Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Biology_of_Turtles

.pdf
Скачиваний:
88
Добавлен:
11.02.2015
Размер:
44.64 Mб
Скачать

306

Biology of Turtles

sumption rates and increases in plasma lactate—begins at about 22 mmHg (Milton, 1994). Recent studies have shown a wide variation in the capacity of neurons to tolerate hypoxia, reflective of the function and the degree of hypoxia normally encountered (Milton & Prentice, in press). In the mammalian brain, the hippocampus, corpus striatum, and Purkinje cells are the most vulnerable to hypoxic or ischemic damage; within the hippocampus, CA1 neurons are more sensitive than cells of the CA3 region (Pulsinelli, 1985). For comparison, let us look first at the sequence of events in the mammalian brain that occur when oxygen is lacking—for example, as might happen in the event of a cardiac arrest, or more locally in the event of a stroke.

12.2The Brain in Crisis

As the end electron acceptor that makes mitochondrial oxidative phosphorylation possible, oxygen is critical to nearly all forms of life on earth. Severe hypoxia or cerebral ischemia forces tissues to rely on the ATP supplied by anaerobic glycolysis, which supplies only 2 moles of ATP per mole of glucose rather than the far greater amounts of ATP generated from oxidative phosphorylation. In theory, each mole of glucose supplies 36 moles of ATP by oxidative phosphorylation, though empirical measurements indicate that true ATP production is closer to 25 moles/mole glucose (Hochachka & Somero, 2002). When forced to rely on less than 1/10th of its normal energy supply, cellular ATP demand rapidly outstrips ATP supply and ATP-driven processes begin to fail. In the neurons, failure of ATP-dependent ion transporters leads to the breakdown of membrane potential; K+ leakage out of the cells triggers membrane depolarization and the release of excitatory amino acids (EAA) such as aspartate and glutamate (Lutz et al., 2003). Elevated levels of glutamate, due to both increased release and decreased reuptake, are generally regarded as a significant factor leading to hypoxic or ischemic cell death. High extracellular glutamate stimulates NMDA receptors and permits an influx of calcium from the extracellular space (Harukuni & Bhardwaj, 2006). Increased intracellular Ca2+ is thought to lead to neurodegenerative events including the generation of excess levels of reactive oxygen species (ROS), lipid peroxidation, and the activation of executioner caspases that trigger apoptosis (programmed cell death). The activation of glutamate receptors (Dugan et al., 1995; Sharp et al., 2005; Kahlert et al., 2005) and elevated intracellular calcium (Sharikabad et al., 2004; Nagy et al., 2004) are associated with free-radical formation, and the direct activation of NMDA receptors results in a massive release of hydroxyl radicals (OH) (Lancelot et al., 1998; Laplanche et al., 2003). High intracellular calcium also activates proteases, lipases, and endonucleases, which in turn destroy cellular integrity (Lutz et al., 2003).

In addition to glutamate related damage, hypoxic or ischemic events cause the release of the monoamine dopamine into the extracellular space; however, in contrast to other excitotoxins, increases are seen before cellular depolarization, even under conditions of mild hypoxia (Globus et al., 1988; Huang et al., 1994) and may be a major cause of hypoxic/ischemic damage (Mitsuyo et al., 2003). Dopamine is thought to contribute to neuronal damage by increasing the release of excitatory neurotransmitters, through the production of oxygen free radicals, by inhibiting Na+/K+ ATPase and by uncoupling glucose metabolism from cerebral blood flow (Lutz et al., 2003). Following hypoxia/ ischemia, additional neuronal damage occurs during reperfusion, thought to be caused by the postischemic release of oxygen radicals, the synthesis of nitric oxide, inflammation, and an imbalance between the excitatory and inhibitory neurotransmitter systems (Berger et al., 2002); this damage can continue for days after the initial hypoxic/ischemic insult.

12.3The Anatomy and Physiology of Anoxic Survival

By contrast, certain species of freshwater turtle can survive extended periods of anoxia and fully recover (Ultsch, 2006), as can the Crucian carp (Carassius carassius), a northern European fish that likewise winters in iced-over, anoxic, ponds (Lutz et al., 2003). Extended anoxic survival involves the downregulation of energy-demanding processes and those that are potentially damaging to

The Physiology and Anatomy of Anoxia Tolerance in the Freshwater Turtle Brain

307

cellular survival, and a concurrent upregulation of a variety of protective mechanisms that decrease energy demand and promote cell survival. For example, factors that are downregulated include the enzymes of both glycolysis and oxidative phosphorylation (Brooks & Storey, 1988; Brooks & Storey, 1989), an overall decrease in protein synthesis (Fraser et al., 2001), the release of excitotoxic neurotransmitters (Nilsson & Lutz, 1991; Milton & Lutz, 1998; Milton et al., 2002; Milton & Lutz, 2005), and the activity of ion channels, all of which decrease electrical activity (Fernandes et al., 1997) and thus lower both energy demand and overall metabolic rate.

Conversely, a number of pro-survival mechanisms are either already constitutively elevated or are clearly upregulated, including increasing glycolytic energy supply through high constitutive tissue levels of glycogen (McDougal et al., 1968; Clark & Miller, 1973), increases in plasma glucose (Milton, 1994), increased brain blood flow (Hylland et al., 1994), the release of inhibitory neuromodulators such as adenosine (Nilsson & Lutz, 1992) and GABA (Nilsson & Lutz, 1991), as well as the maintenance of (Lutz & Manuel, 1999) or increase in (Lutz & Leone-Kabler, 1995) neurotransmitter receptor function. Changes at the molecular level also occur, including increased expression or activation of heat shock proteins (Prentice et al., 2004), members of the mitogen activated kinase families (Greenway & Storey, 2000; Milton, unpublished data), and the recently discovered neuroglobin (Milton et al., 2006).

12.3.1 Energy Supply and the Enzymes of Anaerobic Glycolysis

The upregulation of energy conserving processes, coupled to the downregulation of energy expenditures, leaves the turtle in a state of profound, reversible, metabolic depression. The key to survival is to lower energy demand to meet the reduced energy supplied by anaerobic glycolysis; thus, the ability to carry out widespread glycolysis is critical. A study by Suarez et al. (1989) comparing glucose use in rats and turtles showed that after accounting for temperature differences, there was at least a sixfold difference in metabolic rates between rats and turtles, suggesting that lower rates of ATP use in turtles are inherently protective in anoxia as they are more easily supported by anaerobic glycolysis. However, lower metabolic rates and temperature differences alone are clearly not sufficient for anoxia survival, as the brains of other poikilotherms respond to anoxia in a manner similar to mammals. Remarkably, at their respective biological temperatures, the glycolytic capacity of rats and turtles is similar, as determined by the activities of hexokinase and lactate dehydrogenase (Suarez et al., 1989). This indicates not only a greater reliance on glycolysis by turtles even under normoxic conditions but a ready capacity for anaerobic glycolysis in anoxia. Key enzymes that have been studied in relation to brain metabolism include both those critical to oxidative phosphorylation, primarily cytochrome c oxidase, and those of the glycolytic pathway including hexokinase, lactate dehydrogenase, and glycogen phosphorylase. As is also seen in mammals, there is a detectable general pattern of differences in metabolism between gray and white matter in the turtle brain, such as gray/white matter differences in glucose metabolism indicated by 2-deoxyglucose use and citrate synthase (Sokoloff et al., 1977; Suarez et al., 1989). These differences result in a repeatedly observed rostral-to-caudal gradation of metabolic indicators, including the enzymes of energy production and antioxidants, that in general declines in parallel with the increase in white matter content of the more caudal brain regions (Oke et al., 1987), perhaps because the metabolic rate of the soma (and greater numbers of energy consuming dendrites and synapses) is greater than that of axons and a greater proportion of gray matter is somata.

As the terminal oxidase of the electron transport chain, the activity of cytochrome oxidase (CO) is directly linked to oxygen consumption. As neural activity is constrained, at least in part, by the availability of cellular ATP, and the activity of CO in aerobic metabolism determines the level of ATP available to the cell, then CO activity in turn can serve as a marker of metabolic capacity (Sakata et al., 2005). Hence, CO histochemistry has been used extensively to follow changes in neural metabolism after experimental manipulations; for example, significant decreases in CO activity have been shown to follow experimentally induced decreases in excitatory afferent input

308

Biology of Turtles

in mammalian cells (Hevner et al., 1992; Hevner & Wong-Riley, 1993; Wong-Riley et al., 1998a). On a finer scale, increased CO activity indicative of elevated metabolic capacity in the hypothalamus, nucleus sphericus, and septum is associated with heightened aggressiveness and territoriality in both male and female leopard geckos (Coomber et al., 1997; Sakata et al., 2002). In general, CO activity closely follows that of Na+/K+ ATPase, rather than being limited to a particular neurotransmitter or cell signaling system (Wong-Riley et al., 1998b; Sakata et al., 2005). When Xia et al. (1992) compared CO activity in the adult and newborn rat and the turtle brain, activity levels were found to be highest in the rostral brain areas in both adult rat and turtle compared to the brain stem and spinal cord. Although overall activity in the turtle brain was only 20 to 30% that of the rat brain, in keeping with their overall lower oxygen demand, the trend was similar, with CO activity in the turtle brain highest in the cortex, dorsal ventricular ridge, and the paleostriatum augmentum (Figure 12.3).

A similar pattern was reported by Suarez et al. (1989) for citrate synthase (CS), an enzyme of the citric acid cycle. In the rat, CS activity was highest in the cortex and striatum and lowest in the pons-medulla and hypothalamus. In the turtle, activity was highest in the cerebellum and dorsal ventricular ridge and lowest in the medulla, although activity was also low in the olfactory lobe. Overall oxidative capacity was twoto three-fold lower in turtles than rats (Suarez et al., 1989). As is the general pattern for CO, CS activity could be linked to Na+/K+ ATPase activity; both were highest in the cortex of the rat and lowest in the hypothalamus, whereas both were highest in the dorsal ventricular ridge (DVR) of the turtle—the reptilian structural analogue to the mammalian hippocampus—and lowest in the olfactory bulb (Suarez et al., 1989).

Alternatively, activity of the glycolytic enzyme hexokinase (HK) is actually higher in some parts of the turtle brain compared to the rat, with a very different distribution from that of CO (Xia et al., 1992). In the adult rat, HK levels are still higher in the cortex than the brainstem, as with CO, but there is comparatively less heterogeneity. Cortex levels of HK activity in the turtle brain are indistinguishable from the rat despite innate differences in metabolic demand, whereas levels in the

Cytochome Activity (relative optical density)

0.8

Turtle Cyt c

0.7

Rat Cyt c

 

0.6

 

0.5

 

0.4

 

0.3

 

0.2

 

0.1

 

0

Cortex Hippocampus alamus Cerebellum Brainstem Spinal cord

Figure 12.3  Cytochrome c oxidase activities in the CNS of the adult rat and the freshwater turtle Trachemys scripta. Values are means ± SD of relative optical densities for n = 9 corresponding brain sections in three animals per group. CO activity in all turtle sections is significantly lower than that seen in rat CNS. Data adapted from Xia et al. (1992).

The Physiology and Anatomy of Anoxia Tolerance in the Freshwater Turtle Brain

309

Hexokinase Activity (relative optical densities)

0.3

 

 

 

0.25

 

 

 

0.2

 

 

Turtle

 

 

 

Rat

0.15

 

 

 

*

*

 

*

*

*

0.1

 

0.05

 

 

 

0

Thalamus Cerebellum

Brainstem Spinal cord

Cortex Hippocampus

Figure 12.4  Hexokinase activities in the CNS of the adult rat and the freshwater turtle Trachemys scripta. Values are means ± SD of relative optical densities for n = 12 corresponding brain sections in four animals per group. The asterisk indicates a significant difference between rat and turtle. Data adapted from Xia et al. (1992).

hippocampus, thalamus, and hypothalamus are still approximately half that of the rat (Figure 12.4). Importantly, HK activity in the turtle CNS ranges from 37% to more than 100% higher in the spinal cord, brain stem, and cerebellum compared to activity levels in the adult rat (Xia et al., 1992), presumably conferring a greater glycolytic capacity to these areas and to the turtle brain overall.

Lactate dehydrogenase (LDH) activities follow the same rostral-to-caudal pattern in the rat brain as other metabolic enzymes, being highest in the cortex and striatum and lowest in the ponsmedulla and hypothalamus, whereas turtle LDH activity was highest in the thalamus and optic lobe and lowest in the olfactory lobe and medulla (Suarez et al., 1989). However, when normalized for temperature turtles appear to have more HK and LDH activity overall, such that even lower rates in the medulla are still high compared to the rat (Suarez et al., 1989). High rates of glycolysis in the hindbrain could then support, for example, the continued functioning in anoxia of the Na+/K+ ATPase (Hylland et al., 1997), and thus continued functioning of the lower brain regions. Compared to mammals, anoxia-tolerant turtles have a high brain glycogen content (McDougal et al., 1968; Partata & Marques, 1994), with glycogen located primarily in the neurons rather than in the glia as is seen in mammals (Partata & Achaval-Elena, 1995). Anoxia also produces a marked hyperglycemia in turtles, with plasma glucose increasing from basal values of 3 mM to as high as 25 mM at room temperature (Penney, 1974; Keiver et al., 1992).

The idea that high glycolytic capacity is indicative of continued function during anoxia is supported by the work of Partata et al. (1999). They examined the distribution of both glycogen phosphorylase (GP) and CO in the South American turtle Trachemys dorbigni to create a metabolic map of the turtle brain, finding that GP and CO activities were neither distributed diffusely nor associated exclusively with a particular functional system. As in studies of the rat, the patterns of enzyme distribution partially overlapped with activity of both enzymes high in discrete portions of the cortex (cortex medialis, cortex dorsomedialis), in the striatum and substantia nigra, and in the primordium hippocampi (Partata et al., 1999). In the rat, GP is found primarily in glial cells (Swanson & Choi, 1993), whereas in turtles it is localized to the neurons (Partata & Achaval-Elena, 1995). Of course, as GP is part of the glycogen utilization pathway, the authors correlate GP activity with anaerobic metabolism, and thus suggest that in some parts of the brain both aerobic and anaerobic metabolism contribute to functional energy demand, whereas in other regions one pathway is more

310

Biology of Turtles

dominant. However, an alternative explanation is that it is more critical to maintain function in some regions of the brain during anoxia than in others. For example, in the striatum the release and reuptake of neurotransmitters continues during long-term anoxia (Milton & Lutz, 1998; Milton et al., 2002) despite energetic costs of reuptake transport mechanisms of as high as 1.5 ATP/molecule (Swanson & Duan, 1999). The continued release and re-uptake of these compounds implies a likely function for the neurotransmitters during anoxia, perhaps to maintain neuronal networks or prepare for the eventual upregulation of metabolism during recovery (Milton et al., 2002), but the energetic demands of transport also means a continuous energy supply is required. Thus, the high activity levels of both GP and CO would allow aerobic metabolism under basal conditions, with a change to anaerobiosis during anoxia at levels sufficient to maintain neurotransmitter homeostasis, albeit at a reduced level. Along these lines, Partata et al. (1999) note the remarkable overlap in distribution of GP and CO in the medulla, where HK activity was also found to be high (Xia et al., 1992). The authors also suggested that the CO and HK overlap is critical as CO may regulate the onset of glycolysis; lactate production is known to begin before CO is fully reduced in the anoxic turtle brain (Lutz et al., 1984; Lutz et al., 1985). By contrast, the dorsal ventricular ridge and most layers of the optic tectum expressed significant GP levels but not CO activity, indicating some continued function for the eye during anoxia (Rosenberg & Ariel, 1990; Fan et al., 1995, 1997). However, the BON showed detectable CO but not GP (Partata et al., 1999), though continued normal electrical activity has been shown to occur in the accessory optic system for several days in vitro (Ariel & Fan, 1993; Johnson et al., 1998). The sum of these metabolic enzymes studies then implies that the continued function of the brainstem in anoxia is critical to anoxic survival or recovery.

Alternatively, these studies did not measure enzyme activity levels during anoxia, and it has been known for many years that enzymes of the glycolytic pathway as well as those of aerobic metabolism are strongly downregulated in anoxia both in the brain and in other organs (Brooks & Storey, 1988, 1989). In turtles, as in other good facultative anaerobes, glycolytic activation is attenuated after the first hour or two of anoxia; as the rate of energy consumption gets more and more depressed, so does the rate of energy production (Kelly & Storey, 1988). Despite the widespread presence of glycogen phosphorylase reported by Partata et al. (1999), GP activity in the brain is reduced by 70% over the initial hours of anoxia (Brooks & Storey, 1988). Phosphofructokinase (PFK) shows increased inhibition as well, reflecting its role as the primary locus of glycolytic rate depression (Brooks & Storey, 1988). In most animals, hypoxia induces a sharp increase in anaerobic energy production (the Pasteur Effect) as metabolic supply attempts to keep pace with metabolic demand. The Pasteur Effect results primarily from the release of metabolic control at the PFK locus (Storey, 1988); in the turtle, covalent modification of enzymes results rather in a depression of metabolic rate (Brooks & Storey, 1988, 1989).

Energy production then continues in the anoxic turtle brain, albeit at greatly reduced rates that suffice to meet the reduced energy demands of the hypometabolic state, with some sections of the brain metabolically equipped for extended glycolysis. The distribution of glycolytic enzymes within the brain is likely to indicate critical areas that maintain function even in extended anoxia; the hypometabolism of anoxia is not simply a “shutting down” of the brain but is a selective process presumed to maximize energy savings while allowing critical processes to continue. Studies have revealed that whereas many processes are minimized, others continue in anoxia, including the continued release and reuptake of neurotransmitters, the upregulation of protective molecular pathways, and selective electrical activity (Lutz & Milton, 2004).

12.3.2 Ion Channels

As with all cells of the body, neurons maintain their ion balance across the cell membrane with a higher concentration of K+ ions inside the cell compared to the outside, whereas Na+ ion gradients are the reverse. These ion differences are used by the neuron to generate action potentials, and the restoration of ion gradients after an action potential—plus counteracting continuous leak channels

The Physiology and Anatomy of Anoxia Tolerance in the Freshwater Turtle Brain

311

within the membrane—consumes enormous amounts of energy. Within the brain, as much as 50 to 60% of the total normoxic ATP demand is spent by the Na+-K+-ATPase alone to maintain ion homeostasis (Erecsinska & Silver, 1989), thus any reduction in ion flux across the membrane is likely to have a profound effect on survival by decreasing energy demand. For example, numerous studies indicate that an influx of Ca2+ from the extracellular space into the cytoplasm is the key event in a variety of neurodegenerative processes; cell death results from the calcium-related generation of ROS (Sharikabad et al. 2004; Nagy et al. 2004), lipid peroxidation, and the activation of executioner caspases (Lee et al., 2005). Disruption of cellular calcium homeostasis has been proposed as a critical event in both apoptosis and necrosis, and increases in cytoplasmic free calcium levels precede apoptotic cell death under a variety of circumstances (Kruman & Mattson, 1999), including glutamate stimulation (Stout et al., 1998) and superoxide exposure.

Suarez et al. (1989) examined the abundance of voltage gated calcium channels (VGCC) in the turtle brain. They report no consistent trends in distribution between the rat and turtle brain (Figure 12.5); while VGCC density was only half that in the turtle thalamus compared to the rat, it was a third greater in the turtle cerebellum (Suarez et al., 1989). In the rat, VGCCs were most abundant in the hippocampus, striatum, and cortex and lowest in the hypothalamus and pons, thus following to a major extent the established trend in CO, CS, and glycolytic enzyme activity (Suarez et al., 1989; Partata et al., 1999). By contrast, the abundance of the VGCC in the turtle was highest in the olfactory lobe (which has low metabolic enzyme activity) and the DVR (which has high enzyme activity); VGCCs were lowest in the turtle thalamus (high LDH activity) and the medulla (low LDH and HK activity) (Suarez et al., 1989). Thus, simple differences in VGCC abundance is not the basis for differences between the mammalian and turtle brain in either anoxia tolerance or hypoxia-sensitivity, nor does it appear that VGCC distribution can be linked in the turtle to regional metabolic capacity.

The distributions of relatively few other ion channels have been described in the turtle brain (Xia & Haddad, 1991, 1993), with some additional studies restricted to very limited regions of the brain, e.g., the brainstem and spinal cord (Keifer & Carr, 2000). Xia and Haddad (1993) compared the kinetics and distribution of voltage-sensitive Na+ channels in the adult rat and turtle brain using

Ca2+ Channel Density (pmol/g)

8

 

 

 

7

 

 

 

6

 

 

 

5

 

 

Turtle

4

 

 

Rat

 

 

 

3

 

 

 

2

 

 

 

1

 

 

 

0

Thalamus

Cerebellum

Pons-Medulla

Cortex

Figure 12.5  Voltage-dependent calcium channel densities (pmol/g) in homologous regions of the turtle and rat brains at 25°C. Data are means ± SE, n = 5 and 4 in the rat and turtle brains, respectively. Data adapted from Suarez et al. (1989).

312

Biology of Turtles

autoradiography of bound saxitoxin. As with the voltage-gated calcium channels, Na+ channels in the rat generally followed established metabolic gradients, with a very heterogeneous distribution. Saxitoxin binding in the rat brain was highest in the cortex, hippocampus, amygdala, and cerebellum and low in the brainstem and spinal cord (Xia & Haddad, 1993). Whereas saxitoxin in the turtle had similar binding properties as in the rat, overall binding density was much lower, especially in the cerebellum and rostral areas such as the cortex. Very few areas, including the medullary raphe, had similar or higher channel density in the turtle compared to the rat (Xia & Haddad, 1993). However, the differences in receptor density are unlikely to explain the approximately 100-fold difference in anoxia sensitivity between the turtle and mammalian brains, a conclusion also reached by Edwards et al. (1989). That study found that turtle synaptosomes had a voltage-gated Na+ channel density approximately a third that of rat synaptosomes—but again, this alone is of little significance when compared to the enormous differences in anoxia tolerance (Edwards et al., 1989).

Yet another difference in ion channels between turtles and rats lies in the density of ATPsensitive potassium channels (K-ATP). K-ATP channels were first discovered in cardiac myocytes (Noma, 1983) and have subsequently been found in many other excitable cell types (Ashcroft, 1988), including neurons (Mourre et al., 1989; Krnjevic, 1993). These channels are normally inhibited by physiological levels of ATP but open when ATP levels decrease (Ben Ari et al., 1990; Cameron & Baghdady, 1994). As the open probability of K-ATP channels is thus directly linked to intracellular energy (ATP/ADP) levels, the channels link the metabolic status of a cell to its electrical activity (Liss & Roeper, 2001). Sulfonylurea binding studies indicate the K-ATP are widely expressed throughout different regions of the brain, including pyramidal and striatal neurons, various nuclei of the hypothalamus, and GABAergic and dopaminergic neurons of the substantia nigra (Liss & Roeper, 2001).

K-ATP channels are thought to play a critical protective role in the early stages of brain ischemia in mammals (Ben Ari et al., 1990) through hyperpolarization of the membrane, by increasing brain blood flow (and thus glucose delivery) (Tomiyama et al., 1999), and decreasing the release of excitatory neurotransmitters (Tanaka et al., 1996; Wind et al., 1997). Blockade of K-ATP in the anoxic rat brain increases neuronal depolarization, thus suggesting that open K-ATP can dampen the anoxia-induced depolarization of CNS neurons (Jiang & Haddad, 1991). By contrast, K-ATP activation does not have any significant effect on K+ flux in newborn rats or adult turtles, suggesting that these channels are not a major route of K+ loss in more hypoxia-tolerant animals (Jiang & Haddad, 1991). Using the binding of glibenclamide, a specific inhibitor that binds to the sulfonylurea subunit of K-ATP channels, Xia and Haddad (1991) compared the binding properties and distribution of K-ATP in the rat and turtle brains with autoradiography. Highand low-affinity binding channels were revealed in both, but the rat had far higher binding densities and a more heterogeneous distribution. In the rat, there was a more than a ten-fold difference between the highest and lowest densities, with highest K-ATP levels in the substantia nigra, hippocampus, cerebellum, and a few thalamic nuclei. Intermediate levels were found in the basal ganglia, septum, thalamus, and hypoglossal nuclei, whereas densities were lowest in the cerebellum, brainstem, and spinal cord (Xia & Haddad, 1991). The cortex itself is highly variable; K-ATP are present at high densities in the hippocampal CA3 and CA4 regions and the dentate gyrus but are only intermediate in the CA1 and CA2 regions. Interestingly, the CA3 region has proven to be less sensitive to hypoxic-ischemic brain damage than the CA1 area (Pulsinelli, 1985); perhaps this is related to the protection afforded by higher densities of these ATP-sensitive channels.

In the rat brain, we again see a general rostral-to-caudal gradient, though there is less specific correlation between K-ATP channels and the enzymes of metabolism, indicating perhaps a greater association of K-ATP with specific pathways and nuclei rather than with generalized brain metabolic activity. For example, high densities in the substantia nigra could be related to the theoretical role of K-ATP channels in depressing the excessive release of excitotoxins such as dopamine in the face of an energy deficit; the substantia nigra has a high density of dopaminergic neurons that terminate in the striatum (Parent, 1979). As an aside, stimulation of these channels to increase

The Physiology and Anatomy of Anoxia Tolerance in the Freshwater Turtle Brain

313

hypoxic/ischemic neuronal survival in the mammalian brain has been an interest for nearly two decades with little result.

However, in contrast to high K-ATP channel density in the adult rat, binding densities were very low in the newborn rat and turtle brains, both more hypoxia-tolerant than the adult rat. Average densities of the K-ATP channels were 25 and 36 fmol/mg protein in the turtle and newborn rat, respectively, versus a mean 219 fmol/mg in the adult rat (Xia & Haddad, 1991). Also in sharp contrast to the adult rat was the overall homogeneity of binding density in turtles: densities in most brain regions ranged from 7 to 25 fmol/mg protein in turtles, whereas in the adult rat values ranged from 46 fmol/mg in the spinal cord to around 309 fmol/mg in the substantia nigra. Whereas densities were lower and more homogeneous in the newborn rat than the adult, this trend was even greater in the turtle brain, especially in the more caudal areas. Xia and Haddad (1991) reported densities too low to measure in the spinal cord, intermediate binding in the rostral and cerebellar regions, and values as high as 40 fmol/ng in the striatum. K-ATP densities in the turtle brain are thus about 1/10th the values of the rat brain, in keeping with their lower overall metabolic intensity; the authors conclude that K-ATP channel density then could not be the critical difference that allows for extended anoxia in turtles (Xia & Haddad, 1991).

However, more recent work has shown that even if K-ATP density is not high in the turtle brain, K-ATP channel function is of critical importance in anoxia tolerance, performing many of the same protective functions as have been reported in mammalian neurons (although apparently with more success). In mammalian cells, the opening of K-ATP channels in response to low intracellular ATP is in part protective because the K+ outflow temporarily hyperpolarizes the cell. It is critical to realize that if ATP levels do not recover, the increased conductance will result in a more rapid efflux of K+ and shorten the time to depolarization (Lutz et al., 2003); indeed, the K-ATP channels are thought to be a major route of K+ efflux during anoxic depolarization in mammals (Xia & Haddad, 1991), though they appear of less importance in the depolarizing turtle brain (Pek-Scott & Lutz, 1998). In turtles, there is much evidence that membrane potential is maintained during anoxia (Lutz et al., 2003), and anoxia is associated with a decrease in K+ efflux that can be initially blocked by K- ATP blockers (Pek-Scott & Lutz, 1998). Whereas the addition of glibenclamide blocks this anoxiainduced reduction in K+ efflux during the first hour, the effect is diminished by 2 hours anoxia and completely disappears by 4 hours anoxia; these differences are thought to be associated with a temporary fall and subsequent recovery in cellular energy stores (Pek-Scott & Lutz, 1998). The temporary opening of K-ATP channels is protective in the turtle brain during a period of temporary energy crisis: open K-ATP are associated with the decreased release of both dopamine (Milton & Lutz, 2005) and glutamate (Milton et al., 2002) during a period when the turtle brain has perhaps less energy to spare on reuptake transport mechanisms.

However, whereas the distribution and densities of ion channels may not reveal differences between anoxia-tolerant and hypoxia-sensitive animals, or determine which parts of the brain are more or less active in anoxia, changes in ion channel activity are more revealing. Modulating the conductance of ion channels can result in significant alterations to metabolism. For example, increasing Na+ channel activity with veratridine increases energy consumption by 20% in rat and turtle synaptosomes (Edwards et al., 1989). Conversely, blocking voltage-dependent Na+ channels with tetrodotoxin causes a 22% decrease in the energy consumption of turtle synaptosomes (Edwards et al., 1989) and a 20% decrease in the in situ perfused rat brain (Xie et al., 1994). Thus, the ability to alter ion conductance could be a powerful mechanism to decrease metabolic costs by reducing the costs of ion pumping. Extensive evidence shows that ion flux through membrane channels indeed decreases in the anoxic turtle brain (“channel arrest”), including decreased K+ flux (Pek-Scott & Lutz, 1998) and a reduction in the density of voltage-gated Na+ channels (Perez-Pinzon et al., 1992). In the isolated turtle cerebellum, 4 hours anoxia produced a 42% decrease in the density of voltagegated Na+ channels (Perez-Pinzon et al., 1992).

Anoxia-induced decreases in Ca2+ influx through the glutamate-responsive N-methyl-D- aspartate (NMDA) receptors of the turtle cerebrocortex have also been well described (Bickler &

314

Biology of Turtles

% Normalized NMDAR Currents

200

 

 

 

 

 

 

180

Normoxia

 

 

 

 

 

160

Anoxia

 

 

 

 

 

 

 

 

 

 

 

140

 

 

 

 

 

 

120

 

 

 

 

 

 

100

 

 

 

 

 

 

80

 

 

 

 

 

 

60

 

 

 

 

 

 

40

 

 

 

 

 

 

20

 

 

 

 

 

 

0

0

20

40

60

80

100

 

Time (min)

Figure 12.6  Normalized whole-cell NMDA currents in turtle cortical neurons. Data are mean ± SEM of 5 to 11 independent experiments. Reprinted from Shin et al. (2005), with permission from Elsevier.

Gallego, 1993; Shin et al., 2005) (Figure 12.6). Ionized calcium in the cerebrospinal fluid increases fiveto six-fold during prolonged anoxia, greatly increasing the potential for Ca2+ influx and neurotoxicity. In cortical sheets dissected out from turtles submerged in anoxic water (2 to 4°C) for 2 hours to 6 weeks, the Ca2+ influx via NMDA channels decreased 30 to 40% within the first 2 hours of anoxia and remained stable over 5 further weeks of anoxia (Bickler, 1998). As NMDA receptors are inhibited by both acidity (Traynelis & Cull-Candy, 1990) and high magnesium (Ascher & Nowak, 1987), and anoxia induces both a five-fold increase in Mg2+ and a significant drop in pH, these changes will partly offset the long-term increase in calcium (Bickler, 1998) but do not explain the decrease in NMDA receptor activity within 2 hours of anoxic submergence.

A recent study by Prentice et al. (2003) indicated one mechanism by which the downregulation of ion channels may occur, in demonstrating that transcription of the voltage-dependent K+ channel Kv1 is reversibly regulated by oxygen supply. Four hours of anoxia reduced Kv1 transcription to less than 20% of basal rates; mRNA levels were restored following 4 hours reoxygenation in vivo (Prentice et al., 2003). However, neither actin nor hypoxia-inducible factor (HIF-1) transcription were altered by 4 hours anoxia, indicating that the down-regulation of ion channel transcription is a targeted adaptation rather than a general effect of anoxia.

Of course, other mechanisms may also play a role in channel arrest, such as phosphorylation and dephosphorylation; covalent modification of proteins is one of the simplest methods to control activity, as occurs with metabolic enzymes (Brooks & Storey, 1988). Reversible protein phosphorylation has been shown to play a role in hypometabolism in a variety of enzyme systems and in different animal phyla, indicating its widespread phylogenetic conservation of the mechanism (MacDonald & Storey, 1999; Ramnanan & Storey, 2006). For example, NMDA receptor activity (the primary glutamate receptor of the CNS) is also decreased in the turtle brain during anoxia, associated with a decrease in Ca2+ permeability (Bickler, 1998; Bickler et al., 2000). Bickler and coworkers have recently described several mechanisms by which NMDA receptors are silenced that work at different times during anoxia in the turtle brain, with dephosphorylation occurring in the first few minutes of anoxia; over days to weeks, receptors are actually removed from the cell membrane (Bickler et al., 2000). However, none of these studies on ion channel activity examined regional differences in the brain, so it is unknown what role relative changes in channel arrest play in altering the activity of specific areas of the CNS.

The Physiology and Anatomy of Anoxia Tolerance in the Freshwater Turtle Brain

315

The decreased ion permeability reported in the turtle brain is paralleled by both a dramatic down-regulation of the EEG (Fernandes et al., 1997) and reductions in Na+/K+ ATPase activity (Hylland et al., 1997). During the initial hour of transition to the anoxic hypometabolic state, there is a dramatic decrease in brain electrical activity in turtles, though the pattern of changes suggests a systematic shutdown rather than simply a global depression of nervous activity (Fernandes et al., 1997). In essence, the turtle enters a reversible coma, though the significantly depressed EEG is regularly interrupted (every 0.5 to 2 min) by short bursts (3 to 15 sec) of mixed frequency activity, lending support to the hypothesis that some activities in the brain continue even in prolonged anoxia (Lutz & Milton, 2004). Again, the EEG work by Fernandes et al. (1997) was at the whole brain level, such that the relative activity of different regions, as suggested by metabolic indicators and ion channel distribution, could not be determined. It has been noted that turtle brainstem neurons in vitro have less spontaneous spike activity than is seen in homologous structures in mammals in vivo (Rosenberg & Ariel, 1990; Fan et al., 1995, 1997), though whether this is an effect of lower body temperatures or an adaptation to conserve energy is not known; clearly, lower spike activity would conserve energy and reduce excitotoxic damage. Buck and Bickler (1998) found that anoxia induces increases in the threshold potential of isolated cortical sheets in Chrysemys picta, leading to energy savings by increasing the synaptic input needed to trigger an action potential (“spike arrest”) and thus reducing electrical activity in the neuron.

However, depressed ion channel conductance and reduced electrical activity are reflected in alterations in Na+/K+ ATPase activity that are indeed different in individual regions of the turtle brain. Na+/K+ ATPase activity is reduced by 31% in the turtle cortex (telencephalon) after 24 hours anoxia and by 34% in the cerebellum (Hylland et al., 1997). Na+/K+ ATPase activity in the turtle brainstem is slightly (though not significantly) higher than in the telencephalon or cerebellum under basal conditions; activity also drops by 24 hours anoxia but the difference is not significant (Figure 12.7). Despite the lower densities of ion channels, measures of Na+/K+ ATPase activity that are initially higher and decrease less in the brain stem support the conclusions suggested by the distribution of the enzymes of anaerobic metabolism—that is, that the brainstem in the anoxic turtle remains active. Continued activity may reflect a role in maintaining basal physiological functions such as circulatory control, or perhaps this area is responsible for taking the turtle out of its comatose state upon reoxygenation (Hylland et al., 1997). As an aside, Na+/K+ ATPase activity in the brain of the anoxia-tolerant Crucian carp (Carassius carassius) does not decrease in anoxia (Hylland et al., 1997); there is also evidence that rates of ion flux are maintained (Johansson & Nilsson, 1995).

Activity

g

 

)

 

–1

ATPase

–1

min

+

(µmol

-

 

+

 

/K

 

Na

 

 

Normoxia

 

 

3

Anoxia

 

 

Recovery

 

 

 

 

 

2

*

*

 

 

 

 

1

 

 

 

0

Telencephalon

Cerebellum

Brain Stem

 

Figure 12.7  Na+/K+ ATPase activity in three parts of the normoxic, 24 anoxic, and recovered T. scripta brain. Data are mean ± SEM of n = 6 animals. The asterisk indicates P < 0.05. Reprinted from Hylland et al. (1997), with permission from Elsevier.

Соседние файлы в предмете [НЕСОРТИРОВАННОЕ]