Mitochondria produce metabolic energy, serve as biosensors for oxidative stress, and eventually become effector organelles for cell death through apoptosis. The extent to which these manifold mitochondrial functions are altered by previously unrecognized actions of anesthetic agents seems to explain and link a wide variety of perioperative phenomena that are currently of interest to anesthesiologists from both a clinical and a scientific perspective. In addition, many surgical patients may be at increased perioperative risk because of inherited or acquired mitochondrial dysfunction leading to increased oxidative stress. This review summarizes the essential aspects of the bioenergetic process, presents current knowledge regarding the effects of anesthetics on mitochondrial function and the extent to which mitochondrial state determines anesthetic requirement and potential anesthetic toxicity, and considers some of the many implications that our knowledge of mitochondrial dysfunction poses for anesthetic management and perioperative medicine.

MITOCHONDRIA not only generate and modulate bioenergy but also serve as the final effectors for the termination of cell viability as organisms approach the end of their lifespan. Therefore, the implications of these processes with regard to understanding evolution, disease, aging, and death are profound. Particularly relevant to anesthesiologists is the role of mitochondria in determining the response of the nervous system to anesthetic agents, in initiating mechanisms of cell injury or protection after ischemic, hypoxic, or toxic injuries, and their ability to precipitate critical illness in individuals with inherited or acquired mitochondrial disorders. These aspects of mitochondrial biology and pathophysiology will be briefly summarized in this clinically oriented review.

Mitochondria produce the energy needed for normal cellular function and metabolic homeostasis by oxidative phosphorylation,1a process conducted by a series of five enzyme complexes located on the inner mitochondrial membrane (fig. 1). Four of these complexes comprise the mitochondrial electron transport chain (ETC) and function as a biochemical “conveyor belt” for electrons. Oxidative phosphorylation couples the oxidation of reduced nicotinamide adenine dinucleotide and flavin adenine dinucleotide, generated by the Krebs cycle and by the β-oxidation of fatty acids, to the phosphorylation of adenosine diphosphate (ADP) to adenosine triphosphate (ATP). Electron donation to complex I (reduced nicotinamide adenine dinucleotide–ubiquinone oxidoreductase) initiates this process. Alternatively, electrons originating from succinate and from reduced flavin adenine dinucleotide can be channeled into the ETC through complex II (succinate–ubiquinone oxidoreductase). Electrons are transported from complex I or II to complex III (ubiquinone–cytochrome c  oxidoreductase) via  a mobile electron carrier, coenzyme Q (ubiquinone), and subsequently on to complex IV (cytochrome c  oxidase) via  cytochrome c . Complex IV uses electrons from cytochrome c  to reduce molecular oxygen, the final acceptor of electrons, to water at that site.

Fig. 1. Schematic representation of the mitochondrial components needed for oxidative phosphorylation. Complexes I–IV, located within the inner mitochondrial membrane, are oxidase complexes that, along with coenzyme Q (Co Q) and cytochrome  c (Cyto C), comprise the electron transport chain.  Dotted lines indicate pathway for electron flow. Complexes I, III, and IV also pump hydrogen ions (  dashed lines ) into the intermembrane space and generate the electrochemical gradient that ultimately powers the phosphorylation of adenosine diphosphate (ADP) to adenosine triphosphate (ATP) by ATP synthase. Inner membrane–bound uncoupling protein (UCP) is an alternate return path for hydrogen ions. Adenine nucleotide translocase (ANT) regulates the balance of ATP and ADP within the mitochondrial matrix. FADH2= flavin adenine dinucleotide; H2O = water; NAD = nicotinamide adenine dinucleotide; NADH = reduced nicotinamide adenine dinucleotide; O2= oxygen; Pi = inorganic phosphate. Copyright © 2000 American Diabetes Association. Modified with permission from The American Diabetes Association, from Boss  et al. 4 

Fig. 1. Schematic representation of the mitochondrial components needed for oxidative phosphorylation. Complexes I–IV, located within the inner mitochondrial membrane, are oxidase complexes that, along with coenzyme Q (Co Q) and cytochrome  c (Cyto C), comprise the electron transport chain.  Dotted lines indicate pathway for electron flow. Complexes I, III, and IV also pump hydrogen ions (  dashed lines ) into the intermembrane space and generate the electrochemical gradient that ultimately powers the phosphorylation of adenosine diphosphate (ADP) to adenosine triphosphate (ATP) by ATP synthase. Inner membrane–bound uncoupling protein (UCP) is an alternate return path for hydrogen ions. Adenine nucleotide translocase (ANT) regulates the balance of ATP and ADP within the mitochondrial matrix. FADH2= flavin adenine dinucleotide; H2O = water; NAD = nicotinamide adenine dinucleotide; NADH = reduced nicotinamide adenine dinucleotide; O2= oxygen; Pi = inorganic phosphate. Copyright © 2000 American Diabetes Association. Modified with permission from The American Diabetes Association, from Boss  et al. 4 

Close modal

Intrinsically linked to this process of electron transport is the generation and maintenance of a hydrogen ion gradient across the inner mitochondrial membrane. The inner membrane separates the intermembrane space from the mitochondrial matrix. The gradient is established by proton pumps in ETC complexes I, III, and IV. The F1F0–ATPase (ATP synthase) complex within the inner membrane uses this proton motive force to phosphorylate ADP. This last step in the overall process of oxidative phosphorylation produces the ATP that serves as the fundamental “currency” needed for most energy-requiring biologic transactions. Another membrane-integrated protein, adenine nucleotide translocase, regulates an “antiport” process that moves ADP and ATP in opposite directions across the inner mitochondrial membrane. Adenine nucleotide translocase delivers ATP to energy-requiring sites, mostly in the cytosol, and simultaneously resupplies the ATP synthase complex with new substrate.

The hydrogen ion gradient established by the process of oxidative phosphorylation can also be dissipated by proton leakage back into the matrix through the inner membrane that bypasses the ATP synthase complex. Uncoupling proteins (UCPs) within the inner membrane provide this alternate pathway for proton influx. In effect, UCPs convert some of the electrochemical energy generated by the ETC into heat rather than into ATP. The rate of proton leakage through a UCP seems to be influenced by a variety of conditions, including changes in the magnitude of the hydrogen ion gradient itself, increased catecholamines levels, and variations in fatty acid concentrations.2UCP-1, also called thermogenin, was originally characterized in the mitochondria of brown fat cells3and is now known to play a role in nonshivering thermogenesis in human neonates. Subsequently, additional UCP isoforms were identified in a variety of tissues.4Although their precise metabolic functions have yet to be determined, UCPs may play an important role in adult obesity, diabetes mellitus,5and perhaps other conditions where the regulation of oxidative metabolism seems to be disrupted.

The mitochondrion, unique among mammalian organelles, contains multiple copies of a small circular genome of approximately 16,000 nucleotide base pairs. This mitochondrial DNA (mtDNA) has been completely characterized in humans.6mtDNA encodes for some key subunits needed for electron transport and oxidative phosphorylation, although the majority of mitochondrial proteins needed for normal bioenergetic function are encoded by nuclear DNA (nDNA)7and therefore must be imported into the mitochondrial matrix from the cell cytosol.8Complex IV of the ETC, for example, contains 13 subunits, 10 of which are encoded in nDNA. The expression of the mitochondrial genome itself requires a single mitochondrial transcription factor that arises from the nuclear genome.9 

Overall, the human mitochondrial genome encodes for 13 peptides (subunits of complexes I, III, and IV and the ATP synthase complex), 2 ribosomal ribonucleic acids (RNAs), and 22 transfer RNAs. Nuclear DNA encodes for at least 1,000 proteins that are needed for mitochondrial bioenergetic and metabolic functions and for mtDNA expression and replication.10Although there may be as many as 1,000 copies of mtDNA in most cells, acquired mtDNA point mutations and base pair deletions are extremely rare and are normally found in only a minute proportion of total mtDNA11despite the fact that mtDNA, unlike nDNA, lacks histone protection and is surrounded by potentially damaging oxidative influences.12This observation supports the hypothesis that there must be effective molecular repair and disposal mechanisms for damaged mtDNA within mitochondria.13,14 

Oxidative phosphorylation is the major intracellular source of reactive oxygen species (ROS). ROS or “free radicals” such as superoxide, peroxide, or hydroxyl radicals (O2, H2O2, and OH) are routinely generated as byproducts of the interaction between free electrons and oxygen. ROS are an unavoidable consequence of aerobic metabolism and can be produced anywhere there is leakage of electrons from the ETC. Although only a tiny percentage of metabolically consumed oxygen is converted into ROS, these ephemeral but highly reactive molecules can degrade or destroy mitochondrial enzyme complexes, membranes, and other structural components of cell microarchitecture, either by direct contact or through lipid peroxidation.15ROS such as hydroxyl and reactive nitrogen species such as peroxynitrite (ONO2formed from the interaction of superoxide and nitric oxide) react almost instantly with proteins to generate protein carbonyls16,17and with polyunsaturated fatty acids in membranes to generate a variety of lipid peroxidation products including 4-hydroxynonenal and malondialdehyde.18These reaction products drastically reduce the membrane fluidity needed for normal cell function. With half-lives of minutes to hours,19peroxidation products may impact multiple membrane-bound systems and precipitate a series of damaging “chain reaction” peroxidation sequences in adjacent cells.20 

All obligate aerobes have intrinsic defensive systems to protect against damage by ROS.21This includes several forms of superoxide dismutase, catalase, and glutathione peroxidase. Copper, zinc superoxide dismutase (present in the cytosol),22and manganese superoxide dismutase (found within the mitochondrion)23,24convert superoxide into oxygen and hydrogen peroxide. The bulk of hydrogen peroxide is quickly broken down by catalases into water and oxygen,25although some peroxide in human cells, particularly neurons, is inactivated either by glutathione peroxidase26or by the more recently described peroxiredoxins,27a group of thioredoxin-dependent antioxidant peroxidases that reduce peroxynitrite and also modulate the role of peroxide as a second messenger molecule.28 

Overall, endogenous antioxidant defense systems provide effective homeostasis with regard to suppressing ROS levels within the cell as well as within mitochondria, and there are endogenous “backup” systems to repair ROS-mediated damage if it occurs.29In many disease states, however, and perhaps in normal aging, mechanisms that prevent or limit ROS-mediated damage may become inadequate.30Elevation of ROS beyond normal levels, regardless of etiology, inevitably produces oxidative stress. The insidious onset or sustained low level of oxidative stress can be cytoprotective if it induces or enhances the expression of additional antioxidant or molecular repair systems. In contrast, however, rapid or overwhelming increases in ROS are fundamentally cytotoxic, primarily through disruption of intracellular calcium regulation or by initiating the destructive sequence of events known as apoptosis.

Apoptosis, or programmed cell death, is a genetically controlled event that has biologic value because it permits prompt and orderly disposal of damaged, infected, or aging cells, especially in the nervous system,31and has survival value for the species because it facilitates complex organogenesis and tissue development.32It may also limit the spread of “rogue” or neoplastic cells.33,34It is a biochemical cascade usually mediated by caspases, a large family of proteolytic enzymes that activate the nucleases that digest DNA (fig. 2). Each cell within a multicellular organism seems to possess multiple overlapping mechanisms to accomplish what is, in effect, “cellular suicide.” Because nDNA fragmentation is prominent in this process, apoptosis was originally thought to be solely a function of cell nuclei. Now, however, the central role of mitochondria in cellular apoptosis is universally acknowledged.35 

Fig. 2. Simplified schematic of major pathways for apoptosis. The intrinsic or mitochondrial pathway requires a cascade of events beginning with a mitochondrial permeability transition (MPT) that can be triggered by a variety of stimuli, including oxidative stress, high levels of nitric oxide (NO), acute hypercalcemia, or release of proapoptotic “death proteins.” The intrinsic cascade releases cytochrome  c (Cyto C) from the mitochondrial inner membrane into the cytosol, which triggers caspase activation facilitated by apoptotic protease activation factor 1 (Apaf-1). Activation of nucleases may also occur without caspase activation when an MPT releases apoptosis-inducing factor (AIF) from the mitochondrion. Extrinsic pathways for apoptosis have been described that require activation of cell surface  N -methyl-d-aspartate (NMDA) receptors by neurotransmitters such as glutamate or by “death receptors” triggered by tumor necrosis factor (TNF) or other cytokines.  Dotted lines indicate evidence of protective or antiapoptotic properties for low levels of NO, some Bcl proteins (Bcl-2), and heat shock proteins (HSPs). mtDNA = mitochondrial DNA; nDNA = nuclear DNA. 

Fig. 2. Simplified schematic of major pathways for apoptosis. The intrinsic or mitochondrial pathway requires a cascade of events beginning with a mitochondrial permeability transition (MPT) that can be triggered by a variety of stimuli, including oxidative stress, high levels of nitric oxide (NO), acute hypercalcemia, or release of proapoptotic “death proteins.” The intrinsic cascade releases cytochrome  c (Cyto C) from the mitochondrial inner membrane into the cytosol, which triggers caspase activation facilitated by apoptotic protease activation factor 1 (Apaf-1). Activation of nucleases may also occur without caspase activation when an MPT releases apoptosis-inducing factor (AIF) from the mitochondrion. Extrinsic pathways for apoptosis have been described that require activation of cell surface  N -methyl-d-aspartate (NMDA) receptors by neurotransmitters such as glutamate or by “death receptors” triggered by tumor necrosis factor (TNF) or other cytokines.  Dotted lines indicate evidence of protective or antiapoptotic properties for low levels of NO, some Bcl proteins (Bcl-2), and heat shock proteins (HSPs). mtDNA = mitochondrial DNA; nDNA = nuclear DNA. 

Close modal

The rate-limiting fundamental step in the mitochondrial apoptotic pathway is induction of a mitochondrial permeability transition (MPT),36an electrochemical event characterized by transient influx of solutes through large pores or “megachannels” in the otherwise essentially impermeable mitochondrial inner membrane. During MPT-induced permeabilization, collapse of the transmembrane electrochemical gradient for hydrogen ions stops the process of oxidative phosphorylation. In addition, there is efflux of cytochrome c from the intermembrane space into the cytosol. There, cytochrome c combines with apoptotic protease activation factor 1 and dATP to form a complex that oligomerizes, recruits, and activates procaspase 9. Subsequent procaspase-3 recruitment forms an “apoptosome “and activates caspase 3. This leads to the activation of nucleases that completes the “intrinsic” or mitochondrial-dependent pathway. Destruction of nDNA and mtDNA by nucleases is irreversible and complete within a few hours.

There is also an “extrinsic” pathway that can initiate caspase activation and apoptosis without mitochondrial involvement. Apoptosis can be triggered by the binding of a signaling messenger such as tumor necrosis factor or similar extrinsic cytokine37to a cell surface “death receptor.” It seems that a variety of signal transducers and activators of transcription may play a role in the suppression or activation of apoptosis, especially with neoplastic cells.38In addition, there is at least one apoptotic cascade that uses apoptosis-inducing factor (AIF), a flavoprotein sequestered in the intermembrane region of the mitochondrion, to initiate apoptosis without caspase activation. AIF normally stabilizes mitochondrial membrane permeability and supports oxidative phosphorylation.39However, if released through the outer membrane into the cytosol, AIF can produce terminal damage to nDNA.

The precise regulation of calcium ion flux across mitochondrial membranes is essential both to normal mitochondrial bioenergetic function and to the apoptotic process.40Calcium is of special importance because it is a signaling cation for many preprogrammed cell functions. Calcium concentrations within the cytosol are normally orders of magnitude less than extracellular calcium concentrations. With their capacity for calcium uptake, mitochondria may therefore function as a reservoir or buffer to stabilize calcium concentrations within the cytosol at very low levels. It is also possible that tiny fluctuations in cytosol calcium concentrations are “sensed” within the mitochondrion, providing a control system that links changes in cellular energy demand to the rate of energy production by oxidative phosphorylation.41 

Pathways for apoptosis involve counterbalancing concentrations of antiapoptotic and proapoptotic proteins, many of which are encoded by the Bcl-2 B-cell leukemia gene.33,42Although the precise role and interactions between the members of this large family of proteins are still under intense investigation,37they modulate the likelihood of an MPT initiating permeabilization and triggering apoptosis, probably through their effects on inner membrane stability.43In addition, Bcl-2 family proteins such as Bid may act as a “bridge” between different apoptotic pathways, allowing them to share some, but not all, mediators.44Oxidative stress or high levels of ionized calcium within the mitochondrial matrix can also induce an MPT. During a state of oxidative stress, there is a synergistic relation between MPT and ROS formation that produces an upward spiral of mitochondrial damage and continuing release of ROS and calcium into the cytosol.45 

It has been difficult to distinguish clearly between proapoptotic and antiapoptotic factors. For example, at low intracellular concentrations, nitric oxide is a potent but reversible inhibitor of cellular respiration and oxygen consumption46and inhibits mitochondrial-dependent apoptosis. At higher levels, however, or in the presence of increased intracellular or intramitochondrial calcium, nitric oxide enhances the apoptogenic effects of ROS or may even become a “death messenger” capable of initiating apoptosis.47It has also recently been emphasized that caspases, as do cytochrome c  and AIF, play a prominent role in programmed cell death but are also essential to many vital nonapoptotic cell processes.48Therefore, therapeutic strategies that suppress or block the effects of putative proapoptotic agents may produce unintended interruptions of other cell functions and actually compromise cell viability. Clearly, more investigation is necessary to define the importance and role of apoptosis in the maintenance of normal cellular function and in the pathogenesis of disease.

Although the extent to which they alter mitochondrial function in vivo  is not yet understood, it has long been known that intravenous drugs with anesthetic properties can depress carbohydrate metabolism, oxygen consumption, and energy production in the nervous system.49Early studies of narcotics demonstrated that they inhibit oxidation of glucose, lactate, and pyruvate in neural tissues at clinically relevant concentrations,50and seven decades later, it has been proposed that morphine may actually have a mitochondrial-based mechanism of clinical action.51It has also been recently shown that propofol markedly decreases oxygen consumption and ATP production in brain synaptososmes52and reduces electron flow in cardiac mitochondria.53,54Propofol inhibits complex I of the ETC but may also effect ATPase and UCPs, uncoupling electron transport from ATP production.55,56The primary effect of barbiturates on oxidative phosphorylation in mitochondria obtained from brain, heart, and liver also seems to be inhibition of complex I, and, like propofol, they seem to “uncouple ” metabolic activity from ATP production, further reducing bioenergetic capacity.57 

Inhalational anesthetic agents have similar depressant effects on mitochondrial respiration, at least in vitro .58–62Studies of cardiac mitochondria exposed to halothane, isoflurane, and sevoflurane suggest that the most common site of action is, again, inhibition of complex I.63At concentrations equal to twice minimum alveolar concentration, complex I activity is reduced by 20% after exposure to halothane and isoflurane and by 10% after exposure to sevoflurane. Oxidative phosphorylation in isolated liver mitochondria is also measurably depressed after exposure to halothane.64Concentrations of 0.5–2% halothane lead to reversible inhibition of complex I that is further exacerbated by the addition of nitrous oxide,65although nitrous oxide itself seems to have little effect on ATP production.66 

Local anesthetics also depress bioenergetic capacity and disrupt oxidative phosphorylation67–70in a manner similar to that of the intravenous agents. Given these observations, it is tempting to speculate that reversible inhibition of mitochondrial electron transfer and decreased energy availability within neural tissues provide a unitary and simple hypothesis for the mechanism of anesthesia. However, the effect of any of these drugs on mitochondrial function or apparent bioenergetics may be incidental and does not necessarily explain their anesthetic actions. In addition, given the demonstrated ability of cells to “down-regulate” their metabolic activity under a variety of conditions,71ATP levels do not provide a sensitive measure of the energy state of an intact cell or of its capacity for oxidative phosphorylation. In fact, much of the available data regarding in vivo  ATP levels in various tissues at clinically relevant concentrations of inhalational agents72–74suggests that there is no consistent change. Consequently, the most recent concepts regarding the underlying mechanism of general anesthesia emphasize the complexity, rather than the simplicity, of anesthetic effects and the high probability that there are multiple effect sites for anesthetics, probably involving transmembrane receptor protein structures.75 

Similarly, propofol-induced depression of myocardial bioenergetics at low clinical concentrations is not sufficient to account for observed alterations of contractile function.76Although earlier work proposed that impaired bioenergetics might be a primary factor in anesthetic-induced depression of myocardial function,77the depression seen with inhalational agents seems to be due to the consequences of impaired calcium utilization on excitation–contraction coupling rather than to inadequate myocardial energy availability.78Although anesthetics therefore probably do not exert their obvious clinical effects through limitation of ATP availability, they may interfere with ATP utilization, produce oxidative stress, or impair mitochondrial function in some other manner.

In isolated individual neurons, 5 min of exposure to lidocaine at clinically relevant concentrations initiates an MPT and complete loss of mitochondrial membrane electrochemical potential.79Subsequently, there is release of mitochondrial cytochrome c  into the cytoplasm and activation of caspases. This suggests that even simple exposure of nerve cells to local anesthetics may be sufficient to trigger apoptotic pathways. The myocardial toxicity of bupivacaine seems to be similarly mediated by a mechanism involving mitochondrial bioenergetics.80It has been proposed that anesthetics impair the ability of the mitochondrion to function as a “biosensor” for oxidative stress, disrupting the normal balance between ROS and endogenous antioxidants or antiapoptotic molecules.81 

Whatever their relative importance to the production of the anesthetic state, the effects of anesthetics on mitochondria described above largely occur on the inner membrane of mitochondria. Therefore, they seem to reflect the physicochemical actions of anesthetics82,83on the lipid or protein components of mitochondrial membranes.84,85In general, it seems a reasonable working hypothesis that drugs with anesthetic properties influence bioenergetic activity through disruption of mitochondrial membrane structure,86either by diffuse, or perhaps by highly specific, effects at lipid or protein sites.87A strong correlation between their anesthetic potency and affinity for cytochrome c  oxidase (ETC complex IV) suggests that it may be a discrete target site for local anesthetics.88The consistent relation between inhibition of complex IV and the octanol–water partition coefficient of local anesthetics89also suggests that lipophilic interactions may produce reversible, short-term distortion or perturbation of essential ETC components. Reversible depression of oxidative phosphorylation in mammals has recently been shown to be initiated by molecules as simple as hydrogen sulfide.90 

Manipulation of the nuclear genome of nematodes has shown a direct link between the composition of mitochondrial proteins and anesthetic requirement.91,92A defect in a subunit of complex I of the ETC is associated with depressed mitochondrial bioenergetics and hypersensitivity to volatile anesthetics.93,94In addition, there seems to be a clear, albeit empirical, correlation between increasing age and declining anesthetic requirement in humans from mid-adulthood through senescence,95a time period during which bioenergetics also seem to be progressively depressed.96Finally, and most recently documented, is the increased sensitivity to inhalational anesthesia seen in children who have depressed mitochondrial bioenergetics due to inherited mitochondrial cytopathies.97These observations hint at a fundamental but still undefined relation between anesthetic requirement and mitochondrial function within the nervous system.

N -methyl-d-aspartate antagonism or γ-aminobutyric acid receptor stimulation can initiate neuronal apoptosis, at least in immature brain tissue.98This seems to provide a physiologic mechanism to facilitate brain development or to cull redundant or failing neurons and provide neuroplasticity. The process may become pathologic when immature or minimally stressed neurons are exposed to drugs such as anesthetic agents, which generally have N -methyl-d-aspartate antagonist or γ-aminobutyric acid mimetic properties. In fact, widespread nonphysiologic apoptosis and neurodegeneration have been observed in laboratory rodent fetal brains after short-term anesthetic exposure99as well as in adult brain after prolonged exposure to nitrous oxide.100Even in mature brain, the transition of immature cells into more highly differentiated neurons with the complex synaptic structure needed for learning could be compromised by routine anesthetic exposure. This hypothesis is supported by recent investigations demonstrating that cognitive deficits persist in aged, but not in young adult, laboratory rodents after routine inhalational anesthesia.101,102Exposure to anesthetic agents also measurably depresses mitochondrial bioenergetics in peripheral T lymphocytes, possibly contributing to impairment of perioperative immune competence.103 

Some preliminary clinical data could also be interpreted to support the hypothesis that anesthetics have intrinsic potential neurotoxicity. In elderly surgical patients, for example, deeper levels of inhalational anesthesia are associated with more severe early postoperative cognitive impairment as well as with a significantly decreased probability of postoperative survival.104,105This suggests that in individuals with limited nervous system reserve or impaired tolerance for oxidative stress, prolonged exposure, or higher anesthetic concentrations could be, in effect, neurotoxic.106Anesthetic exposure may increase mitochondrial ROS sufficiently in some individuals to damage cells through a lipid peroxidation pathway.107Therefore, it is possible that both the desired clinical effects of anesthetics as well as their potential to injure neurons may reflect their interaction with mitochondria, although there is obviously need for caution before extrapolating from laboratory observations to clinical practice.108 

The scope of human disease attributable to inherited, acutely acquired, or insidious impairment of mitochondrial function is clearly far greater than had been previously believed.109,110Given the universal role of mitochondrial bioenergetics in sustaining the normal function of cells in every tissue and organ, mitochondrial cytopathy or short-term mitochondrial dysfunction can potentially produce virtually any symptom, in any organ system, at any stage of life. Many presumably unique “diseases” may actually be expressions of progressive organ system dysfunction due to disordered oxidative metabolism or disruption of other aspects of mitochondrial function. In fact, with more than a hundred mtDNA mutations implicated in human disease,111mitochondrial dysfunction is emerging as a primary focus for investigations into the etiology of sepsis, neurodegenerative disorders, diabetes, cardiovascular disease, and various forms of hepatic and metabolic derangement.112Anesthesiologists are therefore in a unique position to observe and to explore the relevance of congenital and acquired cytopathies to perioperative patient care.

The terms mitochondrial myopathy, inherited mitochondrial encephalomyopathy, and mitochondrial cytopathy are generally equivalent. Clinically, they encompass a wide variety of neurologic syndromes, most described only within the past three decades, that are due to errors in the synthesis of mitochondrial proteins caused by defects in nDNA, mtDNA, or mitochondrial transfer RNA ( appendix 1). Symptoms generally reflect inadequate oxidative phosphorylation, usually first apparent in skeletal muscle or in the retina or other parts of the nervous system with high energy requirements.113,114In addition, inherited or acquired respiratory chain enzymatic deficiencies degrade the efficiency of oxidative phosphorylation and can result in excessive levels of ROS.115Subclinical hepatic and renal involvement is common, but the diagnosis of a mitochondrial-based respiratory chain deficiency is often not considered unless associated with evidence of skeletal muscle weakness or encephalopathy.

The phenotypic variability of inherited mitochondrial cytopathies reflects the uneven distribution of mutant mtDNA to different tissues during the early phases of embryogenesis.116Consequently, even when a defined mtDNA mutation is involved, patients with mitochondrial disorders may present with a wide variety of symptoms, many of them extremely vague or subtle. Mitochondrial cytopathy should be included in the differential diagnosis whenever persistent clinical signs and symptoms include muscle pain in conjunction with weakness or fatigue117or if there is diffuse involvement of several organ systems that does not conform to an established pattern of conventional disease.114 

Because mitochondrial cytopathies involve enzymatic defects that lead to organ dysfunction through impaired oxidative phosphorylation, lactic acidosis and abnormalities in glucose metabolism are common sequelae. The diagnostic algorithm for suspected mitochondrial cytopathy investigations therefore should include screening for measurement of serum and spinal fluid lactate and increased lactate/pyruvate as well as ketone body molar ratios. For pediatric patients, the diagnostic process includes both blood and urine testing, although normal lactate and glucose values do not necessarily rule out the diagnosis of mitochondrial disease. When the index of suspicion for mitochondrial cytopathy is very high in children or in adults, skeletal muscle biopsy can confirm the diagnosis if it reveals the characteristic ragged-red fibers on trichrome stain, which are caused by accumulations of defective mitochondria beneath the sarcolemmal membrane, excess glycogen granules, and cytochrome c  oxidase (complex IV) deficient cells.118 

Biopsy of muscle or skin can also provide material for mtDNA analysis and facilitate genetic counseling. Syndromes caused by inherited mtDNA point deletions or insertions such as Leber hereditary optic neuropathy or NARP (neuropathy, ataxia, retinitis pigmentosa) can be detected by a polymerase chain reaction blood test and are generally maternally inherited.119Similarly, mitochondrial encephalopathy with lactic acidosis and stroke-like episodes, myoclonus epilepsy and ragged-red fibers, and maternally inherited disorder with adult-onset myopathy and cardiomyopathy, each of which is the consequence of a single transfer RNA missense mutation, also follow maternal inheritance patterns.120However, Pearson121and Kearns-Sayre122syndromes, both produced by a single mtDNA base pair deletion or insertion, have sporadic inheritance patterns.123Large-scale mtDNA deletions are usually acquired, not inherited, defects.124 

Mutations of nDNA that produce unstable mtDNA can produce mitochondrial cytopathy syndromes that are clinically indistinguishable from those associated with classic mtDNA mutations.125,126One example is an inherited defect in the nuclear gene that encodes for the mitochondrial transcription factor, producing an inevitably fatal mtDNA deficiency syndrome of infancy.127mtDNA depletion syndrome is a severe disease of childhood characterized by liver failure and neurologic abnormalities due to tissue-specific loss of functional mtDNA. This syndrome is thought to be caused by a putative nuclear gene that controls mtDNA replication or stability.128Similarly, children with mitochondrial neurogastrointestinal encephalomyopathy may have multiple mtDNA deletions and/or mtDNA depletion that results from an nDNA mutation.129Regardless of etiology, however, mitochondrial cytopathies of infancy invariably compromise the developing nervous system and are therefore diagnosed early because symptoms are severe and progress rapidly. Nonspecific neurologic signs include lethargy, irritability, hyperactivity, and poor feeding.

Other variants of inherited cytopathy present later in childhood or even in the young and middle adult years. In these syndromes, subclinical decreases in cardiac, skeletal muscle, and nervous system functional reserve probably begin long before the appearance of overt signs or symptoms. Therefore, preoperative assessment of organ system functional reserve such as maximal oxygen uptake is more useful than routine preoperative “screening” tests in defining the extent to which declining mitochondrial energy production has produced clinical compromise. Patients may ultimately be diagnosed during the evaluation of unexplained muscle weakness, ventilatory failure,130or even upper airway obstruction.131Deterioration is gradual but progressive and inevitably leads to incapacitation. Some mtDNA mutations accumulate over time in a single tissue type (e.g. , skeletal muscle) where clinical deterioration during adulthood correlates with an increasing fraction of mutant mtDNA.132In fact, in patients with skeletal muscle mtDNA mutations, the “mutation load” determines the extent of metabolic impairment and therefore the degree of exercise intolerance as indicated clinically by a reduced rate of muscle oxygen extraction in the face of exaggerated cardiopulmonary responses.133Measurement of venous oxygen partial pressure during forearm exercise may therefore be of value, at least in adults, to assess the severity of aerobic compromise due to mitochondrial dysfunction.134Nevertheless, the true incidence of these later-onset syndromes is unclear because of their insidious onset and the diversity of organ systems involved.135,136 

For both childhood- and adult-onset cytopathies, the general principles of perioperative medical management are comprehensive interdisciplinary consultation and the expectation of a need for supportive care to avoid metabolic acidosis or ventilatory and circulatory insufficiency. Informing these patients and their families that they are at increased risk of adverse outcome is an important part of the preoperative evaluation. Many neurologists recommend nutritional supplementation with vitamins or other purported antioxidants as well as treatment with various cofactors needed for oxidative metabolism ( appendix 2), although, except for coenzyme Q,137there is a paucity of data supporting their therapeutic value. Patients with mitochondrial cytopathy are usually conditioned not to fast for long durations and to eat small, frequent meals, a routine that conflicts with typical perioperative fasting guidelines. To avoid metabolic crisis, therefore, especially in children, an intravenous infusion of glucose should be initiated preoperatively. Choice of fluids may also be important intraoperatively, most anesthesiologists choosing to avoid the lactate load intrinsic to Ringer’s solution. Monitoring and controlling blood glucose, body temperature, and acid–base values within normal limits is crucial perioperatively, and as with any anesthetic, electrocardiogram, blood pressure, pulse oximetry, temperature, and exhaled gas concentrations should be continuously monitored. In addition, arterial catheterization should be considered to facilitate frequent sampling for blood glucose, arterial blood gases, and serum lactate levels.

Other unique concerns regarding the design of an anesthetic plan for these patients include the pharmacodynamic implications of mitochondrial cytopathy such as decreased anesthetic requirement97and susceptibility to prolonged drug-induced nervous system depression because of impaired neuronal bioenergetics, as well as intrinsic skeletal muscle hypotonia and cardiomyopathy138with increased risk of sudden death from conduction abnormalities.139Bulbar muscle weakness may predispose to aspiration of gastric contents, suggesting the need for “full stomach” precautions. Skeletal muscle weakness may compromise postoperative ventilation, especially after upper abdominal or thoracic surgery.140Subclinical erosion of hepatorenal reserve may alter clinical pharmacokinetics for intravenous drugs and predispose to delayed recovery from anesthetic agents, muscle relaxants, and opioids.141 

Susceptibility to malignant hyperthermia or myasthenia-like sensitivity to neuromuscular blockade are issues typically considered for patients with the more familiar muscular dystrophies and neurogenic myopathies. There is a case report that describes increased sensitivity to nondepolarizing blockade in a patient with mitochondrial myopathy,142but this observation has not been confirmed for most forms of inherited mitochondrial cytopathy.143,144Although there is understandable caution, especially in children, regarding the use of halogenated inhalational anesthetics,145only the very rare mitochondrial myopathies with “multicore” or “minicore” histology seem to warrant concerns of an increased risk of malignant hyperthermia.146Therefore, at least at the present time, there is inadequate data to support the recommendation of some authors that the anesthetic plan for patients with mitochondrial disease should routinely include malignant hyperthermia precautions.147,148 

The residual effects of nondepolarizing agents in these patients, who commonly have compromised hepatic and renal function, may further exacerbate their intrinsic muscle weakness and increase the risk of ventilatory failure postoperatively. In addition, anesthetic techniques requiring spontaneous ventilation may predispose to intraoperative metabolic exhaustion and airway obstruction and therefore should probably be avoided. Tracheal intubation with positive-pressure ventilation will prevent intraoperative ventilatory failure, but the anesthesiologist must decide whether the patient should be extubated immediately after surgery or remain intubated and receive prolonged recovery in an intensive care unit.149 

Although individual patients with inherited mitochondrial encephalomyopathies have been exposed to many different general anesthetic regimens without apparent adverse consequences,150–153it remains unclear whether there is a “safe” or “best” anesthetic plan for these patients. There are few reports that describe the anesthetic treatment of adult-onset or acquired mitochondrial encephalomyopathy154and only one, for example, dealing with NARP syndrome.155Clinical reports often suggest only that patients with mitochondrial disorders “do well” with regional anesthetics despite the facts that these agents, like those used for general anesthesia, are known to disrupt mitochondrial function and bioenergetics. Specialized textbooks offer some further detail regarding preoperative assessment and anesthetic management of patients with mitochondrial cytopathies.156 

Age-related decline in organ system functional reserve is subtle but progressive during the middle adult years. It eventually becomes clearly apparent, even in the most fit older subjects, during the later years of geriatric senescence. Because functional reserve provides the “safety margin” needed to meet the additional cellular and bioenergetic demands imposed by trauma or disease and by surgery, healing, and convalescence, inadequate reserve contributes substantially to perioperative morbidity and mortality in older surgical patients.95Significant limitations in the availability of energy derived from oxidative phosphorylation would impact normal physiologic function and the capacity for physical activity and also compromise the energy supply needed for maintenance of normal tissue structure and cell microarchitecture.157It is now generally appreciated that disruption of oxidative phosphorylation is intrinsically involved in the age-related decline of organ function and functional reserve,158,159although the precise mechanism remains elusive.160 

Many observations are consistent with the hypothesis that failing bioenergetic capacity is central to, if not the cause of, human aging. Acquired mtDNA mutations accumulate at a rate that is a thousandfold greater than that of acquired nDNA mutations.161This may reflect greater exposure of mtDNA to mutagenic factors, more effective endogenous nDNA repair mechanisms, or the diploid nature of nDNA itself. Whatever the cause, the prevalence of mutagenic mtDNA lesions increases exponentially during late adulthood and senescence,162,163primarily in brain,164skeletal muscle,165,166and heart,167,168where defects accumulate more extensively than in rapidly dividing tissues.169It is less clear whether increased ROS levels in the cytosol of a cell can damage nDNA and disrupt synthesis of bioenergetic proteins encoded in the nuclear genome.170The age-related increase in mtDNA defects is also coincident with decrements of cytochrome c  oxidase activity,171although the overall decline in skeletal muscle bioenergetic capacity seen in older adults is largely due to reduced general physical activity and not simply to loss of functional mtDNA.172,173 

However, any loss of functional mtDNA may itself increase oxidative stress174,175and further predispose to oxidative damage of the polypeptides of the respiratory complex.176Coincident age-related decline in the effectiveness of ROS scavenging177,178or age-related compromise of endogenous mtDNA repair systems such as the base excision repair pathway may further accelerate the adverse consequences of accumulated oxidative stress.179,180Data supporting the possibility of a self-perpetuating cycle of impaired or inefficient mitochondrial bioenergetics and a coincident increase in ROS has recently been presented in insect studies.181Dysfunction of ETC complex I, which is a rate-limiting component of aerobic respiration, affects the entire process of oxidative phosphorylation, further decreasing the efficiency of electron transfer and increasing levels of ROS, especially superoxide. A similar process could occur in the aging mammalian cell (fig. 3). Given the high energy requirements of neural tissue, it is likely that neuronal bioenergetics, in particular, decline significantly with increasing age and may eventually provide biomarkers for physiologic aging.182 

Fig. 3. Reactive oxygen species (ROS) are continually generated as byproducts of oxidative metabolism. A self-perpetuating, physiologic “cycle of aging” has been proposed  96,180 in which oxidative stress within mitochondria slowly degrades the components needed for energy production and self-repair of damage done by ROS. The resulting decrease in bioenergetic capacity could eventually compromise organ system functional reserve and predispose to increased probability of adverse perioperative outcome. 

Fig. 3. Reactive oxygen species (ROS) are continually generated as byproducts of oxidative metabolism. A self-perpetuating, physiologic “cycle of aging” has been proposed  96,180 in which oxidative stress within mitochondria slowly degrades the components needed for energy production and self-repair of damage done by ROS. The resulting decrease in bioenergetic capacity could eventually compromise organ system functional reserve and predispose to increased probability of adverse perioperative outcome. 

Close modal

Study of the genetic factors determining human longevity suggest that inheritable factors determine approximately one quarter of the observed variability in life expectancy,183and the importance of mtDNA in this relation is becoming clearer.184At least in subprimates, there are recent data that seem to support the concept that genetic alterations that increase susceptibility to damage by oxidative stress are primarily responsible for increased frailty185and reduced lifespan.186For example, genetically altered mice that express a proof-reading deficient version of mtDNA polymerase develop into a young adult mouse phenotype with a threefold to fivefold increase in mtDNA point mutations and increased mtDNA deletions.187In these mice, the increased number of somatic mtDNA mutations is also associated with phenotypic stigmata of aging such as reduced subcutaneous fat, hair loss, and osteoporosis during young adulthood. It is not yet clear whether this mutant phenotype simply mimics aging rather than prematurely expressing genuine manifestations of physiologic aging, but these mice also have a significantly reduced lifespan.

The extent to which aging and perhaps age-related pathophysiology may reflect impaired mitochondrial bioenergetics and accumulated oxidative stress is only now becoming fully apparent.188Tangential support for this concept comes from observations that reduced calorie intake increases lifespan in some laboratory animals, presumably by reducing accumulated oxidative damage18or by decreasing the rate, or increasing the efficiency, of oxidative phosphorylation,189although life extension through caloric restriction is not a consistent observation.190An effective antiaging therapy based on antioxidant effects has yet to be demonstrated,191,192and genetically manipulated overexpression of superoxide dismutase and catalase actually reduces, rather than prolongs, lifespan in transgenic Drosophila .193It has also not yet been shown that genetic manipulation of intrinsic mtDNA repair and replication mechanisms will produce animals with decreased mtDNA mutation rates and increased longevity, although there is evidence of less mitochondrial oxidative stress in long-lived than in short-lived mammalian species.194Some investigators simply remain unconvinced that there is compelling proof of a significant decline in electron transport or oxidative phosphorylation during normal aging.195 

Because the nervous system has adequate functional redundancy and structural plasticity, aging does not seem to degrade day-to-day neurologic or cognitive function. In addition, the degenerative effects of prolonged oxidative stress in neurons may be ameliorated by effective scavenging of ROS,196accelerated mtDNA repair, increased production of Bcl-2 protein or other apoptosis-inhibiting substances, generation of neurotrophic factors, and mobilization of neural stem cells to replace damaged neurons.197Nevertheless, inadequate mitochondrial energy production and cumulative oxidative stress leading to apoptosis may also explain many of the stigmata of age-related neurodegeneration and some neurologic diseases.198–201 

Presbyacusis, the hearing loss that inevitably occurs in old age, has been shown to be a consequence of the progressive deterioration of cochlear mtDNA.202Interestingly, a specific mtDNA point mutation has been shown definitively to predispose patients to sensorineural hearing loss after aminoglycoside antibiotic exposure.203In general, mtDNA mutations are thought to contribute to or predispose patients to the development of many common neurodegenerative disorders, although they rarely display the same inheritance characteristics as classic inherited mitochondrial cytopathies described previously. Parkinsonism, caused by selective inhibition of complex I of the ETC, critically compromises energy availability and leads to apoptosis and death of the dopaminergic cells in the substantia nigra.204Although a greater fraction of mtDNA is defective in parkinsonian patients than in age-matched controls, there does not seem to be a consistent mtDNA mutation. This suggests that defects in nDNA that lead to dysfunctional complex I bioenergetics, rather than mtDNA mutations, may hold the key to explaining this disorder. Similarly, patients with Alzheimer dementia exhibit higher-than-normal rates of mtDNA mutation, but mtDNA defects are neither consistent nor invariable findings.205Friedrich ataxia is a consequence of ROS-mediated damage to the respiratory chain initiated by an nDNA mutation that eventually compromises mitochondrial iron homeostasis.111Therapy with antioxidants and coenzyme Q may improve mitochondrial function in patients with Friedrich ataxia and slow the progression of symptoms.137 

Neuronal excitotoxicity is a major cause of neuronal death. Excitotoxicity represents a state of greatly increased neuronal electrical and metabolic activity that produces oxidative stress. Even when many neurons are initially destroyed by primary necrosis after traumatic brain injury, excitotoxicity contributes to the additional neuronal damage that follows such an event.206During an excitotoxic event, the intrinsic neuronal mechanisms that scavenge ROS and repair ROS-induced damage are quickly overwhelmed.207High levels of excitatory neurotransmitters such as glutamate also interact with cell-surface N -methyl-d-aspartate receptors (fig. 2) to generate excess calcium within mitochondria. Loss of calcium homeostasis induces an MPT, collapsing the hydrogen ion gradient and releasing of cytochrome c  and other proapoptotic proteins from mitochondria into the cytosol. The culmination of this sequence is apoptosis and neuronal death.

Recent work suggests that genetic mutations predisposing patients to Alzheimer dementia also make neurons susceptible to excitotoxic apoptosis after exposure to certain inhalational anesthetics. Isoflurane, for example, has been shown to induce cytotoxicity in primary cortical neurons under these circumstances.208The proposed mechanism, again, is an influx of ionized calcium from the endoplasmic reticulum into the cytosol, but another potential trigger of neuronal apoptosis is zinc. Elemental zinc is highly concentrated in neurons. Zinc release from damaged cells, even in nanomolar quantities, during traumatic or ischemic brain injury or in Alzheimer dementia or parkinsonism, can cause apoptosis in neighboring neurons and increase the extent of neurologic damage.209 

Similarly, the high levels of ROS that define oxidative stress may also produce secondary neuronal damage beyond the area of initial injury after traumatic brain injury. Endogenous nitric oxide can combine with superoxide to form lipid-destructive peroxynitrite. Lipid peroxidation by peroxynitrite can damage mitochondria and the cellular microarchitecture directly, leading to apoptosis and cell death.210In addition, oxidized lipoproteins can be taken up by neighboring neurons, generating a penumbra, or expanded zone, of neuronal injury.20Limiting oxidative stress by maintaining normoxia during cardiopulmonary resuscitation, as opposed to imposing hyperoxia, has, in fact, been shown to cause less brain lipid peroxidation and improve neurologic outcome, at least in the laboratory.211Increased nitric oxide and peroxynitrite with glutamate-mediated activation of nitric oxide synthase has been proposed as a mechanism for neurodegenerative disorders as well.212Older individuals have been shown to be at increased risk of damage from membrane peroxidation under conditions of oxidative or nitrosative stress,213perhaps explaining, at least in part, the relation between acquired neurodegenerative disorders and age.214 

Caspase-mediated apoptosis is a complex biochemical cascade that requires ATP. Fatally compromised cells, especially those that have undergone a bioenergetic catastrophe, may not be able to produce ATP in amounts adequate to support apoptosis. These neurons may instead undergo passive, or primary, necrosis.215Unlike apoptosis, where cell loss is contained and tissue injury relatively controlled, necrosis of neurons leads to mitochondrial swelling, cell lysis, and fragmentation and the diffuse release of proinflammatory substances that invoke a vigorous immune response.216The interaction between the aging immune system and necrotic neurons may explain the amyloid deposits seen in Alzheimer dementia.217The progressive age-related decline in the specificity of the immune system218and failure to clearly distinguish between “self” and “nonself” may therefore play an important role not only in infection, neoplasm, and autoimmune disorders but also in age-related neurodegenerative disease. Complex interactions between neuronal mitochondrial dysfunction and the mechanisms that control necrosis and apoptosis are now also suspected in playing a key role in amyotrophic lateral sclerosis, hepatolenticular degeneration, and perhaps many other neurodegenerative phenomena.219 

Cardiac mitochondria are essential to myocardial energy production and myocyte homeostasis and also impact cardiac myocyte viability through their role as oxidative biosensors. Cardiac myocytes, skeletal muscle fibers, and other long-lived postmitotic cells show dramatic age-related alterations in the morphology of their mitochondria, with a generalized loss of mitochondrial volume and numbers.220There seems to be an increase in oxidative stress in aging cardiac myocytes, especially with coincident atherosclerotic disease,221and antioxidants such as coenzyme Q may provide some protection against oxidative stress in senescence.222In fact, many drugs used to treat myocardial ischemia seem to exert their cardioprotective effects via  their actions on cardiac mitochondrial function.223Angiotensin-converting enzyme inhibitors have been shown to contribute to enhancement of antioxidant defenses. Some of the beneficial effects associated with inhibition of the renin–angiotensin system may therefore be due to the ability of enalapril and losartan to attenuate oxidative damage to mitochondria.224Angiotensin-converting enzyme inhibitors may also facilitate vascular remodeling.225 

Chronic hypoxia produces a loss of mitochondrial bioenergetic capacity in the left ventricular myocardium despite increases in myocardial mass.226Accumulating evidence also suggests that ROS play an important role in the development and progression of other forms of heart failure227as well as in acute contractile dysfunction after myocardial infarction.228In addition to their direct detrimental effects on cellular metabolic function, ROS have been implicated in the development of agonist-induced cardiac hypertrophy, cardiac myocyte apoptosis, and the subsequent remodeling of the failing myocardium. These restorative alterations are driven by metabolically sensitive gene expression, and in this way, ROS may act as potent intracellular second messengers.228Therefore, the effects of increasing myocardial ROS seem to be either beneficial or harmful, depending on site, source, and amount of ROS produced, and the overall metabolic status of the myocyte.

Oxidative stress seems to contribute to the pathology of vascular disease in stroke, hypertension, and diabetes.229Observations that mitochondrial function is disturbed in the skeletal muscle of patients with occlusive vascular disease230further supports the concept that mitochondrial processes are involved in the etiology of vascular diseases. Nuclear magnetic resonance spectroscopy has shown a 40% reduction of in vivo  muscle glucose metabolism in insulin-resistant older adults,231although it is not yet established that this is part of the fundamental pathophysiology of diabetic vascular pathology. This form of insulin resistance may actually reflect an inherited mitochondrial defect altering fatty acid metabolism.232Taken together, however, these observations regarding the etiology and treatment of cardiovascular disease suggest that the role of mitochondrial dysfunction will assume progressively greater importance as the molecular mechanisms involved in ischemic cardiovascular disease are more completely understood.

Sepsis, the systemic inflammatory response syndrome, and multiple organ dysfunction syndrome are the leading causes of morbidity and mortality in critically ill surgical patients. Acute-onset cardiovascular, hepatic, and renal insufficiency and failure are common features of these syndromes. Inadequate delivery of oxygen to the mitochondria of affected tissues is a possible explanation for tissue or organ dysfunction under these circumstances, but measures that increase cardiac output or tissue perfusion in septic patients have not been of value in improving outcome.233It is now clear that impaired bioenergetic capacity plays an important role in explaining the diffuse and persistent cellular and organ dysfunction that occurs under these circumstances. The concept of “cytopathic hypoxia” proposes that, during sepsis, many cells become unable to use readily available molecular oxygen to produce ATP,234explaining inconsistencies in reported data regarding cellular ATP levels during sepsis. Even with impaired bioenergetic capacity, ATP levels would remain relatively unchanged if there is a parallel reduction both in ATP supply and ATP demand in a hypoxic environment.

There is experimental support for the concept of mitochondrial-based cytopathic hypoxia as a primary factor in sepsis. Data from cardiac myocytes confirm that the mitochondria can act as a modulating biosensor for oxidative phosphorylation, which can send poorly perfused or hypoxic tissues into what is, in effect, a hibernation-like state.235The mechanism remains unclear, but it could include uncoupling of ATP production from aerobic metabolism or inhibition of any or all of the five protein–enzyme complexes required for oxidative phosphorylation.236It may also reflect changes in ETC enzyme kinetics. Abnormalities in pH, temperature, or inhibitor-induced conformational changes in enzyme structure could also disrupt oxidative metabolism and explain the appearance of cytopathic hypoxia during sepsis. Myocardial cytochrome oxidase is reversibly inhibited early in sepsis but seems to become irreversibly inactivated during the later phase of sepsis.237Possible mediators of mitochondrial enzyme inhibition during sepsis include ROS, nitric oxide, peroxynitrite, and carbon monoxide. High levels of nitric oxide reversibly inhibit complex IV.238 

Impaired functioning of any of the enzymes within the ETC is itself associated with decreased cardiac cytochrome oxidase subunit IV and complex II protein levels.239Recent evidence suggests that sepsis induces reduced expression of both of the genes that encode for glycolytic proteins and those needed for the protein components of the ETC.240The synthesis of messenger RNA could be disrupted by abnormalities of either nuclear or mitochondrial transcription, because the subunits of the five respiratory chain enzymes arise from both nDNA and mtDNA. Similarly, errors in protein synthesis due to faulty messenger RNA translation would compromise the electron transport chain and disrupt ATP production. In fact, the messenger RNA that encodes for cytochrome oxidase subunit I is decreased within myocardial cells as well as in macrophages during both sepsis and sepsis-related disorders.241Sepsis is also associated with increased expression of endogenous protective antiapoptotic proteins known as heat shock proteins (HSPs).242,243HSP synthesis can be induced by hypoxia, ROS, endotoxins, or cytokines, all of which are common in sepsis. HSPs may either reconfigure or isolate electron transport chain proteins that have been damaged by the mechanisms described above.244Failure to adequately express HSPs during sepsis or shock may be directly related to propagation of tissue injury and poor outcome,245although a recent clinical study suggests that glutamine-enhanced parenteral nutrition can restore HSPs to protective levels.246 

Hormesis  refers to a state of low-level chronic stress that presumably induces the expression of protective genes that increase host survival during physiologic extremes. Although the general phenomenon of stress-induced expression of genes that facilitate ROS scavenging and mtDNA repair is well established,197hormesis may be more easily initiated in some species or tissues than in others. Cold stress has been shown to prolong lifespan in Caenorhabditis elegans ,247and heat stress significantly increases longevity of the fruit fly.248Brief periods of sublethal ischemia generates low-level oxidative stress that induces an adaptive form of metabolic self-protection, limiting the necrosis and tissue injury that would normally follow a subsequent ischemic injury. Hormesis seems to involve modulation of intracellular ion flux249to minimize the probability of initiating the MPT that can trigger apoptosis (see second paragraph under “Apoptosis”). This phenomenon, ischemic preconditioning (IPC), seems to be initiated largely by the receptor-triggered activation of multiple protein kinases.250 

It is now also established that exposure to volatile anesthetics can generate a state of hormesis in mammalian tissues that mimics IPC and shares many triggers or modulators with IPC.251Even at subclinical concentrations, previous exposure to halothane, isoflurane, sevoflurane, or desflurane252,253has been shown to provide prolonged neuroprotection. Anesthetic preconditioning (APC) has also been demonstrated in the myocardium, where previous exposure to isoflurane, desflurane, and sevoflurane confers protection in vivo  against a subsequent ischemic injury.254–256Although the mechanisms may very somewhat, mitochondrial bioenergetics seem to be significantly affected by all these agents.257After isoflurane exposure, excess ROS are generated at complex III of the ETC and seem to trigger APC.258Sevoflurane, on the other hand, attenuates complex I but also leads to increased ROS production.259Therefore, endogenous oxidative stress seems to be a trigger for APC, a concept supported by the observation that molecular species that scavenge ROS block the APC effect.260Nitrous oxide does not produce APC, but neither does it block nor alter the APC phenomenon associated with the potent inhalational agents.261 

Although the full mechanism of APC is not yet fully understood, it may, like IPC, involve multiple G protein–coupled receptor triggers that activate protein kinases.262APC and IPC also seem to have many other common essential steps, including modulation of ATP-sensitive potassium channels. ATP-sensitive potassium channels are essential for normal endovascular function and responsiveness to vasodilators, and they have also been shown to be important biosensors for excitotoxicity.263They seem to limit ischemic injury both in neurons264and in cardiac muscle.265Opening of myocardial mitochondrial ATP-sensitive potassium channels may also be an intrinsic step in APC after exposure to inhalational anesthetics266or, as recently demonstrated, in response to δ-opioid receptor agonists.267 

Genomic analysis suggests that IPC and APC each reflect a unique pattern of induced gene expression for the synthesis of proapoptotic and antiapoptotic proteins.268APC, if not IPC, may involve inducible nitric oxide synthase in neurons,252whereas a delayed form of APC seen in the myocardium requires induction of endothelial nitric oxide synthase.269It remains unclear how many patterns of protective gene expression are possible, but isoflurane pretreatment may also protect cardiac myocytes against apoptosis by increasing the expression of the antiapoptotic protein Bcl-2.270In addition, it may be possible to use the protective effect of inhalational anesthetics even after severe oxidative stress has occurred. A recent study demonstrated a significant “postconditioning” effect for isoflurane in cardiac muscle, largely through the inhibition of MPTs during the reperfusion of injured cells.271Using inhalational anesthetics or G protein–coupled receptor agonists to protect organs that have been, or may subsequently be, exposed to an ischemic or hypoxic event is an attractive prophylactic and therapeutic option.272However, the practicality and the clinical effectiveness of APC still remain to be established.273 

In mammals, brief, sublethal periods of ischemia and hypoxia also induce the expression of HSPs and block the AIF-mediated apoptotic pathway. Adenovirus-mediated gene therapy has been shown to increase HSP expression and reduce mortality from experimentally induced, sepsis-related pulmonary injury.274The prophylactic use of artificial liposomes and nonviral transfection to deliver HSP or to provide either the DNA or messenger RNA275needed to enhance the synthesis of HSP in neurons or cardiac myocytes is another promising concept that may provide organ protection perioperatively without the need for anesthetic exposure.276Transfection can quickly increase HSP in patients at risk for ischemic or traumatic brain injury, perhaps through an effect on the ATP-sensitive potassium channels in the cerebral vasculature.277Other approaches to minimizing cellular injury during periods of oxidative stress or postinjury reperfusion include enhancement of endogenous expression of cytoprotective antioxidants.278Resveratrol, found in grape skin and in red wine, demonstrably reduces the ischemic damage associated with myocardial and brain reperfusion injury.279At least some of the cytoprotective effect of this substance is due to increased expression of heme oxygenase (HO), the enzyme that accelerates destruction of heme, a pro-oxidant that accumulates rapidly after ischemia and oxidative stress.280HO-deficient diabetic mice have an increased risk of ischemic injury compared with wild-type diabetic mice, suggesting that reduced expression of HO in response to oxidative stress may play a role in the etiology of diabetes-related sequelae.281 

Heme oxygenase pathways may be essential to several other forms of cellular adaptation to stress. Inhalational anesthetic exposure in hepatocytes induces expression of an HO isoform through a pathway that, like APC in heart and brain, involves protein kinases.282Increased endothelial HO activity due to up-regulation during oxidative stress is further potentiated by the interaction of thiols with nitric oxide,283which suggests that HO gene expression may also protect against nitric oxide–related, or nitrosative, stress.284Carbon monoxide, generated by HO as a heme breakdown product, may also act as a signaling mediator of hypoxic stress. It has been shown to provide protection from anoxic injury in nematodes by inducing a state of suspended animation.285Transitional metal carbonyls which expedite the intracellular release of carbon monoxide could therefore have therapeutic value as cardioprotective agents.286 

Intramitochondrial glutathione, normally approximately 15% of total cellular glutathione pool, is another endogenous antioxidant that seems to protect against ROS damage, and depletion of mitochondrial glutathione has been linked to apoptosis.287Similarly, melatonin is a significant scavenger of ROS and an antioxidant.288If administration of exogenous melatonin can decrease tissue damage and dysfunction related to oxidative stress,289it may be useful if given prophylactically for ischemic or reperfusion injury or to minimize the potential neurotoxicity of anesthetic exposure in patients with decreased neural reserve. Given recent interest and investigation into the role of anesthetic agents in causing postoperative cognitive dysfunction in older adults,290,291the clinical observation that melatonin reduces postoperative delirium in this surgical patient group292is particularly intriguing with regard to potential prevention of anesthesia-related cognitive impairment.

Advancements in our understanding of the role of the mitochondrion in generating and responding to oxidative stress have supplemented awareness of its pivotal function as a cell energy source. The central role of the mitochondrion as the final mediator of cell death makes it particularly important to evolving concepts of hypoxic tissue injury and protection as well as to our understanding of senescence and degenerative disease. These manifold mitochondrial functions generate many possible hypotheses that seem to link a wide range of phenomena that are of interest to anesthesiologists from both a clinical and a scientific perspective (fig. 4).

Fig. 4. Altered mitochondrial bioenergetics seems to be central to a variety of physiologic and pathophysiological states that share oxidative stress as a common characteristic. Despite the protective effects (  dashed lines ) of DNA repair mechanisms, endogenous antioxidants, and antiapoptotic substances such as heat shock proteins (HSPs), some of which may be further stimulated by ischemic or anesthetic preconditioning (IPC/APC), oxidative stress eventually disrupts the mitochondrion and triggers rapid mitochondrial and cellular self-destruction (mitoptosis and apoptosis). 

Fig. 4. Altered mitochondrial bioenergetics seems to be central to a variety of physiologic and pathophysiological states that share oxidative stress as a common characteristic. Despite the protective effects (  dashed lines ) of DNA repair mechanisms, endogenous antioxidants, and antiapoptotic substances such as heat shock proteins (HSPs), some of which may be further stimulated by ischemic or anesthetic preconditioning (IPC/APC), oxidative stress eventually disrupts the mitochondrion and triggers rapid mitochondrial and cellular self-destruction (mitoptosis and apoptosis). 

Close modal

At the present time, it seems that all anesthetic agents are associated with measurable effects on some aspect of mitochondrial function, although causal relations are difficult to establish, and the primary effect of these drugs does not seem to reflect simple depression of bioenergetic activity. Much work remains to be done, but previously unrecognized effects of anesthetics on mitochondrial bioenergetics and apoptotic pathways suggest that they may have both cytoprotective and potentially neurotoxic actions, depending on clinical context. It is now also increasingly apparent that there are many subgroups of the surgical patient population that should be considered to be at increased risk perioperatively because of mitochondrial dysfunction, whether it is inherited, acquired, or a consequence of comorbid disease.

Future investigation might appropriately focus on the mitochondrion as the site of anesthetic action and mediator of anesthetic pharmacodynamics, as well as the likely source of potential anesthetic neurotoxicity. Other obvious areas in need of continuing investigation include establishing more precise guidelines for the perioperative treatment of surgical patients with inherited or acquired mitochondrial cytopathies, in all their many and varied manifestations, defining the full spectrum of mitochondrial pathways that contribute to tissue injury, and use of anesthetics to provide perioperative organ protection.

1.
Saraste M: Oxidative phosphorylation at the fin de siecle . Science 1999; 283:1488–93
2.
Hibara SM, Silveira LR, Alberici LC, Leandro CV, Lambertucci RH, Polimeno GC, Cury Boaventura MF, Procopio J, Vercesi AE, Curi R: Acute effect of fatty acids on metabolism and mitochondrial coupling in skeletal muscle. Biochim Biophys Acta 2006; 1757:57–66
3.
Nicholls DG, Locke RM: Thermogenic mechanisms in brown fat. Physiol Rev 1984; 64:1–64
4.
Boss O, Hagen T, Lowell BB: Uncoupling proteins 2 and 3: Potential regulators of mitochondrial energy metabolism. Diabetes 2000; 49:143–56
5.
Fleury C, Neverova M, Collins S, Raimbault S, Champigny O, Levi-Meyrueis C, Bouillaud F, Seldin MF, Surwit RS, Ricquier D, Warden CH: Uncoupling protein-2: A novel gene linked to obesity and hyperinsulinemia. Nat Genet 1997; 15:269–72
6.
Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, Eperon IC, Nierlich DP, Roe BA, Sanger F, Schreier PH, Smith AJ, Staden R, Young IG: Sequence and organization of the human mitochondrial genome. Nature 1981; 290:457–65
7.
Berg OG, Kurland CG: Why mitochondrial genes are most often found in nuclei. Mol Biol Evol 2000; 17:951–61
8.
Scheffler IE: Mitochondria. New York, Wiley-Liss, 1999, p 367
New York
,
Wiley-Liss
9.
McCulloch V, Seidel-Rogol BL, Shadel GS: A human mitochondrial transcription factor is related to RNA adenine methyltransferases and binds S-adenosylmethionine. Mol Cell Biol 2002; 22:1116–25
10.
Schatz G: Mitochondria: Beyond oxidative phosphorylation. Biochim Biophys Acta 1995; 1271:123–6
11.
Beckman KB, Ames BN: Endogenous oxidative damage of mtDNA. Mutat Res 1999; 424:51–8
12.
Simonetti S, Chen X, DiMauro S, Schon EA: Accumulation of deletions in human mitochondrial DNA during normal aging: Analysis by quantitative PCR. Biochim Biophys Acta 1992; 1180:113–22
13.
Calabrese V, Scapagnini G, Giuffrida Stella AM, Bates TE, Clark JB: Mitochondrial involvement in brain function and dysfunction: Relevance to aging, neurodegenerative disorders and longevity. Neurochem Res 2001; 26:739–64
14.
Bohr VA, Stevnsner T, de Souza-Pinto NC: Mitochondrial DNA repair of oxidative damage in mammalian cells. Gene 2002; 286:127–34
15.
Kelso GF, Porteous CM, Hughes G, Ledgerwood EC, Gane AM, Smith RA, Murphy MP: Prevention of mitochondrial oxidative damage using targeted antioxidants. Ann N Y Acad Sci 2002; 959:263–74
16.
Stadtman ER, Berlett BS: Reactive oxygen-mediated protein oxidation in aging and disease. Drug Metab Rev 1998; 30:225–43
17.
Stull LB, Leppo MK, Szweda L, Gao WD, Marban E: Chronic treatment with allopurinol boosts survival and cardiac contractility in murine postischemic cardiomyopathy. Circ Res 2004; 95:1005–11
18.
Zainal TA, Oberley TD, Allison DB, Szweda LI, Weindruch R: Caloric restriction of rhesus monkeys lowers oxidative damage in skeletal muscle. FASEB J 2000; 14:1825–36
19.
Keller JN, Mattson MP: Roles of lipid peroxidation in modulation of cellular signaling pathways, cell dysfunction, and death in the nervous system. Rev Neurosci 1998; 9:105–16
20.
de Grey AD: The reductive hotspot hypothesis of mammalian aging: Membrane metabolism magnifies mutant mitochondrial mischief. Eur J Biochem 2002; 269:2003–9
21.
Lane L: Oxygen: The Molecule That Made the World. New York, Oxford University Press, 2002, p 374
New York
,
Oxford University Press
22.
McCord JM, Fridovich I: Superoxide dismutase: An enzymic function for erythrocuprein (hemocuprein). J Biol Chem 1969; 244:6049–55
23.
Lebovitz RM, Zhang H, Vogel H, Cartwright J Jr, Dionne L, Lu N, Huang S, Matzuk MM: Neurodegeneration, myocardial injury, and perinatal death in mitochondrial superoxide dismutase-deficient mice. Proc Nat Acad Sci U S A 1996; 93:9782–7
24.
Fosslien E: Review: Mitochondrial medicine—cardiomyopathy caused by defective oxidative phosphorylation. Ann Clin Lab Sci 2003; 33:371–95
25.
Klichko VI, Radyuk SN, Orr WC: Profiling catalase gene expression in Drosophila melanogaster  during development and aging. Arch Insect Biochem Physiol 2004; 56:34–50
26.
Hellmich HL, Garcia JM, Shimamura M, Shah SA, Avila MA, Uchida T, Parsley MA, Capra BA, Eidson KA, Kennedy DR, Winston JH, DeWitt DS, Prough DS: Traumatic brain injury and hemorrhagic hypotension suppress neuroprotective gene expression in injured hippocampal neurons. Anesthesiology 2005; 102:806–14
27.
Rhee SG, Chae HZ, Kim K: Peroxiredoxins: A historical overview and speculative preview of novel mechanisms and emerging concepts in cell signaling. Free Radic Biol Med 2005; 38:1543–52
28.
Immenschuh S, Baumgart-Vogt E: Peroxiredoxins, oxidative stress, and cell proliferation. Antioxid Redox Signal 2005; 7:768–77
29.
Fridovich I: Fundamental aspects of reactive oxygen species, or what’s the matter with oxygen? Ann N Y Acad Sci 1999; 893:13–8
30.
Fridovich I: Oxygen toxicity: A radical explanation. J Exp Biol 1998; 201:1203–9
31.
Hagberg H: Mitochondrial impairment in the developing brain after hypoxia-ischemia. J Bioenerg Biomembr 2004; 36:369–73
32.
Song Z, Steller H: Death by design: Mechanism and control of apoptosis. Trends Cell Biol 1999; 9:M49–52
33.
Green DR, Kroemer G: The pathophysiology of mitochondrial cell death. Science 2004; 305:626–9
34.
Zhivotovsky B, Kroemer G: Apoptosis and genomic instability. Nat Rev Mol Cell Biol 2004; 5:752–62
35.
Brenner C, Marzo I, Kroemer G: A revolution in apoptosis: From a nucleocentric to a mitochondriocentric perspective. Exp Gerontol 1998; 33:543–53
36.
Kroemer G, Dallaporta B, Resche-Rigon M: The mitochondrial death/life regulator in apoptosis and necrosis. Annu Rev Physiol 1998; 60:619–42
37.
Schultz DR, Harrington WJ Jr: Apoptosis: Programmed cell death at a molecular level. Semin Arthritis Rheum 2003; 32:345–69
38.
Battle TE, Frank DA: The role of STATs in apoptosis. Curr Mol Med 2002; 2:381–92
39.
Vahsen N, Cande C, Briere JJ, Benit P, Joza N, Larochette N, Mastroberardino PG, Pequignot MO, Casares N, Lazar V, Feraud O, Debili N, Wissing S, Engelhardt S, Madeo F, Piacentini M, Penninger JM, Schagger H, Rustin P, Kroemer G: AIF deficiency compromises oxidative phosphorylation. EMBO J 2004; 23:4679–89
40.
Gunter TE, Gunter KK, Sheu SS, Gavin CE: Mitochondrial calcium transport: Physiological and pathological relevance. Am J Physiol 1994; 267:C313–39
41.
Gunter TE, Yule DI, Gunter KK, Eliseev RA, Salter JD: Calcium and mitochondria. FEBS Lett 2004; 567:96–102
42.
Kroemer G, Dallaporta B, Resche-Rigon M: The mitochondrial death/life regulator in apoptosis and necrosis. Ann Rev Physiol 1998; 60:619–42
43.
Hirsch T, Susin SA, Marzo I, Marchetti P, Zamzami N, Kroemer G: Mitochondrial permeability transition in apoptosis and necrosis. Cell Biol Toxicol 1998; 14:141–5
44.
Polster BM, Fiskum G: Mitochondrial mechanisms of neural cell apoptosis. J Neurochem 2004; 90:1281–9
45.
Saris N-EL, Carafoli E: A historical review of cellular calcium handling, with emphasis on mitochondria. Biochemistry (Mosc) 2005; 70:187–94
46.
Brunori M, Giuffre A, Sarti P, Stubauer G, Wilson MT: Nitric oxide and cellular respiration. Cell Mol Life Sci 1999; 56:549–57
47.
Boyd CS, Cadenas E: Nitric oxide and cell signaling pathways in mitochondrial-dependent apoptosis. Biol Chem 2002; 383:411–23
48.
Garrido C, Kroemer G: Life’s smile, death’s grin: Vital functions of apoptosis-executing proteins. Curr Opin Cell Biol 2004; 16:639–46
49.
Webb JL, Elliot KAC: Effects of narcotics and convulsants on tissue glycolysis and respiration. J Pharmacol Exp Ther 1951; 103:24–34
50.
Quastel JH, Wheatley AHM: Narcosis and oxidations of the brain. Proc R Soc Lond 1933; 112:60–79
51.
Mastronicola D, Arcuri E, Arese M, Bacchi A, Mercadante S, Cardelli P, Citro G, Sarti P: Morphine but not fentanyl and methadone affects mitochondrial membrane potential by inducing nitric oxide release in glioma cells. Cell Mol Life Sci 2004; 61:2991–7
52.
Marian M, Parrino C, Leo AM, Vincenti E, Bindoli A, Scutari G: Effect of the intravenous anesthetic 2,6-diisopropylphenol on respiration and energy production by rat brain synaptosomes. Neurochem Res 1997; 22:287–92
53.
Stevanato R, Momo F, Marian M, Rigobello MP, Bindoli A, Bragadin M, Vincenti E, Scutari G: Effects of nitrosopropofol on mitochondrial energy-converting system. Biochem Pharmacol 2002; 64:1133–8
54.
Schenkman KA, Yan S: Propofol impairment of mitochondrial respiration in isolated perfused guinea pig hearts determined by reflectance spectroscopy. Crit Care Med 2000; 28:172–7
55.
Rigoulet M, Devin A, Averet N, Vandais B, Guerin B: Mechanisms of inhibition and uncoupling of respiration in isolated rat liver mitochondria by the general anesthetic 2,6-diisopropylphenol. Eur J Biochem 1996; 241:280–5
56.
Acco A, Comar JF, Bracht A: Metabolic effects of propofol in the isolated perfused rat liver. Basic Clin Pharmacol Toxicol 2004; 95:166–74
57.
Cohen PJ: Effect of anesthetics on mitochondrial function. Anesthesiology 1973; 39:153–64
58.
Jowett M, Quastel JH: The effects of ether on brain oxidation. Biochem J 1937; 31:1101–12
59.
Hall GM, Kirtland SJ, Baum H: The inhibition of mitochondrial respiration by inhalational anaesthetic agents. Br J Anaesth 1973; 45:1005–9
60.
Brunner EA, Cheng SC, Berman ML: Effects of anesthesia on intermediary metabolism. Annu Rev Med 1975; 26:391–401
61.
Lee SL, Alto LE, Dhalla NS: Subcellular effects of some anesthetic agents on rat myocardium. Can J Physiol Pharmacol 1979; 57:65–71
62.
Einarsdottir O, Caughey WS: Interactions of the anesthetic nitrous oxide with bovine heart cytochrome c oxidase: Effects on protein structure, oxidase activity, and other properties. J Biol Chem 1988; 263:9199–205
63.
Hanley PJ, Ray J, Brandt U, Daut J: Halothane, isoflurane and sevoflurane inhibit NADH:ubiquinone oxidoreductase (complex I) of cardiac mitochondria. J Physiol 2002; 544:687–93
64.
Vincenti E, Branca D, Varotto ML, Scutari G: Effects of halothane, enflurane and isoflurane on some energy-converting functions of isolated rat liver mitochondria. Agressologie 1989; 30:517–20
65.
Nahrwold ML, Cohen PJ: Additive effect of nitrous oxide and halothane on mitochondrial function. Anesthesiology 1973; 39:534–6
66.
Becker GL, Pelligrino DA, Miletich DJ, Albrecht RF: The effects of nitrous oxide on oxygen consumption by isolated cerebral cortex mitochondria. Anesth Analg 1986; 65:355–9
67.
Garlid KD, Nakashima RA: Studies on the mechanism of uncoupling by amine local anesthetics: Evidence for mitochondrial proton transport mediated by lipophilic ion pairs. J Biol Chem 1983; 258:7974–80
68.
Dabbeni-Sala F, Schiavo G, Palatini P: Mechanism of local anesthetic effect on mitochondrial ATP synthase as deduced from photolabelling and inhibition studies with phenothiazine derivatives. Biochim Biophys Acta 1990; 1026:117–25
69.
Tarba C, Cracium C: A comparative study of the effects of procaine, lidocaine, tetracaine, and dibucaine on the functions and ultrastructure of isolated rat liver mitochondria. Biochem Biophys Acta 1990; 1019:19–28
70.
Floridi A, Barbieri R, Pulselli R, Fanciulli M, Arcuri E: Effect of the local anesthetic bupivacaine on the energy metabolism of Ehrlich ascites tumor cells. Oncol Res 1994; 6:593–601
71.
Levy RJ, Piel DA, Acton PD, Zhou R, Ferrari VA, Karp JS, Deutschman CS: Evidence of myocardial hibernation in the septic heart. Crit Care Med 2005; 33:2752–6
72.
Levitt JD: The biochemical basis of anesthetic toxicity. Surg Clin North Am 1975; 55:801–18
73.
Ueda I, Kamaya H: Molecular mechanisms of anesthesia. Anesth Analg 1984; 63:929–45
74.
Mets B, Janicki PK, James MF, Hickman R: Hepatic energy charge and adenine nucleotide status in rats anesthetized with halothane, isoflurane or enflurane. Acta Anaesthesiol Scand 1997; 41:252–5
75.
Miller KW: The nature of sites of general anaesthetic action. Br J Anaesth 2002; 89:17–31
76.
Branca D, Vincenti E, Scutari G: Influence of the anesthetic 2,6-diisopropylphenol (propofol) on isolated rat heart mitochondria. Comp Biochem Physiol 1995; 110:41–5
77.
Kissin I, Aultman DF, Smith LR: Effects of volatile anesthetics on myocardial oxidation-reduction status assessed by NADH fluorometry. Anesthesiology 1983; 59:447–52
78.
Hanley PJ, ter Keurs HEDJ, Cannell MB: Excitation–contraction coupling of the heart and the negative inotropic action of volatile anesthetics. Anesthesiology 2004; 101:999–1014
79.
Johnson ME, Uhl C, Spittler K-H, Wang H, Gores G: Mitochondrial injury and caspase activation by the local anesthetic lidocaine. Anesthesiology 2004; 101:1184–94
80.
Irwin W, Fontaine E, Agnolucci L, Penzo D, Betto R, Bortolotto S, Reggiani C, Salviati G, Bernardi P: Bupivacaine myotoxicity is mediated by mitochondria. J Biol Chem 2002; 277:1221–7
81.
Wei H, Kang B, Wei W, Liang G, Meng QC, Li Y, Eckenhoff RG: Isoflurane and sevoflurane affect cell survival and BCL-2/BAX ratio differently. Brain Res 2005; 1037:139–47
82.
Hofeler H, Krieglstein J: Solubilization of hexokinase activity by an effect of thiopental on the mitochondrial membrane. Pharmacology 1982; 24:156–61
83.
Curatola G, Mazzanti L, Grilli G, Lenaz G: Molecular mechanism of general anesthesia: I. Fluorescence studies in mitochondrial membranes. Boll Soc Ital Biol Sper 1979; 55:499–505
84.
Lenaz G, Curatola G, Mazzanti L, Parenti-Castelli G, Landi L, Sechi AM: A conformational model of the action of general anesthetics at the membrane level: III. Anesthetics and the properties of membrane-bound enzymes: Mitochondrial ATPase. Ital J Biochem 1978; 27:431–49
85.
Xi J, Liu R, Asbury GR, Eckenhoff MF, Eckenhoff RG: Inhalational anesthetic-binding proteins in rat neuronal membranes. J Biol Chem 2004; 279:19628–33
86.
Mazzanti L, Curatola G, Zolese G, Bertoli E, Lenaz G: Lipid protein interactions in mitochondria: VIII. Effect of general anesthetics on the mobility of spin labels in lipid vesicles and mitochondrial membranes. J Bioenerg Biomembr 1979; 11:17–32
87.
Eckenhoff RG, Johansson JS: Molecular interactions between inhaled anesthetics and proteins. Pharmacol Rev 1997; 49:343–67
88.
Vanderkooi G, Chazotte B: Cytochrome c oxidase inhibition by anesthetics: Thermodynamic analysis. Proc Nat Acad Sci U S A 1982; 79:3749–53
89.
Casanovas AM, Malmary Nebot MF, Courriere P, Oustrin J: Inhibition of cytochrome oxidase activity by local anaesthetics. Biochem Pharmacol 1983; 32:2715–9
90.
Blackstone E, Morrison M, Roth MB: H2S induces a suspended animation-like state in mice. Science 2005; 308:518
91.
Kayser EB, Morgan PG, Sedensky MM: GAS-1: A mitochondrial protein controls sensitivity to volatile anesthetics in the nematode Caenorhabditis elegans . Anesthesiology 1999; 90:545–54
92.
Hartman PS, Ishii N, Kayser EB, Morgan PG, Sedensky MM: Mitochondrial mutations differentially affect aging, mutability and anesthetic sensitivity in Caenorhabditis elegans . Mech Ageing Dev 2001; 122:1187–201
93.
Kayser EB, Sedensky MM, Morgan PG: The effects of complex I function and oxidative damage on lifespan and anesthetic sensitivity in Caenorhabditis elegans . Mech Ageing Dev 2004; 125:455–64
94.
Kayser EB, Morgan PG, Sedensky MM: Mitochondrial complex I function affects halothane sensitivity in Caenorhabditis elegans . Anesthesiology 2004; 101:365–72
95.
Muravchick S: Geroanesthesia: Principles for Management of the Elderly Patient. St. Louis, Mosby, 1997, p 306
St. Louis
,
Mosby
96.
Ozawa T: Genetic and functional changes in mitochondria associated with aging. Physiol Rev 1997; 77:425–64
97.
Morgan PG, Hoppel CL, Sedensky MM: Mitochondrial defects and anesthetic sensitivity. Anesthesiology 2002; 96:1268–70
98.
Olney JW, Wozniak DF, Jevtovic-Todorovic V, Farber NB, Bittigau P, Ikonomidou C: Drug-induced apoptotic neurodegeneration in the developing brain. Brain Pathol 2002; 12:488–98
99.
Jevtovic-Todorovic V, Hartman RE, Izumi Y, Benshoff ND, Dikranian K, Zorumski CF, Olney JW, Wozniak DF: Early exposure to common anesthetic agents causes widespread neurodegeneration in the developing rat brain and persistent learning deficits. J Neurosci 2003; 23:876–82
100.
Jevtovic-Todorovic V, Beals J, Benshoff N, Olney JW: Prolonged exposure to inhalational anesthetic nitrous oxide kills neurons in adult rat brain. Neuroscience 2003; 122:609–16
101.
Culley DJ, Baxter MG, Yukhananov R, Crosby G: Long-term impairment of acquisition of a spatial memory task following isoflurane-nitrous oxide anesthesia in rats. Anesthesiology 2004; 100:309–14
102.
Crosby C, Culley DJ, Baxter MG, Yukhananov R, Crosby G: Spatial memory performance 2 weeks after general anesthesia in adult rats. Anesth Analg 2005; 101:1389–92
103.
Delogu G, Antonucci A, Moretti S, Marandola M, Tellan G, Signore M, Famularo G: Oxidative stress and mitochondrial glutathione in human lymphocytes exposed to clinically relevant anesthetic drug concentrations. J Clin Anesth 2004; 16:189–94
104.
Monk TG, Saini Y, Weldon BC, Sigl JC: Anesthetic management and one-year mortality after noncardiac surgery. Anesth Analg 2005; 100:4–10
105.
Lennmarken C, Lindholm M-L, Greenwald S, Sandin R: Confirmation that low intraoperative BIS™ levels predict increased risk of post-operative mortality (abstract). Anesthesiology 2003; 99:A303
106.
Jevtovic-Todorovic V, Carter LB: The anesthetics nitrous oxide and ketamine are more neurotoxic to old than to young rat brain. Neurobiol Aging 2005; 26:947–56
107.
Zhang JG, Tirmenstein MA, Nicholls-Grzemski FA, Fariss MW: Mitochondrial electron transport inhibitors cause lipid peroxidation-dependent and -independent cell death: Protective role of antioxidants. Arch Biochem Biophys 2001; 393:87–96
108.
Todd MM: Anesthetic neurotoxicity: The collision between laboratory neuroscience and clinical medicine. Anesthesiology 2004; 101:272–3
109.
Wallace DC: Mitochondrial diseases in man and mouse. Science 1999; 283:1482–8
110.
DiMauro S, Schon EA: Mitochondrial respiratory-chain diseases. N Engl J Med 2003; 348:2656–68
111.
Schapira AH: Primary and secondary defects of the mitochondrial respiratory chain. J Inherit Metab Dis 2002; 25:207–14
112.
Treem WR, Sokol RJ: Disorders of the mitochondria. Semin Liver Dis 1998; 18:237–53
113.
Holt IJ, Harding AE, Morgan-Hughes JA: Deletions of muscle mitochondrial DNA in patients with mitochondrial myopathies. Nature 1988; 331:717–9
114.
Walker UA, Collins S, Byrne E: Respiratory chain encephalomyopathies: A diagnostic classification. Eur Neurol 1996; 36:260–7
115.
Nagy IZ: On the true role of oxygen free radicals in the living state, aging, and degenerative disorders. Ann N Y Acad Sci 2001; 928:187–99
116.
Wallace DC, Zheng XX, Lott MT, Shoffner JM, Hodge JA, Kelley RI, Epstein CM, Hopkins LC: Familial mitochondrial encephalomyopathy (MERRF): Genetic, pathophysiological, and biochemical characterization of a mitochondrial DNA disease. Cell 1988; 55:601–10
117.
Griggs RC, Karpati G: Muscle pain, fatigue, and mitochondriopathies. N Engl J Med 1999; 341:1077–8
118.
Johnston W, Karpati G, Carpenter S, Arnold D, Shoubridge EA: Late-onset mitochondrial myopathy. Ann Neurol 1995; 37:16–23
119.
Wallace DC, Singh G, Lott MT, Hodge JA, Schurr TG, Lezza AM, Elsas LJ II, Nikoskelainen EK: Mitochondrial DNA mutation associated with Leber’s hereditary optic neuropathy. Science 1988; 242:1427–30
120.
Wallace DC, Lott MT, Shoffner JM, Brown MD: Diseases resulting from mitochondrial DNA point mutations. J Inherit Metab Dis 1992; 15:472–9
121.
Pearson HA, Lobel JS, Kocoshis SA, Naiman JL, Windmiller J, Lammi AT, Hoffman R, Marsh JC: A new syndrome of refractory sideroblastic anemia with vacuolization of marrow precursors and exocrine pancreatic dysfunction. J Pediatr 1979; 95:976–84
122.
Kearns TP, Sayres GP: Retinitis pigmentosa, external ophthalmoplegia, and complete heart block: Unusual syndrome with histologic study in one of two cases. Arch Ophthalmol 1958; 60:280–9
123.
Graeber MB, Muller U: Recent developments in the molecular genetics of mitochondrial disorders. J Neurol Sci 1998; 153:251–63
124.
Wei YH: Mitochondrial DNA mutations and oxidative damage in aging and diseases: An emerging paradigm of gerontology and medicine. Proc Nat Sci Counc Repub China B 1998; 22:55–67
125.
Suomalainen A, Kaukonen J: Diseases caused by nuclear genes affecting mtDNA stability. Am J Med Genet 2001; 106:53–61
126.
Zeviani M, Spinazzola A, Carelli V: Nuclear genes in mitochondrial disorders. Curr Opin Genet Devel 2003; 13:262–70
127.
Moraes CT, Shanske S, Tritschler HJ, Aprille JR, Andreetta F, Bonilla E, Schon EA, DiMauro S: mtDNA depletion with variable tissue expression: A novel genetic abnormality in mitochondrial diseases. Am J Hum Genet 1991; 48:492–501
128.
Treem WR, Sokol RJ: Disorders of the mitochondria. Semin Liver Dis 1998; 18:237–53
129.
Suomalainen A, Kaukonen J: Diseases caused by nuclear genes affecting mtDNA stability. Am J Med Genet 2001; 106:53–61
130.
Cros D, Palliyath S, DiMauro S, Ramirez C, Shamsnia M, Wizer B: Respiratory failure revealing mitochondrial myopathy in adults. Chest 1992; 101:824–8
131.
Hartley C, Ascott F: Laryngeal involvement in mitochondrial myopathy. J Laryngol Otol 1994; 108:685–7
132.
Weber K, Wilson JN, Taylor L, Brierley E, Johnson MA, Turnbull DM, Bindoff LA: A new mtDNA mutation showing accumulation with time and restriction to skeletal muscle. Am J Hum Genet 1997; 60:373–80
133.
Taivassalo T, Jensen TD, Kennaway N, DiMauro S, Vissing J, Haller RG: The spectrum of exercise tolerance in mitochondrial myopathies: A study of 40 patients. Brain 2003; 126(pt 2):413–23
134.
Taivassalo T, Abbott A, Wyrick P, Haller RG: Venous oxygen levels during aerobic forearm exercise: An index of impaired oxidative metabolism in mitochondrial myopathy. Ann Neurol 2002; 51:38–44
135.
Zupanc ML, Moraes CT, Shanske S, Langman CB, Ciafaloni E, DiMauro S: Deletion of mitochondrial DNA in patients with combined features of Kearns-Sayre and MELAS syndromes. Ann Neurol 1991; 29:680–3
136.
Zeviani M, Bertagnolio B, Uziel G: Neurological presentations of mitochondrial diseases. J Inherit Metab Dis 1996; 19:504–20
137.
Hart PE, Lodi R, Rajagopalan B, Bradley JL, Crilley JG, Turner C, Blamire AM, Manners D, Styles P, Schapira AH, Cooper JM: Antioxidant treatment of patients with Friedreich ataxia: Four-year follow-up. Arch Neurol 2005; 62:621–6
138.
Antozzi C, Zeviani M: Cardiomyopathies in disorders of oxidative metabolism. Cardiovasc Res 1997; 35:184–99
139.
Lauwers MR, Van Lersberghe C, Camu F: Inhalation anaesthesia and the Kearns-Sayre syndrome. Anaesthesia 1994; 49:876–8
140.
Grattan-Smith PJ, Shield LK, Hopkins IJ, Collins KJ: Acute respiratory failure precipitated by general anesthesia in Leigh’s syndrome. J Child Neurol 1990; 5:137–41
141.
Schmiedel J, Jackson S, Schafer J, Reichmann H: Mitochondrial cytopathies. J Neurol 2003; 250:267–77
142.
Naguib M, el Dawlatly AA, Ashour M, al-Bunyan M: Sensitivity to mivacurium in a patient with mitochondrial myopathy. Anesthesiology 1996; 84:1506–9
143.
Wiesel S, Bevan JC, Samuel J, Donati F: Vecuronium neuromuscular blockade in a child with mitochondrial myopathy. Anesth Analg 1991; 72:696–9
144.
D’Ambra MN, Dedrick D, Savarese JJ: Kearns-Sayre syndrome and pancuronium-succinylcholine induced neuromuscular blockade. Anesthesiology 1979; 51:343–5
145.
Farag E, Argalious M, Narouze S, DeBoer GE, Tome J: The anesthetic management of ventricular septal defect (VSD) repair in a child with mitochondrial cytopathy. Can J Anaesth 2002; 49:958–62
146.
Figarella-Branger D, Kozak-Ribbens G, Rodet L, Aubert M, Borsarelli J, Cozzone PJ, Pellissier JF: Pathological findings in 165 patients explored for malignant hyperthermia susceptibility. Neuromuscul Disord 1993; 3:553–6
147.
Itaya K, Takahata O, Mamiya K, Saito T, Tamakawa S, Akama Y, Kubota M, Ogawa H: Anesthetic management of two patients with mitochondrial encephalopathy, lactic acidosis and stroke-like episodes (MELAS). Masui 1995; 44:710–2
148.
Cheam EW, Critchley LA: Anesthesia for a child with complex I respiratory chain enzyme deficiency. J Clin Anesth 1998; 10:524–7
149.
Shipton EA, Prosser DO: Mitochondrial myopathies and anaesthesia. Eur J Anaesthesiol 2004; 21:173–8
150.
Maslow A, Lisbon A: Anesthetic considerations in patients with mitochondrial dysfunction. Anesth Analg 1993; 76:884–6
151.
Matsuno S, Hashimoto H, Matsuki A: Neuroleptanesthesia for a patient with mitochondrial encephalomyopathy. Masui 1994; 43:1038–4
152.
Wallace JJ, Perndt H, Skinner M: Anaesthesia and mitochondrial disease. Paediatr Anaesth 1998; 8:249–54
153.
Sharma AD, Erb T, Schulman SR, Sreeram G, Slaughter TF: Anaesthetic considerations for a child with combined Prader-Willi syndrome and mitochondrial myopathy. Paediatr Anaesth 2001; 11:488–90
154.
Sabate S, Ferrandiz M, Paniagua P, Villamor JM, Vilanova F Villar-Landeira JM: Anesthesia in Kearns-Sayre syndrome mitochondrial myopathy [in Spanish]. Rev Esp Anestesiol Reanim 1996; 43:255–7
155.
Ciccotelli KK, Prak EL, Muravchick S: An adult with inherited mitochondrial encephalomyopathy: Report of a case. Anesthesiology 1997; 87:1240–2
156.
Levy R, Muravchick S: Mitochondrial diseases, Anesthesia and Uncommon Diseases, 5th edition. Edited by Fleisher LA. Philadelphia, Elsevier, 2005, pp 455–67Fleisher LA
Philadelphia
,
Elsevier
157.
Gross NJ, Getz GS, Rabinowitz M: Apparent turnover of mitochondrial deoxyribonucleic acid and mitochondrial phospholipids in the tissues of the rat. J Biol Chem 1969; 244:1552–62
158.
Wallace DC: Diseases of the mitochondrial DNA. Annu Rev Biochem 1992; 61:1175–212
159.
Speakman JR, Selman C, McLaren JS, Harper EJ: Living fast, dying when? The link between aging and energetics. J Nutr 2002; 132:1583S–97S
160.
Huang H, Manton KG: The role of oxidative damage in mitochondria during aging: A review. Front Biosci 2004; 9:1100–17
161.
Cortopassi G, Liu Y, Hutchin T: Degeneration of human oncogenes and mitochondrial genes occurs in cells that exhibit age-related pathology. Exp Gerontol 1996; 31:253–65
162.
Linnane AW, Marzuki S, Ozawa T, Tanaka M: Mitochondrial DNA mutations as an important contributor to ageing and degenerative diseases. Lancet 1989; 1:642–51
163.
Cottrell DA, Blakely EL, Borthwick GM, Johnson MA, Taylor GA, Brierley EJ, Ince PG, Turnbull DM: Role of mitochondrial DNA mutations in disease and aging. Ann N Y Acad Sci 2000; 908:199–207
164.
Corral-Debrinski M, Horton T, Lott MT, Shoffner JM, Beal MF, Wallace DC: Mitochondrial DNA deletions in human brain: Regional variability and increase with advanced age. Nat Genet 1992; 2:324–9
165.
Hsieh RH, Hou JH, Hsu HS, Wei YH: Age-dependent respiratory function decline and DNA deletions in human muscle mitochondria. Biochem Mol Biol Int 1994; 32:1009–22
166.
Melov S, Shoffner JM, Kaufman A, Wallace DC: Marked increase in the number and variety of mitochondrial DNA rearrangements in aging human skeletal muscle. Nucleic Acids Res 1995; 23:4122–6
167.
Sugiyama S, Hattori K, Hayakawa M, Ozawa T: Quantitative analysis of age-associated accumulation of mitochondrial DNA with deletion in human hearts. Biochem Biophys Res 1991; 180:894–9
168.
Hattori K, Tanaka M, Sugiyama S, Obayashi T, Ito T, Satake T, Hanaki Y, Asai J, Nagano M, Ozawa T: Age-dependent decrease in mitochondrial DNA in the human heart: Possible contributory factor to presbycardia. Am Heart J 1991; 121:1735–42
169.
Kowald A, Kirkwood TB: Accumulation of defective mitochondria through delayed degradation of damaged organelles and its possible role in the ageing of post-mitotic and dividing cells. J Theor Biol 2000; 202:145–60
170.
Orgel LE: The maintenance of the accuracy of protein synthesis and its relevance to ageing: A correction. Proc Nat Acad Sci U S A 1970; 67:1476
171.
Brierley EJ, Johnson MA, Lightowlers RN, James OF, Turnbull DM: Role of mitochondrial DNA mutations in human aging: Implications for the central nervous system and muscle. Ann Neurol 1998; 43:217–23
172.
Brierley EJ, Johnson MA, James OF, Turnbull DM: Effects of physical activity and age on mitochondrial function. Q J Med 1996; 89:251–8
173.
Brierley EJ, Johnson MA, James OF, Turnbull DM: Mitochondrial involvement in the ageing process: Facts and controversies. Mol Cell Biochem 1997; 174:325–8
174.
Knight JA: The biochemistry of aging. Adv Clin Chem 2000; 35:1–62
175.
Genova ML, Pich MM, Bernacchia A, Bianchi C, Biondi A, Bovina C, Falasca AI, Formiggini G, Castelli GP, Lenaz G: The mitochondrial production of reactive oxygen species in relation to aging and pathology. Ann N Y Acad Sci 2004; 1011:86–100
176.
Wei YH, Lee HC: Oxidative stress, mitochondrial DNA mutation, and impairment of antioxidant enzymes in aging. Exp Biol Med 2002; 227:671–82
177.
Vanella A, Geremia E, D’Urso G, Tiriolo P, DiSilvestro I, Grimaldi R, Pinturo R: Superoxide dismutase activities in aging rat brain. Gerontology 1982; 28:108–13
178.
Barja G: Endogenous oxidative stress: Relationship to aging, longevity and caloric restriction. Ageing Res Rev 2002; 1:397–411
179.
Bohr VA, Stevnsner T, de Souza-Pinto NC: Mitochondrial DNA repair of oxidative damage in mammalian cells. Gene 2002; 286:127–34
180.
Mandavilli BS, Santos JH, Van Houten B: Mitochondrial DNA repair and aging. Mutat Res 2002; 509:127–51
181.
Ferguson M, Mockett RJ, Shen Y, Orr WC, Sohal RS: Age-associated decline in mitochondrial respiration and electron transport in Drosophila melanogaster.  Biochem J 2005; 390:501–11
182.
Navarro A: Mitochondrial enzyme activities as biochemical markers of aging. Mol Aspects Med 2004; 25:37–48
183.
De Benedictis G, Tan Q, Jeune B, Christensen K, Ukraintseva SV, Bonafe M, Franceschi C, Vaupel JW, Yashin AI: Recent advances in human gene-longevity association studies. Mech Ageing Dev 2001; 122:909–20
184.
Rose G, Passarino G, Franceschi C, De Benedictis G: The variability of the mitochondrial genome in human aging: A key for life and death? Int J Biochem Cell Biol 2002; 34:1449–60
185.
Semenchenko GV, Khazaeli AA, Curtsinger JW, Yashin AI: Stress resistance declines with age: Analysis of data from a survival experiment with Drosophila melanogaster . Biogerontology 2004; 5:17–30
186.
Kayser EB, Sedensky MM, Morgan PG: The effects of complex I function and oxidative damage on lifespan and anesthetic sensitivity in Caenorhabditis elegans . Mech Ageing Dev 2004; 125:455–64
187.
Trifunovic A, Wredenberg A, Falkenberg M, Spelbrink JN, Rovio AT, Bruder CE, Bohlooly YM, Gidlof S, Oldfors A, Wibom R, Tornell J, Jacobs HT, Larsson N-G: Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature 2004; 429:417–23
188.
Goyns MH: Genes, telomeres and mammalian ageing. Mech Ageing Dev 2002; 123:791–9
189.
Wickens AP: Ageing and the free radical theory. Resp Physiol 2001; 128:379–91
190.
Cooper TM, Mockett RJ, Sohal BH, Sohal RS, Orr WC: Effect of caloric restriction on life span of the housefly, Musca domestica. FASEB J 2004; 18:1591–3
191.
Miquel J: Can antioxidant diet supplementation protect against age-related mitochondrial damage? Ann N Y Acad Sci 2002; 959:508–16
192.
Rafique R, Schapira AH, Coper JM: Mitochondrial respiratory chain dysfunction in ageing: Influence of vitamin E deficiency. Free Radic Res 2004; 38:157–65
193.
Bayne AC, Mockett RJ, Orr WC, Sohal RS: Enhanced catabolism of mitochondrial superoxide/hydrogen peroxide and aging in transgenic Drosophila  Biochem J 2005; 391:277–84
194.
Kapahi P, Boulton ME, Kirkwood TB: Positive correlation between mammalian life span and cellular resistance to stress. Free Radic Biol Med 1999; 26:495–500
195.
Maklashina E, Ackrell BA: Is defective electron transport at the hub of aging? Aging Cell 2004; 3:21–7
196.
Lebovitz RM, Zhang H, Vogel H, Cartwright J Jr, Dionne L, Lu N, Huang S, Matzuk MM: Neurodegeneration, myocardial injury, and perinatal death in mitochondrial superoxide dismutase-deficient mice. Proc Nat Acad Sci U S A 1996; 93:9782–7
197.
Mattson MP, Duan W, Chan SL, Cheng A, Haughey N, Gary DS, Guo Z, Lee J, Furukawa K: Neuroprotective and neurorestorative signal transduction mechanisms in brain aging: Modification by genes, diet and behavior. Neurobiol Aging 2002; 23:695–705
198.
Ihara Y, Kibata M, Hayabara T, Katayama S, Konishi H, Miura K, Kohno M, Kawai M, Yokoi I, Mori A: Free radicals in the cerebrospinal fluid are associated with neurological disorders including mitochondrial encephalomyopathy. Biochem Mol Biol Int 1997; 42:937–47
199.
Calabrese V, Scapagnini G, Giuffrida Stella AM, Bates TE, Clark JB: Mitochondrial involvement in brain function and dysfunction: Relevance to aging, neurodegenerative disorders and longevity. Neurochem Res 2001; 26:739–64
200.
Jordan J, Cena V, Prehn JH: Mitochondrial control of neuron death and its role in neurodegenerative disorders. J Physiol Biochem 2003; 59:129–41
201.
Ames BN: Mitochondrial decay, a major cause of aging, can be delayed. J Alzheimers Dis 2004; 6:117–21
202.
Pickles JO: Mutation in mitochondrial DNA as a cause of presbyacusis. Audiol Neurootol 2004; 9:23–33
203.
Hutchin T: Sensorineural hearing loss and the 1555G mitochondrial DNA mutation. Acta Otolaryngol 1999; 119:48–52
204.
Schapira AH, Gu M, Taanman JW, Tabrizi SJ, Seaton T, Cleeter M, Cooper JM: Mitochondria in the etiology and pathogenesis of Parkinson’s disease. Ann Neurol 1998; 44:S89–98
205.
Hutchin T, Cortopassi G: A mitochondrial DNA clone is associated with increased risk for Alzheimer disease. Proc Nat Acad Sci U S A 1995; 92:6892–5
206.
Fiskum G: Mitochondrial participation in ischemic and traumatic neural cell death. J Neurotrauma 2000; 17:843–55
207.
Emerit J, Edeas M, Bricaire FL: Neurodegenerative diseases and oxidative stress. Biomed Pharmacother 2004; 58:39–46
208.
Wei H, Kang B, Wei W, Liang G, Meng QC, Li Y, Eckenhoff RG: Isoflurane and sevoflurane affect cell survival and BCL-2/BAX ratio differently. Brain Res 2005; 1037:139–47
209.
Fraker PJ, Lill-Elghanian DA: The many roles of apoptosis in immunity as modified by aging and nutritional status. J Nutr Health Aging 2004; 8:56–63
210.
Keller JN, Mattson MP: Roles of lipid peroxidation in modulation of cellular signaling pathways, cell dysfunction, and death in the nervous system. Rev Neurosci 1998; 9:105–16
211.
Liu Y, Rosenthal RE, Haywood Y, Miljkovic-Lolic M, Vanderhoek JY, Fiskum G: Normoxic ventilation after cardiac arrest reduces oxidation of brain lipids and improves neurological outcome. Stroke 1998; 29:1679–86
212.
Stewart VC, Heales SJ: Nitric oxide-induced mitochondrial dysfunction: Implications for neurodegeneration. Free Radic Biol Med 2003; 34:287–303
213.
Lambert AJ, Portero-Otin M, Pamplona R, Merry BJ: Effect of ageing and caloric restriction on specific markers of protein oxidative damage and membrane peroxidizability in rat liver mitochondria. Mech Ageing Dev 2004; 125:529–38
214.
Jenner P: Oxidative stress in Parkinson’s disease. Ann Neurol 2003; 53 (suppl 3):S26–36
215.
Nieminen AL: Apoptosis and necrosis in health and disease: Role of mitochondria. Int Rev Cytol 2003; 224:29–55
216.
Bertoni-Freddari C, Fattoretti P, Giorgetti B, Solazzi M, Balietti M, Meier-Ruge W: Role of mitochondrial deterioration in physiological and pathological brain aging. Gerontology 2004; 50:187–92
217.
Floyd RA, Hensley K: Oxidative stress in brain aging: Implications for therapeutics of neurodegenerative diseases. Neurobiol Aging 2002; 23:795–807
218.
Urban L, Bessenyei B, Marka M, Semsei I: On the role of aging in the etiology of autoimmunity. Gerontology 2002; 48:179–84
219.
Orth M, Schapira AH: Mitochondria and degenerative disorders. Am J Med Genet 2001; 106:27–36
220.
Terman A, Brunk UT: Myocyte aging and mitochondrial turnover. Exp Gerontol 2004; 39:701–5
221.
Corral-Debrinski M, Shoffner JM, Lott MT, Wallace DC: Association of mitochondrial DNA damage with aging and coronary atherosclerotic heart disease. Mutat Res 1992; 275:169–80
222.
Rosenfeldt F, Miller F, Nagley P, Rowland M, Ou R, Marasco S, Lyon W, Esmore D: Response of the senescent heart to stress: Clinical therapeutic strategies and quest for mitochondrial predictors of biological age. Ann N Y Acad Sci 2004; 1019:78–84
223.
Monteiro P, Oliveira PJ, Concalves L, Providencia LA: Pharmacological modulation of mitochondrial function during ischemia and reperfusion. Rev Port Cardiol 2003; 22:407–29
224.
de Cavanagh EM, Piotrkowski B, Fraga CG: Concerted action of the renin-angiotensin system, mitochondria, and antioxidant defenses in aging. Mol Aspects Med 2004; 25:27–36
225.
Munzel T, Kurz S, Drexler H: Are alterations of skeletal muscle ultrastructure in patients with heart failure reversible under treatment with ACE-inhibitors? Herz 1993; 18:400–5
226.
Nouette-Gaulain K, Forestier F, Malgat M, Marthan R, Mazat JP, Sztark F: Effects of bupivacaine on mitochondrial energy metabolism in heart of rats following exposure to chronic hypoxia. Anesthesiology 2002; 97:1507–11
227.
Byrne JA, Grieve DJ, Cave AC, Shah AM: Oxidative stress and heart failure. Arch Mal Coeur Vaiss 2003; 96:214–21
228.
Corral-Debrinski M, Stepien G, Shoffner JM, Lott MT, Kanter K, Wallace DC: Hypoxemia is associated with mitochondrial DNA damage and gene induction: Implications for cardiac disease. JAMA 1991; 266:1812–6
229.
Yorek MA: The role of oxidative stress in diabetic vascular and neural disease. Free Radic Res 2003; 37:471–80
230.
Pipinos II, Sharov VG, Shepard AD, Anagnostopoulos PV, Katsamouris A, Todor A, Filis KA, Sabbah HN: Abnormal mitochondrial respiration in skeletal muscle in patients with peripheral arterial disease. J Vasc Surg 2003; 38:827–32
231.
Petersen KF, Befroy D, Dufour S, Dziura J, Ariyan C, Rothman DL, DiPietro L, Cline GW, Shulman GI: Mitochondrial dysfunction in the elderly: Possible role in insulin resistance. Science 2003; 300:1140–2
232.
Petersen KF, Dufour S, Befroy D, Garcia R, Shulman GI: Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes. N Engl J Med 2004; 350:664–71
233.
Fink MP: Cytopathic hypoxia: Mitochondrial dysfunction as mechanism contributing to organ dysfunction in sepsis. Crit Care Clin 2001; 17:219–37
234.
Fink MP: Bench-to-bedside review: Cytopathic hypoxia. Crit Care 2002; 6:491–9
235.
Budinger GR, Duranteau J, Chandel NS, Schumacker PT: Hibernation during hypoxia in cardiomyocytes: Role of mitochondria as the O2 sensor. J Biol Chem 1998; 273:3320–6
236.
Fink M: Cytopathic hypoxia in sepsis. Acta Anaesthesiol Scand 1997; 110 (suppl):87–95
237.
Levy RJ, Vijayasarathy C, Raj NR, Avadhani NG, Deutschman CS: Competitive and noncompetitive inhibition of myocardial cytochrome C oxidase in sepsis. Shock 2004; 21:110–4
238.
Brunori M, Giuffre A, Sarti P, Stubauer G, Wilson MT: Nitric oxide and cellular respiration. Cell Mol Life Sci 1999; 56:549–57
239.
Gellerich FN, Trumbeckaite S, Hertel K, Zierz S, Muller-Werdan U, Werdan K, Redl H, Schlag G: Impaired energy metabolism in hearts of septic baboons: Diminished activities of complex I and complex II of the mitochondrial respiratory chain. Shock 1999; 11:336–41
240.
Callahan LA, Supinski GS: Downregulation of diaphragm electron transport chain and glycolytic enzyme gene expression in sepsis. J Appl Physiol 2005; 99:1120–6
241.
Watts JA, Kline JA, Thornton LR, Grattan RM, Brar SS: Metabolic dysfunction and depletion of mitochondria in hearts of septic rats. J Mol Cell Cardiol 2004; 36:141–50
242.
Martins GA, David CM, Mitchell C, Rosas SL, da Costa Carvalho MG: Expression of heat shock protein 70 in leukocytes of patients with sepsis. Int J Mol Med 1999; 3:401–4
243.
Sheth K, De A, Nolan B, Friel J, Duffy A, Ricciardi R, Miller-Graziano C, Bankey P: Heat shock protein 27 inhibits apoptosis in human neutrophils. J Surg Res 2001; 99:129–33
244.
Bruemmer-Smith S, Stuber F, Schroeder S: Protective functions of intracellular heat-shock protein (HSP) 70-expression in patients with severe sepsis. Intensive Care Med 2001; 27:1835–41
245.
Weiss YG, Bouwman A, Gehan B, Schears G, Raj N, Deutschman CS: Cecal ligation and double puncture impairs heat shock protein 70 (HSP-70) expression in the lungs of rats. Shock 2000; 13:19–23
246.
Ziegler TR, Ogden LG, Singleton KD, Luo M, Fernandez-Estivariz C, Griffith DP, Galloway JR, Wischmeyer PE: Parenteral glutamine increases serum heat shock protein 70 in critically ill patients. Intensive Care Med 2005; 31:1079–86
247.
Lithgow GJ, Walker GA: Stress resistance as a determinate of C. elegans  lifespan. Mech Ageing Dev 2002; 123:765–71
248.
Rattan SI: Ageing, gerontogenes, and hormesis. Indian J Exp Biol 2000; 38:1–5
249.
Zaugg M, Schaub MC: Signaling and cellular mechanisms in cardiac protection by ischemic and pharmacological preconditioning. J Muscle Res Cell Motil 2003; 24:219–49
250.
Fryer RM, Auchampach JA, Gross GJ: Therapeutic receptor targets of ischemic preconditioning. Cardiovasc Res 2002; 55:520–5
251.
Zaugg M, Lucchinetti E, Uecker M, Pasch T, Schaub MC: Anaesthetics and cardiac preconditioning: I. Signaling and cytoprotective mechanisms. Br J Anaesth 2002; 91:551–65
252.
Kapinya KJ, Lowl D, Futterer C, Maurer M, Waschke KF, Isaev NK, Dirnagl U: Tolerance against ischemic neuronal injury can be induced by volatile anesthetics and is inducible NO synthase dependent. Stroke 2002; 33:1889–98
253.
Piriou V, Chiari P, Gateau-Roesch O, Argaud L, Muntean D, Salles D, Loufouat J, Gueugniaud PY, Lehot JJ, Ovize M: Desflurane-induced preconditioning alters calcium-induced mitochondrial permeability transition. Anesthesiology 2004; 100:581–8
254.
Ismaeil MS, Tkachenko I, Gamperl AK, Hickey RF, Cason BA: Mechanisms of isoflurane-induced myocardial preconditioning in rabbits. Anesthesiology 1999; 90:812–21
255.
Toller WG, Montgomery MW, Pagel PS, Hettrick DA, Warltier DC, Kersten JR: Isoflurane-enhanced recovery of canine stunned myocardium: Role for protein kinase C? Anesthesiology 1999; 91:713–22
256.
Toma O, Weber NC, Wolter JI, Obal D, Preckel B, Schlack W: Desflurane preconditioning induces time-dependent activation of protein kinase C epsilon and extracellular signal–regulated kinase 1 and 2 in the rat heart in vivo . Anesthesiology 2004; 101:1372–80
257.
Stowe DF, Kevin LG: Cardiac preconditioning by volatile anesthetic agents: A defining role for altered mitochondrial bioenergetics. Antioxid Redox Signal 2004; 6:439–48
258.
Ludwig LM, Tanaka K, Eells JT, Weihrauch D, Pagel PS, Kersten JR, Warltier DC: Preconditioning by isoflurane is mediated by reactive oxygen species generated from mitochondrial electron transport chain complex III. Anesth Analg 2004; 99:1308–15
259.
Riess ML, Eells JT, Kevin LG, Camara AK, Henry MM, Stowe DF: Attenuation of mitochondrial respiration by sevoflurane in isolated cardiac mitochondria is mediated in part by reactive oxygen species. Anesthesiology 2004; 100:498–505
260.
Riess ML, Kevin LG, McCormick J, Jiang MT, Rhodes SS, Stowe DF: Anesthetic preconditioning: The role of free radicals in sevoflurane-induced attenuation of mitochondrial electron transport in Guinea pig isolated hearts. Anesth Analg 2005; 100:46–53
261.
Weber NC, Toma O, Awan S, Frassdorf J, Preckel B, Schlack W: Effects of nitrous oxide on the rat heart in vivo : Another inhalational anesthetic that preconditions the heart? Anesthesiology 2005; 103:1174–82
262.
Gray JJ, Bickler PE, Fahlman CS, Zhan X, Schuyler JA: Isoflurane neuroprotection in hypoxic hippocampal slice cultures involves increases in intracellular Ca2+and mitogen-activated protein kinases. Anesthesiology 2005; 102:606–15
263.
Miki T, Seino S: Roles of KATP channels as metabolic sensors in acute metabolic changes. J Mol Cell Cardiol 2005; 38:917–25
264.
Yamada K, Inagaki N: Neuroprotection by KATP channels. J Mol Cell Cardiol 2005; 38:945–9
265.
Kane GC, Liu XK, Yamada S, Olson TM, Terzic A: Cardiac KATP channels in health and disease. J Mol Cell Cardiol 2005; 38:937–43
266.
Tanaka K, Ludwig LM, Kersten JR, Pagel PS, Warltier DC: Mechanisms of cardioprotection by volatile anesthetics. Anesthesiology 2004; 100:707–21
267.
Patel HH, Ludwig LM, Fryer RM, Hsu AK, Warltier DC, Gross GJ: Delta opioid agonists and volatile anesthetics facilitate cardioprotection via  potentiation of K(ATP) channel opening. FASEB J 2002; 16:1468–7
268.
Sergeev P, da Silva R, Lucchinetti E, Zaugg K, Pasch T, Schaub MC, Zaugg M: Trigger-dependent gene expression profiles in cardiac preconditioning: Evidence for distinct genetic programs in ischemic and anesthetic preconditioning. Anesthesiology 2004; 100:474–88
269.
Chiari PC, Bienengraeber MW, Weihrauch D, Krolikowski JG, Kersten JR, Warltier DC, Pagel PS: Role of endothelial nitric oxide synthase as a trigger and mediator of isoflurane-induced delayed preconditioning in rabbit myocardium. Anesthesiology 2005; 103:74–83
270.
Jamnicki-Abegg M, Weihrauch D, Pagel PS, Kersten JR, Bosnjak ZJ, Warltier D, Bienengraeber MW: Isoflurane inhibits cardiac myocyte apoptosis during oxidative and inflammatory stress by activating Akt and enhancing Bcl-2 expression. Anesthesiology 2005; 103:1006–14
271.
Feng J, Lucchinetti E, Ahuja P, Pasch T, Perriard J-C, Zaugg M: Isoflurane postconditioning prevents opening of the mitochondrial permeability transition pore through inhibition of glycogen synthase kinase 3β. Anesthesiology 2005; 103:987–95
272.
Asano G, Takashi E, Ishiwata T, Onda M, Yokoyama M, Naito Z, Ashraf M, Sugisaki Y: Pathogenesis and protection of ischemia and reperfusion injury in myocardium. J Nippon Med School 2003; 70:384–92
273.
Zaugg M, Lucchinetti E, Garcia C, Pasch T, Spahn DR, Schaub MC: Anaesthetics and cardiac preconditioning: II. Clinical implications. Br J Anaesth 2003; 91:566–76
274.
Weiss YG, Maloyan A, Tazelaar J, Raj N, Deutschman CS: Adenoviral transfer of HSP-70 into pulmonary epithelium ameliorates experimental acute respiratory distress syndrome. J Clin Invest 2002; 110:801–6
275.
Anderson DM, Hall LL, Ayyalapu AR, Irion VR, Nantz MH, Hecker JG: Stability of mRNA/cationic lipid lipoplexes in human and rat cerebrospinal fluid: Methods and evidence for nonviral mRNA gene delivery to the central nervous system. Hum Gene Ther 2003; 14:191–202
276.
Raeburn CD, Cleveland JC Jr, Zimmerman MA, Harken AH: Organ preconditioning. Arch Surg 2001; 136:1263–6
277.
Armstead WM, Hecker JG: Heat shock protein modulation of KATP and KCa channel cerebrovasodilation after brain injury. Am J Physiol 2005; 289:H1184–90
278.
Zhang JG, Tirmenstein MA, Nicholls-Grzemski FA, Fariss MW: Mitochondrial electron transport inhibitors cause lipid peroxidation-dependent and -independent cell death: Protective role of antioxidants. Arch Biochem Biophys 2001; 393:87–96
279.
Zhuang H, Kim YS, Koehler RC, Dore S: Potential mechanism by which resveratrol, a red wine constituent, protects neurons. Ann N Y Acad Sci 2003; 993:276–86
280.
Kim YS, Zhuang H, Koehler RC, Dore S: Distinct protective mechanisms of HO-1 and HO-2 against hydroperoxide-induced cytotoxicity. Free Radic Biol Med 2005; 38:85–92
281.
Liu X, Wei J, Peng DH, Layne MD, Yet SF: Absence of heme oxygenase-1 exacerbates myocardial ischemia/reperfusion injury in diabetic mice. Diabetes 2005; 54:778–84
282.
Hoetzel A, Leitz D, Schmidt R, Tritschier E, Bauer I, Loop T, Humar M, Geiger KK, Pannen BHJ: Mechanism of hepatic heme oxygenase-1 induction by isoflurane. Anesthesiology 2006; 104:101–9
283.
Motterlini R, Foresti R, Bassi R, Calabrese V, Clark JE, Green CJ: Endothelial heme oxygenase-1 induction by hypoxia: Modulation by inducible nitric-oxide synthase and S-nitrosothiols. J Biol Chem 2000; 275:13613–20
284.
Boyd CS, Cadenas E: Nitric oxide and cell signaling pathways in mitochondrial-dependent apoptosis. Biol Chem 2002; 383:411–23
285.
Nystul TG, Roth MB: Carbon monoxide-induced suspended animation protects against hypoxic damage in Caenorhabditis elegans . Proc Nat Acad Sci U S A 2004; 101:9133–6
286.
Clark JE, Naughton P, Shurey S, Green CJ, Johnson TR, Mann BE, Foresti R, Motterlini R: Cardioprotective actions by a water-soluble carbon monoxide-releasing molecule. Circ Res 2003; 93:e2–8
287.
Shan X, Jones DP, Hashmi M, Anders MW: Selective depletion of mito-chondrial glutathione concentrations by (R,S)-3-hydroxy-4-pentenoate potentiates oxidative cell death. Chem Res Toxicol 1993; 6:75–81
288.
Reiter RJ, Tan DX, Manchester LC, El-Sawi MR: Melatonin reduces oxidant damage and promotes mitochondrial respiration: Implications for aging. Ann N Y Acad Sci 2002; 959:238–50
289.
Reiter RJ, Tan DX, Qi W, Manchester LC, Karbownik M, Calvo JR: Pharmacology and physiology of melatonin in the reduction of oxidative stress in vivo . Biol Signals Recept 2000; 9:160–71
290.
Moller JT, ISOPCD Investigators: Long-term postoperative cognitive dysfunction in the elderly: ISPOCD1 study. Lancet 1998; 351:857–61
291.
Abildstrom H, Rasmussen LS, Rentowl P, Hanning CD, Rasmussen H, Kristensen PA, Moller JT: Cognitive dysfunction 1-2 years after non-cardiac surgery in the elderly. ISPOCD group. International Study of Post-Operative Cognitive Dysfunction. Acta Anaesthesiol Scand 2000; 44:1246–51
292.
Hanania M, Kitain E: Melatonin for the treatment and prevention of postoperative delirium. Anesth Analg 2002; 94:338–9

Appendix 1: Clinical Characteristics of Inherited Mitochondrial Cytopathy

  • For large-scale mtDNA deletions: ataxia, peripheral neuropathy, muscle weakness, ophthalmoplegia, ptosis, pigmentary retinopathy, hypoparathyroidism, cardiomyopathy, cardiac conduction defects, sensorineural hearing loss, Fanconi syndrome, lactic acidosis, ragged-red fibers on muscle biopsy

  • For single mtDNA point mutation with mRNA abnormality: seizures, ataxia, psychomotor regression, dystonia, muscle weakness, pigmentary retinopathy, optic atrophy, cardiomyopathy, lactic acidosis, sensorineural hearing loss

  • For multiple mtDNA point mutations with mRNA abnormality: dystonia, optic atrophy, cardiac conduction defects

Appendix 2: Common Therapeutic Treatments and Supplements Used by Patients with Inherited Mitochondrial Cytopathy and Neurodegenerative Disorders

  • β-Carotene

  • L-carnitine

  • Acetyl-L-carnitine

  • Riboflavin (vitamin B2)

  • Nicotinamide (vitamin B3)

  • Vitamin K

  • Vitamin E

  • Vitamin C

  • Thiamine (vitamin B1)

  • Coenzyme Q

  • Selenium, magnesium

  • Calcium, phosphorous

  • Biotin

  • Succinate

  • Creatine

  • Citrates

  • Prednisone

  • Folic acid

  • Lipoic acid