ArticlePDF AvailableLiterature Review

A role for fatty acids and liver fatty acid binding protein in peroxisome proliferation?

Authors:
Biochemical Society Transactions
824
22.
23.
24
25.
26.
27.
28.
20.
30.
31.
32.
33.
34.
35.
30.
K.
J.
(1987) J. Hiol. Chem. 262, 132 19- 13227
Moncecchi,
L).
M.,
I’astusyzn, A.
&
Scallen, T. J.
(1991) J. Biol. Chem. 266,9885-9892
Haker,
M.
E.,
Billheimer,
J.
T.
&
Strauss,
J.
F.
(1991)
DNA Cell Hiol. 10,695-698
He,
Z.,
Yamamoto,
K.,
Furth,
E.
E.,
Schantz,
1,.
J.,
Naylor,
S.
I,.,
George,
H.
&
Billheimer, J. T. (1991)
DNA Cell Hiol. 10,
559-569
Ossendorp,
B.
C., Van-Heudson,
G.
I’,,
L>e
Iker,
A.
I,,
Hos,
K., Schouten,
G.
I,.
&
Wirtz,
K.
W. (1991) Eur. J.
Hiochem. 201,233-239
Mori, T., Tsukamoto, T., Mori,
H.,
Tashiro,
Y.
&
Fu.jiki,
Y.
(1991)
Proc.
Natl. Acad. Sci. [J.S.A.
88,
4338-4342
Seedorf,
V.
&
Assman,
H.
(1
99
1)
J. Hiol. Chem. 266,
630-636
Hillheimer, J. T., Strehl,
I,.
I,.,
Davis,
G.
1,.
&
Strauss,
J.
F.
(1990) DNA Cell Bid.
9,
1
59- 100
Ossendorp,
H.
C., Van-Heudsen,
G.
1’.
&
Wirtz, ti.
W. (1990) Biochem. Hiophys. Kes. Commun. 168,
Kesav,
S.,
Moncecchi,
L).
M.
&
Scallen,
T.
J. (1990)
FASEH
J.
4,732
Ossendorp,
B.
C., Geijtenbeek,
H.
€1.
&
Wirtz,
K.
W.
A. (1992)
FEBS
Lett. 296, 179-183
Duffaud, G.
L).,
March,
1’.
E.
&
Inouvre, M.
(1
987) in
Methods in Enzymology (Wir,
K.
&
Grossman,
I,..
eds.), Academic Press 153,492-507
Szabo,
1,.
J.,
Small, G.
M.
81
1,azarom.
1’.
H.
(1989)
Gene 75.
1
19-126
Huang, C. W.,
Yam,
K.
&
Takaji,
kl.
(100
1)
Gene
106,
h
1-69
Tan. H., Okazaki, K., Kubota,
I.,
Kaniiryo,
T.
&
I
Jtiyma,
H.
(1
990) Eur.
J.
Biochem. 190, 107-
1
12
Gadella,
T.
W. Jr.
&
Wirtz, K.
W.
(1091) Hiochim.
63
1-636
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
Biophys. Acta 1070,237-245
Wirtz, K.
W.
(1001) Klin. Wochenscher
69,
105-1
11
Schroeder,
I;.,
Hutko,
I’.,
Nemecz, G.
&
Scallen,
T.
J.
(1990) J. Biol. Chem. 265, 15
1
-
157
Butko,
P.,
Hapala,
I.,
Scallen,
T.
J.
&
Schroeder,
F.
(1990) Biochemistry 29,4070-4077
Schroeder,
F.,
Hutko,
I’.,
Hapala,
I.
&
Scallen, T. J.
(1990) Lipids 25, 669-677
Billheimer,
J.
T.
&
Gaylor,
J.
I,.
(1990) Hiochem.
Hiophys Acta 1046, 136-1 43
Mendis-Handagama,
S.
M., Watkins,
1’.
A., Gelber,
S.
J.,
Scallen,
T.
J., Zirkin,
H.
K.
&
Ewing,
1,. 1,.
(1
990)
Endocrinology 127,2947-2954
Thompson,
S.
I,.
&
Krisans,
S.
ti. (1990) J. Hiol.
Chem. 265, 5731-5735
Kennert, H., Amsterdam. A., Hillheimer,
J.
T.
81
Strauss. J.
F.
(1
99
1)
Biochemistry 30,
1
1280-
1
1285
Trezciak, W.
11..
Simpson.
E.
K.,
Scallen,
T.
J.?
Vahouny,
G.
V.
&
Waterman,
M.
K.
(1987) J. Hiol.
Chem. 262,3713-3717
Koff>
C.
F.,
I’astuszyn,
A,,
Strauss,
J.
F.
&
Hillheimer,
J.
T. (1002)
J.
Iliol.
C’hem..
267, 15002-
1
SOOX
Van Heudson,
G.
1’.
Il.,
Hos,
K., Kaetz, C.
K.
H.
&
Wirtz,
K.
W. A. (1000)
J.
Hiol. Chem. 265,
4105-41 10
Suzuki,
Y.,
Yamazuchi,
S..
Orii, T., Tsuneoka, M.
&
Tashioro,
Y.
(1000) Cell Struct. Funct. 15, 301-308
Lyons,
H.
T.,
Kharroubi. A., Wolins, N., Tenner,
S..
Chanderbhan,
K.
F.?
Fiskum,
G.
&
Donaldson,
K.
1’.
(1991) Arch. Hiochem. Hiophys. 285,238-245
Kawata,
S..
Imai.
Y..
Inada. M., Inui,
Y.,
Kakimoto, H.,
Fukuda, ti.. Maeda.
Y.
&
Tarui.
S.
(1991) Clin.
Chem.
Acta 197, 20
1-208
Keceived
24
June
1002
IC
A
role for fatty acids and liver fatty acid binding protein in peroxisome
proliferation?
Isabelle Issemann, Rebecca Prince, Jonathan
Tugwood
and Stephen Green*
:I
PLC, Central Toxicology Laboratory, Cell and Molecular Biology Section, Alderley Park, Macclesfield, Cheshire
SK
I0
4TJ,
U.K.
Introduction
I’eroxisome proliferators are
a
diverse group of
chemicals which when administered to rats and
mice produce liver hyperplasia and hypertrophy,
the latter being predominantly due to an increase in
the size and number
of
hepatic peroxisomes
[l].
Interest in these chemicals stems from the observa-
tion that they are rodent liver carcinogens. In par-
Abbreviations used:
I’I’AK.
peroxisomc proliferator
acti-
vated receptor;
I’PKE.
peroxisonie proliferator response
clement;
1;ARI’.
fatty acid binding protein.
‘To whom correspondence should be addressed.
ticular, they are not directly mutagenic and appear
to use
a
novel carcinogenic mechanism. There are
therefore two important research questions: how
do peroxisome proliferators produce liver tumours
in rodents such as rats and mice and what is the
relevance of these findings to man?
The most potent peroxisome proliferators are
hypolipidaemic drugs that were developed for
the
treatment of coronary heart disease
[2].
These
drugs lower circulating levels of cholesterol but are
more effective at lowering triglycerides. They are
therefore used to treat patients with type
IV
hyper-
lipidaemia who have high levels of circulating tri-
Volume
20
Lipid-Binding Proteins
glycerides. The administration of peroxisome
proliferators to rats and mice increases the level of
several hepatic enzymes. These include the peroxi-
soma1 /?-oxidation enzymes such
as
acyl CoA
oxidase, that is responsible for the metabolism of
long chain fatty acids, and the microsomal cyto-
chrome P450 IVA1 that possesses lauric acid
hydroxylase activity. Of particular interest
is
the
finding that the induction of these enzymes
is
regu-
lation, at least in part, at the transcriptional level
We recently identified a member of the
steroid hormone receptor superfamily that can be
activated by peroxisome proliferators
[
51.
We there-
fore termed this the peroxisome proliferator acti-
vated receptor (PPAR). Such receptors are
ligand-activated transcription factors and our more
recent results have demonstrated that PPAR can
bind to a specific DNA sequence (termed the per-
oxisome proliferator response element,
PPRFC)
located upstream
of
the rat acyl CoA oxidase gene
[6].
Therefore PPAR appears
to
mediate the tran-
scriptional effects of peroxisome proliferators. Some
of the questions that remain to be resolved, how-
ever, are how do peroxisome proliferators activate
PPAR and what
is
its natural ligand?
We
present
here some preliminary data indicating a role for
fatty acids in the regulation of PPAR activity and of
PPAR in the regulation of liver fatty acid binding
protein (FABP) gene expression. These data
suggest that one action of peroxisome proliferators
~3~41.
may be to displace fatty acids from FABP, leading
to the activation of PPAR.
Results
Fatty
acids
activate
PPAR
Certain high fat diets induce peroxisome prolifera-
tion in rats [7]. We were therefore interested to
determine whether fatty acids could activate PPAR.
Hepal cells were transfected with the PPAR
expression vector and the ACO(
-
12731
-
471)G.CAT reporter gene that contains the regu-
latory sequences of the rat peroxisomal acyl CoA
oxidase gene upstream of the rabbit /?-globin pro-
moter and the CAT (chloramphenicol transferase)
coding sequence [6]. As shown previously, the per-
oxisome proliferator Wy- 14,643 strongly activated
PPAR at low concentrations producing an increase
in CAT enzyme activity from the ACO-G.CAT
reporter gene (Fig. 1). A variety of fatty acids were
next tested by adding them every 24 h for a total
period of 48 h. As shown in Fig.
1
a
high concentra-
tion
of
palmitic acid stimulated CAT activity indi-
cating that
it
could activate PPAR. We next tested a
thio-substituted fatty acid (tetrathio decanoic acid
[8])
that is structurally similar
to
palmitic acid
except that
it
has a sulphur atom located between
the
/?
and
y
carbon atoms and
is
therefore not
a
substrate for the /?-oxidation pathway. This com-
pound was far more effective than palmitic acid at
activating PPAR suggesting that palmitic acid
is
a
Fig.
I
Fatty acids activate
PPAR
The PPAR expression vector and the acyl CoA oxidase reporter gene were transiently
transfected in Hepal cells in the presence
of
increasing concentrations of palmitic acid,
tetrathio decanoic acid (TTA) or the peroxisome proliferator Wy-
14.643.
The amount of
induced CAT enzyme was measured and expressed as the amoaunt
of
chloramphenicol
acetylated by the cellular extract. Technical details can be found in
[S].
50
C
2
30-
f!
0
-
5
20
~
V
x
zl
10
-
I
E-9
I
E-8
I
E-7
I
E-6
I
E-5 I
E-4
I
E-3
Concentration
(M)
825
I992
Biochemical Society Transactions
826
poor PPAR activator because
it
is rapidly meta-
bolized.
A
consensus
PPRE
in the
FABP
promoter
Both FABP enzyme activity and mRNA levels are
upregulated in primary hepatocyte cultures by per-
oxisome proliferators
[9].
We were interested to
determine whether FABP was regulated at the tran-
scriptional level and in particular whether the
promoter of the FABP gene contained a PPRE. As
seen in Fig. 2, inspection of the
5'
flanking sequence
of the FABP gene
[
101
reveals an imperfect direct
repeat element between
-68
bp and
-56
bp
(S'TGACCTA
TGGCCT
3') that closely resembles
the PPRE (5'TGACC??'TGTCCT 3') identified in
the rat acyl CoA oxidase gene
[6].
To determine
whether this sequence had any role to play in
mediating the effects of peroxisome proliferators we
cloned a region of the rat liver FABP promoter
(-
565
to
+
21) using the polymerase chain reac-
tion and placed this upstream of the coding
sequence of the bacterial enzyme CAT gene. This
reporter gene, pFABP(
-
56.51
+
21)CAT was trans-
fected into a mouse hepatoma cell line (Hepal) in
the absence or presence of a PPAR expression
vector (pSG5-PPAR) and the potent peroxisome
proliferator Wy- 14,643. Disappointingly, the
presence of PPAR and Wy-14,643 had no effect
upon CAT enzyme activity indicating that the
receptor was unable
to
stimulate the transcription of
the FABP promoter.
To
investigate whether this
was because the PPRE-like sequence was unable to
respond to PPAR or whether
it
was due
to
the con-
text of this sequence in the FABP promoter, we
constructed a second reporter gene pFABP(
-
5651
-5O)G.CAT in which the FABP sequences from
-
565
bp to
-
50
bp were placed upstream of the
heterologous rabbit /3-globin promoter (Fig. 2).
Using this reporter gene in transfection assays
demonstrated that
it
was responsive to the presence
Fig.
2
FABP
reporter
genes
The
5'
regulatory sequences of the rat liver fatty acid binding protein gene are shown schematically. The start of transcription
is
indicated with an arrow at
+
I
and the PPRE-like sequence is underlined between -68bp and -56bp. Two reporter gene
constructs were created using the rabbit /3-globin promoter
(
-
I
10
to
+
20)
and two using the natural FABP promoter. These were
tested in transient transfection assays using Hepal cells in the absence or presence
of
the PPAR receptor expression vector (R) and
the peroxisome proliferator Wy-
14,643
(Wy). The values on the right indicate the percentage
of
chloramphenicol acetylated by the
cell extract. Technical details can be found in [6]. Abbreviation used: TTA, tetrathio decanoic acid.
68
-56
Rat
FABP
gene
and
promoter
CAATCACTWCCTATCATATTT
-6
-565
FABP
-50
Globin
CAT
1% 10%
1% 1%
-565
68
+21
pFABP(-565/+21)CAT
J
665
68
+21
e
pFABP(-565/+21 PPRE)CAT
CAATCACTTCTACiATCATATATTT
-
--
1%
1%
1% 1%
Volume
20
Lipid-Binding Proteins
of PPAR and Wy-14,643 (Fig. 2). In addition, a
third reporter gene, pFABP(
-
565/
-
68)G.CAT,
which does not contain the PPRE-like sequence
was not stimulated by PPAR in the presence of
Wy-14,643. These data therefore suggest that
PPAR can recognize and activate the PPRE-like
sequence present in the FABP promoter but that
this
sequence
is
not functional in the context of the
natural promoter when transfected into Hepa 1 cells.
Discussion
A number of potent peroxisome proliferators are
hypolipidaemic drugs and lower the level of circulat-
ing triglycerides in man. These chemicals induce
enzymes that are important in fatty acid metabolism
such as peroxisomal acyl CoA oxidase and cyto-
chrome P450
IVA1.
Furthermore, high fat diets can
induce peroxisome proliferation in rats and per-
oxisome proliferators can elevate the level of rat
liver FABP. Taken together, therefore, these data
suggest an important role of PPAR in regulating
fatty acid homeostasis.
Examination of the
5’
regulatory sequences of
the rat liver FABP indicated the presence of a
sequence that
is
almost identical to the PPRE iden-
tified in the
5’
regulatory sequences of the rat acyl
CoA oxidase gene. This putative response element
appears to be inactive in the context of the natural
FABP promoter but is active when placed upstream
of the rabbit B-globin promoter (Fig. 2). One reason
for the failure of PPAR to activate the FABP-CAT
reporter gene may be due to TATA binding pro-
teins that bind to the adjacent TATA box sequence
interfering with PPAR binding to the
PPRE.
Alter-
natively,
it
may be that Hepal cells lack a co-factor
required for PPAR activity when the PPRE
is
present in the context of the natural FABP promo-
ter. Clearly, further work
is
required to determine
how peroxisome proliferators regulate FABP
expression.
Of particular interest was the finding that fatty
acids are weak PPAR activators ([ 113, Fig. 1). One
explanation for this weak activity is that fatty acids
are metabolized rapidly by the /?-oxidation path-
way. This
is
supported by the observation that the
thio-substituted fatty acid that is blocked for
/?-
oxidation is a much better PPAR activator (Fig. 1).
We and others have found that a wide range of fatty
acids are able to activate PPAR ([ 113, I. Issemann
and R. Prince, unpublished results). An important
question, therefore, is whether any or all of these
fatty acids can bind directly to PPAR or alterna-
tively whether they are metabolized to the ligand or
induce the ligand. These data also raise the question
of how synthetic peroxisome proliferators activate
PPAR. One possibility is that they bind directly to
the receptor in the same way that antioestrogens
such as tamoxifen bind to the oestrogen receptor.
However, thus far we have been unable to
827
demonstrate any interaction between PPAR and the
peroxisome proliferator nafenopin
[
51.
An interest-
ing alternative mechanism for peroxisome prolifera-
tor action is that these chemicals bind to FABP
displacing fatty acids that activate PPAR. Others
[12] have demonstrated that a number of per-
oxisome proliferators can bind to FABP and dis-
place the oleic acid used to monitor the purification
of FABP. It will therefore be important to determine
whether peroxisome proliferators act by simply
competing with fatty acids for binding to FABP
leading to an elevation of fatty acids that activate
PPAR, or alternatively whether they and fatty acids
can bind directly to PPAR. These experiments and
others that determine the true nature of the PPAR
ligand will be important in understanding the role of
peroxisome proliferators and PPAR in hypo-
lipidaemia and rodent liver cancer.
We thank Jon Bremer
for
providing the tetrathio
decanoic acid used in these studies.
1. Green,
S.
(1992) Biochem. Pharmacol. 43, 393-401
2. Havel, R. J.
&
Kane, J. P. (1973) Annu. Rev. Pharma-
col. 13,287-308
3. Keddy,
J.
K., Goel,
S.
K., Nemali, M. K., Carrino, J.
J.,
Laffler,
T.
G., Reddy, M. K., Sperbeck,
S.
J., Osumi,
T., Hashimoto, T., Lalwani, N.
D.
&
Kao,
M.
S.
(1986)
Proc. Natl. Acad. Sci. U.S.A.
83,
1747-1751
4. Hardwick, J. P., Song,
B.
J., Huberman, E.
&
Gonzalez,
F.
J. (1987) J. Biol. Chem. 262,801-810
5. Issemann,
I.
&
Green,
S.
(1990) Nature 347,645-650
6. Tugwood, J.
D.
Issemann,
I.,
Anderson,
R.
G.,
Bundell, K.
R.,
McPheat, W. L.
&
Green,
S.
(1992)
EMBO J. 11,433-439
7. Brandes,
R.,
Kaikaus, R.
M.,
Lysenko,
N.,
Ockner,
K. K.
&
Bass, N.
M.
(1990) Biochem. Biophys. Acta
1034,5341
8. Sweetser,
D.
A,, Lowe,
J.
R.
&
Gordon, J.
I.
(1986)
J. Hiol. Chem. 261,5553-5561
9. Flatmark,
T.,
Nilsson, A,, Kvannes, J. Eikhom,
T.
S.,
Fukami, M. H., Kryvi, H.
&
Christiansen, E. N.
(1988) Biochem. Biophys. Acta
962,
122-130
10. Spydevold,
0.
&
Bremer, J. (1989) Biochem. Biophys.
Acta 1003,72-79
11. Gottlicher,
M.,
Widmark,
E.,
Li,
Q.
&
Gustafsson,
J.-A. (1992) Proc.
Natl.
Acad. Sci. U.S.A.
89,
4653-4657
12. Cannon, J. R.
&
Eacho,
P.
I.
(1991) Biochem. J. 280,
387-391
Received 25 June 1992
I992
... beyond DR1 was not revealed (4)(5)(6)(7). Subsequently, based on close to 30 PPREs, a 5=-(A)AACT-3= sequence directly upstream of DR1s was identified (8,9). These sporadically identified extended DR1s proved later to be the strongest PPREs in biochemical assays (7,10,11). ...
... Subsequently, based on close to 30 PPREs, a 5=-(A)AACT-3= sequence directly upstream of DR1s was identified (8,9). These sporadically identified extended DR1s proved later to be the strongest PPREs in biochemical assays (7,10,11). More recently, the importance of the 5= extension was also confirmed by biotin-labeled double-stranded DNA pulldown assays (12) and microfluidics-based dimer-DNA interaction assays (13). ...
Article
PPARγ is a nuclear receptor essential for adipocyte development and the maintenance of the alternatively polarized macrophage phenotype. Biochemical studies have established that as an obligate heterodimer with RXR, PPARγ binds directly repeated nuclear receptor half sites spaced by one nucleotide (DR1). However, it has not been systematically and genome-wide analyzed how cis factors such as the sequence of DR1s and adjacent sequences, and trans factors such as co-binding linage-determining transcription factors (LDTFs) contribute to the direct binding of PPARγ in different cellular contexts. We developed a novel motif optimization approach using sequence composition and ChIP-seq densities from macrophages and adipocytes to complement de novo motif enrichment analysis and to define and classify high-affinity binding sites. We found that approximately half of the PPARγ cistrome represents direct DNA binding; both half sites can be extended upstream, and these are typically not of equal strength within a DR1. Strategically positioned LDTFs have greater impact on PPARγ binding than the quality of DR1, and the presence of the extension of DR1 provides a remarkable synergy with LDTFs. This approach of considering not only nucleotide frequencies but also their contribution to protein binding in a cellular context, is applicable to other transcription factors.
... When the lipid set was used as a query to determine which enzymes had altered activities ( Figure S1), predominantly reactions generating CoA were elevated (namely acyl-CoA thioesterases (ACOT2), carnitine phosphotransferases (CPT1), and 3-hydroxymethyl-3-methylglutaryl-CoA lyase) and reactions consuming CoA were diminished. Linoleate and oleate are activating ligands of peroxisome-proliferator activating receptor-alpha (PPARa), which stimulates peroxisome formation and beta-oxidation [14] HMDB0000673 HMDB0000207 HMDB0000220 ...
Article
Full-text available
Pre-eclampsia is the most common pregnancy complication affecting 1 in 20 pregnancies, characterized by high blood pressure and signs of organ damage, most often to the liver and kidneys. Metabolic network analysis of published lipidomic data points to a shortage of Coenzyme A (CoA). Gene expression profile data reveal alterations to many areas of metabolism and, crucially, to conflicting cellular regulatory mechanisms arising from the overproduction of signalling lipids driven by CoA limitation. Adverse feedback loops appear, forming sphingosine-1-phosphate (a cause of hypertension, hypoxia and inflammation), cytotoxic isoketovaleric acid (inducing acidosis and organ damage) and a thrombogenic lysophosphatidyl serine. These also induce mitochondrial and oxidative stress, leading to untimely apoptosis, which is possibly the cause of CoA restriction. This work provides a molecular basis for the signs of pre-eclampsia, why polycystic ovary syndrome is a risk factor and what might be done to treat and reduce the risk of disease.
... As the central factor of research, LvFABP also was inhibited significantly at the environmental concentration CFA. However, a potent PPARα agonist ligand like bezafibrate dramatically upregulates transcription of L-FABP in mice by activating PPARα and subsequent binding of the activated receptor to the PPREs in the promoter region of L-FABP (Issemann et al. 1992;Seedorf et al. 1998;Wolfrum et al. 1999). CFA with the environmental concentration (4 µg/L) in our study may lead to the gene expression inhibition compared with other concentrations of peroxisome proliferators. ...
Article
Full-text available
Clofibric acid (CFA), a drug and personal care product, has been identified as ubiquitous in the aquatic system and surface water, causing pollution to the environment. In this study, after environmental (4 µg/L) levels of CFA challenge, the LvFABP, LvACS gene expressions, total haemocyte count (THC), relative enzymes (SOD1 and GST) activities in Litopenaeus vannamei were observed to decrease. In the meantime LvFATP, LvRXR expression and the level of NEFA were upregulated in L. vannamei body. LvFABP expression in vivo was knocked down by dsRNA-mediated RNA interference (RNAi), which led to significantly decreased levels of PPARα (including LvFATP, LvRXR and LvACS). When exposed to environmental CFA after 4 days, LvFABP knocked down group had a sharp upregulation of LvFATP, LvRXR, LvACS expression, GST activity and NEFA amount, following decreased THC and SOD1 activity. These results suggested that environmental concentration CFA may have some toxicological effect on L. vannamei, following fatty acids metabolism and oxidative stress responses by LvFABP via the PPARα/RXR signaling pathway, including LvFATP, LvRXR and LvACS.
... 65 A toxicodynamic explanation for our observations may be a more effective activation of PPARγ by long-chain FAs. 66 Similar to our finding, Intrasuksri et al. (1988) 65 detected higher PPAR induction potency for FAs in decreasing order from oleic acid (C16) > octanoic acid > octanedioic acid. The neutral form can only passively permeate through the membrane which is impermeable for the anionic form of the fatty acids. ...
Article
Full-text available
Standard ecotoxicological testing of microplastic does not provide insight into the influence that environmental weathering by, e.g., UV light has on related effects. In this study, we leached chemicals from plastic into artificial sea water during simulated UV-induced weathering. We tested largely additive-free pre-production polyethylene, polyethylene terephthalate, polypropylene and polystyrene and two types of plastic obtained from electronic equipment as positive controls. Leachates were concentrated by solid-phase extraction and dosed into cell-based bioassays that cover i) cytotoxicity; ii) activation of metabolic enzymes via binding to the arylhydrocarbon receptor (AhR) and the peroxisome proliferator-activated receptor (PPARγ); iii) specific, receptor-mediated effects (estrogenicity, ER); and iv) adaptive response to oxidative stress (AREc32). LC-HRMS analysis was used to identify possible chain-scission products of polymer degradation, which were then tested in AREc32 and PPARγ. Explicit activation of all assays by the positive controls provided proof-of-concept of the experimental setup to demonstrate effects of chemicals liberated during weathering. All plastic leachates activated the oxidative stress response, in most cases with increased induction by UV-treated samples compared to dark controls. For PPARγ, polyethylene-specific effects were partially explained by the detected dicarboxylic acids. Since the pre-production plastic showed low effects often in the range of the blanks future studies should investigate implications of weathering on end consumer products containing additives.
... In fact, in animals, the deletion of the L-FABP gene resulted in increased expression of E-FABP and cholesterol accumulation [59]. FABPs also facilitate the polyunsaturated FA-induced transactivation of PPARα and PPARγ via direct interaction with the nuclear receptor's ligand binding domain [60][61][62][63][64]. FAs also induce FABP expression [65]. ...
Article
Full-text available
Obesity and hyperlipidemia are the most prevalent independent risk factors of chronic kidney disease (CKD), suggesting that lipid accumulation in the renal parenchyma is detrimental to renal function. Non-esterified fatty acids (also known as free fatty acids, FFA) are especially harmful to the kidneys. A concerted, increased FFA uptake due to high fat diets, overexpression of fatty acid uptake systems such as the CD36 scavenger receptor and the fatty acid transport proteins, and a reduced β-oxidation rate underlie the intracellular lipid accumulation in non-adipose tissues. FFAs in excess can damage podocytes, proximal tubular epithelial cells and the tubulointerstitial tissue through various mechanisms, in particular by boosting the production of reactive oxygen species (ROS) and lipid peroxidation, promoting mitochondrial damage and tissue inflammation, which result in glomerular and tubular lesions. Not all lipids are bad for the kidneys: polyunsaturated fatty acids (PUFA) such as eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) seem to help lag the progression of chronic kidney disease (CKD). Lifestyle interventions, especially dietary adjustments, and lipid-lowering drugs can contribute to improve the clinical outcome of patients with CKD.
Article
Activation of hepatic stellate cells (HSCs) is the main course of liver fibrosis which is positively correlated with adverse clinical outcomes in non-alcoholic steatohepatitis (NASH). Diethyldithiocarbamate (DDC) attenuates NASH related liver fibrosis in mice, but its underlying mechanisms remains unclear. In this study, the data showed that DDC inhibited the activation of HSCs in high fat choline-deficient, L-amino acid-defined (CDAA) diet induced NASH. Double Immunofluorescence analysis showed that the baseline expression of peroxisome proliferator-activated receptor α (PPARα) is high in HSCs in normal mouse liver and notably decreases in the NASH liver, indicating that PPARα might be associated with the activation of HSCs. While, DDC upregulated PPARα in HSCs in the NASH liver. Mixture of free fatty acid was used to induce steatosis of hepatocytes. Human HSCs (LX-2 cells) were activated after co-cultured with steatotic hepatocytes, and DDC inhibited the activation of LX-2 cells. Meanwhile, DDC upregulated PPARα and FABP1, and promoted the accumulation of LDs in LX-2 cells. PPARα small interfering RNA blocked these effect of DDC. These findings suggest that PPARα is associated with the activation of HSCs in the context of NASH. DDC improves NASH related fibrosis through inhibiting the activation of HSCs via PPARα/FABP1.
Preprint
Full-text available
Pre-eclampsia is the most common pregnancy complication affecting 1 in 20 pregnancies, characterized by high blood pressure and signs of organ damage, most often to the liver and kidneys. Metabolic network analysis of published lipidomic data points to a shortage of Coenzyme A (CoA). Gene-expression profile data reveal alterations to many areas of metabolism and, crucially, to conflicting cellular regulatory mechanisms arising from the overproduction of signalling lipids driven by CoA limitation. Adverse feedback loops appear, forming sphingosine-1-phosphate (a cause of hypertension, hypoxia and inflammation), cytotoxic isoketovaleric acid (inducing acidosis and organ damage) and a thrombogenic lysophosphatidyl serine. These also induce mitochondrial and oxidative stress, leading to untimely apoptosis, which is possibly the cause of CoA restriction. This work provides a molecular basis for the signs of pre-eclampsia, why other conditions are risk factors and what might be done to treat and reduce the risk of this and related diseases.
Article
Polyunsaturated fatty acids (PUFA) influence many physiological functions. Associations have been found between single nucleotide polymorphisms (SNP) in the FADS1 (Fatty acid desaturase 1) gene and the relative abundance of PUFA in serum lipids. This study examines the relationship between two SNPs in the FADS1 gene (rs174546, rs174537) and the fatty acid (FA) composition of serum lipids in adolescents (13-18 years). We used DNA samples (670 children; 336 girls and 334 boys) from the Childhood Obesity Prevalence and Treatment (COPAT) project. Genomic DNA was extracted from peripheral blood leukocytes in whole blood samples. For genotype analysis, TaqMan SNP Genotyping assays (Applied Biosystems) were used. Fatty acid composition of serum lipids was assessed using gas chromatography. The T-statistic and regression were used for statistical evaluations. Minor allele T carriers in both SNPs had significant lower level of palmitic acid (16:0, phospholipids) and arachidonic acid (20:4[n-6], phospholipids) in both sexes. In girls, we found a significant positive association between minor allele T carriers and eicosadienoic acid (20:2[n-6], cholesteryl esters) in both SNPs. Being a minor allele T carrier was significantly positively associated with dihomo-γ-linolenic acid (20:3[n-6], phospholipids) in boys in both SNPs. SNPs (including rs174546, rs174537) in the FADS gene cluster should have impacted desaturase activity, which may contribute to different efficiency of PUFA synthesis.
Article
Full-text available
Peroxisome proliferators are a diverse group of chemicals, including several hypolipidaemic drugs, that activate a nuclear hormone receptor termed the peroxisome proliferator activated receptor (PPAR). The peroxisomal enzyme acyl CoA oxidase (ACO) is the most widely used marker of peroxisome proliferator action. We have examined the 5' flanking region of the rat ACO gene for sequences that mediate the transcriptional effect of peroxisome proliferators and have identified an element located 570 bp upstream of the ACO gene that confers responsiveness to the hypolipidaemic peroxisome proliferator Wy-14,643. This peroxisome proliferator response element (PPRE) contains a direct repeat of the sequence motifs TGACCT and TGTCCT and binds PPAR. These data therefore indicate an important role of PPAR in mediating the action of peroxisome proliferators including the induction of ACO.
Article
Full-text available
Lauric acid omega-hydroxylase (cytochrome P-450LA omega) was purified from livers of rats that had been given the hypolipidemic drug clofibrate. Immunoblot analysis with anti-cytochrome P-450LA omega antibody revealed the presence of two clofibrate-induced cytochrome P-450 proteins in both liver and kidney. This antibody was used to screen a lambda gt11 expression library constructed from liver mRNA isolated from rats given clofibrate. The clone, pP-450LA omega, was isolated and found to contain 2062 base pairs with an open reading frame of 509 amino acids (Mr 58,222). The pP-450LA omega cDNA insert was placed in the yeast expression vector pAAH5, and the resultant plasmid expressed a protein of the same size and activity as cytochrome P-450LA omega. The mechanism of the regulation of the cytochrome P-450LA omega gene by hypolipidemic agents was examined. The rate of cytochrome P-450LA omega gene transcription was increased as early as 1 h after administration of clofibrate, and this increase was followed by an elevation of cytochrome P-450LA omega mRNA, immunochemically detectable protein, and cytochrome P-450LA omega activity. In contrast, no increase in transcriptional activity was detected after clofibrate administration for the cytochrome P-450PCN or P-450b/e genes. The cytochrome P-450LA omega has several structural features in common with other cytochrome P-450s, including a conserved cysteine-containing heme-binding fifth ligand fragment and a hydrophobic amino terminus. Overall, cytochrome P-450LA omega shared less than 35% cDNA nucleotide and amino acid similarity with cytochromes P-450c, P-450d, P-450e, and P-450PCN, indicating that it is a member of a new cytochrome P-450 gene family. Southern blot analysis indicates that this family contains two or three genes.
Article
We have cloned a member of the steroid hormone receptor superfamily of ligand-activated transcription factors. The receptor homologue is activated by a diverse class of rodent hepatocarcinogens that causes proliferation of peroxisomes. Identification of a peroxisome proliferator-activated receptor should help elucidate the mechanism of the hypolipidaemic effect of these hepatocarcinogens and aid evaluation of their potential carcinogenic risk to man.
  • P Butko
  • I Hapala
  • T J Scallen
  • F Schroeder
Butko, P., Hapala, I., Scallen, T. J. & Schroeder, F. (1990) Biochemistry 29,4070-4077
  • S I Thompson
  • S Krisans
Thompson, S. I,. & Krisans, S. ti. (1990) J. Hiol.
  • K F Chanderbhan
  • G Fiskum
  • K Donaldson
Chanderbhan, K. F.? Fiskum, G. & Donaldson, K. 1'. (1991) Arch. Hiochem. Hiophys. 285,238-245
  • S Kawata
  • Y Imai
  • M Inada
  • Y Inui
  • H Kakimoto
  • Ti Fukuda
Kawata, S.. Imai. Y.. Inada. M., Inui, Y., Kakimoto, H., Fukuda, ti.. Maeda. Y. & Tarui. S. (1991) Clin. Chem.
  • R J Havel
  • J P Kane
Havel, R. J. & Kane, J. P. (1973) Annu. Rev. Pharmacol. 13,287-308
  • J P Hardwick
  • B J Song
  • E Huberman
  • F J Gonzalez
Hardwick, J. P., Song, B. J., Huberman, E. & Gonzalez, F. J. (1987) J. Biol. Chem. 262,801-810