Oxidative Injury and Apoptosis of Dorsal Root Ganglion ... - Diabetes

3 downloads 0 Views 390KB Size Report
Greene DA, Sima AA, Stevens MJ, Feldman EL, Lattimer SA: Complica- .... Hinokio Y, Suzuki S, Hirai M, Chiba M, Hirai A, Toyota T: Oxidative DNA damage in ...
Oxidative Injury and Apoptosis of Dorsal Root Ganglion Neurons in Chronic Experimental Diabetic Neuropathy Ann M. Schmeichel, James D. Schmelzer, and Phillip A. Low

We evaluated the effects of chronic hyperglycemia on L5 dorsal root ganglion (DRG) neurons using immunohistochemical and electrophysiologic techniques for evidence of oxidative injury. Experimental diabetic neuropathy was induced by streptozotocin. To evaluate the pathogenesis of the neuropathy, we studied peripheral nerve after 1, 3, and 12 months of diabetes. Electrophysiologic abnormalities were present from the first month and persisted over 12 months. 8-Hydroxy-2ⴕdeoxyguanosine labeling was significantly increased at all time points in DRG neurons, indicating oxidative injury. Caspase-3 labeling was significantly increased at all three time points, indicating commitment to the efferent limb of the apoptotic pathway. Apoptosis was confirmed by a significant increase in the percentage of neurons undergoing apoptosis at 1 month (8%), 3 months (7%), and 12 months (11%). These findings support the concept that oxidative stress leads to oxidative injury of DRG neurons, with mitochondrium as a specific target, leading to impaired mitochondrial function and apoptosis, manifested clinically as a predominantly sensory neuropathy. Diabetes 52:165–171, 2003

T

he precise pathogenesis of diabetic neuropathy is unknown. However, a number of putative pathophysiologic mechanisms exist. Hyperglycemia is reported to result in oxidative stress (1,2), polyol pathway overactivity (3), increased advanced glycation end products (4), nerve hypoxia/ischemia (1), deficiency of ␥-linolenic acid (5), increase in protein kinase C, especially ␤-isoform (6), and growth factor(s) deficiency (7). These pathways all converge in producing oxidative stress (8). Oxidative stress seems to be more severe at the dorsal root ganglion (DRG) than at the nerve (9), and recent findings that beyond 6 months of diabetes, diabetic rats develop florid radicular pathology (10) and vacuolar degeneration of DRG neurons (11) have led to the present hypothesis that a primary target of diabetic neural complications is the sensory neuron (1). We wished to evaluate, in chronic experimental diabetic neuropathy (EDN), the role of oxidative injury as a function of duration of diabetes on DRG neurons. To achieve

From the Department of Neurology, Mayo Clinic, Rochester, Minnesota. Address correspondence and reprint requests to Phillip A. Low, Department of Neurology, Mayo Clinic, 200 First St. SW, Rochester, MN 55905. E-mail: [email protected]. Received for publication 25 March 2002 and accepted in revised form 23 September 2002. 8-OHdG, 8-hydroxy-2⬘-deoxyguanosine; DRG, dorsal root ganglion; EDN, experimental diabetic neuropathy; GSH, reduced glutathione; STZ, streptozotocin; TUNEL, TdT-mediated dUTP-biotin nick end labeling. DIABETES, VOL. 52, JANUARY 2003

these goals, we undertook immunohistochemical studies of oxidative injury, mitochondrial function, caspase-3, and TdT-mediated dUTP-biotin nick end labeling (TUNEL) of DRG neurons. RESEARCH DESIGN AND METHODS Animals. Male Sprague-Dawley rats that weighed 250 ⫾ 5 g at the beginning of the study were used. Experimental diabetes was produced by the intraperitoneal injection of streptozotocin (STZ) in 0.05 mol/l citrate buffer at pH 4.5 (65 mg/ml; dose 1.32 ml/kg). The control group received an intraperitoneal injection of citrate buffer alone. Control and diabetic rats had free access to Purina Rodent Laboratory Chow and water. They did not receive insulin treatment. The rats were accepted as diabetic when their fasting blood glucose exceeded 16.7 mmol/l 3 days after injection of STZ and remained ⬎16.7 mmol/l at the time when they were killed. Several groups of rats were used to accomplish these studies. Duration of diabetes was different for three groups: 1. EDN, duration 1 month (n ⫽ 8); age- and gender-matched control littermates (n ⫽ 8) 2. EDN, 3 months (n ⫽ 8); age- and gender-matched control littermates (n ⫽ 8) 3. EDN, 12 months (n ⫽ 9); age- and gender-matched control littermates (n ⫽ 9) 4. EDN, 12 months (n ⫽ 5); age- and gender-matched control littermates (n ⫽ 5). Nerve conduction studies were done on groups 1– 4 and one EDN and one control from group 4. Groups 1–3 were used for immunohistochemical studies (8-hydroxy-2⬘-deoxyguanosine [8-OHdG] and caspase-3) and for TUNEL. Group 4 was used for histochemical studies of cytochrome oxidase reactivity. Immunohistochemistry. We used methods that are standard in our laboratory. Nerve was fixed in situ for 30 min, flash frozen, and cut into 10-␮m sections. Endogenous peroxidase, avidin, and biotin were blocked appropriately, and a 0.05% Tween-20 in phosphate buffer was used. Sections were incubated in the primary antibody for 1 h. After washing, specific immunoreactivity was visualized using a Vectastain Elite ABC kit (Vector Labs, Burlingame, CA) with diaminobenzidine as the chromagen. Negative controls were generated by omission of the primary antibodies and the use of a blocking peptide. Specificity for 8-OHdG was additionally tested by a pretreatment with an RNase solution that did not reduce staining and a DNase-1 (Boehringer Mannheim) treatment that reduced staining to background staining intensity. We used the following antibodies: anti– 8-OHdG (mouse monoclonal, 1:40; Genox, Baltimore, MD) and anti– caspase-3 (rabbit polyclonal, 1:100; Santa Cruz Biotechnology, Santa Cruz, CA). Staining for cytochrome oxidase was performed using methods from Seligman et al. (12). A TUNEL labeling kit (Oncogene, Boston, MA) was used for labeling apoptotic nuclei. Positive and negative controls were generated per kit protocol. An additional positive control was also used by use of DNase-1 to “nick” the DNA. Three levels of DRG were assayed per animal at 100-␮m intervals. An average of 100 neurons/1 month, 104 neurons/3 months, and 125 neurons/12 months were graded per animal. Evaluation. For 8-OHdG and caspase-3, we used a semiquantitative grading system, because cells varied by the severity and presence of staining. To a first approximation, the grades were as follows: 0, light staining and only affecting ⱕ5% of neurons; 1, staining affecting 5–10% of neurons; 2, staining affecting 10 –20% of neurons; 3, staining affecting ⱖ20% of neurons. In the evaluation of TUNEL positivity, neurons were considered positive when the nuclei stained positively and showed chromatin clumping. The number of neurons that fulfilled the criteria was expressed as a percentage. 165

IMCL AND INSULIN RESISTANCE IN ZDF RATS

TABLE 1 Electrophysiology on diabetic (EDN) and aged-matched control (CON) rats

Group

n

CON 1 mo EDN 1 mo

8 8

CON 3 mo EDN 3 mo

8 8

CON 12 mo EDN 12 mo

9 9

Digital nerve Ampl CV (m/s) (␮V)

Sciatic-tibial Ampl CV (m/s) (mV)

Caudal NAP Ampl CV (m/s) (␮V)

Caudal CMAP Ampl CV (m/s) (mV)

42.8 ⫾ 0.7 36.4 ⫾ 0.8 * 53.1 ⫾ 1.1 45.2 ⫾ 0.9 * 47.1 ⫾ 2.0 40.4 ⫾ 1.4 †

45.3 ⫾ 1.6 39.8 ⫾ 1.3 † 49.4 ⫾ 1.0 41.9 ⫾ 0.9 * 51.1 ⫾ 1.2 46.0 ⫾ 1.1 ‡

45.7 ⫾ 1.0 38.9 ⫾ 1.1 ‡ 53.6 ⫾ 1.0 45.5 ⫾ 0.9 * 63.5 ⫾ 0.9 49.1 ⫾ 1.1 *

31.5 ⫾ 1.1 27.2 ⫾ 1.6 † 37.5 ⫾ 1.1 35.6 ⫾ 1.0

19.5 ⫾ 1.3 23.0 ⫾ 2.1 46.1 ⫾ 2.7 55.1 ⫾ 5.1 24.5 ⫾ 2.2 33.0 ⫾ 2.1 †

10.9 ⫾ 0.2 8.8 ⫾ 0.8 † 9.9 ⫾ 0.4 7.3 ⫾ 0.7 ‡ 5.0 ⫾ 1.0 6.1 ⫾ 1.1

78 ⫾ 4 89 ⫾ 5 142 ⫾ 5 140 ⫾ 9 188 ⫾ 8 201 ⫾ 13

46.4 ⫾ 0.3 40.7 ⫾ 1.8 ‡

9.5 ⫾ 0.4 5.6 ⫾ 0.6 * 8.3 ⫾ 0.6 4.6 ⫾ 0.6 ‡ 6.7 ⫾ 0.3 5.1 ⫾ 0.9

Data are means ⫾ SE. *P ⬍ 0.001; †P ⬍ 0.05; ‡P ⬍ 0.01. Ampl, amplitude; CMAP, compound muscle action potential; CV, conduction velocity; NAP, nerve action potential.

Electrophysiology. We used techniques that are standard for our laboratory (13,14). Sensory nerve conduction velocity of digital and caudal nerves was measured using fine stainless steel near-nerve–stimulating and –recording electrodes. Motor nerve conduction velocity was measured in the sciatic-tibial and caudal nerves. The compound muscle action potentials were recorded from the pair of fine stainless steel electrodes in the dorsum of hind paw while stimulated by another pair of electrodes at the level of the sciatic notch and ankle. Recordings were made at 35°C, amplified 1,000 times, stored on computer disks, and analyzed off-line using a Nicolet 4094 digital oscilloscope (Nicolet Instruments, Madison, WI) with associated stimulators and stimulus isolation units. Statistical methods. Means of continuous variables were compared between groups using ungrouped two-tailed t test. Nonparametric relationships were examined with Mann-Whitney U tests. Univariate associations between variables were performed using Spearman rank correlations. Data were expressed as mean ⫾ SE unless otherwise stated. P ⬍ 0.05 was considered significant.

RESULTS

Animals. Final weights for control rats at 1, 3, and 12 months were 401 ⫾ 7, 455 ⫾ 10, and 595 ⫾ 12 g, respectively, and were significantly (P ⬍ 0.001) greater than those of diabetic rats with corresponding values of 243 ⫾ 12, 214 ⫾ 15, and 294 ⫾ 26 g. Blood glucose values for control rats at 1, 3, and 12 months were 9.9 ⫾ 0.3, 8.7 ⫾ 0.4, and 8.3 ⫾ 0.4 mmol/l, respectively, and were significantly (P ⬍ 0.001) lower than values in diabetic rats of 39.4 ⫾ 1.4, 33.3 ⫾ 1.6, and 30.9 ⫾ 1.2 mmol/l. Glycosylated hemoglobin values in control rats at these time points were 3.5 ⫾ 0.2, 3.5 ⫾ 0.1, and 5.0 ⫾ 0.1%, significantly (P ⬍ 0.001) lower than corresponding values in diabetic rats of 15.4 ⫾ 0.5, 15.8 ⫾ 0.4, and 15.2 ⫾ 0.5%. Electrophysiology. At 1 month, sensory conduction velocities were consistently and highly significantly increased (digital P ⬍ 0.001; caudal P ⬍ 0.01), whereas corresponding motor conduction velocities were less significant (sciatic-tibial P ⫽ 0.02) or marginal (caudal P ⫽ 0.045) in controls. At 3 months, both motor and sensory conduction velocities were highly significantly reduced for the limb nerves of EDN (digital sensory P ⬍ 0.001; sciatic-tibial motor P ⬍ 0.001). However, greater involvement of sensory conduction continued to be seen for caudal nerve, where sensory deficit was highly significant (P ⬍ 0.0001) and motor fibers were not significantly reduced in EDN. At 12 months, differences between diabetic and control rats had lessened for limb nerves but persisted in caudal nerve. Both limb and caudal nerves had reduced motor and sensory conduction velocities. For all 166

three time points, there was a more variable reduction in amplitudes (Table 1). Histochemistry. Patchy areas of cytochrome oxidase staining loss, indicating a loss of enzyme activity, were seen (Fig. 1). Succinic dehydrogenase and trichrome staining did not show major differences between diabetic and control rats (data not shown). 8-OHdG. Excess labeling with 8-OHdG was seen at all three time points evaluated (Fig. 2). Immunolabeling was nuclear in a granular multifocal distribution. Labeling index was evaluated semiquantitatively on the basis of the number and intensity of immunolabeling (Fig. 3). Immunolabeling was significantly increased (Mann-Whitney U test) at 1 month (EDN 0.81 ⫾ 0.19 vs. controls 0.19 ⫾ 0.13; P ⬍ 0.001), 3 months (EDN 1.20 ⫾ 0.24 vs. controls 0.00 ⫾ 0.00; P ⬍ 0.001), and 12 months (EDN 1.72 ⫾ 0.29 vs. controls 0.57 ⫾ 0.20; P ⬍ 0.001). There was a progressive increase with duration of diabetes. Caspase-3. An increased number of DRG neurons stained positive for caspase-3 at all three time points evaluated (Fig. 4). Labeling index was evaluated semiquantitatively (Fig. 5) and was significantly increased (Mann-Whitney U test) at 1 month (EDN 1.50 ⫾ 0.19 vs. controls 0.19 ⫾ 0.09;

FIG. 1. Histochemical labeling of cytochrome oxidase activity in DRG neurons of EDN (right; duration 12 months) and age-matched control (left). Arrows indicate areas of reduced activity. Bar ⴝ 50 ␮m. DIABETES, VOL. 52, JANUARY 2003

A.M. SCHMEICHEL, J.D. SCHMELZER, AND P.A. LOW

FIG. 2. Control (left) and diabetic (right) DRG neurons at 1 (top), 3 (middle), and 12 months (bottom) of diabetes or age-matched controls. Labeling with 8-OHdG is increased in EDN when compared with control DRG. Bar ⴝ 20 um.

P ⬍ 0.001), 3 months (EDN 1.21 ⫾ 0.57 vs. controls 0.07 ⫾ 0.19; P ⬍ 0.001), and 12 months (EDN 1.40 ⫾ 0.10 vs. controls 0.60 ⫾ 0.20; P ⬍ 0.001). TUNEL. The number of TUNEL-positive DRG neurons was increased at all three time points evaluated (Fig. 6). TUNEL positivity expressed as a percentage of positive DRG neurons was significantly increased (unpaired t test, two-tailed) at 1 month (EDN 8.00 ⫾ 1.00% vs controls 2.00 ⫾ 0.00; P ⬍ 0.001), 3 months (EDN 7.00 ⫾ 1.00 vs. controls 1.00 ⫾ 0.00; P ⬍ 0.001), and 12 months (EDN 11.00 ⫾ 2.00 vs. controls 1.00 ⫾ 0.00; P ⬍ 0.001). We compared the percentage of small versus large neurons, as previously defined (15), affected by apoptosis. The percentage of small versus large apoptotic DRG neurons at 1 month was 7% large and 10% small, (difference between large and small, P ⫽ 0.05). Corresponding percentages at 3 months were 6% large and 8% small (NS), and at 12 months were 11% large and 12% small (NS). Correlations. The associations between the index of oxidative stress (8-OHdG) and apoptosis (TUNEL, caspase-3) were evaluated using Spearman correlation statistics. At 1 month, 8-OHdG significantly correlated with caspase-3 (R ⫽ 0.71; P ⬍ 0.01) and TUNEL (R ⫽ 0.80; P ⬍ 0.001). At 3 months, the significant association persisted between 8-OHdG, caspase-3 (R ⫽ 0.93; P ⬍ 0.001), and TUNEL (R ⫽ 0.74; P ⬍ 0.01). However, by 12 months, the association was not significant for either caspase-3 (R ⫽ 0.34; NS) or TUNEL (R ⫽ 0.41; NS). For the relationship between caspase-3 and TUNEL, correlation was highly DIABETES, VOL. 52, JANUARY 2003

FIG. 3. Semiquantitative immunohistochemical labeling with 8-OHdG of control and EDN DRG at 1 (A), 3 (B), and 12 months (C) of diabetes. ***P < 0.001, Mann-Whitney U test.

significant at 1 month (R ⫽ 0.80; P ⬍ 0.001) and 3 months (R ⫽ 0.64; P ⫽ 0.01) but was no longer significant by 12 months (R ⫽ 0.45; NS). DISCUSSION

There are several findings of this study. First, there is early development of neuropathy electrophysiologically, with more pronounced involvement of sensory fibers followed by a combined sensorimotor neuropathy. In the chronic model (12 months), mitochondrial function may be impaired (cytochrome oxidase stain). Pathophysiologically, there is oxidative injury to DRG neurons early with persistence and some accentuation with increasing age. Oxidative DNA injury is associated with commitment to 167

IMCL AND INSULIN RESISTANCE IN ZDF RATS

FIG. 4. Control (left) and diabetic (right) DRG neurons at 1 (top), 3 (middle), and 12 months (bottom) of diabetes or age-matched controls. Immunolabeling with antibody to caspase 3. Diabetic DRG show increased caspase-3 expression. Bar ⴝ 20 um.

apoptosis as indicated by caspase-3 expression and TUNEL positivity. Although motor and sensory conduction abnormalities are present in EDN, the most consistent abnormalities in chronic EDN are sensory, and it is sometimes the only abnormality in chronic EDN (13). Early electrophysiologic studies have focused on motor conduction abnormalities (16,17), because sensory recordings are technically more demanding and require miniaturized recording electrodes. The present study demonstrates more severe involvement of sensory conduction at the earlier time point of 1 month with involvement of both motor and sensory conduction at later time points. The presence of combined motor and sensory deficits and their reversibility with treatment (5,6,9,13,18) and the close correlation between the deficits in conduction and perfusion suggest that one component of the neuropathy, especially short term, is related to microvascular pathology. We suggest that the chronic distal sensory deficits seen in chronic progressive EDN might be more closely related to DRG pathology demonstrated in this study and in Kishi et al. (15). There is vacuolar and pigmentary degeneration of DRG neurons in very chronic EDN (14), in association with florid pathological changes of dorsal and ventral roots. Previously, demyelination of ventral root had been described in aged rats, but the sensory roots were unaffected (19). Lipid peroxidation effects are greater in nerve root and DRG because the blood-nerve and perineurial barriers are lower at these sites (20) and may contribute to the 168

FIG. 5. Semiquantitative immunohistochemical labeling with caspase-3 of control and EDN DRG at 1 (A), 3 (B), and 12 months (C) of diabetes. **P < 0.001; ***P < 0.001, Mann-Whitney U test.

development of a radicular myelinopathy (14). Other investigators have confirmed these findings of mitochondrial pathology, including ballooning of mitochondria and disruption of the internal cristae as a result of the neurotoxic effects of glucose (2,21–23). Oxidative stress depends on a balance among free radical generation, pro-oxidant status, and free radical defenses (24). Changes in all three areas occur in EDN (1). Peripheral nerve antioxidant defenses are very low compared with brain and liver (25,26); normal nerve reduced glutathione (GSH) and glutathione containing enzymatic scavengers (glutathione peroxidase and reductase) are only ⬃10% that of brain (25,26). Increased generation of reduced oxygen species occurs in both human and EDN DIABETES, VOL. 52, JANUARY 2003

A.M. SCHMEICHEL, J.D. SCHMELZER, AND P.A. LOW

FIG. 6. TUNEL positivity for DRG neurons. Control (left) and diabetic (right) DRG neurons at 1 (top), 3 (middle), and 12 months (bottom) of diabetes or age-matched controls. Diabetic DRG nuclei show clumping and increased TUNEL positivity. Bar ⴝ 20 um.

(1,27). Diabetic peripheral nerve has increased conjugated dienes (9,28), hydroperoxides, and GSH (9). These changes are more consistent in lumbar dorsal root ganglia and superior cervical ganglion than in nerve (9). The most sensitive index of ongoing oxidative stress is a reduction in GSH (9,13). Pro-oxidant status is altered: diabetic sciatic nerve has increased lipolysis (18) and altered reducing equivalents (29). The diabetic state results in reduced superoxide dismutase (28) and a further reduction in GSH (9) and glutathione peroxidase, the limiting antioxidant enzymatic scavenger in the diabetic mouse (30). These earlier studies demonstrating footprints of oxidative stress have recently been supplanted by immunocytochemical evidence of DNA damage and cellular localization (31,32). Reduced oxygen species cause irreversible DNA damage to specific proteins. In recent years, antibodies have been generated against modified structures specific for reduced oxygen species–induced damage (31,33). The epitopes include 8-hydroxy-2⬘-deoxyguanosine and 4-hydroxy-2-nonenal-modified proteins (34). Urinary 8-hydroxy-2⬘-deoxyguanosine has been reported to be increased in human diabetes (35,36). 8-OHdG staining in the present study confirms oxidative injury to DRG neurons, which, with their large complement of mitochondria and high oxidative metabolism, compose a susceptible target to oxidative injury. The vacuoles observed in our study may be related to dysfunctional mitochondria (2; current study). Mitochondrial DNA is unusually susceptible to oxidative damage (37), and these DIABETES, VOL. 52, JANUARY 2003

FIG. 7. Percentage of DRG neurons showing TUNEL positivity of control and EDN DRG at 1 (A), 3(B), and 12 months (C) of diabetes. ***P < 0.001, unpaired, two-tailed t test.

dysfunctional mitochondria have increased reduced oxygen species leakage (38). A vicious cycle of oxidative damage to inner membrane proteins (of mitochondria), leading to imbalances in the electron transport chain, resulting in increased superoxide and hydrogen peroxide production, which in turn further damages membrane proteins, has been suggested (39) as a major pathogenetic mechanism in other states such as aging (40). The highly significant association between 8-OHdG and caspase-3 immunoreactivity and TUNEL positivity at 1 and 3 months support the notion that oxidative injury is responsible for apoptosis. Support for the importance of apoptosis derives from the close correlation between caspase-3 and TUNEL positivity. The loss of the association between oxidative injury and indexes of apoptosis at 169

IMCL AND INSULIN RESISTANCE IN ZDF RATS

FIG. 8. Pathogenetic model of DRG neuronopathy in EDN.

12 months and the progressive loosening of the relationship between TUNEL and caspase-3 immunoreactivity are of interest. The declining correlations with chronicity of diabetes suggest the evolution of more complex interactions. It has been suggested that apoptosis in certain neurological disorders can be much delayed (apoptosis lente) (41). At short time points, every neuron that enters the efferent limb of the apoptotic pathway also undergoes DNA fragmentation, and hence there is a highly significant correlation between TUNEL positivity and caspase-3 immunoreactivity. With increasing duration of disease, because of apoptosis lente, there is less concordance between caspase-3 immunoreactivity and TUNEL positivity as a result of the variable rate at which TUNEL-positive neurons disappear. We have previously reported a selective loss of the largest DRG neurons (15). The present study lends some support to that observation. Normally large neurons compose 28% (42) or 29% (15), whereas small neurons compose ⬎70% of neurons. When we separately analyzed the percentage of large versus small DRG neurons affected by apoptosis, large were, especially with more chronic diabetes, as frequently affected as small, so the larger neurons do seem to be particularly vulnerable. Reduced cytochrome oxidase activity, although preliminary, supplements the morphological changes in providing additional evidence of dysfunctional mitochondrium. These findings interdigitate well with the in vitro and in vivo studies of recent-onset diabetes by the Michigan group. They demonstrate the reduction of mitochondrial membrane potential, leakage of cytochrome c, and caspase-3 activation (2,21–23). Our study links these changes with oxidative injury and extends the changes from 1 to 12 months of diabetes. Caspase-3 activation commits the neuron to the efferent limb of the apoptotic pathway. TUNEL positivity at all three time points indicates that a small population of sensory neurons is committed to apoptosis. These findings 170

agree with our recent observations of a loss of DRG neurons in EDN, affecting primarily the largest neurons (15). In that study, small DRG neurons were unaffected, but there was a 41% attrition of the largest DRG neurons. These findings now permit a reasonable pathogenetic model of sensory neuropathy of EDN (Fig. 7). Hyperglycemia results in oxidative stress by numerous pathways described in the introduction. Of particular importance to mitochondrial function is the reduction in GSH (43), reduction in NADPH, and increase in NADH (22). The generation of reduced oxygen species, especially superoxide, and peroxynitrite results in single-strand break of DNA (44). Pivotal to the pathogenesis of apoptosis in DRG neurons is the release of cytochrome c from the outer mitochondrial membrane. The egress of cytochrome c requires the opening of mitochondrial pores (44). There is good evidence for this event. Mitochondrial membrane potential falls in hyperglycemic mitochondrium (45) associated with swelling of the organelle. The precise mechanism of pore opening is not known. Most likely, mechanisms are a combination of an alteration in the balance between the protective and proapoptotic components of the bcl family and a reduction in GSH (33). Cytochrome c release has been demonstrated (45), and this binds to Apaf 1, which then recruits procaspase-9 (apoptosome complex) (46). This oligomerization and autoactivation of procaspase-9 to caspase-9 is followed by proteolytic cleavage of procaspase-3 (and other executioner procaspases) activating the executioner caspase-3. It is consistently increased in experimental diabetes (2,21). The net effect is apoptosis of DRG neurons (Fig. 8). ACKNOWLEDGMENTS

This study was supported by grants from the National Institutes of Health (NS22352, NS32352, NS39722, and MO1 RR00585 to P.A.L.), the Juvenile Diabetes Research Foundation, and Mayo Funds. REFERENCES 1. Low PA, Lagerlund TD, McManis PG: Nerve blood flow and oxygen delivery in normal, diabetic, and ischemic neuropathy. Int Rev Neurobiol 31:355– 438, 1989 2. Russell JW, Sullivan KA, Windebank AJ, Herrmann DN, Feldman EL: Neurons undergo apoptosis in animal and cell culture models of diabetes. Neurobiol Dis 6:347–363, 1999 3. Greene DA, Sima AA, Stevens MJ, Feldman EL, Lattimer SA: Complications: neuropathy, pathogenetic considerations. Diabetes Care 15:1902– 1925, 1992 4. Brownlee M, Cerami A, Vlassara H: Advanced products of nonenzymatic glycosylation and the pathogenesis of diabetic vascular disease. Diabetes Metab Rev 4:437– 451, 1988 5. Cameron NE, Cotter MA: Effects of evening primrose oil treatment on sciatic nerve blood flow and endoneurial oxygen tension in streptozotocindiabetic rats. Acta Diabetol 31:220 –225, 1994 6. Cameron NE, Cotter MA, Jack AM, Basso MD, Hohman TC: Protein kinase C effects on nerve function, perfusion, Na(⫹), K(⫹)-ATPase activity and glutathione content in diabetic rats. Diabetologia 42:1120 –1130, 1999 7. Hellweg R, Hartung HD: Endogenous levels of nerve growth factor (NGF) are altered in experimental diabetes mellitus: a possible role for NGF in the pathogenesis of diabetic neuropathy. J Neurosci Res 26:258 –267, 1990 8. Low PA: Pathogenesis of diabetic neuropathy. In Joslin’s Diabetes Mellitus. Kahn CR, Weir GC, King GL, Moses AC, Jacobson AM, Smith RJ, Eds. Philadelphia, Lippincott Williams & Wilkins, 2001. In press 9. Nickander KK, Schmelzer JD, Rohwer DA, Low PA: Effect of ␣-tocopherol deficiency on indices of oxidative stress in normal and diabetic peripheral nerve. J Neurol Sci 126:6 –14, 1994 DIABETES, VOL. 52, JANUARY 2003

A.M. SCHMEICHEL, J.D. SCHMELZER, AND P.A. LOW

10. Tamura E, Parry GJ: Severe radicular pathology in rats with longstanding diabetes. J Neurol Sci 127:29 –35, 1994 11. Sasaki H, Schmelzer JD, Zollman PJ, Low PA: Neuropathology and blood flow of nerve, spinal roots and dorsal root ganglia in longstanding diabetic rats. Acta Neuropathol 93:118 –128, 1997 12. Seligman AM, Karnovsky MJ, Wasserkrug HL, Hanker JS: Nondroplet ultrastructural demonstration of cytochrome oxidase activity with a polymerizing osmiophilic reagent, diaminobenzidine (DAB). J Cell Biol 38:1– 14, 1968 13. Nagamatsu M, Nickander KK, Schmelzer JD, Raya A, Wittrock DA, Tritschler H, Low PA: Lipoic acid improves nerve blood flow, reduces oxidative stress, and improves distal nerve conduction in experimental diabetic neuropathy. Diabetes Care 18:1160 –1167, 1995 14. Sasaki H, Kihara M, Zollman PJ, Nickander KK, Smithson IL, Schmelzer JD, Willner CL, Benarroch EE, Low PA: Chronic constriction model of rat sciatic nerve: nerve blood flow, morphologic and biochemical alterations. Acta Neuropathol 93:62–70, 1997 15. Kishi M, Tanabe J, Schmelzer JD, Low PA: Morphometry of dorsal root ganglion in chronic experimental diabetic neuropathy. Diabetes 51:819 – 824, 2002 16. Eliasson SG: Nerve conduction changes in experimental diabetes. J Clin Invest 43:2353–2358, 1964 17. Gillon KR, Hawthorne JN: Sorbitol, inositol and nerve conduction in diabetes. Life Sci 32:1943–1947, 1983 18. Yao JK, Low PA: Improvement of endoneurial lipid abnormalities in experimental diabetic neuropathy by oxygen modification. Brain Res 362:362–365, 1986 19. Knox CA, Kokmen E, Dyck PJ: Morphometric alteration of rat myelinated fibers with aging. J Neuropathol Exp Neurol 48:119 –139, 1989 20. Olsson Y: Topographical differences in the vascular permeability of the peripheral nervous system. Acta Neuropathol 10:26 –33, 1968 21. Feldman EL, Russell JW, Sullivan KA, Golovoy D: New insights into the pathogenesis of diabetic neuropathy. Curr Opin Neurol 12:553–563, 1999 22. Greene DA, Stevens MJ, Obrosova I, Feldman EL: Glucose-induced oxidative stress and programmed cell death in diabetic neuropathy. Eur J Pharmacol 375:217–223, 1999 23. Greene DA, Obrosova IG, Stevens MJ, Feldman EL: Pathways of glucosemediated oxidative stress in diabetic neuropathy. In Antioxidants in Diabetes Management. Packer L, Rosen P, Tritschler HJ, King GL, Azzi A, Eds. New York, Marcel Dekker, 2000, p. 111–119 24. Halliwell B, Gutteridge JMC: Oxygen radicals and the nervous system. Trends Neurosci 8:22–26, 1985 25. Romero FJ, Segura-Aguilar J, Monsalve E, Hermenegildo C, Nies E, Puertas FJ, Roma J: Antioxidant and glutathione-related enzymatic activities in rat sciatic nerve. Neurotoxicol Teratol 12:603– 605, 1990 26. Romero FJ, Monsalve E, Hermenegildo C, Puertas FJ, Higueras V, Nies E, Segura-Aguilar J, Roma J: Oxygen toxicity in the nervous tissue: comparison of the antioxidant defense of rat brain and sciatic nerve. Neurochem Res 16:157–161, 1991 27. Low PA, Suarez GA: Diabetic neuropathies. Br J Pharmacol 115:373–379, 1995 28. Low PA, Nickander KK: Oxygen free radical effects in sciatic nerve in experimental diabetes. Diabetes 40:873– 877, 1991

DIABETES, VOL. 52, JANUARY 2003

29. Low PA, Ward K, Schmelzer JD, Brimijoin S: Ischemic conduction failure and energy metabolism in experimental diabetic neuropathy. Am J Physiol 248:E457–E462, 1985 30. Hermenegildo C, Raya A, Roma J, Romero FJ: Decreased glutathione peroxidase activity in sciatic nerve of alloxan-induced diabetic mice and its correlation with blood glucose levels. Neurochem Res 18:893– 896, 1993 31. Toyokuni S: Reactive oxygen species-induced molecular damage and its application in pathology. Pathol Int 49:91–102, 1999 32. Kruman II, Mattson MP: Pivotal role of mitochondrial calcium uptake in neural cell apoptosis and necrosis. J Neurochem 72:529 –540, 1999 33. Ji C, Amarnath V, Pietenpol JA, Marnett LJ: 4-hydroxynonenal induces apoptosis via caspase-3 activation and cytochrome c release. Chem Res Toxicol 14:1090 –1096, 2001 34. Toyokuni S, Tanaka T, Hattori Y, Nishiyama Y, Yoshida A, Uchida K, Hiai H, Ochi H, Osawa T: Quantitative immunohistochemical determination of 8-hydroxy-2⬘-deoxyguanosine by a monoclonal antibody N45.1: its application to ferric nitrilotriacetate-induced renal carcinogenesis model. Lab Invest 76:365–374, 1997 35. Leinonen J, Lehtimaki T, Toyokuni S, Okada K, Tanaka T, Hiai H, Ochi H, Laippala P, Rantalaiho V, Wirta O, Pasternack A, Alho H: New biomarker evidence of oxidative DNA damage in patients with non-insulin-dependent diabetes mellitus. FEBS Lett 417:150 –152, 1997 36. Hinokio Y, Suzuki S, Hirai M, Chiba M, Hirai A, Toyota T: Oxidative DNA damage in diabetes mellitus: its association with diabetic complications. Diabetologia 42:995–998, 1999 37. Richter C, Park JW, Ames BN: Normal oxidative damage to mitochondrial and nuclear DNA is extensive. Proc Natl Acad Sci U S A 85:6465– 6467, 1988 38. Turrens JF, Boveris A: Generation of superoxide anion by the NADH dehydrogenase of bovine heart mitochondria. Biochem J 191:421– 427, 1980 39. Shigenaga MK, Hagen TM, Ames BN: Oxidative damage and mitochondrial decay in aging. Proc Natl Acad Sci U S A 91:10771–10778, 1994 40. Wilson PD, Franks LM: The effect of age on mitochondrial ultrastructure and enzymes. Adv Exp Med Biol 53:171–183, 1975 41. Broccolini A, Engel WK, Askanas V: Localization of survival motor neuron protein in human apoptotic-like and regenerating muscle fibers, and neuromuscular junctions. Neuroreport 10:1637–1641, 1999 42. Tandrup T: A method for unbiased and efficient estimation of number and mean volume of specified neuron subtypes in rat dorsal root ganglion. J Comp Neurol 329:269 –276, 1993 43. Lee AY, Chung SS: Contributions of polyol pathway to oxidative stress in diabetic cataract. FASEB J 13:23–30, 1999 44. Soriano FG, Virag L, Jagtap P, Szabo E, Mabley JG, Liaudet L, Marton A, Hoyt DG, Murthy KGK, Salzman AL, Southan GJ, Szabo C: Diabetic endothelial dysfunction: the role of poly(ADP-ribose) polymerase activation. Nat Med 7:108 –113, 2001 45. Srinivasan S, Stevens M, Wiley JW: Diabetic peripheral neuropathy: evidence for apoptosis and associated mitochondrial dysfunction. Diabetes 49:1932–1938, 2000 46. Hengartner MO: The biochemistry of apoptosis. Nature 407:770 –776, 2000

171