation is the
upregulation of large groups of mitochondrial and nuclear-encoded Acknowledgements
genes involved in respiration and β-fatty-acid oxidation. In the liver,
PGC-1α activates gluconeogenesis and fasting-related genes. PGC-1α We would like to acknowledge the Medical Research Council and
is subject to complex regulation at the transcriptional and post- European Intensive Care Society for research funding.
translational level (reviewed in [110]).
References
3.3.3. Mitochondrial biogenesis in sepsis
Cytokines increase stability and activity of PGC-1α through MAP [1] R.C. Bone, R.A. Balk, F.B. Cerra, R.P. Dellinger, A.M. Fein, W.A. Knaus, R.M. Schein,
kinase-dependent phosphorylation in muscle cells [111]]. In the same W.J. Sibbald, Definitions for sepsis and organ failure and guidelines for the use of
innovative therapies in sepsis. The ACCP/SCCM Consensus Conference Commit-
study, 12 h after administration of LPS, transgenic mice over- tee, Am. College of Chest Physicians/Soci. Critical Care Med. Chest 101 (1992)
expressing PGC-1α linked to the muscle creatine kinase (MCK) 1644–1655.
promoter showed a near-doubling of whole body oxygen consump- [2] D.C. Angus, R.S. Wax, Epidemiology of sepsis: an update, Crit Care Med. 29 (2001)
109–116.
tion with both increased state 3 and state 4 respiration in isolated [3] G.S. Martin, D.M. Mannino, S. Eaton, M. Moss, The epidemiology of sepsis in the
muscle mitochondria. These increases in respiration were related to United States from 1979 through 2000, N. Engl. J. Med. 348 (2003) 1546–1554.
elevated expression of some, but not all, of the PGC-1α downstream [4] D.A. Harrison, C.A. Welch, J.M. Eddleston, The epidemiology of severe sepsis in
England, Wales and Northern Ireland, 1996 to 2004: secondary analysis of a high
targets measured. Levels of the PGC-1α protein itself increased in
quality clinical database, the ICNARC Case Mix Programme Database, Crit. Care
response to LPS, but it is unclear to what extent this reflects the (London, England) 10 (2006) R42.
endogenous protein or effects on the MCK promoter. [5] J.C. Marshall, D.J. Cook, N.V. Christou, G.R. Bernard, C.L. Sprung, W.J. Sibbald,
To investigate a potential role of mitochondrial biogenesis in recovery Multiple organ dysfunction score: a reliable descriptor of a complex clinical
outcome, Crit. Care Med. 23 (1995) 1638–1652.
from sepsis, a 3 day mouse model of bacterial sepsis was developed that [6] C. Brun-Buisson, F. Doyon, J. Carlet, P. Dellamonica, F. Gouin, A. Lepoutre, J.C.
exhibited a reduction in whole body oxygen consumption and REE, Mercier, G. Offenstadt, B. Régnier, Incidence, risk factors, and outcome of severe
together with increased mRNA levels of mitochondrial superoxide sepsis and septic shock in adults. A multicenter prospective study in intensive
care units. French ICU Group for Severe Sepsis, JAMA 274 (1995) 968–974.
dismutase and a depression in mitochondrial DNA copy number [7] T.M. Perl, L. Dvorak, T. Hwang, R.P. Wenzel, Long-term survival and function after
(mtDNA) [47]. Recovery was associated with increased expression of suspected gram-negative sepsis, JAMA 274 (1995) 338–345.
NRF-1, Tfam and PGC-1α transcripts and a progressive increase in [8] E. Abraham, M. Singer, Mechanisms of sepsis-induced organ dysfunction, Crit.
Care Med. 35 (2007) 2408–2416.
cytochrome and mtDNA, followed by recovery of oxygen consumption [9] R.S. Hotchkiss, I.E. Karl, The pathophysiology and treatment of sepsis, N. Engl. J.
and REE. This study suggests a link between recovery from sepsis and Med. 348 (2003) 138–150.
mitochondrial biogenesis. Interestingly, our preliminary (unpublished) [10] B. Ellger, Y. Debaveye, G. Van den Berghe, Endocrine interventions in the ICU, Eur.
J. Intern. Med. 16 (2005) 71–82.
data in muscle biopsies taken from critically ill patients indicate that
[11] E.S. Nylen, B. Muller, Endocrine changes in critical illness, J. Intensive Care Med.
transcript levels of PGC1α, Tfam and NRF-1 were significantly elevated in (2004).
patients on entry to the ICU, but only in those who went on to survive. A [12] H.R. Michie, Metabolism of sepsis and multiple organ failure, World J. Surgery 20
(1996) 460–464.
similar pattern was seen in ATP and phosphocreatine content of these
[13] K. Träger, D. DeBacker, P. Radermacher, Metabolic alterations in sepsis and
biopsies [Brealey 2003 and unpublished data]. With the inherent vasoactive drug-related metabolic effects, Curr. Opinion in Crit. Care 9 (2003)
difficulties in obtaining samples from “vital” organs, it is uncertain 271–278.
whether this situation is unique to muscle but together, these data [14] R.R. Wolfe, Sepsis as a modulator of adaptation to low and high carbohydrate and
low and high fat intakes, European J. Clin. Nutrition 53 (Suppl 1) (1999) 136–142.
suggest that an ability to maintain functional mitochondria early in critical [15] D.D. Rees, S. Cellek, R.M. Palmer, S. Moncada, Dexamethasone prevents the
illness, or an ability to rapidly regenerate, may be crucial to recovery. induction by endotoxin of a nitric oxide synthase and the associated effects on
vascular tone: an insight into endotoxin shock, Biochem. Biophys. Res. Commun.
173 (1990) 541–547.
4. Conclusions/future perspectives [16] J. Carré, M. Singer, S. Moncada, Nitric Oxide, in: J.-L. Vincent (Ed.), Mechanisms of
sepsis-induced organ dysfunction and recovery, vol. 44, Springer, Heidelberg,
The controversy of to what extent tissue hypoxia, mitochondrial 2006, pp. 77–95.
[17] G. Kreymann, S. Grosser, P. Buggisch, C. Gottschall, S. Matthaei, H. Greten, Oxygen
dysfunction and ATP depletion contribute to organ dysfunction in sepsis consumption and resting metabolic rate in sepsis, sepsis syndrome, and septic
has been ongoing for many years. From the discussion above, it is clear shock, Crit. Care Med. 21 (1993) 1012–1019.
that cellular energetic metabolism in the pathophysiological syndrome [18] P.M. Rothwell, P.G. Lawler, Prediction of outcome in intensive care patients using
endocrine parameters, Crit. Care Med. 23 (1995) 78–83.
of sepsis could be affected at multiple sites. However, understanding the
[19] E. Levy, R.J. Slusser, B.H. Ruebner, Hepatic changes produced by a single dose of
ATP supply-demand interaction is crucial for determining the relative endotoxin in the mouse. Electron microscopy, Am. J. Pathol. 52 (1968) 477–502.
importance of these factors in organ failure, yet few data are available [20] M. Singer, D. Brealey, Mitochondrial dysfunction in sepsis, Biochem. Soc. Symp.
66 (1999) 149–166.
that describe the effect of sepsis on the cellular demand for ATP. Limited
[21] D. Brealey, M. Brand, I. Hargreaves, S. Heales, J. Land, R. Smolenski, N.A. Davies, C.E.
information can be gained from measurement of steady state metabolite Cooper, M. Singer, Association between mitochondrial dysfunction and severity
content or metabolic activities alone. Instead, a systems approach is and outcome of septic shock, Lancet 360 (2002) 219–223.
required to determine the relative importance of each of the factors [22] K. Fredriksson, F. Hammarqvist, K. Strigård, K. Hultenby, O. Ljungqvist, J.
Wernerman, O. Rooyackers, Derangements in mitochondrial metabolism in
discussed above in the development and severity of MOF, and in the intercostal and leg muscle of critically ill patients with sepsis-induced multiple
subsequent recovery period. organ failure, Am. J. Physiol. Endocrinol. Metab. 291 (2006) 1044–1050.
770 J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771

[23] D. Brealey, S. Karyampudi, T.S. Jacques, M. Novelli, R. Stidwill, V. Taylor, R.T. [54] D.F. Rolfe, A.J. Hulbert, M.D. Brand, Characteristics of mitochondrial proton leak
Smolenski, M. Singer, Mitochondrial dysfunction in a long-term rodent model of and control of oxidative phosphorylation in the major oxygen-consuming tissues
sepsis and organ failure, Am. J. Physiol. Regul. Integr. Comp. Physiol. 286 (2004) of the rat, Biochim. Biophys. Acta 1188 (1994) 405–416.
491–497. [55] J.H. James, C.H. Fang, S.J. Schrantz, P.O. Hasselgren, R.J. Paul, J.E. Fischer, Linkage of
[24] S.P. Kantrow, D.E. Taylor, M.S. Carraway, C.A. Piantadosi, Oxidative metabolism aerobic glycolysis to sodium–potassium transport in rat skeletal muscle. Implica-
in rat hepatocytes and mitochondria during sepsis, Arch. Biochem. Biophys. tions for increased muscle lactate production in sepsis, J. Clin. Invest. 98 (1996)
345 (1997) 278–288. 2388–2397.
[25] Y. Mizobata, D. Prechek, J.D. Rounds, V. Robinson, D.W. Wilmore, D.O. Jacobs, The [56] J.D. Campbell, R.J. Paul, The nature of fuel provision for the Na+,K+-ATPase in
duration of infection modifies mitochondrial oxidative capacity in rat skeletal porcine vascular smooth muscle, J. Physiol. (Lond) 447 (1992) 67–82.
muscle, J. Surg. Res. 59 (1995) 165–173. [57] D. Atkinson, Cellular energy metabolism and its regulation, Academic Press,
[26] M.P. Fink, Bench-to-bedside review: cytopathic hypoxia, Crit. Care (London, New York, 1977.
England) 6 (2002) 491–499. [58] F. Buttgereit, M.D. Brand, A hierarchy of ATP-consuming processes in mammalian
[27] U.B. Fläring, O.E. Rooyackers, C. Hebert, T. Bratel, et al., Temporal changes in cells, Biochem. J. 312 (Pt 1) (1995) 163–167.
whole-blood and plasma glutathione in ICU patients with multiple organ failure, [59] R.M. Subramanian, N. Chandel, G.R. Budinger, P.T. Schumacker, Hypoxic confor-
Intensive Care Med. (2005). mance of metabolism in primary rat hepatocytes: a model of hepatic hibernation,
[28] U.B. Fläring, O.E. Rooyackers, J. Wernerman, F. Hammarqvist, Temporal Hepatology 45 (2007) 455–464.
changes in muscle glutathione in ICU patients, Intensive Care Med. 29 (2003) [60] W. Wieser, G. Krumschnabel, Hierarchies of ATP-consuming processes: direct
2193–2198. compared with indirect measurements, and comparative aspects, Biochem. J. 355
[29] D.A. Svistunenko, N. Davies, D. Brealey, M. Singer, C.E. Cooper, Mitochondrial (2001) 389–395.
dysfunction in patients with severe sepsis: an EPR interrogation of individual [61] D.G. Hardie, J.W. Scott, D.A. Pan, E.R. Hudson, Management of cellular energy by
respiratory chain components, Biochim. Biophys. Acta 1757 (2006) 262–272. the AMP-activated protein kinase system, FEBS Lett. 546 (2003) 113–120.
[30] G.H. Metnitz, M. Fischer, C. Bartens, H. Steltzer, T. Lang, W. Druml, Impact of acute [62] U. Schlattner, M. Tokarska-Schlattner, T. Wallimann, Mitochondrial creatine
renal failure on antioxidant status in multiple organ failure Acta anaesthesiolo- kinase in human health and disease, Biochim. Biophys. Acta 1762 (2006)
gica Scandinavica, 44, 2000, pp. 236–240. 164–180.
[31] O. Huet, R. Obata, C. Aubron, A. Spraul-Davit, J. Charpentier, C. Laplace, T. Nguyen- [63] B. Levy, S. Gibot, P. Franck, A. Cravoisy, P.E. Bollaert, Relation between muscle
Khoa, M. Conti, E. Vicaut, J.P. Mira, J. Duranteau, Plasma-induced endothelial Na+K+ ATPase activity and raised lactate concentrations in septic shock: a
oxidative stress is related to the severity of septic shock, Crit. Care Med. 35 (2007) prospective study, Lancet 365 (2005) 871–875.
821–826. [64] D.F. Rolfe, G.C. Brown, Cellular energy utilization and molecular origin of standard
[32] E.D. Crouser, Mitochondrial dysfunction in septic shock and multiple organ metabolic rate in mammals, Physiol. Rev. 77 (1997) 731–758.
dysfunction syndrome, Mitochondrion 4 (2004) 729–741. [65] V. Nogueira, L. Walter, N. Avéret, E. Fontaine, M. Rigoulet, X.M. Leverve, Thyroid
[33] R.S. Hotchkiss, P.E. Swanson, B.D. Freeman, K.W. Tinsley, J.P. Cobb, G.M. status is a key regulator of both flux and efficiency of oxidative phosphorylation
Matuschak, T.G. Buchman, I.E. Karl, Apoptotic cell death in patients with sepsis, in rat hepatocytes, J. Bioenerg. Biomembr. 34 (2002) 55–66.
shock, and multiple organ dysfunction, Crit. Care Med. 27 (1999) 1230–1251. [66] D.G. Nicholls, Forty years of Mitchell's proton circuit: From little grey books to
[34] M. Singer, V. De Santis, D. Vitale, W. Jeffcoate, Multiorgan failure is an adaptive, little grey cells, Biochim. Biophys. Acta — Bioenergetics. 1777 (2008) 550–556.
endocrine-mediated, metabolic response to overwhelming systemic inflamma- [67] C. Affourtit, P.G. Crichton, N. Parker, M.D. Brand, Novel uncoupling proteins,
tion, Lancet 364 (2004) 545–548. Novartis Found Symp. 287 (2007) 70–80 discussion 80–91.
[35] D. Nicholls, Mitochondrial bioenergetics, aging, and aging-related disease Science [68] M.D. Brand, J.L. Pakay, A. Ocloo, J. Kokoszka, D.C. Wallace, P.S. Brookes, E.J.
of aging knowledge environment: SAGE KE, 2002, 2002, pe12. Cornwall, The basal proton conductance of mitochondria depends on adenine
[36] G.C. Brown, Control of respiration and ATP synthesis in mammalian mitochondria nucleotide translocase content, Biochem. J. 392 (2005) 353–362.
and cells, Biochem. J. 284 (Pt 1) (1992) 1–13. [69] D. Arsenijevic, H. Onuma, C. Pecqueur, S. Raimbault, B.S. Manning, B. Miroux, E.
[37] R. Heinrich, T.A. Rapoport, Linear theory of enzymatic chains; its application for Couplan, M.C. Alves-Guerra, M. Goubern, R. Surwit, F. Bouillaud, D. Richard, S.
the analysis of the crossover theorem and of the glycolysis of human Collins, D. Ricquier, Disruption of the uncoupling protein-2 gene in mice reveals a
erythrocytes, Acta Biol. Med. Ger. 31 (1973) 479–494. role in immunity and reactive oxygen species production, Nat. Genet. 26 (2000)
[38] H. Kacser, J.A. Burns, The control of flux, Symp. Soc. Exp. Biol. 27 (1973) 65–104. 435–439.
[39] M.D. Brand, Top down metabolic control analysis, J. Theor. Biol.182 (1996) 351–360. [70] S. Rousset, Y. Emre, O. Join-Lambert, C. Hurtaud, D. Ricquier, A.M. Cassard-
[40] S. Schuster, D. Kahn, H.V. Westerhoff, Modular analysis of the control of complex Doulcier, The uncoupling protein 2 modulates the cytokine balance in innate
metabolic pathways, Biophys. Chem. 48 (1993) 1–17. immunity, Cytokine 35 (2006) 135–142.
[41] G.C. Brown, R.P. Hafner, M.D. Brand, A qtop–downq approach to the determination [71] M.P. Murphy, Slip and leak in mitochondrial oxidative phosphorylation, Biochim.
of control coefficients in metabolic control theory, Eur. J. Biochem. 188 (1990) Biophys. Acta 977 (1989) 123–141.
321–325. [72] B. Kadenbach, Intrinsic and extrinsic uncoupling of oxidative phosphorylation,
[42] A.K. Groen, R.J. Wanders, H.V. Westerhoff, R. van der Meer, J.M. Tager, Biochim. Biophys. Acta 1604 (2003) 77–94.
Quantification of the contribution of various steps to the control of mitochondrial [73] M.D. Brand, The efficiency and plasticity of mitochondrial energy transduction,
respiration, J. Biol. Chem. 257 (1982) 2754–2757. Biochem. Soc. Trans. 33 (2005) 897–904.
[43] E.K. Ainscow, M.D. Brand, The responses of rat hepatocytes to glucagon and [74] M.D. Brand, M.E. Harper, H.C. Taylor, Control of the effective P/O ratio of oxidative
adrenaline. Application of quantified elasticity analysis, Eur. J. Biochem. 265 phosphorylation in liver mitochondria and hepatocytes, Biochem. J. 291 (Pt 3)
(1999) 1043–1055. (1993) 739–748.
[44] M.E. Harper, M.D. Brand, Hyperthyroidism stimulates mitochondrial proton leak [75] J. Ciapaite, G. van Eikenhorst, S.J. Bakker, M. Diamant, R.J. Heine, M.J. Wagner, H.V.
and ATP turnover in rat hepatocytes but does not change the overall kinetics of Westerhoff, K. Krab, Modular kinetic analysis of the adenine nucleotide
substrate oxidation reactions, Can. J. Physiol. Pharmacol. 72 (1994) 899–908. translocator-mediated effects of palmitoyl-CoA on the oxidative phosphorylation
[45] F. Buttgereit, M.D. Brand, M. Müller, Effects of methylprednisolone on the energy in isolated rat liver mitochondria, Diabetes 54 (2005) 944–951.
metabolism of quiescent and ConA-stimulated thymocytes of the rat, Biosci. Rep. [76] C. Szabó, B. Zingarelli, M. O'Connor, A.L. Salzman, DNA strand breakage, activation
13 (1993) 41–52. of poly (ADP-ribose) synthetase, and cellular energy depletion are involved in the
[46] W.J. Hubbard, M. Choudhry, M.G. Schwacha, J.D. Kerby, L.W. Rue, K.I. Bland, I.H. cytotoxicity of macrophages and smooth muscle cells exposed to peroxynitrite,
Chaudry, Cecal ligation and puncture, Shock 24 (Suppl 1) (2005) 52–57. Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 1753–1758.
[47] D.W. Haden, H.B. Suliman, M.S. Carraway, K.E. Welty-Wolf, A.S. Ali, H. Shitara, H. [77] R.G. Boutilier, Mechanisms of cell survival in hypoxia and hypothermia, J. Exp.
Yonekawa, C.A. Piantadosi, Mitochondrial biogenesis restores oxidative metabo- Biol. 204 (2001) 3171–3181.
lism during Staphylococcus aureus sepsis, Am. J. Respir. Crit. Care Med. 176 (2007) [78] C. Bauer, A. Kurtz, Oxygen sensing in the kidney and its relation to erythropoietin
768–777. production, Annu. Rev. Physiol. 51 (1989) 845–856.
[48] S. Berg, P.L. Sappington, L.J. Guzik, R.L. Delude, M.P. Fink, Proinflammatory [79] D.I. Edelstone, M.E. Paulone, I.R. Holzman, Hepatic oxygenation during arterial
cytokines increase the rate of glycolysis and adenosine-5'-triphosphate turnover hypoxemia in neonatal lambs, Am. J. Obstet. Gynecol. 150 (1984) 513–518.
in cultured rat enterocytes, Crit. Care Med. 31 (2003) 1203–1212. [80] J. Lutz, H. Henrich, E. Bauereisen, Oxygen supply and uptake in the liver and the
[49] M.T. Frost, Q. Wang, S. Moncada, M. Singer, Hypoxia accelerates nitric oxide- intestine, Pflugers Arch. 360 (1975) 7–15.
dependent inhibition of mitochondrial complex I in activated macrophages, Am. [81] K. Nagano, S. Gelman, D.A. Parks, E.L. Bradley, Hepatic oxygen supply-uptake
J. Physiol. Regul. Integr. Comp. Physiol. 288 (2005) 394–400. relationship and metabolism during anesthesia in miniature pigs, Anesthesiology
[50] A. Jacob, M. Zhou, R. Wu, V.J. Halpern, T.S. Ravikumar, P. Wang, Pro-inflammatory 72 (1990) 902–910.
cytokines from Kupffer cells downregulate hepatocyte expression of adrenome- [82] D.P. Jones, H.S. Mason, Gradients of O2 concentration in hepatocytes, J. Biol. Chem.
dullin binding protein-1, Biochim. Biophys. Acta 1772 (2007) 766–772. 253 (1978) 4874–4880.
[51] I. Belikova, A.C. Lukaszewicz, V.F. Ass, C. Damoisel, M. Singer, D. Payen, Oxygen [83] T. Kashiwagura, D.F. Wilson, M. Erecińska, Oxygen dependence of cellular
consumption of human peripheral blood mononuclear cells in severe human metabolism: the effect of O2 tension on gluconeogenesis and urea synthesis in
sepsis, Crit. Care Med. (2007). isolated rat hepatocytes, J. Cell Physiol. 120 (1984) 13–18.
[52] R.P. Hafner, G.C. Brown, M.D. Brand, Analysis of the control of respiration rate, [84] V. Borutaite, S. Moncada, G.C. Brown, Nitric oxide from inducible nitric oxide
phosphorylation rate, proton leak rate and protonmotive force in isolated synthase sensitizes the inflamed aorta to hypoxic damage via respiratory inhibition,
mitochondria using the 'top–down' approach of metabolic control theory, Eur. J. Shock 23 (2005) 319–323.
Biochem. 188 (1990) 313–319. [85] A. Jekabsone, J.J. Neher, V. Borutaite, G.C. Brown, Nitric oxide from neuronal nitric
[53] G.C. Brown, P.L. Lakin-Thomas, M.D. Brand, Control of respiration and oxidative oxide synthase sensitises neurons to hypoxia-induced death via competitive
phosphorylation in isolated rat liver cells, Eur. J. Biochem. 192 (1990) 355–362. inhibition of cytochrome oxidase, J. Neurochem. 103 (2007) 346–356.
J.E. Carré, M. Singer / Biochimica et Biophysica Acta 1777 (2008) 763–771 771

[86] P. Mander, V. Borutaite, S. Moncada, G.C. Brown, Nitric oxide from inflammatory- [100] P. Boekstegers, S. Weidenhöfer, G. Pilz, K. Werdan, Peripheral oxygen availability
activated glia synergizes with hypoxia to induce neuronal death, J. Neurosci. Res. within skeletal muscle in sepsis and septic shock: comparison to limited infection
79 (2005) 208–215. and cardiogenic shock, Infection 19 (1991) 317–323.
[87] P.T. Schumacker, N. Chandel, A.G. Agusti, Oxygen conformance of cellular respi- [101] M. Sair, P.J. Etherington, C. Peter Winlove, T.W. Evans, Tissue oxygenation and
ration in hepatocytes, Am. J. Physiol. 265 (1993) 395–402. perfusion in patients with systemic sepsis, Crit. Care Med. 29 (2001) 1343–1349.
[88] P.T. Schumacker, Current paradigms in cellular oxygen sensing, Adv. Exp. Med. [102] A. Protti, J. Carré, M.T. Frost, V. Taylor, R. Stidwill, A. Rudiger, M. Singer, Succinate
Biol. 543 (2003) 57–71. recovers mitochondrial oxygen consumption in septic rat skeletal muscle, Crit.
[89] G.L. Wang, G.L. Semenza, General involvement of hypoxia-inducible factor 1 in Care Med. (2007) Publish Ahead of Print.
transcriptional response to hypoxia, Proc. Natl. Acad. Sci. U. S. A. 90 (1993) [103] E.D. Crouser, M.W. Julian, D.V. Blaho, D.R. Pfeiffer, Endotoxin-induced mitochon-
4304–4308. drial damage correlates with impaired respiratory activity, Crit. Care Med. 30
[90] D.G. Hardie, The AMP-activated protein kinase pathway—new players upstream (2002) 276–284.
and downstream, J. Cell Sci. 117 (2004) 5479–5487. [104] E. Nisoli, E. Clementi, S. Moncada, M.O. Carruba, Mitochondrial biogenesis as a
[91] G. Broder, M.H. Weil, Excess lactate: an index of reversibility of shock in human cellular signaling framework, Biochem. Pharmacol. (2004).
patient, Science 143 (1964) 1457–1459. [105] E. Nisoli, E. Clementi, C. Paolucci, V. Cozzi, C. Tonello, C. Sciorati, R. Bracale, A.
[92] U.G. Hodgin, J.P. Sanford, Gram-negative rod bacteremia. An analysis of 100 Valerio, M. Francolini, S. Moncada, M.O. Carruba, Mitochondrial biogenesis in
patients, Am. J. Med. 39 (1965) 952–960. mammals: the role of endogenous nitric oxide, Science 299 (2003) 896–899.
[93] L. Gattinoni, L. Brazzi, P. Pelosi, R. Latini, G. Tognoni, A. Pesenti, R. Fumagalli, A trial [106] H.B. Suliman, K.E. Welty-Wolf, M.S. Carraway, L. Tatro, et al., Lipopolysaccharide
of goal-oriented hemodynamic therapy in critically ill patients. SvO2 Collabora- induces oxidative cardiac mitochondrial damage and biogenesis, Cardiovascular
tive Group, N. Engl. J. Med. 333 (1995) 1025–1032. Res. (2004).
[94] M.A. Hayes, A.C. Timmins, E.H. Yau, M. Palazzo, C.J. Hinds, D. Watson, Elevation of [107] H. Esterbauer, H. Oberkofler, F. Krempler, W. Patsch, Human peroxisome
systemic oxygen delivery in the treatment of critically ill patients, N. Engl. J. Med. proliferator activated receptor gamma coactivator 1 (PPARGC1) gene: cDNA
330 (1994) 1717–1722. sequence, genomic organization, chromosomal localization, and tissue expres-
[95] A. Dyson, R. Stidwill, V. Taylor, M. Singer, Tissue oxygen monitoring in rodent sion, Genomics 62 (1999) 98–102.
models of shock, Am. J. Physiol. Heart Circ. Physiol. 293 (2007) 526–533. [108] D. Knutti, A. Kaul, A. Kralli, A tissue-specific coactivator of steroid receptors,
[96] D. De Backer, M.J. Dubois, Assessment of the microcirculatory flow in patients in identified in a functional genetic screen, Mol. Cell Biol. 20 (2000) 2411–2422.
the intensive care unit, Curr. Opin. Crit. Care 7 (2001) 200–203. [109] Z. Wu, P. Puigserver, U. Andersson, C. Zhang, G. Adelmant, V. Mootha, A. Troy, S.
[97] D.M. Rosser, M. Manji, H. Cooksley, G. Bellingan, Endotoxin reduces maximal Cinti, B. Lowell, R.C. Scarpulla, B.M. Spiegelman, Mechanisms controlling mito-
oxygen consumption in hepatocytes independent of any hypoxic insult, Intensive chondrial biogenesis and respiration through the thermogenic coactivator PGC-1,
Care Med. 24 (1998) 725–729. Cell 98 (1999) 115–124.
[98] D.M. Rosser, R.P. Stidwill, D. Jacobson, M. Singer, Cardiorespiratory and [110] C. Handschin, B.M. Spiegelman, Peroxisome proliferator-activated receptor
tissue oxygen dose response to rat endotoxemia, Am. J. Physiol. 271 (1996) gamma coactivator 1 coactivators, energy homeostasis, and metabolism, Endocr.
891–895. Rev. 27 (2006) 728–735.
[99] T.J. VanderMeer, H. Wang, M.P. Fink, Endotoxemia causes ileal mucosal acidosis in [111] P. Puigserver, J. Rhee, J. Lin, Z. Wu, J.C. Yoon, C.Y. Zhang, S. Krauss, V.K. Mootha, B.B.
the absence of mucosal hypoxia in a normodynamic porcine model of septic Lowell, B.M. Spiegelman, Cytokine stimulation of energy expenditure through p38
shock, Crit. Care Med. 23 (1995) 1217–1226. MAP kinase activation of PPARgamma coactivator-1, Mol. Cell 8 (2001) 971–982.

Вам также может понравиться