Download REGULATION OF KETONE BODY AND COENZYME A

Document related concepts

Proteolysis wikipedia , lookup

Oligonucleotide synthesis wikipedia , lookup

Deoxyribozyme wikipedia , lookup

Point mutation wikipedia , lookup

Peptide synthesis wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Butyric acid wikipedia , lookup

Specialized pro-resolving mediators wikipedia , lookup

Glucose wikipedia , lookup

Nucleic acid analogue wikipedia , lookup

Basal metabolic rate wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Isotopic labeling wikipedia , lookup

Hepoxilin wikipedia , lookup

Amino acid synthesis wikipedia , lookup

Biosynthesis wikipedia , lookup

Citric acid cycle wikipedia , lookup

Glycolysis wikipedia , lookup

Biochemistry wikipedia , lookup

Metabolism wikipedia , lookup

Glyceroneogenesis wikipedia , lookup

Fatty acid synthesis wikipedia , lookup

Fatty acid metabolism wikipedia , lookup

Ketosis wikipedia , lookup

Transcript
REGULATION OF KETONE BODY AND COENZYME A
METABOLISM IN LIVER
by
SHUANG DENG
Submitted in partial fulfillment of the requirements
For the Degree of Doctor of Philosophy
Dissertation Adviser: Henri Brunengraber, M.D., Ph.D.
Department of Nutrition
CASE WESTERN RESERVE UNIVERSITY
August, 2011
SCHOOL OF GRADUATE STUDIES
We hereby approve the thesis/dissertation of
__________________
Shuang Deng ____________
_
_
Doctor of Philosophy
candidate for the ________________________________degree
*.
Edith Lerner, PhD
(signed) ________________________________________________
(chair of the committee)
Henri Brunengraber, MD, PhD
________________________________________________
Colleen Croniger, PhD
________________________________________________
Paul Ernsberger, PhD
________________________________________________
Janos Kerner, PhD
________________________________________________
Michelle Puchowicz, PhD
________________________________________________
June 23, 2011
(date) _______________________
*We also certify that written approval has been obtained for any
proprietary material contained therein.
I dedicate this work to my parents,
my son and my husband
TABLE OF CONTENTS
Table of Contents………………………………………………………………….
iv
List of Tables……………………………………………………………………….
viii
List of Figures………………………………………………………………………
ix
Acknowledgements……………………………………………………………….
xii
List of Abbreviations……………………………………………………………….
xiv
Abstract……………………………………………………………………………..
xvii
CHAPTER 1: KETONE BODY METABOLISM
1.1 Overview………………………………………………………………………..
1
1.1.1 General introduction of ketone bodies………………………………..
1
1.1.2 Ketogenesis is stimulated by fasting, stress and diabetes…………
1
1.1.3 Ketone body utilization in peripheral tissues……………………….
2
1.2 Adipose tissue lipolysis and regulation……………………………………..
3
1.2.1 General introduction of lipolysis……………………………………….
3
1.2.2 Lipolysis is regulated by dietary, hormonal and neurological
factors…………………………………………………………………...
3
1.2.3 Enzymes involved in lipolysis………………………………………….
4
1.2.4 Mechanism of lipolysis………………………………………………….
5
1.2.5 Measurement of lipolysis……………………………………………….
7
1.2.6 Adipocyte lipolysis provides substrates for ketogenesis……………
8
1.3 Fatty acid β-oxidation and its regulation…………………………………….
9
1.3.1 Formation of acyl-CoA and its regulation…………………………….
9
1.3.2 CPT system and its regulation by malonyl- CoA, dietary and
iv
hormonal factors………………………………………………………...
10
1.3.3 The mitochondrial β-oxidation cycle and its regulation……………..
12
1.3.4 Generation and utilization of acetyl-CoA……………………………..
13
1.4 The β-hydroxy-β-methylglutaryl-CoA (HMG-CoA) cycle and its
regulation.………………………………………………………………..
14
1.4.1 Overview of the HMG-CoA cycle……………………………………...
14
1.4.2 Function of the mitochondria HMG-CoA……………………………..
14
1.4.3 Regulation of the HMG-CoA cycle…………………………………….
14
1.5 C5-ketogenesis and its regulation……………………………………………
15
1.5.1 Overview…………………………………………………………………
15
1.5.2 Sources of odd-chain fatty acids in animals………………………….
17
1.5.3 C5-ketone body formation and its regulation…………………………
17
1.6 The ketone body utilization…………………………………………………...
18
1.6.1 The ketone body utilization pathway………………………………….
18
1.6.2 The ketone body utilization in fetal liver……………………………..
20
1.6.3 The ketone body utilization in peripheral tissues……………………
21
1.7 The role of ketone bodies in mammalian metabolism……………………..
24
1.8 The measurement of ketone body turnover………………………………...
25
1.8.1 The measurement of ketone body turnover with isotopic methods..
25
1.8.2 Pseudoketogenesis……………………………………………………..
27
CHAPTER 2: ANAPLEROSIS
2.1 Overview………………………………………………………………………..
29
2.2 The cataplerosis………………………………………………………………
29
v
2.3 The anaplerosis………………………………………………………………..
31
2.3.1 The significance of anaplerosis………………………………………..
31
2.3.2 Anaplerotic substrates………………………………………………….
32
2.3.3 The measurement of anaplerosis……………………………………..
35
2.3.4 The anaplerotic diet therapy…………………………………………...
36
CHAPTER 3: THE SYNTHESES OF ADENINE NUCLEOTIDES,
COENZYME A (CoA) AND DEOXYRIBONUCLEIC ACID (DNA) IN RAT
LIVER
3.1 Overview of the biosynthesis pathways of adenine nucleotides, CoA
and DNA………………………………………………………………………...
38
3.1.1 Overview…………………………………………………………………
38
3.1.2 The biosynthesis pathway of adenine nucleotides…………………
38
3.1.3 The CoA biosynthesis pathway………………………………………..
42
3.1.4 The DNA biosynthesis pathway……………………………………….
45
3.2 The techniques used for tracing the biosynthesis pathways……………..
46
3.2.1 The techniques used for tracing ATP synthesis from ADP…………
47
3.2.2 The echniques used for tracing CoA synthesis……………………...
48
3.2.3 The techniques used for tracing DNA synthesis…………………….
48
3.3 Tracing the syntheses of biopolymers with [2H]water……………………..
49
3.3.1 Overview…………………………………………………………………
49
3.3.2 Principle of the use of [2H]water to trace the biosyntheses of
(pseudo)-biopolymers…………………………………………………..
50
3.3.3 General protocols of application of [2H]water………………………..
53
vi
3.3.4 The pros and cons of using [2H]water………………………………...
54
3.3.5 Tracing the syntheses of biopolymers………………………………..
56
CHAPTER 4: RESEARCH PROPOSAL
4.1 Project 1. C4- and C5-ketogenesis in rat liver………………………………
62
4.2 Project 2. Tracing the syntheses of adenine nucleotides, CoA and DNA
65
in rat liver…………………………………………………………..
4.3 Publications……………………………………………………………………
68
4.3.1 Deng S., Zhang G.F., Kasumov T., Roe C.R., and Brunengraber
H. Interrelations between C4-ketogenesis, C5-ketogenesis, and
anaplerosis In the perfused rat liver. J Biol Chem 284:2779927807,2009.
……………………………………………………………………………
68
4.3.2 Deng S., Zhang G.F., Kombu R.S. Harris S.R. DeSantis D.,
Vasquez E.J., Puchowicz M.A., Anderson V.E., Brunengraber H.
Tracing the syntheses of adenine nucleotides, CoA and DNA in
rat liver. To be submitted to J Biol Chem.
……………………………………………………………………………
105
CHAPTER 5: IMPLICATIONS AND FUTURE DIRECTIONS
5.1 C4-and C5-ketogenesis in rat liver…………………………………………...
146
5.1.1 Results and Discussion……………………………………………….
146
5.1.2 Future directions………………………………………………………..
148
5.2 Tracing the syntheses of adenine nucleotides, CoA and DNA in rat liver
…………………………………………………………………………...
150
5.2.1 Discussion and conclusions…………………………………………...
150
5.2.2 Future directions………………………………………………………..
152
LITERATURE CITED……………………………………………………………..
154
vii
LIST OF TABLES
Table 4.1. Apparent kinetics of the labeling of CoA and its components
from 2H-enriched body water in rats over 10 and 31 days……….
viii
129
LIST OF FIGURES
Figure 1.1. Mechanism of adipocyte lipolysis regulation mediated by PKA…
6
Figure 1.2. Formation of ketone bodies from fatty acid partial β-oxidation….
16
Figure 1.3. Pathway of ketone-body utilization in peripheral tissues…………
19
Figure 1.4. Pseudoketogensis in extrahepatic tissues…………………………
28
Figure 2.1. Main anaplerotic processes feeding into citric acid cycle………..
30
Figure 3.1. Biosynthesis pathways of adenine nucleotides, CoA and DNA
…………………………………………………………………………
39
Figure 3.2. Pathway of Coenzyme A synthesis………………………………...
43
Figure 3.3. Protocol for using [2H]water to trace the rates of synthesis and
degradation of biopolymers in vivo…………………………………
55
Figure 4.1. Scheme of C4-ketogenesis and C5-ketogenesis in the liver……..
95
Figure 4.2. Comparison between the uptake of octanoate (A), heptanoate
(B), or propionate (C) and the production of C4-ketone bodies
and C5-ketone bodies…………………………………………………
96
Figure 4.3. Competition between octanoate and heptanoate for uptake by
perfused rat livers…………………………………………………….
97
Figure 4.4. Competition between C4-ketogenesis from octanoate and C5ketogenesis from heptanoate in perfused rat livers………………
98
Figure 4.5. Profiles of concentrations of octanoate (●) and propionate (▲)
in the effluent perfusate………………………………………………
Figure 4.6. Labeling pattern of effluent β-hydroxybutyrate (BHB) and tissue
acetyl-CoA from livers perfused with increasing concentrations
ix
99
of [1-13C]octanoate (A) or [8-13C]octanoate (B)…………………..
100
Figure 4.7 Sharing of acetyl groups between C4- and C5-ketogenesis
reflected by the mass isotopomer distribution of BHB and BHP…
101
Figure 4.8 Mass isotopomer distribution of HMG-CoA (A) and HEG-CoA
(B) in livers perfused with constant 1 mM [1-13C]heptanoate and
increasing concentrations of unlabeled octanoate………………..
102
Figure 4.9 Mass isotopomer distribution of BHB-CoA and AcAc-CoA in
livers perfused with increasing concentrations of [1-13C]octanoate
………………………………………………………………………….
103
Figure 4.10 Anaplerosis and glucose labeling from increasing
concentrations of [13C3]propionate (♦) or [5,6,7-13C3]heptanoate
(■,▲)………………………………………………………………….
104
Figure 4.11. Sites of labeling of nucleotides, CoA and DNA from [2H]water
………………………………………………………………………... 132
Figure 4.12. In vivo labeling of liver CoA from [2H]water over 31 days……… 133
Figure 4.13. In vivo labeling of liver dR-DNA [2H] from water over 10 days… 134
Figure 4.14. Mass isotopomer distribution of liver perfusate glucose in
100% 2H2O buffer…………………………………………………...
135
Figure 4.15. Mass isotopomer distribution of glucose-6-P, ribose-5-P and
PEP in livers perfused with 4 mM unlabeled glucose in 100%
2
H2O buffer…………………………………………………………..
136
Figure 4.16. M1 enrichment of AMP, ADP, ATP and CoA in livers perfused
with 4 mM unlabeled glucose in 100% 2H2O buffer……………… 137
x
Figure 4.17. M1 enrichment of AMP, ADP, ATP in livers perfused with
4 mM unlabeled glucose in buffers enriched 0 to 100% with
2
H2O…………………………………………………………………...
138
Figure 4.18. M1 enrichment of CoA and its components in livers perfused
with 4 mM unlabeled glucose in buffers enriched 0 to 100%
with 2H2O……………………………………………………………..
139
Figure 4.19. Mass isotopomer distribution of perfusate glucose in liver
perfusion experiments starting with 4 mM [13C6]glucose………..
140
Figure 4.20. Mass isotopomer distribution of glucose-6-P, ribose-5-P and
glycerate-3-P from livers perfused with 4 mM [13C6]glucose……
141
Figure 4.21. Mass isotopomer distribution of AMP, ATP and CoA in livers
perfused with 4 mM [13C6]glucose………………………………… 142
Figure 4.22. Mass isotopomer distribution of perfusate glucose in liver
perfusion experiments starting with 4 mM unlabeled glucose +
2 mM [13C5]ribose…………………………………………………..
143
Figure 4.23. Mass isotopomer distribution of glucose-6-P, ribose-5-P, PEP
and glycerate-3-P from livers perfused with 4 mM unlabeled
glucose + 2 mM [13C5]ribose………………………………………..
144
Figure 4.24. Mass isotopomer distribution of AMP, ADP, ATP and CoA in
livers perfused with 4 mM unlabeled glucose + 2 mM
[13C5]ribose…………………………………………………………..
xi
145
ACKNOWLEDGEMENTS
I owe many thanks and appreciations to those people who are with me during the
past four years.
First and foremost, I would like to thank my adviser, Dr. Henri Brunengraber. I
sincerely thank him for giving me the chance to continue my education. I thank
him for his scientific guidance, support, patience and humor on the 5,000 li road.
He is an adviser who not only teaches me critical thinking, but also is always
willing to sit and discuss every detail of my experiments, presentations and
scientific writings. His guidance gives me enough confidence to face the
challenge of my future endeavors. Undoubtedly, this four-year training
experience has changed my life for ever.
I would also give my sincere gratitude to my dissertation committee: Dr. Edith
Lerner, Dr. Colleen Croniger, Dr. Paul Ernsberger, Dr. Janos Kerner and Dr.
Michelle Puchowicz. Their unlimited advices and kindness helped me get through
frustrating times. They witnessed my scientific growth and I deeply appreciate it.
I also want to thank all my colleagues who made my life here really enjoyable. I
would like to thank Dr. Guo-Fang and Dr. Rajan for offering their help whenever it
was needed. I started with my first living perfusion with France David and John
Koshy four years ago; This was the best team I ever worked with. I thank
Stephanie Harris for sharing so many things with me. She made me feel that I
am always on the “short-cut”. I had such great time work with Sharon, lan, Qinlin
and Edwin. I will never forget it! I am so grateful to Fred, Brian and David, without
xii
their help, I could not finish my experiments. I thank Asha for cleaning my messy
stuffs. Last but not least, I give my deepest thanks to Sophie for spending
countless hours on my papers and figures.
Finally, I want to give my special thanks to my family and my friends. My parents,
my son and my husband supported me with their love in the long way. My friends
were always there when I needed them.
All of these friends and colleagues made me enjoy my life in this country!
xiii
LIST OF ABBREVIATIONS
AcAc
Acetoacetate
ACC
Acetyl-CoA carboxylase
ADP
Adenosine diphosphate
ACBP
Acyl-CoA binding protein
AMP
Adenosine monophosphate
ATLG
Adipose triglyceride lipase
ATP
Adenosine triphosphate
BBB
Blood brain barrie
BHB
β-Hydroxybutyrate
BHP
β-Hydroxypentanoate
BKP
β-Ketopentanoate
BrdU
5’- Bromo - 2’- deoxyuridine
CAC
Citric acid cycle
CDP
Cytidine diphosphate
CoA
Coenzyme A
CPT
Carnitine-palmitoyltransferae
DAG
Diacylglycerol
dATP
Deoxyadenosine triphosphate
dCTP
Deoxycytidine triphosphate
DNA
Deoxyribonucleic acid
dNTP
Deoxyribonucleotide-triphosphate
dGTP
Deoxyguanosine triphosphate
xiv
dR
Deoxyribose
dTTP
Deoxythymidine triphosphate
GC-MS
Gas chromatography – mass spectrometry
GNG
Gluconeogenesis
FABP
Fatty acid binding protein
FOD
Fatty acid oxidation disorders
GABA
γ-Aminobutyric acid
GDP
Guanosine diphosphate
HEG-CoA
β-Hydroxy-β-ethylglutaryl-CoA
HMG-CoA
β-Hydroxy-β-methylglutaryl-CoA
HPLC
High performance liquid chromatorgraphy
HSL
Hormone-sensitive lipase
MCD
Malonyl-CoA decarboxylase
MCT
Monocarboxylate transporter
MGL
Monoacylglycerol lipase
MPE
Molar percent enrichment
OAA
Oxaloacetate
Ra
Rate of appearance
PC
Pyruvate carboxylase
PEPCK
Phosphoenolpyruvate carboxylase
PKA
Protein kinase A
PRPP
Phosphoribosylpyrophosphate
TDP
Thymidine diphosphate
xv
TG
Triglycerides
xvi
Regulation of Ketone Body and Coenzyme A Metabolism in Liver
Abstract
by
SHUANG DENG
The dietary treatment of patients with long-chain fatty acid oxidation disorders is
in a transition period from classical medium-even-chain trioctanoin to mediumodd-chain triheptanoin. In my first project, I investigated the interrelations
between C4- ketogenesis (production of β-hydroxybutyrate + acetoacetate), C5ketogenesis (production of β-hydroxypentanoate + ketopentanoate), and
anaplerosis in isolated rat livers perfused with 13C-labeled octanoate, heptanoate,
or propionate. Although the uptakes of octanoate and heptanoate by the liver are
similar, the rate of C5-ketogenesis from heptanoate is much lower than the rate of
C4-ketogenesis. This results from the channeling of the propionyl moiety of
heptanoate
into
anaplerosis
and
gluconeogenesis.
C5-ketogenesis
from
propionate is virtually nil because the kinetics of acetoacetyl-CoA thiolase do not
favor the formation of β-ketopentanoyl-CoA from propionyl-CoA and acetyl-CoA.
Anaplerosis and gluconeogenesis from heptanoate are inhibited by octanoate.
The data have implications for the design of diets for the treatment of long-chain
fatty acid oxidation disorders, such as the triheptanoin-based diet.
xvii
The goal of my second project was to sort out some of the mechanisms by which
2
H from 2H-enriched body fluids becomes incorporated into C-H bonds during the
biosyntheses of adenine nucleotides, coenzyme A and DNA in rat liver. I
perfused rat livers with 2H-enriched water, or buffer containing [13C6]glucose, or
[13C5]ribose. The mass isotopomer distribution of glucose and glycolytic
intermediates reveals intense cycling and redistribution of label through
glycolysis, gluconeogenesis, the citric acid cycle and the pentose phosphate
pathway. In the presence of 2H-enriched water, most of the 2H found in adenine
nucleotides and CoA is incorporated between ribose-5-P and AMP. Most of the
turnover of adenine nucleotides is supported by salvage pathways. In rat livers
perfused with [13C6]glucose, I showed that the non-oxidative branch of the
pentose phosphate pathway is more active than the oxidative branch in the
fasted state. The fractional turnover rate of liver CoA is 33%/h in rat liver
perfused with 100% 2H2O buffer. The fractional turnover rate of liver adenine
nucleotides is 7%/h.
xviii
CHAPTER 1: KETONE BODY METABOLISM
1.1 Overview
1.1.1 General introduction of ketone bodies
In the metabolic literature, the expression “ketone bodies” refers to three watersoluble compounds: acetoacetate (AcAc), D-3-hydroxybutyrate (BHB) and
acetone (for extensive reviews of ketone body metabolism, see (1-3). Ketone
bodies are strongly acidic and water-soluble compounds derived from the partial
beta-oxidation of long-chain fatty acids in liver. Acetone is formed by the
decarboxylation of AcAc, catalyzed by free amino groups of proteins (4). The
concentrations of ketone bodies in plasma range from 0.1- 0.3 mM in the fed
state (5; 6), 1-3 mM in fasting (7-11) and up to 25 mM in decompensated
diabetes (12; 13). In most cases, ketone bodies are derived from even-chain fatty
acids and have 4 carbons (C4-ketone bodies). When ketone bodies are derived
from odd-chain fatty acids, they have 5 carbons (C5-ketone bodies, which are
described in section 1.5).
1.1.2 Ketogenesis is stimulated by fasting, stress and diabetes
Ketone body production is stimulated by fasting, stress or diabetes. Mobilization
of lipids in white adipose tissue lipolysis releases free fatty acids, which travel
through the circulating system, bound to plasma albumin and enter into the liver
(14-16). The liver is the only organ that synthesizes ketone bodies (1). Under
some special conditions (such as suckling), ketogenesis also occurs in the small
intestine, kidney and brain astrocytes (17). Long-chain fatty acids are converted
to long-chain acyl-CoAs in the hepatocyte cytosol. These CoA esters are then
1
transported into the mitochondria via the carnitine palmitoyltransferase (CPT)
system (1; 14; 18). Malonyl-CoA inhibits CPT I and regulates the entrance of
long-chain acyl-CoAs into the mitochondria (19; 20). Soluble medium-chain fatty
acids (8-10 carbons) enter directly into the mitochondria and their oxidation is
independent of the CPT system (1). In the mitochondria, fatty acids are oxidized
to acetyl-CoA. The liver, a small organ (2-3% of body weight), cannot oxidize all
the acetyl-CoA derived from fatty acid β-oxidation. The excess acetyl-CoA (80%
of the nonesterified fatty acids carbons going down β-oxidation) enters the βhydroxy-β-methylglutaryl-CoA (HMG-CoA) cycle and produces ketone bodies (1).
1.1.3 Ketone body utilization in peripheral tissues
For a long time, ketone bodies were regarded as harmful because high
concentrations of acidic ketone bodies accumulate in decompensated diabetics
(12; 13). However, ketone bodies are good fuels for peripheral tissues if the
capacity of these tissues to utilize ketone bodies is not inhibited by a large supply
of competing substrates such as glucose (2). Ketone body utilization occurs by
two pathways: oxidation in the mitochondria of non-hepatic tissues (21) and
lipogenesis in the cytosol of liver, lactating mammary gland, and developing brain
(22; 23). Ketone bodies provide alternative fuels to peripheral tissues when
glucose is in short supply or cannot be efficiently used. This is especially
important for the brain because glucose is its main energy substrate (2). Owen
et al. (24) found that the brain can also use ketone bodies during extended
starvation. The developing brain uses ketone bodies to synthesize lipids (25-27).
2
Ketone body utilization spares muscle proteins in starvation (28). The lactating
mammary gland uses ketone bodies to synthesize milk lipids (29).
1.2 Adipose tissue lipolysis and regulation
1.2.1 General introduction of lipolysis
In mammals, about 95% of body fats are stored in adipose tissue during
abundant energy supply. A small amount of fat is found in muscle, liver and
pancreas (30). The fats are stored mostly as hydrophobic lipids – triglycerides
(TGs). In adipose tissue, the fats are mobilized when there is a shortage of
dietary substrates (31). Body fat distribution is different between man and
woman. Females have more total body fat and a higher percentage of lower body
fat than males do (32). In addition to supplying stored energy for the body,
adipose tissue is also an endocrine organ which secretes adipokines (31).
Adipose tissue has two main metabolic functions: lipogenesis and lipolysis.
Lipolysis is a process by which triglycerides are hydrolyzed to fatty acids and
glycerol. It occurs in extracellular space, regulated by lipoprotein lipase in
postprandial conditions, and in intracellular space mediated by hormonesensitive lipase. Adipocyte lipolysis controls the release of fatty acids into the
circulating system and is a major regulator of lipid metabolism (33).
1.2.2 Lipolysis is regulated by dietary, hormonal and neurological factors
Lipolysis is highly reglulated by diet, hormonal and neural factors. Variations in
dietary composition have obvious effects on TG lipolysis. Ingestion of a high fat
diet increases glycerol release in gut. A high carbohydrate decreases glycerol
release in gut (34). Leptin, glucocorticoids, glucagon, catecholamines, T3 and T4
3
are lipolytic hormones that stimulate lipolysis (31). Insulin, prostaglandins, IL-6
are antilipolytic and reduce lipolysis (34). The most important hormonal factors
are lipolytic catecholamines and antilipolytic insulin. Catecholamines are “fight or
flight” hormones produced by the adrenal glands and sympathetic nerves: they
respond to stress. The most abundant catecholamines are epinephrine,
norepinephrine and dopamine. Catecholamines trigger lipolysis by increasing
cAMP and activating protein kinase A (PKA) (30). Another potent lipolytic
compound, atrial natriuretic peptide, triggers lipolysis during repeated bouts of
exercise (35). Catecholamine resistance is found in obese patients, caused by
the decreased number of β2-adrenoreceptors. Catecholamine resistance
contributes to reduced lipolysis both in vivo and in vitro (36). In contrast, the most
important antilipolytic hormone insulin blocks lipolysis via degrading cAMP and
activating protein phosphatase (37).
1.2.3 Enzymes involved in lipolysis
Hormone-sensitive lipase (HSL)
Hormone-sensitive lipase (HSL) is the rate-limiting enzyme of lipolysis. Its activity
is regulated by catecholamines and insulin. The structure of HSL from rat
adipose tissue was first identified by Holm’s group (38). It consists of 768 amino
acids and the molecular weight is 84 kDa (39). The phosphorylation sites
responding to lypolytic stimulations are within 150 amino acids close to the Cterminal (40). HSL is mainly expressed in adipose tissue (41), but it is also found
in muscle (42), macrophages (43) and pancreatic β-cells (44). Obese people
have decreased adipocyte HSL activity and expression. This contributes to the
4
decrease of lipolysis and the increases in adipose tissue mass (45). Fatty acids
also inhibit the HSL hydrolytic function; this inhibition is reduced by HSL binding
to adipocyte lipid-binding protein both in vivo and in vitro (46).
Adipose triglyceride lipase (ATGL)
Another lipase with triacylglycerol hydrolase activity is adipose triglyceride lipase
(ATGL) (also called desnutrin or phospholipase A2)(47). ATGL is highly
expressed in adipose tissue with a higher activity to hydrolyze triacylglycerols
than HSL. ATGL predominates in basal lipolysis (48). ATGL-null mice show
disruption of energy homeostasis by the defective lipolysis (49). ATGL does not
respond to catecholamines. Unlike HSL, ATGL cannot be phosphorylated by
PKA (50): it is under the control of peroxisome proliferator-activated receptor γ
and insulin (51).
Monoacylglyerol lipase (MGL)
Monoacylglycerol
lipase
(MGL)
is
a
necessary
enzyme
to
hydrolyze
monoacylglycerols to free fatty acids and glycerol (30). This enzyme is not under
hormonal regulation and is not the rate-limiting step in lipolysis (52).
1.2.4 Mechanism of lipolysis
A well known control mechanism of lipolysis is exerted by the regulation of HSL
(Fig 1.1). This regulation is mediated by cAMP and protein kinase A (PKA) (53).
β-adrenergic catecholamines, the release of which is triggered by the
sympathetic nervous system, increase during fasting, physical exercise or
stressful conditions. They bind to G-protein coupled β-adrenoreceptors, which
5
Figure 1.1: Mechanism of adipocyte lipolysis regulation mediated by PKA.
Modified from Fukao et al. (54) Prostaglandins, Leukotrienes and Essential Fatty
Acids 70 (2004): 243-251.
6
activate adenyl cyclase and increase intracellular cAMP concentration (55). The
and Ser660 residues (41). PKA also phosphorylates perilipin, which is a protein
located on the surface of lipid droplets and blocking the contact of HSL with lipids
(56). There are two possible mechanisms by which phosphorylated perilipins
influence lipolysis. Londos et al. (57) reported that once perilipins are
phosphorylated, they disassociate from the lipid droplets and leave the lipid
surface for the activated HSL. Another mechanism proposed by the Granneman
group (58; 59) is that perilipins are “scaffold proteins”. The phosphorylated
perilipins change their structures and recruit HSL to access the lipid surface.
ATGL and HSL together hydrolyze 95% of TGs to diacylglycerol (DAG) in mouse
adipose tissue (60). HSL continues to hydrolyze DAG to MAG. MGL is the last
enzyme that converts MAG to one glycerol and one fatty acid molecule. The
released fatty acid and glycerol bound to transporters are exported into the blood
(30).
1.2.5 Measurement of lipolysis
Adipocyte lipolysis is measured both in vivo and in vitro. In vivo measurements
provide more accurate and reliable results than in vitro measurements because
all physiological influence come to play (61). Triglycerides are hydrolyzed to free
fatty acids and glycerol in a 3/1 ratio. If the ratio is lower than 3:1, this means that
some free fatty acids have been reesterified in adipose tissue (62). Plasma free
fatty acid and glycerol concentration are indicators of adipose tissue lipolysis.
Radioactive or stable tracers are reliable techniques to determine systemic
lipolysis in vivo (33). Stable isotopes techniques use gas-chromatography, HPLC
7
and mass spectrometry techniques. Tracers of free fatty acids (63; 64) or glycerol
(65) are infused in vivo, and the appearance rate (Ra) of free fatty acids and
glycerol are determined from their steady-state labeling in plasma. Palmitate or
oleate are good tracers because the Ra of these single fatty acids reflects the Ra
of plasma total free fatty acids. Local lipolysis is mostly determined by
arteriovenous (A-V) differences or microdialysis (33). A-V differences are based
on the measurement of metabolite concentrations across subcutaneous adipose
tissue (66). Microdialysis uses semipermeable dialysis catheters implanted in a
tissue, and analyzes the metabolite concentrations in the effluent dialysis fluid.
However, this method can only measure the interstitial glycerol not fatty acid
concentration because fatty acids are not dialyzed via standard dialysis
membranes (67).
1.2.6 Adipocyte lipolysis provides substrate for ketogenesis
TG lipolysis in adipose tissue (adipocyte lipolysis) is the first key step for
ketogenesis since it provides ketogenic substrates (free fatty acids) to the liver.
Scow et al. (68) found that in pancreatectomized rats depleted of adipose tissue,
there was no ketosis, even at extremely low level of insulin. Under modest
starvation conditions, plasma free fatty acid concentration is less than 1 mM, and
ketone body level is lower than 5 mM. However, in decompensated diabetes, the
free fatty acid level ranges from 2 - 4 mM, which increases ketone body
concentration to 20 mM (69; 70). Although
free fatty acids and ketone bodies
are “precursor and product”, the nonparallel relation between the fatty acid and
ketone body levels results from variations in the esterification rate (71).
8
1.3 Fatty acid β-oxidation and its regulation
1.3.1 Formation of acyl-CoA and its regulation
It is still not clear whether long-chain fatty acids enter hepatic cells via diffusion
(72) or protein-mediated fatty acid transporters. Transported insoluble long-chain
fatty acids (normally C14-18) bind to cytosolic fatty acid binding proteins (FABP),
which decrease the potential toxicity of free fatty acids (17). Inhibition of the
binding potentially decreases fatty acid flux in liver (73).
Long-chain acyl-CoAs are formed by acyl-CoA synthetases, which are present in
mitochondria (74), peroxisomes (75) and endoplasmic reticulum (76). The ratio
of NAD/NADH is a major control of β-oxidation. Even if the formation of acyl-CoA
is not a major control site of β-oxidation, the ratio of [acyl-CoA]/[CoASH] affects
β-oxidation flux via regulation of acyl-CoA synthetase in vitro (17). Before acylCoA binding protein (ACBP) was isolated from mammalian tissues, acyl-CoAs
were found binding to FABPs (17). ACBP has a very high affinity for long-chain
acyl-CoAs; this
minimizes the hydrolysis of the esters (77). It is generally
accepted that the activated fatty acids are used by two metabolic pathways in
liver: cytosolic lipid synthesis via glycerol-3-phosphate, and mitochondrial
oxidation. The two processes are highly regulated. Three proposed pathways for
delivering fatty acids from plasma to mitochondria were summarized by Eaton
(17): first, fatty acids enter the cytosol via transporter proteins and bind to
FABPs. They are either converted to acyl-CoAs in the plasma membrane or in
the cytosol. Second mechanism is that the ACBP-bound acyl-CoAs move to the
mitochondrial outer membrane. The third mechanism makes fatty acids bind to
9
FABP and migrate to the mitochondrial membrane, where they are converted to
acyl-CoAs.
1.3.2 The CPT system and its regulation by malonyl - CoA, dietary and
hormonal factors
The mitochondrial carnitine palmitoyltransferase (CPT) system and its
regulation
The transport of long-chain fatty acyl-CoAs from cytosol to mitochondria is highly
dependent on the CPT system. Acyl-CoAs are first converted by carnitine
palmitoyl-transferase I (CPT I) at the outer mitochondrial membrane to
acylcarnitines, which are transferred to the mitochondrial matrix via a
translocase. Finally, mitochondrial carnitine-palmitoyl-transferase II (CPT II)
catalyzes the reconversion of acylcarnitines back to acyl-CoAs. Unlike long-chain
fatty acids, short-chain or medium-chain fatty acids bypass the carnitine system
and directly enter into mitochondria, where they are acylated to acyl-CoAs by
mitochondrial acyl-coA synthetase (17). As a result, the oxidation of short-chain
or medium-chain fatty acids is independent of CPT I regulation.
The concept that CPT I is the control site for hepatic β-oxidation originated from
data generated in rat livers perfused with octanoate (1). McGarry (1) et al. found
that the octanoate oxidation rate is similar in livers from fed and fasted rats.
However, the oleate β-oxidation rate varies with nutritional conditions (1). This is
because octanoate directly enters mitochondria independent of the CPT I
system. This finding lead to the understanding that CPT I is subject to metabolic
regulation. CPT I has two isoforms: L-CPT I in liver and M-CPT I in muscle,
10
which have different kinetic properties (54). CPT I is regulated by fuel status
(fatty acids, ketone bodies, glucose) via controlling malonyl-CoA and hormonal
levels (78).
Malonyl-CoA regulation
In the liver, malonyl-CoA is a signaling metabolite as well as an intermediate of
fatty acid synthesis (79). The concentration of hepatic malonyl-CoA level is about
13 nmol•g-1 in the fed state; it decreases to half in the fasted state (80). MalonylCoA is the product of acetyl-CoA carboxylase (ACC), the activity of which is
highly up-regulated by insulin and down-regulated by glucagon and AMPK. The
increased glucogon or epinephrine raises cAMP levels which lead to the
activation AMPK. AMPK phosphorylates ACC and inhibit its enzyme activity.
Insulin activates a phosphatase to dephosphorylate ACC and makes it inactive.
Thus, ACC has for a long time been considered as the major enzyme
responsible for the changes in malonyl-CoA level (81). Another relevant enzyme
is malonyl-CoA decarboxylase (MCD) which converts malonyl-CoA to acetyl-CoA
and provides a way to dispose of malonyl-CoA. Whether MCD is regulated by
AMPK are in debate. Park et al. (82) showed that MCD are positively regulated
by AMPK in rat liver, muscle and adipose tissue after exercise. The opposite
opionin was proposed by Habinowski et al. (83). They found that MCD is not a
substrate in fast twitch muscle and 832/13 ISN-1 islet cell line. The inhibition of
MCD increases malonyl-CoA concentration and decreases β-oxidation (84). ACC
and MCD work together to control cytosolic malonyl-CoA which regulates CPT I
activity.
11
Malonyl-CoA plays an important role during the transition from the fed to the
fasted states by switching hepatic fatty acid metabolism from synthesis to
oxidation and ketogenesis. In the normal fed state, malonyl-CoA inhibits CPT I
and thus decreases the transport of long-chain fatty acyl-CoAs into mitochondria:
this leads to the decrease in acyl-CoA oxidation. However, in the fasted state,
liver CPT I is less sensitive to malonyl-CoA because of an increased Ki (85).
Upon refeeding, insulin restores the sensitivity of CPT I to malonyl-CoA within 24
hours (86; 87). Other studies concluded that the desensitization may be caused
by changes in the CPT I membrane environment (79).
Hormonal regulation
The [glucagon]:[insulin] ratio affects the carnitine acyltransferase reaction. The
liver carnitine content increases with the increase of the [glucagon]:[insulin] ratio
(88). Insulin and glucagon regulate the L-CPT I gene expression (89). The
addition of glucagon to cultured fetal hepatocytes increases L-CPT I mRNAs. In
contrast, insulin depresses CPT-I mRNA abundance to about 10-fold in 8 hr in
H4IIE rat hepatoma cells (90).
1.3.3 The mitochondrial β-oxidation cycle and its regulation
The β-oxidation cycle in mitochondria catalyzes the sequential removal of twocarbon units from the acyl-CoA chain. By this process, fatty acids are converted
into acetyl-CoA to produce energy. Four intramitochondrial enzymes regulate this
cycle. Acyl-CoA dehydrogenases catalyze the first step of β-oxidation cycle and
convert acyl-CoAs to enoyl-CoAs. There are three isozymes of acyl-CoA
dehydrogenases, each of them is responsible for specific fatty acyl chain lengths.
12
Acyl-CoA dehydrogenase has the lowest activity of the β-oxidation enzymes in
rat tissues (91; 92). One of the most common fatty acid oxidation disorder (FOD)
affects very long-chain acyl-CoA dehydrogenase. Those patients accumulate
toxic long-chain or very long-chain fatty acyl-CoA in cells. This damages cell
membrane (93). The second step of β-oxidation is the formation of βhydroxyacyl-CoAs from enoyl-CoAs catalyzed by enoyl-CoA hydratase. This step
is not a control point of β-oxidation (17). In the third step, β-hydroxyacyl-CoAs
are dehydrogenated to β-ketoacyl-CoA by β-hydroxyacyl-CoA dehydrogenase.
The final step of the β-oxidation cycle is catalyzed by 3-ketoacyl-CoA thiolase,
which releases one acetyl-CoA molecule. The last three enzymes are linked
together in a trifunctional protein. The complete deficiency of this enzyme is rare
and causes sudden death (94).
1.3.4 Generation and utilization of acetyl-CoA
Early studies claimed that oxaloacetate (OAA), a citric acid cycle (CAC)
intermediate, is a primary regulator of ketogenesis in vivo (95). The low level of
anaplerosis would decrease acetyl-CoA flux to the CAC, and increase the
conversion of acetyl-CoA to ketone bodies. Williamson et al. (96) disagreed with
this mechanism. By perfusing radioactive labeled octanoate in rat livers under
different physiological conditions, they showed that enhanced hepatic fatty acid
β-oxidation produces more acetyl-CoA than can be oxidized by the CAC. The
acetyl-CoA builds up and flows to the ketogenic pathway. The latter mechanism
is generally accepted today.
13
1.4
The
β-hydroxy-β-methylglutaryl-CoA
(HMG-CoA)
cycle
and
its
regulation
1.4.1 Overview of the HMG-CoA cycle
HMG-CoA is produced by different cytosolic and mitochondrial HMG-CoA
synthases. It is a key branch point for hepatic cholesterogenesis in the cytosol
and for ketogenesis in the mitochondria (97). In 1950’s, Rudney proved that
acetyl-CoA and AcAc-CoA react to form HMG-CoA in rat and beef liver
mitochondria (98). HMG-CoA then undergoes a cleavage to form acetoacetate
and acetyl-CoA (139). The ketogenic pathway is also called the HMG-CoA cycle
(97). In the cytosol, HMG-CoA is reduced to mevalonate, an intermediate of
cholesterogenesis (97).
1.4.2 Function of the mitochondrial HMG-CoA
The formation of ketone bodies in the liver occurs via the HMG-CoA cycle. The
old view was that thiolase catalyzes the condensation of two molecules of acetylCoA to produce acetoacetyl-CoA (AcAc-CoA), which reacts with another acetylCoA molecule to form HMG-CoA, the latter is cleaved to AcAc. However, AcAcCoA is actually derived from the last four carbons of long-chain fatty acids that
undergo β-oxidation. The rate limiting step of ketogenesis is the formation of
HMG-CoA (see figure 1.2).
1.4.3 Regulation of the HMG-CoA cycle
Hepatic mitochondrial HMG-CoA synthase and HMG-CoA lyase are responsible
for ketogenesis. The absence of one or both enzymes causes the incapability to
synthesize ketone bodies in liver. Mitochondrial HMG-CoA synthase in avian
14
liver is a homodimer consisting of two 53-57 kDa monomers (99). This enzyme
was first identified as a control site for ketogenesis by Williamson’s group in 1968
(23). In 1975, Lane’s group proved that mitochondrial and cytosolic HMG-CoA
synthases have different chemical structures and functions
(97) Covalent
modification by succinyl-CoA is the main mechanism of regulation of
mitochondrial HMG-CoA synthase activity (100). Succinyl-CoA binds reversibly to
HMG-CoA synthase to form an inactive succinyl-enzyme complex. This inactive
enzyme converts to active enzyme via the release of succinate (101). In the
fasted state, increased glucagon lowers mitochondrial succinyl-CoA level and
hepatic ketogenesis is stimulated via desuccinylation of HMG-CoA synthase
(102). Moreover, a high fat or a high carbohydrate diet (103) decreases HMGCoA synthase activity.
1.5 C5-Ketogenesis and its regulation
1.5.1 Overview
β-Hydroxypentanoate (BHP) and β-ketopentanoate (BKP) are homologues of the
physiological ketone bodies BHB and AcAc. They derive from the partial-oddchain fatty acid β-oxidation. Odd-chain fatty acids are absent from the diet of
non-ruminant mammals, thus only trace amounts of C5-ketone bodies are found
in human plasma. The oxidation of 1,3-pentanediol, a potential animal nutritent,
in dog liver produce C5- ketone bodies (104). In the clinical field, BHP and BKP
were found in the urine of patients who have inherited deficiency of propionylCoA carboxylase (105). Instead of acetyl-CoA, the accumulated body propionylCoA is carboxylased by ACC to form odd-chain fatty acids, which are oxidized to
15
Fatty acids
β-oxidation
O
(AcAc-CoA)
O
CH3-C-CH2-C-SCoA
O
(Ac-CoA)
HMG-CoA Synthase
CH3-C-SCoA
(HMG-CoA)
O
OH
O
C
CH2
CH2
C
C
CH3
O
SCoA
CH3-C-SCoA
HMG-CoA lyase
O
O
-OOC-CH2-C-CH3
(Acetoacetate)
Acetoacetate
decarboxylase
O
D-β-hydroxybutyrate
dehydrogenase
CO2
CH3-C-CH3
OH
-OOC-CH2-CH-CH3
D-β-hydroxybutyrate
Acetone
Figure 1.2: Formation of ketone bodies from fatty acid partial β-oxidation.
16
form C5-ketone bodies via HMG-CoA cycle. The clinical names of BHP and BKP
are 3-hydroxyvelerate and 3-oxovalerate which are considered as abnormal
metabolites (105). Plasma C5-ketone bodies are also reported in patients who
have long-chain FOD, and are treated with anaplerotic triheptanoin diet (106).
1.5.2 Sources of odd-chain fatty acids in animals
Odd-chain fatty acids are absent or present in only trace amounts in most plants
(107). Humans acquire odd-chain fatty acids mainly from ruminant milk. They are
produced by rumen bacteria via de novo synthesis (108). Thus, the concentration
of
milk odd- and branched-chain fatty acids serve as a biomarker of the
duodenal microbial flora (109). A small amount of odd-chain fatty acids is
produced from de novo synthesis in the ruminant mammary gland. Odd-chain
fatty acids are synthesized using propionyl-CoA as the primer instead of acetylCoA (110).
1.5.3 C5-ketone body formation and its regulation
Like C4-ketone bodies, C5-ketone bodies are synthesized in liver mitochondria
from the partial β-oxidation of odd-chain fatty acids. After several β-oxidation
cycles, odd-chain fatty acids generate BKP-CoA. C5-ketone body production
uses the HMG-CoA cycle. The counterpart of HMG-CoA in C5- ketogenesis is 3hydroxy-3-ethylglutaryl-CoA
(HEG-CoA).
Incubation
of
liver
extract
with
propionyl-CoA and [1-14C]acetyl-CoA produces [14C]HEG-CoA (111). BHP and
BKP are interconverted by mitochondrial BHB dehydrogenase (112). In patients
with deficiency of biotin (113) or vitamin B12 (114), C5-ketone bodies are found in
body fluids. The formation of C5-ketone bodies involves either the conversion of
17
propionyl-CoA to BKP-CoA via 3-oxoacyl-CoA thiolase or the β-oxidation of oddchain fatty acids (115).
1.6 The ketone body utilization
1.6.1 The ketone body utilization pathway
Under normal physiological conditions, the production of ketone bodies in liver is
well balanced by their utilization in peripherial tissues. Ketone bodies are
produced during starvation as alternative substrates. Ketone bodies are used as
fuels for energy and lipogenesis (2).
AcAc is the ketone body directly used in peripherial tissues. BHB is first
converted into AcAc by 3-hydroxybutyrate dehydrogenase (116). The utilization
of AcAc includes 2 steps: the first step is the conversion of AcAc to its CoA ester,
the second step is the formation of acetyl-CoA (see figure 1.3).
The formation of acetoacetyl-CoA is catalyzed by two different enzymes:
(i) AcAc + succinyl-CoA ↔ acetoacetyl-CoA + succinate
This reaction is catalyzed by 3-oxoacid-CoA transferase, which is a mitochondrial
enzyme that is absent in liver (19), but is present in peripherial tissues. This
reaction is reversible.
(ii) AcAc + ATP + CoA → AcAc-CoA + AMP + pyrophosphate
Acetoacetyl-CoA synthetase catalyzes this irreversible reaction. This cytosolic
enzyme has lower enzymatic activity than 3-oxoacid-CoA transferase (117).
The second step is the cleavage of acetoacetyl-CoA into 2 acetyl-CoA by
acetoacetyl-CoA thiolase.
Acetoacetyl-CoA + CoA ↔ 2 acetyl-CoA
18
Figure 1.3: Pathway of ketone-body utilization in peripheral tissues. Modified
from Williamson DH et al. (2) Physiol Rev 60: 13-187, 1980.
a:
3-hydroxybutyrate
dehydrogenase;
b:
3-oxoacid-CoA
transferase;
c:
acetoacetyl-CoA thiolase; d: citrate synthase; e: acetoacetyl-CoA synthetase; f:
ATP citrate lyase; g: acetoacetyl-CoA reductase or β-ketoacyl-ACP reductase; h:
enoyl-CoA hydratase or β-hydroxyacyl-ACP dehydrase; i: enoyl-CoA reductase,
enoyl-ACP reductase, or enoyl-ACP reductase (NADPH); j; 3-hydroxybutyrylCoA synthetase or butyryl-CoA synthetase.
19
This reversible reaction occurs in mitochondria. It strongly favors the formation of
acetyl-CoA rather than acetoacetyl-CoA. Acetoacetyl-CoA thiolase is also found
in the cytosol (118), where it is part of the synthesis of lipids (119).
C5-ketone bodies are metabolized in a similar way to C4-ketone bodies (104).
The administration of BHP and BKP to conscious dogs in a dose corresponding
to 75% of caloric requirement leads to only 1.3 mM plasma C5-ketone body
concentration (104). This shows the rapid metabolism of C5-ketone bodies. BHP
is first oxidized to BKP by BHB dehydrogenase present in most cells, followed by
converting of BKP to its CoA ester: BKP-CoA in peripheral tissues by 3-oxoacidCoA transferase. The last step is the cleavage of BKP-CoA to acetyl-CoA and
propionyl-CoA for energy utilization.
1.6.2 The ketone body utilization in fetal liver
The normal adult liver is traditionally considered as a ketone body producer but
not a ketone utilizer because of the absence of 3-oxoacid-CoA transferase (120).
A study reported a low activity of 3-oxoacid-CoA transferase in normal rat liver
(23). This report was disputed based on labeling patterns of ketone bodies in
liver which are incompatible with the presence of 3-oxoacid transferase (121).
However, the fetal liver uses ketone bodies for lipogenesis. The increased insulin
dephosphorylates ACC via a phosphatase and increase the lipids synthesis in
fetal liver. The mammalian fetus uses glucose supplied by the mother as the
main energy substrate (122). During the late pregnancy period, the mother has
increased ketone body levels (123) and pregnant rats develop hyperketonemia
20
very fast (124). The increased ketone bodies are transported from mother to
fetus; the fetal liver acts like a peripherial tissue as reflected by the increased
activity of ketone-utilizing enzymes (125). Yokoo et al. (126) found that ketone
bodies are oxidized in isolated hepatocytes from fetal liver, and that BHB is
utilized faster than AcAc.
1.6.3 The ketone body utilization in peripheral tissues
The ketone body utilization in brain (developing and adult)
The brain and nervous tissues use glucose as the primary energy substrate for
energy production (127; 128). A shift in energy substrate utilization (129) (129;
129)from glucose to ketone bodies occurs under certain nutritional conditions
during fasting, feeding a high-fat diet (130) and early development and
throughout the suckling period (131). Ketone bodies have become a potential
therapeutic agent for the treatment of seizure disorders, Alzheimer’s and
Parkinson’s diseases, which are caused by the malfunctions of brains (132).
As monocarboxylates, the uptake of ketone bodies by brain is highly dependent
on their transport across the blood brain barrier (BBB), and thus any therapeutic
strategy involving the use of KB’s should include a component for conditions
when KB uptake by brain is maximal. The first step to transport ketone bodies is
via monocarboxylase transporter (MCT-1), which is a primary carrier-mediated
transporter and locates in the endothelial cell membranes of BBB (133). The
exact location of this transporter in rat brain endothelium and glia has been
identified by Gerhart et al. by using immunoelectron microscopy technique (134).
MCT-1 is regulated in brain by nutritional status such as with feeding the
21
ketogenic diet (135). It has been shown that MCT-1 significantly increased in the
brain of the rats treated with ketogenic diet compared to the rats on a regular
chow (135; 136).
“NOTES Summary on the use of diet or any conditions such as with infusion, the
transport capacity is limiting must be considered before the treatment can be
efficient ore effective.” Tracer conditions are not an issue but transport capacity
is.
The mature brain uses glucose as the main energy substrate in the fed state.
The circulating ketone body concentration is a main regulator of ketone body
utilization in the brain (137). The decreased permeability to BHB of the adult
brain shows that compared to the suckling rat brain, the mature brain relies more
on glucose than on ketone bodies (138) . Ketone bodies in adult brains are
mostly used for oxidation to produce energy and not for the syntheses of lipids
(139; 140).
The developing rat brains are different from adult brains. After birth, circulating
ketone bodies in the suckling pups reach between 1-2 mM (hyperketonemic
condition) (141) because of the high fat content of maternal milk (142). The
capacity of the ketone body transporter is high during the suckling period (138).
In the developing rat brain, ketone bodies are lipogenic. In suckling rat brains,
ketone bodies are also important precursors for the synthesis of cholesterol,
which is very important for the brain growth (143). AcAc is rapidly incorporated
into brain cholesterol during the first week of life (25). Ketone bodies are also
22
used to synthesize other types of lipids such as phospholipids and sphingolipids,
which are required for brain development (144).
The ketone body utilization in the lactating mammary gland
During lactation, substrates such as glucose, lactate, pyruvate (145) and ketone
bodies are used for lipogenesis in the lactating mammary gland. In ruminants,
over 90% of milk fat consists of fatty acids or triglycerides. Humans’ breastmilk
have even more fat than the rumints. One of the sources of the fatty acids is the
de novo synthesis from acetate and BHB. The latter is produced by the rumen
epithelial from absorbed butyrate (146).
The ketone-body utilization enzymes show high activities in the lactating
mammary gland of fed lactating rats (147). Incubation experiments with ketone
bodies and [3H]water showed that ketone bodies increase the fatty acid synthesis
rate in slices of mammary gland of lactating rats (148). The incorporation of [314
C]BHB into lipid in the mammary gland was 5-fold higher than in the liver.
Insulin increased this lipid incorporation (29). Given that the oxidation of [314
C]AcAc accounted for 65% of the net ketone-body utilization in vitro (147), the
mammary gland is regarded as the major tissue of ketone-body utilization in fed
lactating rats. However in starved lactating rats, ketone-bodies are not used as
an alternative substrate to glucose (169).
The ketone body utilization in kidney and heart
The
kidney is
an
important
user of
ketone
bodies.
Experiments
in
nephrectomized rats showed that the kidney removes about 30% of total body
ketone bodies (149). During prolonged starvation, obese subjects have increased
23
renal reabsorption rate of AcAc and BHB. This minimizes ammonium loss
through urine and prevents the breakdown of body proteins(150).
The heart uses glucose and fatty acids as main energy substrates in the fed
state. However, during starvation, this organ has the highest ketone body
utilization rate (2). The uptake of plasma ketone bodies is proportional to their
circulating concentrations (151). The perfused rat heart completely oxidizes
ketone bodies. This process can account for up to 75% of the organ’s oxygen
consumption (152).
1.7 The role of ketone bodies in mammalian metabolism
Ketone bodies were originally considered to be harmful because of their
association with uncontrolled diabetes. The ketone bodies accumulate up to 25
mM in diabetic patients’ plasma (1). It causes the changed body pH, disrupted
electrolyts balance and cardiac arrest. On the other hand, this wide range of
concentrations shows that ketone body concentration is a sensitive physiological
index of metabolic changes. Ketone bodies are alternative fuels during fasting
and hypoglycemic conditions (12). In peripherial tissues, especially in the
nervous system, ketone bodies decrease glucose utilization and inhibit
proteolysis (153). They are important lipogenic precursors in fetal liver, lactating
mammary gland and developing brain (2).
Nowadays, ketone bodies are not considered only as energy substrates in
peripherial tissues. Ketone bodies are involved in whole-body metabolism
because they are the signal of lack of carbohydrates. They may also play
important role in the regulation of proteolysis (2). Although there is no direct
24
evidence to show the inhibition of ketone body on protein breakdown, the use of
ketone bodies spares the muscle proteins is a good physiological consideration.
1.8 The measurement of ketone body turnover
1.8.1 The measurement of ketone body turnover with isotopic methods
Overview of measurement of ketone bodies
Ketone body turnover is generally assessed by two methods in vivo: hepatic
arteriovenous difference (154) and isotopic method (155). Compared to the
arteriovenous difference technique, isotopic-dilution technique is more applicable
because it is noninvasive. Thus, this technique is suitable for humans. The
isotopic method is used to measure the steady-state ketone body turnover rate
(155). Two isotopic methods have been applied to trace ketone body turnover:
single-isotope and double (dual)-isotope techniques. The general principle is to
infuse radioactive or stable-isotope labeled ketone bodies into a vein, and then
measure the specific activity or labeling of ketone bodies in the arterial plasma.
The dilution of specific activity reflects the rate of hepatic ketogenesis. Different
isotope labeled tracers including 2H-, 3H-,
13
C-,14C-ketone bodies have been
used (7; 156; 157).
Single-Isotopic method
The single-isotope technique used either [14C]AcAc or [14C]BHB to measure
ketone body production or utilization rate, using the “total ketone body specific
activity” describled by McGarry et al. (158). However, the disequilibrium of
labeling between the two ketone bodies was not taken into account by the
calculation. Although AcAc and BHB are interconverted very fast, the specific
25
radioactivity of the two substrates takes a long time to reach equilibrium (159).
Despite the above problem, ketone body kinetics has been extensively used in
animals and humans. In diabetic dogs, total ketone body production rate is 38.1 ±
5.6 μmol/kg•min, which is much higher than that of the control group: 11.4 ± 1.9
μmol/kg•min (155). Balasse (190) showed that in obese fasting patients, infusion
of [14C]AcAc or [14C] BHB results in similar production rate of total ketone bodies:
1908 ± 80 μmol/min. This result is almost double as that from Wolfe’s group
(160).
Double-isotopic method
A double-isotope (dual-isotope) technique was also used to evaluate ketogenesis
in vivo. The general equation used for the measurement is following (161):
Ra = [I- (C x V x dSR / dt)] / SR
Ra: is the rate of appearance for the individual ketone body; I: is the tracer
infusion rate; C: is the concentration; V: is the assumed volume distribution (35%
of body weight); SR: is the specific radioactivity.
Reed et al. (161) used [4-3H] and
14
C-labeled ketone bodies to study the kinetics
of ketone-bodies in the rat. They simultaneously infused [3-14C]AcAc and [43
H]BHB into rats. The two different tracers yielded similar kinetics of ketone
bodies. Miles et al. (162) compared single-isotope and dual isotope techniques.
When they infused [14C]BHB alone or simultaneously infused [13C]AcAc and
[14C]BHB in mongrel dogs, basal ketone body turnover were similar: 2.2 ± 0.2
μmol/kg•min and 2.7 ± 0.2 μmol/kg•min, respectively. But infusion of [13C]AcAc
26
overestimated the turnover by 55%. This shows that the choice of a single tracer
affects the measurement.
1.8.2 Pseudoketogenesis
Isotopic artifacts may overestimate ketone body turnover. When labeled AcAc or
BHB is infused and enters extrahepatic tissues such as heart or muscle (163).
AcAc is activated to AcAc-CoA via 3-oxoacid-CoA transferase. Unlabeled
glucose and fatty acids produce cold AcAc because of the reversal of the
reaction catalyzed by 3-oxoacid-CoA transferase. This results in the decrease of
the specific activity of plasma ketone bodies. This dilution is caused only by
labeling exchange via reversal of 3-oxoacid-CoA transferase in extrahepatic
tissues and not by net ketogenesis. This process is called pseudoketogenesis
because it results from an isotopic exchange (see figure 1.4). The minimal
pseudoketogenesis ranges from 19-32% of the uptake of ketone body
metabolism in heart (163). Pseudoketogenesis occurrs in the isolated working rat
heart (164) and in the hepatectomized dog (163).
27
Figure 1.4: Pseudoketogensis in extrahepatic tissues. Modified from Des Rosiers
et al (165). Am J Physiol 258: E519-E528, 1990. (1) 3-oxoacid-CoA transferase;
(2) acetoactyl-CoA thiolase; (3) 3-hydroxybutyrate dehydrogenase.
28
CHAPTER 2: ANAPLEROSIS
2.1 Overview
Anaplerosis is a process that refills the catalytic intermediates of the CAC. The
CAC involves eight reactions which oxidize acetyl groups to carbon dioxide and
regenerate the acceptor of the acetyl groups. The total pool size of the eight
intermediates is small and the individual intermediates have very different sizes
(See Figure 2.1) (166). The moderate leakage of intermediates through the
mitochondrial membrane and the cell membrane is a physiological process. The
leakage of intermediates from the CAC is called cataplerosis. Anaplerosis and
cataplerosis are two reciprocal processes working together to balance the CAC
flux and sustain cell homeostasis (167).
2.2 The cataplerosis
Cataplerosis is a component of gluconeogenesis and glyceroneogenesis.
Cataplerosis via phosphoenolpyruvate carboxykinase (PEPCK) contributes to
gluconeogenesis in liver and kidney (167). During starvation, the liver and kidney
use PEP for gluconeogenesis. PEP is derived from the CAC intermediate
oxalacetate formed by pyruvate carboxylase. In adipose tissue, OAA and PEP
are used for glyceroneogenesis (168). In normal healthy subjects, cataplerosis is
well balanced by anaplerosis. During metabolic transitions, the CAC intermediate
concentrations vary in response to these changes. Afterward, anaplerotic influx
be balanced by cataplerotic outflux (169). Some pathological conditions lead to
excessive loss of CAC intermediates. One typical example is FOD. Patients with
FOD have deficiency of long-chain or very long-chain acyl-CoA dehydrogenase.
29
Figure 2.1: Main anaplerotic processes feeding into citric acid cycle. Modified
from Brunengraber H et al .(166): J Inherit Metab Dis (2006) 29: 327-331.
30
This causes the accumulation of toxic long-chain or very long-chain acyl-CoAs.
As a result, the mitochondrial membrane is damaged and large amounts of CAC
intermediates leak out (63).
2.3 The anaplerosis
2.3.1 The significance of anaplerosis
Anaplerosis plays an important role in mammalian tissues. In the brain,
transmitter glutamate, γ-aminobutyric acid (GABA) in neurons and glutamine
from glia are all derived from α-ketoglutarate. The loss of α-ketoglutarate from
the CAC would decrease the energy production by the cells if the cataplerosis
was not balanced by anaplerosis (170). The heart is another organ where
anaplerosis is especially important. The heart can oxidize almost all energy
substrates such as glucose, fatty acids, ketone bodies, acetate and even amino
acids under special conditions. This is because the heart needs to work non-stop
(171).
When rat hearts are perfused only with C4-ketone bodies, heart
contraction is severely impaired. The decrease is reversed by addition of an
anaplerotic substrate such as pyruvate (172). The most effective anaplerotic
substrates for the heart are pyruvate and propionate (173). Anaplerosis
contributes to the maintenance of energy production. It also works as a signal to
regulate insulin release in pancreatic β-cells. The increase of anaplerosis :
cataplerosis ratio in pancreas raises insulin release by insulinoma cells (174).
The increased blood glucose leads to the increased CAC intermediates and thus
increase β-cell ATP:ADP ratio, which triggers insulin release (175). It has been
shown that pyruvate carboxylase activity, an anaplerotic enzyme, tightly coupled
31
with glucose-induced insulin secretion. Anaplerosis is especially important in
many pathological conditions related to energy metabolism reviewed below.
2.3.2 Anaplerotic substrates
Overview
As mentioned before, the loss of CAC intermediates must be replenished via
anaplerotic processes. Figure 2.1 demonstrates that three main anaplerotic
substrates enter the CAC: pyruvate, glutamine/glutamate and propionyl-CoA.
Pyruvate is the most studied anaplerotic substrate. It enters the CAC via OAA or
malate formed by pyruvate carboxylase or malic enzyme, respectively.
Glutamate and glutamine are the precursors of α-ketoglutarate. They enter the
CAC via glutamate dehydrogenase and/or aminotransferases (166). The third
important entrance point is via the CAC intermediate succinyl-CoA. PropionylCoA is converted to methylmalonyl-CoA, which is continuously converted to
succinyl-CoA. A number of precursors of propionyl-CoA enter the CAC through
this pathway, such as odd-chain fatty acids, propionylcarnitine and C5-ketone
bodies (166).
Pyruvate
Two enzymes are responsible for the anaplerosis from pyruvate: pyruvate
carboxylase (PC) and malic enzyme. PC was first identified as a gluconeogenic
enzyme in liver and kidney in 1959 (176). Later, this enzyme was found in nongluconeogenic tissues such as adipose tissue (177), brain (178) and pancreatic
islets (179). PC catalyzes the conversion of pyruvate to oxaloacetate in the
mitochondria, and serves as an anaplerotic enzyme. It has been reported that
32
50% of pyruvate is carboxylated to form OAA in the β-cells (180). Pyruvate also
enters the CAC as malate via malic enzyme. Inhibition of malic enzyme by
hydroxymalonate decreased the incorporation of
14
C from [1-14C]pyruvate into
malate and impaired heart contraction (181). Malic enzyme activity increases in
hypertrophied rat heart (182). In contrast, PC activity does not change compared
to control. This shows that in this model, malic enzyme is more anaplerotic than
PC.
The physiological concentration of pyruvate is 0.1 mM to 0.2 mM (183). The
conversion of pyruvate to OAA is important for cell proliferation and repair (184).
High concentrations of pyruvate increased myocardium function in isolated
guinea-pigs heart (185). The Infusion of dipyruvyl-acetyl-glyerol for 2 hr in
anesthetized pigs decreased the coronary infarct size after temporary coronary
clamping (186). In PC knockdown β-cells, the incorporations of
14
C from [U-
14
C]glucose into lipids and the acid pellet were decreased 20 – 30% compared to
control cells within 45-min incubation. This is because deficiency of PC reduced
the conversion of pyruvate to OAA. Thus less OAA was transported from
mitochondria to cytosol for further biosynthesis (180). Isolated rat pancreatic
islets perifused with buffer containing phenylacetate, an inhibitor of PC, showed a
decrease in insulin secretion. This resulted from the disruption of pyruvate
cycling (187).
Glutamine/glutamate
The glutamate/glutamine couple is an effective anaplerotic system. Glutamate
enters the CAC through α-ketoglutarate formed by glutamate dehydrogenase and
33
glutamate aminotransferase. Glutamine is first converted to glutamate via
glutaminase and then enters the CAC. The onset of exercise causes the increase
of CAC intermediate concentrations up to 4 fold in skeletal muscle (248). This
happens mainly via anaplerotic glutamate. Muscle glutamate content decreases
about 60% (188) and glutamate provides anaplerotic carbon at the level of αketoglutarate (189). Subjects who ingested glutamine before exercise increased
CAC intermediates pool size (190). In contrast, ingestion of ornithine αketoglutarate did not expand the CAC pool size. After ischemia, the perfused
working rat heart is rescued by glutamine, but not by glutamate and αketoglutarate. If the exogenous glutamine concentration reaches 2.5 mM, the
heart goes to complete post-ischaemic cardiac functional recovery (106). This is
because the transport of gluctamine into cells is much efficient than the transport
of gluctamate. As a result, glutamine is a better anaplerotic substrate than
glutamate in the heart (191). However, it is difficult to evaluate the anaplerotic
efficiency of the two substrates by stable isotopic techniques because of their
interconversion.
Precursors of propionyl-CoA
Anaplerotic propionyl-CoA is very effective even at a low concentration. This is
because of the irreversible conversion of propionyl-CoA to succinyl-CoA. Oddchain fatty acids, C5-ketone bodies and branched-chain amino acids are all
precursors of propionyl-CoA. These anaplerotic substrates enter the CAC via the
same pathway. These substrates are metabolized to propionyl-CoA, which is
34
continuously converted to succinyl-CoA (a CAC intermediate) by propionyl-CoA
carboxylase, methylmalonyl-CoA racemase and methylmalonyl-CoA mutase.
The circulating propionate concentration is only 0.05 mM (192). This low
concentration results from the rapid uptake of 99% of the gut propionate by the
liver (192). In perfused rat heart, 0 – 2 mM [13C3]propionate increased total heart
anaplerosis (192). The only fate of propionate in the perfused heart is
anaplerosis. This showed the high efficiency of this anaplerotic substrate. C5ketone bodies are effective anaplerotic substrates in heart, brain and kidney
(157). C5-ketone bodies are rapidly metabolized and produce propionyl-CoA.
Finally, the oxidation of heptanoate, a medium-chain fatty acid, produces
propionyl-CoA as well as acetyl-CoA. The important therapeutic role of
triheptanoin will be discussed in section 2.3.4.
2.3.3 The measurement of anaplerosis
The contribution of anaplerotic substrates to the CAC intermediates has been
calculated by isotopic techniques combined with mass spectrometry and NMR
techniques (193). There are two related ways to quantify anaplerosis: relative
anaplerosis and absolute anaplerosis.
Relative anaplerosis
Relative anaplerosis represents the contribution of anaplerotic substrates to the
turnover of CAC intermediates. It is calculated by two ways: (i) the ratio of
labeling of a CAC intermediate versus the labeling of the anaplerotic substrate, or
(ii) the uptake of anaplerotic substrate divided by the rate of CAC flux, which is
related to oxygen comsumption (194). When [13C3]propionate is infused in
35
anesthetized pigs for 1 hr, the heart anaplerotic flux ratio is 8.9 % calculated by
the enrichment ratio (M3 succinate) / ( M3 proionate) (194). The use of the M3
isotopomer is because M3 succinate derives only from M3 propionate. When
Kasumov et al. (192) perfused rat hearts with [3-13C]propionate instead of
[13C3]propionate, the apparent relative anaplerotic ratio was 3-fold higher in the
[3-13C]propionate perfusion ((M1 succinyl-CoA) / (M1 propionyl-CoA)) than that in
the [13C3]propionate perfusion ((M3 succinyl-CoA) / (M3 propioyl-CoA)). This
overestimation resulted from the contribution of M1 succinyl-CoA via αketoglutarate dehydrogenase. If one used [13C3]pyruvate as the anaplerotic
substrate,
13
C labeled citrate shows the highest M3 labeling of CAC
intermediates. As a result, the ratio (M3 OAA moiety of citrate) / (M3 pyruvate) is
also used to evaluate relative anaplerosis from pyruvate(195).
Absolute anaplerosis
When the measurement of relative anaplerosis is coupled to the measurement of
metabolic CAC activity (oxygen consumption), one calculates absolute
anaplerosis expressed as μmol•g-1•min-1 (194). In pigs infused with [13C]pyruvate,
the absolute rates of pyruvate carboxylation in the heart is 1.6 μmol•g-1•min-1; this
accounts for about 6% of the total turnover of CAC intermediates.
2.3.4 The anaplerotic diet therapy
FODs are inherited metabolic disorders associated with energy deprivation. FOD
can effect the carnitine cycle as well as the mitochondrial β-oxidation spiral.
Patients with FOD often suffer from muscle weakness, hyponia and cardiac
problems. Acute stress such as enteritis triggers the sudden shock or death. The
36
most common disorder affects very long-chain acyl-CoA dehydrogenase (63).
Since the 1980’s, the classical dietary treatment is to use medium-eventriglycerides (trioctanoin) as part of the energy supply, and to prevent the
ingestion of long-chain triglycerides, thus avoiding the accumulation of toxic longchain acyl-CoAs (196). In 2002, Roe et al (106) replaced trioctanoin with oddmedium-chain triheptanoin. As reviewed before, heptanoate, the fatty acid
component of triheptanoin, is oxidized to produce acetyl-CoA as well as the
anaplerotic and gluconeogenic precursor propionyl-CoA.
When the patients’
dietary treatments were switched from octanoin to triheptanoin, there was
remarkable improvement in muscle strength, endurance and activity.
Pyruvate carboxylase (PC) is a key enzyme of gluconeogenesis in liver and
kidney. PC deficiency is a very rare and severe autosomal recessive disease
characterized with impaired gluconeogenesis and lactic acidosis (197). A six-dayold girl was diagnosed as biotin-unresponsive pyruvate carboxylase deficiency
type B. When she was given triheptanoin from day 7, her major hepatic failure
was reversed within 48 hr. At 3 months of age, her psychomotor development,
EEG and MRI reached to the normal level. At the age of 6 months, ageappropriate myelination had developed (197).
37
CHAPTER 3: THE SYNTHESES OF ADENINE NUCLEOTIDES, COENZYME A
(CoA) AND DEOXYRIBONUCLEIC ACID (DNA) IN RAT LIVER
3.1 Overview of the biosynthesis pathways of adenine nucleotides, CoA
and DNA
3.1.1 Overview
The biosynthesis pathways of adenine nucleotides, CoA and DNA are
interrelated via the synthesis of adenine nucleotides (See figure 3.1). The
synthesis of adenine nucleotides involves the de novo synthesis pathway and the
purine salvage pathways. The two processes are demonstrated in figure 3.1.
Both processes use phosphoribosylpyrophosphate (PRPP) as the key precursor.
PRPP is formed via the phosphorylation of ribose-5-phosphate, which is mainly
derived from glucose-6-phosphate produced by glycolysis, gluconeogenesis
(GNG) and glycogen breakdown. The cycling of pentose groups between RNA,
ribose, ribose-1-phosphate and ribose-5-phosphate also partially contributes to
the formation of ribose-5-phosphate (198). The synthesis of CoA is related to the
adenine nucleotide synthesis pathway via ATP. ATP contributes to the formation
of CoA through two ways: phosphorylation of CoA precursors, and the transfer of
the adenine-ribose nucleus to dephospho-CoA. The DNA synthesis pathway is
shown in figure 3.1.
3.1.2 The biosynthesis pathway of adenine nucleotides
The de novo synthesis pathway
38
Figure 3.1: Biosynthesis pathways of adenine nucleotides, CoA and DNA.
39
The purine de novo synthesis pathway appears almost identical in all living
organisms. The pathway was discovered by Buchanan et al. in the 1950’s (199).
The de novo synthesis pathway begins with non-purine precursors: amino acids,
ribose-5-phosphate, carbon dioxide and ammonium. The rate-limiting step is the
formation of PRPP by PRPP amidotransferase. After ten sequential steps,
inosinate, the first intermediate with a complete purine ring, is formed. The
conversion of inosinate to AMP requires aspartate and GTP (200). The
incorporation of radioactive glycine into purine is most frequently used to
measure the rate of purine de novo synthesis (201). Tullson et al. (202) used [114
C]glycine as a tracer to study the rate of de novo synthesis of adenine
nucleotides in isolated rat hindquarters. Rates of de novo synthesis were highest
in fast-twitch red muscles (57 nmol / h x g) and lowest in fast-twitch white
muscles (26 nmol / h x g). The rate in slow-twitch muscles was between those of
the two types of fast-twitch muscles. The de novo synthesis of adenine
nucleotides accounts for a small fraction of adenine nucleotide turnover in rat
heart (203). The turnover rate of the ATP nucleus is 0.08 nmol / min x g in rat
heart (204). This rate increases 10-fold by administration of ribose (205). This is
because ribokinase and ribose-phosphate pyrophosphokinase increase the
availability of PRPP for adenine nucleotide synthesis.
Not all the tissues can synthesize purines de novo. Bone marrow (206),
leukocytes (207) and blood cells (208) are not able to synthesize purines de
novo. In these cells, the purine salvage pathway is the only source of purines.
40
Abnormal high rates of purine de novo synthesis result in gout. This disease of
joints is clinically expressed as acute arthritis. The deposition of crystals of
sodium urate in the joints leads to inflammation and pain. The precise cause of
gout is still not identified, but it results partly from purine overproduction via the
de novo synthesis (209).
The purine salvage pathways
Compared to the de novo synthesis pathway, the purine salvage pathways are
much simpler. The salvage pathways recycle free bases (hypoxanthine, guanine
and adenine) and nucleosides (adenosine and guanosine) from nucleic acid
breakdown (210). About 90% of free purines are recycled in liver (211). The
regulation of this recycling is related to the production of ribose-5-phosphate and
PRPP (211). The salvage pathway is also very important for the de novo
pathway. This is because four enzymatic steps in the de novo pathway use ATP
(211). The purine salvage pathways are especially important for those tissues
that are not able to synthesize purines de novo such as bone marrow (206),
leukocytes (207) and blood cells (208).
The liver is the major organ that salvages purines (212). The salvage pathways
also regulate the heart nucleotide pools (210). The rate of purine salvage in heart
is evaluated at 0.2 nmol / mg wt x min (213). The most important salvage
enzyme is hypoxanthine/guanine phosphoribosyltransferse. The total quantity of
purine salvaged by this enzyme was estimated by measuring the amount of uric
acid excreted in the urine. A healthy adult on a purine-free diet excretes around
400 mg uric acid per day. This corresponds to 6 mg purine / kg body weight per
41
day (214). A healthy child produces about 10 mg purine / kg body weight per day
calculated from the average uric acid excretion (215). The difference in rates of
purine synthesis between adults and children shows the high needs of purine for
child growth.
Lesch-Nyhan syndrome is a disease caused by a defect in purine salvage
pathways. This sex-linked recessive disorder results from the lack of
hypoxanthine/guanine phosphoribosyltransferase. This enzyme which salvages
hypoxanthine is highly expressed in brain (216). This is why patients with LeschNyhan syndrome show major neurological symptoms.
3.1.3 The CoA biosynthesis pathway
General CoA synthesis pathway
CoA was first discovered by Lipmann in 1946 (217). This heat-stable cofactor
was named for its function in acetylation. CoA plays an important role as an acylgroup carrier and carboxyl-activating group in more than 100 biochemical
reactions. About 4% of enzymes use CoA as an cofactor (218). The biosynthesis
of CoA is a five - step pathway ubiquitous in bacteria, plants and mammals (see
figure 3.2). CoA synthesis starts with pantothenate, also called vitamin B5, which
cannot be de novo synthesized in mammalian cells. The only source of
pantothenate for mammals is the diet (218). Some bacteria can de novo
synthesize pantothenate from β-alanine and α-ketoisovalerate (219) . The first
and rate-limiting step of CoA synthesis is the phosphorylation of pantothenate to
4’-phosphopantothenate by pantothenate kinase. After 4’-phosphopantothenate
reacts with cysteine, a decarboxylation step forms 4’-phosphopantetheine. Next,
42
Figure 3.2: Pathway of Coenzyme A synthesis. Modified from Tahiliani AG et al.
(220) J Mol Cell Cardiol 1987 (19): 1161-1167.
43
ATP transfers its AMP moiety to 4’-phosphopantetheine to produce dephosphoCoA, which is finally phosphorylated to CoA by ATP (218). Labeled CoA was
synthesized
by
incubating
murine
hepatocytes
in
media
containing
[13C3,15N1]pantothenate. The cellular uptake of labeled pantothenate from the
media initiates the formation of [13C3,15N1]CoA (221).
Regulation of CoA contents in tissues
CoA is unequally distributed between cell cytosol and mitochondria (220). The
cytosolic concentration of CoA is between 0.02 mM to 0.14 mM in animal tissues.
Mitochondrial CoA has a much higher concentration ranging from 2 mM to 5 mM
(218). The hepatic CoA content changes in response to extracellular signals.
Some hormones regulate the CoA contents. In rats fasted for 48 hr, the decrease
in insulin and the increase in glucagon raise the total hepatic CoA contents from
544 to 870 nmol / g dry wt (222). The hepatic CoA content in type I diabetic rat
liver is twice that of controls, possibly because of the depletion of insulin. Feeding
a high fat diet to the rats also increases hepatic CoA content (223). The heart
CoA content is lower than that of liver. Myocardial CoA total content (free +
esterified) is normally 500 nmol / g dry wt (3). The isolated perfused heart from
diabetes rats has high level of total CoA (750 nmol / g dry wt). The addition of
insulin to the perfusate significantly decreased the CoA contents to 670 nmol / g
dry wt (224).
CoA synthesis as an antimicrobial drug target
The treatment of infectious diseases faces increasing challenges. The selection
of more and more drug-resistant pathogens leads to treatment failure (219). The
44
characteristics of the CoA biosynthesis pathway make it a potential antimicrobial
drug target. However, the disruption of genes encoding the enzymes in the
pathway of CoA synthesis results in lethal phenotypes (219). The inhibition of the
rate-limiting
utilization
of
pantothenate
stops
CoA
synthesis
in
some
microorganisms. Pantothenate analogues have been synthesized as antibacterial
and antifungal agents. Pantoyltaurine was the first synthesized pantothenate
analogue (225). This sulphonic acid analogue inhibits bacteria growth both in
vitro (225) and in vivo (226). Pantoyltaurine works for those bacteria that require
exogenous pantothenate. However, it has no effect on bacteria that can
synthesize pantothenate de novo. More and more potential inhibitors of
pantothenate utilization and biosynthesis have been discovered (for an extensive
review, see (219)).
3.1.4 The DNA biosynthesis pathway
The true precursors of DNA synthesis are four deoxyribonucleotide-triphosphates
(dNTPs). They are dATP, dGTP, dCTP and dTTP (227). The proper synthesis of
DNA requires the balanced supply of the four dNTPs in mammalian cells (see
figure 3.1).
The de novo synthesis pathway
The de novo synthesis pathway is irreversible. Four ribonucleotides (ADP, GDP,
CDP and TDP) are first synthesized via the nucleotide synthesis pathway. Then
three deoxyribonucleotides, i.e. dADP, dCDP and dGDP are formed from ADP,
CDP and GDP respectively via ribonucleotide reductase. This enzyme is the first
de novo synthetic enzyme that is allosterically regulated. The formation of dTDP ,
45
besides the use of ribonucleotide reductase, also uses another allosterically
controlled enzyme dCMP deaminase, followed by dTMP synthase and
thymidylate synthase (227). The phosphorylation of the four dNDPs by
nucleoside diphosphate kinase produces dNTPs for DNA synthesis.
The DNA salvage pathway
The cell reuses ribonucleosides and deoxyribonucleosides released from the
degradation of DNA. This process is called the salvage pathway. This pathway
helps to maintain dNTP levels by regulating the flux of ribonucleosides and
deoxyribonucleosides to the dNTP pools (227).
Significance of the dNTP pools
The deoxyribonucleotide-triphosphate pools are small (228). They reach their
largest sizes during the S phase of the cell cycle (229). The dGTP pool has the
smallest size among the four dNTPs (230). The size of the pool is balanced by
DNA synthesis and degradation. Ultraviolet rays (231) and mutagens (232)
disturb dNTP pools and thus impair DNA synthesis. The depletion of dNTP pools
causes severe consequences. dTTP deprivation, also called thymineless death,
first reported in bacteria, is also found in mammalian cells (233). The depletion of
dTTP leads to the breakage of DNA strands and to cell death (234). Inherited
deficiencies in adenosine deaminase and purine nucleoside phosphorylase lead
to accumulation of dATP and dGTP in the immune system. The accumulation of
dATP and dGTP interferes with ribonucleotide reductase in lymphocytes and
thus causes DNA damage and cell death (235).
3.2 The techniques used for tracing the biosynthesis pathways
46
3.2.1 The techniques used for tracing ATP synthesis from ADP
ATP is a main energy source for biological reactions. Thus the measurement of
the rate of ATP synthesis (i.e. of ADP phosphorylation) is generally related to the
mitochondrial energy generating system. This rate only reflects the oxidative
phosphorylation and not the turnover of the adenine-ribose nucleus. The
production of ATP and the ratio of [ATP]/[ADP] are the biomarkers to evaluate
the mitochondrial energy generating system capacity (236). The dynamic of
cellular ATP synthetic activity is quantitatively assessed by luciferin-luciferase
reactions (237). The oxidative synthesis of ATP follows fast kinetics evaluated by
stop-flow rapid – mixing experiments in rat liver mitochondria (238) . After the
supply of oxygen to rat liver mitochondria, the rapid oxidation of NADH via the
respiratory chain initiates the synthesis of ATP within 20 ms. This synthetic
reaction is half-completed within 100 ms. The liver and brain ATP production in
isolated mitochondria of Fisher - 344 rats are not affected by age and caloric
restriction (239).
The basal rate of ATP synthesis from ADP is 180 μmol / g muscle wt / min in
human skeletal muscles (240). In human muscle, the aerobic ATP synthesis rate
reaches to the maximum after 9 seconds of maximal exercise. This was
evaluated by Walter et al. (241) from the resynthesis rate of phosphocreatine.
However, in perfused pig intercostal muscles, anaerobic ATP turnover rate is 6.1
μmol / min•g during the first minute of tonic stimulation of the nerve. In contrast,
the anaerobic ATP synthesis rate drops to 0.4 μmol / min•g in unperfused
muscles (242).
47
3.2.2 The techniques used for tracing CoA synthesis
The methods used to trace CoA synthesis have been developed using labeled or
unlabeled synthetic intermediates. [14C]Pantothenate is the most used tracer to
evaluate CoA synthesis in vivo and in vitro. Reibel et al. (222) injected
[14C]pantothenate into control, fasting and diabetic rats. After 90 min, they
measured pantothenate incorporation into CoA from rat organs. Another strategy
is to perfuse an isolated organ such as the heart with [14C]pantothenate (243).
The calculation of CoA synthesis rate is based on two parameters measured by
liquid scintillation counting: (i) the amount of [14C]pantothenate incorporated into
CoA, and (ii) the average tissue [14C]pantothenate specific activity (244). Isolated
control hearts show high CoA synthesis rates measured with [14C]pantothenate.
The addition of exogenous energy substrates (such as glucose and fatty acids) to
the
perfusate
decreases
CoA
synthesis
(245).
The
incorporation
of
[14C]pantothenate into CoA is almost undetectable in the perfused heart from
type I diabetics rats (246). 4’-Phosphopantetheine is another CoA precursor used
to evaluate the CoA synthesis rate in subcellular preparations. Tahiliani et al.
(220) incubated isolated mitochondria from rat hearts with 4’-phosphopantetheine
in buffer containing ATP. The apparent Vmax for the CoA synthesis was 0.02 nmol
/ mg protein x min. The increase of pH from 7.4 to 8.5 causes an increase in
synthesis.
3.2.3 The techniques used for tracing DNA synthesis
Classical techniques for tracing the rate of DNA synthesis use thymidine analogs
(247). Thymidine analogs become incorporated into DNA during the cell division,
48
and have been used to measure cell proliferation rates. Two commonly used
thymidine analogs are [3H]thymidine and bromodeoxyuridine (BrdU: 5-bromo-2’deoxyuridine). [3H]Thymidine incorporated into DNA strand is detected by
autoradiographic techniques (248). This process normally takes several weeks of
contact with the film. BrdU is a halopyrimidine. The incorporation of BrdU into
DNA is detected by immunohistochemistry using a monoclonal antibody against
a DNA strand containing BrdU (249). BrdU is delivered to an animal by
intraperitoneal or intravenous injection (247). Multiple injections are required to
label the total proliferating population of cells over several days or weeks. The
i.p. dose used for the measurement of rodent adult neurogenesis is 50 – 100
mg/kg; 200 mg/kg is the maximal dose. The animal is sacrificed 1-3 hr after
injection of BrdU (247). A high dose of BrdU triggers neurocyte death (250).
Compared to [3H]thymidine, the BrdU technique is faster and nonradioactive.
However, both precursors are toxic and mutagenic. They impair cell division and
kill susceptible cells (251; 252). The addition of a bromine atom into the DNA
molecule changes the stability of DNA-double strands resulting in chain breakage
and mutations. As a result, these techniques cannot be safely used in humans.
[2H]Water has been extensively used to trace DNA synthesis. This will be
reviewed in section 3.3.
3.3 Tracing the syntheses of biopolymers with [2H]water
3.3.1 Overview
Since Von Hevest (253) first used a radioactive isotope for biological
investigation in 1926, more and more isotopes of C, H, N and O have been used
49
in metabolic studies. In the 1930’s, deuterium was discovered and began to be
used to study metabolic processes (253). Stable isotopes do not decay and can
be used for tracing organic syntheses and metabolic processes. Over the past 30
years, the use of stable isotopes has extensively increased, driven by safety
concerns and the development of mass spectrometric and NMR techniques
(254).
3.3.2 Principle of the use of [2H]water to trace the biosyntheses of (pseudo)biopolymers
Overview
Biopolymers refer to compounds made of repeated building blocks. Pseudobiopolymers refer to compounds made of different building blocks. Fatty acids,
proteins and DNA are all considered as polymers. CoA and adenine nucleotides
are pseudo-biopolymers. The general principle of using [2H]water to evaluate the
syntheses of biopolymers is based on the incorporation of 2H from [2H]water into
the building blocks of biopolymers (254). During the synthesis of many
biopolymers, a number of H atoms in C-H bonds are derived from H2O. If water is
enriched with
2
H, the building blocks become
2
H-labeled before they are
incorporated into biopolymers. When the polymer is synthesized, the labeled
building blocks are assembled. The
2
H-enrichment of the building blocks,
measured by mass spectrometry, reflects the rate of biosynthesis of the polymer.
One assumption, proved by many experiments, is that once the polymer is
assembled, the labeling of the building blocks does not change (no new labeling
50
into the block; no loss of label from the block) (255). For an extensive review, see
(256).
Mechanism
of incorporation of [2H]water
into building blocks
of
biopolymers
The incorporation of [2H]water into building blocks of molecules occurs through
different mechanisms: hydration of double bonds (257), reduction of carbonyl
groups by NADH + H+ (258; 259), reduction of double bonds by NADPH + H+,
isotopic exchanges (260) (keto-enol tautomerism, reversible transaminations and
direct exchange C-H for C-2H) and splitting of C-C bond (261).
Hydration of double bonds
Water participates in biological reactions. When an enzyme catalyzes the
reaction of water with biomolecules containing double bonds, those molecules
are hydrated and form new C-H bonds. The hydrogen in the new C-H bonds
derives from water. When body water is
2
H-enriched, those biomolecules
become 2H-enriched because of the formation of C-2H bonds. In the CAC, the
conversion of cis-aconitate to isocitrate, and the conversion of fumarate to malate
introduce water into the products (257).
Reduction of carbonyl groups by NADH + H+
The NAD+ / NADH + H+ couple works as hydride-accepting and hydride-donating
(262). NAD+ dehydrogenases catalyze the interconversion of the members of the
redox couples. The two hydrogen atoms (position A and position B) on C-4 of
NADH are not equal. Two classes of NAD+ dehydrogenases are responsible for
the transfer of one of the two hydrogen atoms on C-4 of NADH to the substrates.
51
There
are
A-stereospecific
and
B-stereospecific
enzymes.
Alcohol
dehydrogenase is an A-stereospecific enzyme (258). It transfers the hydrogen at
position
A
from
NADH
to
acetaldehyde.
Glyceraldehyde
phosphate
dehydrogenase is a B-stereospecific enzyme (259). It transfers the hydrogen at
position B from NADH to glyceraldehyde 3-phosphate.
The NADH-dependent reduction reaction transfers two hydrogen atoms to the
product in the form of a hydride ion and a proton. The hydride ion is added
directly to the carbonyl group to form new C-H bond. The hydride ion of NADH
transferred to the substrate is either non-labeled or 2H-lableled. It depends on the
original labeling of the NADH. The proton directly becomes 2H labeled by
equilibrium with [2H]water solvent. The 2H+ only forms labile O-2H bond.
Reduction of double bonds by NADPH + H+
Like NAD+ dehydrogenases, NADP+ dehydrogenases also include A-specific and
B-specific enzymes. NADPH + H+ is generated in the conversion of glucose-6phosphate into ribulose-5-phosphate (263). The oxidation of malate to pyruvate
by malic enzyme also produces NADPH + H+. NADPH + H+ are used as the
reduced agents in biosynthetic reactions. In the fatty acid synthesis process,
NADPH transfers the hydrogen to the double bonds of the synthetic
intermediates and forms the saturated C-H bonds. If NADPH is 2H-labeled, the
newly synthesized fatty acids become 2H-labeled.
Isotopic exchanges
Isotopic exchanges occur during the incorporation of 2H into the molecules. A
typical isotopic exchange is via keto-enol tautomerism. This refers to a chemical
52
equilibrium between a keto and an enol group. The interconversion of the two
forms introduces 2H on C-H bonds adjacent to the carbonyl group. Another
isotopic exchange is via some reversible transaminations. The isotopic
equilibrium between alanine and pyruvate derives from transamination (260).
Alanine is 2H labeled very fast by transamination in the presence of [2H]water.
Pyruvate becomes 2H-labeled via alanine transamination.
Splitting of C-C bonds
The split of C-C bonds introduces hydrogen from the water solvent. A typical
example is the aldol cleavage. The splitting of fructose 1,6 – bisphosphate into
glyceraldehyde 3 – phosphate and dihydroxyacetone phosphate is catalyzed by
aldolase. The new C-H bonds are formed in the two three carbon compounds. If
the body water is 2H-enriched, then the two products are labeled by [2H]water
(261). Another example is the cleavage of citrate to acetyl-CoA and cytosolic
oxaloacetate by ATP citrate lyase. One 2H is introduced on C-2 of acetyl-CoA in
the presence of [2H]water.
3.3.3 General protocols of application of [2H]water
The use of [2H]water to trace the synthesis rate of polymers has been well
developed (264-266). The protocols (see figure 3.3) are mature and safe both for
rodents and humans. The animals first receive an intraperitoneal injection of
normal saline made in 100% 2H-enriched water. The calculation of the injected
amount is based on the body weight, body water content and the target of body
water 2H-enrichment (body water 2H-enriched up to 5% for the rodents). Then the
animal is provided 2H-enriched drinking water to compensate for the production
53
of unlabeled water from the oxidation of nutrients (267). Human subjects,
normally ingest 2H2O in amounts calculated to enrich body fluid at expecting body
water 2H- enrichment (up to 2%) (268). Then they drink some [2H]water daily for
the duration of the experiments. The 2H-enrichment of body water is assyed by (i)
isotope ratio mass spectrometry after reduction of water to H2, and a (ii) by GCMS of acetone after equilibration with water at high pH (268).
3.3.4 The pros and cons of using heavy water
[2H]Water is one of the mostly used tools to measure the turnover rate of
biomolecules in vivo. [2H]Water is a relatively inexpensive tracer. Compared to
tritiated water, [2H]water is non toxic at low enrichment range. For humans, the
safe enrichment ranges from 0.5% up to 2% (268); for rats, this working
enrichment is up to 5%. As a result, [2H]water can be safely administrated to
humans over long periods. The oral administration of [2H]water instead of
intravenous infusion makes taking [2H]water an easy process. When [2H]water
enters the body, it equilibrates rapidly with total body water. As a result, a
homogenously labeled precursor pool is formed. The newly synthesized products
contain deuterium in proportion to the enrichment of body water. This makes
[2H]water a good tracer because of the stably-labeled homogenous water pool.
Another significant advantage is that when using the heavy water method, one
can simultaneously trace several different biosynthetic reactions (fatty acid
synthesis, cholesterol synthesis, gluconeogenesis and DNA synthesis ) (254).
54
Figure 3.3:
Protocol for using [2H]water to trace the rates of synthesis and
degradation of biopolymers in vivo. Modified from Dufner D. et al. (254) Curr
Opin Clin Nutr Metab Care 6: 511-517.
55
A potential problem is that deuterium has a relatively long biological half-life (2
days in rats, 15 days in humans). So if one tries to measure the degradation of
biopolymers via the decay of deuterium, the biopolymers should have longer half
life than body water.
3.3.5 Tracing the syntheses of biopolymers
Lipid (Fatty acids) biosynthesis
[2H]Water is used to measure lipid synthesis via two ways. Schoenheimer et al.
(269) used [2H]water to synthesize lipids in some mice, and fed these 2H-labeled
lipids to other mice. Then they measured the conversion of labeled lipids to
different products. The rate of lipid turnover is calculated from the incorporation
of 2H from 2H-enriched body water into the lipids. To prove that the incorporation
only derives from the new synthesis, Hutman et al. (270) incubated adipose
tissue from rats fasted for 3 days (depletion of glycogens) with tritiated
water.
This is a condition that should not have fatty acid synthesis. They found no
labeling incorporation into fatty acids as expected. The labeling of fatty acids
does not occur by isotopic exchange between water and C-H bonds of fatty
acids.
Lipid synthesis is evaluated either by the numbers of incorporated 2H or as
fractional biosynthesis. The fractional biosynthesis is expressed as follows: (2Henrichment in the lipid) / (theoretical 2H-enrichment in the lipids). The maximum
number of 2H atoms incorporated per molecule (N) is calculated based on a
simple binomial distribution. The theoretical 2H enrichment is N x p, p is 2Henrichment in body water (271).
56
Jungas (272) quantified the total rate of fatty acid synthesis in incubations of rat
adipose tissue. The incorporation of 2H from [2H]water into fatty acids confirmed
that 23 of the hydrogen atoms of the palmitate were derived from 2H2O water.
Wadke et al. (273) found that in rat liver perfused with 100% 2H2O, the number of
deuterium atoms incorporated in [2H]water is 22 and 24 per molecule of newly
synthesized palmitate and stearate, respectively. More recently, Lee et al. (271)
measured the rates of fatty acid and cholesterol synthesis in rat liver and nervous
tissues by using [2H]water. They found that the maximum total incorporation of 2H
into liver palmitate, stearate and cholesterol is 22, 24 and 30, respectively. The
lipids in the nervous system had lower labeling than in the liver. This shows that
the rate of lipid synthesis varies between tissues. The [2H]water has also been
safely used to evaluate human lipogenesis. When subjects ingested [2H]water,
the 2H-enrichment of palmitate of plasma triglycerides (TG) plateaued at 0.6%
under the condition of 0.3% 2H-enrichment of body water. The newly synthesized
TG contributed about 8% to the plasma TG pools (274).
Protein biosynthesis
In 1941, Ussing (275) proposed the use of [2H]water to measure protein turnover
in vivo. He proposed that free amino acids become 2H-labeled at α carbon via
transamination. Free nonessential amino acids equilibrate quickly with 2H-labeled
body water via intermediary metabolic pathways before they are incorporated
into newly synthesized proteins. Once the protein is assembled, 2H is not able to
enter the amino acid residues in the protein backbones. The measurement of the
labeling of protein-bound amino acids yields the evaluation of protein turnover.
57
In 2004, Previs et al. (265) first practically used [2H]water to measure the rates of
protein synthesis in humans. They chose alanine as the precursor of protein
synthesis. Alanine turns over rapidly in vivo. The labeling of the α-hydrogens of
alanine equilibrates quickly with body water. As a result, the rate of protein
synthesis is determined by measuring the incorporation of 2H-labeled alanine into
proteins. The general calculation of protein synthesis rate is: (2H-labeled proteinderived alanine (%)) / (2H-labeled-body water (%)) x 3.7 x time(h)). On average,
3.7 out of 4 carbon-bound hydrogens of alanine exchange with body water (255).
Previs et al. (276) studied the influence of feeding on protein synthesis. They
found that about 50% of the plasma albumin had been synthesized within 5 hr
after feeding in rats.
There was no change in the rates of cardiac protein
synthesis between acute fasting or chronic food restriction. However, the protein
synthesis rate decreases in the liver and skeletal muscle in response to the
change of nutritional state. Hellerstein’s group also used glycine or glutamine
labeled from [2H]water to quantify protein synthesis (268).
DNA biosynthesis
Overview
Cell division and death are basic physiological processes. Cell division, defined
by DNA replication, occurs during the S-phase of the cell cycle. In the presence
of [2H]water, 2H is incorporated in DNA strands. The labeled DNA strands are
distributed equally in daughter cells. The labeling of DNA strands reflects the
fraction of newly synthesized cells formed via cell division. Deoxyribonucleotidetriphosphates are the immediate precursors of DNA synthesis. The deoxyribose
58
(dR) moiety of dNTPs becomes labeled by 2H from [2H]water via the de novo
nucleotide synthesis pathway (266). The C-2H bonds in the dR moiety are very
stable: there is no loss of labeling under physiological conditions. Not all the
carbon bonds of the dR moiety of DNA become 2H- labeled. Mass spectrometry
data show that up to six out of seven hydrogen atoms are labeled in the
presence of [2H]water. The M1 dR labeling is a nearly linear sum of the 2Henrichment of each C-H bonds. The labeling of dR, isolated from DNA, is
measured by GC-MS using different derivatives (277; 278).
Calculation of the rate of DNA turnover
The rate of DNA turnover is normally expressed as a fractional synthesis rate
based on the precursor – product relationship. There are two ways to calculate
fractional synthesis. If a fully turned-over tissue is available as the reference for
comparison, the calculation of fractional rate is expressed as (dR enrichment of
sample cell) / (dR enrichment of fully turned-over cells). In the rat, bone marrow
cells turn over very fast and are used as reference cells (256) In humans, blood
monocytes
and
granulocytes
are
normally
chosen
as
reference
cells
(256).However, reference cells are not available in many cases. Hellerstein et al.
(256) calculated the maximal dR labeling based on 2H-enrichment in body water.
This value is used to substitute for that of fully turned-over cells. Thus the
fractional synthesis is expressed as (dR enrichment of sample cells) / (maximal
dR enrichment) when fully turned-over cells are not available.
The application of [2H]water in normal and cancer cells
59
The [2H]water method is suitable for the measurement of the rate of DNA
synthesis both in slow turning-over cells and in fast turning-over cells. Human
monocytes and granulocytes turn over very fast: more than 20% of cells turn over
within one day (266). Rodents’ colonocytes are fully turned over after two days
as shown by 2H incorporation (279). In contrast, normal rat colon epithelial cells,
mouse mammary epithelial cells and vascular smooth muscle cells have very
slow turnover rates (266). Less than 3% of these cells turn over in one day.
Cell proliferation is the driving force of carcinogenesis and a prognostic
biomarker for cancer progression (280). Hellerstein’s group used [2H]water to
trace the synthesis rate of different cell types. When rats were given dietary
cholic acid, a cancer promoter, the DNA synthesis rates of colon epithelial cells
showed dose-dependent increases in the colonic crypts. The fractional synthesis
rate of DNA in the proliferative zone of the crypts increased from 44 % to 69% in
proportion to the dietary cholic acid content of the diet. The injection of the
carcinogen azoxymethane to rats increased DNA synthesis and the proliferation
of colon epithelial cells in all fractions of the crypts (281). [2H]Water also allows to
trace tumor cells over time. Breast tissue biopsies from women who have breast
cancer and undergo mastectomy show that 2H-labeled breast epithelial cells
remain in proliferation process after the cell apoptotic peak has occurred (282).
To examine the proliferation kinetics of endothelial cells and tumor cells in the
growing prostate cancer, Kim et al. (283) injected prostate cancer cell, PC-3
cells, into mice and used [2H]water to trace the proliferation of endothelial cells
and tumor cells for up to six weeks. The data showed that an increase in
60
endothelial cell proliferation precedes that of the tumor cells in the early phase of
tumor growth. This confirmed that the increase in endothelial cells corresponds to
the growth of tumor cells (283).
61
CHAPTER 4: RESEARCH PROPOSAL
4.1 Project I: C4- and C5-ketogenesis in rat liver
Overview
Compared with the extensive investigation of C4-ketone body regulation, there is
little information available on C5-ketone body metabolism. C5-ketone bodies refer
to β-hydroxypentanoate (BHP) and β-ketopentanoate (BKP). Their clinical names
are 3-hydroxyvalerate and 3-ketovalerate (113). C5-ketone bodies are derived
from the partial oxidation of odd-chain fatty acids. Body fluids contain only traces
of C5-ketone bodies because odd-chain fatty acids are absent from the diet of
non-ruminant mammals. C5-ketone bodies are found in body fluids of patients
with disorders of the anaplerotic pathway, propionyl-CoA → methylmalonyl-CoA
→ succinyl-CoA, such as deficiency in propionyl-CoA carboxylase and
methylmalonyl-CoA mutase as well as biotin or vitamin B12 deficiency (105; 113;
114). Peripheral tissues can use C5-ketone bodies as energy substrates (104).
My initial interest in C5-ketone body metabolism arose from an ongoing clinical
trial of dietary triheptanoin for the treatment of patients who have long-chain fatty
acids disorders (FOD) (106). Patients with FOD suffer from muscle weakness
and rhabdomyolysis. The accumulation of long-chain acyl-CoAs and long-chain
acyl-carnitines damages mitochondria and cell membranes. This leads to
pathological leakage of CAC intermediates which carry acetyl groups as they are
oxidized. As a consequence, the citric acid cycle does not operate optimally. The
classical treatment for FOD uses the even-chain triglyceride (284). The
catabolism of octanoate yields C4-ketone bodies in liver and provides the patients
62
with substitute substrates. However, the loss of CAC intermediates cannot be
compensated via trioctanoin treatment. It was hypothesized that boosting
anaplerosis would compensate for chronic cataplerosis and improve heart and
muscle function. Since 2002, some patients with FOD are treated with dietary
triheptanoin instead of trioctanoin (106). The catabolism of heptanoate produces
anaplerotic propionyl-CoA and C5-ketone bodies. In peripherial tissues, C5ketone bodies are converted to propionyl-CoA, which is used for anaplerosis.
The marked improvement of the patients’ conditions after switching from a
trioctanoin- to a triheptanoin-based diet supports the hypothesis.
Patients with FOD who ingest dietary triheptanoin accumulate both C4- and C5ketone bodies in their plasma. This suggested that acetyl-CoA groups derived
from heptanoate can be used for the synthesis of C4-and C5-ketone bodies.
Based on the above considerations, this project aims to investigate the following
questions:
1) What is the mechanism of regulation of C5-ketogenesis in the perfused rat
liver? Does the regulation reflect the metabolism of C5-ketone bodies in vivo?
2) Why do the patients with deficiency in propionyl-CoA carboxylase and
methylmalonyl-CoA accumulate only low level of C5-ketone bodies?
3) What are the interrelations between C4- and C5-ketogenesis in rat livers
perfused with octanoate and/or heptanoate?
4) What are the fates of acetyl groups of both fatty acids and of the propionylCoA moiety of heptanoate?
63
5) Does 3-Hydroxyl-3-ethylglutaryl-CoA (HEG-CoA) replace 3-Hydroxyl-3methylglutaryl-CoA (HMG-CoA) when C5-ketone boides are formed?
In order to answer the above questions, my experimental strategies are designed
with two types of perfusion experiments: i) perfuse single fatty acid alone, and ii)
perfuse two fatty acids together.
Experimental procedures
Rat livers will be perfused with 4 mM glucose and i) increasing concentrations (0
to 1 mM) of [13C3]propionate, [1-13C]octanoate, [8-13C]octanoate or [5,6,713
C3]heptanoate, or ii) 1mM octanoate ([1-13C]octanoate or [8-13C]octanoate) + 0
– 1 mM heptanoate ( or vice versa) with different labeling patterns ([113
C3]heptanoate or [5,6,7-13C3]heptanoate). The concentrations and labeling
pattern of C4- and C5-ketone bodies in the effluent perfusate will be assayed as
their tert-butyldimethylsilyl derivatives by GC-MS. Labeling of glucose will be
assayed on the pentaacetate derivative. The assay of the mass isotopomers
distribution of acyl-CoA esters from frozen liver powder will be analyzed by liquid
chromatography - mass spectrometry. Acyl-CoA esters will be identified and their
mass isotopomer distribution will be calculated.
Relative
rates
of
anaplerosis
from
[13C3]propionyl-CoA
precursors
([13C3]propionate, [5,6,7-13C3]heptanoate) will be calculated (18) as the ratio (m3
succinyl-CoA) / (m3 propionyl-CoA). In experiments with [1-13C]octanoate or 813
C]octanoate, calculation of the distribution of label between the two acetyl of
C4-ketone bodies will be based on the enrichment of the whole BHB molecule
and its fragments (164).
64
4.2 Project 2: Tracing the syntheses of adenine nucleotides, CoA and DNA
in rat liver
Overview
[2H]Water has been extensively used to trace the syntheses of fatty acids (269;
272), sterol (264), glucose (285), amino acids, peptides (286), proteins (265) and
nucleic acids (266; 287). The initial goal on this project was to measure the rate
of CoA turnover by using [2H]water. Because CoA shares the same backbone of
adenine-ribose nucleus with adenine nucleotides and DNA, later I expanded the
goal of the present study is to sort out the mechanisms by which 2H from the
aqueous medium becomes incorporated into C-H bonds during the biosyntheses
of adenine nucleotides, CoA and DNA in rat liver. In addition, [13C]-labeled
substrates will be used to complement the results from [2H]water. Another goal is
to evaluate the kinetics of the sequential steps of tracer incorporation. This
involves the interconnections between different pathways. Based on above, the
following questions are investigated:
1) What is the labeling pattern of CoA and its fragments?
2) How do adenine nucleotides and DNA become labeled in vivo?
3) How do glycolysis, gluconeogenesis, pentose phosphate pathway and the
three targeted pathways, interconnect?
4) Do the purine salvage pathways contribute to the syntheses of adenine
nucleotides, CoA and DNA?
Based on above questions, the following experiments are designed:
65
In vivo experiments: rats will receive an introperitoneal injection of normal saline
made up in 100% 2H2O in amounts calculated to achieve a 2.5% 2H-enrichment
of body water, assuming that total body water accounts for 66% of body weight.
The rats will be provided 3.25% 2H-enriched drinking water to compensate for
the producton of unlabeled water from the oxidation of foodstuffs and
endogenous substrates (271). At different time (2h to 31 days) the rats will be
anesthetized with isoflurane before freezing a liver lobe and sampling aortic
blood. In vitro experiments: isolated rat livers will be perfused with recirculating
bicarbonate buffer containing 4% dialyzed, fatty acid-free, bovine serum albumin,
4 mM glucose. In some experiments, (i) unlabeled glucose will be replaced by
[13C6]glucose, (ii) 2 mM [13C5]ribose will be added to 4 mM unlabeled glucose, (iii)
4 mM glucose will be added to perfusate made up in 100% 2H2O. Livers will be
quick-frozen at 2 h. One-half of the perfusion experiments will be conducted for 2
h. In other experiments, lobes of the livers will be tied off, cut out and quickfrozen at 30, 60, 90 and 120 min.
The concentration and mass isotopmer distribution of glucose will be assayed by
GC-MS of the permethyl and the pentaacetate derivatives, respectively. Lactate,
formate and dR moiety of DNA of MID will be assayed as the pentaflurobenzyl
derivative. 2H-enriched water will be assayed after equilibration with acetone in
alkaline medium. LC-MS analysis will be used to assay the MID of CoA, glucose6-phosphate, ribose-5-phosphate, ATP, ADP, AMP, PEP and 3-phosphoglycerate.
66
4.3 Publications
4.3.1 Deng S., Zhang G.F, Kasumove T., Roe C.R., and Brunengraber H.
Interrelations between C4-ketogenesis, C5-ketogenesis, and anaplerosis in the
perfused rat liver. J Biol Chem 284: 27799-27807, 2009.
4.3.2 Deng S., Zhang G.F., Kombu R.S. Harris S.R. DeSantis D., Vasquez E.J.,
Puchowicz M.A., Anderson V.E., Brunengraber H. Tracing the syntheses of
adenine nucleotides, CoA and DNA in rat liver. To be submitted to J Biol Chem.
67
This research was originally published in The Journal of Biological Chemistry.
Interrelations between C4-ketogenesis, C5-ketogenesis, and Anaplerosis in the
Perfused Rat Liver. The Journal of Biological Chemistry. 2009; (284): 2779927807. © The American Society for Biochemistry and Molecular Biology.
68
4.3.1
INTERRELATIONS BETWEEN C4-KETOGENESIS, C5-KETOGENESIS
AND ANAPLEROSIS IN THE PERFUSED RAT LIVER
Shuang Deng1, Guo-Fang Zhang1, Takhar Kasumov1,
Charles R. Roe2, and Henri Brunengraber1
From the Department of Nutrition1, Case Western Reserve University, Cleveland
OH 44106, and Institute of Metabolic Disease2, Baylor University Medical Center,
Dallas TX 75226
Running Title: C4- and C5-ketogenesis in liver
Address correspondence to: Henri Brunengraber, Department of Nutrition, Case
Western Reserve Univ., School of Medicine, WG 48, 10900 Euclid Ave.,
Cleveland,OH,44106-4954.
Tel.: 216. 368.6548; Fax: 216.368.6560; E-mail: [email protected]
Abstract
We investigated the interrelations between C4-ketogenesis (production of βhydroxybutyrate
+
acetoacetate),
C5-ketogenesis
(production
of
β-
hydroxypentanoate + β-ketopentanoate) and anaplerosis in isolated rat livers
perfused with
13
C-labeled octanoate, heptanoate or propionate.
Mass
isotopomer analysis of C4- and C5-ketone bodies and of related acyl-CoA esters
reveal that C4- and C5-ketogenesis share the same pool of acetyl-CoA. Although
the uptake of octanoate and heptanoate by the liver are similar, the rate of C5-
69
ketogenesis from heptanoate is much lower than the rate of C4-ketogenesis from
octanoate. This results from the disposal of the propionyl moiety of heptanoate
by anaplerosis of the citric acid cycle. C5-ketogenesis from propionate is virtually
nil because acetoacyl-CoA thiolase does not favor the formation of βketopentanoyl-CoA from propionyl-CoA and acetyl-CoA. Anaplerosis and
gluconeogenesis from heptanoate are inhibited by octanoate by competition. The
data have implications for the design of diets for the treatment of long-chain fatty
acid oxidation disorders, such as the triheptanoin-based diet.
Introduction
The regulation of the metabolism of C4-ketone bodies, i.e., β-hydroxybutyrate
(BHB) and acetoacetate (AcAc) has been extensively investigated in vivo, in
isolated livers, hepatocytes and subcellular preparations (for reviews, see (1-4)).
In contrast, very little information is available on the metabolism of C5-ketone
bodies, i.e., β-hydroxypentanoate (BHP) and β-ketopentanoate (BKP), which are
known in the clinical literature as 3-hydroxyvalerate and 3-ketovalerate (5,6).
The C5-ketone bodies are formed in liver from the partial oxidation of odd-chain
fatty acids (Fig 4.1, central column). C5-Ketogenesis uses the same enzymes of
the 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA) cycle as C4-ketogenesis. The
counterpart of HMG-CoA in C5-ketogenesis is 3-hydroxy-3-ethylglutaryl-CoA
(HEG-CoA). We only found one report on the formation of [14C]HEG-CoA in liver
extract incubated with propionyl-CoA and [1-14C]acetyl-CoA (7).
Because odd-chain fatty acids are absent from the diet of non-ruminant
mammals, body fluids contain only traces of C5-ketone bodies. However, C5-
70
ketone bodies and hydroxyethylglutarate are found in body fluids of patients with
disorders of the anaplerotic pathway: propionyl-CoA ➔ methylmalonyl-CoA ➔
succinyl-CoA, such as deficiency in propionyl-CoA carboxylase, methylmalonylCoA mutase, as well as biotin or vitamin B12 deficiency (5,6,8). The formation of
C5-ketone bodies in these pathological states involves either the conversion of
propionyl-CoA to BKP-CoA via 3-ketoacyl-CoA thiolase (Fig 4.1, reaction 1), or
the β-oxidation of odd-chain fatty acids synthesized in these patients (9) using
propionyl-CoA as a primer (10).
Like their C4 counterparts, the C5-ketone bodies are interconverted by
mitochondrial BHB dehydrogenase (11). In peripheral tissues, C5-ketone bodies
are converted to propionyl-CoA (which is anaplerotic) + acetyl-CoA via 3oxoacid-CoA transferase (12) and 3-ketoacyl-CoA thiolase. Peripheral tissues
have a high capacity to utilize exogenous C5-ketone bodies (13), especially
heart, kidney and brain which have high activities of 3-oxoacid-CoA transferase
(14,15).
Our interest in C5-ketone body metabolism arose from an ongoing clinical trial
where patients with long-chain fatty acid oxidation disorders are treated with a
diet containing triheptanoin (16,17), instead of the classical treatment with the
even-chain triglyceride trioctanoin. These patients suffer from muscle weakness
and rhabdomyolysis, manifested by the release of creatine kinase in plasma. It
was hypothesized that the accumulation of long-chain acyl-CoAs and long-chain
acylcarnitines results in membrane damage with release of large and small
molecules from cells. The leakage of small molecules would deplete
71
intermediates of the citric acid cycle (CAC) which carry acetyl groups as they are
oxidized. It was further hypothesized that boosting anaplerosis with a suitable
substrate would compensate for the chronic cataplerosis and improve heart and
muscle function. The catabolism of heptanoate yields propionyl-CoA which can
be used for anaplerosis in most tissues, and C5-ketone bodies in liver. C5-ketone
bodies are converted to propionyl-CoA which can be used for anaplerosis in
peripheral tissues. The marked improvement of the patients’ condition after
switching from trioctanoin to triheptanoin-based diet supported the hypothesis.
After ingestion of meals containing triheptanoin as the only lipid component, both
C5-ketone bodies and C4-ketone bodies accumulated in plasma of patients that
have been diagnosed with disorders of long-chain fatty acid oxidation (16). This
suggested that acetyl groups derived from heptanoate can be used for the
synthesis of C4- and C5-ketone bodies. Alternatively, the accumulation of C4ketone bodies after triheptanoin ingestion might result from the inhibition of the
utilization of C4-ketone bodies in peripheral tissues by C5-ketone bodies.
The aim of the present study was to investigate the interaction between C4- and
C5-ketogenesis in rat livers perfused with octanoate and/or heptanoate. To gain
insight on the fates of the acetyl groups of both fatty acids, and on the fate of the
propionyl-CoA moiety of heptanoate, we conducted the experiments with a series
of
13
labeled
substrates:
[1-13C]octanoate,
[8- 13 C]octanoate,
[5,6,7-
C 3 ]heptanoate, [1- 1 3 C]heptanoate and [ 13 C 3 ]propionate. The outcome
of the propionyl-CoA moiety of [5,6,7-13C3]heptanoate and [13C3]propionate was
traced by measurements of anaplerosis and glucose labeling by mass
72
isotopomer1 analysis (18). In previous studies on the metabolism of odd-chain
fatty acids in liver or hepatocytes (19,20), ketone bodies were assayed with BHB
dehydrogenase. This assay does not differentiate C4- from C5-ketone bodies. In
the present study, we used gas chromatography-mass spectrometry to
specifically assay C4- and C5-ketone bodies (13).
EXPERIMENTAL PROCEDURES
Materials- General chemicals were purchased from Sigma-Aldrich. [13C3]propionic,
[2H5]propionic, [1-13C]octanoic, [8-13C]octanoic, [2H15]octanoic, [2H13]heptanoic, and
[5,6,7-13C3]heptanoic acids,
2
H2O (99%), NaO2H, and
13
CO2 gas were obtained
from Isotec. [1-13C]Heptanoic acid was prepared by reacting hexylmagnesium
bromide with
13
CO2. The purity of the vacuum-distilled product was assessed by
GC-MS and NMR. Internal standards of β-hydroxy-[2,2,3,4,4,4-2H6]butyrate and
R,S-β-hydroxy-[2,2,3,4,4-2H5]pentanoate were prepared (13,21) by (i) incubating
ethyl acetoacetate or ethyl β-ketopentanoate in 2H2O + NaO2H overnight, (ii)
reacting with NaB2H4, (iii) acidification to destroy excess NaB2H4, and (iv)
neutralization and lyophilization. β-Hydroxypentanoate was prepared by
overnight hydrolysis of the ethyl ester with 1.1 equivalents of NaOH, followed by
neutralization
and
lyophilization.
Reagents
for
trimethylsilyl
and
tert-
butyldimethylsilyl derivatization were purchased from Pierce.
Liver perfusion experiments- Male Sprague-Dawley rats (130-140g) were fed ad
libitum for 8-12 days with standard laboratory chow. Livers from overnight-fasted
rats were perfused with non-recirculating bicarbonate buffer (30 ml/min) and
4mM glucose containing 0 to 1 mM of either a single fatty acid ([1-
73
13
C]octanoate, [8-13 C]octanoate, [5,6,7-13C3]heptanoate, [13C3]propionate),
or (1mM of octanoate + 0 to 1 mM of heptanoate (or vice versa)) with different
labeling patterns. At 14 and 18.5 min, samples of influent and effluent perfusate
(collected over 1 min) were treated with NaB2H4 and frozen. This procedure
stabilizes the unstable ketoacids by converting them to stable monodeuterated
hydroxyacids (13,21). Other samples of perfusate were frozen without treatment.
After 20 min, the livers were quick-frozen and kept in liquid nitrogen until
analysis.
Analytical Procedures In order to assay the uptake of fatty acids and the
production of C4- and C5-ketone bodies, 0.1 ml samples of NaB2H4- treated
perfusate (influent and effluent) were (i) spiked with internal standards (30 nmol
[2H13] heptanoate, 30 nmol [2H15]octanoate, 50 nmol [2H5] propionate, 45 nmol βhydroxy-[2H6]butyrate and 34 nmol β-hydroxy-[2H5]pentanoate), (ii) acidified (to
destroy excess NaB2H4), and (iii) deproteinized with 0.7 ml acetonitrile/methanol
(7:3). Supernatants from samples from experiments with octanoate and/or
heptanoate were dried with N2 and the residue treated to form the trimethylsilyl
derivatives of the analytes. Supernatants from samples from experiments with
propionate were dried with N2 and the residue treated to form the tertbutyldimethylsilyl derivatives of the analytes. The treatment of the samples with
NaB2H4 prevents assaying the individual concentrations of the labeled C4-ketone
bodies (BHB and AcAc) and C5-ketone bodies (BHP and BKP), because of
overlapping of some mass isotopomers. Therefore the data of this assay yield
total concentrations of C4- or C5-ketone bodies (13,21).
74
For the assay of the labeling pattern of individual C4- and C5-ketone bodies, 0.5
ml samples of effluent perfusate (not treated with NaB2H4) were incubated with
50 µmol methoxylamine-HCl (adjusted to pH 9) and incubated at 60°C for 40 min.
After acidifying with HCl to pH = 1, the samples were extracted 3 times with 5 ml
diethyl ether. The combined extracts were dried with N2, and the residue treated
to form the tert-butyldimethylsilyl derivatives of the ketone bodies. GC-MS assay
under electron ionization conditions yielded the mass isotopomer distribution of
the total C4- and C5-ketone body molecules, and of the fragments corresponding
to C-3+4 of C4-ketone bodies, and C-3+4+5 of C5-ketone bodies (13,21).
Labeling of glucose was assayed on the penta-acetate derivative.
For the assay of the mass isotopomer distribution of acyl-CoA esters, 250 mg of
frozen liver powder was extracted with 4 ml of methanol/H2O (1:1) containing 5%
acetic acid. After 1 min extraction at 0°C with a Polytron, and centrifugation at
8,000 g for 30 min at 4°C, the supernatant was run through a solid phase
extraction ion exchange cartridge packed with 300 mg of 2-2(pyridyl)ethyl silica
gel (Sigma-Aldrich). The cartridge was pre-activated with 3 ml of methanol
followed by 3 ml of methanol/H2O (1:1) containing 5% acetic acid. After washing
the cartridge with 3 ml of methanol/H2O (1:1) containing 5% acetic acid, the acylCoA esters were eluted out by (i) 3 ml of methanol/H2O (1:1) containing 50 mM
ammonium formate, (ii) 3 ml methanol/H2O (3:1) containing 50 mM ammonium
formate, and (iii) 3 ml methanol. The combined effluent was dried under N2 and
the residue stored at -80ºC until LC-MS analysis.
75
LC-MS assays- After dissolving the acyl-CoAs in 100 µl of HPLC buffer A (5%
acetonitrile in 100 mM ammonium formate, pH 5.0), 10 µl of solution was injected
on a Thermo Electron Hypersil GOLD column (C18, 100 × 2.1 mm, 3 µm particle
size) protected by a guard column (Hypersil Gold, C18, 10 × 2.1 mm , 3 µm
particle size). Gradient elution at constant flow rate of 0.2 ml/min was: (i) 98%
buffer A + 2% buffer B (5 mM ammonium formate in 95% acetonitrile) for 7 min,
(ii) from 2% to 60% B from 7 to 25 min, (iii) from 60% to 90% B from 25 to 26
min, (iv) 90% B from 26 to 30 min, and (v) re-equilibration with initial buffer for 9
min before next injection. The order of acyl-CoA elution (min) was malonyl-CoA
(2.2), methylmalonyl-CoA (2.7), succinyl-CoA (3.5), HMG-CoA (3.9), acetyl-CoA
(6.7), AcAc-CoA (6.9), BHB-CoA (8.1), HEG-CoA (9.1), BKP-CoA (14.0),
propionyl-CoA (14.2), BHP-CoA (14.4), pentanoyl-CoA (17.7), hexanoyl-CoA
(19.5), heptanoyl-CoA (20.9), octanoyl-CoA (22.2).
The liquid chromatograph was coupled to a 4000 QTrap mass spectrometer
(Applied Biosystems, Foster City, CA) operated under positive electrospray
ionization mode with the following parameters: the source temperature was set at
600ºC with gas 1 and gas 2 at 65 and 55 psi, respectively. The curtain gas was
at 30 psi and the collision-activated dissociation gas pressure was held at high.
The turbo ion-spray voltage, declustering potential, entrance potential, and
collision cell exit potential were 4500, 70, 10, and 50 V, respectively. Multiple
reaction monitoring mode was used for quantitation and isotope enrichment
analysis. The analyst software (version 1.4.2, Applied Biosystems) was used for
data collection and analysis.
76
Calculations- Correction of measured mass isotopomer distributions for natural
enrichment was performed using the CORMAT software (22). Relative rates of
anaplerosis from [13C3]propionyl-CoA precursors ([13C3]propionate, [5,6,713
C3]heptanoate) were calculated (18) as the ratio (m3 succinyl-CoA)/(m3
propionyl-CoA).
These ratios refer to the contribution of the anaplerotic
substrates to the catalytic intermediates of the CAC which carry acetyl units as
they are oxidized. Note that, when label enters the CAC only via propionyl-CoA,
M3 succinyl-CoA is only formed from the sequence: M3 propionyl-CoA ➔ M3
methylmalonyl-CoA ➔ M3 succinyl-CoA.
(Recycling of label in the CAC
cannot form M3 succinyl-CoA).
In experiments with [1-13C]octanoate or [8-13C]octanoate, the distribution of label
between the two acetyl of C4-ketone bodies was calculated using (i) the m1 and
m2 enrichments of the whole BHB molecule, and (ii) the m1 enrichment of the C3+4 fragment of BHB. The m1 enrichment of the C-1+2 acetyl of BHB was
calculated as: m1 of C-1+2 = [(2 m2 + m1) of C-1→4] - (m1 of C-3+4).
Data presentation and Statistics- Herein, we present data from ∼70 liver
perfusion experiments. For each of the conditions chosen, we ran 6 perfusions
in the presence of selected unlabeled or
13
C-labeled substrate(s) with the
concentration parameters being allowed to vary. The data points shown in the
figures represent means of duplicate gas chromatography-mass spectrometry or
liquid chromatography-mass spectrometry injections, which differed by < 2%.
The statistical differences between some profiles were tested using a paired t
test (Graph Pad Prism Software, version 3).
77
RESULTS AND DISCUSSION
Relationship between the uptake of C8-, C7-, and C3-fatty acids and the formation
of C4- and C5-ketone bodies. Fig 4.2A shows the uptake of octanoate and its
conversion to C4-ketone bodies. Because 1 molecule of octanoate can yield up
to two molecules of C4-ketone bodies, the yield of C4-ketogenesis from
exogenous octanoate ranged from 80 to 90%. This calculation assumes that
basal C4-ketogenesis, at zero octanoate concentration, was not inhibited by
increasing concentrations of octanoate. Similar data were reported by McGarry
and Foster (23). Fig. 4.2B shows the uptake of heptanoate and its conversion to
both C5- and C4-ketone bodies. The uptake of heptanoate (Fig. 4.2B) was very
similar to the uptake of octanoate (Fig. 4.2A).
Because the oxidation of 1
molecule of heptanoate yields 1 molecule of propionyl-CoA and 2 molecules of
acetyl-CoA, one calculates that (i) only about 40% of the propionyl moiety of
heptanoate was converted to C5-ketone bodies, and (ii) about 75% of the acetyl
moiety of heptanoate was converted to C5- and C4-ketone bodies. Thus, as
outlined in Fig. 4.1, acetyl-CoA derived from heptanoate seems to be used to
form both C5- and C4-ketone bodies. This will be confirmed by the labeling data
presented below.
The low yield of conversion of the propionyl moiety of
heptanoate to C5-ketone bodies results from its diversion to anaplerosis and
gluconeogenesis (Fig. 4.1) as will be shown below.
Fig. 4.2C shows the uptake of propionate and the release of C5- and C4-ketone
bodies. The yield of C5-ketogenesis from propionate was extremely low (about
0.1%). The much lower yield of C5-ketogenesis from propionate compared to
78
heptanoate (Fig. 4.2B) reflects the properties of 3-ketoacyl-CoA thiolase (Fig.
4.1, reaction 1), which had been described for the interconversion of AcAc-CoA
and acetyl-CoA by this enzyme (24-27).
Although the thiolase reaction is
reversible in vitro, its kinetic properties prevent the formation of AcAc-CoA from
acetyl-CoA. This explains why C4-ketogenesis cannot be fueled by acetyl-CoA
derived from glucose metabolism. The virtually nil C5-ketogenesis from
propionate demonstrates that, in the intact liver, thiolase does not allow the
formation of BKP-CoA from propionyl-CoA + acetyl-CoA. In contrast, thiolase
allows the cleavage of BKP-CoA derived from heptanoate to propionyl-CoA +
acetyl-CoA. Fig. 4.2C also shows that propionate inhibits C4-ketogenesis from
endogenous fatty acids, as reported by Brass and Beyerinck (20).
To test the interaction between the uptakes of octanoate and heptanoate, as well
as the formation of C4- and C5-ketone bodies, we perfused livers with 1 mM of
one fatty acid and increasing concentrations of the second fatty acid, and viceversa (Figs. 4.3 and 4.4). Figs. 4.3A and 4.3B show that, starting with similar
uptakes of 1 mM octanoate or heptanoate alone, the effects of octanoate on
heptanoate metabolism are different from the effects of heptanoate on octanoate
metabolism. The competition favors octanoate uptake over heptanoate uptake
(Figs. 4.3A and 4.3B; p< 0.05 for both comparisons). Fig. 4.4A shows that, when
increasing concentrations of heptanoate are added to a constant 1 mM
octanoate, the production of C4-ketone bodies was not inhibited (slope was not
significantly different from zero). Also, the production of C5-ketone bodies was
much lower than in the presence of heptanoate alone (Fig. 4.2B). Fig. 4.4B
79
shows that, when increasing concentrations of octanoate were added to a
constant 1 mM heptanoate, the production of C4-ketone bodies increased
markedly, almost as much as in the presence of octanoate alone
(Fig. 4.2A).
Furthermore, the production of C5-ketone bodies from heptanoate decreased
compared to what occurred in the presence of 1 mM heptanoate alone (Fig.
4.2B).
Labeling of ketone bodies. Before presenting the labeling pattern of C4- and C5ketone bodies labeled from [13C]octanoate or/and [13C]heptanoate, we want to
stress that the interpretation of the data must take into account (i) the concept of
zonation of liver metabolism (28), and (ii) the factors that influence the
distribution of label between the two acetyl moieties of C4-ketone bodies. First,
Fig. 4.5 shows that at low influent concentrations of octanoate (up to 0.2 mM),
almost no substrate left the liver in the effluent perfusate. In perfusions with
increasing concentrations of heptanoate, the profile of effluent heptanoate
concentrations (not shown) was identical to what was measured in perfusions
with octanoate. Under these conditions, the pericentral cells of the liver lobule
were in contact with much lower octanoate or heptanoate concentrations than the
periportal cells.
The zonation of propionate concentrations (Fig. 4.5, upper
curve) was less pronounced than in perfusions with octanoate, but the profiles
were not significantly different. Second, the distribution of label between the two
acetyl moieties of C4-ketone bodies depends on whether label enters the HMGCoA cycle as acetyl-CoA or as the C-3+4 moiety of AcAc-CoA (Fig. 4.1). The
latter is formed in one of the final steps of fatty acid β-oxidation. Hüth has shown
80
that, in the reversible reaction catalyzed by 3-ketoacyl-CoA thiolase (AcAc-CoA +
CoA ↔ 2 acetyl-CoA (Fig. 4.1 (left side), Reaction 1), the C-3+4 moiety of AcAcCoA exchanges much more slowly with the free acetyl-CoA pool than its C-1+2
moiety (25).
For livers perfused with increasing concentrations of [1-13C]octanoate or [813
C]octanoate, Figs. 4.6, A and B show the labeling of BHB, its acetyl moieties
and of liver acetyl-CoA. We calculated the distribution of label between the two
acetyl moieties of C4-ketone bodies from the electron ionization mass spectrum
of the TBDMS derivative of BHB, which includes ions corresponding to carbons 1
to 4, and carbons 3 to 4 of BHB (21). The two acetyl moieties of BHB had very
different labeling. In the presence of [1-13C]octanoate, most of the label of BHB
was on the C-1+2 acetyl, with relatively little label on the C-3+4 acetyl (Fig. 4.6A).
The opposite labeling pattern of BHB was observed in perfusions with [813
C]octanoate, where 90% of the BHB labeling was on the C-3+4 acetyl, the
labeling of which plateaued at 37%. This results from the fact that the C-3+4
acetyl of BHB derives from the C-7+8 acetyl of octanoate via the C-3+4 acetyl of
AcAc-CoA. Because the C-3+4 acetyl of AcAc-CoA exchanges poorly with the
free acetyl-CoA pool, the two acetyl moieties of BHB have very unequal labeling.
Katz (29) had suggested that the labeling of C-1+2 of BHB could be taken as a
proxy, i.e., an indicator, of that of liver mitochondrial acetyl-CoA (29). Fig. 4.6A
shows that, in the presence of [1-13C]octanoate, the enrichment of acetyl-CoA
plateaued close to 25%, which is the maximal possible enrichment because [113
C]octanoate was labeled on only one acetyl moiety. At high [1-13C]octanoate
81
concentrations, the labeling of acetyl-CoA and of the C-1+2 acetyl of BHB were
very close. However, at low [1-13C]octanoate concentrations (0.1 to 0.2 mM), the
C-1+2 acetyl of BHB was much more labeled than acetyl-CoA. There are two
possible explanations of this discrepancy: metabolic zonation (28) and/or
metabolic channeling (166). First, in livers perfused with 0.1 to 0.2 mM [113
C]octanoate, the pericentral cells of the liver lobule were in contact with
perfusate containing little or no labeled octanoate, as inferred from the very low
concentration of octanoate in the effluent perfusate (Fig. 4.5). In the pericentral
cells, acetyl-CoA was unlabeled or minimally labeled. Thus, when acetyl-CoA
was assayed in a total liver extract, its enrichment was a composite of labeled
periportal and unlabeled pericentral acetyl-CoA. Moreover, because ketogenesis
predominates in the periportal cells, there were little unlabeled ketone bodies
produced in the pericentral cells, thus little dilution of the ketone bodies in the
total liver extract. Second, metabolic channeling transfers labeled acetyl-CoA
derived from fatty acid oxidation directly to the HMG-CoA cycle without
equilibration with the total pool of mitochondrial acetyl-CoA.
We previously
showed (31) that, in livers perfused with 0.2 mM [1-13C]octanoate, the labeling of
the C1+2 moiety of BHB is greater than that of the acetyl moiety of citrate (a
proxy of mitochondrial acetyl-CoA).
In the perfusions with [8-13C]octanoate (Fig. 4.6B), the enrichment of the C-1+2
acetyl of BHB was greater than the almost nil enrichment of acetyl-CoA (p =
0.0005). This is a consequence of the poor equilibration of label between [413
C]AcAc-CoA derived from [8-13C]octanoate and free acetyl-CoA via thiolase.
82
Therefore, the labeling of the C1+2 acetyl of BHB cannot, in most case, be used
as a proxy of the labeling of mitochondrial acetyl-CoA.
C4-ketogenesis and C5-ketogenesis share a pool of acetyl-CoA. The labeling
pattern of ketone bodies, HMG-CoA and HEG-CoA show that, when octanoate
and heptanoate are used by the liver, acetyl-CoA derived from each substrate is
available to both C4- and C5-ketogenesis.
Consider the labeling of ketone
bodies. First, in liver perfused with [1-13C]heptanoate alone (Fig. 4.7A, left side),
BHB was M1 and M2 labeled. Therefore, labeled acetyl-CoA derived from [113
C]heptanoate was incorporated into C4-ketone bodies. As increasing
concentrations of unlabeled octanoate were added to 1 mM [1-13C]heptanoate,
the M1 and M2 enrichments of BHB decreased, as expected. Second, in livers
perfused with 1 mM [5,6,7-13C3]heptanoate alone (left side of Fig. 4.7B), BHP
was only M3 labeled, as expected.
When increasing concentrations of [1-
13
C]octanoate were added to [5,6,7-13C3]heptanoate, BHP became up to 20% M4
labeled, while M3 labeling decreased from 90% to 70% (Fig. 4.7B). Thus, in this
experiment, M4 BHP was formed from a M3 propionyl-CoA derived from
[5,6,7- 13C3]heptanoate, and a M1 acetyl-CoA derived from [1-13C]octanoate.
However, when increasing concentrations of [8-13C]octanoate were added to 1
mM [5,6,7-13C3]heptanoate (Fig. 4.7C), the M4 labeling of BHP was very low (up
to 2%), while the M3 enrichment of BHP barely decreased from 90% (not
shown). This is because [8-13C]octanoate does not substantially label acetylCoA, as mentioned above (Fig. 4.6B). Third, in perfusions with constant 1 mM [113
C]octanoate and increasing concentrations of unlabeled heptanoate, M1 BHP
83
did accumulate (Fig. 4.7D). This M1 BHP was formed from unlabeled propionyl
(derived from heptanoate), and M1 acetyl-CoA (derived from [1-13C]octanoate).
These three set of experiments show that C4- and C5-ketogenesis share the
same acetyl-CoA pool (Fig. 4.1).
Consider now the labeling patterns of HMG-CoA (Fig. 4.8A) and HEG-CoA (Fig.
8B). In a liver perfused only with [1-13C]heptanoate (Fig. 4.8, A and B, left sides),
HMG-CoA was M1, M2, and M3 labeled, while HEG-CoA was M1 and M2
labeled. The M1 and M2 enrichments of HMG-CoA were not significantly different
from the corresponding enrichments of HEG-CoA. This also indicates that the
syntheses of HMG-CoA and HEG-CoA share the same pool of acetyl-CoA, in this
case
([1-13C]acetyl-CoA
derived
from
[1-13C]heptanoate).
As
increasing
concentrations of unlabeled octanoate were added to the 1 mM [113
C]heptanoate, the enrichments of (i) the M1, M2, and M3 isotopomers of HMG-
CoA, and (ii) M1 and M2 isotopomers of HEG-CoA decreased. This
demonstrates that HMG-CoA synthesis can use acetyl-CoA derived from
heptanoate oxidation, and that HEG-CoA synthesis can use acetyl-CoA derived
from octanoate oxidation. This does confirm that acetyl-CoA derived from
octanoate oxidation was used for C5-ketogenesis.
Reversibility of the BHB-CoA dehydrogenase reaction. When [1-13C]octanoate
undergoes β-oxidation, carbons 5 to 8 of octanoate (which will go to BHB-CoA
and AcAc-CoA) are initially unlabeled. Although as predicted, hexanoyl-CoA and
butyryl-CoA were unlabeled (not shown), BHB-CoA and AcAc-CoA were M1 and
M2 labeled (Fig. 4.9). The labeling of AcAc-CoA results from the partial isotopic
84
equilibration of AcAc-CoA and acetyl-CoA via 3-ketoacyl-CoA thiolase (Fig. 4.1,
Reaction 1). The M1 and M2 labeling of BHB-CoA (Fig. 4.9) shows that the
BHB-CoA dehydrogenase reaction is reversible in the intact liver in spite of the
high rate of β-oxidation. A similar equilibration occurs in the catabolism of [113
C]heptanoate. In one perfusion with 1 mM [1-13C]heptanoate, although
pentanoyl-CoA was unlabeled, BHP-CoA and acetyl-CoA were 12.7% and 40%.
M1 labeled, respectively (BKP-CoA was undetectable).
Anaplerosis from heptanoate and propionate.
Anaplerosis can be expressed as a relative or an absolute flux. Relative
anaplerosis is the fractional contribution of an anaplerotic substrate to the fourcarbon component of CAC intermediates which carry acetyl groups as they are
oxidized. For example, in Fig 4.10A, [13C3]propionate contributes up to 0.37 of
the catalytic intermediates of the CAC. The remaining fraction (1 – 0.37 = 0.63)
derives from recycling of catalytic intermediates. Absolute anaplerosis is a flux of
an anaplerotic substrate into the CAC, expressed as μmol•(g dry wt)-1•min-1.
In experiments with precursors of M3 propionyl-CoA ([5,6,7-13C3]heptanoate and
[13C3]propionate),
we calculated relative anaplerosis as the m3 enrichment ratio
(succinyl-CoA)/(propionyl-CoA) (Fig. 4.10A).This ratio represents the contribution
of the propionyl-CoA precursor to CAC catalytic intermediates which carry acetyl
units as they are oxidized. Assuming that propionate and heptanoate in the liver
are channeled only to anaplerosis and C5-ketogenesis, we calculated absolute
anaplerosis from each propionyl-CoA precursor as (uptake of substrate minus
C5-ketogenesis)/(fractional anaplerosis from this substrate). This rate, which is
85
up to 15 μmol⋅min−1⋅(g dry wt)−1 for [5,6,7-13C3]heptanoate
(Fig. 4.10B), does
not represent a flux of acetyl-CoA through the CAC. It does represent the rate of
the sections of the CAC going from succinyl-CoA to oxaloacetate, the
cataplerotic intermediate leading to PEP and glucose.
Relative anaplerosis from [13C3]propionate increased faster with substrate
concentration than relative anaplerosis from [5,6,7-13C3]heptanoate. The opposite
occurred with absolute anaplerosis (compare Figs. 4.10A and 4.10B).
This
probably results in a faster CAC flux in the presence of [5,6,7-13C3]heptanoate
(which supplies 2 acetyl-CoA) than in the presence of [13C3]propionate (which
supplies no acetyl-CoA).
Absolute anaplerosis from [5,6,7- 13C3]heptanoate was significantly decreased
in the presence of octanoate (p <0.05), although the latter is not anaplerotic (Fig.
4.10B, top vs. bottom curve).
This results from the inhibition of heptanoate
uptake by octanoate (Figs. 4.2B and 4.3A). Thus, anaplerosis from an odd-chain
fatty acid is modulated by the presence of an even-chain fatty acid. If all absolute
anaplerosis from [5,6,7-13C3]heptanoate or [13C3]propionate (Fig. 4.10B) were
used for gluconeogenesis, one calculates that the glucose in the effluent
perfusate should be about 1.4 to 1.7% labeled in M2 + M3 isotopomers. This
calculation takes into account the m3 enrichment of succinyl-CoA, the absolute
anaplerosis and the supply of unlabeled glucose in the influent perfusate. In fact,
glucose labeling from [5,6,7-13C3]heptanoate and [13C3]propionate plateaued at
0.8 to 1%. This dilution (by a factor of about 2) results from isotopic exchanges
between CAC intermediates and related compounds, as shown by Hetenyi (32).
86
Quite striking is the absence of labeling of glucose from [5,6,7-13C3]heptanoate in
the presence of unlabeled octanoate (Fig. 4.10C, lower curve, ■). This, in spite
of the fact that anaplerosis from [5,6,7-13C3]heptanoate was not fully abolished by
octanoate (Fig. 4.10B).
This inhibition may result from the increase in the
[NADH]/[NAD+] ratios induced by the rapid oxidation of octanoate.
One can wonder whether zonation of liver metabolism influences calculated rates
of relative anaplerosis, especially at low [5,6,7-13C3]heptanoate concentrations,
when the pericentral hepatocytes are not in contact with the labeled substrate
(Fig. 4.5). The pericentral cells must have pools of unlabeled propionyl-CoA,
methylmalonyl-CoA and succinyl-CoA derived from aminoacid catabolism and
recycling of succinyl-CoA in the CAC. Hence, in the extract of a whole liver
perfused with a low concentration of [5,6,7-13C3]heptanoate, the m3 enrichments
of propionyl-CoA and succinyl-CoA formed in periportal cells are diluted by
unlabeled substrates formed in the pericentral cells. Moreover, the relative
anaplerosis calculated at low [5,6,7-13C3]heptanoate concentration may either be
over- or underestimated, depending on the relative sizes and enrichments of the
propionyl-CoA and succinyl-CoA pools in the pericentral cells.
Concluding
remarks.
measurements
of
In
the
substrate
above
fluxes
investigations,
and
mass
the
association
isotopomer
analysis
of
of
intermediates provides a wealth of information on the interrelation between C4ketogenesis, C5-ketogenesis and anaplerosis. Our data extend to C5-ketogenesis
the concept that the mitochondrial 3-ketoacyl-CoA thiolase reaction (Fig. 4.1,
Reaction 1) does not allow a net acyl-CoA condensation flux (33). This is shown
87
by the absence of C5-ketogenesis from propionate (Fig. 4.2C), and the absence
of C4-ketogenesis from acetate (34) or from glucose. The absence of net acylCoA condensation does not prevent partial isotopic equilibration between acetylCoA and AcAc-CoA, as shown by (i) the incorporation of label from [113
C]octanoate into C4-ketone bodies (Fig. 4.6A), and C5-ketone bodies (in the
presence of unlabeled heptanoate, Fig. 4.7D), and (ii) the incorporation of label
from [1-14C]acetate into C4-ketone bodies (see Fig. 4.1 of (34)).
Also, the
absence of net condensation between propionyl-CoA derived from propionate
and acetyl-CoA, resulting in only traces of C5-ketone body production (Fig. 4.2C),
does not prevent incorporation of label from [13C3]propionate into HEG-CoA
which was 90% M3 labeled (not shown). The absence of C5-ketogenesis from
propionate or propionyl-CoA strongly suggests that the concentration of C5ketone bodies found in body fluids of patients with disorders of the propionyl-CoA
pathway are formed via β-oxidation of odd-long-chain fatty acids synthesized
from propionyl-CoA in these patients.
This is similar to C5-ketogenesis from
heptanoate (Fig. 4.1, middle column).
Our data confirm Hüth et al. (25) finding that the rate of equilibration of the C3+4
acetyl moiety of AcAc-CoA with acetyl-CoA via AcAc-CoA thiolase is much
smaller than the rate of equilibration of the C1+2 moiety. This is clearly illustrated
by the difference in the enrichments of the C1+2 and C3+4 moieties of C4-ketone
bodies in the presence of [1-13C]octanoate vs. [8-13C]octanoate (Fig. 4.6).
Although the liver takes up octanoate or heptanoate at similar rates (Fig. 4.2, A
and B), the flux of C4-ketogenesis is more rapid than that of C5-ketogenesis. This
88
results from the diversion of the propionyl moiety of heptanoate to anaplerosis of
the CAC and gluconeogenesis (Figs. 4.1, 10). Also, in the presence of octanoate
and heptanoate, the uptake of octanoate and C4-ketogenesis prevails over the
uptake of heptanoate and C5-ketogenesis.
The inhibition of anaplerosis and
gluconeogenesis from heptanoate by octanoate (Fig. 4.10, B and C) has
implications for the design of diets, which are both anaplerotic and
gluconeogenic, for the treatment of some metabolic diseases such as disorders
of long-chain fatty acid oxidation. Physicians may be tempted to progressively
modify the patients’ diet to replace medium-even-chain triglycerides (which have
been used since the 1980s to treat such patients (35)) by triheptanoin (16,17).
This would not be advisable because octanoate inhibits heptanoate uptake (Fig
4.3B), C5-ketogenesis from heptanoate (Fig 4.4B) and anaplerosis from
heptanoate (Figs 4.10A, 4.10B).
REFERENCES
1. McGarry, J. D. and Foster, D. W. (1980) Annu. Rev. Biochem. 49, 395-420
2. Robinson, A. M. and Williamson, D. H. (1980) Physiol Rev. 60, 143-187
3. Balasse, E. O. and Féry, F. (1989) Diabetes Metab Rev. 5, 247-270
4.
Fukao, T., Lopaschuk, G. D., and Mitchell, G. A. (2004) Prostaglandins
Leukot. Essent. Fatty Acids 70, 243-251
5.
Sweetman, L., Weyler, W., Nyhan, W. L., de Cespedes, C., Loria, A. R., and
Estrada, Y. (1978) Biomed. Mass Spectrom. 5, 198-207
6.
Kuhara, T., Inoue, Y., Shinka, T., Matsumoto, I., and Matsuo, M. (1983)
Biomed. Mass Spectrom. 10, 629-632
7.
Baker, F. C. and Schooley, D. A. (1981) Biochim. Biophys. Acta 664, 356372
8.
Duran, M., Gompertz, D., Bruinvis, L., Ketting, D., and Wadman, S. K.
(1978) Clin. Chim. Acta 82, 93-99
89
9.
Hommes, F. A., Kuipers, J. R., Elema, J. D., Jansen, J. F., and Jonxis, J. H.
(1968) Pediatr. Res. 2, 519-524
10. Wakil, S. J. (1989) Biochemistry 28, 4523-4530
11. Bergmeyer, H. U., Gawehn, K., Klotzsch, H., Krebs, H. A., and Williamson,
D. H. (1967) Biochem. J. 102, 423-431
12. Jencks, W. P. (1973) In Boyer, P. D., editor. The Enzymes, Academic
Press, New York
13. Leclerc, J., Des Rosiers C., Montgomery, J. A., Brunet, J., Ste-Marie, L.,
Reider, M. W., Fernandez, C. A., Powers, L., David, F., and Brunengraber,
H. (1995) Am. J. Physiol 268, E446-E452
14. Williamson, D. H., Bates, M. W., Page, M. A., and Krebs, H. A. (1971)
Biochem. J. 121, 41-47
15. Fukao, T., Song, X. Q., Mitchell, G. A., Yamaguchi, S., Sukegawa, K., Orii,
T., and Kondo, N. (1997) Pediatr. Res. 42, 498-502
16. Roe, C. R., Sweetman, L., Roe, D. S., David, F., and Brunengraber, H.
(2002) J. Clin. Invest 110, 259-269
17. Roe, C. R., Yang, B. Z., Brunengraber, H., Roe, D. S., Wallace, M., and
Garritson, B. K. (2008) Neurology 71, 260-264
18. Kasumov, T., Cendrowski, A. V., David, F., Jobbins, K. A., Anderson, V. E.,
and Brunengraber, H. (2007) Arch. Biochem. Biophys. 463, 110-117
19. Krebs, H. A. and Hems, R. (1970) Biochem. J. 119, 525-533
20. Brass, E. P. and Beyerinck, R. A. (1988) Biochem. J. 250, 819-825
21. Des Rosiers C., Montgomery, J. A., Desrochers, S., Garneau, M., David, F.,
Mamer, O. A., and Brunengraber, H. (1988) Anal. Biochem. 173, 96-105
22. Fernandez, C. A., Des Rosiers, C., Previs, S. F., David, F., and
Brunengraber, H. (1996) J. Mass Spectrom. 31, 255-262
23. McGarry, J. D. and Foster, D. W. (1971) J. Biol. Chem. 246, 1149-1159
24. Hüth, W., Dierich, C., von Oeynhausen, V., and Seubert, W. (1974)
Biochem. Biophys. Res. Commun. 56, 1069-1077
25. Hüth, W., Dierich, C., von, Oeynhausen, V., and Seubert, W. (1973) Hoppe
Seylers. Z. Physiol Chem. 354, 635-649
26. Hüth, W. and Alves, F. (1985) Biochim. Biophys. Acta 830, 274-281
27. Balzer, I. and Hüth, W. (1986) Biochim. Biophys. Acta 870, 350-356
90
28. Jungermann, K. and Kietzmann, T. (1996) Annu. Rev. Nutr. 16, 179-203
29. Katz, J. (1985) Am. J. Physiol 248, R391-R399
30. Srere, P. A. (1987) Annu. Rev. Biochem. 56, 89-124
31. Des Rosiers, C., David, F., Garneau, M., and Brunengraber, H. (1991) J
Biol. Chem 266, 1574-1578
32. Hetenyi, G., Jr. (1982) Fed. Proc. 41, 104-109
33. Hüth, W. and Menke, R. (1982) Eur. J. Biochem. 128, 413-419
34. Endemann, G., Goetz, P. G., Tomera, J. F., Rand, W. M., Desrochers, S.,
and Brunengraber, H. (1987) Biochem Cell Biol. 65, 989-996
35. Bougnères, P. F., Saudubray, J. M., Marsac, C., Bernard, O., Odièvre, M.,
and Girard, J. (1981) J. Pediatr. 98, 742-746
FOOTNOTES
This work was supported by the NIH (Road Map grant 5R33DK070291 and grant
5R01DK069752) and the Cleveland Mt. Sinai Health Care Foundation.
1
Mass isotopomers are designated as M1, M2…Mn, where n is the number of
heavy atoms in the molecule. The mol percent enrichment of each isotopomer is
designated as m1, m2…mn.
The abbreviations used are: AcAc, acetoacetate; BHB, β-hydroxybutyrate; BKP,
β-ketopentanoate
BHP,
(3-ketovalerate);
β-hydroxypentanoate
(3-
hydroxyvalerate); CAC, citric acid cycle; HEG-CoA, β-hydroxy-β-ethylglutarylCoA;
HMG-CoA,
β-hydroxy-β-methylglutaryl-CoA;
enrichment.
91
MPE,
molar
percent
FIGURE LEGENDS
Figure 4.1 Scheme of C4-ketogenesis and C5-ketogenesis in the liver. Numbers
refer to the following enzymes: 3-ketoacyl-CoA thiolase (1), HMG-CoA synthase
(2), HMG-CoA lyase (3), β-hydroxybutyrate dehydrogenase (4). The figure also
shows the link between propionyl-CoA and the citric acid cycle (CAC) via
anaplerosis.
Figure 4.2 Comparison between the uptake of octanoate (A), heptanoate (B), or
propionate (C) and the production of C4-ketone bodies (β-hydroxybutyrate +
acetoacetate) and C5-ketone bodies (β-hydroxypentanoate + β-ketopentanoate).
(fatty acid uptake, ●; C4-ketogenesis, ■; C5-ketogenesis, ▲). A, no C5-ketone
bodies were detected in the presence of octanoate. A, B, C, n = 6 at zero
concentration of influent fatty acid; n = 1 for the other concentrations. All liver
perfusions reported in this manuscript were conducted for 20 min. All rates of
substrate uptake and production (Figs 2 to 5) were assayed on samples of
influent and effluent perfusate taken at 14 and 18.5 min (collecting the samples
over 1 min). All reported values are the means of the 14 and 18.5 min
measurements.
Figure 4.3 Competition between octanoate and heptanoate for uptake by
perfused rat livers.
A, constant 1 mM octanoate + increasing heptanoate
concentration in influent perfusate. B, constant 1 mM heptanoate + increasing
octanoate concentration in influent perfusate. The vertical scale shows the
uptake of octanoate (▲) and heptanoate (△).
92
Figure 4.4 Competition between C4-ketogenesis from octanoate and C5ketogenesis from heptanoate in perfused rat livers. A, constant 1 mM octanoate
+ increasing heptanoate concentration in influent perfusate. B, constant 1 mM
heptanoate + increasing octanoate concentration in influent perfusate. The
vertical scale shows the production of C4-ketone bodies (▲) and C5-ketone
bodies (△).
Figure 4.5 Profiles of concentrations of octanoate (●) and propionate (▲) in the
effluent perfusate. The data refer to perfusions with increasing concentrations of
a single fatty acid, and are plotted as a function of the influent concentration of
each fatty acid. The dotted line is the theoretical identity of influent and effluent
concentrations.
Figure 4.6 Labeling pattern of effluent β-hydroxybutyrate (BHB) and tissue
acetyl-CoA
from
livers
perfused
with
increasing
concentrations
of
[1-
13
C]octanoate (A) or [8-13C]octanoate (B). The figures show the molar percent
enrichments (MPE) of the M1 (■) and M2 (□) mass isotopomers of BHB, the MPE
of the C-1+2 (▲) and C-3+4 (△) acetyls of BHB, and the MPE of liver acetyl-CoA
(●). The labeling patterns of ketone bodies (Figs 6, 7) were assayed in samples
of effluent perfusate taken at 14 and 18.5 min (collecting the samples over 1
min). All reported values are the means of the 14 and 18.5 min measurements.
Figure 4.7 Sharing of acetyl groups between C4- and C5-ketogenesis reflected by
the mass isotopomer distribution of BHB and BHP. A, labeling pattern of BHB in
perfusions with constant 1 mM [1-13C]heptanoate + increasing concentrations of
unlabeled octanoate.
B, labeling pattern of BHP in perfusions with constant
93
[5,6,7-13C3]heptanoate and increasing concentrations of [1-13C]octanoate. C,
labeling pattern of BHP in perfusions with constant [5,6,7-13C3]heptanoate and
increasing concentrations of [8-13C]octanoate.
D, labeling pattern of BHP in
perfusions with constant [1-13C]octanoate and increasing concentrations of
unlabeled heptanoate.
Figure 4.8 Mass isotopomer distribution of HMG-CoA (A) and HEG-CoA (B) in
livers perfused with constant 1 mM [1-13C]heptanoate and increasing
concentrations of unlabeled octanoate.
Figure 4.9 Mass isotopomer distribution of BHB-CoA and AcAc-CoA in livers
perfused with increasing concentrations of [1-13C]octanoate.
Figure 4.10 Anaplerosis and glucose labeling from increasing concentrations of
[13C3]propionate (♦) or [5,6,7-13C3]heptanoate (■,▲). The perfusions with
increasing [5,6,7-13C3]heptanoate concentrations were conducted in the absence
(▲) or presence (■) of constant 1 mM [1-13C]octanoate. A, relative anaplerosis
expressed as the m3 enrichment ratio (succinyl-CoA)/(propionyl-CoA).
absolute anaplerosis. C, M2 + M3 labeling of glucose in the effluent perfusate.
94
B,
Figure 4.1
95
Figure 4.2
96
Figure 4.3
97
Figure 4.4
98
Figure 4.5
99
Figure 4.6
100
Figure 4.7
101
Figure 4.8
102
Figure 4.9
103
Figure 4.10
104
4.3.2
Tracing the syntheses of adenine nucleotides, CoA and DNA in liver
Shuang Deng, Guo-Fang Zhang, Rajan S. Kombu, Stephanie R. Harris, David A.
DeSantis, Edwin J. Vasquez, Michelle A. Puchowicz, Vernon E. Anderson, and
Henri Brunengraber
Depts of Nutrition and Biochemistry, Case Western Reserve University,
Cleveland OH 44109
Running title: Syntheses of adenine nucleotides, CoA and DNA in liver
Corresponding author: Henri Brunengraber, Department of Nutrition, Case
Western Reserve University, 10900 Euclid Ave. WG-48; Cleveland OH 441064954. Tel: (216) 368-6548. Fax: (216) 368-6560. Email: [email protected]
105
Abstract
The syntheses of the carbon skeletons of adenine nucleotides and coenzyme A
were monitored in livers from overnight-fasted rats perfused with buffer
containing either 4 mM [13C6]glucose, 4 mM unlabeled glucose + 2 mM
[13C5]ribose,
or with buffer made up in 0 to 100% deuterated water and
containing 4 mM unlabeled glucose. With all labeled substrates, the mass
isotopomer distribution of glucose and glycolytic intermediates reveals intense
cycling and redistribution of label through glycolysis, gluconeogenesis, the citric
acid cycle and the pentose phosphate pathway. This cycling affects the
isotopomer distribution of ribose-5-phosphate, adenine nucleotides and CoA. In
the presence of 2H-enriched water, most of the 2H found in adenine nucleotides
and CoA is incorporated between ribose-5-P and AMP. Most of the turnover of
adenine nucleotides and CoA is supported by salvage pathways. The fractional
turnover rates of the adenine moiety of adenine nucleotides and of CoA are 7%/h
and 33%/h, respectively. Our findings open the way to studies on the modulation
of adenine nucleotides and CoA turnover in physiological and pathological
conditions.
106
Introduction
2
H-enriched water has been used to measure the synthesis of biomolecule since
the 1930s (1). In the course of biosyntheses, H+ from the aqueous medium
become incorporated into C-H bonds by various mechanisms: hydration,
reduction (2), splitting of C-C bonds and isotopic exchanges (3). If the aqueous
medium is 2H-enriched, the product of the synthesis becomes 2H-enriched in
some proportion to the net synthesis of the compound. This technique has been
used to trace the syntheses of fatty acids (1,4), sterols (5,6), glucose (7),
aminoacids (8), peptides (9,10), proteins (8,11-13) and nucleic acids (14,15). In
most cases, it is difficult or impossible to convert the 2H-enrichment of the
product into a rate of synthesis expressed in chemical units, e.g., μmol⋅h−1⋅kg−1.
This is because it is often impossible to predict the number of
2
H atoms
incorporated into a synthesis from the stoichiometry of the reactions of the
synthesis. In favorable cases, empirical H/C incorporation ratios can be
calculated by comparing the incorporation of labeled H (3H or 2H) and the
incorporation of labeled C (14C or
13
C) into the product. For example, the H/C
ratios for fatty acid and cholesterol synthesis from glucose were calculated from
experiments where the products were synthesized in the presence of 3H- or 2Henriched water and [U-14C]glucose (4,16). The H/C ratios are expected to be
different when lipogenesis is fueled from glucose or from acetate (17).
In most cases, relative rates of synthesis are calculated by comparing the 2Henrichment of the product at a given time with the 2H-enrichment at infinite time
(14,15). This technique has led to many studies on the syntheses of proteins (18)
107
and DNA (19), as well as to studies on the kinetics of cell proliferation in health
and disease (14,15,20).
The goal of the present study was to sort out some of the mechanisms by which
2
H from the aqueous medium becomes incorporated into C-H bonds during the
biosyntheses of adenine nucleotides, CoA and DNA in liver (Fig 4.11). To
achieve this goal, we used two experimental models with different labeling
dynamics. The first model is the isolated rat liver perfused for 2 h with
recirculating buffer
enriched with 2H (0 to 100%) (16). To test whether 2H
incorporation into adenine nucleotides and CoA reflects net synthesis vs isotopic
exchanges, we also perfused rat livers with buffer made up in unlabeled water
and containing [13C6]glucose or [13C5]ribose. We view the perfused rat liver as a
closed system where the organ is exposed to an initial amount of substrate which
is not replenished. This model allows for extensive equilibration of labeling
patterns of metabolites. Also, because of the short duration of the experiments,
metabolites salvaged from nucleic acids are not labeled. The second model is
the live rat the body fluids of which are kept chronically enriched with 2H for up to
30 days. We view the live rat as an open system where unlabeled carbon
substrates are constantly supplied by the diet. Although the 2H-enrichment of
body fluids of the live rat is kept constant, the supply of carbon substrates
prevents building blocks of biopolymers to reach the same degree of labeling as
would occur in a closed system. However, because of the long duration of the
experiments, metabolites salvaged from nucleic acids can be labeled. In the
context of this report, the turnover of adenine nucleotides and CoA is the
108
turnover of the adenosine moiety of these metabolites (synthesized via de novo
or/and salvage pathways).
Experimental Procedures
Materials.Sigma-Aldrich-Isotec supplied most chemicals, enzymes and the
following
isotopically
labeled
compounds:
2
H2O
(99.8%),
[13C6]glucose,
[13C5]ribose.
Perfused liver experiments.
Livers from overnight-fasted male rats (200-250 g) were perfused (21) with
recirculating bicarbonate buffer containing 4% dialyzed, fatty acid-free, bovine
serum albumin and 4 mM glucose. In some experiments, (i) unlabeled glucose
was replaced by [13C6]glucose, (ii) 2 mM [13C5]ribose was added to 4 mM
unlabeled glucose, and (iii) 4 mM unlabeled glucose was added to perfusate
made up in 100% 2H2O. One-half of the experiments were conducted for 2 h and
the livers were quick-frozen. In other experiments, lobes of the livers were tied
off, cut out and quick-frozen at 30, 60, 90 and 120 min. To test for interference of
high 2H-enrichment of perfusate water, a series of perfusions was conducted with
buffer made up in 0 to 100% 2H2O + 4 mM glucose.
In vivo experiments.
Nineteen rats (200 ± 10 g) received an intraperitoneal injection of normal saline
made up in 99% 2H2O in amounts calculated to achieve a 2.5% 2H-enrichment of
body water, assuming that total body water accounts for 66% of body weight.
109
The rats were fed regular chow and provided drinking water 3.25% enriched in
2
H to compensate for the production of unlabeled water from the oxidation of
foodstuffs and endogenous substrates (6). One control rat was injected with
unlabeled saline. At various times after 2H loading (2 h to 31 days) the rats were
anesthetized with isoflurane before freezing a liver lobe and sampling aortic
blood. The turnover of plasma and liver glutathione, labeled from body water was
previously reported (9).
Analytical Procedures
The concentration and mass isotopomer distribution (MID) of glucose were
assayed by GC-MS of the pentaacetate and permethyl derivatives, respectively.
Lactate and formate MID were assayed as the pentafluorobenzyl derivative
(22,23). The 2H-enrichment of water was assayed after isotopic equilibration with
unlabeled acetone in alkaline medium (24). The enrichment of deoxyribose,
isolated from DNA, was assayed as the pentaacetate derivative (15).
LC-MS analysis was used to assay the MID of CoA (25), glucose-6-P, ribose-5P, ATP, ADP, AMP, PEP and 3-P-glycerate (26). Fragmentation of CoA allowed
to calculate the labeling of following components: adenine, ribose, adenosine,
panthetheine and cysteine (27). Enzymatic assays were used for the
concentrations of CoA and adenine nucleotides.
Calculations and statistics. Measured mass isotopomer distributions were
corrected for natural enrichment as in (28). Statistical differences were assayed
by t test using the Prism software.
110
Results
In vivo experiments.
The body water of two groups of rats (6 and 19 rats) was kept 2.5% 2H-enriched
for up to 10 and 31 days, respectively (9). We assayed the M1 enrichment of liver
CoA and its building blocks (Fig. 4.12, for the 31 day group), as well as the 2Henrichment of plasma formate (Fig. 4.12, lower curve). The data of the labeling
patterns were fitted to monoexponential saturation curves (Table 4.1). The
apparent fractional turnover rates (k in day-1) were different for the CoA
components. Also, the apparent fractional turnover rates were lower in the 31
day experiment compared to the 10 day experiment. The number of 2H atoms
incorporated into CoA and its components was calculated by dividing the
extrapolated enrichment at infinite time by the 2H-enrichment of body water. This
number of 2H atoms was greater in the 31-day than in the 10-day experiment,
although no statistical significance on differences in 2H incorporation could be
calculated. The data clearly show that the labeling kinetics of CoA and its
components cannot be calculated from monoexponential saturated curves.
We assayed the 2H-enrichment of deoxyribose from DNA isolated from the livers
of the second groups of rats (body water 2.5% 2H-enriched for 10 days). Fig 4.13
shows that the labeling profile of dR-DNA was well fitted to a monoexponential
saturation curve (k = 0.17 ± 0.02; R2 = 0.98). Dividing the extrapolated dR-DNA
enrichment at infinite time (1.26%) by the 2H-enrichment of body water (2.5%)
yielded an apparent 0.5 2H atom incorporated per dR-DNA residue. Hellerstein’s
111
group had reported that the number of 2H atoms incorporated per dR-DNA
residue is 3.5 (14,15). Thus, although the monoexponential saturation curve
yielded an excellent fitting of our data, this excellent fitting does not justify the
use of monoexponential labeling kinetics.
Rat livers perfused in 100% 2H2O buffer.
Two series of livers were perfused with 4 mM unlabeled glucose in buffer made
up in 100% 2H2O. In one series, livers were frozen at 120 min. In the 2nd series,
samples of liver were taken every 30 min and quick frozen. The first series
allowed following the MID of metabolites in the perfusate modified by a constant
liver mass. The second series allowed following the MID of intracellular
metabolites in livers the mass of which decreased every 30 min. In both series,
the MID of perfusate glucose was assayed throughout the experiment. Fig 4.14
shows, for the series without biopsies, the MID of perfusate glucose
characterized by the almost linear accumulation of M1 to M7 mass isotopomers.
In the M7 mass isotopomers, all C-H bonds of glucose are labeled. The M4 to
M7 isotopomers must be labeled in the two triose moieties of glucose (29). These
mass isotopomers must be formed by a combination of glycolysis, citric acid
cycle and gluconeogenesis. The M1 to M3 mass isotopomers could be labeled in
one or both triose moieties of glucose. The most abundant mass isotopomer, M1,
corresponds to a population of 7 positional isotopomers in which one of the 7 CH groups is 2H-labeled. The most abundant of the positional isotopomers is most
likely [2-2H]glucose because 2H-labeling on C-2 of glucose-6-P occurs via both
glycogenolysis and gluconeogenesis before glucose-6-P is hydrolyzed by
112
glucose-6-Pase (7,29). In support of this interpretation, Fig 4.15 shows the
labeling patterns of glucose-6-P, ribose-5-P and PEP which show almost all
possible deuterated mass isotopomers (except for M4 of ribose-5-P which was
not detected). The MID of glucose-6-P (in plateau during the 2nd hour of the
experiment) shows a clear precursor-to-product relationship with the MID of the
large pool of perfusate glucose which becomes progressively labeled with time.
The MID of ribose-5-P is fairly stable during the 2nd hour with a predominance of
the M1 isotopomer.
Fig 4.16 shows that the M1 labeling of the 3 adenine nucleotides plateaued at
about 60% while that of CoA increased slowly over the 2 hr experiment. The data
illustrate the rapid interconversion of the 3 adenine nucleotides, and the
precursor-to-product relationship between AMP and CoA. Note that the total 2Hlabeling of AMP is much greater than that of its precursor ribose-5-P (see
Discussion). We were unable to measure the enrichment of the very small pool
of phosphoribosylpyrophosphate. As one follows the isotopomer patterns from
glucose-6-P to ribose-5-P, to adenine nucleotides, one notes the increasingly
dominance of the M1 isotopomers.
The labeling of the deoxyribose moiety of
DNA was not detected at the end of the 2 hr experiment.
To test whether in livers perfused with 100% 2H2O, the syntheses of adenine
nucleotides and CoA could be inhibited by the high 2H-enrichment, we conducted
perfusions with buffer made up in 0 to 100% 2H2O + 4 mM unlabeled glucose. Fig
4.17 shows that the labeling of adenine nucleotides and of CoA increased almost
linearly with the 2H-enrichment of perfusate water. Most of the label in CoA was
113
in the ribose moiety. Very low labeling was detected in the adenine moiety. Small
fractions of CoA labeling were found on the pantetheine moiety (Fig 4.18).
Because all labeled glucose isotopomers were released in the perfusate via
glucose-6-P and glucose-6-phosphatase, we used 2 calculations to estimate a
minimal flux through this enzyme in intact perfused livers. The first calculation is
based on the accumulation of labeled glucose isotopomers in the perfusate:
Flux1 = [(total mol fraction of M1 to M7 glucose isotopomers at 120 min)(total
glucose in perfusate)]/[(liver dry weight)(120 min)]. This calculation does not
include the release of unlabeled glucose from the unlabeled fraction of glucose6-P. The second calculation uses the precursor-to-product relationship between
the M1 enrichments of glucose-6-P and glucose:
Flux2 = [(mol fraction of M1 glucose at 120 min)(total glucose in perfusate)]/[(mol
fraction of M1 G6P)(liver dry weight)(120 min)]
The second calculation was conducted using the most abundant isotopomers of
glucose and glucose-6-P, i.e., M1. The two calculated fluxes are 0.55 ± 0.01 and
0.60 ± 0.1 μmol⋅min−1⋅(g dry wt)−1, respectively (SE, n = 5). These are minimal
fluxes because some labeled glucose molecules must have cycled more than
once through the glucose ➔ glucose-6-P ➔ glucose cycle. These values are
similar to those reported by Katz et al (30) who measured glucose-6-P
dephosphorylation
in
hepatocytes
incubated
with
[2-3H]glucose
[26
μmol⋅hr−1⋅(100 mg protein)−1 which is equivalent to 0.43 μmol⋅min−1⋅(g dry wt)−1].
114
Rat livers perfused with [13C6]glucose.
Two series of livers were perfused with 4 mM [13C6]glucose in buffer made up in
normal water. In one series, livers were frozen at 120 min. In the 2nd series,
samples of liver were taken every 30 min and quick frozen. Fig 4.19 shows the
isotopomer distribution of perfusate glucose. Because the enrichment of
commercial [13C6]glucose is 99% for each carbon, the M6 and M5 enrichments of
perfusate glucose at zero time were 95 and 3%, respectively as expected. The
M6 enrichment of glucose decreased from 95 to 78% over 2 h. During the same
period, the proportion of unlabeled glucose (M) in the perfusate increased from
almost zero to 8%. This reflects the dilution of labeled perfusate glucose by
unlabeled glucose derived from glycogenolysis. In parallel
glucose became
slightly enriched with M1 to M4 isotopomers. These variations in isotopomer
distribution of glucose reflect the cycling between perfusate glucose and
glycolytic, gluconeogenic, pentose phosphate pathway and citric acid cycle
intermediates. This interpretation is supported by the MIDs of glucose-6-P,
ribose-5-P and glycerate-3-P (Fig 4.20). The total proportion of labeled mass
isotopomers of glucose-6-P (M1 to M6) is about 50% during the 2nd hour of the
experiment. Thus, about one-half of the glucose-6-P derived from unlabeled
glycogen. The presence of M1 to M5 mass isotopomers of glucose-6-P results
from the loss of label from [13C6]glucose in (i) the citric acid cycle, and (ii) the
pyruvate ➔ oxaloacetate ➔ PEP ➔ pyruvate cycle, before labeled triose units
are recombined into glucose-6-P. Indeed, the MID of glycerate-3-P shows similar
proportions of the M1 to M3 mass isotopomers (Fig 4.20C).
115
During the last 30 min of the perfusions, the labeling pattern of ribose-5-P (Fig
4.20) was fairly stable, with a total sum of labeled mass isotopomers (M1 to M5)
of about 22%. This labeled ribose-5-P is the precursor for the synthesis of
adenine nucleotides. At 120 min, the total labeling of AMP and ATP (Fig 4.21A,
4.21B) was about 23% of the total labeling of ribose-5-P. Because ATP is the
precursor of CoA synthesis, the total labeling of CoA (Fig 4.21C) at 120 min was
lower (3%) than the total labeling of ATP, as expected from the precursor-toproduct relationship.
Rat livers perfused with unlabeled glucose + [13C5]ribose.
One series of livers were perfused with 4 mM unlabeled glucose + 2 mM
[13C5]ribose in buffer made up in normal water. Samples of liver were taken every
30 min and quick frozen. Fig 4.22 shows the mass isotopomer distribution of
perfusate glucose which becomes progressively labeled with all possible mass
isotopomers (M1 to M6).
Fig 4.23B shows the isotopomer distribution of ribose-5-P, the first metabolite
labeled from [13C5]ribose. The main isotopomer is M5 the enrichment of which
stabilizes at about 33% during the 2nd hour of the perfusion. The M1 to M4
isotopomers, which account for about 10% of all ribose-5-P, derive most likely
from the reversible reactions of the non-oxidative branch of the pentose
phosphate pathway. The total percentage of the labeled isotopomers of ribose-5P is about 45%. Thus, slightly more than half of ribose-5-P was unlabeled.
116
The two trioses, glycerate-3-P and PEP are M1 to M3 labeled, with M3 being the
main mass isotopomer. Glucose-6-P shows M1 to M3 and M6 mass
isotopomers. Clearly, the mass isotopomer distribution of glucose-6-P results
from the combination of labeled triose units. The labeling pattern of glucose-6-P
explains the accumulation of the M1 to M6 isotopomers of glucose.
Discussion
In vivo experiments
When a metabolite becomes labeled with a time profile suggesting a saturation
curve, it is tempting to fit the label vs time data to a monoexponential saturation
curve. This is justified when the metabolite becomes labeled in a single net
synthetic process (not by isotopic exchange) from a precursor with constant
isotopic enrichment. During the synthesis of complex molecules made of different
building blocks, e.g., proteins, adenine nucleotides, CoA or DNA, each building
block can become labeled at single or multiple sites. Our data show that when
liver CoA becomes labeled from 2H-enriched body water, each of the building
blocks of CoA becomes labeled with different kinetics (Fig 4.22, Table 4.1). Also,
the computed kinetics parameters change with the duration of the experiment
(Table 4.1). In addition, when conducting long-term in vivo experiments with 2Henriched body water, one must realize that the experimental model is open.
Although the 2H-enrichment of body water can easily be kept constant by labeling
drinking water, the constant supply of unlabeled substrates from the diet dilutes
the labeling of intermediates and impacts on the kinetics. Lastly, when the
117
synthesis of the complex molecule uses a combination of de novo synthesis and
salvage pathways, the incorporation of label into the product via some of the
salvage pathways can be delayed until the intermediates of the salvage
pathways become labeled from the degradation of the labeled polymer that
accumulated in the early phase of the experiments. Thus, kinetic parameters of
complex molecules should be interpreted with caution. This was recognized by
Hellerstein who pioneered the use of
2
H-enriched water to monitor DNA
synthesis and cell turnover (14,15). Hellerstein proposed to calculate fractional
cell division by the
2
H labeling ratio (deoxyribose in DNA of cells of
interest)/(deoxyribose in DNA of fully-turnovered reference cells). This ratio
provides a useful yardstick for estimating cell turnover under physiological and
pathological conditions. Of particular interest is the use of this ratio to monitor the
effectiveness of the treatment of malignancies (20,31).
Liver perfusion experiments.
We recently found that, in rat livers perfused with buffer made in 100% 2H2O,
CoA was substantially labeled after 2 hours. Because CoA, adenine nucleotides
and nucleic acids share the same early steps of synthesis via de novo and
salvage pathways, we decided to examine in detail the dynamics of adenine
nucleotides and CoA labeling from
2
H-enriched water and from relevant
[13C]substrates (glucose and ribose). We selected the isolated liver as a closed
system with a limited pool of substrates. In this model, we followed the labeling
patterns of the precursors of adenine nucleotides and CoA synthesis. This led to
estimates of fractional turnover rates of these compounds.
118
The time profiles of the mass isotopomer distribution of perfusate glucose,
labeled from any of 3 different isotopic substrates (2H2O, [13C6]glucose or
[13C5]ribose) provide evidence of constant glycolysis and gluconeogenesis. This,
in spite of the near constancy of the glucose concentration in the perfusate (4
mM at zero time and 3.7 mM at 120 min). The presence of all possible labeled
glucose mass isotopomers (M1 to M7 from
2
H2O and M1 to M6 from
[13C6]glucose or [13C5]ribose) demonstrates that the labeling pattern of trioses is
modified by reactions of the citric acid cycle and of the substrate cycle pyruvate
➔ oxaloacetate ➔ PEP ➔ pyruvate. For example, in livers perfused with
[13C6]glucose, trioses which must be initially M3, become M1 + M2 + M3 labeled
via citric acid cycle and pyruvate cycle reactions (Fig 4.20). In citric acid cycle
intermediates, exchanges of
13
C for
12
C result from (i) reversibility of
decarboxylating reactions (32) and (ii) randomization of the labeling of
oxaloacetate
via
the
rapidly
reversible
reactions
catalyzed
by
malate
dehydrogenase and fumarase (33). Recombination of two trioses with M to M3
mass isotopomers yields glucose molecules that are M to M6. Although the
release of unlabeled (M) glucose is not visible from the mass isotopomer
distribution of perfusate glucose (Fig 4.19), it is inferred from the fraction of
glucose-6-P that is unlabeled (about 55%, Fig 4.19). A similar rationale can be
formulated for the mass isotopomer distribution of glucose (i) in livers perfused in
100% 2H2O buffer, or (ii) in livers perfused with regular buffer + 4 mM unlabeled
glucose + 2 mM [13C5]ribose.
119
The re-distribution of label [13C6]glucose or [13C5]ribose administered in vivo or
added to in vitro preparations had been described previously. The mass
isotopomer distribution of glucose labeled from highly 2H-enriched water has not
been previously described to our best review of the literature. We followed the
mass isotopomer distribution of glucose and related compounds to assess
whether the MID of ribose-5-P would be sufficiently stable to be used to calculate
the turnover of adenine nucleotides and of CoA. Our data show that the labeling
pattern of ribose-5-P became fairly constant after 30 min of perfusion under
conditions when label reaches ribose-5-P from
2
H2O, [13C6]glucose or
[13C5]ribose (Figs 4.15, 4.20, 4.23).
In perfusions conducted with [13C6]glucose,
the comparison of the MIDs of
glucose-6-P and ribose-5-P (Figs 4.20A and 4.20B) provides information on the
conditions of operation of the pentose phosphate pathway in these livers. The M6
enrichment of glucose-6-P (30 - 35% after 90 min) is much higher that the M5
enrichment of ribose-5-P (4 - 4.5%). If all ribose-5-P had been formed in the
oxidative branch of the pentose phosphate pathway, the M5 enrichment of
ribose-5-P should be close to the M6 enrichment of glucose-6-P. This is because
the removal of C-1 of [13C6]glucose-6-P would result in the formation of a M5
pentose-P. This was not the case. Therefore, in these livers, most of the ribose5-P was formed in the reversal of the non-oxidative branch of the pentose
phosphate pathway. The reversal of this branch involves multiple combinations
between intermediates with 3, 4, 5, 6 and 7 carbons. The probability that these
multiple combinations will yield M5 ribose-5-P is low. This explains the low M5
120
enrichment of ribose-5-P. In these livers from overnight-fasted rats perfused with
a low glucose concentration (4 mM), rates of fatty acid synthesis are low: about
15% of rates measured in livers from fed rats perfused with a high glucose
concentration (21). Therefore, in the livers used in the present study, the
requirement for NADPH production via the pentose phosphate pathway was low.
This explains why the MID of ribose-5-P is compatible with it being formed mostly
by the reversal of the non-oxidative branch of the pentose phosphate pathway.
In livers perfused with 100% 2H2O buffer, the MIDs of ribose-5-P and of AMP
(Figs 4.15, 4.16) were not in an usual precursor-to-product relationship where the
labeling of the product would be lower than that of the precursor. In this case, the
much higher M1 labeling of AMP compared to ribose-5-P shows that one 2H
atom from the buffer was incorporated between ribose-5-P and AMP. Multiple
reaction monitoring of the mass spectrum of AMP revealed that most of the 2Hlabeling was in the ribose moiety. We were unable to detect PRPP and measure
its labeling. To interpret the difference in the MIDs of ribose-5-P and AMP, one
must consider the different sources of the adenosine nucleus of AMP. Note that
the 3 adenine nucleotides had almost identical labeling patterns, reflecting their
rapid interconversion via ADP re-phosphorylation and the adenylate kinase
reaction. The relevant sources of AMP in short-term experiments where DNA and
RNA are unlabeled are: (i) de novo synthesis with construction of the adenine
moiety,
(ii) salvage of adenine from RNA and DNA breakdown followed by
reaction with PRPP catalyzed by adenine phosphoribosyl transferase, and (iii)
salvage of nucleoside (adenine-ribose) from the breakdown of RNA.
121
The de novo synthesis of AMP with construction of the adenine moiety involves
multiple steps of 2H incorporation. First, one 2H atom is incorporated at the
inversion of configuration between α-PRPP and 5-phospho-β-D-ribosylamine.
Second, two 2H are incorporated via two reactions involving N10-formyl H4 folate
derived from formate. Thus, in a liver perfused with 100% 2H2O buffer, the de
novo synthesis of AMP with construction of the adenine moiety should yield AMP
molecules labeled on the ribose moiety and the adenine moiety. However, in our
experiments, the labeling of AMP showed only small proportions of mass
isotopomers heavier than M1.
The synthesis of AMP via the salvage of adenine from RNA and DNA breakdown
followed by reaction with PRPP catalyzed by APRT should yield only M1 AMP
when the perfusion is conducted in 100% 2H2O buffer. This is because of the
inversion of configuration of C-1 of the ribose moiety of α-PRPP when it is
converted to adenosine. Lastly, the synthesis of AMP by phosphorylation of
adenosine (adenine-ribose) should yield unlabeled AMP. This is because, in
short-term experiments, adenosine derived from RNA breakdown is not labeled.
Because the M1 enrichment of AMP remains stable at about 60% during the 2nd
hour of the experiment, about 40% of the AMP produced was unlabeled, and
presumably was formed via phosphorylation of salvaged adenosine. The 60%
M1 AMP was made from ribose-5-P which was about 15% M1 labeled, probably
mostly via reaction of PRPP with adenine. It is not clear whether the 15% M1
enrichment of ribose-5-P is a component of the 60% M1 enrichment of AMP. This
is because we do not know the distribution of 2H-labeling between the carbons of
122
ribose-5-P. Because the M2/M5 abundance ratio is equally low in ribose-5-P and
in AMP, it is likely that very little if any AMP was synthesized with building up the
adenine nucleus during the 2 h of the experiment.
In perfusions with 2 mM [13C5]ribose, the main isotopomer of ribose-5-P is M5
(about 32% during the 2nd hour (Fig 4.23). One would expect that in the presence
of 2 mM [13C5]ribose, most ribose-5-P would derive from the phosphorylation of
[13C5]ribose by ribokinase. In fact, about 60% of ribose-5-P was unlabeled. This,
and the dominance of the M5 isotopomer of ribose-5-P compared with the
multiple mass isotopomer distribution of glucose-6-P and of gluconeogenic
intermediates (PEP, glycerate-3-P) suggests that little label reached ribose-5-P
via the two branches of the pentose phosphate pathway. This suggests that most
unlabeled ribose-5-P arises by processes that would not be fed back by a high
concentration of ribose-5-P. A likely source of unlabeled ribose-5-P is the
nucleoside phosphorylases which form nucleobases and ribose-1-P. The latter is
converted to ribose-5-P by phosphopentomutase (34).
The action of nucleosidase on nucleosides derived from RNA degradation
releases free unlabeled ribose which, after phosphorylation, somewhat dilutes
the M5 enrichment of exogenous ribose-5-P. Overall, our data suggest that, in
perfusions with [13C5]ribose, the bulk of ribose-5-P arises from the salvage
pathways
of
RNA
degradation
via
nucleoside
phosphorylases
and
nucleosidases. This is also likely in the absence of exogenous ribose since in the
presence of 2H2O or [13C6]glucose, most ribose-5-P is unlabeled.
123
In perfusions in 100% 2H2O buffer, the stability of the M1 enrichment of adenine
nucleotides (60%, Fig 4.16 A-C) and the fairly linear increase in M1 labeling of
CoA (0.56%/min from 60 to 120 min, Fig 4.16 D) allowed calculating the
fractional turnover rate of CoA of 0.93%/min or 56%/h. Because free CoA (the
actual analyte) exchanges probably rapidly with all CoA esters, the fractional
turnover rate of 56%/h probably applies to the whole CoA pool (free and
esterified).
In perfusions with 4 mM unlabeled glucose + 2 mM [13C5]ribose, the stability of
the M5 enrichment of ribose-5-P (33%, Fig 4.23) and the fairly linear increase in
M5 labeling of AMP (2.2%/h, Fig 4.24) allowed calculating a fractional turnover
rate of AMP of 6.7%/h). This fractional turnover rate applies to the three adenine
nucleotides which are in isotopic equilibrium (Fig 4.24).
In conclusion, our study revealed the rapid turnover of the nucleus of adenine
nucleotides and of CoA in rat liver. It appears that most of these turnover are
fueled by salvage pathways that recycle purine bases and nucleosides. Our
findings open the way to studies on the modulation of adenine nucleotides and
CoA turnover in physiological and pathological conditions.
124
Abbreviations: GC-MS, gas chromatography-mass spectrometry; LC-MS, liquid
chromatography-mass spectrometry; MID, mass isotopomer distribution
Footnote.
1. Mass isotopomers are designated as M, M1, M2,...Mn where n is the number
of heavy atoms in the molecule. In the context of this paper, mass isotopomers
are refered to as “isotopomers”. The isotopic enrichment of each isotopomer is
expressed as mol percent.
Acknowledgments
This work was supported by the NIH (Roadmap grant R33DK070291 and grant
R01ES013925.). We thank the Case Mouse Metabolic Phenotyping Center for
help with the in vivo experiments.
Reference List
1. Rittenberg, D. and Schoenheimer, R. (1937) J. Biol. Chem. 117, 485-490
2. Levy, L. M. and Betts, G. F. (1989) Biochim. Biophys. Acta 997, 331-333
3. Babu, U. M. and Johnston, R. B. (1974) Biochem. Biophys. Res. Commun.
58, 460-466
125
4. Jungas, R. L. (1968) Biochemistry 7, 3708-3717
5. Wong, W. W., Hachey, D. L., Feste, A., Leggitt, J., Clarke, L. L., Pond, W
G., and Klein, P. D. (1991) J Lipid Res. 32, 1049-1056
6. Lee, W. N., Bassilian, S., Ajie, H. O., Schoeller, D. A., Edmond, J., Bergner,
E. A., and Byerley, L. O. (1994) Am. J Physiol 266, E699-E708
7. Landau, B. R., Wahren, J., Chandramouli, V., Schumann, W. C., Ekberg,
K., and Kalhan, S. C. (1996) J Clin Invest 98, 378-385
8. Previs, S. F., Fatica, R., Chandramouli, V., Alexander, J. C., Brunengraber,
H., and Landau, B. R. (2004) Quantifying rates of protein synthesis in humans by
use of 2H2O: application to patients with end-stage renal disease.
9. Kombu, R. S., Zhang, G. F., Abbas, R., Mieyal, J. J., Anderson, V. E.,
Kelleher, J. K., Sanabria, J. R., and Brunengraber, H. (2009) Am. J Physiol
Endocrinol. Metab 297, E260-E269
10. Wang, B., Sun, G., Anderson, D. R., Jia, M., Previs, S., and Anderson, V. E.
(2007) Isotopologue distributions of peptide product ions by tandem mass
spectrometry: quantitation of low levels of deuterium incorporation.
11. Busch, R., Kim, Y. K., Neese, R. A., Schade-Serin, V., Collins, M., Awada,
M., Gardner, J. L., Beysen, C., Marino, M. E., Misell, L. M., and Hellerstein, M. K.
(2006) Biochim. Biophys. Acta 1760, 730-744
12. Xiao, G. G., Garg, M., Lim, S., Wong, D., Go, V. L., and Lee, W. N. (2008) J
Appl. Physiol 104, 828-836
13. Dufner, D. A., Bederman, I. R., Brunengraber, D. Z., Rachdaoui, N., IsmailBeigi, F., Siegfried, B. A., Kimball, S. R., and Previs, S. F. (2005) Am. J Physiol
Endocrinol. Metab 288, E1277-E1283
14. Neese, R. A., Misell, L. M., Turner, S., Chu, A., Kim, J., Cesar, D., Hoh, R.,
Antelo, F., Strawford, A., McCune, J. M., Christiansen, M., and Hellerstein, M. K.
(2002) Proc Natl Acad Sci U S A 99, 15345-15350
15. Voogt, J. N., Awada, M., Murphy, E. J., Hayes, G. M., Busch, R., and
Hellerstein, M. K. (2007) Nat. Protoc. 2, 3058-3062
16. Wadke, M., Brunengraber, H., Lowenstein, J. M., Dolhun, J. J., and
Arsenault, G. P. (1973) Biochemistry 12, 2619-2624
17. Endemann, G., Goetz, P. G., Tomera, J. F., Rand, W. M., Desrochers, S.,
and Brunengraber, H. (1987) Biochem Cell Biol. 65, 989-996
126
18. Yuan, C. L., Sharma, N., Gilge, D. A., Stanley, W. C., Li, Y., Hatzoglou, M.,
and Previs, S. F. (2008) Am. J Physiol Endocrinol. Metab 295, E216-E222
19. Collins, M. L., Eng, S., Hoh, R., and Hellerstein, M. K. (2003) J Appl. Physiol
94, 2203-2211
20. Mohri, H., Perelson, A. S., Tung, K., Ribeiro, R. M., Ramratnam, B.,
Markowitz, M., Kost, R., Hurley, A., Weinberger, L., Cesar, D., Hellerstein, M. K.,
and Ho, D. D. (2001) J Exp. Med. 194, 1277-1287
21. Brunengraber, H., Boutry, M., and Lowenstein, J. M. (1973) Fatty acid and
3-beta-hydroxysterol synthesis in the perfused rat liver.
22. Hazey, J. W., Yang, D., Powers, L., Previs, S. F., David, F., Beaulieu, A. D.,
Puchowicz, M. A., Potter, J. L., Palmquist, D. L., and Brunengraber, H. (1997)
Anal Biochem 248, 158-167
23. Tomcik, K., Ibarra, R. A., Sadhukhan, S., Han, Y., Tochtrop, G. P., and
Zhang, G. F. (2011) Anal Biochem 410, 110-117
24. Yang, D., Diraison, F., Beylot, M., Brunengraber, D. Z., Samols, M. A.,
Anderson, V. E., and Brunengraber, H. (1998) Anal Biochem 258, 315-321
25. Zhang, G. F., Kombu, R. S., Kasumov, T., Han, Y., Sadhukhan, S., Zhang,
J., Sayre, L. M., Ray, D., Gibson, K. M., Anderson, V. A., Tochtrop, G. P., and
Brunengraber, H. (2009) J Biol. Chem 284, 33521-33534
26. Yang, W. C., Sedlak, M., Regnier, F. E., Mosier, N., Ho, N., and Adamec, J.
(2008) Anal. Chem 80, 9508-9516
27. Dalluge, J. J., Gort, S., Hobson, R., Selifonova, O., Amore, F., and Gokarn,
R. (2002) Anal. Bioanal. Chem. 374, 835-840
28. Fernandez, C. A., Des Rosiers, C., Previs, S. F., David, F., and
Brunengraber, H. (1996) J Mass Spectrom. 31, 255-262
29. Chandramouli, V., Ekberg, K., Schumann, W. C., Wahren, J., and Landau,
B. R. (1999) Am. J Physiol 277, E717-E723
30. Katz, J., Wals, P. A., and Rognstad, R. (1978) J Biol Chem 253, 4530-4536
31. Kim, S. J., Turner, S., Killion, S., and Hellerstein, M. K. (2005) Biochem.
Biophys. Res. Commun. 331, 203-209
32. Yang, L., Kasumov, T., Kombu, R. S., Zhu, S. H., Cendrowski, A. V., David,
F., Anderson, V. E., Kelleher, J. K., and Brunengraber, H. (2008) Metabolomic
and mass isotopomer analysis of liver gluconeogenesis and citric acid cycle: II.
Heterogeneity of metabolite labeling pattern.
127
33. Magnusson, I., Schumann, W. C., Bartsch, G. E., Chandramouli, V.,
Kumaran, K., Wahren, J., and Landau, B. R. (1991) J. Biol. Chem. 266, 69756984
34. Tozzi, M. G., Camici, M., Mascia, L., Sgarrella, F., and Ipata, P. L. (2006)
FEBS J 273, 1089-1101
128
Table 4.1. Apparent kinetics of the labeling of CoA and its components from 2Henriched body water in rats over 10 and 31 days
2H
( 10 d)
2H
( 31 d)
k (10 d)
k (31 d)
Adenine
0.24 ± 0.1
0.19 ± 0.03
1.1
1.2
Ribose
0.46 ± 0.08
0.41 ± 0.07
2.7
3.7
Nucleotide
0.79 ± 0.18
0.32 ± 0.04
3.8
4.9
Pantetheine
0.43 ± 0.06
0.41 ± 0.06
2.7
3.1
β-Alanine cysteamine
0.72 ± 0.23
0.33 ± 0.05
1.5
2
Whole CoA
0.45 ± 0.06
0.23 ± 0.04
6.6
8.5
E (t) = E inf (1- e-kt), k: rate constant (day-1), number of 2H = Einf % /2.5 %
129
Figure legends
Figure 4.11. Sites of labeling of nucleotides, CoA and DNA from [2H]water.
Figure 4.12. In vivo labeling of liver CoA from [2H]water over 31 days.
Figure 4.13. In vivo labeling of liver dR-DNA [2H]water over 10 days
Figure 4.14 Mass isotopomer distribution of liver perfusate glucose in 100% 2H2O
buffer.
Figure 4.15. Mass isotopomer distribution of glucose-6-P, ribose-5-P and PEP in
livers perfused with 4 mM unlabeled glucose in 100% 2H2O buffer.
Figure 4.16. M1 enrichment of AMP, ADP, ATP and CoA in livers perfused with 4
mM unlabeled glucose in 100% 2H2O buffer.
Figure 4.17. M1 enrichment of AMP, ADP and ATP in livers perfused with 4 mM
unlabeled glucose in buffer enriched 0 to 100% with 2H2O.
Figure 4.18. M1 enrichment of CoA and its components in livers perfused with 4
mM unlabeled glucose in buffer enriched 0 to 100% with 2H2O.
Figure 4.19. Mass isotopomer distribution of perfusate glucose in liver perfusion
experiments starting with 4 mM [13C6]glucose.
Figure 4.20. Mass isotopomer distribution of glucose-6-P, ribose-5-P and
glycerate-3-P from livers perfused with 4 mM [13C6]glucose.
130
Figure 4.21. Mass isotopomer distribution of AMP, ATP and CoA in livers
perfused with 4 mM [13C6]glucose.
Figure 4.22. Mass isotopomer distribution of perfusate glucose in liver perfusion
experiments starting with 4 mM unlableled glucose + 2 mM [13C5]ribose.
Figure 4.23. Mass isotopomer distribution of glucose-6-P, ribose-5-P, PEP and
glycerate-3-P from livers perfused with 4 mM unlableled glucose + 2 mM
[13C5]ribose.
Figure 4.24. Mass isotopomer distribution of AMP, ADP, ATP and CoA in livers
perfused with 4 mM unlableled glucose + 2 mM [13C5]ribose.
131
Figure 4.11
132
Figure 4.12
2
Figure 4.13
134
Figure 4.14
135
Figure 4.15
136
Figure 4.16
A
B
AMP
70
ADP
70
60
50
M1
40
M2
30
M3
20
M4
10
M5
Mass Isotopomer
Distribution (%)
Mass Isotopomer
Distribution (%)
60
0
M1
M2
M3
M4
M5
50
40
30
20
10
0
0
30
60
90
120
0
30
Time (min)
Mass Isotopomer
Distribution (%)
D
ATP
70
90
120
Time (min)
CoA
70
60
M1
M3
M4
M5
50
40
30
20
10
Mass Isotopomer
Distribution (%)
C
60
0
M1
M3
M4
M5
60
50
40
30
20
10
0
0
30
60
Time (min)
90
0
120
137
30
60
Time (min)
90
120
Figure 4.17
80
A
70
AMP
60
Mass Isotopomer Distribution (%)
50
40
M1
30
M2
20
M3
10
M4
0
0
80
20
B
70
40
60
80
100
ADP
60
50
M1
40
M3
30
M4
20
M5
10
0
0
80
20
60
80
100
ATP
C
70
40
60
50
40
M1
30
M3
20
M4
10
0
0
20
40
60
2H-enrichment
80
100
(%)
138
Figure 4.18
Mass Isotopomer Distribution (%)
50
45
40
M1 whole CoA
35
M1
nucleotide
30
25
M1 ribose
20
15
10
M1
adenine
M1 pantetheine
5
β-alanine+
cysteamine
0
0
20
40
60
2H-enrichment of
perfusate (%)
139
80
100
Figure 4.19
100
Mass Isotopomer Distribution (%)
90
80
70
60
50
M1
40
M3
M2
M4
30
M5
20
M6
M0
10
0
0
15
30
45
60
Time (min)
140
75
90
105
120
Figure 4.20
141
Figure 4.21
142
Figure 4.22
143
Figure 4.23
G6P labeling
R5P labeling
M1
20
M2
10
M3
0
0
30
60
90
120
MID ( %)
MID (%)
30
M6
50
40
30
20
10
0
M1
M2
M3
M4
M5
0
30
Time (min)
M1
20
M2
10
M3
0
90
30
MID ( %)
MID (%)
30
60
120
Glycerate-3-P labeling
40
30
90
Time (min)
PEP labeling
0
60
20
M1
10
M2
0
120
0
Time (min)
30
60
Time (min)
144
90
120
M3
Figure 4.24
5
Mass Isotopomer
Distribution (%)
6
AMP labeling
4
M3
3
M4
2
M5
1
Mass Isotopomer
Distribution (%)
6
ADP labeling
4
M3
M4
2
M5
0
0
0
30
60
90
0
120
30
90
120
Time (min)
Time (min)
6
60
ATP labeling
CoA labeling
Mass Isotopomer
Distribution (%)
6
4
M2
4
M3
M3
M4
M4
2
2
M5
M5
0
0
0
30
60
Time (min)
90
120
0
30
60
Time (min)
145
90
120
CHAPTER 5: IMPLICATIONS AND FUTURE DIRECTIONS
5.1 Project 1: C4- and C5-ketogenesis in rat liver
5.1.1 Results and discussion
First, I studied the relationship between the uptake of C8, C7 and C3 fatty acids
and the formation of C4- and C5-ketone bodies. The uptake of octanoate (Fig.
4.2A) was very similar to the uptake of heptanoate (Fig.4.2B). Octanoate
catabolism only yields C4-ketone bodies. The yield of C4-ketone bodies from
exogenous octanoate ranged from 80 to 90%. Acetyl-CoA derived from
heptanoate is used to form both C5- and C4-ketone bodies. Only 40% of the
propionyl moiety of heptanoate was converted to C5-ketone bodies. This results
from the diversion of propionyl-CoA to anaplerosis and gluconeogenesis
(Fig.4.1). Propionate is taken up by rat liver (Fig. 4.2C). However the virtually nil
C5-ketogenesis from propionate demonstrates that, in the intact liver, thiolase
does not allow the formation of BKP-CoA from propionyl-CoA + acetyl-CoA
(Fig.4.1). I also perfused livers with two fatty acids to test the interaction between
the uptakes of octanoate and heptanoate as well as the formation of C4- and C5ketone bodies. The data show that the competition favors octanoate uptake over
heptanoate uptake. The rate of C5-ketogenesis from heptanoate is much lower
than the rate of C4-ketogenesis from octanoate (Figs 4.3 and 4.4).
Second, I assayed the labeling pattern of ketone bodies from livers perfused with
increasing concentrations of [1-13C]octanoate (Fig.4.6A), or [8-13C]octanoate
(Fig.4.6B). In the presence of [1-13C]octanoate, most of the labeling of BHB was
on the C-1+2 acetyl, with relatively little label on the C-3+4 acetyl. The opposite
146
labeling was observed in perfusions with [8-13C]octanoate. This results from the
fact that the C-3+4 acetyl of BHB derives from the C-7+8 acetyl of octanoate via
the C-3+4 acetyl of AcAc-CoA.
Third, the data shows that C4-ketogenesis and C5-ketogenesis share a pool of
acetyl-CoA. The labeling patterns of ketone bodies, HMG-CoA and HEG-CoA
show that when octanoate and heptanoate are used by the liver, acetyl-CoA
derived from each substrate is available to both C4- and C5-ketogenesis.
Fourth, I demonstrated the reversibility of the BHB-CoA dehydrogenase reaction
in the intact liver. Initially unlabeled BHB-CoA and AcAc-CoA derived from βoxidation of [1-13C]octanoate become M1- and M2-labeled. This labeling results
from the partial isotopic equilibration of AcAc-CoA and acetyl-CoA via 3-ketoacylCoA thiolase (Fig. 4.1, reaction 1). The M1 and M2 labeling of BHB-CoA (Fig.
4.9) show that the BHB-CoA dehydrogenase and the thiolase reactions are
reversible in the intact liver despite the high rate of β-oxidation.
Fifth, the data showed that heptanoate and propionate are anaplerotic
substrates. Relative anaplerosis from [13C3]propionate increased faster with
substrate concentration than relative anaplerosis from [5,6,7-13C3]heptanote
(Fig.4.10A). This is because propionyl-moeity from heptanoate also goes to C5ketogenesis. The opposite occurred with absolute anaplerosis (Fig.4.10B). This
probably results in a faster CAC flux in the presence of [5,6,7-13C3]heptanote
(which supplies 2 acetyl-CoA) than in the presence of [13C3]propionate (which
supplies no acetyl-CoA). Absolute anaplerosis from [5,6,7-13C3]heptanoate was
significantly decreased in the presence of octanoate (Fig. 4.10B). This results
147
from the inhibition of heptanoate uptake by octanoate (Fig. 4.2B and 4.3A). Quite
striking is the inhibition of labeling of glucose from [5,6,7-13C3]heptanote by
unlabeled octanoate (Fig. 4.10C). This may results from the increase in the
[NADH]/[NAD+] ratios induced by the rapid oxidation of octanoate, thus inhibiting
gluconeogenesis.
Concluding remarks
overall, my data explored the interrelation between C4-
ketogenesis, C5-ketogenesis, and anaplerosis. In the presence of octanoate and
heptanoate, the uptake of octanoate and C4-ketogenesis prevail over the uptake
of heptanoate and C5-ketogenesis. The virtual absence of C5-ketogenesis from
propionate demonstrates that the mitochondrial 3-ketoacyl-CoA thiolase does not
allow a net acyl-CoA condensation flux. These findings have clinical implications:
i) the concentration of C5-ketone bodies in body fluids of patients with disorders
of the propionyl-CoA pathway are formed via β-oxidation of odd-long-chain fatty
acids synthesized from propionyl-CoA, and ii) the inhibition of anaplerosis and
gluconeogenesis from heptanoate by octanoate has implications for the design of
diets for the treatment of some metabolic diseases such as disorders of longchain fatty acid oxidation. It would not be advisable to progressively modify the
patients’ diet to replace trioctanoin by triheptanoin because octanoate inhibits
heptanoate uptake (Fig. 4.3B), C5-ketogenesis from heptanoate(Fig. 4.4B), and
anaplerosis from heptanoate (Fig 4.10 A and B).
5.1.1 Future directions
The present study shows that odd-chain fatty acids are anaplerotic substrates
and even-chain fatty acids inhibit the uptake of odd-chain fatty acids in isolated
148
rat liver. Based on this study, I expect that odd-chain fatty acids could be used
alone for the dietary treatment of patients with FODs.
The decompensated patients with FODs have severe clinical symptoms under
stress such as fasting, fever, infection or trauma. Those stresses stimulate
catecholamine secretion and lipolysis. They also inhibit insulin secretion. This
causes the patients’ shock or death. The traditional acute treatments for the
decompensated patients involves infusing glucose and insulin to provide energy
substrates. This treatment is not always effective. The use of triheptanoin is a
good addition to the treatment of patients with FODs.
Several knockout mice models with FODs have been developed. These models
include inherited deficiencies in VLCAD (288; 289), LCAD (290), SCAD (291)
and so on. I would choose VLCAD-/- mice as the animal model. Put the mice
under stress conditions by fasting or cold exposure.
The following treatment will be used to evaluate the recovery of VLCAD-/- mice
from stress conditions: i) infusion of triheptanoin, ii) dietary treatment with
triheptanoin alone, iii) dietary treatment with trioctanoin alone, iv) mix of
triheptanoin and trioctanoin, V) possibly treatment with insulin and/or high
glucose diet. I hypothesize that the administration of triheptanoin alone will most
improve VLCAD -/- mice conditions under stress. Tissues including the liver,
heart, muscle and brain will be isolated from mice for the analyses. C4- and C5ketone bodies, glucose will be analyzed by GC-MS. Free CoA and acyl-CoA
esters will be analyzed by HPLC-MS.
149
5.2 Project 2: Tracing the syntheses of adenine nucleotides, CoA and DNA
in rat liver
5.2.1 Results and discussion
First, when rats were perfused with 100%
2
H2O buffer, [13C6]glucose, or
[13C5]ribose, the changes of the labeling pattern of glycolytic intermediates and
ribose-5-P provided evidence of constant glycolysis, gluconeogenesis and CAC
operation. The presence of all possible labeled glucose mass isotopomers (M1 to
M7 from 2H2O and M1 to M6 from [13C6]glucose or [13C5]ribose) demonstrates
that the labeling pattern of trioses is modified by reactions of citric acid cycle and
of the substrate cycle pyruvate → oxaloacetate → PEP → pyruvate (Figs 4.14
and 4.22).
Second, in perfusions conducted with 4 mM [13C6]glucose, the comparison of the
MIDs of glucose-6-P and ribose-5-P (Fig 4.20) provides information on the
conditions of operation of the pentose phosphate pathway in these livers. The M6
enrichment of glucose-6-P (30-35% after 90 min) is much higher than the M5
enrichment of ribose-5-P (4-4.5%). This is because most of the ribose-5-P in
these livers was formed in the reversal of the non-oxidative branch of the
pentose phosphate pathway. The low enrichment of ribose-5-P resulted from the
multiple combinations between intermediates with 3,4,5,6 and 7 carbons, with
low probability to yield M5 ribose-5-P. These fasted rat livers had low fatty acid
synthesis rates and the requirement for NADPH production via the oxidative
branch of the pentose phosphate pathway was low.
150
Third, in livers perfused with 100% 2H2O buffer, the much higher M1 labeling of
AMP compared to ribose-5-P shows that one 2H atom from the buffer was
incorporated between ribose-5-P and AMP in the ribose moiety (Figs 4.15 and
4.16). ATP was labeled mainly from salvage pathways because it’s labeling
shows only small proportions of mass isotopomers heavier than M1. The
synthesis of M1 AMP via the salvage of adenine from RNA and DNA breakdown
involves probably the inversion of configuration of C-1 of the ribose moiety of αPRPP when it is converted to adenosine by adenine phosphoribosyl transferase.
This leads to the incorporation of 2H+ from the buffer on C-1 of the ribose moiety
of AMP.
Fourth, my data allow to calculate the fractional synthesis rates of adenine
nucleotides and CoA. In perfusions in 100% 2H2O buffer, the fractional turnover
rate of CoA is 0.56% / min or 33% / h. This rate probably applies to the whole
CoA pool (free and esterified). In perfusions with 4 mM unlabeled glucose + 2
mM [13C5]ribose, the fractional turnover rate of AMP is 7%/h. This rate applies to
the three adenine nucleotides which are in isotopic equilibrium (Fig 4.24)
Concluding remarks:
1. My data show that during a short 2hr liver perfusion period, the rates of de
novo synthesis of the adenine nucleotides and CoA in rat livers are very low. The
measurements of the de novo synthesis rates need longer-term experiment. The
adenine nucleotides and CoA are mainly synthesized from salvage pathways in
rat livers perfused for 2hr.
151
2. In the fasted liver, glycolysis, gluconeogenesis, citric acid cycle and the
pentose phosphate pathway are constantly cycling and changing the labeling
patterns of their intermediates.
3. The non-oxidative branch of the pentose phosphate pathway provides more
ribose-5-P than the oxidative branch in the fasted rat liver.
4. [2H]water is a good tracer that can be used in different time-scale experiments
(in vivo for a long-term experiment, perfused isolated organ for a short-term
experiment) to diagnose variations in ATP and DNA metabolism.
5.2.2 Future directions
This project provides a useful model and method to measure the synthesis of
adenine nucleotides, CoA and DNA in rat livers. Future work should include the
following:
1. Using NMR, identify the rates of 2H incorporation into the ribose moiety of
adenine nucleotides, CoA and DNA during long-term experiments. I hypothesize
that the incorporation of 2H at different carbon positions (C-2-5) reflect the rate of
de novo synthesis; 2H on C-1 reflects both of the de novo synthesis and the
purine salvage pathways.
2. Perfusion of different isolated rat organs with 2H2O buffer during short-term
experiments will evaluate the purine salvage pathways.
152
3. Clinical applications in patients with cancer or disorders of purine metabolism
monitor the rates of synthesis of DNA labeling from [2H]water as an index of the
rate of cell division (before and after chemotherapy).
153
Reference List
1. McGarry JD and Foster DW. Regulation of hepatic fatty acid oxidation
and ketone body production. Annu Rev Biochem 49: 395-420, 1980.
2. Robinson AM and Williamson DH. Physiological roles of ketone bodies
as substrates and signals in mammalian tissues. Physiol Rev 60: 143-187,
1980.
3. Balasse EO and Fery F. Ketone body production and disposal: effects of
fasting, diabetes, and exercise. Diabetes Metab Rev 5: 247-270, 1989.
4. Pederson KJ. The ketonic decomposition of beta-carboxylic acids. J Am
Chem Soc 51: 2098-2107, 1929.
5. Wildenhoff KE. Diurnal variations in the concentrations of blood
acetoacetate, 3-hydroxybutyrate and glucose in normal persons. Acta Med
Scand 191: 303-306, 1972.
6. Wildenhoff KE, Johansen JP, Karstoft H, Yde H and Sorensen NS.
Diurnal variations in the concentrations of blood acetoacetate and 3hydroxybutyrate. The ketone body peak around midnight and its
relationship to free fatty acids, glycerol, insulin, growth hormone and
glucose in serum and plasma. Acta Med Scand 195: 25-28, 1974.
7. Bates MW, Krebs HA and Williamson DH. Turnover rates of ketone
bodies in normal, starved and alloxan-diabetic rats. Biochem J 110: 655661, 1968.
8. Hawkins RA, Williamson DH and Krebs HA. Ketone-body utilization by
adult and suckling rat brain in vivo. Biochem J 122: 13-18, 1971.
154
9. Owen OE, Felig P, Morgan AP, Wahren J and Cahill GF, Jr. Liver and
kidney metabolism during prolonged starvation. J Clin Invest 48: 574-583,
1969.
10. Owen OE and Reichard GA, Jr. Human forearm metabolism during
progressive starvation. J Clin Invest 50: 1536-1545, 1971.
11. Reichard GA, Jr., Owen OE, Haff AC, Paul P and Bortz WM. Ketonebody production and oxidation in fasting obese humans. J Clin Invest 53:
508-515, 1974.
12. Bates MW. Kinetics of ketone body metabolism in fasted and diabetic
rats. Am J Physiol 221: 984-991, 1971.
13. Genuth SM and Castro J. Effect of oral alanine on blood betahydroxybutyrate and plasma glucose, insulin, free fatty acids, and growth
hormone in normal and diabetic subjects. Metabolism 23: 375-386, 1974.
14. Fritz IB. Factors influencing the rates of long-chain fatty acid oxidation
and synthesis in mammalian systems. Physiol Rev 41: 52-129, 1961.
15. Hanson PG, Johnson RE and Zaharko DS. Correlation between ketone
body and free fatty acid concentrations in the plasma during early
starvation in man. Metabolism 14: 1037-1040, 1965.
16. Ontko JA and Zilversmit DB. Correlation between concentrations of
circulating free fatty acids and ketone bodies. Proc Soc Exp Biol Med 121:
319-321, 1966.
17. Eaton S. Control of mitochondrial beta-oxidation flux. Prog Lipid Res 41:
197-239, 2002.
155
18. Shepherd D, Yates DW and Garland PB. The relationship between the
rates of conversion of palmitate into citrate or acetoacetate and the acetylcoenzyme A content of rat-liver mitochondria. Biochem J 97: 38C-40C,
1965.
19. McGarry JD, Mannaerts GP and Foster DW. A possible role for malonylCoA in the regulation of hepatic fatty acid oxidation and ketogenesis. J
Clin Invest 60: 265-270, 1977.
20. McGarry
JD,
Leatherman
GF
and
Foster
DW.
Carnitine
palmitoyltransferase I. The site of inhibition of hepatic fatty acid oxidation
by malonyl-CoA. J Biol Chem 253: 4128-4136, 1978.
21. MacKay EM. The significance of ketosis. J Clin Endocrionol 3: 101-110,
1943.
22. Endemann G, Goetz PG, Edmond J and Brunengraber H. Lipogenesis
from ketone bodies in the isolated perfused rat liver. Evidence for the
cytosolic activation of acetoacetate. J Biol Chem 257: 3434-3440, 1982.
23. Williamson DH, Bates MW, Page MA and Krebs HA. Activities of
enzymes involved in acetoacetate utilization in adult mammalian tissues.
Biochem J 121: 41-47, 1971.
24. Owen OE, Morgan AP, Kemp HG, Sullivan JM, Herrera MG and Cahill
GF, Jr. Brain metabolism during fasting. J Clin Invest 46: 1589-1595,
1967.
156
25. Cremer JE and Heath DF. The estimation of rates of utilization of glucose
and ketone bodies in the brain of the suckling rat using compartmental
analysis of isotopic data. Biochem J 142: 527-544, 1974.
26. Edmond J. Ketone bodies as precursors of sterols and fatty acids in the
developing rat. J Biol Chem 249: 72-80, 1974.
27. Webber RJ and Edmond J. Utilization of L(+)-3-hydroxybutyrate, D(-)-3hydroxybutyrate, acetoacetate, and glucose for respiration and lipid
synthesis in the 18-day-old rat. J Biol Chem 252: 5222-5226, 1977.
28. Palaiologos G and Felig P. Effects of ketone bodies on amino acid
metabolism in isolated rat diaphragm. Biochem J 154: 709-716, 1976.
29. Robinson AM and Williamson DH. Utlization of D-3-hydroxy[314C]butyrate for lipogenesis in vivo in lactating rat mammary gland.
Biochem J 176: 635-638, 1978.
30. Jocken JW and Blaak EE. Catecholamine-induced lipolysis in adipose
tissue and skeletal muscle in obesity. Physiol Behav 94: 219-230, 2008.
31. Fonseca-Alaniz MH, Takada J, onso-Vale MI and Lima FB. Adipose
tissue as an endocrine organ: from theory to practice. J Pediatr (Rio J )
83: S192-S203, 2007.
32. Votruba SB and Jensen MD. Regional fat deposition as a factor in FFA
metabolism. Annu Rev Nutr 27: 149-163, 2007.
33. Coppack SW, Jensen MD and Miles JM. In vivo regulation of lipolysis in
humans. J Lipid Res 35: 177-193, 1994.
157
34. Kather H, Wieland E, Scheurer A, Vogel G, Wildenberg U and Joost
C. Influences of variation in total energy intake and dietary composition on
regulation of fat cell lipolysis in ideal-weight subjects. J Clin Invest 80:
566-572, 1987.
35. Moro C, Polak J, Hejnova J, Klimcakova E, Crampes F, Stich V,
Lafontan M and Berlan M. Atrial natriuretic peptide stimulates lipid
mobilization during repeated bouts of endurance exercise. Am J Physiol
Endocrinol Metab 290: E864-E869, 2006.
36. Reynisdottir S, Wahrenberg H, Carlstrom K, Rossner S and Arner P.
Catecholamine resistance in fat cells of women with upper-body obesity
due to decreased expression of beta 2-adrenoceptors. Diabetologia 37:
428-435, 1994.
37. Stralfors P and Honnor RC. Insulin-induced dephosphorylation of
hormone-sensitive lipase. Correlation with lipolysis and cAMP-dependent
protein kinase activity. Eur J Biochem 182: 379-385, 1989.
38. Holm C, Kirchgessner TG, Svenson KL, Fredrikson G, Nilsson S,
Miller CG, Shively JE, Heinzmann C, Sparkes RS, Mohandas T and .
Hormone-sensitive lipase: sequence, expression, and chromosomal
localization to 19 cent-q13.3. Science 241: 1503-1506, 1988.
39. Li Z, Sumida M, Birchbauer A, Schotz MC and Reue K. Isolation and
characterization of the gene for mouse hormone-sensitive lipase.
Genomics 24: 259-265, 1994.
158
40. Contreras JA, Karlsson M, Osterlund T, Laurell H, Svensson A and
Holm
C.
Hormone-sensitive
lipase
is
structurally
related
to
acetylcholinesterase, bile salt-stimulated lipase, and several fungal
lipases. Building of a three-dimensional model for the catalytic domain of
hormone-sensitive lipase. J Biol Chem 271: 31426-31430, 1996.
41. Anthonsen MW, Ronnstrand L, Wernstedt C, Degerman E and Holm
C. Identification of novel phosphorylation sites in hormone-sensitive lipase
that are phosphorylated in response to isoproterenol and govern activation
properties in vitro. J Biol Chem 273: 215-221, 1998.
42. Holm C, Belfrage P and Fredrikson G. Immunological evidence for the
presence of hormone-sensitive lipase in rat tissues other than adipose
tissue. Biochem Biophys Res Commun 148: 99-105, 1987.
43. Khoo JC, Reue K, Steinber D and Schotz MC. Expression of hormonesensitive lipase mRNA in macrophages. J Lipid Res 34: 1969-1974, 1993.
44. Mulder H, Holst LS, Svensson H, Degerman E, Sundler F, Ahren B,
Rorsman P and Holm C. Hormone-sensitive lipase, the rate-limiting
enzyme in triglyceride hydrolysis, is expressed and active in beta-cells.
Diabetes 48: 228-232, 1999.
45. Large V, Reynisdottir S, Langin D, Fredby K, Klannemark M, Holm C
and Arner P. Decreased expression and function of adipocyte hormonesensitive lipase in subcutaneous fat cells of obese subjects. J Lipid Res
40: 2059-2066, 1999.
159
46. Shen WJ, Liang Y, Hong R, Patel S, Natu V, Sridhar K, Jenkins A,
Bernlohr DA and Kraemer FB. Characterization of the functional
interaction of adipocyte lipid-binding protein with hormone-sensitive lipase.
J Biol Chem 276: 49443-49448, 2001.
47. Bezaire V and Langin D. Regulation of adipose tissue lipolysis revisited.
Proc Nutr Soc 68: 350-360, 2009.
48. Langin D, Dicker A, Tavernier G, Hoffstedt J, Mairal A, Ryden M,
Arner E, Sicard A, Jenkins CM, Viguerie N, van H, V, Gross RW, Holm
C and Arner P. Adipocyte lipases and defect of lipolysis in human obesity.
Diabetes 54: 3190-3197, 2005.
49. Haemmerle G, Lass A, Zimmermann R, Gorkiewicz G, Meyer C,
Rozman J, Heldmaier G, Maier R, Theussl C, Eder S, Kratky D,
Wagner EF, Klingenspor M, Hoefler G and Zechner R. Defective
lipolysis and altered energy metabolism in mice lacking adipose
triglyceride lipase. Science 312: 734-737, 2006.
50. Zimmermann R, Strauss JG, Haemmerle G, Schoiswohl G, BirnerGruenberger R, Riederer M, Lass A, Neuberger G, Eisenhaber F,
Hermetter A and Zechner R. Fat mobilization in adipose tissue is
promoted by adipose triglyceride lipase. Science 306: 1383-1386, 2004.
51. Kim JY, Tillison K, Lee JH, Rearick DA and Smas CM. The adipose
tissue triglyceride lipase ATGL/PNPLA2 is downregulated by insulin and
TNF-alpha in 3T3-L1 adipocytes and is a target for transactivation by
PPARgamma. Am J Physiol Endocrinol Metab 291: E115-E127, 2006.
160
52. Fredrikson G, Tornqvist H and Belfrage P. Hormone-sensitive lipase
and monoacylglycerol lipase are both required for complete degradation of
adipocyte triacylglycerol. Biochim Biophys Acta 876: 288-293, 1986.
53. Blaak EE, van Baak MA, Kemerink GJ, Pakbiers MT, Heidendal GA
and Saris WH. Beta-adrenergic stimulation of energy expenditure and
forearm skeletal muscle metabolism in lean and obese men. Am J Physiol
267: E306-E315, 1994.
54. Fukao T, Lopaschuk GD and Mitchell GA. Pathways and control of
ketone
body
metabolism:
on
the
fringe
of
lipid
biochemistry.
Prostaglandins Leukot Essent Fatty Acids 70: 243-251, 2004.
55. Greenberg AS, Shen WJ, Muliro K, Patel S, Souza SC, Roth RA and
Kraemer FB. Stimulation of lipolysis and hormone-sensitive lipase via the
extracellular signal-regulated kinase pathway. J Biol Chem 276: 4545645461, 2001.
56. Greenberg AS, Egan JJ, Wek SA, Garty NB, Blanchette-Mackie EJ
and Londos C. Perilipin, a major hormonally regulated adipocyte-specific
phosphoprotein associated with the periphery of lipid storage droplets. J
Biol Chem 266: 11341-11346, 1991.
57. Londos C, Brasaemle DL, Schultz CJ, Segrest JP and Kimmel AR.
Perilipins, ADRP, and other proteins that associate with intracellular
neutral lipid droplets in animal cells. Semin Cell Dev Biol 10: 51-58, 1999.
161
58. Moore HP, Silver RB, Mottillo EP, Bernlohr DA and Granneman JG.
Perilipin targets a novel pool of lipid droplets for lipolytic attack by
hormone-sensitive lipase. J Biol Chem 280: 43109-43120, 2005.
59. Granneman JG, Moore HP, Granneman RL, Greenberg AS, Obin MS
and Zhu Z. Analysis of lipolytic protein trafficking and interactions in
adipocytes. J Biol Chem 282: 5726-5735, 2007.
60. Schweiger M, Schreiber R, Haemmerle G, Lass A, Fledelius C,
Jacobsen P, Tornqvist H, Zechner R and Zimmermann R. Adipose
triglyceride lipase and hormone-sensitive lipase are the major enzymes in
adipose tissue triacylglycerol catabolism. J Biol Chem 281: 40236-40241,
2006.
61. Mersmann HJ. Comparison of plasma free-fatty-acid and blood-glycerol
concentrations with measurement of lipolysis in porcine adipose tissue in
vitro. J Anim Sci 63: 757-769, 1986.
62. Vaughan M. The production and release of glycerol by adipose tissue
incubated in vitro. J Biol Chem 237: 3354-3358, 1962.
63. Gu L, Zhang GF, Kombu RS, Allen F, Kutz G, Brewer WU, Roe CR and
Brunengraber H. Parenteral and Enteral Metabolism of Anaplerotic
Triheptanoin in Normal Rats. II. Effects on lipolysis, glucose production
and liver acyl-CoA profile. Am J Physiol Endocrinol Metab 2009.
64. Mittendorfer B, Liem O, Patterson BW, Miles JM and Klein S. What
does the measurement of whole-body fatty acid rate of appearance in
162
plasma by using a fatty acid tracer really mean? Diabetes 52: 1641-1648,
2003.
65. Horton TJ, Dow S, Armstrong M and Donahoo WT. Greater systemic
lipolysis in women compared with men during moderate-dose infusion of
epinephrine and/or norepinephrine. J Appl Physiol 107: 200-210, 2009.
66. Coppack SW, Frayn KN, Humphreys SM, Whyte PL and Hockaday
TD. Arteriovenous differences across human adipose and forearm tissues
after overnight fast. Metabolism 39: 384-390, 1990.
67. Spector AA. Fatty acid binding to plasma albumin. J Lipid Res 16: 165179, 1975.
68. Scowro and Chernick SS. Hormonal control of protein and fat
metabolism in the pancreatectomized rat. Recent Prog Horm Res 16: 497545, 1960.
69. Meier JM, McGarry JD, Faloona GR, Unger RH and Foster DW.
Studies of the development of diabetic ketosis in the rat. J Lipid Res 13:
228-233, 1972.
70. Alberti KG, Hockaday TD and Turner RC. Small doses of intramuscular
insulin in the treatment of diabetic "coma". Lancet 2: 515-522, 1973.
71. Mayes PA and Felts JM. Regulation of fat metabolism of the liver. Nature
215: 716-718, 1967.
72. Hamilton JA and Kamp F. How are free fatty acids transported in
membranes? Is it by proteins or by free diffusion through the lipids?
Diabetes 48: 2255-2269, 1999.
163
73. Luxon BA. Inhibition of binding to fatty acid binding protein reduces the
intracellular transport of fatty acids. Am J Physiol 271: G113-G120, 1996.
74. Norum KR, Farstad M and Bremer J. The submitochondrial distribution
of acid:CoA ligase (AMP) and palmityl-CoA:carnitine palmityltransferase in
rat liver mitochondria. Biochem Biophys Res Commun 22: 797-804, 1966.
75. Shindo Y and Hashimoto T. Acyl-Coenzyme A synthetase and fatty acid
oxidation in rat liver peroxisomes. J Biochem 84: 1177-1181, 1978.
76. Kornberg A and Pricer WE, Jr. Enzymatic synthesis of the coenzyme A
derivatives of long chain fatty acids. J Biol Chem 204: 329-343, 1953.
77. Gossett RE, Schroeder F, Gunn JM and Kier AB. Expression of fatty
acyl-CoA binding proteins in colon cells: response to butyrate and
transformation. Lipids 32: 577-585, 1997.
78. Cook GA and Park EA. Expression and regulation of carnitine
palmitoyltransferase-Ialpha and -Ibeta genes. Am J Med Sci 318: 43-48,
1999.
79. Saggerson D. Malonyl-CoA, a key signaling molecule in mammalian cells.
Annu Rev Nutr 28: 253-272, 2008.
80. Guynn RW, Veloso D and Veech RL. The concentration of malonylcoenzyme A and the control of fatty acid synthesis in vivo. J Biol Chem
247: 7325-7331, 1972.
81. Hardie DG. Regulation of fatty acid and cholesterol metabolism by the
AMP-activated protein kinase. Biochim Biophys Acta 1123: 231-238,
1992.
164
82. Park H, Kaushik VK, Constant S, Prentki M, Przybytkowski E,
Ruderman NB and Saha AK. Coordinate regulation of malonyl-CoA
decarboxylase, sn-glycerol-3-phosphate acyltransferase, and acetyl-CoA
carboxylase by AMP-activated protein kinase in rat tissues in response to
exercise. J Biol Chem 277: 32571-32577, 2002.
83. Habinowski SA, Hirshman M, Sakamoto K, Kemp BE, Gould SJ,
Goodyear LJ and Witters LA. Malonyl-CoA decarboxylase is not a
substrate of AMP-activated protein kinase in rat fast-twitch skeletal muscle
or an islet cell line. Arch Biochem Biophys 396: 71-79, 2001.
84. Lopaschuk GD and Stanley WC. Malonyl-CoA decarboxylase inhibition
as a novel approach to treat ischemic heart disease. Cardiovasc Drugs
Ther 20: 433-439, 2006.
85. Cook GA. Differences in the sensitivity of carnitine palmitoyltransferase to
inhibition by malonyl-CoA are due to differences in Ki values. J Biol Chem
259: 12030-12033, 1984.
86. Grantham BD and Zammit VA. Restoration of the properties of carnitine
palmitoyltransferase I in liver mitochondria during re-feeding of starved
rats. Biochem J 239: 485-488, 1986.
87. Faye A, Borthwick K, Esnous C, Price NT, Gobin S, Jackson VN,
Zammit VA, Girard J and Prip-Buus C. Demonstration of N- and Cterminal
domain
intramolecular
interactions
in
rat
liver
carnitine
palmitoyltransferase 1 that determine its degree of malonyl-CoA
sensitivity. Biochem J 387: 67-76, 2005.
165
88. McGarry JD, Robles-Valdes C and Foster DW. Role of carnitine in
hepatic ketogenesis. Proc Natl Acad Sci U S A 72: 4385-4388, 1975.
89. Louet JF, Le MC, Pegorier JP, Decaux JF and Girard J. Regulation of
liver carnitine palmitoyltransferase I gene expression by hormones and
fatty acids. Biochem Soc Trans 29: 310-316, 2001.
90. Park EA, Mynatt RL, Cook GA and Kashfi K. Insulin regulates enzyme
activity, malonyl-CoA sensitivity and mRNA abundance of hepatic
carnitine palmitoyltransferase-I. Biochem J 310 ( Pt 3): 853-858, 1995.
91. Jackson S, Bartlett K, Land J, Moxon ER, Pollitt RJ, Leonard JV and
Turnbull DM. Long-chain 3-hydroxyacyl-CoA dehydrogenase deficiency.
Pediatr Res 29: 406-411, 1991.
92. Jackson S, Kler RS, Bartlett K, Briggs H, Bindoff LA, Pourfarzam M,
Gardner-Medwin D and Turnbull DM. Combined enzyme defect of
mitochondrial fatty acid oxidation. J Clin Invest 90: 1219-1225, 1992.
93. Kinman RP, Kasumov T, Jobbins KA, Thomas KR, Adams JE,
Brunengraber LN, Kutz G, Brewer WU, Roe CR and Brunengraber H.
Parenteral and enteral metabolism of anaplerotic triheptanoin in normal
rats. Am J Physiol Endocrinol Metab 291: E860-E866, 2006.
94. Ibdah JA, Paul H, Zhao Y, Binford S, Salleng K, Cline M, Matern D,
Bennett MJ, Rinaldo P and Strauss AW. Lack of mitochondrial
trifunctional protein in mice causes neonatal hypoglycemia and sudden
death. J Clin Invest 107: 1403-1409, 2001.
166
95. Lehninger AL. A quantitative study of the products of fatty acid oxidation
in liver suspensions. J Biol Chem 164: 291-306, 2000.
96. Williamson JR, Scholz R and Browning ET. Control mechanisms of
gluconeogenesis and ketogenesis. II. Interactions between fatty acid
oxidation and the citric acid cycle in perfused rat liver. J Biol Chem 244:
4617-4627, 1969.
97. Clinkenbeard KD, Reed WD, Mooney RA and Lane MD. Intracellular
localization of the 3-hydroxy-3-methylglutaryl coenzme A cycle enzymes in
liver. Separate cytoplasmic and mitochondrial 3-hydroxy-3-methylglutaryl
coenzyme A generating systems for cholesterogenesis and ketogenesis. J
Biol Chem 250: 3108-3116, 1975.
98. Rudney H. The biosynthesis of beta-hydroxy-beta-methylglutaric acid. J
Biol Chem 227: 363-377, 1957.
99. Reed WD, Clinkenbeard D and Lane MD. Molecular and catalytic
properties
of
mitochondrial
(ketogenic)
3-hydroxy-3-methylglutaryl
coenzyme A synthase of liver. J Biol Chem 250: 3117-3123, 1975.
100. Casals N, Roca N, Guerrero M, Gll-Gomez G, Ayte J and Ciudad
C.J.and Hegardt F.G. Regulation of the expression of mitochondrial 3hydroxy-3-methylglutaryl-CoA synthase gene. Biochem J 283: 261-264,
1992.
101. Lowe DM and Tubbs PK. Succinylation and inactivation of 3-hydroxy-3methylglutaryl-CoA synthase by succinyl-CoA and its possible relevance
to the control of ketogenesis. Biochem J 232: 37-42, 1985.
167
102. Siess EA, Fahimi FM and Wieland OH. Decrease by glucagon in hepatic
succinyl-CoA. Biochem Biophys Res Commun 95: 205-211, 1980.
103. Quant PA, Robin D, Robin P, Ferre P, Brand MD and Girard J. Control
of hepatic mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase during
the foetal/neonatal transition, suckling and weaning in the rat. Eur J
Biochem 195: 449-454, 1991.
104. Leclerc J, Des Rosiers C., Montgomery JA, Brunet J, Ste-Marie L,
Reider MW, Fernandez CA, Powers L, David F and Brunengraber H.
Metabolism of R-beta-hydroxypentanoate and of beta-ketopentanoate in
conscious dogs. Am J Physiol 268: E446-E452, 1995.
105. Sweetman L, Weyler W, Nyhan WL, de CC, Loria AR and Estrada Y.
Abnormal metabolites of isoleucine in a patient with propionyl-CoA
carboxylase deficiency. Biomed Mass Spectrom 5: 198-207, 1978.
106. Roe CR, Sweetman L, Roe DS, David F and Brunengraber H.
Treatment of cardiomyopathy and rhabdomyolysis in long-chain fat
oxidation disorders using an anaplerotic odd-chain triglyceride. J Clin
Invest 110: 259-269, 2002.
107. Diedrich M and Henschel K.P. The natural occurrence of unusual fatty
acids. 1. Odd numbered fatty acids. Nahrung 34: 935-943, 1990.
108. Vlaeminck B, Fievez V, Tamminga S, Dewhurst RJ, van VA, De BD
and Demeyer D. Milk odd- and branched-chain fatty acids in relation to
the rumen fermentation pattern. J Dairy Sci 89: 3954-3964, 2006.
168
109. Vlaeminck B, Dufour C, van Vuuren AM, Cabrita AR, Dewhurst RJ,
Demeyer D and Fievez V. Use of odd and branched-chain fatty acids in
rumen contents and milk as a potential microbial marker. J Dairy Sci 88:
1031-1042, 2005.
110. Keneda T. Iso- and anteiso-fatty acids in bacteria: biosynthesis, function,
and taxonomic significance. Microbiol rev 55: 288-302, 91 A.D.
111. Baker FC and Schooley DA. Biosynthesis of 3-hydroxy-3-methylglutarylCoA, 3-hydroxy-3-ethylglutaryl-CoA, mevalonate and homomevalonate by
insect corpus allatum and mammalian hepatic tissues. Biochim Biophys
Acta 664: 356-372, 1981.
112. Bergmeyer HU, Gawehn K, Klotzsch H, Krebs HA and Williamson DH.
Purification and properties of crystalline 3-hydroxybutyrate dehydrogenase
from Rhodopseudomonas spheroides. Biochem J 102: 423-431, 1967.
113. Kuhara T, Inoue Y, Shinka T, Matsumoto I and Matsuo M. Identification
of 3-hydroxy-3-ethylglutaric acid in urine of patients with propionic
acidaemia. Biomed Mass Spectrom 10: 629-632, 1983.
114. Duran M, Gompertz D, Bruinvis L, Ketting D and Wadman SK. The
variability of metabolite excretion in propionicacidaemia. Clin Chim Acta
82: 93-99, 1978.
115. Hommes FA, Kuipers JR, Elema JD, Jansen JF and Jonxis JH.
Propionicacidemia, a new inborn error of metabolism. Pediatr Res 2: 519524, 1968.
169
116. Green DE, Dewan JG and Leloir LF. The beta-hydroxybutyric
dehydrogenase of animal tissues. Biochem J 31: 934-949, 1937.
117. Fenselau A and Wallis K. Comparative studies on 3-oxo acid coenzyme
A transferase from various rat tissues. Biochem J 142: 619-627, 1974.
118. Middleton B. The oxoacyl-coenzyme A thiolases of animal tissues.
Biochem J 132: 717-730, 1973.
119. Buckley BM and Williamson DH. Acetoacetate and brain lipogenesis:
developmental pattern of acetoacetyl-coenzyme A synthetase in the
soluble fraction of rat brain. Biochem J 132: 653-656, 1973.
120. Tildon JT and Sevdalian DA. CoA transferase in the brain and other
mammalian tissues. Arch Biochem Biophys 148: 382-390, 1972.
121. Lincoln BC, Des Rosiers C and Brunengraber H. Metabolism of S-3hydroxybutyrate in the perfused rat liver. Arch Biochem Biophys 259: 149156, 1987.
122. Seccombe DW, Harding PG and Possmayer F. Fetal utilization of
maternally derived ketone bodies for lipogenesis in the rat. Biochim
Biophys Acta 488: 402-416, 1977.
123. Paterson P, Sheath J, Taft P and Wood C. Maternal and foetal ketone
concentrations in plasma and urine. Lancet 1: 862-865, 1967.
124. Cheng KK and Yang MM. Study of pregnancy ketosis in the rat. Q J Exp
Physiol Cogn Med Sci 55: 83-92, 1970.
170
125. Shambaugh GE, Mrozak SC and Freinkel N. Fetal fuels. I. Utilization of
ketones by isolated tissues at various stages of maturation and maternal
nutrition during late gestation. Metabolism 26: 623-635, 1977.
126. Shambaugh GE, III, Koehler RR and Yokoo H. Fetal fuels. III: Ketone
utilization by fetal hepatocyte. Am J Physiol 235: E330-E337, 1978.
127. Siesjo BK. In: Brain energy metabolism. New York: Wiley 70: 101-125,
1978.
128. Sokoloff L. Metabolism of ketone bodies by the brain. Annu Rev Med 24:
271-280, 1973.
129. Vannucci SJ and Simpson IA. Developmental switch in brain nutrient
transporter expression in the rat. Am J Physiol Endocrinol Metab 285:
E1127-E1134, 2003.
130. Nebeling LC and Lerner E. Implementing a ketogenic diet based on
medium-chain triglyceride oil in pediatric patients with cancer. J Am Diet
Assoc 95: 693-697, 1995.
131. Nehlig A and Pereira d, V. Glucose and ketone body utilization by the
brain of neonatal rats. Prog Neurobiol 40: 163-221, 1993.
132. Veech RL, Chance B, Kashiwaya Y, Lardy HA and Cahill GF, Jr.
Ketone bodies, potential therapeutic uses. IUBMB Life 51: 241-247, 2001.
133. Enerson BE and Drewes LR. Molecular features, regulation, and function
of monocarboxylate transporters: implications for drug delivery. J Pharm
Sci 92: 1531-1544, 2003.
171
134. Gerhart DZ, Enerson BE, Zhdankina OY, Leino RL and Drewes LR.
Expression of monocarboxylate transporter MCT1 by brain endothelium
and glia in adult and suckling rats. Am J Physiol 273: E207-E213, 1997.
135. Leino RL, Gerhart DZ, Duelli R, Enerson BE and Drewes LR. Dietinduced ketosis increases monocarboxylate transporter (MCT1) levels in
rat brain. Neurochem Int 38: 519-527, 2001.
136. Puchowicz MA, Xu K, Sun X, Ivy A, Emancipator D and LaManna JC.
Diet-induced ketosis increases capillary density without altered blood flow
in rat brain. Am J Physiol Endocrinol Metab 292: E1607-E1615, 2007.
137. Dahlquist G, Persson U and Persson B. The activity of D- hydroxybutyrate dehydrogenase in fetal, infant and adult rat brain and the
influence of starvation. Biol Neonate 20: 40-50, 1972.
138. Daniel PM, Love ER, Moorhouse SR and Pratt OE. The transport of
ketone bodies into the brain of the rat (in vivo). J Neurol Sci 34: 1-13,
1977.
139. Yeh YY, Streuli VL and Zee P. Ketone bodies serve as important
precursors of brain lipids in the developing rat. Lipids 12: 957-964, 1977.
140. Cremer JE, Teal HM and Cunningham VJ. Inhibition, by 2-oxo acids that
accumulate in maple-syrup-urine disease, of lactate, pyruvate, and 3hydroxybutyrate transport across the blood-brain barrier. J Neurochem 39:
674-677, 1982.
141. Schroeder H, Bomont L and Nehlig A. Influence of early chronic
phenobarbital treatment on cerebral arteriovenous differences of glucose
172
and ketone bodies in the developing rat. Int J Dev Neurosci 9: 453-461,
1991.
142. Dysza HA, Czajka DM and Miller SA. Influence of artifical diet on weight
gain and body composition of the neonatal rat. J Nutr 84: 100-106, 1964.
143. Lopes-Cardozo M and Klein W. Ketone-body utilization and lipid
synthesis by developing rat brain-a comparison between in vivo and in
vitro experiments. Neurochem Int 6: 459-466, 1984.
144. Yeh YY and Sheehan PM. Preferential utilization of ketone bodies in the
brain and lung of newborn rats. Fed Proc 44: 2352-2358, 1985.
145. Hawkins RA and Williamson DH. Measurements of substrate uptake by
mammary gland of the rat. Biochem J 129: 1171-1173, 1972.
146. Bernard L, Leroux C and Chilliard Y. Expression and nutritional
regulation of lipogenic genes in the ruminant lactating mammary gland.
Adv Exp Med Biol 606: 67-108, 2008.
147. Williamson DH, McKeown SR and Ilic V. Interactions of glucose,
acetoacetate and insulin in mammary-gland slices of lactating rats.
Biochem J 150: 145-152, 1975.
148. Bartley JC. Modulation by ketone bodies of the rate of fatty acid synthesis
in mammary gland slices from lactating rats. Lipids 11: 774-777, 1976.
149. Angielski S and Lukowicz J. The role of the kidney in the removal of
ketone bodies under different acid-base status of the rat. Am J Clin Nutr
31: 1635-1641, 1978.
173
150. Sapir DG and Owen OE. Renal conservation of ketone bodies during
starvation. Metabolism 24: 23-33, 1975.
151. Randle PJ, Newsholme EA and Garland PB. Regulation of glucose
uptake by muscle. 8. Effects of fatty acids, ketone bodies and pyruvate,
and of alloxan-diabetes and starvation, on the uptake and metabolic fate
of glucose in rat heart and diaphragm muscles. Biochem J 93: 652-665,
1964.
152. Hall LM. Preferential oxidation of acetoacetate by the perfused heart.
Biochem Biophys Res Commun 6: 177-179, 1961.
153. Sherwin RS, Hendler RG and Felig P. Effect of diabetes mellitus and
insulin on the turnover and metabolic response to ketones in man.
Diabetes 25: 776-784, 1976.
154. Garber AJ, Menzel PH, Boden G and Owen OE. Hepatic ketogenesis
and gluconeogenesis in humans. J Clin Invest 54: 981-989, 1974.
155. Keller U, Sonnenberg GE and Stauffacher W. Validation of a tracer
technique to determine nonsteady-state ketone body turnover rates in
man. Am J Physiol 240: E253-E262, 1981.
156. Bougneres PF, Balasse EO, Ferre P and Bier DM. Determination of
ketone body kinetics using a D-(-)-3-hydroxy[4,4,4-2H3]butyrate tracer. J
Lipid Res 27: 215-220, 1986.
157. Balasse EO. Kinetics of ketone body metabolism in fasting humans.
Metabolism 28: 41-50, 1979.
174
158. McGarry JD, Guest MJ and Foster DW. Ketone body metabolism in the
ketosis of starvation and alloxan diabetes. J Biol Chem 245: 4382-4390,
1970.
159. Barton RN. Isotopic study of ketone body kinetics: invalidity of
calculations based upon specific radioactivity of total ketone bodies.
Metabolism 29: 392-396, 1980.
160. Wolfe BM, Havel JR, Marliss EB, Kane JP, Seymour J and Ahuja SP.
Effects of a 3-day fast and of ethanol on splanchnic metabolism of FFA,
amino acids, and carbohydrates in healthy young men. J Clin Invest 57:
329-340, 1976.
161. Reed WD, Baab PJ, Hawkins RL and Ozand PT. A double-isotope
method for the measurement of ketone-body turnover in the rat. Effect of
L-alanine. Biochem J 219: 15-24, 1984.
162. Miles JM, Schwenk WF, McClean KL and Haymond MW. A dualisotope technique for determination of in vivo ketone body kinetics. Am J
Physiol 251: E185-E191, 1986.
163. Fink G, Desrochers S, Des Rosiers C, Garneau M, David F, Daloze T,
Landau BR and Brunengraber H. Pseudoketogenesis in the perfused rat
heart. J Biol Chem 263: 18036-18042, 1988.
164. Des Rosiers C, Montgomery JA, Desrochers S, Garneau M, David F,
Mamer OA and Brunengraber H. Interference of 3-hydroxyisobutyrate
with measurements of ketone body concentration and isotopic enrichment
175
by gas chromatography-mass spectrometry. Anal Biochem 173: 96-105,
1988.
165. Des Rosiers C, Montgomery JA, Garneau M, David F, Mamer OA,
Daloze P, Toffolo G, Cobelli C, Landau BR and Brunengraber H.
Pseudoketogenesis in hepatectomized dogs. Am J Physiol 258: E519E528, 1990.
166. Brunengraber H and Roe CR. Anaplerotic molecules: current and future.
J Inherit Metab Dis 29: 327-331, 2006.
167. Owen OE, Kalhan SC and Hanson RW. The key role of anaplerosis and
cataplerosis for citric acid cycle function. J Biol Chem 277: 30409-30412,
2002.
168. Ballard FJ, Hanson RW and Leveille GA. Phosphoenolpyruvate
carboxykinase and the synthesis of glyceride-glycerol from pyruvate in
adipose tissue. J Biol Chem 242: 2746-2750, 1967.
169. Brunengraber H, Boutry M and Lowenstein JM. Fatty acid and 3- hydroxysterol synthesis in the perfused rat liver. Including measurements
on
the
production
of
lactate,
pyruvate,
-hydroxy-butyrate,
and
acetoacetate by the fed liver. J Biol Chem 248: 2656-2669, 1973.
170. Hassel B. Pyruvate carboxylation in neurons. J Neurosci Res 66: 755762, 2001.
171. Bing RJ. The metabolism of the heart. Trans Am Coll Cardiol 5: 8-14,
1955.
176
172. Taegtmeyer H, Hems R and Krebs HA. Utilization of energy-providing
substrates in the isolated working rat heart. Biochem J 186: 701-711,
1980.
173. Gibala MJ, Young ME and Taegtmeyer H. Anaplerosis of the citric acid
cycle: role in energy metabolism of heart and skeletal muscle. Acta
Physiol Scand 168: 657-665, 2000.
174. Boucher A, Lu D, Burgess SC, Telemaque-Potts S, Jensen MV,
Mulder H, Wang MY, Unger RH, Sherry AD and Newgard CB.
Biochemical mechanism of lipid-induced impairment of glucose-stimulated
insulin secretion and reversal with a malate analogue. J Biol Chem 279:
27263-27271, 2004.
175. Zoppi CC, Silveira LR, Oliveira CA, Boschero AC, Curi R and
Carneiro EM. Insulin release, peripheral insulin resistance and muscle
function in protein malnutrition: a role of tricarboxylic acid cycle
anaplerosis. Br J Nutr 103: 1237-1250, 2010.
176. Utter MF and Keech DB. Formation of oxaloacetate from pyruvate and
carbon dioxide. J Biol Chem 235: C17-C18, 1960.
177. Mackall JC and Lane MD. Role of pyruvate carboxylase in fatty acid
synthesis: alterations during preadipocyte differentiation. Biochem Biophys
Res Commun 79: 720-725, 1977.
178. Cooper AJ and Plum F. Biochemistry and physiology of brain ammonia.
Physiol Rev 67: 440-519, 1987.
177
179. MacDonald MJ. Feasibility of a mitochondrial pyruvate malate shuttle in
pancreatic islets. Further implication of cytosolic NADPH in insulin
secretion. J Biol Chem 270: 20051-20058, 1995.
180. Hasan NM, Longacre MJ, Stoker SW, Boonsaen T, Jitrapakdee S,
Kendrick MA, Wallace JC and MacDonald MJ. Impaired anaplerosis
and insulin secretion in insulinoma cells caused by small interfering RNAmediated suppression of pyruvate carboxylase. J Biol Chem 283: 2804828059, 2008.
181. Russell RR, III and Taegtmeyer H. Pyruvate carboxylation prevents the
decline in contractile function of rat hearts oxidizing acetoacetate. Am J
Physiol 261: H1756-H1762, 1991.
182. Pound KM, Sorokina N, Ballal K, Berkich DA, Fasano M, Lanoue KF,
Taegtmeyer H, O'Donnell JM and Lewandowski ED. Substrate-enzyme
competition attenuates upregulated anaplerotic flux through malic enzyme
in hypertrophied rat heart and restores triacylglyceride content: attenuating
upregulated anaplerosis in hypertrophy. Circ Res 104: 805-812, 2009.
183. Panchal AR, Comte B, Huang H, Kerwin T, Darvish A, Des Rosiers C,
Brunengraber H and Stanley WC. Partitioning of pyruvate between
oxidation and anaplerosis in swine hearts. Am J Physiol Heart Circ Physiol
279: H2390-H2398, 2000.
184. Xu J, Han J, Long YS, Epstein PN and Liu YQ. The role of pyruvate
carboxylase in insulin secretion and proliferation in rat pancreatic beta
cells. Diabetologia 51: 2022-2030, 2008.
178
185. Bunger R, Swindall B, Brodie D, Zdunek D, Stiegler H and Walter G.
Pyruvate attenuation of hypoxia damage in isolated working guinea-pig
heart. J Mol Cell Cardiol 18: 423-438, 1986.
186. Stanley WC, Kivilo KM, Panchal AR, Hallowell PH, Bomont C,
Kasumov T and Brunengraber H. Post-ischemic treatment with
dipyruvyl-acetyl-glycerol decreases myocardial infarct size in the pig.
Cardiovasc Drugs Ther 17: 209-216, 2003.
187. Fransson U, Rosengren AH, Schuit FC, Renstrom E and Mulder H.
Anaplerosis via pyruvate carboxylase is required for the fuel-induced rise
in the ATP:ADP ratio in rat pancreatic islets. Diabetologia 49: 1578-1586,
2006.
188. Constantin-Teodosiu D, Simpson EJ and Greenhaff PL. The
importance of pyruvate availability to PDC activation and anaplerosis in
human skeletal muscle. Am J Physiol 276: E472-E478, 1999.
189. Ahmed A, Maxwell DL, Taylor PM and Rennie MJ. Glutamine transport
in human skeletal muscle. Am J Physiol 264: E993-1000, 1993.
190. Bruce M, Constantin-Teodosiu D, Greenhaff PL, Boobis LH, Williams
C and Bowtell JL. Glutamine supplementation promotes anaplerosis but
not oxidative energy delivery in human skeletal muscle. Am J Physiol
Endocrinol Metab 280: E669-E675, 2001.
191. Rennie MJ, Bowtell JL, Bruce M and Khogali SE. Interaction between
glutamine availability and metabolism of glycogen, tricarboxylic acid cycle
intermediates and glutathione. J Nutr 131: 2488S-2490S, 2001.
179
192. Kasumov T, Cendrowski AV, David F, Jobbins KA, Anderson VE and
Brunengraber H. Mass isotopomer study of anaplerosis from propionate
in the perfused rat heart. Arch Biochem Biophys 463: 110-117, 2007.
193. Sherry AD, Malloy CR, Roby RE, Rajagopal A and Jeffrey FM.
Propionate metabolism in the rat heart by 13C n.m.r. spectroscopy.
Biochem J 254: 593-598, 1988.
194. Martini WZ, Stanley WC, Huang H, Rosiers CD, Hoppel CL and
Brunengraber H. Quantitative assessment of anaplerosis from propionate
in pig heart in vivo. Am J Physiol Endocrinol Metab 284: E351-E356,
2003.
195. Comte B, Vincent G, Bouchard B and and Des Rosiers C. Probing the
origin of acetyl-CoA and oxaloacetate entering the citric acid cycle from
the 13C labeling of citrate released by perfused rat hearts. J Biol Chem
272: 26117-26124, 1997.
196. Roe DS, Vianey-Saban C, Sharma S, Zabot MT and Roe CR. Oxidation
of unsaturated fatty acids by human fibroblasts with very-long-chain acylCoA dehydrogenase deficiency: aspects of substrate specificity and
correlation with clinical phenotype. Clin Chim Acta 312: 55-67, 2001.
197. Mochel F, DeLonlay P, Touati G, Brunengraber H, Kinman RP, Rabier
D, Roe CR and Saudubray JM. Pyruvate carboxylase deficiency: clinical
and biochemical response to anaplerotic diet therapy. Mol Genet Metab
84: 305-312, 2005.
180
198. Tozzi MG, Camici M, Mascia L, Sgarrella F and Ipata PL. Pentose
phosphates in nucleoside interconversion and catabolism. FEBS J 273:
1089-1101, 2006.
199. Sonne JC, Lin I and Buchanan JM. Biosynthesis of the purines. IX.
Precursors of the nitrogen atoms of the purine ring. J Biol Chem 220: 369378, 1956.
200. Hartman SC and Buchanan JM. The biosynthesis of the purines. Ergeb
Physiol 50: 75-121, 1959.
201. Chen S, Turner S, Tsang E, Stark J, Turner H, Mahsut A, Keifer K,
Goldfinger M and Hellerstein MK. Measurement of pancreatic islet cell
proliferation by heavy water labeling. Am J Physiol Endocrinol Metab 293:
E1459-E1464, 2007.
202. Tullson PC, John-Alder HB, Hood DA and Terjung RL. De novo
synthesis of adenine nucleotides in different skeletal muscle fiber types.
Am J Physiol 255: C271-C277, 1988.
203. Dow JW, Nigdikar S and Bowditch J. Adenine nucleotide synthesis de
novo in mature rat cardiac myocytes. Biochim Biophys Acta 847: 223-227,
1985.
204. Zimmer HG, Steinkopff G and Gerlach E. [Studies on regulation of denovo synthesis of adenine nucleotides in the heart]. Verh Dtsch Ges
Kreislaufforsch 40: 348-352, 1974.
205. Zimmer HG and Gerlach E. Stimulation of myocardial adenine nucleotide
biosynthesis by pentoses and pentitols. Pflugers Arch 376: 223-227, 1978.
181
206. Smellie RM, Thomson RY and Davidson JN. The nucleic acid
metabolism of animal cells in vitro. I. The incorporation of 14C-formate.
Biochim Biophys Acta 29: 59-74, 1958.
207. Scott JL. Human leukocyte metabolism in vitro. I. Incorporation of
adenine-8-C-14 and formate-C-14 into the nucleic acids of leukemic
leukocytes. J Clin Invest 41: 67-79, 1962.
208. Holmsen H and Rozenberg MC. Adenine nucleotide metabolism of blood
platelets. 3. Adenine phosphoribosyl transferase and nucleotide formation
from exogenous adenine. Biochim Biophys Acta 157: 266-279, 1968.
209. Wyngaarden JB. The metabolic basis of inherited disease. New York:
McGraw 1: 1-7, 1966.
210. Manfredi JP and Holmes EW. Purine salvage pathways in myocardium.
Annu Rev Physiol 47: 691-705, 1985.
211. Kim YA, King MT, Teague WE, Jr., Rufo GA, Jr., Veech RL and
Passonneau JV. Regulation of the purine salvage pathway in rat liver. Am
J Physiol 262: E344-E352, 1992.
212. Murray AW. The biological significance of purine salvage. Annu Rev
Biochem 40: 811-826, 1971.
213. Namm DH. Myocardial nucleotide synthesis from purine bases and
nucleosides. Comparison of the rates of formation of purine nucleotides
from various precursors and identification of the enzymatic routes for
nucleotide formation in the isolated rat heart. Circ Res 33: 686-695, 1973.
182
214. Kelley WN, Rosenbloom FM and Seegmiller JE. The effects of
azathioprine (imuran) on purine synthesis in clinical disorders of purine
metabolism. J Clin Invest 46: 1518-1529, 1967.
215. Lesch M and Nyhan WL. A familial disorder of uric acid metabolism and
central nervous system function. Am J Med 36: 561-570, 1964.
216. Rosenbloom FM, Kelley WN, Miller J, Henderson JF and Seegmiller
JE. Inherited disorder of purine metabolism. Correlation between central
nervous system dysfunction and biochemical defects. JAMA 202: 175177, 1967.
217. Lipmann F. Enzymatic acetylation and the coenzyme of acetylation. Biol
Bull 91: 239, 1946.
218. Leonardi R, Zhang YM, Rock CO and Jackowski S. Coenzyme A: back
in action. Prog Lipid Res 44: 125-153, 2005.
219. Spry C, Kirk K and Saliba KJ. Coenzyme A biosynthesis: an
antimicrobial drug target. FEMS Microbiol Rev 32: 56-106, 2008.
220. Tahiliani AG and Neely JR. Mitochondrial synthesis of coenzyme A is on
the external surface. J Mol Cell Cardiol 19: 1161-1167, 1987.
221. Basu SS, Mesaros C, Gelhaus SL and Blair IA. Stable isotope labeling
by essential nutrients in cell culture for preparation of labeled coenzyme A
and its thioesters. Anal Chem 83: 1363-1369, 2011.
222. Reibel DK, Wyse BW, Berkich DA, Palko WM and Neely JR. Effects of
diabetes and fasting on pantothenic acid metabolism in rats. Am J Physiol
240: E597-E601, 1981.
183
223. Horie S, Isobe M and Suga T. Changes in CoA pools in hepatic
peroxisomes of the rat under various conditions. J Biochem 99: 13451352, 1986.
224. Lopaschuk GD and Neely JR. Stimulation of myocardial coenzyme A
degradation by fatty acids. Am J Physiol 253: H41-H46, 1987.
225. Snell EE and Mitchell HK. Growth inhibition by N-(dihydroxydimethylbutyryl)taurine and its reversal by pantothenic aicd. J Biol Chem
141: 121-128, 1941.
226. McIlwain H and Hawking F. Chemotherapy of blocking bacterial nutrients
Antistreptococcal activity of pantoyltaurine. Lancet 241: 449-451, 1943.
227. Reichard
P.
Interactions
between
deoxyribonucleotide
and
DNA
synthesis. Annu Rev Biochem 57: 349-374, 1988.
228. Skoog L. An enzymatic method for the determination of dCTP and dGTP
in picomole amounts. Eur J Biochem 17: 202-208, 1970.
229. Walters RA, Tobey RA and Ratliff RL. Cell-cycle-dependent variations of
deoxyribonucleoside triphosphate pools in Chinese hamster cells. Biochim
Biophys Acta 319: 336-347, 1973.
230. Skoog KL, Nordenskjold BA and Bjursell KG. Deoxyribonucleosidetriphosphate pools and DNA synthesis in synchronized hamster cells. Eur
J Biochem 33: 428-432, 1973.
231. Newman CN and Miller JH. Mechanism of UV-induced deoxynucleoside
triphosphate pool imbalance in CHO-K1 cells. Mutat Res 145: 95-101,
1985.
184
232. Das SK and Berger NA. Alterations in deoxynucleoside triphosphate
metabolism in DNA damaged cells: identification and consequences of
poly(ADP-ribose) polymerase dependent and independent processes.
Biochem Biophys Res Commun 137: 1153-1158, 1986.
233. Cohen SS and Barner HD. STUDIES ON UNBALANCED GROWTH IN
ESCHERICHIA COLI. Proc Natl Acad Sci U S A 40: 885-893, 1954.
234. Ayusawa D, Koyama H, Shimizu K, Kaneda S, Takeishi K and Seno T.
Induction, by thymidylate stress, of genetic recombination as evidenced by
deletion of a transferred genetic marker in mouse FM3A cells. Mol Cell
Biol 6: 3463-3469, 1986.
235. Martin DW, Jr. and Gelfand EW. Biochemistry of diseases of
immunodevelopment. Annu Rev Biochem 50: 845-877, 1981.
236. Janssen AJ, Trijbels FJ, Sengers RC, Wintjes LT, Ruitenbeek W,
Smeitink JA, Morava E, van Engelen BG, van den Heuvel LP and
Rodenburg RJ. Measurement of the energy-generating capacity of
human muscle mitochondria: diagnostic procedure and application to
human pathology. Clin Chem 52: 860-871, 2006.
237. Hara KY and Mori H. An efficient method for quantitative determination of
cellular ATP synthetic activity. J Biomol Screen 11: 310-317, 2006.
238. Lemasters JJ and Hackenbrock CR. Continuous measurement and
rapid kinetics of ATP synthesis in rat liver mitochondria, mitoplasts and
inner membrane vesicles determined by firefly-luciferase luminescence.
Eur J Biochem 67: 1-10, 1976.
185
239. Drew B and Leeuwenburgh C. Method for measuring ATP production in
isolated mitochondria: ATP production in brain and liver mitochondria of
Fischer-344 rats with age and caloric restriction. Am J Physiol Regul
Integr Comp Physiol 285: R1259-R1267, 2003.
240. Szendroedi J, Schmid AI, Chmelik M, Toth C, Brehm A, Krssak M,
Nowotny P, Wolzt M, Waldhausl W and Roden M. Muscle mitochondrial
ATP synthesis and glucose transport/phosphorylation in type 2 diabetes.
PLoS Med 4: e154, 2007.
241. Walter G, Vandenborne K, McCully KK and Leigh JS. Noninvasive
measurement of phosphocreatine recovery kinetics in single human
muscles. Am J Physiol 272: C525-C534, 1997.
242. Pedersen BL, Arendrup H, Secher NH and Quistorff B. Metabolism of
perfused pig intercostal muscles evaluated by 31P-magnetic resonance
spectroscopy. Exp Physiol 91: 755-763, 2006.
243. Reibel DK, Wyse BW, Berkich DA and Neely JR. Regulation of
coenzyme A synthesis in heart muscle: effects of diabetes and fasting. Am
J Physiol 240: H606-H611, 1981.
244. Robishaw JD, Berkich D and Neely JR. Rate-limiting step and control of
coenzyme A synthesis in cardiac muscle. J Biol Chem 257: 10967-10972,
1982.
245. Robishaw JD and Neely JR. Pantothenate kinase and control of CoA
synthesis in heart. Am J Physiol 246: H532-H541, 1984.
186
246. Beinlich CJ, Robishaw JD and Neely JR. Metabolism of pantothenic
acid in hearts of diabetic rats. J Mol Cell Cardiol 21: 641-649, 1989.
247. Taupin P. BrdU immunohistochemistry for studying adult neurogenesis:
paradigms, pitfalls, limitations, and validation. Brain Res Rev 53: 198-214,
2007.
248. Taupin P. Protocols for studying adult neurogenesis: insights and recent
developments. Regen Med 2: 51-62, 2007.
249. Gratzner HG. Monoclonal antibody to 5-bromo- and 5-iododeoxyuridine: A
new reagent for detection of DNA replication. Science 218: 474-475, 1982.
250. Nagao T, Kuwagata M and Saito Y. Effects of prenatal exposure to 5bromo-2'-deoxyuridine on the developing brain and reproductive function
in male mouse offspring. Reprod Toxicol 12: 477-487, 1998.
251. Asher E, Payne CM and Bernstein C. Evaluation of cell death in EBVtransformed lymphocytes using agarose gel electrophoresis, light
microscopy and electron microscopy. II. Induction of non-classic apoptosis
("para-apoptosis") by tritiated thymidine. Leuk Lymphoma 19: 107-119,
1995.
252. Rocha B, Penit C, Baron C, Vasseur F, Dautigny N and Freitas AA.
Accumulation of bromodeoxyuridine-labeled cells in central and peripheral
lymphoid organs: minimal estimates of production and turnover rates of
mature lymphocytes. Eur J Immunol 20: 1697-1708, 1990.
253. Schoenheimer R and Rittenberg D. Deuterium as an indicator in the
study of intermediary metabolism. Science 82: 156-157, 1935.
187
254. Dufner D and Previs SF. Measuring in vivo metabolism using heavy
water. Curr Opin Clin Nutr Metab Care 6: 511-517, 2003.
255. Dufner DA, Bederman IR, Brunengraber DZ, Rachdaoui N, IsmailBeigi F, Siegfried BA, Kimball SR and Previs SF. Using 2H2O to study
the influence of feeding on protein synthesis: effect of isotope equilibration
in vivo vs. in cell culture. Am J Physiol Endocrinol Metab 288: E1277E1283, 2005.
256. Busch R, Neese RA, Awada M, Hayes GM and Hellerstein MK.
Measurement of cell proliferation by heavy water labeling. Nat Protoc 2:
3045-3057, 2007.
257. Cook PF, Blanchard JS and Cleland WW. Primary and secondary
deuterium isotope effects on equilibrium constants for enzyme-catalyzed
reactions. Biochemistry 19: 4853-4858, 1980.
258. Billeter SR, Webb SP, Agarwal PK, Iordanov T and Hammes-Schiffer
S. Hydride transfer in liver alcohol dehydrogenase: quantum dynamics,
kinetic isotope effects, and role of enzyme motion. J Am Chem Soc 123:
11262-11272, 2001.
259. Levy LM and Betts GF. Stereospecificity of C4 nicotinamide hydrogen
transfer
of
the
NADP-dependent
glyceraldehyde-3-phosphate
dehydrogenase. Biochim Biophys Acta 997: 331-333, 1989.
260. Babu UM and Johnston RB. D2O-alanine exchange reactions catalyzed
by alanine racemase and glutamic pyruvic transminase. Biochem Biophys
Res Commun 58: 460-466, 1974.
188
261. Bock G, Schumann WC, Basu R, Burgess SC, Yan Z, Chandramouli
V, Rizza RA and Landau BR. Evidence that processes other than
gluconeogenesis may influence the ratio of deuterium on the fifth and third
carbons of glucose: implications for the use of 2H2O to measure
gluconeogenesis in humans. Diabetes 57: 50-55, 2008.
262. Belenky P, Bogan KL and Brenner C. NAD+ metabolism in health and
disease. Trends Biochem Sci 32: 12-19, 2007.
263. Lee WN, Boros LG, Puigjaner J, Bassilian S, Lim S and Cascante M.
Mass isotopomer study of the nonoxidative pathways of the pentose cycle
with [1,2-13C2]glucose. Am J Physiol 274: E843-E851, 1998.
264. Brunengraber H, Sabine JR, Boutry M and Lowenstein JM. 3- Hydroxysterol synthesis by the liver. Arch Biochem Biophys 150: 392-396,
1972.
265. Previs SF, Fatica R, Chandramouli V, Alexander JC, Brunengraber H
and Landau BR. Quantifying rates of protein synthesis in humans by use
of 2H2O: application to patients with end-stage renal disease. Am J
Physiol Endocrinol Metab 286: E665-E672, 2004.
266. Neese RA, Misell LM, Turner S, Chu A, Kim J, Cesar D, Hoh R, Antelo
F, Strawford A, McCune JM, Christiansen M and Hellerstein MK.
Measurement in vivo of proliferation rates of slow turnover cells by 2H2O
labeling of the deoxyribose moiety of DNA. Proc Natl Acad Sci U S A 99:
15345-15350, 2002.
189
267. Varady KA, Hudak CS and Hellerstein MK. Modified alternate-day
fasting and cardioprotection: relation to adipose tissue dynamics and
dietary fat intake. Metabolism 58: 803-811, 2009.
268. Busch R, Kim YK, Neese RA, Schade-Serin V, Collins M, Awada M,
Gardner JL, Beysen C, Marino ME, Misell LM and Hellerstein MK.
Measurement of protein turnover rates by heavy water labeling of
nonessential amino acids. Biochim Biophys Acta 1760: 730-744, 2006.
269. Schoenheimer R. The Investigation of Intermediary Metabolism with the
Aid of Heavy Hydrogen: Harvey Lecture, January 21, 1937. Bull N Y Acad
Med 13: 272-295, 1937.
270. GUTMAN A and SHAFRIR E. Metabolic influences on enzymes of
glycogen synthesis and breakdown in adipose tissue. Am J Physiol 207:
1215-1220, 1964.
271. Lee WN, Bassilian S, Ajie HO, Schoeller DA, Edmond J, Bergner EA
and Byerley LO. In vivo measurement of fatty acids and cholesterol
synthesis using D2O and mass isotopomer analysis. Am J Physiol 266:
E699-E708, 1994.
272. Jungas RL. Fatty acid synthesis in adipose tissue incubated in tritiated
water. Biochemistry 7: 3708-3717, 1968.
273. Wadke M, Brunengraber H, Lowenstein JM, Dolhun JJ and Arsenault
GP. Fatty acid synthesis by liver perfused with deuterated and tritiated
water. Biochemistry 12: 2619-2624, 1973.
190
274. Diraison F, Pachiaudi C and Beylot M. Measuring lipogenesis and
cholesterol synthesis in humans with deuterated water: use of simple gas
chromatographic/mass spectrometric techniques. J Mass Spectrom 32:
81-86, 1997.
275. Ussing HH. The rate of protein renewal in mice and rats studied by
means of heavy hydrogen. Acta Physiol Scand 2: 209-221, 1941.
276. Yuan CL, Sharma N, Gilge DA, Stanley WC, Li Y, Hatzoglou M and
Previs SF. Preserved protein synthesis in the heart in response to acute
fasting and chronic food restriction despite reductions in liver and skeletal
muscle. Am J Physiol Endocrinol Metab 295: E216-E222, 2008.
277. Neese RA, Siler SQ, Cesar D, Antelo F, Lee D, Misell L, Patel K,
Tehrani S, Shah P and Hellerstein MK. Advances in the stable isotopemass spectrometric measurement of DNA synthesis and cell proliferation.
Anal Biochem 298: 189-195, 2001.
278. McCune JM, Hanley MB, Cesar D, Halvorsen R, Hoh R, Schmidt D,
Wieder E, Deeks S, Siler S, Neese R and Hellerstein M. Factors
influencing T-cell turnover in HIV-1-seropositive patients. J Clin Invest
105: R1-R8, 2000.
279. Kim SJ, Cheung S and Hellerstein MK. Isolation of nuclei from labelretaining cells and measurement of their turnover rates in rat colon. Am J
Physiol Cell Physiol 286: C1464-C1473, 2004.
280. Hayes GM, Busch R, Voogt J, Siah IM, Gee TA, Hellerstein MK,
Chiorazzi N, Rai KR and Murphy EJ. Isolation of malignant B cells from
191
patients with chronic lymphocytic leukemia (CLL) for analysis of cell
proliferation: validation of a simplified method suitable for multi-center
clinical studies. Leuk Res 34: 809-815, 2010.
281. Kim SJ, Turner S, Killion S and Hellerstein MK. In vivo measurement of
DNA synthesis rates of colon epithelial cells in carcinogenesis. Biochem
Biophys Res Commun 331: 203-209, 2005.
282. Turner SM, Linfoot PA, Neese RA and Hellerstein MK. Sources of
plasma glucose and liver glycogen in fasted ob/ob mice. Acta Diabetol 42:
187-193, 2005.
283. Kim J, Choi YL, Vallentin A, Hunrichs BS, Hellerstein MK, Peehl DM
and Mochly-Rosen D. Centrosomal PKCbetaII and pericentrin are critical
for human prostate cancer growth and angiogenesis. Cancer Res 68:
6831-6839, 2008.
284. Gillingham MB, Connor WE, Matern D, Rinaldo P, Burlingame T,
Meeuws K and Harding CO. Optimal dietary therapy of long-chain 3hydroxyacyl-CoA dehydrogenase deficiency. Mol Genet Metab 79: 114123, 2003.
285. Landau BR, Wahren J, Chandramouli V, Schumann WC, Ekberg K
and Kalhan SC. Contributions of gluconeogenesis to glucose production
in the fasted state. J Clin Invest 98: 378-385, 1996.
286. Wang B, Sun G, Anderson DR, Jia M, Previs S and Anderson VE.
Isotopologue distributions of peptide product ions by tandem mass
192
spectrometry: quantitation of low levels of deuterium incorporation. Anal
Biochem 367: 40-48, 2007.
287. Voogt JN, Awada M, Murphy EJ, Hayes GM, Busch R and Hellerstein
MK. Measurement of very low rates of cell proliferation by heavy water
labeling of DNA and gas chromatography/pyrolysis/isotope ratio-mass
spectrometric analysis. Nat Protoc 2: 3058-3062, 2007.
288. Exil VJ, Roberts RL, Sims H, McLaughlin JE, Malkin RA, Gardner CD,
Ni G, Rottman JN and Strauss AW. Very-long-chain acyl-coenzyme a
dehydrogenase deficiency in mice. Circ Res 93: 448-455, 2003.
289. Cox KB, Hamm DA, Millington DS, Matern D, Vockley J, Rinaldo P,
Pinkert CA, Rhead WJ, Lindsey JR and Wood PA. Gestational,
pathologic and biochemical differences between very long-chain acyl-CoA
dehydrogenase deficiency and long-chain acyl-CoA dehydrogenase
deficiency in the mouse. Hum Mol Genet 10: 2069-2077, 2001.
290. Kurtz DM, Rinaldo P, Rhead WJ, Tian L, Millington DS, Vockley J,
Hamm DA, Brix AE, Lindsey JR, Pinkert CA, O'Brien WE and Wood
PA. Targeted disruption of mouse long-chain acyl-CoA dehydrogenase
gene reveals crucial roles for fatty acid oxidation. Proc Natl Acad Sci U S
A 95: 15592-15597, 1998.
291. Kelly CL and Wood PA. Cloning and characterization of the mouse shortchain acyl-CoA dehydrogenase gene. Mamm Genome 7: 262-264, 1996.
193