Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Exp Physiol. Author manuscript; available in PMC 2020 Nov 4.
Published in final edited form as:
PMCID: PMC7640985
NIHMSID: NIHMS956381
PMID: 29604234

Increased skeletal muscle mitochondrial free radical production in peripheral arterial disease despite preserved mitochondrial respiratory capacity

Abstract

Skeletal muscle mitochondrial dysfunction, which is not fully explained by disease-related arterial occlusion, has been implicated in the pathophysiology of peripheral arterial disease (PAD). Therefore, this study comprehensively assessed mitochondrial respiratory function in biopsies from the gastrocnemius of 10 patients with PAD (Fontaine Stage II) and 12 healthy controls (HC). Intramuscular and systemic inflammation, mitochondrial-derived reactive oxygen species (ROS) production, and oxidative stress were also assessed to better understand the mechanisms responsible for the proposed PAD-induced mitochondrial dysfunction. Interestingly, mitochondrial respiratory capacity, assessed as complex I (CI) and complex II (CII)-driven State 3 respiration, measured separately and in combination (State 3:CI+II), revealed no difference between the patients with PAD and the HC. However, mitochondrial-derived ROS production was significantly elevated in PAD (HC: 1.0 ± 0.9; PAD: 4.3 ± 1.0 AU/mg tissue). Furthermore, patients with PAD exhibited significantly greater concentrations of the pro-inflammatory markers, tumor necrosis factor (TNF-α) in plasma (HC: 0.9 ± 0.4; PAD: 2.0 ± 0.3 pg/ml) and Interleukin 6 (IL-6) in both plasma (HC: 2.3 ± 0.4; PAD: 4.3 ± 0.5 pg/ml) and muscle (~75% greater). Intramuscular oxidative stress, assessed by protein carbonyls and 4-HNE, was significantly greater in PAD compared to HC. Ankle brachial index (ABI) was significantly correlated with intramuscular inflammation, oxidative stress, and mitochondrial-derived ROS production. Thus, elevated intramuscular inflammation, oxidative stress, and mitochondrial-derived ROS production likely contribute to the pathophysiology of the skeletal muscle dysfunction associated with PAD, even in the presence of preserved mitochondrial respiratory function in this population.

Keywords: PAD, ROS, oxidative capacity

Introduction

Mitochondrial dysfunction has been implicated in the pathophysiology of the symptoms associated with skeletal muscle dysfunction in PAD, which is not fully explained by the disease-related reduction in limb blood flow during exercise in this population (Brass & Hiatt, 2000; Pipinos et al., 2007; Pipinos, Judge, et al., 2008; Pipinos et al., 2006; Regensteiner et al., 1993; Weiss et al., 2013). Indeed, evidence suggests that an acquired mitochondrial myopathy may contribute to the progression of PAD (Brass & Hiatt, 2000; Pipinos et al., 2006; Pipinos et al., 2003). Alterations in mitochondrial volume, including both an increase (Elander, Sjostrom, Lundgren, Schersten, & Bylund-Fellenius, 1985) and decrease (Baum et al., 2016) have been documented in patients with PAD. However, very few studies have actually examined intrinsic mitochondrial respiratory function in patients with PAD. One such study, by Hou and colleagues (Hou et al., 2002), reported no difference in skeletal muscle ATP production rates, assessed in isolated mitochondria, between patients with PAD and healthy controls. In contrast, utilizing permeabilized muscle fibers, Pipinos and colleagues (Pipinos et al., 2006; Pipinos et al., 2003; Pipinos, Swanson, et al., 2008), documented impaired respiratory capacity in mitochondrial complexes I, II, and IV. Although the permeabilized fiber technique likely better represents in vivo physiology (Picard et al., 2010) it should be noted that the muscle assessed by Pipinos and colleagues (Pipinos et al., 2003) was typically taken from patients experiencing the most severe and debilitating form of occlusive disease, including patients with critical limb ischemia (Pipinos et al., 2006) and patients undergoing amputation (Pipinos et al., 2003). Therefore, the degree to which changes in mitochondrial function impact skeletal muscle physiology as a consequence of PAD, independent of sarcopenia resulting from muscle disuse and physical inactivity, is still unclear.

Attenuated mitochondrial and skeletal muscle function in various pathologies has commonly been linked to both inflammation and oxidative stress (Ha et al., 2016; Pipinos et al., 2007; Pipinos, Judge, et al., 2008; Pipinos et al., 2006; Signorelli et al., 2012; Signorelli, Fiore, & Malaponte, 2014; Signorelli et al., 2016; Signorelli et al., 2003; Weiss et al., 2013). Specifically, in a number of chronic clinical disorders such as sepsis, heart failure, and chronic obstructive pulmonary disease (Gifford et al., 2015; Quoilin, Mouithys-Mickalad, Lecart, Fontaine-Aupart, & Hoebeke, 2014; Stride et al., 2013), mitochondrial dysfunction has been associated with increased systemic inflammation which influences muscle protein synthesis and impairs both mitochondrial and muscle function (Supinski & Callahan, 2007). These detrimental processes appear to be predominantly a consequence of oxidative stress, secondary to reactive oxygen species (ROS), which may, somewhat ironically, be mitochondrial in origin (Supinski & Callahan, 2007). In PAD, augmented inflammation has been associated with an increased severity of the disease and compromised muscle function, as determined by ankle brachial index (ABI) and walking capacity, respectively (Nylaende et al., 2006; Signorelli et al., 2003). Furthermore, in addition to potentially abnormal mitochondrial function (Pipinos et al., 2006; Pipinos et al., 2003), studies in patients with PAD have revealed structural alterations in skeletal muscle, including a selective loss of type II fibers relative to type I fibers, general fiber atrophy, increased fibrosis, changes in capillarity, fatty deposition, and evidence of increased oxidative stress-induced damage (Brass, Hiatt, Gardner, & Hoppel, 2001; Cluff et al., 2013; Ha et al., 2016; Koutakis et al., 2015; Pipinos et al., 2006; Regensteiner et al., 1993). However, the degree to which mitochondrial ROS production is linked to inflammation and oxidative stress within the skeletal muscle of patients with PAD and the impact on mitochondrial respiratory capacity remains unknown.

Thus, the primary purpose of this study was to comprehensively assess mitochondrial respiratory function in a cohort of patients with PAD and HC matched for age and physical activity. Additionally, to better understand the mechanisms responsible for the proposed PAD-induced mitochondrial dysfunction, intramuscular and systemic inflammation, mitochondrial-derived ROS production, and oxidative stress were also assessed in both subject groups. We hypothesized that 1) skeletal muscle from patients with PAD would exhibit attenuated mitochondrial respiratory function compared to HC and 2) inflammation, mitochondrial ROS production, and oxidative stress, would be greater in patients with PAD compared to HC, implying a mechanistic role for these factors in the mitochondrial dysfunction associated with this population.

Methods

Ethical approval

The experimental protocol was approved by the Human Research Protection Program at the University of Utah and the Salt Lake City Veterans Affairs Medical Center (ethics approval #30810) and conforms to the Declaration of Helsinki, except for registration in a database. Written informed consent was obtained from all volunteers prior to their participation in the study.

Subjects

A total of 22 subjects, 12 HC and 10 patients with mild PAD, classified as Fontaine Stage II (Norgren et al., 2007), were recruited to participate in this study. Inclusion criteria for patients with PAD were a clinical diagnosis with evidence of femoropopliteal occlusive disease (ABI ≤ 0.90), intermittent claudication, and no prior PAD-related surgical interventions. Arterial lesions in the patients with PAD were localized to the distal femoral and/or popliteal arteries and characterized by moderate calcification. All HC were normotensive (<140/90 mmHg), not taking any medication recognized to alter blood flow or metabolism, and no evidence of overt cardiovascular disease. All subjects were recruited based upon being moderately physically active (assessed by both interview and accelerometry). Females were postmenopausal and not taking any form of estrogen-replacement therapy, minimizing the hormonal influences. Subjects reported to the laboratory for testing in a fasted state (>8 h postprandial) and refrained from caffeine or strenuous exercise prior to the studies (>24 h). All subjects performed a 10-m walk test as a clinical assessment of overall physical function during the initial screening visit, as previously described (McCully, Leiper, Sanders, & Griffin, 1999; Peters, Fritz, & Krotish, 2013).

Muscle Biopsy

A percutaneous needle biopsy was obtained from the anteromedial aspect of the gastrocnemius muscle belly, ~10 cm distal to the tibial tuberosity at a depth of ~1.0 cm under sterile conditions. In the patients with PAD, biopsies were performed on the most symptomatic leg, whereas biopsies in the HC were performed on the right leg. Immediately following removal of the muscle sample (~150 mg) from the leg, ~20% of the sample was immersed in ice-cold biopsy preservation fluid (BIOPS) for respiratory analysis (Park et al., 2014), and the remaining muscle sample was snap frozen and stored at −80°C for histological and biochemical analyses, as well as electron paramagnetic resonance (EPR) spectroscopy.

Mitochondrial Respiration

Muscle samples were prepared and permeabilized, as described by Park et al. (Park et al., 2014). Briefly, BIOPS-immersed fibers were carefully separated with fine-tip forceps and subsequently bathed in a BIOPS-based saponin solution (50μg saponin/ml BIOPS) for 30 minutes. Following saponin treatment, muscle fibers were rinsed twice in ice-cold mitochondrial respiration fluid (MIR05) for 10 minutes each rinse. The wet weight of the muscle sample (~3–4 mg) was then assessed on a calibrated scale.

The muscle fibers were then placed in two temperature-controlled respiration chambers (Oxytherm, Hansatech Instruments. Norfolk, UK) in 2 ml of MIR05 solution and warmed to 37°C, allowing duplicate measurements. After allowing the muscle 10 minutes to equilibrate, State 2 respiration was assessed with malate (2 mM) + glutamate (10 mM), followed by State 3 complex I (State 3:CI) respiration with ADP (5 mM) (State 3 CI). Subsequently, succinate (10 mM) was added to assess State 3 complex I and II glycolytic oxidation rate (State 3 S CI+II) followed by octanoyl carnitine (1.5 mM) to assess State 3:CI+CII fatty acid oxidation rate (State 3 O CI+II). Next, cytochrome C (10 μM) was added to the bath to verify membrane integrity, as recommended by Pesta et al. (Pesta & Gnaiger, 2012). Finally, rotenone (0.5mM) was added to inhibit CI for the assessment of State3:CII respiration (State 3 CII). Step changes in respiration lasted as long as required to produce a steady state respiration rate, typically ~3 minutes in duration. Background respiration inherent to the experimental setup was measured and taken into account when assessing overall mitochondrial respiration. The respiration data for each of the two separately assessed muscle fiber samples was then averaged. The rate of O2 consumption was measured as picomoles of O2 per second and then expressed relative to muscle sample mass (pmol·s−1·mg wet wt−1). Control ratios, including the respiratory control ratio (RCR), and substrate control ratios for the addition of octanoyl carnitine (SCRO) and succinate (SCRS), were calculated from the respiratory flux measurements, as described previously (Kraunsoe et al., 2010).

CS activity

Citrate synthase (CS) activity was assessed as an indicator of mitochondrial content and volume (Larsen et al., 2012; Meinild Lundby et al., 2018; Pipinos, Swanson, et al., 2008). After the respiration measurements, the same muscle samples (~3–4 mg wet weight) were homogenized in a buffer containing 250 mM sucrose, 40 mM KCl, 2 mM EGTA, and 20 mM Tris·HCl (Qiagen, Hilden, Germany) and a subsequent CS activity assay was performed using a spectrophotometer (Biotek Instruments, Winooski, VT), as previously described (Picard et al., 2008).

Muscle fiber type

Immunofluorescence was used to determine muscle fiber type (Myosin Heavy Chain I: primary antibody BA-F8, secondary antibody Alexa Fluor 350; Myosin Heavy Chain IIa/IIx: primary antibody SC-71, secondary antibody Alexa Fluor 488, Abcam, Cambridge, MA) and capillarity (ab96884, Abcam, Cambridge, MA) with an Imager A2 microscope with Axiocam and accompanying software (Zeiss, Jena, Germany), as described in detail by Bloemberg and Quadrilatero (Bloemberg & Quadrilatero, 2012). Primary antibodies used for fiber typing were from Developmental Studies Hybridoma Bank (University of Iowa, Iowa City, IA). Secondary antibodies were from Invitrogen (Thermo Fisher Scientific, Waltham, MA).

Free radicals, inflammation, and oxidative stress

Free radical measurements were assessed by EPR spectroscopy on muscle tissue and whole blood samples using an EMX X-band spectrometer (Bruker, MA). The frozen muscle tissue samples were placed in a micro-centrifuge tube containing 150 μL of the spin trap 1-hydroxy-4-[2-(triphenylphosphonio)-acetamido]-2,2,6,6-tetramethylpiperidine (mitoTempo-H) (0.5 mmol/L) (Enzo Life Sciences San Diego, CA) for 60 minutes at 37° C. After 60 minutes, the samples were placed on ice and 50 μL of the solution was loaded into a capillary tube for EPR spectroscopy analysis. The EPR spectroscopy scan was run with a center field at approximately g = 2.004 and the area under the curve of the EPR spectroscopy signal was calculated by double integration. Whole blood samples were collected and spin-trapped (Richardson et al., 2007). Briefly, 1.5 ml of venous blood was collected into a vacutainer containing 0.5 ml of the spin trap α-phenyl-N-tert-butylnitrone (PBN) (0.0140 mol/L). After centrifugation, the PBN adduct (300 μl) was pipetted into a precision-bore quartz EPR sample tube (Wilmad, Vineland, NJ). EPR spectroscopy was then performed at 21°C. The area under the curve of the EPR spectroscopy signal was calculated by double integration.

Venous blood samples, for the analysis of blood born markers of inflammation and oxidative stress, were centrifuged to allow the collection of the plasma and stored at −80°C until analysis. Concentrations of the pro-inflammatory cytokines C-reactive protein (CRP), TNFα, IL-6, and the anti-inflammatory cytokine IL-10, were determined by solid phase sandwich ELISA (R&D Systems, Minneapolis, MN, USA). Superoxide dismutase (SOD) activity was assessed as previously described (Wheeler, Salzman, Elsayed, Omaye, & Korte, 1990). Colorimetric ELISA assays were used to detect lipid peroxidation by plasma malondialdehyde (MDA) levels (Bioxytech LPO-586, Foster City, CA) and protein oxidation by protein carbonyl (PC) levels (R&D Systems, Minneapolis, MN).

To analyze skeletal muscle protein oxidation, tissue PC content was measured after treatment with 2.4-dinitrophenylhydraziine (DNPH), which forms labeled protein hydrazine derivatives (Mecocci et al., 1999). Briefly, two 200 μl samples were transferred to two 2 ml plastic tubes, a sample tube and a control tube. 800 μl of DNPH was then added to the sample tube and 800 μl of 2.5M hydrochloride was added to the control tube. Both the control and sample tube were incubated in the dark at room temperature for one hour and vortexed every 15 min. 1 ml of 20% and 10% trichloroacetic acid (TCA) was added and afterwards both tubes were centrifuged at 10,000 g for 10 min at 4°C. The pellets formed were then washed three times with 1 ml of ethanol/ethyl acetate mixture (1:1). The final pellets were dissolved in 500 μl of guanidine hydrochloride by vortexing and the solution was centrifuged at 10,000 g for an additional 10 min. PC content was determined spectrophotometrically at 370 nm and calculated using a molar absorption coefficient of 22,000 M−1 cm−1 (Pansarasa, Bertorelli, Vecchiet, Felzani, & Marzatico, 1999).

The relative abundance of the target proteins for IL-6 and 4-HNE was determined in skeletal muscle samples by immunoblotting. Briefly, tissue samples were homogenized 1:12 (wt/vol) in an ice-cold buffer containing 5 mM Tris (pH 7.5) and 5 mM EDTA (pH 8.0) with a protease inhibitor cocktail (1). Homogenates were centrifuged at 1500 g for 10 min at 4°C. After centrifugation, the supernatant was collected and the protein concentration was determined using the Bradford technique. Proteins from the supernatant fraction were separated via polyacrylamide gel electrophoresis, transferred onto a polyvinylidene difluoride membrane, and incubated with primary and secondary antibodies specific to the proteins of interest (Kwon, Tanner, et al.). Membranes were exposed on a ChemiDoc XRS (Bio-Rad) and quantified with Image Lab software (Bio-Rad). The specific antibodies used to detect skeletal muscle proteins IL-6 and 4-HNE included: Anti-Interleukin 6 (IL6) (ab6672, Abcam, Cambridge, MA) and 4-hydroxynonenal (4-HNE) (ab46545, Abcam, Cambridge, MA). The protein abundance of each protein was normalized to beta-actin (ab8227, Abcam, Cambridge, MA), which served as a loading control (Kwon, Smuder, et al.).

Lower Leg Volume and Muscle Mass

Lower leg volume was calculated based on lower leg circumference (three sites: distal, middle, and proximal), lower leg length, and skinfold measurements, with muscle mass then being calculated based upon the expected volume ratio of the major muscles of the lower leg (Jones & Pearson, 1969). This method has previously been confirmed to provide a valid estimate for muscle mass across a spectrum of individuals with normal muscle mass and severe muscle atrophy (Layec, Venturelli, Jeong, & Richardson, 2014).

Statistical Analysis

Differences in all measured variables between HC and PAD were determined with either independent t-tests or nonparametric Mann-Whitney tests, where appropriate (SigmaStat 11.0; Systat Software, San Jose, California). Potential between variable relationships were assessed using the Pearson test or the nonparametric Spearman rank-order correlation. A preliminary analysis performed with a two-factor ANOVA, with gender and age as factors, was used to determine if there was evidence of a significant gender or age specific effect in the main variables of this study. Statistical significance was accepted at P < 0.05. All data in the text and tables are expressed as mean ± S.D.

Results

Subject Characteristics

Subject characteristics and the physical activity assessments are displayed in Table 1, while subject blood, comorbidity, and medication data are documented in Table 2. The patients with PAD were reasonably well-matched with the HC for gender, age, height, weight, BMI, mean arterial pressure, lower leg volume and muscle mass, and hematology with the exception of red blood cells, which were significantly higher in the in the HC compared to patients with PAD (p < 0.05). While not statistically significant, fasting blood glucose concentrations tended to be higher in the patients with PAD compared to the HC (Table 2). In accordance with their disease status, patients with PAD possessed a significantly lower ABI (p < 0.05), lower HDL, and possessed a greater number of co-morbidities and medications compared to the HC (Table 2). Surprisingly, LDL was significantly elevated in the HC, while total cholesterol also tended to be higher when compared to the patients with PAD. Furthermore, although subjects were reasonably well-matched for physical activity, the HC had a significantly faster 10-m walk time compared to the patients with PAD (p < 0.05). In addition, the patients with PAD tended to spend more time being physically inactive (Table 1). Skeletal muscle morphological characteristics in the HC and patients with PAD are displayed in Table 3. Patients with PAD had a significantly greater percentage, surface area, and total number of type 1 muscle fibers compared to the HC (p < 0.05), although there was no difference between groups for the average number of capillaries around each muscle fiber.

Table 1:

Subject Characteristics and Physical Activity/Functional Assessments

HCPADP Value
N (Female/Male)11 (2/9)10 (2/8)
Age (y)62 ± 865 ± 90.42
Height (cm)173 ± 8168 ± 80.25
Weight (kg)82 ± 1979 ± 130.40
BMI (kg/m2)27 ± 528 ± 40.78
MAP (mmHg)106 ± 11107 ± 190.79
ABI1.18 ± 0.100.67 ± 0.10*2.98E-10
Smoker (former/present)(2/0)(4/5)
Lower leg volume (L)2.4 ± 0.82.4 ± 0.50.97
Lower leg muscle mass (kg)1.1 ± 0.21.1 ± 0.20.97
Physical activity
 Steps (count/day)6491 ± 29994838 ± 39820.29
 Sedentary activity (min/day)1239 ± 911383 ± 2540.11
 Light activity (min/day)97 ± 52100 ± 450.91
 Moderate activity (min/day)83 ± 4065 ± 370.35
 Vigorous activity (min/day)36 ± 2122 ± 220.19
10-meter walk time (s)6.8 ± 1.19.2 ± 2.1*0.01

Notes: BMI = body mass index; ABI = ankle/brachial index, MAP = mean arterial pressure. Values expressed as means ± SD.

*Significant difference between groups (p < 0.05).

Table 2:

Subject Blood, Comorbidity, and Medication Data

HCPADP Value
Glucose (mg/dL)88 ± 16110 ± 300.06
Cholesterol (mg/dL)199 ± 31165 ± 430.07
Triglycerides (mg/dL)101 ± 51147 ± 880.20
HDL (mg/dL)58 ± 1742 ± 7*0.03
LDL (mg/dL)130 ± 1995 ± 44*0.04
WBC (K/μL)6.3 ± 2.17.9 ± 1.80.10
RBC (K/μL)4.9 ± 0.34.5 ± 0.2*0.02
Hemoglobin (g/dL)14.2 ± 3.014.5 ± 1.40.82
Hematocrit (%)45.5 ± 2.743.2 ± 3.90.15
Comorbidity (%)
 Coronary artery disease040
 Myocardial infraction020
 Hypertension030
 Diabetes Mellitus010
 Renal insufficiency010
Medications (%)
 Statins medication040
 Aspirin/Clopidogrel20100

Notes: HDL = high-density cholesterol; LDL = low-density cholesterol; WBC = white blood cells; RBC = red blood cells. Values expressed as means ± SD.

*Significant difference between groups (p < 0.05).

Table 3:

Morphological Characteristics

HCPADP Value
Type 1 muscle fiber (%)53 ± 770 ± 2*0.03
Type 1 fiber area (μM2)415,315 ± 49,090633,346 ± 53,678*0.009
Type 2 fiber area (μM2)407,572 ± 74,915237,158 ± 29,1570.07
Number of type 1 fibers36 ± 556 ± 5*0.01
Number of type 2 fibers33 ± 623 ± 20.15
Average Ncaf2.9 ± 0.33.0 ± 0.20.22

Notes: Ncaf = number of capillaries around a single muscle fiber. Values expressed as means ± SD.

*Significant difference between groups (p < 0.05).

Mitochondrial function

Mitochondrial respiration in the HC and patients with PAD is displayed in Figure 1. There were no differences between groups for State 2 (HC: 8 ± 1; PAD 8 ± 2 pMol.mg.s−1), State 3CI (HC: 18 ± 1; PAD 20 ± 1 pMol.mg.s−1), State 3:CII (HC: 16 ± 1; PAD 16 ± 1 pMol.mg.s−1), or State 3:CI+II respiration using either octanoyl carnitine (HC: 19 ± 1; PAD 21 ± 2 pMol.mg.s−1) or succinate (HC: 25 ± 2; PAD 26 ± 2 pMol.mg.s−1). Furthermore, no differences between the HC and patients with PAD were observed in CS activity (HC: 38 ± 6; PAD: 40 ± 7 nmol.min.mg tissue−1) (p = 0.84). The RCR (HC: 4.7 ± 1.9; PAD: 4.7 ± 1.9), SCRS (HC: 1.3 ± 0.2; PAD: 1.3 ± 0.3), and SCRO (HC: 1.1 ± 0.3; PAD: 1.1 ± 0.3) were not different between groups.

An external file that holds a picture, illustration, etc.
Object name is nihms-956381-f0001.jpg
Mitochondrial respiration in healthy controls (HC) and patients with peripheral arterial disease (PAD).

CI, complex I; CII, complex II; O, octanoyl carnitine; S, succinate. Data presented as mean ± SD.

Inflammation, free radicals, and oxidative stress

Plasma markers of inflammation, free radical production, and oxidative stress are displayed in Table 4. Patients with PAD exhibited significantly greater plasma concentrations of TNF-α and IL-6 compared to the HC (p < 0.05). There were no significant differences between groups for the EPR spectroscopy detection of ROS in whole blood or plasma concentrations of CRP, IL-10, SOD, MDA, or PC. Patients with PAD had significantly increased levels of intramuscular pro-inflammatory cytokine IL-6 (~75% greater) (Figure 2A) and the intramuscular lipid peroxidation marker 4-HNE (~94% greater) (Figure 2B), expressed as a percentage of the HC (p < 0.05). Intramuscular PC were significantly elevated in patients with PAD compared to the HC (HC: 2.12 ± 0.10; PAD: 3.44 ± 0.15 nmol/mg protein, p < 0.05) (Figure 2C). In addition, the EPR spectroscopy signal representative of intramuscular, mitochondrial-derived ROS production was significantly greater in patients with PAD compared to the HC (HC: 1.00 ± 0.36; PAD 4.31 ± 0.97 AU per mg tissue, p < 0.05) (Figure 2D).

An external file that holds a picture, illustration, etc.
Object name is nihms-956381-f0002.jpg
The presence of intramuscular inflammatory marker IL-6 (panel A), oxidative stress markers 4-HNE and protein carbonyls (panels B and C, respectively), and mitochondrial ROS production (panel D) in healthy controls (HC) and patients with peripheral arterial disease (PAD).

Representative Western Blot images are presented for IL-6 and 4-HNE (panels A and B, respectively). Panel A, HC (n = 9) and PAD (n = 7); Panels B and C, HC (n = 9) and PAD (n = 7); Panel D, HC (n = 12) and PAD (n = 10). Data presented as mean ± SEM. *Significant difference between groups, (p < 0.05).

Table 4:

Plasma Markers of Inflammation, Free Radical Production, and Oxidative Stress

HCPADP Value
TNF-α (pg/ml)0.9 ± 0.42.0 ± 0.3*0.02
IL-6 (pg/ml)2.3 ± 0.44.3 ± 0.5*0.008
CRP (mg/L)1.3 ± 0.11.1 ± 0.10.54
IL-10 (pg/ml)0.3 ± 0.10.5 ± 0.20.12
EPR (A.U.)5.8 ± 1.74.0 ± 1.10.44
SOD (U/ml)2.7 ± 0.32.4 ± 0.30.51
MDA (μM)1.4 ± 0.11.6 ± 0.50.17
PC (nM/mg)0.04 ± 0.010.04 ± 0.020.90

Notes: TNF-a = tumor necrosis factor alpha, IL-6 = interleukin 6, CRP = C-reactive protein, IL-10 = interleukin 10, EPR = electron paramagnetic resonance spectroscopy detection of ROS, SOD = superoxide dismutase, MDA = malondialdehyde, PC = protein carbonyls. Values expressed as means ± SD.

*Significant difference between groups (p < 0.05).

Relationships between mitochondrial function, inflammation, free radicals, and oxidative stress

The hypothesis that an elevation in mitochondrial-derived ROS production, inflammation, and oxidative stress in patients with PAD would imply a mechanistic link between mitochondrial dysfunction led to the assessment of the relationships between these variables across all patients with PAD and the HC. State 2 respiration was negatively correlated to intramuscular 4-HNE (Figure 3A) (p < 0.05). When the groups were analysed separately, this correlation persisted in the patients with PAD (r = −0.883, p = 0.008), but not in the HC (r = 0.300, p = 0.47). A statistically significant correlation revealed that intramuscular PC were positively related to plasma concentrations of TNF-α (Figure 3B). Again, this relationship remained within the patients with PAD (r = 0.912, p = 0.01), but was absent within the HC (r = −0.615, p 0.10). Statistically significant positive correlations (p < 0.05) were evident between intramuscular 4-HNE and plasma concentration of IL-6 (Figure 3C), as well as mitochondrial-derived ROS production and intramuscular PC (Figure 3D) (p < 0.05), although these relationships were not present within either groups alone. Interestingly, statistically significant correlations (p < 0.05) with ABI were evident with mitochondrial ROS production, intramuscular inflammation, intramuscular oxidative stress, and systemic inflammation (Table 5). However, the statistical significance of these correlations was lost when the HC and patients with PAD were analyzed separately, likely due to a small number of relatively homogeneous subjects within each group. There were no statistically significant correlations with any measured variables and ROS, assessed by EPR, in whole blood. Furthermore, there were no relationships between capillarity and mitochondrial function.

An external file that holds a picture, illustration, etc.
Object name is nihms-956381-f0003.jpg
Correlations between mitochondrial state 2 respiration, inflammation, mitochondrial-derived ROS production, and oxidative stress.

Evidence of a relationship between mitochondrial state 2 respiration and intramuscular 4-HNE (panel A); plasma concentrations of TNF-α and intramuscular PC (panel B); plasma concentrations of IL-6 and intramuscular 4-HNE (panel C); and mitochondrial ROS production and intramuscular protein carbonyls (panel D). HC (n = 9) and PAD (n = 7). *Significant correlation, (p < 0.05).

Table 5:

Correlations between Inflammation, Free Radical Production, and Oxidative Stress

RelationshiprP Value
Inflammation
 Plasma IL-6/Intramuscular PC (nmol/mg protein)0.640.01
 Plasma IL-6/Plasma MDA (nM)0.550.02
ABI
 ABI/Intramuscular PC (nmol/mg protein)−0.800.001
 ABI/Intramuscular 4-HNE (fold beta actin)−0.660.006
 ABI/Intramuscular IL-6 (fold beta actin)−0.700.003
 ABI/Mitochondrial-derived ROS production (U/mg tissue)−0.550.03
 ABI/Plasma TNF-α Concentrations (pg/ml)−0.560.01
 ABI/Plasma IL-6 Concentrations (pg/ml)−0.540.02

Notes: TNF-α = tumor necrosis factor alpha, IL-6 = interleukin 6, CRP = C-reactive protein, MDA = malondialdehyde, PC = protein carbonyls, r = Pearson correlation.

Discussion

As the role of mitochondrial dysfunction in the pathophysiology of PAD is unclear, this study sought to comprehensively assess mitochondrial respiratory function in a group of patients with PAD and reasonably well-matched HC. In addition, this study aimed to assess the level of inflammation, mitochondrial-derived ROS production, and oxidative stress in an attempt to uncover the mechanisms responsible for the proposed mitochondrial dysfunction in patients with PAD. In contrast to previous studies utilizing the permeabilized muscle fiber technique in patients with severe PAD, this study revealed that mitochondrial respiratory function was preserved in patients with a milder form of PAD compared to the HC. However, the patients with PAD exhibited greater levels of inflammation, mitochondrial-derived ROS production, and intramuscular oxidative stress compared to the HC. Therefore, there was no evidence of impaired mitochondrial respiratory function in the patients with mild PAD despite elevated inflammation, mitochondrial-derived ROS production, and oxidative stress, suggesting that these factors, although likely important in the pathophysiology of PAD, do not compromise mitochondrial function in this population.

Mitochondrial respiratory function in patients with PAD

Mitochondrial DNA damage has been detected in patients with PAD (Brass & Hiatt, 2000), and is proposed to reflect systemic inflammation and oxidative stress, however, conclusions from prior studies regarding the impact of PAD on mitochondrial function have been equivocal. For example, Hou and colleagues (Hou et al., 2002) reported preserved peak mitochondrial ATP production rates in isolated mitochondria from patients with PAD with similar average ABIs (~0.67) as the patients assessed in the current study (Table 1). However, evidence suggests that the mitochondrial isolation procedure itself may selectively damage the organelles of interest, potentially confounding the assessment of mitochondrial function (Picard et al., 2011). More recently, Lindegaard-Pedersen and colleagues (Lindegaard Pedersen, Baekgaard, & Quistorff, 2017) found no differences in mitochondrial respiratory function assessed in permeabilized skeletal muscle fibers, a method documented to better represent in vivo physiology, between healthy controls and patients with PAD, with the exception of increased sensitivity to the substrate succinate in the patients with PAD. In contrast, Pipinos and colleagues (Pipinos et al., 2006; Pipinos et al., 2003) reported attenuated mitochondrial respiratory capacity from patients with PAD in permeabilized skeletal muscle fibers. However, it should be noted that the subjects had advanced PAD, critical limb ischemia, an average ABI of ~0.37, pain at rest, and a loss of muscle mass in the affected limb of the majority of patients. In fact, as nearly half of the patients in the study of Pipinos and colleagues (Pipinos et al., 2003) were undergoing surgical amputations of the symptomatic limb, the attenuated mitochondrial respiratory capacity may have been a consequence of end-stage sarcopenia and tissue necrosis (Anderson et al., 2009), rather than PAD per se.

Using a similar approach to Pipinos and colleagues (Pipinos et al., 2006; Pipinos et al., 2003), this study revealed preserved mitochondrial respiratory function in permeabilized muscle fibers from patients with PAD compared to reasonably well-matched HC (Figure 1). While these findings conflict with those of Pipinos and colleagues (Pipinos et al., 2006; Pipinos et al., 2003), as already eluded to, the disparity may be attributed to differences in disease severity. Specifically, the patients in the present study were classified as Fontaine stage IIa based on symptoms of intermittent claudication that appeared after walking >200 meters, an average ABI of 0.67 ± 0.10 (Table 1), and the absence of pain at rest (Gardner & Afaq, 2008; Norgren et al., 2007). Using a similar cohort of patients with PAD, Lindegaard-Pedersen and colleagues (Lindegaard Pedersen et al., 2017) reported no differences in State 3 and 4 mitochondrial respiration between healthy controls and patients with PAD. However, the same study reported significantly lower mitochondrial respiratory function in a group of patients with PAD combined with the diagnosis of type 2 diabetes, suggesting the potential role of insulin resistance in impairing compensatory mechanisms that preserve skeletal muscle function in patients with early stage PAD. Thus, the present findings demonstrate that mitochondrial respiratory function, assessed in permeabilized skeletal muscle fibers, is not an obligatory accompaniment to the early stages of PAD.

Differences in physical activity may also contribute to the previously documented decline in mitochondrial function in patients with PAD compared to their healthy counterparts. In this study, participants were carefully matched for physical activity using both accelerometry (Table 1) and questionnaires. Prior studies have associated a decline in mitochondrial function with walking performance in older adults (Coen et al., 2013; Santanasto et al., 2016) and in patients with PAD (Hou et al., 2002), while also demonstrating the beneficial effects of exercise to off-set this age-related decline in mitochondrial function (Joseph, Adhihetty, & Leeuwenburgh, 2015). Furthermore, the efficacy of physical activity to preserve mitochondrial function with age would most likely translate into preserved mitochondrial function with maintained physical activity in patients with PAD. Indeed, prior evidence strongly supports a link between the benefits of exercise therapy, which enhances mitochondrial synthesis and metabolic responses, and quality of life, which, likely in combination, ultimately increases physical activity in patients with PAD (Haas, Lloyd, Yang, & Terjung, 2012). Thus, based on the present findings, it appears that physical activity may play an important role in the preservation of mitochondrial respiratory capacity in the patients with PAD evaluated in this study.

Inflammation in the pathogenesis of PAD

Emerging evidence suggests that elevated systemic inflammation and oxidative stress collectively contribute to muscle dysfunction and sarcopenia in patients with PAD (Ha et al., 2016; Koutakis et al., 2015; Pipinos et al., 2006; Pipinos et al., 2003; Pipinos, Swanson, et al., 2008; Signorelli et al., 2012; Signorelli et al., 2014; Signorelli et al., 2016; Signorelli et al., 2003; Weiss et al., 2013). Furthermore, similar mechanisms involving inflammation, elevated ROS production, and oxidative stress contribute to the vascular sequela associated with impaired macro and microvascular dysfunction in atherosclerosis (Steven, Daiber, Dopheide, Münzel, & Espinola-Klein, 2017). Indeed, the pathogenesis of PAD appears to begin with chronic, repeated cycles of ischemia-reperfusion induced during skeletal muscle contraction as blood flows through an arterial stenosis, eliciting an exaggerated inflammatory response and systemic free radical production (Pipinos, Judge, et al., 2008). Elevations in free radicals trigger numerous pathophysiological pathways that result in systemic oxidative stress, leading to the clinical manifestations of muscle dysfunction and sarcopenia associated with PAD (Hiatt, Armstrong, Larson, & Brass, 2015). Although it is becoming increasingly apparent that the vasculature is a significant source of free radicals (e.g. NADPH oxidase) (Berry et al., 2000; Pipinos, Judge, et al., 2008; Walker et al., 2016), many have proposed that damaged skeletal muscle mitochondria may be the primary source of ROS in PAD (Pipinos, Judge, et al., 2008; Pipinos et al., 2006; Pipinos, Swanson, et al., 2008). As mitochondrial dysfunction may limit muscle function and lead to loss of muscle mass in patients with PAD, understanding the pathophysiological consequences of inflammation, mitochondrial-derived ROS production, and oxidative stress is clinically important.

In accordance with the proposed inflammation and oxidative stress-mediated pathway for the pathogenesis of PAD, this study revealed evidence of elevated systemic inflammation, as evidenced by ~2-fold greater plasma concentrations of TNF-α and IL-6 (Table 4) and a ~75% increase in intramuscular levels of IL-6 (Figure 2A) in patients with PAD compared to the HC. In accordance with these findings, Signorelli and colleagues (Signorelli et al., 2012; Signorelli et al., 2003) have documented elevations in plasma concentrations of TNF-α and IL-6 in patients with PAD at rest and following treadmill walking, although the influence on mitochondrial function and oxidative stress was not assessed. Using magnetic resonance spectroscopy to assess mitochondrial function in vivo, Tecilazich and colleagues (Tecilazich et al., 2013) reported slower phosphocreatine recovery kinetics associated with an increased pro-inflammatory state in diabetics with PAD. However, this impairment in mitochondrial function was evident in the diabetics whether or not they had PAD. To our knowledge, the present study is the first to document an increased pro-inflammatory state within both plasma and skeletal muscle of patients with mild PAD, supporting a systemic inflammatory profile associated with the pathophysiology of the disease.

Mitochondrial-derived ROS and oxidative stress in the pathogenesis of PAD

The heightened inflammatory profile observed in patients with PAD was accompanied by a ~4-fold greater mitochondrial-derived ROS production and a nearly two-fold increase in intramuscular oxidative stress compared to the HC, as evidence by increased lipid peroxidation (4-HNE) and protein oxidation (PC) (Figure 2). To our knowledge, the present study is the first to provide direct evidence of mitochondrial-derived ROS in patients with PAD. Furthermore, there was a significant positive relationship between mitochondrial-derived ROS production and intramuscular protein carbonyl concentrations (Figure 3D), suggesting a link between mitochondrial-derived ROS production and skeletal muscle oxidative stress. In contrast, patients with PAD and the HC exhibited similar levels of systemic free radicals, assessed by EPR spectroscopy in whole blood, corresponding to no difference in systemic levels of oxidative stress, as evidenced by similar plasma markers of lipid peroxidation (MDA) and protein oxidation (protein carbonyls) (Table 4). Thus, the present data suggest that the redox balance within skeletal muscle mitochondria may be more susceptible to increased systemic inflammation, resulting in elevated ROS production, although causality cannot be determined from the current study design. Of note, significant relationships between plasma concentrations of TNF-α and intramuscular protein carbonyl concentrations (Figure 3B), as well as between intramuscular levels of IL-6 and 4-HNE (Figure 3C), suggest a link between systemic inflammation and skeletal muscle oxidative stress, which is potentially mediated by excess mitochondrial-derived ROS production. Conversely, increased mitochondrial-derived ROS during the early stages of PAD may potentially serve to preserve mitochondrial function, mediated by an upregulation of the transcription factors NRF-1, NRF-2, and the PPARgamma coactivator-1alpha (PGC-1α), which are involved in mitochondrial biogenesis (Irrcher, Ljubicic, & Hood, 2009; Lee, Yin, Chi, & Wei, 2002; St-Pierre et al., 2003). While significant interactions between the ABI, mitochondrial-derived ROS, and markers of inflammation and oxidative stress were evident when both groups were combined in the present study, future studies that include a broader range of ABI’s in patients with PAD are needed to determine the relationship between ROS and mitochondrial function with increasing disease severity.

Interestingly, while there were no differences in mitochondrial respiratory function between patients with PAD and the HC, a significant negative correlation suggests a link between State 2, uncoupled, respiration and intramuscular 4-HNE (Figure 3A). Indeed, increased levels of superoxide and lipid peroxides have been documented to activate mitochondrial uncoupling to protect against further damage from ROS (Harper, Bevilacqua, Hagopian, Weindruch, & Ramsey, 2004) and the lipids composing the mitochondrial membrane are highly susceptible to oxidative damage (Chen & Yu, 1994), leading to the production of 4-HNE. In the present study, patients with PAD had greater levels of 4-HNE despite similar State 2 respiration compared to the HC. Therefore, the lipid oxidative damage may be a consequence of an insufficient adaptive response in patients with PAD to increase mitochondrial uncoupling in the face of increased levels of ROS. However, in recognition that our tissue samples were of inadequate mass to provide measurements of 4-HNE in all subjects (HC = 9, PAD = 7), future studies are needed to further elucidate the relationship between lipid peroxidation and the control of uncoupled mitochondrial respiration in patients with PAD.

Skeletal muscle fiber type and capillarity in patients with PAD

Alterations in skeletal muscle fiber isoforms and morphology have been previously associated with PAD (Cluff et al., 2013; McGuigan et al., 2001a, 2001b). A shift toward the more oxidative type 1 muscle fiber accompanies prolonged PAD and has been thought to be an adaptation driven by a need for greater oxygen delivery to the chronically ischemic muscle, but the fiber type transition may also be a consequence of disease-related neuropathy during the denervation-reinnervation process (Pipinos, Judge, et al., 2008). In the present study, patients with PAD exhibited a significantly greater percentage of type 1 muscle fibers, accompanied by a ~52% greater surface area and ~55% greater number of type 1 fibers compared to the HC (Table 3). The type 2 fiber area tended to be lower in the patients with PAD (Table 3), although statistical significance was not achieved (p = 0.07). A progressive loss in type 2 fibers has previously been documented in aging, cancer, diabetes, and heart failure (Miljkovic, Lim, Miljkovic, & Frontera, 2015). The proposed mechanisms that lead to a selective loss of type 2 fibers in these age-related chronic diseases include the activation of the FoxO and TGFβ family, autophagy inhibition, and NF-κB signalling related to increased TNF-α (Wang & Pessin, 2013). Prolonged, moderate ischemia has been demonstrated to increase FoxO signalling in animal models of PAD (Roudier et al., 2013). However, the influence of FoxO and NF-κB signalling on type 2 fibers in patients with PAD has not been fully elucidated.

Alterations in capillarity and mitochondria have been previously documented in patients with PAD (Baum et al., 2016; Elander et al., 1985). Capillarity, determined by the number of capillaries around a muscle fiber, was similar between the HC and patients with PAD. Capillary density, measured from biopsies of the medial gastrocnemius in patients with PAD, has been previously documented to positively correlate with whole body VO2 peak, peak walking time, and claudication onset time assessed during treadmill walking (Robbins et al., 2011), such that patients with reduced capillary density tend to exhibit lower exercise capacity and experience greater physical impairments. Furthermore, improvements in capillary density have been documented in patients with PAD following exercise training (Duscha et al., 2011; McGuigan et al., 2001b), demonstrating the influence of physical activity to potentially improve the pathophysiological decrements typically observed in the microcirculation of patients with PAD. Previous evidence suggests that the increase in angiogenic capacity following exercise training is mediated by the transcriptional coactivator PGC-1α (Chinsomboon et al., 2009), which, interestingly, is also recognized to regulate oxidative capacity by a concomitant, training-induced increase in mitochondrial biogenesis and capillarity (Haas et al., 2012). While the morphological shift to a predominant oxidative type 1 muscle fiber in patients with PAD in the present study is in agreement with previously documented findings, the observed preservation of capillary density may have been influenced by the careful matching of physical activity in the HC and the patients with PAD (Table 2).

Conclusion

The primary purpose of this study was to evaluate alterations in skeletal muscle physiology by comprehensively assessing mitochondrial respiratory function, oxidative stress, and inflammation in patients with PAD and HC of a similar age and reasonably well-matched for physical activity. There was no evidence of impaired mitochondrial respiratory function in this particular cohort of patients with PAD despite elevated inflammation, mitochondrial-derived ROS production, and oxidative stress compared to the HC. The present findings suggest that increased intramuscular inflammation, oxidative stress, and mitochondrial-derived ROS production precede impaired mitochondrial respiratory function in patients with mild PAD, although these factors likely contribute to the pathophysiology of muscle dysfunction with increased disease severity over time.

New Findings

The degree to which skeletal muscle mitochondrial-derived ROS production is linked to impaired skeletal muscle function in patients with early-stage peripheral arterial disease (PAD) and the impact on mitochondrial respiratory capacity remains unknown. This is the first study to document increased mitochondrial-derived reactive oxygen species (ROS) production associated with elevated intramuscular oxidative stress, despite preserved mitochondrial respiratory function, in patients with PAD. Furthermore, systemic inflammation, mitochondrial-derived ROS production, and skeletal muscle oxidative stress were strongly correlated to disease severity, as indicated by Ankle-Brachial Index (ABI), in patients with PAD.

Acknowledgements

The authors wish to thank all the subjects who partook in this study for their dedicated participation in this research.

Grants

This work was funded, in part, by the National Heart, Lung, and Blood Institute at the National Institute of Health (K99HL125756, PO1 HL1091830); the Flight Attendant Medical Research Institute (FAMRI), the Veteran’s Administration Rehabilitation Research and Development Service (E6910-R, E1697-R, E1433-P and E9275-L), Senior Research Career Scientist Award (E9275-L), and the Spire Award E1572P.

Footnotes

Disclosures

The authors declare that there are no competing interests associated with the manuscript.

References

  • Anderson EJ, Lustig ME, Boyle KE, Woodlief TL, Kane DA, Lin C-T, … Neufer PD (2009). Mitochondrial H(2)O(2) emission and cellular redox state link excess fat intake to insulin resistance in both rodents and humans. The Journal of Clinical Investigation, 119(3), 573–581. doi: 10.1172/JCI37048 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Baum O, Torchetti E, Malik C, Hoier B, Walker M, Walker PJ, … Hellsten Y (2016). Capillary ultrastructure and mitochondrial volume density in skeletal muscle in relation to reduced exercise capacity of patients with intermittent claudication. American Journal of Physiology-Regulatory, Integrative and Comparative Physiology, 310(10), R943–R951. doi: 10.1152/ajpregu.00480.2015 [PubMed] [CrossRef] [Google Scholar]
  • Berry C, Hamilton CA, Brosnan MJ, Magill FG, Berg GA, McMurray JJ, & Dominiczak AF (2000). Investigation into the sources of superoxide in human blood vessels: angiotensin II increases superoxide production in human internal mammary arteries. Circulation, 101(18), 2206–2212. [PubMed] [Google Scholar]
  • Bloemberg D, & Quadrilatero J (2012). Rapid determination of myosin heavy chain expression in rat, mouse, and human skeletal muscle using multicolor immunofluorescence analysis. PLoS One, 7(4), e35273. doi: 10.1371/journal.pone.0035273 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Brass EP, & Hiatt WR (2000). Acquired skeletal muscle metabolic myopathy in atherosclerotic peripheral arterial disease. Vasc Med, 5(1), 55–59. [PubMed] [Google Scholar]
  • Brass EP, Hiatt WR, Gardner AW, & Hoppel CL (2001). Decreased NADH dehydrogenase and ubiquinol-cytochrome c oxidoreductase in peripheral arterial disease. Am J Physiol Heart Circ Physiol, 280(2), H603–609. [PubMed] [Google Scholar]
  • Chen JJ, & Yu BP (1994). Alterations in mitochondrial membrane fluidity by lipid peroxidation products. Free Radic Biol Med, 17(5), 411–418. [PubMed] [Google Scholar]
  • Chinsomboon J, Ruas J, Gupta RK, Thom R, Shoag J, Rowe GC, … Arany Z (2009). The transcriptional coactivator PGC-1alpha mediates exercise-induced angiogenesis in skeletal muscle. Proc Natl Acad Sci U S A, 106(50), 21401–21406. doi: 10.1073/pnas.0909131106 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Cluff K, Miserlis D, Naganathan GK, Pipinos II, Koutakis P, Samal A, … Casale GP (2013). Morphometric analysis of gastrocnemius muscle biopsies from patients with peripheral arterial disease: objective grading of muscle degeneration. Am J Physiol Regul Integr Comp Physiol, 305(3), R291–299. doi: 10.1152/ajpregu.00525.2012 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Coen PM, Jubrias SA, Distefano G, Amati F, Mackey DC, Glynn NW, … Goodpaster BH (2013). Skeletal muscle mitochondrial energetics are associated with maximal aerobic capacity and walking speed in older adults. J Gerontol A Biol Sci Med Sci, 68(4), 447–455. doi: 10.1093/gerona/gls196 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Duscha BD, Robbins JL, Jones WS, Kraus WE, Lye RJ, Sanders JM, … Annex BH (2011). Angiogenesis in skeletal muscle precede improvements in peak oxygen uptake in peripheral artery disease patients. Arterioscler Thromb Vasc Biol, 31(11), 2742–2748. doi: 10.1161/atvbaha.111.230441 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Elander A, Sjostrom M, Lundgren F, Schersten T, & Bylund-Fellenius AC (1985). Biochemical and morphometric properties of mitochondrial populations in human muscle fibres. Clin Sci (Lond), 69(2), 153–164. [PubMed] [Google Scholar]
  • Gardner AW, & Afaq A (2008). MANAGEMENT OF LOWER EXTREMITY PERIPHERAL ARTERIAL DISEASE. Journal of cardiopulmonary rehabilitation and prevention, 28(6), 349–357. doi: 10.1097/HCR.0b013e31818c3b96 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Gifford JR, Trinity JD, Layec G, Garten RS, Park SY, Rossman MJ, … Richardson RS (2015). Quadriceps exercise intolerance in patients with chronic obstructive pulmonary disease: the potential role of altered skeletal muscle mitochondrial respiration. J Appl Physiol (1985), 119(8), 882–888. doi: 10.1152/japplphysiol.00460.2015 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ha DM, Carpenter LC, Koutakis P, Swanson SA, Zhu Z, Hanna M, … Casale GP (2016). Transforming growth factor-beta 1 produced by vascular smooth muscle cells predicts fibrosis in the gastrocnemius of patients with peripheral artery disease. J Transl Med, 14(1), 39. doi: 10.1186/s12967-016-0790-3 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Haas TL, Lloyd PG, Yang HT, & Terjung RL (2012). Exercise training and peripheral arterial disease. Compr Physiol, 2(4), 2933–3017. doi: 10.1002/cphy.c110065 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Harper ME, Bevilacqua L, Hagopian K, Weindruch R, & Ramsey JJ (2004). Ageing, oxidative stress, and mitochondrial uncoupling. Acta Physiol Scand, 182(4), 321–331. doi: 10.1111/j.1365-201X.2004.01370.x [PubMed] [CrossRef] [Google Scholar]
  • Hiatt WR, Armstrong EJ, Larson CJ, & Brass EP (2015). Pathogenesis of the limb manifestations and exercise limitations in peripheral artery disease. Circ Res, 116(9), 1527–1539. doi: 10.1161/circresaha.116.303566 [PubMed] [CrossRef] [Google Scholar]
  • Hou XY, Green S, Askew CD, Barker G, Green A, & Walker PJ (2002). Skeletal muscle mitochondrial ATP production rate and walking performance in peripheral arterial disease. Clin Physiol Funct Imaging, 22(3), 226–232. [PubMed] [Google Scholar]
  • Irrcher I, Ljubicic V, & Hood DA (2009). Interactions between ROS and AMP kinase activity in the regulation of PGC-1alpha transcription in skeletal muscle cells. Am J Physiol Cell Physiol, 296(1), C116–123. doi: 10.1152/ajpcell.00267.2007 [PubMed] [CrossRef] [Google Scholar]
  • Jones PR, & Pearson J (1969). Anthropometric determination of leg fat and muscle plus bone volumes in young male and female adults. J Physiol, 204(2), 63p–66p. [PubMed] [Google Scholar]
  • Joseph AM, Adhihetty PJ, & Leeuwenburgh C (2015). Beneficial effects of exercise on age-related mitochondrial dysfunction and oxidative stress in skeletal muscle. J Physiol. doi: 10.1113/jp270659 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Koutakis P, Myers SA, Cluff K, Ha DM, Haynatzki G, McComb RD, … Pipinos II (2015). Abnormal myofiber morphology and limb dysfunction in claudication. J Surg Res, 196(1), 172–179. doi: 10.1016/j.jss.2015.02.011 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kraunsoe R, Boushel R, Hansen CN, Schjerling P, Qvortrup K, Stockel M, … Dela F (2010). Mitochondrial respiration in subcutaneous and visceral adipose tissue from patients with morbid obesity. J Physiol, 588(Pt 12), 2023–2032. doi: 10.1113/jphysiol.2009.184754 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kwon OS, Smuder AJ, Wiggs MP, Hall SE, Sollanek KJ, Morton AB, … Powers SK AT1 receptor blocker losartan protects against mechanical ventilation-induced diaphragmatic dysfunction. J Appl Physiol (1985), 119(10), 1033–1041. [PMC free article] [PubMed] [Google Scholar]
  • Kwon OS, Tanner RE, Barrows KM, Runtsch M, Symons JD, Jalili T, … Drummond MJ MyD88 regulates physical inactivity-induced skeletal muscle inflammation, ceramide biosynthesis signaling, and glucose intolerance. Am J Physiol Endocrinol Metab, 309(1), E11–21. [PMC free article] [PubMed] [Google Scholar]
  • Larsen S, Nielsen J, Hansen CN, Nielsen LB, Wibrand F, Stride N, … Hey-Mogensen M (2012). Biomarkers of mitochondrial content in skeletal muscle of healthy young human subjects. J Physiol, 590(14), 3349–3360. doi: 10.1113/jphysiol.2012.230185 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Layec G, Venturelli M, Jeong EK, & Richardson RS (2014). The validity of anthropometric leg muscle volume estimation across a wide spectrum: from able-bodied adults to individuals with a spinal cord injury. J Appl Physiol (1985), 116(9), 1142–1147. doi: 10.1152/japplphysiol.01120.2013 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lee HC, Yin PH, Chi CW, & Wei YH (2002). Increase in mitochondrial mass in human fibroblasts under oxidative stress and during replicative cell senescence. J Biomed Sci, 9(6 Pt 1), 517–526. [PubMed] [Google Scholar]
  • Lindegaard Pedersen B, Baekgaard N, & Quistorff B (2017). Mitochondrial dysfunction in calf muscles of patients with combined peripheral arterial disease and diabetes type 2. Int Angiol, 36(5), 482–495. doi: 10.23736/s0392-9590.17.03824-x [PubMed] [CrossRef] [Google Scholar]
  • McCully K, Leiper C, Sanders T, & Griffin E (1999). The effects of peripheral vascular disease on gait. J Gerontol A Biol Sci Med Sci, 54(7), B291–294. [PubMed] [Google Scholar]
  • McGuigan MR, Bronks R, Newton RU, Sharman MJ, Graham JC, Cody DV, & Kraemer WJ (2001a). Muscle fiber characteristics in patients with peripheral arterial disease. Med Sci Sports Exerc, 33(12), 2016–2021. [PubMed] [Google Scholar]
  • McGuigan MR, Bronks R, Newton RU, Sharman MJ, Graham JC, Cody DV, & Kraemer WJ (2001b). Resistance training in patients with peripheral arterial disease: effects on myosin isoforms, fiber type distribution, and capillary supply to skeletal muscle. J Gerontol A Biol Sci Med Sci, 56(7), B302–310. [PubMed] [Google Scholar]
  • Mecocci P, Fano G, Fulle S, MacGarvey U, Shinobu L, Polidori MC, … Beal MF (1999). Age-dependent increases in oxidative damage to DNA, lipids, and proteins in human skeletal muscle. Free Radic Biol Med, 26(3–4), 303–308. [PubMed] [Google Scholar]
  • Meinild Lundby AK, Jacobs RA, Gehrig S, de Leur J, Hauser M, Bonne TC, … Lundby C (2018). Exercise training increases skeletal muscle mitochondrial volume density by enlargement of existing mitochondria and not de novo biogenesis. Acta Physiol (Oxf), 222(1). doi: 10.1111/apha.12905 [PubMed] [CrossRef] [Google Scholar]
  • Miljkovic N, Lim J-Y, Miljkovic I, & Frontera WR (2015). Aging of Skeletal Muscle Fibers. Annals of Rehabilitation Medicine, 39(2), 155–162. doi: 10.5535/arm.2015.39.2.155 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Norgren L, Hiatt WR, Dormandy JA, Nehler MR, Harris KA, Fowkes FG, … Rosenfield K (2007). Inter-Society Consensus for the Management of Peripheral Arterial Disease (TASC II). Eur J Vasc Endovasc Surg, 33 Suppl 1, S1–75. doi: 10.1016/j.ejvs.2006.09.024 [PubMed] [CrossRef] [Google Scholar]
  • Nylaende M, Kroese A, Stranden E, Morken B, Sandbaek G, Lindahl AK, … Seljeflot I (2006). Markers of vascular inflammation are associated with the extent of atherosclerosis assessed as angiographic score and treadmill walking distances in patients with peripheral arterial occlusive disease. Vasc Med, 11(1), 21–28. [PubMed] [Google Scholar]
  • Pansarasa O, Bertorelli L, Vecchiet J, Felzani G, & Marzatico F (1999). Age-dependent changes of antioxidant activities and markers of free radical damage in human skeletal muscle. Free Radic Biol Med, 27(5–6), 617–622. [PubMed] [Google Scholar]
  • Park SY, Gifford JR, Andtbacka RH, Trinity JD, Hyngstrom JR, Garten RS, … Richardson RS (2014). Cardiac, skeletal, and smooth muscle mitochondrial respiration: are all mitochondria created equal? Am J Physiol Heart Circ Physiol, 307(3), H346–352. doi: 10.1152/ajpheart.00227.2014 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Pesta D, & Gnaiger E (2012). High-resolution respirometry: OXPHOS protocols for human cells and permeabilized fibers from small biopsies of human muscle. Methods Mol Biol, 810, 25–58. doi: 10.1007/978-1-61779-382-0_3 [PubMed] [CrossRef] [Google Scholar]
  • Peters DM, Fritz SL, & Krotish DE (2013). Assessing the reliability and validity of a shorter walk test compared with the 10-Meter Walk Test for measurements of gait speed in healthy, older adults. J Geriatr Phys Ther, 36(1), 24–30. doi: 10.1519/JPT.0b013e318248e20d [PubMed] [CrossRef] [Google Scholar]
  • Picard M, Csukly K, Robillard ME, Godin R, Ascah A, Bourcier-Lucas C, & Burelle Y (2008). Resistance to Ca2+-induced opening of the permeability transition pore differs in mitochondria from glycolytic and oxidative muscles. Am J Physiol Regul Integr Comp Physiol, 295(2), R659–668. doi: 10.1152/ajpregu.90357.2008 [PubMed] [CrossRef] [Google Scholar]
  • Picard M, Ritchie D, Wright KJ, Romestaing C, Thomas MM, Rowan SL, … Hepple RT (2010). Mitochondrial functional impairment with aging is exaggerated in isolated mitochondria compared to permeabilized myofibers. Aging Cell, 9(6), 1032–1046. doi: 10.1111/j.1474-9726.2010.00628.x [PubMed] [CrossRef] [Google Scholar]
  • Picard M, Taivassalo T, Ritchie D, Wright KJ, Thomas MM, Romestaing C, & Hepple RT (2011). Mitochondrial structure and function are disrupted by standard isolation methods. PLoS One, 6(3), e18317. doi: 10.1371/journal.pone.0018317 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Pipinos II, Judge AR, Selsby JT, Zhu Z, Swanson SA, Nella AA, & Dodd SL (2007). The myopathy of peripheral arterial occlusive disease: part 1. Functional and histomorphological changes and evidence for mitochondrial dysfunction. Vasc Endovascular Surg, 41(6), 481–489. doi: 10.1177/1538574407311106 [PubMed] [CrossRef] [Google Scholar]
  • Pipinos II, Judge AR, Selsby JT, Zhu Z, Swanson SA, Nella AA, & Dodd SL (2008). The myopathy of peripheral arterial occlusive disease: Part 2. Oxidative stress, neuropathy, and shift in muscle fiber type. Vasc Endovascular Surg, 42(2), 101–112. doi: 10.1177/1538574408315995 [PubMed] [CrossRef] [Google Scholar]
  • Pipinos II, Judge AR, Zhu Z, Selsby JT, Swanson SA, Johanning JM, … Dodd SL (2006). Mitochondrial defects and oxidative damage in patients with peripheral arterial disease. Free Radic Biol Med, 41(2), 262–269. doi: 10.1016/j.freeradbiomed.2006.04.003 [PubMed] [CrossRef] [Google Scholar]
  • Pipinos II, Sharov VG, Shepard AD, Anagnostopoulos PV, Katsamouris A, Todor A, … Sabbah HN (2003). Abnormal mitochondrial respiration in skeletal muscle in patients with peripheral arterial disease. J Vasc Surg, 38(4), 827–832. [PubMed] [Google Scholar]
  • Pipinos II, Swanson SA, Zhu Z, Nella AA, Weiss DJ, Gutti TL, … Casale GP (2008). Chronically ischemic mouse skeletal muscle exhibits myopathy in association with mitochondrial dysfunction and oxidative damage. Am J Physiol Regul Integr Comp Physiol, 295(1), R290–296. doi: 10.1152/ajpregu.90374.2008 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Quoilin C, Mouithys-Mickalad A, Lecart S, Fontaine-Aupart MP, & Hoebeke M (2014). Evidence of oxidative stress and mitochondrial respiratory chain dysfunction in an in vitro model of sepsis-induced kidney injury. Biochim Biophys Acta, 1837(10), 1790–1800. doi: 10.1016/j.bbabio.2014.07.005 [PubMed] [CrossRef] [Google Scholar]
  • Regensteiner JG, Wolfel EE, Brass EP, Carry MR, Ringel SP, Hargarten ME, … Hiatt WR (1993). Chronic changes in skeletal muscle histology and function in peripheral arterial disease. Circulation, 87(2), 413–421. [PubMed] [Google Scholar]
  • Richardson RS, Donato AJ, Uberoi A, Wray DW, Lawrenson L, Nishiyama S, & Bailey DM (2007). Exercise-induced brachial artery vasodilation: role of free radicals. Am J Physiol Heart Circ Physiol, 292(3), H1516–1522. doi: 10.1152/ajpheart.01045.2006 [PubMed] [CrossRef] [Google Scholar]
  • Robbins JL, Jones WS, Duscha BD, Allen JD, Kraus WE, Regensteiner JG, … Annex BH (2011). Relationship between leg muscle capillary density and peak hyperemic blood flow with endurance capacity in peripheral artery disease. J Appl Physiol (1985), 111(1), 81–86. doi: 10.1152/japplphysiol.00141.2011 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Roudier E, Milkiewicz M, Birot O, Slopack D, Montelius A, Gustafsson T, … Haas TL (2013). Endothelial FoxO1 is an intrinsic regulator of thrombospondin 1 expression that restrains angiogenesis in ischemic muscle. Angiogenesis, 16(4), 759–772. doi: 10.1007/s10456-013-9353-x [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Santanasto AJ, Coen PM, Glynn NW, Conley KE, Jubrias SA, Amati F, … Newman AB (2016). The relationship between mitochondrial function and walking performance in older adults with a wide range of physical function. Exp Gerontol, 81, 1–7. doi: 10.1016/j.exger.2016.04.002 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Signorelli SS, Anzaldi M, Fiore V, Simili M, Puccia G, Libra M, … Neri S (2012). Patients with unrecognized peripheral arterial disease (PAD) assessed by ankle-brachial index (ABI) present a defined profile of proinflammatory markers compared to healthy subjects. Cytokine, 59(2), 294–298. doi: 10.1016/j.cyto.2012.04.038 [PubMed] [CrossRef] [Google Scholar]
  • Signorelli SS, Fiore V, & Malaponte G (2014). Inflammation and peripheral arterial disease: the value of circulating biomarkers (Review). Int J Mol Med, 33(4), 777–783. doi: 10.3892/ijmm.2014.1657 [PubMed] [CrossRef] [Google Scholar]
  • Signorelli SS, Li Volsi G, Pitruzzella A, Fiore V, Mangiafico M, Vanella L, … Li Volti G (2016). Circulating miR-130a, miR-27b, and miR-210 in Patients With Peripheral Artery Disease and Their Potential Relationship With Oxidative Stress: A Pilot Study. Angiology. doi: 10.1177/0003319716638242 [PubMed] [CrossRef] [Google Scholar]
  • Signorelli SS, Mazzarino MC, Di Pino L, Malaponte G, Porto C, Pennisi G, … Virgilio V (2003). High circulating levels of cytokines (IL-6 and TNFalpha), adhesion molecules (VCAM-1 and ICAM-1) and selectins in patients with peripheral arterial disease at rest and after a treadmill test. Vasc Med, 8(1), 15–19. [PubMed] [Google Scholar]
  • St-Pierre J, Lin J, Krauss S, Tarr PT, Yang R, Newgard CB, & Spiegelman BM (2003). Bioenergetic analysis of peroxisome proliferator-activated receptor gamma coactivators 1alpha and 1beta (PGC-1alpha and PGC-1beta) in muscle cells. J Biol Chem, 278(29), 26597–26603. doi: 10.1074/jbc.M301850200 [PubMed] [CrossRef] [Google Scholar]
  • Steven S, Daiber A, Dopheide JF, Münzel T, & Espinola-Klein C (2017). Peripheral artery disease, redox signaling, oxidative stress – Basic and clinical aspects. Redox Biology, 12, 787–797. doi: 10.1016/j.redox.2017.04.017 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Stride N, Larsen S, Hey-Mogensen M, Sander K, Lund JT, Gustafsson F, … Dela F (2013). Decreased mitochondrial oxidative phosphorylation capacity in the human heart with left ventricular systolic dysfunction. Eur J Heart Fail, 15(2), 150–157. doi: 10.1093/eurjhf/hfs172 [PubMed] [CrossRef] [Google Scholar]
  • Supinski GS, & Callahan LA (2007). Free radical-mediated skeletal muscle dysfunction in inflammatory conditions. Journal of Applied Physiology, 102(5), 2056–2063. doi: 10.1152/japplphysiol.01138.2006 [PubMed] [CrossRef] [Google Scholar]
  • Tecilazich F, Dinh T, Lyons TE, Guest J, Villafuerte RA, Sampanis C, … Veves A (2013). Postexercise phosphocreatine recovery, an index of mitochondrial oxidative phosphorylation, is reduced in diabetic patients with lower extremity complications. J Vasc Surg, 57(4), 997–1005. doi: 10.1016/j.jvs.2012.10.011 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Walker MA, Hoier B, Walker PJ, Schulze K, Bangsbo J, Hellsten Y, & Askew CD (2016). Vasoactive enzymes and blood flow responses to passive and active exercise in peripheral arterial disease. Atherosclerosis, 246, 98–105. doi: 10.1016/j.atherosclerosis.2015.12.029 [PubMed] [CrossRef] [Google Scholar]
  • Wang Y, & Pessin JE (2013). Mechanisms for fiber-type specificity of skeletal muscle atrophy. Current opinion in clinical nutrition and metabolic care, 16(3), 243–250. doi: 10.1097/MCO.0b013e328360272d [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Weiss DJ, Casale GP, Koutakis P, Nella AA, Swanson SA, Zhu Z, … Pipinos II (2013). Oxidative damage and myofiber degeneration in the gastrocnemius of patients with peripheral arterial disease. J Transl Med, 11, 230. doi: 10.1186/1479-5876-11-230 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Wheeler CR, Salzman JA, Elsayed NM, Omaye ST, & Korte DW Jr. (1990). Automated assays for superoxide dismutase, catalase, glutathione peroxidase, and glutathione reductase activity. Anal Biochem, 184(2), 193–199. [PubMed] [Google Scholar]
-