OBJECTIVE

The contribution of mitochondrial dysfunction to skeletal muscle insulin resistance remains elusive. Comparative proteomics are being applied to generate new hypotheses in human biology and were applied here to isolated mitochondria to identify novel changes in mitochondrial protein abundance present in insulin-resistant muscle.

RESEARCH DESIGN AND METHODS

Mitochondria were isolated from vastus lateralis muscle from lean and insulin-sensitive individuals and from obese and insulin-resistant individuals who were otherwise healthy. Respiration and reactive oxygen species (ROS) production rates were measured in vitro. Relative abundances of proteins detected by mass spectrometry were determined using a normalized spectral abundance factor method.

RESULTS

NADH- and FADH2-linked maximal respiration rates were similar between lean and obese individuals. Rates of pyruvate and palmitoyl-dl-carnitine (both including malate) ROS production were significantly higher in obesity. Mitochondria from obese individuals maintained higher (more negative) extramitochondrial ATP free energy at low metabolic flux, suggesting that stronger mitochondrial thermodynamic driving forces may underlie the higher ROS production. Tandem mass spectrometry identified protein abundance differences per mitochondrial mass in insulin resistance, including lower abundance of complex I subunits and enzymes involved in the oxidation of branched-chain amino acids (BCAA) and fatty acids (e.g., carnitine palmitoyltransferase 1B).

CONCLUSIONS

We provide data suggesting normal oxidative capacity of mitochondria in insulin-resistant skeletal muscle in parallel with high rates of ROS production. Furthermore, we show specific abundance differences in proteins involved in fat and BCAA oxidation that might contribute to the accumulation of lipid and BCAA frequently associated with the pathogenesis of insulin resistance.

Defining the role of mitochondrial dysfunction in the pathogenesis of skeletal muscle insulin resistance has been challenging. If mitochondria were less able to oxidize fatty acids, intramyocellular triacylglycerol and its metabolites, such as long-chain fatty acyl-CoAs, would rise, leading to impaired insulin signaling and insulin resistance (1). Mitochondrial mass is consistently 14–38% lower in muscle in obesity and type 2 diabetes (2,,5). In vivo studies, using magnetic resonance spectroscopy, concluded that basal and insulin-stimulated ATP synthesis rates were lower in insulin-resistant offspring of patients with type 2 diabetes (6,7). Although these data suggest intrinsic mitochondrial impairment, it may be that differences in mitochondrial content and cellular ATP demand underlie these observations.

Measurements on isolated mitochondria show that maximal respiration is lower in muscle from type 2 diabetics (8). Studies using saponin-permeabilized isolated human skeletal muscle (SM) fibers reported normal (9) or impaired (10) respiration in type 2 diabetic subjects. Nair et al., in 2009, showed dissociation between insulin sensitivity and mitochondrial ATP synthesis rates while demonstrating an effect of age on isolated mitochondrial activity (11). Given the diversity of these results, the contribution of mitochondrial oxidative capacity to insulin resistance is poorly defined.

Besides an intrinsic impairment in oxidative capacity, increased oxidative stress could produce insulin resistance (12). The superoxide anion (−O2·) is derived from several cellular sources, but the main contributor in SM may be electron leakage from mitochondria (13). At rest, ∼0.1–0.2% of consumed oxygen is converted to reactive oxygen species (ROS) (14). Insulin-resistant animal models have higher levels of superoxide production (15), and healthy individuals fed a high-fat diet had increased ROS generation in permeabilized muscle fibers without a change in mitochondrial respiration (16).

This study was undertaken to identify abnormalities in mitochondria isolated from insulin-resistant muscle by assessing mitochondrial bioenergetics and ROS production rates combined with proteomic assessment of abundance of mitochondrial proteins.

Subjects.

The purpose, nature and potential risks of the study were explained to the participants, and written consent was obtained. The protocol was approved by the institutional review board of Arizona State University. Additional information is provided in an online appendix (http://diabetes.diabetesjournals.org/cgi/content/full/db10-0174/DC1).

Muscle biopsy and hyperinsulinemic-euglycemic clamp.

Participants underwent a euglycemic-hyperinsulinemic clamp with basal muscle biopsies, regarding which the online appendix provides additional information.

Myosin heavy-chain isoform distribution.

Myosin heavy-chain (MHC) isoform composition was determined as previously described (17). MHC I, IIx, and IIa protein bands were visualized using Coomasie stain and quantified using densitometry.

Small scale isolation of mitochondria.

The online appendix provides information regarding the small-scale isolation of mitochondria.

Mitochondrial energetics.

Oxygen uptake by isolated mitochondria was measured using a DW1 Electrode Chamber (Hansatech Instruments, Norfolk, U.K.) in 250 μl respiration medium containing (in millimolar): 100 KCl, 50 3-[N-morpholino] propanesulfonic acid, 10 K.PO4, 20 glucose, 10 MgCl2, 1 EGTA, and 0.2% BSA, pH 7.0, at 37°C. Indexes of mitochondrial function were assessed using the following:: 1 mmol/l pyruvate, 10 mmol/l glutamate, 10 μmol/l palmitoyl-dl-carnitine (PC), and 5 mmol/l succinate. All substrates included 1 mmol/l malate to prime the citrate cycle. Mitochondria (10–50 μg) and substrates were added to respiration medium, and maximal (state 3) O2 consumption rate (Jo2) was stimulated with 125 nmol ADP (0.5 mmol/l). Complete phosphorylation of this ADP bolus elicited transition to resting Jo2 (state 4), which allowed calculation of the ADP-to-O ratio and the state 3–to–state 4 ratio (respiratory control ratio).

Citrate synthase activity.

Citrate synthase activity was measured using the method of Srere (18). The online appendix provides further information.

Sensitivity of mitochondria to metabolic control signals.

The sensitivity of mitochondrial respiration to kinetic and thermodynamic control signals (free ADP and ATP free energy, respectively) was assessed using the “creatine kinase energy clamp” of Messer et al. (19). Mitochondria (40–60 μg protein) were added to respiration medium containing 1.22 mmol/l phosphocreatine (PCr), 5.0 mmol/l creatine (Cr), and 5.0 mmol/l ATP. Creatine kinase was added in excess (30 units creatine kinase; Sigma Chemical, St. Louis, MO) to maintain thermodynamic equilibrium. Free [ADP] ([ADP]f) was calculated using the equilibrium constant of the creatine kinase reaction (KCK), assumed to be 150 (20):

formula

The [ADP]f with the known phosphate ([Pi]) of the medium provided the extramitochondrial free energy of ATP (ΔGATP): ΔGATP = RTLn(Γ/KATP), where R is the gas constant, T is degrees Kelvin, Γ is the mass action ratio of the ATP hydrolysis reaction ([ADP]f[Pi]/[ATP]), and KATP is the equilibrium constant of ATP hydrolysis (225,000, based on the study of Teague and Dobson [20]).

Stepwise additions of PCr established progressively lower [ADP]f and higher (more negative) ΔGATP, suppressing mitochondrial respiration (supplemental Fig. 1). Eadie-Hofstee analysis of the [ADP]f:Jo2 relation yielded the KMADP for respiration. Plots of ΔGATP:Jo2 were near linear (thermodynamic force-flow relation) and were extrapolated through the abscissa to calculate the theoretical ΔGATP at zero respiration (static head energy) (Fig. 3 A).

ROS measurement.

Substrates (see mitochondrial energetics for concentrations of substrates [10 μmol/l rotenone and 1 μg/μl oligomycin]) were added to respiration medium. Horseradish peroxidase (20 mU) and 10-Acetyl-3,7-dihydroxyphenoxazine (17 ng/μl) were added to a 96-well plate. ROS production was initiated by the addition of 10 μg mitochondrial protein. Optical density was measured every 5 min for 30 min (λex: 550 and λem: 590 nm). Due to limited material, 10 μg rat mitochondria were added to the H2O2 standard curve to correct for signal quenching by mitochondria.

Mass spectrometry sample preparation.

Forty micrograms of mitochondrial protein were separated by one-dimensional SDS polyacrylamide gel (12%). Proteins were visualized with Coomassie blue (Sigma Chemical, St. Louis, MO). Gel lanes (supplemental Fig. 2) were cut into 20 slices. Each slice was destained, dehydrated, and digested with trypsin as previously described in detail (21). The volume of the extracted peptide solution was reduced to 5 μl by vacuum centrifugation, and 20 μl of 0.1% trifluoroacetic acid/1% formic acid/2% acetonitrile was added.

Mass spectrometry.

High-performance liquid chromatography–electrospray ionization–tandem mass spectrometry was performed on a hybrid linear ion trap–Fourier Transform ion cyclotron resonance mass spectrometer (LTQ FT; Thermo Fisher; San Jose, CA) fitted with a PicoView nanospray source (New Objective, Woburn, MA) (refer to Supplemental Information).

Data analysis and bioinformatics.

Assignment of individual proteins was completed using Mascot (version 2.2; Matrix Science, London, U.K.) and Scaffold (version 02_00_06; Proteome Software, Portland, OR) as previously described (21) (refer to Supplemental Information).

Normalized spectral abundance factor calculations and immunoblotting.

Normalized spectral abundance factors (NSAFs) were calculated for every protein detected in each isolated mitochondria sample as previously described by Hwang et al. (22), who confirmed that this methodology can be applied to complex protein mixtures in a reproducible fashion. In addition, loading 30–120 μg protein lysate generates similar NSAF values (22). Briefly, NSAF values were log2 transformed to yield a Gaussian distribution. For every protein, the average lean log2 NSAF values were subtracted from all log2 NSAF values (both lean and obese groups), resulting in a log2-fold change for the obese group relative to the lean group. The online appendix provides information regarding immunoblotting.

Metabolic and skeletal muscle characteristics.

Lean and obese participants were matched for age, sex, and aerobic capacity (Vo2max) (Table 1). Obese individuals had higher BMI (P < 0.01), fasting plasma insulin (P < 0.01), A1C (P < 0.01), and fasting plasma glucose (P < 0.05). Insulin-stimulated glucose disposal was lower in the obese group (7.9 ± 0.6 vs. 3.7 ± 0.3 mg · kg−1 · min−1; P < 0.01); suppression of endogeneous glucose output was complete in both groups (0.1 ± 0.1 vs. 0.0 ± 0.1 mg · kg−1 · min−1). Obese participants had a higher proportion of myosin heavy-chain IIx (MYHI) (P < 0.05). No differences in MHC IIa (MYH2) or I (MYH7) were observed. Total mitochondrial protein and citrate synthase activity isolated per muscle wet weight were lower in obese individuals (P < 0.01), although the proportion of mitochondria recovered in both lean and obese groups was similar (∼65%).

TABLE 1

Clinical, metabolic, and skeletal muscle characteristics

LeanObese
N 20 14 
Male/female 11/9 7/7 
Age (years) 38 ± 3 38 ± 4 
BMI (kg/m225 ± 1 36 ± 1* 
Weight (kg) 77.6 ± 3.6 103.1 ± 4.2 
Fat mass (kg) 20.4 ± 2.8 37.4 ± 2.5* 
Plasma glucose (mg/dl) 86 ± 2 93 ± 2 
Serum insulin (pmol/l) 30 ± 6 90 ± 12* 
A1C (%) 5.4 ± 0 5.7 ± 0.1* 
HDL cholesterol (mmol/l) 51 ± 3 46 ± 2 
LDL cholesterol (mmol/l) 109 ± 9 113 ± 8 
Plasma triglycerides (mmol/l) 85 ± 8 126 ± 21 
OGTT (2-h glycemia) 105 ± 6 135 ± 8* 
Rd basal (mg · kg−1 · min−12.1 ± 0.1 1.8 ± 0.1* 
Rd clamp (mg · kg−1 · min−17.9 ± 0.6 3.7 ± 0.3* 
HGP clamp (mg · kg−1 · min−10.1 ± 0.1 0.0 ± 0.1 
Pulse (bpm) 67 ± 2 71 ± 2 
Systolic/diastolic blood pressure (mmHg) 124 ± 5/82 ± 3 130 ± 3/87 ± 2 
Vo2max (ml · min−1 · kg−131.5 ± 1.9 25.9 ± 1.8 
Vo2max (ml · min−1 · kg FFM−142.9 ± 2.4 38.8 ± 2.0 
% MHC I 31 ± 4 24 ± 3 
% MHC IIx 21 ± 2 31 ± 4* 
% MHC IIa 48 ± 6 45 ± 6 
Isolated mitochondrial protein (μg/μl) 6.5 ± 0.3 5.4 ± 0.2* 
Isolated citrate synthase activity (μmol · min−1 · g wet weight−18.2 ± 0.6 6.0 ± 0.2* 
Mitochondria recovery (%) 66 ± 3 64 ± 3 
LeanObese
N 20 14 
Male/female 11/9 7/7 
Age (years) 38 ± 3 38 ± 4 
BMI (kg/m225 ± 1 36 ± 1* 
Weight (kg) 77.6 ± 3.6 103.1 ± 4.2 
Fat mass (kg) 20.4 ± 2.8 37.4 ± 2.5* 
Plasma glucose (mg/dl) 86 ± 2 93 ± 2 
Serum insulin (pmol/l) 30 ± 6 90 ± 12* 
A1C (%) 5.4 ± 0 5.7 ± 0.1* 
HDL cholesterol (mmol/l) 51 ± 3 46 ± 2 
LDL cholesterol (mmol/l) 109 ± 9 113 ± 8 
Plasma triglycerides (mmol/l) 85 ± 8 126 ± 21 
OGTT (2-h glycemia) 105 ± 6 135 ± 8* 
Rd basal (mg · kg−1 · min−12.1 ± 0.1 1.8 ± 0.1* 
Rd clamp (mg · kg−1 · min−17.9 ± 0.6 3.7 ± 0.3* 
HGP clamp (mg · kg−1 · min−10.1 ± 0.1 0.0 ± 0.1 
Pulse (bpm) 67 ± 2 71 ± 2 
Systolic/diastolic blood pressure (mmHg) 124 ± 5/82 ± 3 130 ± 3/87 ± 2 
Vo2max (ml · min−1 · kg−131.5 ± 1.9 25.9 ± 1.8 
Vo2max (ml · min−1 · kg FFM−142.9 ± 2.4 38.8 ± 2.0 
% MHC I 31 ± 4 24 ± 3 
% MHC IIx 21 ± 2 31 ± 4* 
% MHC IIa 48 ± 6 45 ± 6 
Isolated mitochondrial protein (μg/μl) 6.5 ± 0.3 5.4 ± 0.2* 
Isolated citrate synthase activity (μmol · min−1 · g wet weight−18.2 ± 0.6 6.0 ± 0.2* 
Mitochondria recovery (%) 66 ± 3 64 ± 3 

Data are presented as means ± SE. Student's t test was performed.

*P < 0.01 and

P < 0.05 vs. lean participants.

‡Mitochondria pellets were resuspended in 100 μl of mannitol-sucrose buffer. HGP, hepatic glucose production.

In vitro mitochondrial oxidative capacity.

ADP-to-O and respiratory control ratios were within previously published ranges for isolated mitochondria from human muscle and did not differ between groups (Table 2). Maximal and resting respiration in response to NADH-linked (pyruvate, malate, and glutamate) and FADH2-linked (succinate) substrates did not differ between the groups (Fig. 1 A and B). PC oxidation also was similar between both groups.

TABLE 2

Respiratory control and ADP-to-O ratios

LeanObese
RCR   
    P + M 20.5 ± 1.8 17.4 ± 1.3 
    PC + M 9.1 ± 1.6 11.4 ± 1.7 
    P + M + GLU 12.9 ± 2.4 14.1 ± 1.4 
    GLU + MAL 16.1 ± 1.4 13.7 ± 2.3 
    P + M + GLU + SUCC 10.0 ± 0.6 10.7 ± 1.2 
ADP/O   
    P + M 2.30 ± 0.03 2.26 ± 0.05 
    PC + M 1.94 ± 0.05 1.88 ± 0.06 
    P + M+ GLU 2.14 ± 0.06 2.03 ± 0.06 
    GLU + MAL 2.12 ± 0.08 2.02 ± 0.07 
    P + M + GLU + SUCC 1.77 ± 0.07 1.65 ± 0.03 
LeanObese
RCR   
    P + M 20.5 ± 1.8 17.4 ± 1.3 
    PC + M 9.1 ± 1.6 11.4 ± 1.7 
    P + M + GLU 12.9 ± 2.4 14.1 ± 1.4 
    GLU + MAL 16.1 ± 1.4 13.7 ± 2.3 
    P + M + GLU + SUCC 10.0 ± 0.6 10.7 ± 1.2 
ADP/O   
    P + M 2.30 ± 0.03 2.26 ± 0.05 
    PC + M 1.94 ± 0.05 1.88 ± 0.06 
    P + M+ GLU 2.14 ± 0.06 2.03 ± 0.06 
    GLU + MAL 2.12 ± 0.08 2.02 ± 0.07 
    P + M + GLU + SUCC 1.77 ± 0.07 1.65 ± 0.03 

Data are presented as means ± SE. Lean/obese n: pyruvate (P) plus malate (M), 18/12; palmitoyl-dl-carnitine (PC) plus malate, 16/10; pyruvate plus malate plus glutamate (G), 5/11; glutamate plus malate, 7/5; pyruvate plus malate plus glutamate plus succinate (Succ), 7/7.

FIG. 1.

In vitro oxidative capacity of isolated mitochondria. Mitochondrial fractions were rapidly isolated from basal vastus lateralis biopsies (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants). A: State 3 respiration rate was measured in response to 0.5 mmol/l ADP. B: State 4 respiration rate was measured following the consumption of 0.5 mmol/l ADP. (See research design and methods for concentrations.) Data are presented as means ± SEM in units of nmol min−1 mg protein−1. Lean/obese n: pyruvate (P) plus malate (M), 18/12; palmitoyl-dl-carnitine (PC) plus malate, 16/10; pyruvate plus malate plus glutamate (G), 5/11; glutamate plus malate, 7/5; pyruvate plus malate plus glutamate plus succinate (Succ), 7/7.

FIG. 1.

In vitro oxidative capacity of isolated mitochondria. Mitochondrial fractions were rapidly isolated from basal vastus lateralis biopsies (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants). A: State 3 respiration rate was measured in response to 0.5 mmol/l ADP. B: State 4 respiration rate was measured following the consumption of 0.5 mmol/l ADP. (See research design and methods for concentrations.) Data are presented as means ± SEM in units of nmol min−1 mg protein−1. Lean/obese n: pyruvate (P) plus malate (M), 18/12; palmitoyl-dl-carnitine (PC) plus malate, 16/10; pyruvate plus malate plus glutamate (G), 5/11; glutamate plus malate, 7/5; pyruvate plus malate plus glutamate plus succinate (Succ), 7/7.

Close modal

ROS production.

Rates of ROS production were measured in the presence of oligomycin in a situation where the ETC complexes were highly reduced by the influx of electrons provided by NADH (pyruvate and PC) and FADH2 (PC). Theoretical sites of ROS production are shown in Fig. 2,A. With pyruvate and PC (both including malate), insulin-resistant muscle mitochondria produced ROS at higher rates than did insulin-sensitive muscle (Fig. 2 B). Rotenone addition to inhibit passage of electrons from complex I to ubiquinone typically results in increased ROS production (23). However, rotenone raised ROS production only in the mitochondria from insulin-sensitive individuals.

FIG. 2.

In vitro ROS production rates. A: Diagram indicating the electron transport chain complexes along with their substrates (NADH, FADH2, ubiquinone, or cytochrome c) and the inhibitors that target them (red text: rotenone, antimycin A, myxothyazole, and oligomycin). Complexes I and III contain ROS-generating sites (red arrows). Black full arrows show the path of electron flow within the inner mitochondrial membrane (IMM), and black dashed arrows show the directional path of protons across the IMM. Electrons can flow from complex II to I (reverse electron flow) and produce ROS; rotenone blocks this flow from complex II. ROS production by complex I is enhanced in the presence of rotenone because rotenone's binding site is downstream of the ROS-producing site in this complex. B: ROS production rates were quantified by incubating isolated mitochondria (10 μg) with various substrate combinations, Amplex Red, and horseradish peroxidase (see research design and methods for detailed procedures). (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants.) Data are presented as means ± SEM; n = 7 per substrate combination. **P < 0.05 vs. lean participants. G, glutamate; M, malate; O, 1 μg/ml oligomycin; P, pyruvate; PC, palmitoyl-dl-carnitine; R, 10 μmol/l rotenone; Succ, succinate.

FIG. 2.

In vitro ROS production rates. A: Diagram indicating the electron transport chain complexes along with their substrates (NADH, FADH2, ubiquinone, or cytochrome c) and the inhibitors that target them (red text: rotenone, antimycin A, myxothyazole, and oligomycin). Complexes I and III contain ROS-generating sites (red arrows). Black full arrows show the path of electron flow within the inner mitochondrial membrane (IMM), and black dashed arrows show the directional path of protons across the IMM. Electrons can flow from complex II to I (reverse electron flow) and produce ROS; rotenone blocks this flow from complex II. ROS production by complex I is enhanced in the presence of rotenone because rotenone's binding site is downstream of the ROS-producing site in this complex. B: ROS production rates were quantified by incubating isolated mitochondria (10 μg) with various substrate combinations, Amplex Red, and horseradish peroxidase (see research design and methods for detailed procedures). (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants.) Data are presented as means ± SEM; n = 7 per substrate combination. **P < 0.05 vs. lean participants. G, glutamate; M, malate; O, 1 μg/ml oligomycin; P, pyruvate; PC, palmitoyl-dl-carnitine; R, 10 μmol/l rotenone; Succ, succinate.

Close modal

Mitochondrial sensitivity to metabolic control signals.

Steady-state respiration rates spanning the aerobic range were established by experimentally manipulating extramitochondrial energy phosphate levels (ADP) with the creatine kinase energy clamp. Supplemental Fig. 1 shows a typical experiment using stepwise addition of phosphocreatine. Eadie-Hofstee plots of Jo2 vs. Jo2/[ADP] yielded similar values for the KMADP for respiration (35 ± 2 μmol/l for lean vs. 31 ± 4 for obese; P = 0.33). Figure 3,A shows a typical relationship between ΔGATP and oxygen consumption. The ΔGATP:Jo2 (force flow) relation extrapolated through the abscissa indicated a higher (more negative) static head ΔGATP in insulin-resistant muscle mitochondria (P = 0.003) (Fig. 3 B). The slope of the force-flow relationship in the lean group showed a trend to be higher than that in the obese group (95.4 ± 7.1 vs. 75.4 ± 8.8 nmol O2 · min−1 · kcal−1 · mol−1 (P ≤ 0.10) (supplemental Fig. 3).

FIG. 3.

Static head ΔGATP. Isolated mitochondria were challenged with pyruvate plus malate at different PCr-to-Cr ratios. A: schematic representation of the Jo2 vs. ΔGATP plot. B: The static head ΔGATP was calculated by extrapolating the curve of the Jo2 vs. ΔGATP curve to Jo2 = 0. (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants.) Data are presented as means ± SEM; n = 7 per group. Student's t test was performed. *P < 0.01 vs. lean participants.

FIG. 3.

Static head ΔGATP. Isolated mitochondria were challenged with pyruvate plus malate at different PCr-to-Cr ratios. A: schematic representation of the Jo2 vs. ΔGATP plot. B: The static head ΔGATP was calculated by extrapolating the curve of the Jo2 vs. ΔGATP curve to Jo2 = 0. (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants.) Data are presented as means ± SEM; n = 7 per group. Student's t test was performed. *P < 0.01 vs. lean participants.

Close modal

Differences in mitochondrial protein abundance.

The number of unique proteins assigned per mitochondrial isolate was similar between the groups (lean 425 ± 19 vs. obese 422 ± 29 proteins; P = n.s.), as were the number of total known mitochondrial proteins (lean 248 ± 8 vs. obese 236 ± 13 mitochondrial proteins; P = n.s.). A composite list assigned 691 unique proteins. In addition, the total spectra generated solely by known mitochondrial proteins did not differ between groups (lean 19,683 ± 1,421 vs. obese 18889 ± 1,238; P = n.s.). Approximately 60% of total spectra were assigned to known mitochondrial proteins.

Electron chain complexes.

Table 3 depicts the number of assigned subunits in at least half of all participants. Complex II (succinate dehydrogenase) and III (ubiquinone: cytochrome c reductase) subunits were present in equal numbers per mitochondrial mass in insulin-sensitive and insulin-resistant muscle. A number of complex I subunits (P ≤ 0.05: NDUFC2, NDUFA8, NDUFA11, NDUFS2, and NDUFB8; P ≤ 0.10: NDUFS5, NDUFB3, and MTND1) were observed to be lower in abundance in mitochondria from the obese group. The iron-sulfur cluster scaffold protein (NFU1) was also lower in the obese group (P < 0.05). Only one subunit from complex IV was significantly decreased in obesity: COX5B (Table 4).

TABLE 3

Detected ETC subunits by tandem mass spectrometry

ComplexNumber of known subunitsAssigned in 16 participantsAssigned in 8 participants
46 27 34 
II 
III 
IV 16 (plus SURF1 and SCO1) 
19 (plus ATPAF1/2) 10 13 
ComplexNumber of known subunitsAssigned in 16 participantsAssigned in 8 participants
46 27 34 
II 
III 
IV 16 (plus SURF1 and SCO1) 
19 (plus ATPAF1/2) 10 13 

Number of electron transport chain subunits confidently assigned in half or all subjects.

TABLE 4

Protein abundance differing significantly in isolated mitochondria from insulin-resistant skeletal muscle of obese subjects

Protein name (gene name)Fold change in mitochondria from obese (log2)P (Student's t test)Lean/obese subjects with protein assigned (n)Total spectra per Group
LeanObese
Myosin-2 (MYH20.92 ± 0.16 <0.01 8/8 2,664 4,821 
Myosin-1 (MYH11.74 ± 0.23 <0.01 8/8 1,691 4,239 
Methylmalonate-semialdehyde dehydrogenase [acylating], mitochondrial precursor (ALDH6A1−1.05 ± 0.12 <0.01 7/5 64 22 
Isoform SERCA1A/B of sarcoplasmic/endoplasmic reticulum calcium ATPase 1 (ATP2A10.75 ± 0.15 <0.01 8/8 2,433 3,748 
Glycogen phosphorylase, muscle form (PYGM1.40 ± 0.20 <0.01 6/5 19 38 
Peroxiredoxin-2 (PRDX20.93 ± 0.20 <0.01 8/7 72 113 
Prenylcysteine oxidase 1 precursor (PCYOX11.07 ± 0.18 0.01 4/6 11 32 
Calsequestrin-2 precursor (CASQ2−0.71 ± 0.11 0.01 8/8 295 169 
Dihydropyridine receptor α2 subunit (CACNA2D11.17 ± 0.32 0.02 7/7 47 102 
Apolipoprotein A-I precursor (APOA11.04 ± 0.16 0.02 7/8 56 100 
NADH dehydrogenase [ubiquinone] 1 subunit C2 (NDUFC2−0.58 ± 0.16 0.02 8/8 57 37 
Carnitine O-palmitoyltransferase I, muscle isoform (CPT1B−0.81 ± 0.32 0.03 8/7 114 62 
NADH dehydrogenase [ubiquinone] 1α subcomplex subunit 8 (NDUFA8−0.30 ± 0.12 0.03 8/8 1,487 1,152 
NADH dehydrogenase [ubiquinone] 1α subcomplex subunit 11 (NDUFA11−0.50 ± 0.16 0.04 8/8 259 176 
Similar to CG7601-PA (DHRS7C0.63 ± 0.14 0.04 8/7 64 82 
Isoform 1 of tripartite motif-containing protein 72 (TRIM72−0.85 ± 0.37 0.04 8/6 32 15 
Cytochrome c oxidase subunit 5B, mitochondrial precursor (COX5B−0.33 ± 0.09 0.04 8/8 576 428 
NADH dehydrogenase [ubiquinone] iron-sulfur protein 2 (NDUFS2−0.20 ± 0.05 0.05 8/8 426 349 
Propionyl-CoA carboxylase β chain, mitochondrial precursor (PCCB−0.80 ± 0.21 0.05 5/5 11 
NADH dehydrogenase;NADH dehydrogenase [ubiquinone] 1β subcomplex subunit 8 (NDUFB8−0.31 ± 0.13 0.05 8/8 244 188 
Erythrocyte band 7 integral membrane protein (STOM1.24 ± 0.42 0.05 6/8 24 78 
Protein name (gene name)Fold change in mitochondria from obese (log2)P (Student's t test)Lean/obese subjects with protein assigned (n)Total spectra per Group
LeanObese
Myosin-2 (MYH20.92 ± 0.16 <0.01 8/8 2,664 4,821 
Myosin-1 (MYH11.74 ± 0.23 <0.01 8/8 1,691 4,239 
Methylmalonate-semialdehyde dehydrogenase [acylating], mitochondrial precursor (ALDH6A1−1.05 ± 0.12 <0.01 7/5 64 22 
Isoform SERCA1A/B of sarcoplasmic/endoplasmic reticulum calcium ATPase 1 (ATP2A10.75 ± 0.15 <0.01 8/8 2,433 3,748 
Glycogen phosphorylase, muscle form (PYGM1.40 ± 0.20 <0.01 6/5 19 38 
Peroxiredoxin-2 (PRDX20.93 ± 0.20 <0.01 8/7 72 113 
Prenylcysteine oxidase 1 precursor (PCYOX11.07 ± 0.18 0.01 4/6 11 32 
Calsequestrin-2 precursor (CASQ2−0.71 ± 0.11 0.01 8/8 295 169 
Dihydropyridine receptor α2 subunit (CACNA2D11.17 ± 0.32 0.02 7/7 47 102 
Apolipoprotein A-I precursor (APOA11.04 ± 0.16 0.02 7/8 56 100 
NADH dehydrogenase [ubiquinone] 1 subunit C2 (NDUFC2−0.58 ± 0.16 0.02 8/8 57 37 
Carnitine O-palmitoyltransferase I, muscle isoform (CPT1B−0.81 ± 0.32 0.03 8/7 114 62 
NADH dehydrogenase [ubiquinone] 1α subcomplex subunit 8 (NDUFA8−0.30 ± 0.12 0.03 8/8 1,487 1,152 
NADH dehydrogenase [ubiquinone] 1α subcomplex subunit 11 (NDUFA11−0.50 ± 0.16 0.04 8/8 259 176 
Similar to CG7601-PA (DHRS7C0.63 ± 0.14 0.04 8/7 64 82 
Isoform 1 of tripartite motif-containing protein 72 (TRIM72−0.85 ± 0.37 0.04 8/6 32 15 
Cytochrome c oxidase subunit 5B, mitochondrial precursor (COX5B−0.33 ± 0.09 0.04 8/8 576 428 
NADH dehydrogenase [ubiquinone] iron-sulfur protein 2 (NDUFS2−0.20 ± 0.05 0.05 8/8 426 349 
Propionyl-CoA carboxylase β chain, mitochondrial precursor (PCCB−0.80 ± 0.21 0.05 5/5 11 
NADH dehydrogenase;NADH dehydrogenase [ubiquinone] 1β subcomplex subunit 8 (NDUFB8−0.31 ± 0.13 0.05 8/8 244 188 
Erythrocyte band 7 integral membrane protein (STOM1.24 ± 0.42 0.05 6/8 24 78 

Data are means ± SE unless otherwise indicated. Total spectra = sum of spectra across participants within the group.

Fatty acid oxidation and amino acid metabolism.

Carnitine palmitoyltransferase 1B (CPT1B) was significantly lower in the obese group. Methyl-malonate-semialdehyde dehydrogenase (ALDH6A1) and propionyl-CoA carboxylase β were also lower in mitochondria from the obese group. These proteins are involved in valine and valine/isoleucine oxidation.

Fiber type and related proteins.

MYH2, coding for myosin type IIa, was higher in the obese versus lean participants; MYH7 (myosin type I) was unchanged. MYH1, coding for myosin type IIx, was higher in the obese participants. In agreement with the myosin isoforms, the Ca2+ ATPase fast twitch isoform (ATP2A1) was significantly higher and the slow twitch isoform (ATP2A2) tended to be lower (P = 0.07) in mitochondria from obese individuals. CASQ2, the major calsequestrin isoform in cardiac muscle, comprises 50% of the calsequestrin population in slow oxidative muscle fibers. CASQ2 relative abundance was lower in mitochondria from obese individuals.

Comparison of protein abundance differences identified by mass spectrometry with immunoblot analysis.

Mitochondrial lysates were probed for COXIV abundance, and no difference was observed between lean and obese groups (P = 0.4), as predicted by the mass spectrometry data (Fig. 4). ALDH6A1 abundance normalized to COXIV (P = 0.06) (ALDH6A1 abundance without normalization to the COXIV loading control resulted in P = 0.04) showed a 59% lower abundance in obesity by immunoblot—comparable with the 52% lower abundance determined by mass spectrometry.

FIG. 4.

Comparison of protein abundances determined by mass spectrometry to immunoblotting. Mitochondrial lysates (20 μg) (n = 3–4) were probed for ALDH6A1 and COXIV abundances. ALDH6A1 abundance was normalized to a loading control (COXIV). (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants.) *P = 0.06.

FIG. 4.

Comparison of protein abundances determined by mass spectrometry to immunoblotting. Mitochondrial lysates (20 μg) (n = 3–4) were probed for ALDH6A1 and COXIV abundances. ALDH6A1 abundance was normalized to a loading control (COXIV). (□, lean insulin-sensitive participants; ■, obese insulin-resistant participants.) *P = 0.06.

Close modal

It has been proposed that excess lipid availability combined with a lower muscular capacity to oxidize fat (24) may lead to accumulation of intramyocellular lipids (25). This in turn is thought to lead to insulin-signaling abnormalities and insulin resistance (1). Lower oxidative capacity could be due to lower mitochondrial content, a defect intrinsic to the individual mitochondrion, or both. Mitochondrial content was significantly lower in the obese individuals, consistent with previous reports (2,,5). To determine the presence of intrinsic defects, mitochondria were isolated from lean, insulin-sensitive and obese, insulin-resistant human vastus lateralis muscle. The lower mitochondrial content but similar Vo2max values in the obese versus lean volunteers may appear contradictory. However, Vo2max primarily reflects cardiovascular, rather than mitochondrial, parameters (26,27).

With regard to intrinsic characteristics of the mitochondria, we used the substrate combinations of pyruvate, glutamate, and pyruvate plus glutamate (all including malate), which rely on complex I activity, whereas the addition of succinate recruits complex II activity. Fatty acid oxidation requires complexes I and II and the electron transfer flavoprotein shuttling electrons from dehydrogenases within the β-oxidation pathway to the electron transport chain (ETC). Basal and maximal respiration rates, regardless of substrates, did not differ in the obese participants—a result similar to that in a recent study (11). These findings are in accordance with some (9,28) but not all (8,10) studies assessing mitochondrial function in insulin resistance and type 2 diabetes. Differences between the present results and findings of decreased mitochondrial function in diabetes could be due to the possibility that insulin resistance itself is not associated with gross changes in mitochondrial function or mitochondrial protein abundance; rather, mitochondrial dysfunction may be more associated with hyperglycemia. Results of a whole muscle proteomics comparison of lean and obese normoglycemic and type 2 diabetic individuals showed that decreases in ETC components and other mitochondrial proteins are most pronounced in the diabetics, with more moderate changes in the obese group (22). In addition, most of the earlier studies did not report aerobic fitness measurements (8,9,28), whereas one study did match the control and diabetic subjects for whole-body aerobic power (Vo2max) (10). The lean and obese individuals in the present study had comparable Vo2max and glycemia, and their muscle mitochondria displayed similar maximal substrate oxidation rates, coupling, and respiratory control—in contrast to the study of Phielix et al. (10). These differences suggest that the progression from insulin resistance to type 2 diabetes may result in functional defects in SM mitochondria. On the other hand, our data suggest that reports of decreased in vivo ATP turnover in resting insulin-resistant muscle (6,7,29) most likely reflect decreased cellular ATP demand, given that 1) in vivo mitochondria rest at <2% of their oxidative potential (30), 2) the differences in observed mitochondrial content were rather small, and 3) insulin resistance appears to improve, not degrade, the cellular energy state defended by muscle mitochondria. If higher ΔGATP had any influence on resting ATP turnover, it would be predicted to be in the opposite direction (31). Our results show that there are no intrinsic defects in mitochondrial oxidative capacity in insulin resistance without hyperglycemia. Taken together, these data indicate that with regard to substrate oxidation, the findings of decreased ATP turnover in vivo (6,7,29) likely are explained on a gross level by a decrease in mitochondrial content in insulin-resistant muscle.

Nevertheless, there were more subtle intrinsic differences in the function of mitochondria from insulin-resistant muscle. In resting muscle, mitochondrial oxidative flux is extremely low (32), with very low [ADP]f and robust (highly negative) ΔGATP (31). Contractile activity can elicit extremely high oxidative flux, revealing a cellular aerobic scope spanning two orders of magnitude (30). The metabolic signals that control respiration over this range include both kinetic (e.g., [ADP] or KMADP) and thermodynamic (ΔGATP) components. The creatine kinase energy clamp provides an in vitro evaluation of the mitochondrial sensitivity to these signals. In the present study, the KMADP determined using the creatine kinase clamp for respiration was not different between groups. However, extrapolation of the thermodynamic force-flow relation to the hypothetical zero Jo2 resulted in significantly more negative (higher) static head ΔGATP in mitochondria isolated from insulin-resistant muscle. This finding is consistent with the observation of Petersen et al. (6), who used 31P–nuclear magnetic resonance to show that the inorganic phosphate–to–phosphocreatine ratio is lower in insulin-resistant than in insulin-sensitive muscle. Lower inorganic phosphate and more robust ΔGATP are characteristics that differentiate type II from type I mammalian muscle cells (33). It is therefore noteworthy that insulin-resistant muscle showed trends toward greater expression of type II myocyte proteins, e.g., MYH2 (type IIa), MYH1 (type IIx), ATP2A1 (Ca2+ ATPase fast twitch isoform), and higher PYGM (glycogen phosphorylase).

Taken together, these data suggest that, at the low metabolic turnover rates associated with rest, mitochondria from insulin-resistant muscle maintain higher thermodynamic driving forces. Because the protonmotive force and the oxidation-reduction energy (redox pressure) down the electron transport chain are the energetic precursors to ΔGATP, higher ROS production rates by mitochondria from the obese group would be the predicted outcome. This result was experimentally observed. In the absence of rotenone (Fig. 2 A), ROS production was higher in the obese individuals, suggesting that muscle mitochondria from obese individuals produce more ROS at complex I and/or complex III. Complex III has been labeled as a higher contributor of ROS in SM mitochondria (14). Secondly, the level of long-chain fatty acid oxidation intermediates within the matrix, which have been postulated to be elevated in insulin resistance due to incomplete β-oxidation (34), dictates the amount of ROS produced (35) specifically at the level of complex III and the electron transfer flavoprotein (35). Higher rates of ROS production in insulin resistance have been reported previously (16) and precede the onset of the diabetic state in mice on a high-fat diet (36). Lower rates of ROS production in obese individuals, but not type 2 diabetic individuals, have been reported (37). However, this study did not report maximal ROS production rates because mitochondria were not incubated with oligomycin during the measurements. On the other hand, the insulin-sensitizing effect of exercise training is abrogated upon consumption of antioxidants, supporting a positive role for ROS in the modulation of insulin sensitivity (38). Additionally, mice lacking glutathione peroxidase are protected from insulin resistance induced by high-fat diets (39). Therefore, it is possible that higher ROS production rates observed in isolated mitochondria from obese individuals may be a compensatory mechanism to increase insulin action; this possibility warrants further investigation because ROS have also been identified as a causal factor of insulin resistance (12).

In order to guide future research efforts studying the role of the mitochondrion in insulin resistance, we performed an analysis of the abundance of individual proteins in mitochondria (22,40). We used equal amounts of isolated mitochondrial protein for proteomic analysis, resulting in a comparison of relative protein abundance per mitochondrial mass. Several individual subunits of ETC complex I were significantly reduced in obese, insulin-resistant individuals. Complex I is a 980 kDa conglomerate comprised of 46 subunits in mammals (41). Fourteen of these 46 subunits are the core subunits, responsible for the location of the redox centers (flavin mononucleotide and eight Fe-S cluters) (41). The remaining 32 subunits are less well characterized: they may be involved in maintaining the structure of the complex, in linkage of the complex to the inner mitochondrial membrane, or even in altering ROS generation (41). Five complex I subunits were significantly lower in the obese individuals. These subunits were distributed among the four subcomplexes of complex I (NDUFB3/B8/C2 in subcomplex Iβ, ND1 in Iγ, NDUFA8/A11/S5 in Iα, and NDUFS2 in Iλ) (41). A negative correlation has previously been found between the degree of insulin resistance and complex I activity (4). NDUFS2 and ND1 are two subunits found within the core of complex I. A lower abundance of these two subunits did not translate to a detectable decrease in NADH-linked respiration.

The present study also shows a lower abundance of the muscle outer-mitochondrial membrane carnitine palmitoyltransferase 1B (CPT1B). Cpt1B transesterifies long-chain fatty acyl-CoAs to fatty acylcarnitines, which are transported into the mitochondrial matrix for β-oxidation (42). Its expression is dependent on peroxisome proliferator–activated receptor coactivator-1α (PGC-1α) (43) transcription factor activity. Reduced Cpt1B protein level could be a direct consequence of lower PGC-1α gene expression reported in insulin-resistant muscle (44,45). Furthermore, increasing Cpt1B protein levels in SM increased fatty acid oxidation and lowered the severity of insulin resistance in rats on a high-fat diet (46). In the present study, the utilization of PC in the assessment of fatty acid–mediated mitochondrial respiration bypassed the highly regulated CPT system. Our results therefore show that β-oxidation activity per se is not impaired in insulin-resistant muscle but, rather, that the protein determining the rate of fatty acid entry into mitrochondria, Cpt1B, was decreased.

The last set of proteins altered in the proteomics analysis involved enzymes involved in BCAA catabolism. MMS dehydrogenase and propionyl-CoA carboxylase are required for valine and isoleucine metabolism before entry into the tricarboxylic acid (TCA) cycle as succinyl-CoA. Decreased BCAA oxidation, together with increased proteolysis in insulin resistance (47), could lead to increased intracellular BCAA concentrations. Such an increase could activate mammalian target of rapamycin signaling, which can inhibit insulin signaling and produce insulin resistance (48). Moreover, recent metabolomic studies in humans and BCAA/high-fat feeding studies in rodents indicate that alterations in BCAA metabolism are associated with insulin resistance (49,50). Alternatively, or in addition, it is possible that decreased entry of BCAA into the TCA cycle could reduce TCA cycle flux and lead to accumulation of incompletely oxidized fatty acids, as suggested by Koves and Muoio (34). A decreased CPT1B abundance then could be a compensatory effort to protect the cell from accumulation of incompletely oxidized fatty acids and other fatty acid metabolites. A caveat is that this scenario would imply a rate-determining role of BCAA-derived TCA intermediates.

In conclusion, mitochondria isolated from skeletal muscle from obese and overweight insulin-resistant volunteers do not have an intrinsic abnormality in fuel oxidation. However, measurements of citrate synthase activity and mitochondrial protein indicated a decrease in mitochondrial mass. Therefore, in order to maintain similar respiration rates, we propose that mitochondria in insulin-resistant muscle undergo adaptations that allow them to generate a more negative ΔGATP. A consequence of this higher ΔGATP would be increased ROS production rates at rest (51) as found here. We propose a model for these data shown in Fig. 5.

FIG. 5.

Proposed bioenergetic model occurring in skeletal muscle insulin resistance. In insulin-resistant skeletal muscle, high catalytic potential in the redox-producing block (i.e., NADH levels) owing to increased citric acid cycle dehydrogenase activity; (isocitrate dehydrogenase [when considering its three subunits] abundance is higher in mitochondria from insulin-resistant muscle [data not shown]) relative to the redox-utilizing block (i.e., ETC, which is particularly impaired by complex I in insulin resistance), which would lead to the following: 1) higher static head energy (at low flux, high-redox pressure developed by the dehydrogenase is kinetically allowed to achieve near equilibration with ATP energy downstream), 2) lower ΔGATP:Jo slope (the impediment to electron flow is exposed by relaxation of downstream ΔGATP energy and transition to higher flux), and 3) higher ROS production due to higher driving force (more negative redox potential of electron transfer centers). The protonmotive force (Δp) was not measured and was included for completeness of the model. No differences in the abundance of ATP synthase F1 component or the adenine nucleotide translocator (ANT) were observed in mitochondria from insulin-resistant muscle by mass spectrometry.

FIG. 5.

Proposed bioenergetic model occurring in skeletal muscle insulin resistance. In insulin-resistant skeletal muscle, high catalytic potential in the redox-producing block (i.e., NADH levels) owing to increased citric acid cycle dehydrogenase activity; (isocitrate dehydrogenase [when considering its three subunits] abundance is higher in mitochondria from insulin-resistant muscle [data not shown]) relative to the redox-utilizing block (i.e., ETC, which is particularly impaired by complex I in insulin resistance), which would lead to the following: 1) higher static head energy (at low flux, high-redox pressure developed by the dehydrogenase is kinetically allowed to achieve near equilibration with ATP energy downstream), 2) lower ΔGATP:Jo slope (the impediment to electron flow is exposed by relaxation of downstream ΔGATP energy and transition to higher flux), and 3) higher ROS production due to higher driving force (more negative redox potential of electron transfer centers). The protonmotive force (Δp) was not measured and was included for completeness of the model. No differences in the abundance of ATP synthase F1 component or the adenine nucleotide translocator (ANT) were observed in mitochondria from insulin-resistant muscle by mass spectrometry.

Close modal

We also defined new abnormalities in mitochondria from human insulin-resistant muscle using mass spectrometry. A number of complex I subunits were found to be lower in mitochondria from insulin-resistant muscle, but the lack of characterization of function of most complex I subunits makes it difficult to interpret the consequences of this finding. However, these differences may be related to differences in ROS production. Mitochondria from insulin-resistant muscle also was found to contain less carnitine palmitoyltransferase 1β, offering an explanation for the intramuscular lipid accumulation observed in insulin-resistant muscle (25). In addition, enzymes involved in BCAA metabolism were lower in abundance in insulin-resistant muscle. In addition to the lower mitochondrial content in insulin-resistant muscle, perturbations intrinsic to mitochondria could alter ROS production, limit the entry of fatty acids into the matrix and affect the overall oxidation rate of BCAA.

The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

This study was supported in part by National Institutes of Health grants R01DK47936 (to L.J.M.), R01DK66483 (to L.J.M.), R01DK081750 (to Z.Y.), and the American Diabetes Association.

No potential conflicts of interest relevant to this article were reported.

N.L. researched data, contributed to discussion, wrote the manuscript, and reviewed and edited the manuscript. B.G. researched data. B.B. researched data, contributed to discussion, wrote the manuscript, and reviewed and edited the manuscript. W.T.W. researched data, wrote the manuscript, and reviewed and edited the manuscript. Z.B. researched data. E.A.D.F. researched data. C.B. researched data. C.M. researched data. K.H. researched data. Z.Y. researched data, wrote the manuscript, and reviewed and edited the manuscript. L.J.M. researched data, contributed to discussion, wrote the manuscript, and reviewed and edited the manuscript.

The authors gratefully acknowledge the technical assistance of Kenneth Kirschner, Arizona State University, and the excellent nursing skills of Alexandra Meyer, Julie Robbins, and Dianne Denardo, Arizona State University. The authors are particularly indebted to their enthusiastic participants for willingness to participate in this study.

1.
Szendroedi
J
,
Roden
M
.
Ectopic lipids and organ function
.
Curr Opin Lipidol
2009
;
20
:
50
56
2.
Vondra
K
,
Rath
R
,
Bass
A
,
Slabochova
Z
,
Teisinger
J
,
Vitek
V
.
Enzyme activities in quadriceps femoris muscle of obese diabetic male patients
.
Diabetologia
1977
;
13
:
527
529
3.
Morino
K
,
Petersen
KF
,
Dufour
S
,
Befroy
D
,
Frattini
J
,
Shatzkes
N
,
Neschen
S
,
White
MF
,
Bilz
S
,
Sono
S
,
Pypaert
M
,
Shulman
GI
.
Reduced mitochondrial density and increased IRS-1 serine phosphorylation in muscle of insulin-resistant offspring of type 2 diabetic parents
.
J Clin Invest
2005
;
115
:
3587
3593
4.
Kelley
DE
,
He
J
,
Menshikova
EV
,
Ritov
VB
.
Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes
.
Diabetes
2002
;
51
:
2944
2950
5.
Ritov
VB
,
Menshikova
EV
,
He
J
,
Ferrell
RE
,
Goodpaster
BH
,
Kelley
DE
.
Deficiency of subsarcolemmal mitochondria in obesity and type 2 diabetes
.
Diabetes
2005
;
54
:
8
14
6.
Petersen
KF
,
Dufour
S
,
Befroy
D
,
Garcia
R
,
Shulman
GI
.
Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes
.
N Engl J Med
2004
;
350
:
664
671
7.
Petersen
KF
,
Dufour
S
,
Shulman
GI
.
Decreased insulin-stimulated ATP synthesis and phosphate transport in muscle of insulin-resistant offspring of type 2 diabetic parents
.
PLoS Med
2005
;
2
:
e233
8.
Mogensen
M
,
Sahlin
K
,
Fernstrom
M
,
Glintborg
D
,
Vind
BF
,
Beck-Nielsen
H
,
Hojlund
K
.
Mitochondrial respiration is decreased in skeletal muscle of patients with type 2 diabetes
.
Diabetes
2007
;
56
:
1592
1599
9.
Boushel
R
,
Gnaiger
E
,
Schjerling
P
,
Skovbro
M
,
Kraunsoe
R
,
Dela
F
.
Patients with type 2 diabetes have normal mitochondrial function in skeletal muscle
.
Diabetologia
2007
;
50
:
790
796
10.
Phielix
E
,
Schrauwen-Hinderling
VB
,
Mensink
M
,
Lenaers
E
,
Meex
R
,
Hoeks
J
,
Kooi
ME
,
Moonen-Kornips
E
,
Sels
JP
,
Hesselink
MK
,
Schrauwen
P
.
Lower intrinsic ADP-stimulated mitochondrial respiration underlies in vivo mitochondrial dysfunction in muscle of male type 2 diabetic patients
.
Diabetes
2008
;
57
:
2943
2949
11.
Karakelides
H
,
Irving
BA
,
Short
KR
,
O'Brien
P
,
Nair
KS
:
Age, obesity, and sex effects on insulin sensitivity and skeletal muscle mitochondrial function
.
Diabetes
2009
;
59
:
89
97
12.
Houstis
N
,
Rosen
ED
,
Lander
ES
.
Reactive oxygen species have a causal role in multiple forms of insulin resistance
.
Nature
2006
;
440
:
944
948
13.
Droge
W
.
Free radicals in the physiological control of cell function
.
Physiol Rev
2002
;
82
:
47
95
14.
Tahara
EB
,
Navarete
FD
,
Kowaltowski
AJ
.
Tissue-, substrate-, and site-specific characteristics of mitochondrial reactive oxygen species generation
.
Free Radic Biol Med
2009
;
46
:
1283
1297
15.
Blendea
MC
,
Jacobs
D
,
Stump
CS
,
McFarlane
SI
,
Ogrin
C
,
Bahtyiar
G
,
Stas
S
,
Kumar
P
,
Sha
Q
,
Ferrario
CM
,
Sowers
JR
.
Abrogation of oxidative stress improves insulin sensitivity in the Ren-2 rat model of tissue angiotensin II overexpression
.
Am J Physiol Endocrinol Metab
2005
;
288
:
E353
E359
16.
Anderson
EJ
,
Lustig
ME
,
Boyle
KE
,
Woodlief
TL
,
Kane
DA
,
Lin
CT
,
Price
JW
 3rd
,
Kang
L
,
Rabinovitch
PS
,
Szeto
HH
,
Houmard
JA
,
Cortright
RN
,
Wasserman
DH
,
Neufer
PD
.
Mitochondrial H2O2 emission and cellular redox state link excess fat intake to insulin resistance in both rodents and humans
.
J Clin Invest
2
Feb
2009
[
Epub ahead of print
]
17.
Rockl
KS
,
Hirshman
MF
,
Brandauer
J
,
Fujii
N
,
Witters
LA
,
Goodyear
LJ
.
Skeletal muscle adaptation to exercise training: AMP-activated protein kinase mediates muscle fiber type shift
.
Diabetes
2007
;
56
:
2062
2069
18.
Srere
PA
.
Citrate synthase
.
Methods Enzymol
1969
;
13
:
3
5
19.
Messer
JI
,
Jackman
MR
,
Willis
WT
.
Pyruvate and citric acid cycle carbon requirements in isolated skeletal muscle mitochondria
.
Am J Physiol Cell Physiol
2004
;
286
:
C565
C572
20.
Teague
WE
 Jr
,
Dobson
GP
.
Effect of temperature on the creatine kinase equilibrium
.
J Biol Chem
1992
;
267
:
14084
14093
21.
Lefort
N
,
Yi
Z
,
Bowen
B
,
Glancy
B
,
De Filippis
EA
,
Mapes
R
,
Hwang
H
,
Flynn
CR
,
Willis
WT
,
Civitarese
A
,
Hojlund
K
,
Mandarino
LJ
.
Proteome profile of functional mitochondria from human skeletal muscle using one-dimensional gel electrophoresis and HPLC-ESI-MS/MS
.
J Proteomics
2009
;
72
:
1046
1060
22.
Hwang
H
,
Bowen
BP
,
Lefort
N
,
Flynn
CR
,
De Filippis
EA
,
Roberts
C
,
Smoke
CC
,
Meyer
C
,
Hojlund
K
,
Yi
Z
,
Mandarino
LJ
.
Proteomics analysis of human skeletal muscle reveals novel abnormalities in obesity and type 2 diabetes
.
Diabetes
2009
;
59
:
33
42
23.
Liu
Y
,
Fiskum
G
,
Schubert
D
.
Generation of reactive oxygen species by the mitochondrial electron transport chain
.
J Neurochem
2002
;
80
:
780
787
24.
Kelley
DE
,
Goodpaster
B
,
Wing
RR
,
Simoneau
JA
.
Skeletal muscle fatty acid metabolism in association with insulin resistance, obesity, and weight loss
.
Am J Physiol
1999
;
277
:
E1130
E1141
25.
Goodpaster
BH
,
He
J
,
Watkins
S
,
Kelley
DE
.
Skeletal muscle lipid content and insulin resistance: evidence for a paradox in endurance-trained athletes
.
J Clin Endocrinol Metab
2001
;
86
:
5755
5761
26.
Saltin
B
,
Strange
S
.
Maximal oxygen uptake: “old” and “new” arguments for a cardiovascular limitation
.
Med Sci Sports Exerc
1992
;
24
:
30
37
27.
Wagner
PD
.
Determinants of maximal oxygen transport and utilization
.
Annu Rev Physiol
1996
;
58
:
21
50
28.
Bjorntorp
P
,
Schersten
T
,
Fagerberg
SE
.
Respiration and phosphorylation of mitochondria isolated from the skeletal muscle of diabetic and normal subjects
.
Diabetologia
1967
;
3
:
346
352
29.
Befroy
DE
,
Petersen
KF
,
Dufour
S
,
Mason
GF
,
de Graaf
RA
,
Rothman
DL
,
Shulman
GI
.
Impaired mitochondrial substrate oxidation in muscle of insulin-resistant offspring of type 2 diabetic patients
.
Diabetes
2007
;
56
:
1376
1381
30.
Richardson
RS
,
Knight
DR
,
Poole
DC
,
Kurdak
SS
,
Hogan
MC
,
Grassi
B
,
Wagner
PD
.
Determinants of maximal exercise VO2 during single leg knee-extensor exercise in humans
.
Am J Physiol
1995
;
268
:
H1453
H1461
31.
Jeneson
JA
,
Westerhoff
HV
,
Kushmerick
MJ
.
A metabolic control analysis of kinetic controls in ATP free energy metabolism in contracting skeletal muscle
.
Am J Physiol Cell Physiol
2000
;
279
:
C813
C832
32.
Kirkwood
SP
,
Zurlo
F
,
Larson
K
,
Ravussin
E
.
Muscle mitochondrial morphology, body composition, and energy expenditure in sedentary individuals
.
Am J Physiol
1991
;
260
:
E89
E94
33.
Kushmerick
MJ
,
Moerland
TS
,
Wiseman
RW
.
Mammalian skeletal muscle fibers distinguished by contents of phosphocreatine, ATP, and Pi
.
Proc Natl Acad Sci U S A
1992
;
89
:
7521
7525
34.
Koves
TR
,
Ussher
JR
,
Noland
RC
,
Slentz
D
,
Mosedale
M
,
Ilkayeva
O
,
Bain
J
,
Stevens
R
,
Dyck
JR
,
Newgard
CB
,
Lopaschuk
GD
,
Muoio
DM
.
Mitochondrial overload and incomplete fatty acid oxidation contribute to skeletal muscle insulin resistance
.
Cell Metab
2008
;
7
:
45
56
35.
Seifert
EL
,
Estey
C
,
Xuan
JY
,
Harper
ME
.
Electron transport chain-dependent and -independent mechanisms of mitochondrial H2O2 emission during long-chain fatty acid oxidation
.
J Biol Chem
2009
;
283
:
5748
5758
36.
Bonnard
C
,
Durand
A
,
Peyrol
S
,
Chanseaume
E
,
Chauvin
MA
,
Morio
B
,
Vidal
H
,
Rieusset
J
.
Mitochondrial dysfunction results from oxidative stress in the skeletal muscle of diet-induced insulin-resistant mice
.
J Clin Invest
2008
;
118
:
789
800
37.
Abdul-Ghani
MA
,
Jani
R
,
Chavez
A
,
Molina-Carrion
M
,
Tripathy
D
,
Defronzo
RA
.
Mitochondrial reactive oxygen species generation in obese non-diabetic and type 2 diabetic participants
.
Diabetologia
2009
;
52
:
574
582
38.
Ristow
M
,
Zarse
K
,
Oberbach
A
,
Kloting
N
,
Birringer
M
,
Kiehntopf
M
,
Stumvoll
M
,
Kahn
CR
,
Bluher
M
.
Antioxidants prevent health-promoting effects of physical exercise in humans
.
Proc Natl Acad Sci U S A
2009
;
106
:
8665
8670
39.
Loh
K
,
Deng
H
,
Fukushima
A
,
Cai
X
,
Boivin
B
,
Galic
S
,
Bruce
C
,
Shields
BJ
,
Skiba
B
,
Ooms
LM
,
Stepto
N
,
Wu
B
,
Mitchell
CA
,
Tonks
NK
,
Watt
MJ
,
Febbraio
MA
,
Crack
PJ
,
Andrikopoulos
S
,
Tiganis
T
.
Reactive oxygen species enhance insulin sensitivity
.
Cell Metab
2009
;
10
:
260
272
40.
Lefort
N
,
Yi
Z
,
Bowen
B
,
Glancy
B
,
De Filippis
EA
,
Mapes
R
,
Hwang
H
,
Flynn
CR
,
Willis
WT
,
Civitarese
A
,
Hojlund
K
,
Mandarino
LJ
.
Proteome profile of functional mitochondria from human skeletal muscle using one-dimensional gel electrophoresis and HPLC-ESI-MS/MS
.
J Proteomics
2009
;
72
:
1046
1060
41.
Janssen
RJ
,
Nijtmans
LG
,
van den Heuvel
LP
,
Smeitink
JA
.
Mitochondrial complex I: structure, function and pathology
.
J Inherit Metab Dis
2006
;
29
:
499
515
42.
Wang
D
,
Harrison
W
,
Buja
LM
,
Elder
FF
,
McMillin
JB
.
Genomic DNA sequence, promoter expression, and chromosomal mapping of rat muscle carnitine palmitoyltransferase I
.
Genomics
1998
;
48
:
314
323
43.
Kleiner
S
,
Nguyen-Tran
V
,
Bare
O
,
Huang
X
,
Spiegelman
B
,
Wu
Z
.
PPAR{delta} agonism activates fatty acid oxidation via PGC-1{alpha} but does not increase mitochondrial gene expression and function
.
J Biol Chem
2009
;
284
:
18624
18633
44.
Mootha
VK
,
Lindgren
CM
,
Eriksson
KF
,
Subramanian
A
,
Sihag
S
,
Lehar
J
,
Puigserver
P
,
Carlsson
E
,
Ridderstrale
M
,
Laurila
E
,
Houstis
N
,
Daly
MJ
,
Patterson
N
,
Mesirov
JP
,
Golub
TR
,
Tamayo
P
,
Spiegelman
B
,
Lander
ES
,
Hirschhorn
JN
,
Altshuler
D
,
Groop
LC
.
PGC-1alpha-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes
.
Nat Genet
2003
;
34
:
267
273
45.
Patti
ME
,
Butte
AJ
,
Crunkhorn
S
,
Cusi
K
,
Berria
R
,
Kashyap
S
,
Miyazaki
Y
,
Kohane
I
,
Costello
M
,
Saccone
R
,
Landaker
EJ
,
Goldfine
AB
,
Mun
E
,
DeFronzo
R
,
Finlayson
J
,
Kahn
CR
,
Mandarino
LJ
.
Coordinated reduction of genes of oxidative metabolism in humans with insulin resistance and diabetes: Potential role of PGC1 and NRF1
.
Proc Natl Acad Sci U S A
2003
;
100
:
8466
8471
46.
Bruce
CR
,
Hoy
AJ
,
Turner
N
,
Watt
MJ
,
Allen
TL
,
Carpenter
K
,
Cooney
GJ
,
Febbraio
MA
,
Kraegen
EW
.
Overexpression of carnitine palmitoyltransferase-1 in skeletal muscle is sufficient to enhance fatty acid oxidation and improve high-fat diet–induced insulin resistance
.
Diabetes
2009
;
58
:
550
558
47.
Bell
JA
,
Volpi
E
,
Fujita
S
,
Cadenas
JG
,
Sheffield-Moore
M
,
Rasmussen
BB
.
Skeletal muscle protein anabolic response to increased energy and insulin is preserved in poorly controlled type 2 diabetes
.
J Nutr
2006
;
136
:
1249
1255
48.
Tzatsos
A
,
Kandror
KV
.
Nutrients suppress phosphatidylinositol 3-kinase/Akt signaling via raptor-dependent mTOR-mediated insulin receptor substrate 1 phosphorylation
.
Mol Cell Biol
2006
;
26
:
63
76
49.
Newgard
CB
,
An
J
,
Bain
JR
,
Muehlbauer
MJ
,
Stevens
RD
,
Lien
LF
,
Ha
AM
,
Shah
SH
,
Arlotto
M
,
Slentz
CA
,
Rochon
J
,
Gallup
D
,
Ilkayeva
O
,
Wenner
BR
,
Yancy
WS
 Jr
,
Eisenson
H
,
Musante
G
,
Surwit
RS
,
Millington
DS
,
Butler
MD
,
Svetkey
LP
.
A branched-chain amino acid-related metabolic signature that differentiates obese and lean humans and contributes to insulin resistance
.
Cell Metab
2009
;
9
:
311
326
50.
Tremblay
F
,
Jacques
H
,
Marette
A
.
Modulation of insulin action by dietary proteins and amino acids: role of the mammalian target of rapamycin nutrient sensing pathway
.
Curr Opin Clin Nutr Metab Care
2005
;
8
:
457
462
51.
Korshunov
SS
,
Skulachev
VP
,
Starkov
AA
.
High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria
.
FEBS Lett
1997
;
416
:
15
18
Readers may use this article as long as the work is properly cited, the use is educational and not for profit, and the work is not altered. See http://creativecommons.org/licenses/by-nc-nd/3.0/ for details.