ArticlePDF Available

Combined Hyperinsulinemia and Hyperglycemia, But Not Hyperinsulinemia Alone, Suppress Human Skeletal Muscle Lipolytic Activity in Vivo

Authors:

Abstract and Figures

Effects of circulating insulin and glucose concentrations on skeletal muscle and adipose tissue lipolytic activity were investigated in 10 type 1 diabetes patients with no endogenous insulin secretion. Microdialysis measurements of interstitial glycerol and determination of fractional glycerol release were carried out during standardized combinations of relative hypoinsulinemia/moderate hyperglycemia (11 mmol/liter), hyperinsulinemia/ normoglycemia (5 mmol/liter), and hyperinsulinemia/moderate hyperglycemia, respectively. Local tissue blood flow rates were measured with the (133)Xe clearance technique. In response to the change from hypo- to hyperinsulinemia, the fractional release of glycerol decreased from 159.6 +/- 17.8 to 85.1 +/- 13.7 micromol/liter (P < 0.0001) in adipose tissue, whereas it remained unchanged in skeletal muscle (44.6 +/- 6.4 vs. 36.0 +/- 7.4 micromol/liter; not significant). When hyperinsulinemia was combined with hyperglycemia, fractional glycerol release was further reduced in adipose tissue (64.5 +/- 12.2 micromol/liter; P < 0.05), and in this situation it was also markedly decreased in skeletal muscle (18.1 +/- 4.8 micromol/liter; P < 0.0001). Skeletal muscle blood flow was unaltered over the respective study periods. Adipose tissue blood flow decreased by 50% in response to hyperinsulinemia (P < 0.0005), but no further change was seen when hyperinsulinemia was combined with hyperglycemia. It is concluded that in patients with type 1 diabetes, insulin does not exert an antilipolytic effect in skeletal muscle during normoglycemia. However, in response to combined hyperinsulinemia and hyperglycemia, the lipolytic activity in skeletal muscle is restrained in a similar way as in adipose tissue. This may be explained by a glucose-mediated potentiation of the antilipolytic effectiveness of insulin.
Content may be subject to copyright.
Combined Hyperinsulinemia and Hyperglycemia, But
Not Hyperinsulinemia Alone, Suppress Human Skeletal
Muscle Lipolytic Activity in Vivo
VERONICA QVISTH, EVA HAGSTRO
¨M-TOFT, STAFFAN ENOKSSON, ROBERT S. SHERWIN,
STEFAN SJO
¨BERG, AND JAN BOLINDER
Departments of Medicine (V.Q., E.H.-T., S.S., J.B.) and Vascular Surgery (S.E.), Karolinska University Hospital-Huddinge,
Karolinska Institute, S-141 86 Stockholm, Sweden; and Department of Internal Medicine (R.S.S.), Section of Endocrinology,
Yale University School of Medicine, New Haven, Connecticut 06520
Effects of circulating insulin and glucose concentrations on
skeletal muscle and adipose tissue lipolytic activity were in-
vestigated in 10 type 1 diabetes patients with no endogenous
insulin secretion. Microdialysis measurements of interstitial
glycerol and determination of fractional glycerol release were
carried out during standardized combinations of relative hy-
poinsulinemia/moderate hyperglycemia (11 mmol/liter), hy-
perinsulinemia/ normoglycemia (5 mmol/liter), and hyperin-
sulinemia/moderate hyperglycemia, respectively. Local tissue
blood flow rates were measured with the
133
Xe clearance tech-
nique. In response to the change from hypo- to hyperinsulin-
emia, the fractional release of glycerol decreased from 159.6
17.8 to 85.1 13.7
mol/liter (P<0.0001) in adipose tissue,
whereas it remained unchanged in skeletal muscle (44.6 6.4
vs. 36.0 7.4
mol/liter; not significant). When hyperinsulin-
emia was combined with hyperglycemia, fractional glycerol
release was further reduced in adipose tissue (64.5 12.2
mol/liter; P<0.05), and in this situation it was also markedly
decreased in skeletal muscle (18.1 4.8
mol/liter; P<0.0001).
Skeletal muscle blood flow was unaltered over the respective
study periods. Adipose tissue blood flow decreased by 50% in
response to hyperinsulinemia (P<0.0005), but no further
change was seen when hyperinsulinemia was combined with
hyperglycemia. It is concluded that in patients with type 1
diabetes, insulin does not exert an antilipolytic effect in skel-
etal muscle during normoglycemia. However, in response to
combined hyperinsulinemia and hyperglycemia, the lipolytic
activity in skeletal muscle is restrained in a similar way as in
adipose tissue. This may be explained by a glucose-mediated
potentiation of the antilipolytic effectiveness of insulin.
(J Clin Endocrinol Metab 89: 4693– 4700, 2004)
SKELETAL MUSCLE INSULIN resistance, as reflected by
attenuated insulin-mediated glucose utilization, is a
characteristic feature of type 2 diabetes mellitus, obesity,
dyslipidemia, and hypertension (i.e. the metabolic syn-
drome). Of several possible pathogenetic mechanisms, free
fatty acids (FFA) are considered to play a significant role in
the development of insulin resistance. Thus, there is a re-
ciprocal relationship between lipid and carbohydrate utili-
zation in muscle, and increased levels of FFA in the circu-
lation impair insulin-stimulated skeletal muscle glucose
disposal (1–7).
In humans, the source of FFA is generally assumed to be
hydrolysis (lipolysis) of triglycerides in adipose tissue de-
posits. However, FFA generated from hydrolysis of im trig-
lycerides may also be of importance, because several inves-
tigators have demonstrated a significant lipolytic activity in
human skeletal muscle (8–11). Recent data also suggest that
skeletal muscle lipolysis rates differ between muscle groups
with varying fiber composition (12). Moreover, an inverse
relationship between im triglyceride content and insulin sen-
sitivity has been observed in many studies, as recently re-
viewed (13). This might imply a link between dysregulation
of the local lipolytic activity and development of insulin
resistance in skeletal muscle.
The hormonal regulation of human skeletal muscle lipolysis
is still not completely elucidated. For instance, findings in initial
studies that measured interstitial glycerol levels in vivo sug-
gested a similar antilipolytic effect of insulin in skeletal muscle
and adipose tissue (9, 10, 14). However, in more recent inves-
tigations in which the fractional release of glycerol was assessed
in situ, we (15, 16) and others (11) have not been able to confirm
that insulin exerts an inhibitory effect on skeletal muscle lipo-
lytic activity. To further evaluate this seemingly controversial
issue, we evaluated the effect of insulin as well as the possible
influence of the prevailing glucose level on skeletal muscle and
adipose tissue lipolysis, bearing in mind that circulating glucose
concentrations may affect overall lipolytic activity in vivo (17).
For this purpose, we investigated 10 male patients with insulin-
dependent (type 1) diabetes without endogenous insulin se-
cretion. By infusing insulin and glucose, the patients were
clamped at standardized combinations of relative hypoinsu-
linemia/hyperglycemia, hyperinsulinemia/normoglycemia,
and hyperinsulinemia/hyperglycemia, respectively. Microdi-
alysis measurements of glycerol concentrations in situ and cal-
culation of the fractional release of glycerol (i.e. difference be-
tween interstitial and arterialized venous plasma glycerol) were
carried out simultaneously in the two tissues together with
determinations of local blood flow rates using the
133
Xe clear-
ance technique.
Abbreviations: FFA, Free fatty acid; I-A, interstitial and arterialized;
LPL, lipoprotein lipase.
JCEM is published monthly by The Endocrine Society (http://www.
endo-society.org), the foremost professional society serving the en-
docrine community.
0021-972X/04/$15.00/0 The Journal of Clinical Endocrinology & Metabolism 89(9):4693– 4700
Printed in U.S.A. Copyright © 2004 by The Endocrine Society
doi: 10.1210/jc.2003-030656
4693
Subjects and Methods
Subjects
Ten men (age, 43 3 yr; body mass index, 23.8 0.4 kg/m
2
) with
insulin-dependent type 1 diabetes mellitus were investigated. All sub-
jects were plasma C peptide negative and were receiving continuous
insulin treatment (total insulin dose, 0.8 0.1 U/kg24 h). Apart from
the insulin therapy, they took no other medication. The duration of
diabetes was 19 2 yr, and the metabolic control was fairly good, as
evidenced by their hemoglobin A
1c
levels (6.9 0.4%; normal range,
3.45.0%). Eight of the patients had background diabetic retinopathy;
otherwise, they had no other signs of diabetic micro- or macrovascular
complications.
The ethics committee of the Karolinska Institute approved the study.
The subjects were given a detailed description of the study before they
gave their informed consent.
Microdialysis device
The principles of the microdialysis technique and the microdialysis
device have been described in detail previously (18). Briefly, the mi-
crodialysis catheter (CMA/60, CMA Microdialysis AB, Stockholm, Swe-
den) consists of a semipermeable polyamide membrane (30 0.62 mm;
molecular cut-off, 20,000 Da), which is glued to the end of double-lumen
polyurethane tubing. The probe is connected to a microinfusion pump
(CMA/100 microinjection pump, CMA Microdialyis AB) and is contin-
uously perfused with a sterile perfusion fluid. The perfusate solution
enters the catheter through the outer lumen, an exchange of metabolites
takes place over the microdialysis membrane, and the solution leaves
through the inner lumen from which it is collected. The composition of
the sampled dialysate then mirrors the interstitial fluid.
Study protocol
The subjects were admitted to the hospital on the evening before the
investigation. They were fasting from midnight and received a variable
iv insulin infusion to maintain the blood glucose level at approximately
11 mmol/liter during the night. Venous blood glucose was monitored
at the bedside every hour (HemoCue, Angelholm, Sweden). The subjects
remained in a supine position throughout the study period.
The experiment started at 0730 h. A retrograde catheter was inserted
in a dorsal vein in the hand that was placed in a heated box (63 C) for
sampling of arterialized venous plasma. Arterialization was confirmed
by blood gas analysis (95% O
2
saturation). Retrograde catheters were
also placed in the cubital veins bilaterally for blood sampling and in-
fusion of insulin and glucose, respectively. After superficial skin anes-
thesia (EMLA, Astra, Sodertalje, Sweden), one microdialysis catheter
was inserted percutaneously into the medial portion of the gastrocne-
mius muscle in one leg. A second microdialysis catheter was inserted
into the periumbilical sc adipose tissue about 8 cm lateral to the um-
bilicus. The catheters were continuously perfused with Ringers solution
(Apoteksbolaget, Umea, Sweden) containing 147 mmol/liter sodium, 4
mmol/liter potassium, 2.3 mmol/liter calcium, and 156 mmol/liter chlo-
ride at a flow rate of 0.3
l/min. Previous studies have shown that the
recovery of glycerol is almost complete (95%) in skeletal muscle as well
as in adipose tissue at this flow rate (9).
After a 120-min equilibration period, a 60-min basal sampling period
was started during which the variable insulin infusion was continued to
keep the arterialized venous blood glucose at about 11 mmol/liter.
Thereafter, the insulin infusion was increased to 0.075 U/kgh for 90 min
to make blood glucose fall in a linear fashion from 11 to 5.5 mmol/liter.
This rate of iv insulin administration was thereafter maintained through-
out the rest of the experimental period. A variable iv glucose infusion
was then added for the next 60 min to maintain the blood glucose at 5.5
mmol/liter. Over the following 90 min, the glucose infusion rate was
increased to raise the blood glucose level gradually to 11 mmol/liter. In
some patients, the insulin infusion rate had to be slightly decreased to
raise the blood glucose concentration to 11 mmol/liter within the 90-min
period. The last 60 min of the experiment, blood glucose was maintained
at 11 mmol/liter with a variable glucose infusion. Thus, with this ex-
perimental protocol three separate 60-min steady state periods were
achieved (Fig. 1). The first period (A) reflected a state of relative insulin
deficiency and moderate hyperglycemia. The second period (B) reflected
a combination of hyperinsulinemia and normoglycemia, whereas the
last period (C) showed sustained hyperinsulinemia coupled with mod-
erate hyperglycemia.
Throughout the experiment, dialysate samples were collected in 10-
min fractions for analysis of glycerol. Arterialized venous plasma and
whole blood were drawn every 10 min for analysis of glycerol and
glucose during the three steady state periods. Plasma free insulin was
analyzed every 20 min during the three steady state periods. Blood
glucose was monitored bedside (HemoCue) for adjustment of glucose
and/or insulin infusion rates every 10 min throughout the investigation.
Adipose tissue blood flow and skeletal muscle blood flow rates were
determined with the
133
Xe clearance technique (19) as previously de-
scribed in detail (15). In adipose tissue,
133
Xe (1 MBq in 0.1 ml saline;
Mallinckrodt, Petten, The Netherlands) was injected percutaneously
into the sc tissue in the contralateral abdominal side of the umbilicus 90
min before the first sampling period. After 30 min of equilibration, the
residual activity was continuously monitored externally with a scintil-
lation detector (Mediscint, Oakfield Instruments Ltd., Oxford, UK). The
fractional decay per minute was assessed for consecutive 30-min periods
throughout the study period. In skeletal muscle,
133
Xe (0.3 MBq in 0.1
ml saline) was injected in the medial part of the contralateral gastroc-
nemius muscle after 25 min during each steady state period. After 5-min
FIG. 1. Study protocol.
4694 J Clin Endocrinol Metab, September 2004, 89(9):46934700 Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis
equilibration, the residual activity was monitored for 10 min. In the
skeletal muscle, the
133
Xe decay curve gradually becomes multiexpo-
nential (20). Therefore, the
133
Xe clearance technique cannot be used for
estimating muscle blood flow continuously over extended time periods,
but it can be correctly assessed from the initial part of the
133
Xe washout
curve (20, 21). Adipose tissue and muscle blood flow (TBF) rates were
calculated according to the following formula: TBF k
100 ml/100
gmin, where kdenotes the rate constant of the decay of the residual
activity, and
is the tissue to blood partition coefficient. The values for
were set at 10 ml/g for adipose tissue and 0.7 ml/g for muscle (19, 21).
Glycerol kinetics
In methodological experiments, skeletal muscle and adipose tissue
glycerol uptake and release were simultaneously investigated in four
normal weight, healthy subjects (three women and one man; age, 2528
yr). The investigations were carried out at the General Clinical Research
Center, Yale University School of Medicine, and were approved by the
human investigations committee at Yale. At 0800 h, after an overnight
fast, a 6-h primed continuous infusion of [1,1,2,3,3-
2
H
5
]glycerol (0.02
mg/m
2
body surface areamin) was administered via an antecubital
vein. A retrograde cannula was inserted into a dorsal vein in the con-
tralateral hand that was warmed for sampling of arterialized venous
blood. A small volume of 0.9% NaCl was allowed to flow through the
sampling cannula to maintain potency. Microdialysis catheters were
inserted percutaneously into the medial head of the gastrocnemius mus-
cle and into the periumbilical sc adipose tissue, respectively. The cath-
eters were continuously perfused (0.3
l/min) with an artificial extra-
cellular fluid, as described above. After a minimum of 120 min of
equilibration, dialysate samples were collected in 30-min fractions dur-
ing the 4-h study period for analysis of glycerol and glycerol isotope
enrichment. Arterialized venous blood samples were obtained at 15-min
intervals for measurement of plasma glycerol and glycerol isotope en-
richment. Conversion of interstitial (I) to venous (V) glycerol concen-
trations was made according to the equation: [glycerol]
V
([glyc-
erol)]
I
[glycerol]
a
)(1 e
PS/Q
)[glycerol]
a
where [glycerol]
a
is the
glycerol concentration in arterialized venous plasma, Q represents the
plasma flow rate, and PS denotes the permeability surface product area.
The latter value was approximated to 5 ml/100 gmin in both skeletal
muscle and adipose tissue, as previously suggested (2224) and dis-
cussed in detail (16). The fractional extraction of the glycerol isotope, and
rates of glycerol uptake and release were calculated according to the
formula (25, 26):
Fractional extraction 共关glycerolaEa
共关glycerolVEd/共关glycerolaEa兲共%
Glycerol uptake 共关glycerolaQ
1glycerolVEd/共关glycerolaEa]] nmol/100 gmin
Glycerol release Q([glycerol
V
[glycerol]
a
)glycerol uptake
(nmol/100 gmin), where E
a
and E
d
represent the enrichment of glycerol
in arterialized venous plasma and tissue dialysate, respectively.
Analyses
Dialysate glycerol was measured with an enzymatic fluorometric
method using an automatic tissue dialysate sample analyzer (CMA/600,
CMA Microdialysis). Plasma glycerol was determined with biolumi-
nescence (27). Free insulin in serum was measured with a commercial
RIA kit (Pharmacia Biotech, Uppsala, Sweden). The hospitals routine
clinical chemistry department determined plasma glucose and
hematocrit.
Gas chromatography-mass spectrometry analysis of enrichments of
[1,1,2,3,3-
2
H
5
]glycerol in plasma, infusate, and dialysates was performed
using the triacetate derivative of glycerol (28), as described in detail
previously (29). Gas chromatography-mass spectrometry analysis was
performed with a Hewlett-Packard 5971A Mass Selective Detector (Wil-
mington, DE), operating in the electron ionization mode. Ions with m/z
145, 146, and 148 were monitored for molar percent enrichment in
glycerol.
Statistics
Data are presented as the mean sem. Comparisons of plasma,
muscle, and adipose tissue data were performed by factorial ANOVA.
Variations over time in the same individuals were calculated by one-
factor ANOVA, corrected for repeated measurements. Post hoc analyses
were performed using Scheffe´’s F test. Linear regression analysis was
performed using the method of least squares. P0.05 was considered
statistically significant.
Results
Arterialized venous blood glucose and plasma free insulin
levels are shown in Fig. 2. Plasma free insulin was 87 12
FIG. 2. Arterialized venous plasma concentrations of glucose and insulin during standardized iv administration of glucose and/or insulin in
10 patients with type 1 diabetes without endogenous insulin secretion. Values are the mean SEM.
Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis J Clin Endocrinol Metab, September 2004, 89(9):46934700 4695
pmol/liter during the first sampling period (A). During the
following two periods (B and C), insulin levels were 515
25 and 428 18 pmol/liter, respectively (P0.05). The blood
glucose concentration was initially 11.3 0.2 mmol/liter (A),
decreased to 5.9 0.1 mmol/liter during the second sam-
pling period (B), and finally returned to 11.3 0.2 mmol/
liter during the last steady state sampling period (C).
In the fasting hypoinsulinemic state, glycerol levels were
about 3.5 times higher in adipose tissue and 1.6 times higher
in muscle compared with the arterialized venous plasma
glycerol concentrations (P0.05, by factorial ANOVA; Fig.
3A). Glycerol concentrations in plasma, adipose tissue, and
muscle were all significantly reduced during hyperinsulin-
emia and normoglycemia (P0.0001, by ANOVA). In skel-
etal muscle, glycerol concentrations continued to decrease
when circulating glucose was raised from normoglycemic to
hyperglycemic levels (P0.01, by ANOVA). Conversely, in
adipose tissue and plasma, glycerol concentrations were
maximally decreased during the normoglycemic, hyperin-
sulinemic period, with no further reduction during the hy-
perglycemic, hyperinsulinemic period.
The fractional release of glycerol, the difference between
the interstitial glycerol concentration in adipose and muscle
tissue and that in the arterialized venous plasma, is shown
in Fig. 3B. In skeletal muscle, the interstitial-plasma differ-
ence in glycerol remained unchanged between steady state
periods A and B. It was then significantly reduced by 53%
during the last steady state period when hyperinsulinemia
was combined with hyperglycemia (P0.0001). In adipose
tissue, the decrease in the interstitial-plasma difference in
FIG. 4. Relationship between the I-A glycerol difference in gastroc-
nemius muscle and the arterialized venous plasma glycerol concen-
tration during steady-state periods of hyperglycemia/hyperinsulin-
emia (A), normoglycemia/hyperinsulinemia (B), and hyperglycemia/
hyperinsulinemia (C), respectively. See Fig. 2 for additional details.
FIG. 3. Effect of circulating insulin and glucose levels on glycerol
concentrations in arterialized venous plasma, adipose tissue, and
skeletal muscle (A) and on the difference between the interstitial
tissue glycerol concentration and the arterialized venous plasma glyc-
erol level (I-A difference; B). Statistical differences between study
periods were calculated with ANOVA, repeated measurements, and
post hoc analysis by Scheffe´s F test. *, P0.05 between study periods
A and B; f,P0.05 between study periods B and C. See Fig. 2 for
additional details.
4696 J Clin Endocrinol Metab, September 2004, 89(9):46934700 Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis
glycerol values was most apparent in the hyperinsulinemic,
normoglycemic state (47%; P0.0001, by ANOVA), but a
further reduction (13%; P0.05, by ANOVA) was also
registered in the hyperinsulinemic, hyperglycemic state.
Using individual data, no significant relations were found
between the fractional release of glycerol in skeletal muscle
and the arterialized venous plasma glycerol concentration in
steady state periods AC (Fig. 4).
For comparative purposes, the same type of data presen-
tation was also calculated from a methodological study that
was recently presented by our group (12). Briefly, in the latter
study circulating plasma glycerol was raised about 9-fold
above basal levels in six healthy subjects by a continuous 3-h
infusion of a triacylglycerol emulsion. In the basal postab-
sorptive state, no significant correlation was observed be-
tween the fractional glycerol release in the gastrocnemius
muscle and the arterialized venous plasma glycerol concen-
tration (Fig. 5A). During the triacylglycerol infusion, nega-
tive interstitial and arterialized (I-A) glycerol difference val-
ues were registered in five of the six subjects, indicating
definite uptake of glycerol in the muscle. In these five sub-
jects a strong negative relationship (r ⫽⫺0.84, 0.05) was
evident between the I-A glycerol difference and the plasma
glycerol concentration (Fig. 5B).
Figure 6 depicts adipose tissue and muscle blood flow
rates. The blood flow rate was 4.4 times higher in adipose
tissue compared with muscle tissue during the baseline re-
cordings. In adipose tissue, blood flow decreased by approx-
imately 30% during hyperinsulinemia and normoglycemia
(P0.0005, by ANOVA). No further change was seen when
hyperinsulinemia was coupled with hyperglycemia. The
blood flow rate remained unchanged during the hyperinsu-
linemic, normoglycemic and hyperinsulinemic, hyperglyce-
mic periods in the skeletal muscle.
The results of the methodological experiments, in which
regional glycerol kinetics were investigated in the basal state
in healthy subjects, are summarized in Table 1. After the
120-min equilibration period, the enrichment of glycerol in
arterialized venous plasma (8.0 1.2%) and in skeletal mus-
cle and adipose tissue dialysates (3.4 0.2% and 2.8 0.4%,
respectively) remained stable during the 4-h sampling period
(data not shown). In skeletal muscle, there was significant
fractional extraction of the glycerol tracer (30%), whereas
in adipose tissue, the corresponding value did not differ
significantly from zero. In skeletal muscle, there was an ap-
preciable calculated uptake of glycerol, which amounted to
about 40% of the calculated release of glycerol. In adipose
tissue, the release of glycerol was approximately 3 times
higher than that in skeletal muscle, whereas no uptake of
glycerol was registered.
FIG. 5. Relationship between the I-A glycerol difference in gastrocnemius muscle and the arterialized venous plasma glycerol concentration
in six healthy subjects in the fasting state (A) and in response to triacylglycerol infusion (B). Muscle dialysate glycerol was measured in 15-min
samples over 60 min in the fasting state and during a 3-h infusion of triacylglycerol emulsion (200 mg/ml) given at a rate of 1.85 ml/kgh.
Arterialized venous plasma glycerol was determined every 15 min. Data in B relate to the last 60 min of the triacylglycerol infusion when
steady-state conditions were achieved. Individual data from one subject were excluded from the statistical calculation in B. In this subject, a
positive I-A glycerol difference value was registered, showing net glycerol release (and not uptake) in response to triacylglycerol infusion.
FIG. 6. Effect of circulating insulin and glucose levels on blood flow
rates in adipose tissue and skeletal muscle. See Figs. 2 and 3 for
additional details.
TABLE 1. Glycerol kinetics in skeletal muscle and adipose tissue
Muscle Adipose tissue
Fractional extraction of
[
2
H
5
]glycerol (%)
32 9241
Glycerol uptake (nmol/100 gmin) 33.8 10.9 10.4 67.1
Glycerol release (nmol/100 gmin) 80.7 15.3 255.5 34.7
Regional glycerol kinetics in skeletal muscle and adipose tissue
were investigated in four healthy, normal weight subjects in the basal,
postabsorptive state. [
2
H
5
]Glycerol was infused systemically, and
glycerol isotope enrichment was measured in arterialized venous
plasma and in skeletal muscle and adipose tissue microdialysates. For
further details, see Subjects and Methods. Values are the means
SEM.
Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis J Clin Endocrinol Metab, September 2004, 89(9):46934700 4697
Discussion
In this study we also investigated whether insulin possess
an antilipolytic effect in human skeletal muscle by focusing
on a possible role of the level of glycemia. This was carried
out in patients with type 1 diabetes with no residual endog-
enous insulin secretion, thereby enabling circulating insulin
and glucose levels to be experimentally controlled in a stan-
dardized way by iv administration of insulin and glucose. In
the first part of the experiment, the patients were clamped at
moderate hyperglycemia (11 mmol/liter) by a variable, low
rate infusion of insulin to simulate a state of relative insulin
deficiency. Thereafter and for the remaining period of the
experiment, circulating insulin was increased well above the
concentration of the hormone that possess a maximally ef-
fective antilipolytic effect (30, 31). Plasma glucose was set at
normoglycemia (5 mmol/liter) and moderate hyperglycemia
(11 mmol/liter), respectively.
In response to the change from hypo- to hyperinsulinemia,
the interstitial glycerol concentration decreased significantly
in both adipose tissue and skeletal muscle. Similar findings
have previously been reported and were interpreted as being
indicative of an antilipolytic effect of insulin in muscle (9, 10,
14). However, the interstitial concentration of glycerol de-
pends not only on lipolytic activity in the tissue, but also on
the delivery of glycerol from other sources via the arterial
circulation. Consequently, the fractional release of glycerol
(i.e. the I-A difference between plasma glycerol concentra-
tions) should represent a more accurate measure of the local
lipolytic activity. Thus, as expected, we found that hyper-
insulinemia mediated a marked reduction (50%) in the I-A
glycerol difference in adipose tissue. In skeletal muscle, in
contrast, it remained unchanged in response to insulin. Con-
sistent with recent observations (11, 15, 16), this finding def-
initely argues against an inhibitory effect of insulin on muscle
lipolysis in humans at physiological, normoglycemic condi-
tions. Instead, the decrease in skeletal muscle interstitial glyc-
erol after insulin stimulation is best explained by reduced
arterial inflow of glycerol to the tissue bed as a result of
insulin-induced antilipolysis in adipose tissue.
When hyperinsulinemia was combined with moderate hy-
perglycemia, in contrast, a significant decrease in the frac-
tional release of glycerol was registered in both adipose
tissue and skeletal muscle compared with the corresponding
data obtained during hyperinsulinemia/normoglycemia.
For practical reasons it was not possible to infuse insulin and
glucose in such a way that study periods B and C could be
evaluated in a randomized way. Therefore, we cannot ex-
clude the possibility that the observed decrease in fractional
glycerol release in the two tissue compartments between the
respective study periods was due to the extended time of
hyperinsulinemia, rather than being the result of the change
from normo- to hyperglycemia. This appears less likely,
however, because the glycerol concentrations in arterialized
venous plasma, adipose tissue, and skeletal muscle remained
stable during each of the experimental periods. Hence, our
findings may indicate a different regulation of lipolytic ac-
tivity in adipose tissue as well as in skeletal muscle during
combined hyperinsulinemia and hyperglycemia. The molec-
ular mechanism underlying this phenomenon is not known.
However, many years ago we observed that the antilipolytic
effect of insulin in isolated adipocytes was potentiated by
glucose regardless of whether the cells had been exposed to
high insulin and glucose concentrations in vitro (32) or in vivo
(33); this was associated with increased insulin receptor bind-
ing affinity as well as stimulation of postbinding events. It
has also been shown that the effectiveness of the inhibitory
action of insulin on adipose tissue lipolysis is dependent on
the prevailing lipolytic activity, so the antilipolytic effect of
the hormone is more pronounced when the rate of lipolysis
is augmented (34, 35), probably due to increased insulin
receptor autophosphorylation and signal transduction
through an insulin receptor substrate-1- and phosphatidyl-
inositol 3-kinase-dependent pathway (35). Furthermore, at
least in adipocytes, glucose enhances the lipolysis rate (36,
37), which may be attributed to increased expression of the
rate-limiting enzyme for lipolysis, hormone-sensitive lipase
(38, 39). This enzyme is also present in skeletal muscle (40).
Thus, it is possible that our observed decrease in lipolytic
activity in adipose tissue as well as in skeletal muscle in
response to combined hyperinsulinemia and hyperglycemia
was the result of a more pronounced antilipolytic effective-
ness of insulin because of an augmented, glucose-mediated,
elevation of lipolytic activity in both tissues. In theory, the
decrease in fractional glycerol release in the two tissues could
also have been due to increased local glycerol reutilization
during combined insulin and glucose stimulation. To date
there has been no method available that has allowed unam-
biguous differentiation between in vivo glycerol release and
uptake in a single skeletal muscle in humans. Although var-
ious tracer techniques have been used for this purpose, these
methods, even when combined with elaborate catherization
techniques, can only determine the net substrate release or
uptake in the various muscles (and other tissues) drained by
the venous effluent. Bearing in mind that lipolytic activity
may vary even between different skeletal muscle groups in
humans (12), no definite conclusions about glycerol kinetics
in a separate muscle tissue compartment can be made. In the
present study, however, we introduced a novel approach to
simultaneously investigate local tissue glycerol uptake and
release using a combination of systemic administration of a
stable glycerol tracer and measurement of enrichment of
glycerol in tissue interstitial fluid sampled by microdialysis.
By this method we registered significant fractional extraction
of the glycerol tracer and definite calculated glycerol uptake
in skeletal muscle. Hence, these findings are the first to show
with certainty that uptake of glycerol takes place in a defined
skeletal muscle group in humans, as previously shown in
rodents (41). In adipose tissue, in contrast, we found no
appreciable uptake of glycerol. This may be expected, be-
cause the enzyme responsible for glycerol utilization, glyc-
erokinase, is present in skeletal muscle, but not in adipose
tissue in humans (42). Moreover, in skeletal muscle, the cal-
culated uptake of glycerol was approximately 40% the cor-
responding calculated release of glycerol. This is similar to
that reported by Jensen (26), who estimated regional glycerol
uptake and release across the thigh. Although caution should
be exercised in extrapolating data, our findings relating to the
calculated regional rates of glycerol uptake and release argue
against the idea that the decrease in the interstitial-plasma
4698 J Clin Endocrinol Metab, September 2004, 89(9):46934700 Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis
difference in glycerol observed in both skeletal muscle and
adipose tissue during combined hyperinsulinemia and hy-
perglycemia was due to an increase in glycerol uptake rather
than suppression of the lipolytic activity. Hence, assuming a
comparable relation between skeletal muscle glycerol uptake
and release in the patients with type 1 diabetes and healthy
subjects, the uptake of glycerol should have been increased
at least 2- to 3-fold to account for the more than 50% decrease
in fractional glycerol release found in skeletal muscle in the
former study group. This appears unlikely because, if any-
thing, the uptake of glycerol in muscle tissues is supposed to
be reduced in response to meal ingestion and hyperinsulin-
emia (26). Furthermore, as mentioned above, no uptake of
glycerol takes place in adipose tissue. Moreover, using data
from a previously published methodological investigation
(12) in which circulating plasma glycerol levels were elevated
well above the basal interstitial glycerol concentration in
skeletal muscle by a continuous triacylglycerol infusion, we
found a strong negative correlation between the I-A glycerol
difference in the gastrocnemius muscle and the arterialized
venous plasma glycerol concentration (i.e. when uptake of
glycerol constituted the main direction of glycerol movement
in muscle tissue). By contrast, no apparent relationship be-
tween these variables was observed during any of the three
experimental study periods in the present investigation.
Thus, taken together, our findings are best explained by a
suppression of lipolytic activity in both skeletal muscle and
adipose tissue in response to combined hyperinsulinemia
and hyperglycemia.
Differences in lipoprotein lipase (LPL)-mediated intravas-
cular triglyceride hydrolysis in the two tissues may also be
of importance. It is not possible, using microdialysis, to dif-
ferentiate between intracellular (mediated by hormone-
sensitive lipase) and extracellular, LPL-induced lipolysis. In
response to insulin, however, the activity of LPL is stimu-
lated in adipose tissue, whereas the opposite occurs in skel-
etal muscle (43). Therefore, because the reduction in frac-
tional release of glycerol was evident in both tissues during
combined hyperinsulinemia and hyperglycemia, it seems
unlikely that differences in insulin-mediated LPL activity
influenced our findings in a major way.
Local tissue blood flow is another determinant of the in-
terstitial glycerol level (44), via delivery and removal of the
metabolite by the microcirculation. In this study, we used the
133
Xe washout technique for simultaneous determinations of
absolute blood flow rates in adipose tissue and skeletal mus-
cle. This method has been widely applied for recordings of
nutritive blood flow in both tissues (19, 21). Accordingly, in
a previous methodological study (16), we found the repro-
ducibility and precision of this technique to be satisfactory.
In the present study, no effect of insulin on the blood flow
rate was registered in skeletal muscle, whereas in adipose
tissue it decreased significantly, by about 30%, in response to
insulin. These findings were unexpected, because previous
data in healthy, nonobese subjects have shown an increase in
muscle blood flow, but no change in adipose tissue blood
flow after insulin stimulation (16, 45). However, defective
insulin-mediated blood flow regulation has been demon-
strated in various insulin-resistant conditions (4547).
Hence, our present data may be explained by disturbed
vascular effects of insulin related to the diabetic state. Dif-
ferences in study design are probably also of importance. For
example, during the basal part of the present experiment
(hyperglycemia/relative hypoinsulinemia), the recorded
blood flow rate in adipose tissue (6 ml/100 gmin) was
approximately 3 times higher than that previously observed
by our group in the postabsorptive state in healthy subjects
(15, 16). Therefore, comparison of present and previous
blood flow data should be performed with caution. Never-
theless, the fact that muscle blood flow rates remained un-
changed in response to insulin imply that the observed
changes in the fractional release of glycerol in skeletal muscle
should not have been influenced by microcirculatory events.
It should be noted that the lipolytic activity might vary
between different muscle groups, probably depending on the
fiber composition (12). In the present study we preferred to
examine gastrocnemius muscle, which possesses a high li-
polysis rate (12). In consequence, we cannot exclude the
possibility that the lipolytic activity in other muscle groups
may be regulated differently in response to insulin and glu-
cose stimulation. Moreover, from the data in this study we
cannot define with certainty the level of glycemia that po-
tentiates the antilipolytic effect of insulin in skeletal muscle
or whether this phenomenon is confined to diabetic patients.
However, in a previous study, no suppressive effect of in-
sulin on skeletal muscle lipolysis rates was registered in
either nonobese or obese subjects with normal glucose tol-
erance in response to elevated circulating levels of insulin
and glucose (maximum arterialized venous plasma glucose
concentration, 10 mmol/liter) after an oral glucose load
(15). Thus, considering previous and present findings to-
gether, it would appear that elevation of circulating glucose
levels at least above the normal postprandial range is needed
to influence the antilipolytic effect of insulin in skeletal
muscle.
In summary, the results of this experimental study in pa-
tients with type 1 diabetes mellitus add further evidence for
the supposition that insulin does not exert an antilipolytic
effect in skeletal muscle in vivo in humans during normal
glycemic conditions. This is in contrast to the insulin effect
on the regulation of adipose tissue lipolysis. In response to
combined hyperinsulinemia and moderate hyperglycemia,
in contrast, the lipolytic activity in skeletal muscle is reduced
in a similar way as in adipose tissue, which may be explained
by a glucose-mediated potentiation of the antilipolytic ef-
fectiveness of the hormone. This mechanism may be impor-
tant for increased triglyceride deposition in skeletal muscle
(and adipose tissue), which is well documented in various
insulin-resistant conditions with subtle abnormalities in glu-
cose control.
Acknowledgments
We acknowledge the excellent technical assistance of B.-M. Leijon-
huvfud, K. Hertel, E. Sjo¨lin, and K. Wåhle´n. We also thank Gary W. Cline
and Frances Rife (Yale General Clinical Research Center) for assistance
with the stable isotope preparation and analyses.
Received April 15, 2003. Accepted June 4, 2004.
Address all correspondence and requests for reprints to: Dr. Jan
Bolinder, Department of Medicine, M 63, Karolinska University Hos-
pital-Huddinge, S-141 86 Stockholm, Sweden. E-mail: jan.bolinder@hs.se.
Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis J Clin Endocrinol Metab, September 2004, 89(9):46934700 4699
This work was supported by grants from the Swedish Research Coun-
cil, the Swedish Diabetes Association, Karolinska Institute, NIH (RR-
00125), and the Yale Diabetes Endocrinology Research Center
(DK-45735).
References
1. Randle PJ, Garland PB, Hales CN, Newsholme EA 1963 The glucose-fatty acid
cycle. Its role in insulin sensitivity and the metabolic disturbances of diabetes
mellitus. Lancet 1:785789
2. Yki-Ja¨ rvinen H, Puhakainen I, Koivisto VA 1991 Effect of free fatty acids on
glucose uptake and nonoxidative glycolysis across human forearm tissues in
the basal state and during insulin stimulation. J Clin Endocrinol Metab 72:
12681277
3. Kelley DE, Mokan M, Simoneau J-A, Mandarino LJ 1993 Interaction between
glucose and free fatty acid metabolism in human skeletal muscle. J Clin Invest
92:9198
4. Boden G, Chen X, Ruiz J, White JV, Rossetti L 1994 Mechanisms of fatty-
acid-induced inhibition of glucose uptake. J Clin Invest 93:24382446
5. Boden G, Chen X 1995 Effects of fat on glucose uptake and utilization in
patients with non-insulin-dependent diabetes. J Clin Invest 96:12611268
6. Roden M, Price TB, Perseghin G, Petersen KF, Rothman DL, Cline GW,
Shulman GI 1996 Mechanism of free fatty acid-induced insulin resistance in
humans. J Clin Invest 97:28592865
7. Roden M, Krssak M, Stingl H, Gruber S, Hofer A, Furnsinn C, Moser E,
Waldhausl W 1999 Rapid impairment of skeletal muscle glucose transport/
phosphorylation by free fatty acids in humans. Diabetes 48:358364
8. Maggs DG, Jacob R, Rife F, Lange R, Leone P, During MJ, Tamborlane WV,
Sherwin RS 1995 Interstitial fluid concentrations of glycerol, glucose and
amino acids in human quadriceps muscle and adipose tissue. J Clin Invest
96:370377
9. Hagstro¨ m-Toft E, Enoksson S, Moberg E, Bolinder J, Arner P 1998 Absolute
concentrations of glycerol and lactate in human skeletal muscle, adipose tissue
and blood. Am J Physiol 273:E585E592
10. Jacob S, Hauer B, Becker R, Artzner S, Grauer P, Loblein K, Nielsen M, Renn
W, Rett K, Wahl HG, Stumvoll M, Haring HU 1999 Lipolysis in skeletal
muscle is rapidly regulated by low physiological doses of insulin. Diabetologia
42:11711174
11. Sjo¨ strand M, Gudbjo¨ rnsdottir S, Holma¨ng A, Strindberg L, Ekberg K,
Lo¨ nnroth P 2002 Measurements of interstitial muscle glycerol in normal and
insulin-resistant subjects. J Clin Endocrinol Metab 87:22062211
12. Hagstro¨ m-Toft E, Qvisth V, Nennesmo I, Ryden M, Bolinder H, Enoksson
S, Bolinder J, Arner P 2002 Marked heterogeneity of human skeletal muscle
lipolysis at rest. Diabetes 51:33763383
13. Kelley DE, Goodpasture BH 2001 Skeletal muscle triglyceride. An aspect of
regional adiposity and insulin resistance. Diabetes Care 24:933941
14. Stumvoll M, Jacob S, Wahl HG, Hauer B, Loblein K, Grauer P, Becker R,
Nielsen M, Renn W, Haring H 2000 Suppression of systemic, intramuscular,
and subcutaneous adipose tissue lipolysis by insulin in humans. J Clin En-
docrinol Metab 85:37403745
15. Bolinder J, Kerckhoffs DAJM, Moberg E, Hagstro¨ m-Toft E, Arner P 2000
Rates of skeletal muscle and adipose tissue glycerol release in nonobese and
obese subjects. Diabetes 49:797802
16. Moberg E, Sjo¨ berg S, Hagstro¨ m-Toft E, Bolinder J 2002 No apparent sup-
pression by insulin of in vivo skeletal muscle lipolysis in nonobese women.
Am J Physiol 283:E295E301
17. Wolfe RR, Peters EJ 1987 Lipolytic response to glucose infusion in human
subjects. Am J Physiol 252:E218E223
18. Tossman U, Ungerstedt U 1986 Microdialysis in the study of extracellular
levels of amino acids in the rat brain. Acta Physiol Scand 128:914
19. Larsen OA, Lassen NA, Quaade F 1966 Blood flow through human adipose
tissue determined with radioactive xenon. Acta Physiol Scand 66:337345
20. Sejrsen P, To¨ nnesen KH 1968 Inert gas diffusion method for measurement of
blood flow using saturation techniques: comparison with directly measured
blood flow in isolated gastrocnemius muscle of the cat. Clin Res 22:679693
21. Lassen NA, Lindbjerg IF, Munck O 1964 Measurement of blood flow through
skeletal muscle by intramuscular injection of
133
xenon. Lancet 1:686689
22. Paaske WP 1977 Capillary permeability in skeletal muscle. Acta Physiol Scand
101:114
23. Paaske WP, Nielsen L 1976 Capillary permeability in adipose tissue. Acta
Physiol Scand 98:116122
24. Crone C, Levitt DG 1984 Capillary permeability to small solutes. In: Handbook
of physiology: the cardiovascular system. Bethesda, MD: Am Physiol Soc, sect
2, vol IV, pt 1, chapt 10, p 411466
25. Landau BR 1999 Glycerol production and utilization measured using stable
isotopes. Proc Nutr Soc 58:973978
26. Jensen MD 1999 Regional glycerol and free fatty acid metabolism before and
after meal ingestion. Am J Physiol 276:E863E869
27. Hellme´ r J, Arner P, Lundin A 1989 Automatic luminometric kinetic assay of
glycerol for lipolysis studies. Anal Biochem 177:132137
28. Wolfe RR 1992 Radioactive and stable isotope tracers in biomedicine: prin-
ciples and practice of kinetic analysis. New York: Wiley
29. Robinson C, Tamborlane WW, Maggs DG, Enoksson S, Sherwin RS, Silver
D, Shulman GI, Caprio S 1998 Effect of insulin on glycerol production in obese
adolescents. Am J Physiol 274:E737E743
30. Rabinowitz D, Zierler KL 1962 Forearm metabolism in obesity and its re-
sponse to intra-arterial insulin. Characterization of insulin resistance and ev-
idence for adaptive hyperinsulinism. J Clin Invest 41:21732181
31. Nurjhan N, Campbell PJ, Kennedy FP, Miles JM, Gerich JE 1986 Insulin
dose-response characteristics for suppression of glycerol release and conver-
sion to glucose in humans. Diabetes 35:13261331
32. Arner P, Bolinder J, O
¨stman J 1983 Glucose stimulation of the antilipolytic
effect of insulin in man. Science 220:10571058
33. Arner P, Bolinder J, O
¨stman J 1983 Marked increase in insulin sensitivity of
human fat cells one hour after glucose ingestion. J Clin Invest 73:709714
34. Thomas SHL, Wisher MH, Brandenburg D, So¨ nksen PH 1979 Evidence that
the anti-lipolytic and lipogenic effects of insulin are mediated by the same
receptor. Biochem J 184:355360
35. Zierath JR, Livingston JN, Tho¨ rne A, Bolinder J, Reynisdottir S, Lo¨ nnqvist
F, Arner P 1998 Regional differences in insulin inhibition of non-esterified fatty
acid release from human adipocytes: relation to insulin receptor phosphory-
lation and intracellular signalling through the insulin receptor substrate-1
pathway. Diabetologia 41:13431354
36. Moussalli C, Downs RW, May JM 1986 Potentiation by glucose of lipolytic
responsiveness of human adipocytes. Diabetes 35:759763
37. Szkudelski T, Szkudelska K 2000 Glucose as a lipolytic agent: studies on
isolated rat adipocytes. Physiol Res 49:213217
38. Raclot T, Dauzats M, Langin D 1998 Regulation of hormone-sensitive lipase
expression by glucose in 3T3F442A adipocytes. Biochem Biophys Res Com-
mun 245:510513
39. Botion LM, Green A 1999 Long-term regulation of lipolysis and hormone-
sensitive lipase by insulin and glucose. Diabetes 48:16911697
40. Oscai LB, Essig DA, Palmer WK 1990 Lipase regulation of muscle triglyceride
hydrolysis. J Appl Physiol 69:15711577
41. Guo Z, Jensen MD 1999 Blood glycerol is an important precursor for intra-
muscular triacylglycerol synthesis. J Biol Chem 274:2370223706
42. Watford M 2000 Functional glycerol kinase activity and the possibility of a
major role for glycerogenesis in mammalian skeletal muscle. Nutr Rev 58:
145148
43. Zechner R 1997 The tissue-specific expression of lipoprotein lipase: implica-
tions for energy and lipoprotein metabolism. Curr Opin Lipidol 8:7788
44. Enoksson S, Nordenstro¨ m J, Bolinder J, Arner P 1995 Influence of local blood
flow on glycerol levels in human adipose tissue. Int J Obes 19:350354
45. Baron AD 1993 Cardiovascular actions of insulin in humans. Implications for
insulin sensitivity and vascular tone. Baillieres Clin Endocrinol Metab 7:962
987
46. Laakso M, Edelman SV, Brechtel G, Baron AD 1990 Decreased effect of
insulin to stimulate skeletal muscle blood flow in obese man: a novel mech-
anism for insulin resistance. J Clin Invest 85:18441852
47. Karpe F, Fielding BA, Ilic V, Mcdonald IA, Summers LKM, Frayn KN 2002
Impaired postprandial adipose tissue blood flow response is related to aspects
of insulin sensitivity. Diabetes 51:24672473
JCEM is published monthly by The Endocrine Society (http://www.endo-society.org), the foremost professional society serving the
endocrine community.
4700 J Clin Endocrinol Metab, September 2004, 89(9):46934700 Qvisth et al. Effect of Insulin and Glucose on Skeletal Muscle Lipolysis
... It is also applicable for m D after conversion of the (absolute) interstitial concentration to the venous concentration, according to the following formula: venous concentration = interstitial concentration -arterial concentration × (1 -e -PS/Q ) + arterial concentration, where PS denotes the permeability surface product area and Q represents the plasma fl ow rate. 56 A limitation with the Fick ' s principle is that while it quantifi es the net substrate fl ux across the tissue bed, it cannot separate simultaneous uptake and release of the metabolite in the tissue. This distinction has been accomplished recently with the use of stable, nonradioactive isotopes of the metabolites of interest. ...
... Together with determination of local blood fl ow, the rate of tissue uptake and/or release can then be calculated. 56 By applying a protocol with in situ perfusion of various target cell receptor agonists and antagonists, the regulation of cellular metabolism may be investigated at the molecular level. The in situ perfusion technique can also be performed in conjunction with systemic administration of hormones or regulating compounds, or under experimental conditions where the release of endogenous hormones is stimulated. ...
Article
This chapter discusses the role of microdialysis in the discovery and development of drugs that act on the central nervous system (CNS). The focus is on strategies that will help to select high-quality compounds for development and that contribute to reducing the current high attrition rate in the pharmaceutical industry. Pharmacodynamic and pharmacokinetic microdialysis applications are discussed to illustrate how and when microdialysis can be used in the different stages of the discovery and development process. Since microdialysis data are likely to become an important part of new drug submissions, and thus may potentially contribute to the FDA Critical Path Initiative to facilitate innovation in drug development, some regulatory aspects will be discussed as well.
... Furthermore, in humans at least, the inhibitory action of insulin on AT lipolysis is dependent on the prevailing lipolytic activity, such that the antilipolytic effect of the hormone is more pronounced when the rate of lipolysis is augmented, probably due to increased insulin receptor and signal transduction activity (Zierath et al., 1998). Finally, the shutdown of lipolysis during the IVGTT challenge of ruminants may be partly directly regulated by glucose, which is the metabolic driver of the degree of NEFA suppression in human AT cells (Arner et al., 1983;Qvisth et al., 2004). ...
Article
High energy intake in the dry period has reportedly had adverse effects on mobilization of body reserves, DMI, and productivity of dairy cows. We investigated whether grass silage (GS) fed ad libitum (HEI; 141% of daily metabolizable energy requirements) in an 8-wk dry period affects metabolic adaptation—specifically, peripheral insulin resistance—compared with a total mixed ration consisting of GS, wheat straw, and rapeseed meal (55/40/5%; CEI; 108% of metabolizable energy/d) fed ad libitum. Multiparous Ayrshire dairy cows (n = 16) were used in a randomized complete block design until 8 wk after parturition. Commercial concentrates were fed 1 and 2 kg/d during the last 10 to 6 and 5 to 0 d before the expected calving date, respectively. Postpartum, a similar lactation diet with ad libitum access to GS and increasing concentrate allowance (maximum of 16 kg/d) was offered to all. The HEI group gained more body weight and had higher plasma insulin, glucose, and β-hydroxybutyrate concentrations than the CEI group prepartum. Postpartal plasma glucose tended to be higher and milk yield was greater from wk 5 onward for HEI compared with CEI cows. Intravenous glucose tolerance test (IVGTT) was performed at −13 ± 5 d and 9 ± 1 d relative to calving. The HEI cows had greater insulin response to glucose load and smaller area under the response curve for glucose than CEI cows in prepartal IVGTT. Thus, compensatory insulin secretion adapted to changes in insulin sensitivity of the peripheral tissues, preserving glucose tolerance of HEI cows. Higher insulin levels were needed in HEI cows than in CEI cows to elicit a similar decrement of nonesterified fatty acid concentration in prepartal IVGTT, suggesting reduced inhibition of lipolysis by insulin in HEI cows before parturition. In conclusion, high energy intake of moderately digestible GS with low concentrate feeding in the close-up dry period did not have adverse effects on metabolic adaptation, insulin sensitivity, and body mobilization after parturition. Instead, this feeding regimen was more beneficial to early-lactation performance than GS-based total mixed ration diluted with wheat straw.
... Therefore we found FFA serum concentration to be still in physiological normal range in our study population . (7,8) Our study showed that there was significant correlation between FFA and FABP4. The recent expansion of mechanistic insights into FABP action demonstrated that these lipid chaperones serve as upstream and critical modulators of many lipidsignaling cascades. ...
Article
Full-text available
BACKGROUND: The prevalence of obesity has increased dramatically in recent years. It is commonly associated with type 2 diabetes, coronary artery disease, and hypertension. White adipose tissue (WAT) is a major site of energy storage and is important for energy homeostasis. WAT has been increasingly recognized as an important endocrine organ that secretes a number of biologically active “adipokines”. The resultant higher FFA, FABP4, FABP5 concentration; and lower concentration of adiponectin is known to be correlated with inflammation. The aim of this study was to observe the correlation between FFA, FABP4, FABP5 and adiponectin with TNF-α and Interleukin-6 as markers of inflammation. METHODS: The study was observational with a cross sectional design. The analysis was done on 69 male subjects aged 30-60 years with non diabetic abdominal obesity which is characterized by waist circumference (WC) 98.7±6.5 cm and fasting blood glucose 87.1±9.7 mg/dL. FFA testing was performed by enzymatic colorimetric assay; whereas FABP4, FABP5, TNF-α, adiponectin and IL-6 were performed by ELISA. All statistical calculations were performed with the SPSS 11.5 statistical software package. We used the Pearson or Spearman’s rho correlation coefficient to assess the correlation between various anthropometric and biochemical measures. We also used path analysis Lisrel 8.30 for Windows. RESULTS: This study revealed that there was no correlation between FFA, FABP4 and adiponectin with TNF-α and Interleukin-6, whereas there was correlation between FABP5 with TNF-α and Interleukin-6. This study also showed there were correlations between WC and hsCRP (r=0.314, p=0.000), WC and IL-6 (r=0.276, p=0.022), FFA and FABP4 (r=0.263, p=0.029), FABP4 and WC (r=0.249, p=0.039), FABP4 and BMI (r=0.311, p=0.009), FABP5 and TNF-α (r=0.408, p=0.000), FABP5 and FABP4 (r=0.296, p=0.014), FABP5 and Interleukin-6 (r=0.248, p=0.04), Adiponectin and HDL-Cholesterol (r=0.301, p=0.012). CONCLUSIONS: Abdominal obesity might contribute to inflammation in obese nondiabetic males. This study indicated that in abdominal obesity, FFA may induce inflammation through FABP4 and FABP5. Advancing our understanding of the function and measurement of FABP4 and FABP5 serum concentration will give insight into the clinical diagnosis of obesity-related metabolic disorders. KEYWORDS: obesity, waist circumference, free fatty acid (FFA), fatty acid binding protein (FABP), adiponectin, TNF-α, interleukin-6, inflammation.
... The molecular mechanism for this difference is unknown. In skeletal muscle, only combined hyperinsulinemia and hyperglycemia suppress lipolytic activity (35). Lipoprotein lipase activity, which hydrolyzes lipoproteins to FFAs and glycerol mainly in the epithelium of capillary beds, is reduced by insulin in skeletal muscle (36). ...
Article
Full-text available
Atrial natriuretic peptide (ANP) regulates arterial blood pressure. In addition, ANP has recently been shown to promote human adipose tissue lipolysis through cGMP-mediated hormone-sensitive lipase activation. We hypothesized that ANP increases postprandial free fatty acid (FFA) availability and energy expenditure while decreasing arterial blood pressure. We infused human ANP (25 ng . kg(-1) . min(-1)) in 12 men (age 32 +/- 0.8 years, BMI 23.3 +/- 0.4 kg/m(2)) before, during, and 2 h after ingestion of a standardized high-fat test meal in a randomized, double-blind, cross-over fashion. Cardiovascular changes were monitored by continuous electrocardiogram and beat-by-beat blood pressure recordings. Metabolism was monitored through venous blood sampling, intramuscular and subcutaneous abdominal adipose tissue microdialysis, and indirect calorimetry. ANP infusion decreased mean arterial blood pressure by 4 mmHg during the postprandial phase (P < 0.01 vs. placebo). At the same time, ANP induced lipolysis systemically (P < 0.05 vs. placebo) and locally in subcutaneous abdominal adipose tissue (P < 0.0001 vs. placebo), leading to a 50% increase in venous glycerol (P < 0.01) and FFA (P < 0.05) concentrations compared with placebo. The increase in FFA availability with ANP was paralleled by a 15% increase in lipid oxidation rates (P < 0.05 vs. placebo), driving a substantial increase in postprandial energy expenditure (P < 0.05 vs. placebo). Our data identify the ANP system as a novel pathway regulating postprandial lipid oxidation, energy expenditure, and concomitantly arterial blood pressure. The findings could have therapeutic implications.
Chapter
Clinical metabolic research needs tools to get a footprint of human physiology and pathophysiology in vivo. The novel idea of monitoring interstitial fluid by microdialysis in the brain of laboratory animals was a complement to ordinary blood sampling. Semipermeable hollow fibers were implemented in the brain-tissue mimicking characteristics of artificial blood vessels communicating with freely diffusing molecules in situ. Subcutaneous microdialysis for measurements of interstitial concentrations of any low molecular compound was introduced in 1987. Each microdialysis catheter required calibration in situ by the “no net flux” method to get the “true” interstitial glucose, lactate, adenosine, and glycerol concentrations. Addition of 133Xe-clearance assessments of adipose tissue blood flow (ATBF) allowed adoption of Fick’s principle to get an estimate of substrate release. By the time, technical improvements of probes and recovery techniques have made it possible to monitor interstitial insulin and cytokines/chemokines in both sc adipose tissue and muscle. In fact, almost every organ in the human body may be accessible for monitoring of small and large molecules in the interstitial fluid. Thus, the future for the microdialysis technique in metabolic research looks bright.
Article
Insulin administration during insulin-modified intravenous glucose tolerance test (IM-IVGTT) can induce transient hypoglycemia in healthy insulin-sensitive subjects. This triggers counterregulatory responses (CRR), which influence the kinetics of glucose and non-esterified fatty acids (NEFA) and undermines the accuracy of mathematical modeling methods that do not explicitly account for CRR. The aim of this study is to evaluate mathematical models of glucose and NEFA kinetics against experimental data in presence or absence of CRR. Thirteen healthy non-diabetic subjects underwent a standard IM-IVGTT and a modified test (GC-IM-IVGTT) with a variable glucose infusion preventing hypoglycemia. While model predictions fit very well glucose and NEFA data from GC-IM-IVGTT, they lagged behind observations from IM-IVGTT during recovery from hypoglycemia, independently of insulinemia, which did not differ significantly between protocols. A modification to the glucose minimal model, using glucose concentration below a threshold as signal for CRR, improves model predictions for both glucose and NEFA. The associated increase in endogenous glucose production correlates, among various CRR hormones, mainly with the dynamics of glucagon concentration. The modified minimal models introduce new parameters that quantify strength and duration of CRR following hypoglycemia. Although CRR represents an unwanted side effect in IM-IVGTT occurring only in insulin-sensitive subjects, this study provides new insights leading to improved procedures for estimating insulin sensitivity from IM-IVGTT, which may also allow for assessing the individual capacity of recovery from hypoglycemic events in patients treated with insulin or insulin-releasing drugs.
Chapter
Full-text available
Modeling has contributed to the advancement in animal health and animal production with the aid of two types of models; research models and diet formulation models. The former models admit the inclusion of detailed mathematical accounts of digestive and metabolic processes, and the latter deploy extensive mathematical specifications of the nutritional value of feeds in regard to the animal host. Each has been used to help maintain and enhance health, welfare and production of farm animals. Data for models comes from various sources but, until recent times, the detailed investigation of animal responses to metabolic challenges has been limited to empirical quantification. However, we point out that there remain very useful indices and measures afforded from modeling such responses if special care is taken in mathematically teasing out the metabolic information. In this article we describe how this is done, and we demonstrate the methods with reference to models of the glucose and the NEFA responses of the glucose challenge. Keywordsglucose–non-esterified fatty acids (NEFA)–metabolic model–minimal model–model partitioning
Article
Full-text available
1. The dose-response relationships of insulin stimulation of lipogenesis and inhibition of lipolysis were studied simultaneously by using rat adipocytes to determine whether these different effects of insulin are mediated through the same or different sets of receptors. 2. The sensitivity (defined as the concentration of insulin required to produce a half-maximal effect) of the stimulated lipogenic response to insulin was not significantly different from the sensitivity of the anti-lipolytic response to insulin. The addition of different adrenaline and glucose concentrations did not alter the half-maximal concentration of insulin required to inhibit lipolysis. 3. The specificities of the lipogenic and antilipolytic responses were studied by using insulin analogues. The sensitivities of the lipogenic and anti-lipolytic responses were the same for five chemically modified insulins and hagfish insulin, which have potencies compared with bovine insulin of between 3 and 90%. 4. Starving rats for 48h significantly increased the sensitivities of both the antilipolytic and lipogenic responses to insulin, but the changes in the sensitivities of both lipogenesis and anti-lipolysis returned to that of fed rats. 5. We conclude that insulin stimulates lipogenesis and inhibits lipolysis over the same concentration range. These observations provide powerful evidence that the different effects of insulin are mediated through the same set of receptors.
Article
Full-text available
Obesity is characterized by decreased rates of skeletal muscle insulin-mediated glucose uptake (IMGU). Since IMGU equals the product of the arteriovenous glucose difference (AVGd) across muscle and blood flow into muscle, reduced blood flow and/or tissue activity (AVGd) can lead to decreased IMGU. To examine this issue, we studied six lean (weight 68 +/- 3 kg, mean +/- SEM) and six obese (94 +/- 3 kg) men. The insulin dose-response curves for whole body and leg IMGU were constructed using the euglycemic clamp and leg balance techniques over a large range of serum insulin concentrations. In lean and obese subjects, whole body IMGU, AVGd, blood flow, and leg IMGU increased in a dose dependent fashion and maximal rates of all parameters were reduced in obese subjects compared to lean subjects. The dose-response curves for whole body IMGU, leg IMGU, and AVGd were right-shifted in obese subjects with an ED50 two- to threefold higher than that of lean subjects for each parameter. Leg blood flow increased approximately twofold from basal 2.7 +/- 0.2 to 4.4 +/- 0.2 dl/min in lean, P less than 0.01, and from 2.5 +/- 0.3 to 4.4 +/- 0.4 dl/min in obese subjects, P less than 0.01. The ED50 for insulin's effect to increase leg blood flow was about fourfold higher for obese (957 pmol/liter) than lean subjects (266 pmol/liter), P less than 0.01. Therefore, decreased insulin sensitivity in human obesity is not only due to lower glucose extraction in insulin-sensitive tissues but also to lower blood flow to these tissues. Thus, in vivo insulin resistance can be due to a defect in insulin action at the tissue level and/or a defect in insulin's hemodynamic action to increase blood flow to insulin sensitive tissues.
Article
We measured splanchnic and leg glycerol [and free fatty acid (FFA)] uptake and release in 11 healthy volunteers before and after meal ingestion to assess whether regional FFA-to-glycerol release ratios mirror systemic release ratios. Basal splanchnic triglyceride release was also assessed. Although basal splanchnic glycerol uptake (111 +/- 18 mu mol/min) accounted for most of systemic glycerol rate of appearance (156 +/- 20 mu mol/min), leg glycerol uptake was also noted. The basal, systemic FFA-to-glycerol release ratio was less (2.6 +/- 0.2, P < 0.05) than the splanchnic ratio of 6.1 +/- 1.3, and the leg FFA-to-glycerol release ratio under fed conditions was less than the systemic ratio (0.9 +/- 0.1 vs. 1.6 +/- 0.2, respectively, P < 0.05). Basal splanchnic triglyceride production rates were 74 +/- 20 mu mol/min, which could produce equimolar amounts of glycerol in the peripheral circulation via lipoprotein lipase action. In summary, 1) regional FFA-to-glycerol release ratios do not mirror systemic ratios, 2) leg glycerol uptake occurs in humans, and 3) splanchnic triglyceride production rates are substantial relative to systemic glycerol appearance. Glycerol appearance rates may not be a quantitative index of whole body lipolysis.
Article
Clearance of the radioactive inert gas 133Xenon injected muscularly was studied in 46 healthy subjects and 13 patients with radiologically verified occlusive disease of the main arteries of the legs and intermittent claudication. In contrast with 24Na or 131 I ions 133Xenon diffused freely across cell membranes and local muscle blood-flow can therefore readily be calculated in ml per 100 g per minute from the 133Xenon clearance-rate. Muscle blood flow (MBF) averaged 1.5 ml per 100 g per minute at rest in patients with vascular disease. This was not significantly lower than the average of 2.0 ml per 100 g per minute in healthy persons over age 50. Healthy subjects younger than 50 years had an average resting MBF of 2.2 ml per 100 g per minute. After work induced ischemia maximal MBF inpatients with vascular disease averaged 17 ml per 100 g per minute. This value was significantly below the corresponding normal values of 55 and 52 ml per 100 g per minute in the normal subjects. In the same patients the average time until maximal reactive hyperemia was reached was significantly longer--2.3 minutes in contrast to 0.4 minutes in normal subjects. The 133Xenon values accorded well with the results usually found by means of venous occlusion plethysmography but unlike this method 133Xenon can be used to measure MBF directly during muscular work with little interference from non-muscular tissues.
Article
Capillary permeability in skeletal muscle was investigated by applying the single injection, external registration method to the autoperfused cat gastrocnemius preparation. 51Cr-EDTA (MW: 341.2) and 57Co-B12 (MW: 1353.5) served as simultaneously injected indicators. At a plasma flow of 7 ml/100 g×min and an assumed capillary surface area, S, of 70 cma/g the permeability coefficients Pd(51Cr-EDTA) and Pd(57Co-B12) were 0.74 × 10-5 cm/s and 0.45×10-6 cm/s, respectively. However, correction for the recruitment of perfused capillaries yields Pd(51C-EDTA)- 1.5 × 10-5 cm/s and Pd(57Co-B12) = 0.9 × 10-5 cm/s. The results indicate that the permeability of a muscle capillary for hydrophilic solutes is similar to that of continuous capillaries in other tissues. The Pappenheimer pore radius estimate of 30 Å, the Karnovsky 40 Å interendothelial slit width, and pinocytotic transport are defective in accounting for the results as 51Cr-EDTA and 57Co-B12 diffuse across the capillary membrane at rates proportional to their free diffusion coefficients in water (CDC(51Cr-EDTA)/CDC(57Co-B12) = D(51Cr-EDTA)/D(57Co-B12)= 1.79). The intraendothelial patent channel system of fused vesicles might be the pathway for transcapillary exchange of hydrophilic molecules as this system alone could account for the experimental results. Only 1–2 open channels per μm capillary length are necessary to account for the observed permeabilities. For both indicators CDC was a linear function of plasma flow and CDC increased 3-3.5 fold when plasma flow increased from 4 to 52 ml/100 g× min. Assuming constant permeability coefficients, the CDC increase with plasma flow reflects an increase in capillary surface area indicating a 3-3.5 fold recruitment of capillaries. A constant ratio of CDC(51Cr-EDTA)/CDC(57Co-B12) within the perfusion range indicates that 51Cr-EDTA and 57Co-B12 can be employed at plasma flow > 4 ml/100 g× min without risk of early back-diffusion resulting in underestimation of permeability.
Article
A method for measurement of capillary permeability using external registration of gamma emitting isotopes after close arterial bolus injection was applied to the isolated inguinal fat pad in slightly fasting rabbits. An average extraction of 26 per cent for 51Cr-EDTA was found at a plasma flow of about 7 ml/100 g-min. This corresponds to a capillary diffusion capacity of 2.0 ml/100 g-min which is half the value reported for vasodilated skeletal muscle having approximately twice as great capillary surface area. Thus, adipose tissue has about the same capillary permeability during slight metabolic activation as vasodilated skeletal muscle.
Article
We tested whether FFAs influence glucose uptake by human peripheral tissues in vivo. Whole body glucose uptake, FFA turnover, energy expenditure and substrate oxidation rates, forearm glucose and FFA uptake, and nonoxidative glycolysis (net release of alanine and lactate) were measured in 14 normal male subjects in the basal state (0-240 min; serum insulin, approximately 5 microU/mL) and during euglycemic hyperinsulinemia (240-360 min; approximately 75 microU/mL) on 2 separate occasions, once during elevation of plasma FFA by infusions of Intralipid and heparin (plasma FFA, 4.6 +/- 0.1 vs. 4.2 +/- 0.4 mmol/L; 180-240 vs. 300-360 min) and once during infusion of saline (plasma FFA, 0.50 +/- 0.07 vs. 0.02 +/- 0.07 mmol/L, respectively). In the basal state, whole body glucose disposal remained unchanged, but the fate of glucose was significantly altered toward diminished oxidation (7.3 +/- 0.8 vs. 5.6 +/- 0.5 mumol/kg.min; P less than 0.05, saline vs. Intralipid) and increased nonoxidative glycolysis (P less than 0.05). Elevation of plasma FFA significantly increased forearm glucose uptake (1.0 +/- 0.6 vs. 2.4 +/- 0.7 mumol/kg.min; P less than 0.01) and nonoxidative glycolysis (net release of alanine and lactate, 0.4 +/- 0.5 vs. 1.2 +/- 0.4 mumol glucose equivalents/kg.min; P less than 0.05). During hyperinsulinemia, FFA decreased whole body glucose disposal (38 +/- 2 vs. 30 +/- 3 mumol/kg.min; P less than 0.001) due to a decrease in glucose oxidation (13 +/- 1 vs. 7 +/- 1 mumol/kg.min; P less than 0.01, saline vs. Intralipid), and forearm glucose uptake (31 +/- 4 vs. 24 +/- 6 mumol/kg.min; P less than 0.01, saline vs. Intralipid). Under these conditions, 7 +/- 2% and 3 +/- 1% (P less than 0.05) of forearm glucose uptake could be accounted for by nonoxidative glycolysis in the Intralipid and saline studies, respectively. In summary, 1) elevation of plasma FFA concentrations suppresses the rate of carbohydrate oxidation to a rate that, both basally and during hyperinsulinemia, is similar to that reported for insulin-independent glucose oxidation in the brain; 2) basally, forearm glucose uptake is increased by FFA; and 3) during hyperinsulinemia, FFA inhibit glucose uptake by forearm tissues. We conclude that the interaction between glucose and FFA fuels in human forearm tissues is dependent upon the ambient insulin concentration; the increase in basal glucose uptake would be compatible with the increase need of glucose for FFA reesterification; the decrease in insulin-stimulated glucose uptake supports operation of the glucose-FFA cycle in human forearm tissues.
Article
The cellular control of intramuscular triglyceride (TG) metabolism involves two major identified lipases: hormone-sensitive lipase (HSL) and lipoprotein lipase (LPL). Recently, the presence of HSL in muscle has been unequivocally demonstrated. However, although it is thought that HSL is responsible for intramuscular TG lipolysis, direct evidence for this is lacking. There is evidence to suggest that HSL and LPL are simultaneously activated under a variety of conditions. The two muscle lipases appear to be turned on by the same signal and function as a coordinated unit in meeting the energy demands of muscle. At a time when HSL is presumably hydrolyzing endogenous TG, LPL is sent to the capillary beds in search of substrate. TG uptake from circulation is highly related to muscle LPL activity. Exercise training increases LPL activity in plasma and in parenchymal cells in muscle. These results suggest that training may increase the capacity to clear TG from circulation and that LPL might have a role in replenishing muscle TG stores that have been decreased with exercise.