Ramasamy 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Clinica Chimica Acta 454 (2016) 143–185

Contents lists available at ScienceDirect

Clinica Chimica Acta

journal homepage: www.elsevier.com/locate/clinchim

Update on the molecular biology of dyslipidemias


I. Ramasamy
Department of Biochemistry, Worcester Royal Hospital, Worcester, United Kingdom

a r t i c l e i n f o a b s t r a c t

Article history: Dyslipidemia is a commonly encountered clinical condition and is an important determinant of cardiovascular
Received 2 September 2015 disease. Although secondary factors play a role in clinical expression, dyslipidemias have a strong genetic compo-
Received in revised form 24 October 2015 nent. Familial hypercholesterolemia is usually due to loss-of-function mutations in LDLR, the gene coding for low
Accepted 30 October 2015
density lipoprotein receptor and genes encoding for proteins that interact with the receptor: APOB, PCSK9 and
Available online 4 November 2015
LDLRAP1. Monogenic hypertriglyceridemia is the result of mutations in genes that regulate the metabolism of tri-
Keywords:
glyceride rich lipoproteins (eg LPL, APOC2, APOA5, LMF1, GPIHBP1). Conversely familial hypobetalipoproteinemia
Dyslipidemia is caused by inactivation of the PCSK9 gene which increases the number of LDL receptors and decreases plasma
Familial hypercholesterolemia cholesterol. Mutations in the genes APOB, and ANGPTL3 and ANGPTL4 (that encode angiopoietin-like proteins
Autosomal recessive hypercholesterolemia which inhibit lipoprotein lipase activity) can further cause low levels of apoB containing lipoproteins.
Lipoprotein lipase deficiency Abetalipoproteinemia and chylomicron retention disease are due to mutations in the microsomal transfer protein
Hypertriglyceridemia and Sar1b-GTPase genes, which affect the secretion of apoB containing lipoproteins. Dysbetalipoproteinemia
Dysbetalipoproteinemia stems from dysfunctional apoE and is characterized by the accumulation of remnants of chylomicrons and
Familial hypobetalipoproteinemia
very low density lipoproteins. ApoE deficiency can cause a similar phenotype or rarely mutations in apoE can
Abetalipoproteinemia
be associated with lipoprotein glomerulopathy. Low HDL can result from mutations in a number of genes regu-
Familial combined hypolipidemia
Chylomicron retention disease lating HDL production or catabolism; apoAI, lecithin: cholesterol acyltransferase and the ATP-binding cassette
Tangier disease transporter ABCA1. Patients with cholesteryl ester transfer protein deficiency have markedly increased HDL cho-
lesterol. Both common and rare genetic variants contribute to susceptibility to dyslipidemias. In contrast to rare
familial syndromes, in most patients, dyslipidemias have a complex genetic etiology consisting of multiple genet-
ic variants as established by genome wide association studies. Secondary factors, obesity, metabolic syndrome,
diabetes, renal disease, estrogen and antipsychotics can increase the likelihood of clinical presentation of an in-
dividual with predisposed genetic susceptibility to hyperlipoproteinemia. The genetic profiles studied are far
from complete and there is room for further characterization of genes influencing lipid levels. Genetic assessment
can help identify patients at risk for developing dyslipidemias and for treatment decisions based on ‘risk allele’
profiles. This review will present the current information on the genetics and pathophysiology of disorders
that cause dyslipidemias.
© 2015 Elsevier B.V. All rights reserved.

Contents

1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
1.1. Overview of lipid metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
1.2. Regulation of triglyceride production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
1.2.1. Hormonal control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
1.2.2. Insulin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
1.2.3. Glucagon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
1.3. Nuclear hormone receptors and lipid metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
2. Inherited dyslipidemias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
2.1. Increase in concentrations of lipid or lipoprotein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
2.1.1. Familial hypercholesterolemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
2.2. Primary hypertriglyceridemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
2.2.1. Triglycerides and risk of disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
2.3. Primary monogenic hypertriglyceridemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
2.3.1. LPL deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
2.3.2. ApoCII deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
2.3.3. LMF1 deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

http://dx.doi.org/10.1016/j.cca.2015.10.033
0009-8981/© 2015 Elsevier B.V. All rights reserved.
144 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

2.3.4. GPIHBP1 deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155


2.3.5. ApoAV deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
2.3.6. LPL inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
2.3.7. Polygenic causes of hypertriglyceridemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
2.3.8. Novel rare genetic diseases associated with hypertriglyceridemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
2.4. ApoE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
2.4.1. ApoE and lipoprotein metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
2.4.2. Pathophysiology of dysbetalipoproteinemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
2.4.3. Clinical presentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
2.4.4. Lipoprotein glomerulopathy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
2.5. Decrease in concentration of lipid or lipoprotein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
2.5.1. Familial hypobetalipoproteinemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
2.5.2. Abetalipoproteinemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
2.5.3. Familial combined hypolipidemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
2.5.4. Other candidate genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
2.5.5. Chylomicron retention disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
2.5.6. ApoCIII deficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
3. Disorders of HDL metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
3.1. HDL and cardiovascular disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
3.2. Decrease in HDL concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
3.2.1. Monogenic disorders associated with decreased HDL levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
3.3. Increased HDL concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
3.3.1. Monogenic disorders associated with increased HDL levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
3.4. Polygenic causes of HDL variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
4. Secondary dyslipidemias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
4.1. Obesity, metabolic syndrome and diabetes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
4.2. Hypothyroidism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
4.3. Renal disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
4.4. Pregnancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
5. Treatment of dyslipidemias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
6. Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

1. Background ATP-binding cassette (ABC) transporter protein, ABCA1 is the key cellu-
lar membrane protein for the generation of new HDL particles. Further
1.1. Overview of lipid metabolism cholesterol enrichment of these particles is through ABCG1 and scaven-
ger receptor class B type I (SR-BI). In the case of SR-BI the flux of choles-
Understanding the mechanisms of lipoprotein metabolism has terol is bidirectional. In humans, HDL cholesterol can be returned to the
important clinical indications as lipoproteins are risk factors for liver either by direct uptake by SR-BI or through cholesteryl ester trans-
atherosclerosis. In both liver and intestine, apolipoprotein B (apoB) is fer protein (CETP) exchange of cholesteryl ester for triglycerides (TG) in
cotranslationally translocated to the endoplasmic reticulum (ER) apoB lipoproteins, followed by hepatic uptake of apoB containing parti-
lumen where facilitated by microsomal transfer protein (MTP), lipid is cles. In the circulation several proteins and enzymes modulate HDL. Lec-
added to apoB to form primordial apoB containing particles. In humans ithin: cholesterol acyltransferase (LCAT) esterifies free cholesterol into
apoB48 is secreted exclusively by the intestine in chylomicrons and cholesteryl esters that are hydrophobic and therefore sequestered into
apoB100 is secreted exclusively by the liver in very low density lipopro- the core of particles. LCAT is activated primarily by apoAI but can also
teins (VLDL). ApoB100 has 4536 amino acids. ApoB48 is identical to the be activated by apoAIV, apoCI and apoE [1]. It has been postulated that
amino-terminal 48% of the apoB100 and is produced by a unique mRNA upon esterification of cholesterol in HDL the gradient of free cholesterol
editing process. Cotranslational targeting of apoB for degradation is reg- between the cellular membrane and the surface of the HDL particle is
ulated by the availability of lipids. Newly formed but defective apoB100 maintained. This is thought to generate a continuous flow of cholesterol
particles are cleared by mechanisms including ER associated degrada- from the cell to lipoproteins. CETP transfers cholesteryl esters from HDL
tion or autophagosome formation. Both VLDL and chylomicrons (CM) to apoB containing proteins. Phospholipid transfer protein (PLTP) trans-
require specialized vesicles to traffic from ER through to the Golgi fers phospholipid between HDL and VLDL as well as different HDL par-
prior to being secreted. ticles (Fig. 4). The reader is referred to several reviews including those
The lipase maturation factor 1 (LMF1), the glycosylphosphati by the author for a more detailed discussion on physiological lipopro-
dylinositol-anchored high density lipoprotein binding protein 1 tein metabolism [2,3].
(GPIHBP1) and an angiopoietin like protein play a role in lipoprotein li- Experimental data suggest that the exchangeable lipoproteins not
pase (LPL) mediated hydrolysis of TG so that the right amount of fatty only play a role extracellularly (in the circulation) in modulating
acids is delivered to the right tissue at the right time. Hydrolysis of CM lipoprotein clearance but also have an impact on the rate and effi-
and VLDL by LPL results in chylomicron remnants, intermediate density ciency of lipoprotein assembly and secretion. Exchangeable lipopro-
lipoproteins (IDL) and low density lipoproteins (LDL) particles, which teins are those capable of moving from one lipoprotein particle to
bind to specific receptors in the liver, the LDL receptor (LDLR), VLDL another (apoAI, AII, AIV, AV, CI, CII, CIII) and as opposed to the non-
receptor and LDL receptor related proteins (LRP) prior to removal exchangeable apolipoproteins that remain attached to one lipopro-
from the plasma (Figs. 1, 2 and 3). tein molecule from synthesis and catabolism (apoB100 and apoB48)
Reverse cholesterol transport occurs when lipid free apoAI and apoE (Table 1). The exchangeability of the apolipoproteins is attributed
recruit cholesterol and phospholipid to assemble HDL particles. The to the high content of their amphipathic α-helical structure which
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 145

Fig. 1. Pathways of triglyceride rich lipoprotein metabolism. Intestinal enterocytes absorb dietary lipids and package most of it into chylomicrons. Endothelial bound LPL hy-
drolyses triglycerides in CM to generate chylomicron remnants which are rapidly cleared by the liver. Intrahepatic lipids are repackaged and secreted as VLDL which are sub-
strates for LPL. Glycosylphosphatidylinositol-anchored high density lipoprotein binding protein 1 (GPIHBP1) plays a role in LPL mediated hydrolysis of TG. Remnants formed
by hydrolysis of VLDL are either taken up by the liver or converted to LDL. The hepatic LDL receptor (LDLR) is responsible for clearing LDL from human plasma. The LDL re-
ceptor related lipoprotein (LRP) has a role in triglyceride rich lipoprotein clearance. In addition heparan sulfate proteoglycans (HSPG) act in the clearance of triglyceride-rich
lipoproteins.

displays different affinities towards phospholipid on the surface of regulating whole body TG flux when dietary TG are entering the sys-
lipoproteins [4]. Accumulating evidence suggests that exchangeable temic circulation via chylomicrons.
lipoproteins play a role in lipid mobilization and recruitment during Acute increases in insulin can decrease the secretion of apoB by
lipoprotein assembly and trafficking through the ER-Golgi secretory targeting the protein for intracellular degradation. Insulin suppresses he-
compartments and that this role is independent of their role in lipo- patic VLDL secretion through activation of phosphatidylinositol 3 kinase
protein clearance [5]. Recent suggestions are that variations in sin- (PI3K). Binding of insulin to hepatic insulin receptors, leads to activation
gle nucleotide polymorphisms (SNP) in the ApoAI/CIII/AIV/AV gene of insulin receptor tyrosine kinase and tyrosine phosphorylation of in-
cluster are associated with altered postprandial lipid metabolism sulin receptor substrates. Tyrosine phosphorylated insulin receptor
[6]. substrates complex with and activate PI3K. Activated PI3K interacts
with its substrate phosphotidylinositide [4,5] bisphosphate with the
1.2. Regulation of triglyceride production formation of phosphatidylinositide [3,4,5]triphosphate. Results suggest
that insulin signaling through PI3K inhibited the maturation phase of
1.2.1. Hormonal control VLDL [25]. Insulin may upregulate genes responsible for shifting TG
In humans the production of intestinal and hepatic lipoproteins is from VLDL formation to cytosolic lipid formation [26]. Another site for
subject to both substrate supply and hormonal regulation. Regulation insulin action is the MTP. Insulin inhibits MTP gene transcription
of TG production by substrate availability and summarized in Figs. 5 through the mitogen activated protein kinase (MAPKerk) cascade but
and 6 [22,23]. not through the PI3K pathway [27], other studies suggest that insulin
regulates hepatic MTP expression through the forkhead box 01 (Fox01)
1.2.2. Insulin transcription factor [28]. Besides regulating MTP gene transcription,
Lipoprotein secretion is a complex process in which substrate avail- both MAPKerk and insulin upregulate LDL-receptor gene transcription.
ability and insulin play key roles. Studies in cells, small animals and Evidence suggests that sterol responsive element binding proteins
humans have demonstrated a key role for insulin in hepatic VLDL secre- (SREBPs) are the downstream effectors of MAPKerk to transactivate
tion. Insulin acts as an anabolic hormone stimulating energy storage in the LDL-receptor gene [29,30].
liver and adipose tissue. Insulin signaling acutely suppresses both lipol- Insulin regulates metabolic activity and gene transcription by modu-
ysis and VLDL secretion. Although insulin acutely promotes hepatic lipo- lating the activity of several intracellular signaling pathways. Evidence
genesis it inhibits VLDL secretion. Insulin may indirectly regulate VLDL suggests that insulin regulates lipoprotein metabolism through acute
production by reducing plasma FA as a consequence of inhibiting and adaptive methods. Activation of the MAPKerk in the liver has a ben-
adipose tissue lipolysis. In healthy individuals acute inhibition of VLDL eficial effect on plasma lipid profiles as this would induce LDL-receptor
secretion is only partly due the suppression of FA and further FA inde- and suppress hepatic MTP expression. In addition there is potential for
pendent processes may play a part [24]. Decreasing endogeneous cross talk between the signaling pathways which may act as fine tuning
VLDL production during postprandial conditions is important for of the lipid profile in response to several other stimuli [31].
146 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

Fig. 2. Biosynthesis of chylomicrons. In the intestinal cell newly synthesized apoB48 follows one of two itineraries. In the presence of phospholipids, cholesterol and triacylglycerol, apoB48
is chaperoned by microsomal transfer protein (MTP), to form newly synthesized prechylomicron particles. In the absence of lipids it is rapidly degraded. The prechylomicron particle buds
from the endoplasmic reticulum and is transported to the Golgi as the prechylomicron transport vesicle (PCTV). Within the Golgi complex the prechylomicron forms the mature chylo-
micron particles. The mature chylomicron exits the Golgi complex in large transport vesicles to fuse with the basolateral membrane and is secreted. The mature chylomicrons contain
apoAI, apoAIV and apoB48. TG = triglycerides, PL = phospholipid, Chol: cholesterol.

Acute elevations of plasma FA stimulates not only hepatic but also 1.2.3. Glucagon
intestinally derived apoB48 containing lipoprotein particles in fed Glucagon is a pancreatic hormone that opposes the action of insulin,
humans [32]. Intestinal apoB48 containing lipoprotein production is and stimulates gluconeogenesis and glycogenolysis to promote glucose
acutely suppressed by insulin, and this may involve insulin's direct ef- liberation in hepatocytes. Glucagon actions are transduced through the
fects as well as insulin-mediated suppression of circulating FA [33]. glucagon receptor (Gcgr). In mice, it has been suggested that glucagon
This suggests that factors involved in the regulation of hepatic VLDL action is responsible for multiple pathways regulating lipid homeostasis
may regulate chylomicron production. [34]. The role of glucagon in regulating hepatic and intestinal
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 147

Fig. 3. Biosynthesis of VLDL. Biosynthesis of VLDL starts when apoB100 is translocated to the lumen of the ER and interacts with MTP. In the absence of lipids and microsomal transfer
protein (MTP), apoB100 is targeted for proteosomal degradation. MTP driven lipidation gives rise to pre-VLDL particles. Interaction with molecular chaperones enhance/modify
apoB100 degradation. Pre-VLDL is transported to the Golgi apparatus in a COPII dependent fashion. Pre-VLDL is converted to VLDL in the Golgi prior to secretion. Pre-VLDL which is
not converted to mature lipoprotein are sorted out for autophagy through post-ER presecretory proteolysis (PERPP). TG = triglycerides, PL = phospholipid, Chol: cholesterol.

lipoprotein metabolism is not clear. Though recent studies suggest particle production and clearance with no net effect on concentra-
that glucagon acutely regulates hepatic but not intestinal lipoprotein tion. The mechanism of its effect on apoB100 containing particles is
particle metabolism in humans, both by decreasing lipoprotein unknown [35].
148 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

Fig. 4. Biosynthesis of HDL. Biosynthesis of HDL begins with the interaction of lipid poor apoAI with ABCA1. Lipidated apoAI is the major acceptor for SR-BI and ABCG1 pathways. CETP
promotes the transfer of cholesteryl esters from HDL to VLDL and LDL. PLTP transfers phospholipid between VLDL and HDL and remodels HDL into larger and smaller particles with
the dissociation of lipid poor or lipid free apoAI. Solid arrows indicate cholesterol/cholesteryl ester transfer. chylomicrons (CM); very low density lipoproteins (VLDL); high density lipo-
protein (HDL); apoprotein (apo); lipoprotein lipase(LPL), Hepatic Lipase (HL), ATP-binding cassette (ABC); ABC transporters (ABCA1 and ABCG1), lecithin cholesterol acyl transferase
(LCAT); acyl-CoA:cholesterol acyltransferase (ACAT), cholesteryl ester transfer protein (CETP); Phospholipid transfer protein (PLTP); scavenger receptor BI (SR-BI).

Triglyceride metabolism at the whole body level is summarized in PPARs regulate several genes involved in metabolic processes. PPARα
Fig. 6. is expressed mainly in the liver, digestive tract and kidney, PPARβ is
ubiquitously expressed, PPARγ is mainly expressed in the adipose tis-
1.3. Nuclear hormone receptors and lipid metabolism sue. PPARs bind to specific DNA response elements, the PPAR response
elements as heterodimers with RXR. The fact that fatty acids are activa-
Dietary fats regulate gene expression in a hormone-independent tors of PPARs suggests that lipoproteins serve as ligand carriers for
manner. Since the description of dietary fat as a regulator of gene PPARs. Studies suggest that PPARγ regulates adipose tissue triglyceride
expression, many transcription factors were identified as prospective lipase (ATGL) in adipocytes [37]. In the liver PPARα directly regulates
targets for fatty acid regulation including peroxisome proliferator acti- genes involved in fatty acid uptake and oxidation [38]. Rakhshanderoo
vated receptors (PPARα, β, γ1 and γ2), SREBP-1c, hepatic nuclear fac- et al. [39] report that several genes involved in lipid metabolism were
tors (HNF-4α and γ), retinoid X receptor (RXRα), LXRα and others PPARα targets as were several putative novel genes which included
[36]. Fatty acids bind directly to PPARs, LXR and HNF-4α leading to CREBL3 (cAMP responsive element binding protein 3-like 3). CREBL3
changes in the trans activating activity of these transcription factors. encodes the liver expressed mammalian transcription factor CREBH
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 149

Table 1
Apolipoproteins.

Apolipoprotein Molecular weight Lipoproteins Metabolic functions Synthesis Reference

ApoAI 28,016 HDL, chylomicrons Structural component of HDL, LCAT activator Liver, intestine [3]
ApoAII 17,414 HDL, chylomicrons Liver [7]
ApoAIV 46,465 HDL, chylomicrons Involved in chylomicron assembly and secretion Intestine in humans [8,9]
ApoAV HDL, VLDL, chylomicrons Effects on plasma triglyceride concentrations are Predominantly in the liver [10,11]
complex and variable.
Activator of intravascular hydrolysis by LPL.
Modulates hepatic triglyceride metabolism
ApoB48 264,000 chylomicrons Necessary for assembly and secretion of Intestine [3]
chylomicrons from the small intestine
ApoB100 540,000 VLDL, IDL, LDL Necessary for assembly and secretion of VLDL from Liver [3]
liver. Structural protein of VLDL, IDL and LDL.
Ligand for LDL receptor
ApoCI 6630 Chylomicrons, VLDL, IDL, HDL ApoCI inhibits lipoprotein binding to its receptors. Liver [12,13]
Potent inhibitor of cholesteryl ester transfer
protein.
ApoCII 8900 Chylomicrons, VLDL, IDL, HDL Activator of lipoprotein lipase [19]
ApoCIII 8800 Chylomicrons, VLDL, IDL, HDL Inhibits lipoprotein lipase; increases VLDL Synthesized in the liver and to [12,14,15,16]
secretion. a lesser extent in the intestine
ApoCIII can also stimulate several processes
involved in atherogenesis and vascular
inflammation.
Interferes with remnant lipoprotein clearance.
ApoE 34,145 Chylomicrons, VLDL, IDL, HDL LDL receptor ligand for LDL and chylomicron Predominantly in the liver [17,18]
remnants. Ligand for LRP. Role in reverse
cholesterol transport.
Apo(a) 250,000–800,000 Lp(a) Unknown Liver [20]
ApoM 26,000 Predominantly HDL and also Synthesis of HDL Liver and kidney [21]
found in CM, VLDL, LDL

(cAMP-responsive element binding protein, hepatocyte specific). and degradation of the LDLR [41]. Hepatic lipogenic genes known to
Proteolysis results in CREBH activation and nuclear translocation, be under direct LXR transcriptional control include fatty acid syn-
where it increases expression of the hepatic gluconeogenetic enzyme, thase, steoryl CoA desaturase and SREBP1c. LXR activation mediated
phosphoenol pyruvate carboxykinase [40]. changes in insulin receptor substrate phosphorylation as well as in-
LXRs are activated in response to cellular cholesterol levels, creases in protein tyrosine phosphatase (PTP)1B. Tyrosine phospha-
allowing them to act as cholesterol sensors. The two isoforms of tases including PTP1B can regulate the biological effects of insulin.
LXR, α and β are highly homologous and form obligate heterodimers LXR activation can dysregulate hepatic insulin signaling and lead to
with the RXR. Once activated LXRs induce the expression of an array an increase in VLDL particles. [42,43]. Tumor suppressor phosphatase
of genes involved in cholesterol absorption, efflux, transport and ex- with tensin homology (PTEN) may further regulate insulin sensitivity
cretion. They also induce the expression of inducible degrader of as in its absence the PI3-kinase pathway can become constitutively
LDLR (IDOL), an E3 ubiquitin ligase that catalyzes the ubiquitination active [44].

Fig. 5. Hepatic fatty acid trafficking. In the postabsorptive state, FAs formed from lipolysis in the adipose tissue enter the liver and mix with fatty acids from the cytosolic TG pool and DNL.
FAs are then diverted to the mitochondrial oxidative pathway to form acetyl CoA. FAs are also esterified and enter the cytosolic TG storage pool. TG are assembled from FAs derived from
plasma NEFAs, chylomicron remnants or DNL and are either secreted as VLDL or stored in the cytosol to be used at a later stage. The postprandial increase in insulin shifts the cellular me-
tabolism of FAs away from oxidative pathways towards esterification, and increased DNL. DNL—denovo lipogenesis; ER: endoplasmic reticulum.
150 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

Fig. 6. Metabolism of TG. In the postprandial state CM from the enterocyte enters into systemic circulation. CM are hydrolyzed by LPL and the FA is taken up by muscle (for oxidation),
adipose tissue (for storage) and liver. There is insulin dependent suppression of FA release from the adipose tissue, and VLDL release from the liver. In the postprandial state hepatic
DNL is increased. In the postabsorptive phase there is net release of FAs from the adipose tissue, that is taken up by the liver and muscle. In the liver FA is partitioned for oxidation or
re-esterification and storage or VLDL secretion. CM: chylomicron; CMR: chylomicron remnant; DNL: de novo lipogenesis; LPL: lipoprotein lipase. : VLDL; CM.

In mammals the cholesterol content of the cell membranes is 2. Inherited dyslipidemias


controlled by a family of membrane bound transcription factors des-
ignated as SREBPs. Fatty acids do not bind to SREBP proteins but in- 2.1. Increase in concentrations of lipid or lipoprotein
duce changes in the nuclear abundance of the transcription factor
[45,46]. 2.1.1. Familial hypercholesterolemia
In vivo studies suggest that multiple PPARα and SREBP1c target
genes are regulated by fatty acids. These mechanisms control hepatic 2.1.1.1. Clinical presentation. Elevated plasma levels of LDL cholesterol
lipid composition and the lipids available for VLDL synthesis [47]. have been shown to be a risk factor for the development of atheroscle-
The synthesis of cholesterol and fatty acids begins with their precur- rosis and associated ischemic heart disease (IHD). In men, a rise in total
sor acetyl-CoA. Acetyl-CoA is converted to cholesterol in a pathway cholesterol from 5.2 to 6.2 mmol/L is associated with a threefold in-
involving at least 23 enzymes. Acetyl-CoA forms fatty acids in a path- creased risk of death from IHD [50]. Raised serum cholesterol and LDL
way involving up to 12 enzymatic steps Many of these pathways are cholesterol are characteristic of familial hypercholesterolemia (FH,
regulated by the SREBP family of transcription factors [48]. The three OMIM 143890). In the pre-genomic era the commonly used system
SREBP isoforms are designated SREBP1a, SREBP1c and SREBP2. The for classification of dyslipidemias was based on the Frederickson or
relative activities of three isoforms differ. SREBP1a is a potent activa- World Health Organization International Classification of Diseases of li-
tor of all SREBP-responsive genes, owing to its long transactivation poprotein phenotypes [51]. FH presents as hyperlipoproteinemia (HLP)
domain encoded by the first domain. In SREBP1c this exon encodes type 2A, elevated LDL cholesterol and normal TG levels. Most HLP type
a shorter transactivation domain that is less potent than SREBP1a. 2A patients have no defined monogenic defect, only about ~ 10% are
Although they differ in potency, SREBP1a and 1c both activate monogenic. Causal alleles for monogenic disorders are highly penetrant
genes involved in the synthesis of fatty acids and their incorporation and often lead to severe phenotypes. Alleles that cause complex poly-
into triglycerides and phospholipids. SREBP2 encoded by a different genic traits have a more subtle effect on disease. It is likely that the poly-
gene preferentially activates genes for cholesterol synthesis and genic form involves a multiplicity of loci with a range of effects. The
LDLR [49]. The challenge in the future will be to integrate these monogenic syndromes are generally distinguished from the polygenic
different pathways for a rational model for further understanding of forms by the severity of LDL cholesterol elevation, by the inheritance
lipid metabolism. pattern and the severity of IHD risk [52].
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 151

Raised cholesterol levels initiate changes that lead to the develop- around the world, the proportion of variants distributed equally
ment of atherosclerotic lesions. FH leads often lead to the accumulation among the exons. Pathogenic predictions and experimental/family
of cholesterol in the skin, xanthomas in the tendons, xanthelasmata studies were used to study the consequences of these changes, some
around the eyes and presenile corneal arcus around the cornea [53].A of which are dominant and fully penetrant causes of hypercholesterol-
number of molecular defects have been shown to underlie extreme emia while others seem to have little or no effect on the protein [70].
levels of LDL cholesterol. Defects in the LDLR gene is the most frequent Ideally, the putative mutant LDLR should be expressed in heterologous
cause [54,55]. Defects in other genes, the gene for apoB, proprotein cell in vitro and its activity compared with normal activity. Though the
convertase subtilisin kexin type 9 (PCSK9), and a protein that causes in- absence of the variant from a large number of normolipidemic individ-
ternalization of the LDLR protein can cause a similar phenotype. Not all uals, its cosegregation with raised cholesterol in a patient's family and
clinically diagnosed patients have a known defect, suggesting that other the nature of the amino acid variation may provide sufficient evidence
genes remain to be discovered [56]. In a majority of cases monogenic FH for its pathogenicity [71]. Mutations can either affect LDLR activity,
is inherited as an autosomal dominant disorder. The gene causing the LDLR transcription [66] or prevent the internalization of the receptor
rare autosomal recessive hypercholesterolemia (ARH), LDLRAP1 which through the clathrin coated pit machinery [72].
encodes an adapter protein named LDLRAP1 that is required for the in-
ternalization of the LDLR has recently been characterized [57]. The clin- 2.1.1.3. Polygenic hypercholesterolemia. The Global Lipid Genetic Consor-
ical phenotype from these mutations is variable, the apoB mutation is tium meta-analysis of genome wide association studies (GWAS, up-
less severe [57,58]. The homozygous forms are more severely affected dated in the website www.ebi.ac.uk/gwas) identified several loci
than the heterozygous forms [59]. Carriers of mutations assumed to re- where common variants affect LDL cholesterol concentration and the
sult in complete obliteration of LDLR activity had a more severe pheno- results of another study showed that a proportion of individuals carry-
type with higher LDL cholesterol and higher prevalence of tendon ing several LDL cholesterol raising single nucleotide polymorphisms
xanthomatosis and coronary heart disease than those with residual (SNPs) have LDL cholesterol concentrations that exceed the diagnostic
LDLR activity [60]. Gradient of severity of phenotype presentation has LDL cholesterol threshold for the diagnosis of FH. It has been suggested
been stated as mutations in LDLR = PCSK9 N APOB [61]. The clinical phe- that in a number of patients with FH without a known mutation, raised
notype of ARH is milder than receptor negative LDLR and resembles that LDL cholesterol may have a polygenic cause [73]. Studies have identified
observed in the receptor defective LDLR patients [62], and is more re- loci with evidence for association between common variants (those of
sponsive to lipid lowering treatment [63]. ApoB, ApoE, ApoCI, ApoCIV, ApoCII, PCSK9, ABCA1, ApoAI-ApoCIII-ApoAIV-
There is no single internationally accepted set of criteria for the clin- ApoAV, CETP, LIPC (hepatic lipase), LPL and ANGTPL4) and lipid or lipo-
ical diagnosis of FH. The most commonly used are the US (Make Early protein concentrations [74,75]. In a further study a combination of
Diagnosis to Prevent Early Death) MEDPED, the UK (The Simon Broome rare and common variants in the SORT1 (sortilin), APOB, LDLR, APOE
register) and the Dutch Lipid Clinic criteria. Criteria developed by the and PCSK9 genes were found to contribute to variation in LDL cholester-
groups in the UK and The Netherlands include family and personal his- ol levels [76]. In patients with FH with a known causative mutation, an
tory, physical signs and genetic studies in addition to cholesterol levels, additional polygenic contribution can explain a highly increased LDL
and are classified as definite, probable and possible diagnosis of FH, cholesterol concentration. A polygenic component can explain the
while the US criteria only uses lipid levels [64,65]. The estimated preva- large overlap in LDL cholesterol concentration between mutation carrier
lence worldwide is 1/500 for the heterozygous form and 1 in 1 million and non-carrier relatives [77].
for the homozygous form [66,67]. Recently, the prevalence of the molec-
ular defined homozygous variant was estimated at 1: 300,000 in The 2.1.1.4. Apolipoprotein B. Defects in other genes can cause a similar phe-
Netherlands with a more variable clinical phenotype [68]. The frequen- notype to the LDLR defect. Familial defective apoB100 (FDB, OMIM
cy of the APOB mutation varies from 2–10% of patients with clinical 144010), characterized by a mutation in the APOB gene, the gene coding
criteria for monogenic hypercholesterolemia compared to 52–76% of for apoB the ligand for the receptor on the surface of LDL is probably the
LDLR mutations and 2% PCSK9 and 2% LDLRAP1 mutations [69]. second most frequent cause of the FH phenotype. The 43 kb APOB gene
is located on the short arm of chromosome 2. It consists of 29 exons of
2.1.1.2. Low density lipoprotein receptor. The classical cause of FH is mu- which exon 26, with 7572 bp, is the largest and encodes more than
tations in the LDLR (OMIM 606945) gene. Pathogenic changes in LDLR one half of the full length protein. ApoB100 appears to be arranged as
result in impaired uptake and processing of LDL particles which leads a series of amphipathic α-helices and β-sheet domains (NH2-βα1-β1-
to elevated serum cholesterol levels. The LDLR gene which is located at α2-β2-α3-COOH). ApoB48 comprises the first two domains. It is thought
19p13.2 is composed of 18 exons spanning 45 kb, the transcript is that the βα1 domain binds MTP and that the carboxy terminal binds
5.3 kb long and encodes a peptide of 860 residues. Functional domains LDLR. The mature protein contains 25 cysteine residues of which 16
of the peptide correspond with the exons as follows: signal sequence- are involved in intramolecular linkage [78,79]. Studies of the three
exon 1, ligand binding domain-exons 2–6, epidermal growth factor dimensional structure of apoB100 by immunoelectron microscopy sug-
(EGF) precursor like domain-exons 7–14, O-linked carbohydrate gests that apoB100 enwraps LDL like a belt, completing the encircle-
domain-exon 15, transmembrane domain-exon 16 and 41 bp of exon ment by about amino acid 4050 and that the carboxy terminal forms a
17 and cytoplasmic domain-remainder of exon 17 and exon 18. LDLR bow that crosses backwards over the chain between residues 3000–
is expressed ubiquitously under the control of sterol regulated negative 3500, the suggested binding site for the LDLR [80]. ApoB is subject to
feedback, mediated by three 16 bp imperfect repeats (sterol regulatory both transcriptional and posttranscriptional control. APOB transcription
element) and a TATA like sequence in the promoter. [66]. As of 2012, the is regulated by a promoter in combination with either the liver-specific
LDLR variant database contained over 1288 LDLR gene events. The regulatory region or the intestinal enhancer. Thus two separate regula-
variants include 55% exonic substitutions, 22% exonic small rearrange- tory elements operate to control the expression of apoB in liver and in-
ments (b100 bp), 11% large rearrangements (N100 bp), 2% promoter testine [81]. ApoB mRNA levels remain relatively stable despite the
variants, 10% intronic variants and 1 variant in the 3ˈ untranslated se- presence of a number of factors known to influence apoB secretion,
quence. The LDLR variant database is recorded in the University College suggesting post-translational regulation of apoB. [82]. The ability of
London database and has been transferred to the Leiden Open Source apoB100 to interact with the LDLR depends on both sequence and con-
Variation Database (LOVD) platform. (https://grenada.lumc.nl/LOVD2/ formation [79]. In patients with FDB, LDL had a lower LDLR receptor
UCL-Heart/home.php?select_db=LDLR) and a further database is main- binding capacity than normal LDL [83].
tained by INSERM (http://www.umd.be/LDLR/). FH associated variants In contrast to the large heterogeneity of the LDLR locus only a few
continue to be reported along the length of the LDLR gene by groups mutations near the vicinity of codon 3527 of APOB were known to
152 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

prevent LDL from binding to LDLR. A predominant mutation CCG to CAG directs the trafficking of LDLR to lysosomes for degradation [99]. It has
mutation in the amino acid 3500 (R3500Q or now termed R3527Q) is on been suggested that amyloid precursor protein-like protein 2 is in-
a chromosome with a rare haplotype at the apoB locus suggesting that volved in mediating postendocytic delivery of PCSK9 to lysosomes
most probands descended from a common ancestor who lived in [100].
Europe about 7000 years ago [84]. However, different haplotypes have In 2003 gain of function mutations in PCSK9 (OMIM 603776) were
been reported suggesting an independent R3500Q mutation [85]. The shown to cosegregate in two FH families in which mutations to LDLR
frequency of the R3500Q mutation has been estimated from a four cen- and APOB had been excluded. The mutation was mapped to the 1p34-
ter study to be 1 in 543 [86]. Three further mutations R3500W, R3531C, 32 region [101]. An online database of PCSK9 mutation is now available
R3480W located within a stretch of 51 amino acids and result in the loss PCSK9 @ www.ucl.ac.uk/ldlr/LOVDv.1.1.0/. The clinical features of FH
of an arginine have been linked to defective LDLR binding. One hypoth- linked to PCSK9 mutations are similar to LDLR mutations, though excep-
esis is that arginine 3500 interacts with tryptophan 4369 and this inter- tions exist. PCSK9 patients with the D374Y variant had a more severe
action is essential for the correct conformation of the carboxy terminal phenotype than the FH patients with known mutations in LDLR, even
tail of apoB100. Disruption of this interaction results in a conformational those selected as having null mutations. The D374Y mutations showed
change that disrupts receptor binding. One model suggests that trypto- higher baseline and post statin treatment cholesterol levels [102]. Bio-
phan 4369 interacts with arginines in addition to 3500 in apoB100 dur- chemical and cellular analysis revealed that the gain of function muta-
ing the conversion of VLDL to LDL and that this functions as a modulator tion D374Y resulted in a 10–25 fold higher affinity for LDLR [103]. The
element that inhibits VLDL from interacting with the LDLR [79]. Muta- reported co-crystal structure of PCSK9 with the EGF-A domain of LDLR
tions with less of a severe impairment of binding can cause less of an revealed that at acidic pH Asp374 in PCSK9 forms a hydrogen bond
increase in serum cholesterol [86]. A new mutation H3543Y with a with His306 within the EGF-A domain. The mutation D374Y would po-
prevalence of heterozygotes as 0.47% compared to 0.12% for the sition a hydroxyl group of tyrosine to form a more favorable hydrogen
R3500Q mutation has been reported in German patients with coronary bond [104]. A natural gain of function mutation L108R enhances the
artery disease [87]. The advances in DNA enrichment and next genera- natural interaction PCSK9 for LDLR [105]. It has been suggested
tion sequencing technology have made it possible to sequence the (i) that LDL from PCSK9 mutant patients have an impaired capacity to
whole ‘exome’ ie the protein coding portion of the genome. Whole se- compete for binding to normal LDLRs [102] and that (ii) the secretion
quencing of APOB in FDB families identified R1164T, Q4494 del, of apoB100 containing particles is increased from cells expressing the
R3059C, K3394N mutations that can attenuate cellular LDL uptake. pathogenic PCSK9 variant [106]. It has been suggested that PCSK9
These results emphasize the need to study the whole APOB to identify plays a role in TG rich lipoprotein metabolism at least in certain patho-
new patients with FDB [88,89]. In some cases mutations are susceptibil- physiological conditions [107]. It is likely that in patients with mutations
ity mutations and enhance the effect of other mutations (eg an LDLR de- in the PCSK9 gene, as is the case with patients with FH attributable to
fect) and there is an absence of cosegregation of the APOB mutation with mutations in the LDLR, the nature of the molecular defect has an impact
hypercholesterolemia [90]. on the severity of hypercholesterolemia. Two further mutations S127R
The rare mutations in APOB confer a risk of IHD in the mutation car- and D129G segregated with the FH phenotype. Both mutations have re-
riers, their effect at the population level is minimal. Conversely, single duced autocatalytic activity and were not secreted from cells [108]. The
nucleotide polymorphisms (SNPs) may have a population effect that is S127R mutant has only a modest increase in affinity for LDLR. It has been
far from negligible despite a weak effect at the individual level. ApoB suggested that the S127R mutant may interfere with intracellular traf-
is a highly polymorphic gene and in individuals without rare mutations, ficking of LDLR to the cell surface [109].
polymorphisms may have a role in determining apoB and cholesterol The PCSK9 mutations causing FH are rare, however more frequent
level. Large epidemiological studies have indicated that the increased polymorphism variants have been associated with cholesterol levels
apoB serum levels are strong predictors for IHD [91]. Using isotope la- and IHD in population studies [110]. In patients the use of steroids or
beling techniques both mutations and polymorphisms have been mutations in the LDLR gene were additive with the PCSK9 mutation.
shown to affect LDL metabolism in vivo. However, APOB polymorphisms Mutations are associated with variable phenotypes in families where
studied so far have shown no association with IHD, suggesting other lipid disorders show variable degrees of severity suggesting that the
functionally important SNPs have not been identified [92]. Metabolism phenotype of hypercholesterolemia may result from a combination of
of apoB containing particles are dependent on gender, APOE genotype, environmental and genetic factors [111,112].
LDLR mutations, estrogen levels and LPL activity which may contribute
to inter individual variation in apoB levels. 2.1.1.6. Autosomal recessive hypercholesterolemia. Over the years a num-
ber of rare cases of autosomal recessive rather than autosomal domi-
2.1.1.5. Gain of function mutations in PCSK9. PCSK9 is mainly of hepatic nant form of FH have been described [113]. Interestingly no defect in
origin but is also well expressed in the intestine and kidneys. PCSK9 LDLR function could be detected, though subsequently a recessive null
contains a signal sequence (1–30 amino acids), a prodomain (amino mutation in a novel gene called the LDL receptor adaptor protein 1
acids 31–152), a catalytic domain (amino acids 153–425) and a (LDLRAP1 also known as ARH) was observed to cosegregate with hyper-
C-terminal domain (amino acids 426–694) [93]. In the ER, proPCSK9 is cholesterolemia in members of these rare families [114]. This was con-
autocatalytically cleaved at Q152 resulting in a secreted enzymatically firmed by restoring LDLR function in ARH patient lymphocytes with
inactive complex of PCSK9 with its inhibitory prosegment [94]. PCSK9 the expression of normal ARH protein [115]. To date, autosomal reces-
gene transcription has been shown to be under the control of SREBPs, sive hypercholesterolemia (ARH, OMIM 603813) is caused by mutation
HNF-1 [95] and PPARγ [96]. in the LDLRAP1 gene, though as other recessive forms of hypercholester-
LDL bound to LDLR is internalized in clathrin coated pits and subse- olemia emerge this term may need to change. The LDLRAP1 protein con-
quently undergoes lysosomal degradation. Following dissociation of tains a phosphotyrosine-binding domain that interacts with LDLR via
LDL, the LDLR is recycled back to the plasma membrane where it can the internalization sequence (NPVY) in the cytoplasmic tail of LDLR en-
bind more LDL. Internalization and shuttling of the receptor towards abling the receptor to engage with the clathrin coated pit machinery for
the plasma membrane is a continuous process. PCSK9 plays a role in endocytosis [116]. The conserved motifs of the C-terminal portion of
this process because it promotes the degradation of the LDLR [97]. LDLRAP1 bind to clathrins and to the β2-adaptin subunit of AP-2
PCSK9 binds to the extracellular EGF-A domain of LDLR in a post ER [117]. Most mutations of LDLRAP1 are null mutations though a mutation
compartment and prevents the receptor recycling to the surface [98]. lacking the phosphotyrosine binding domain has been reported
Models suggest that circulating PCSK9 in sera bind to LDLR. Acidification [116]. Whether the role of LDLRAP1 extends beyond sorting of the
of the pH in the endosomes increases PCSK9 affinity to the LDLR and LDLR into clathrin vesicles and its subsequent internalization remains
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 153

Fig. 7. Defects in LDLR cholesterol processing which cause FH. Dominant mutations in LDLR can affect ligand binding and receptor transcription and recycling [1]. Dominant mutations in
the receptor binding domain of apoB reduce its affinity for LDLR [2]. Dominant mutations in PCSK9 reduce the amount of LDLR at the cell surface [3]. Recessive mutations in LDLRAP1 pre-
vent internalization of the receptor [4].

to be solved. It is however involved in escorting LDLR on its complex campesterol, and stigmasterol) in addition to cholesterol accumulate in
recycling route. Additional mutations in LDLRAP1 may account for the blood and tissues of patients with this disorder. Cholesterol levels
more severe phenotype in families with LDLR mutation [118,119] are more variable (normal or mild or moderately elevated plasma choles-
(Fig. 7). terol levels) in sitosterolemia than in other genetic hyperlipidemias, but
can be extremely elevated in some patients. Clinical features may vary
2.1.1.7. Other candidate genes. A recessive mutation in the CYP7A1 gene from the presence of tendon xanthomas, atherosclerosis and its compli-
which encodes the enzyme cholesterol 7α-hydroxylase which catalyzes cations to a milder phenotype with very few specific symptoms or
the initial step in cholesterol metabolism and bile acid synthesis has signs. Disease is suspected in patients with elevated levels of plant sterols.
been identified in a single family. High levels of cholesterol were seen Plasma sterols levels in normal or heterozygotes should not be N1 mg/dL
in three homozygous subjects [120]. The enzyme 3hydroxy-3-methyl- as opposed to N20–30 mg/dL in patients with sitosterolemia, though
glutaryl-coenzyme A reductase (HMGCR) is a rate limiting enzyme in higher levels are found in some heterozygotes [126,127]. Several muta-
cholesterol biosynthesis. Burkhardt et al. [121] identified variants in tions have been described. A few of the mutations are over represented
HMGCR that contribute to LDL cholesterol levels, across populations. in certain ethnic groups eg ABCG5 R389H in Japanese and ABCG8
Both SREBP and the SREBP cleavage-activating protein (SCAP) which Y361X in Caucasians [128].
are involved in cholesterol and LDLR biosynthesis are candidate genes
for the development of hypercholesterolemia. Though variants in both 2.2. Primary hypertriglyceridemia
genes have been associated with higher LDL cholesterol concentrations
the evidence that they cause the phenotype is not strong [122]. 2.2.1. Triglycerides and risk of disease
Sterols in the intestinal lumen enter enterocytes through the The scientific statement from the American Heart Association states
Niemann–Pick C1 like-1 (NPC1L1) transporter and are secreted by that TG represent an important biomarker of IHD risk as it is associated
the heterodimeric ATP-binding cassette transporter ABCG5/ABCG8. with atherogenic remnant lipoprotein particles [129]. Because many
Humans express ABCG5/ABCG8 on the apical border of intestinal human cells can degrade TG, but none cholesterol, it seems more plau-
enterocytes (though ABCG5 has an additional cytoplasmic component) sible that it is the cholesterol content of remnants that is causal for ath-
and in hepatocytes framing the biliary canaliculi cells [123]. Recent pub- erosclerosis development. The specific role for TG as an independent
lications suggest that human NPC1L1 variants may modulate serum LDL risk factor for IHD has not been straightforward to make. Elevated TG
cholesterol levels. Rare NPC1L1 alleles with significantly higher LDL typically exists with a range of comorbidities such as obesity, metabolic
cholesterol levels were associated with increased IHD risk [124]. Con- syndrome and type 2 diabetes, which obscures a direct relationship be-
versely, variants classified as dysfunctional were identified from low tween IHD and TG concentration. Further, there is an inverse relation-
cholesterol absorbers. Dysfunction of the variants associated with ship between HDL and TG concentration [130]. Remnant cholesterol is
poor cholesterol uptake was due to impaired recycling, subcellular lo- the cholesterol content of TG rich lipoproteins composed of VLDL
calization, glycosylation or stabilization of NPC1L1 [125]. (following partial lipolysis by LPL) and IDL in the fasting state and addi-
The rare, recessive disorder, sitosterolemia, (OMIM 210250) is tionally, with chylomicron remnants in the non-fasting state. Several
caused by mutations in either of ABCG5 or ABCG8, that heterodimerize meta-analyses indicate a link between fasting/non-fasting TG values
to form a duplex that promotes sterol excretion. Mutations in either and IHD risk [131]. Both case control and prospective studies have
the ABCG5 or ABCG8 result in increased fractional absorption and de- found that elevated remnant cholesterol is associated with increased
creased biliary excretion of neutral sterols. Plant sterols (eg sitosterols, risk of IHD. Elevated nonfasting plasma triglyceride is a marker of
154 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

elevated nonfasting remnant cholesterol and is associated with in- mutations in LPL, APOCII, APOAV, LMF1, and GHIHBP1 genes. Familial
creased risk for IHD. It has been suggested that a 1 mmol/L increase in chylomicronemia is a rare autosomal disorder, diagnosed predomi-
nonfasting remnant cholesterol is associated with 2.8 fold causal risk nantly in childhood. The frequency of individuals with these elevat-
for IHD [132,133]. In a recent study, with a Mendelian randomization ed TG syndromes is less than 1:1,000,000 and they fit the definition
design, genetically reduced concentrations of nonfasting plasma TGs of HLP type 1 (OMIM 238600) with elevated chylomicronemia, skin
were associated with reduced all cause mortality [134]. and eye abnormalities and pancreatitis. HLP type 1 as a monogenic
TG rich lipoprotein remnants can drive atherogenesis either by disorder is frequently caused by mutations in LPL. Mutations in
direct mechanisms contributing to plaque formation or by indirect GPIHBP1, APOCII and APOAV are rare. In one study among patients
mechanisms associated with impaired vasodilation and enhanced in- with HLP type 1, 51% had mutations in LPL, 16% had a loss of function
flammatory reaction [135]. Mechanistically the explanation for a causal mutation in APOCII, GPIHBP1 or APOAV, 21% had common single nu-
effect of elevated levels of nonfasting remnants on IHD is that remnant cleotide polymorphisms (SNPs) in LPL or APOAV, and 12% of patients
cholesterol, like LDL, can enter and get trapped in the intima of the arte- had no genetic mutation [143]. This would suggest that the molecular
rial wall. In line with this, patients with mutations in the LPL gene caus- diagnosis of familial chylomicronemia requires analysis of LPL gene as
ing chylomicronemia with large CMs and VLDLs have an increased risk the first choice.
of pancreatitis but not an increased risk of IHD. However, some patients
with chylomicronemia due to LPL deficiency have developed significant 2.3.1. LPL deficiency
atherosclerosis [136]. It has been suggested that elevated LDL levels LPL deficiency is associated with an absence or decreased LPL activ-
cause atherosclerosis without a major inflammatory component, ity in post heparin plasma. Nearly a hundred clinically significant LPL
whereas the inflammatory component of atherosclerosis is driven by mutations have been described. Most LPL deficiency mutations are
nonfasting remnant cholesterol levels [137]. The increasing evidence missense mutations, although nonsense, deletions, insertion, splicing,
to support non-fasting TG levels to predict IHD risk identifies a need and some mutations localized to the peptide signal sequence have
for accurate and standardized methodology to measure lipoprotein been reported [144]. Some abolish the production of LPL transcripts or
remnants. However, as lipoprotein remnants are different in composi- lead to unstable transcripts; certain missense mutations prevent secre-
tion of lipids and apolipoproteins as a result of different stages of TG li- tion of LPL from the cells, others interfere with the stability of LPL
polysis and exchange of lipoproteins, assays are difficult to standardize. homodimers; and others yield LPL homodimers that are stable yet
An important question is whether all forms of hypertriglyceridemia are catalytically inactive [145]. The majority of LPL mutations are in the
equally atherogenic. In addition there is a lack of specific target goals for amino terminal catalytic domain and have been isolated to exon 5 and
each component that is applicable to large scale clinical trials. Data from 6 (Human Gene Mutation Database; HGMD® home page). Both homo-
randomized clinical trials on the clinical benefits of lowering TG are zygotes and compound heterozygotes have been demonstrated by
much less robust than the benefits of lowering LDL-cholesterol with classical cloning techniques. Two LPL missense mutations C418Y and
statins, thus hampering recommendations for the management of in- E421K abolish LPLs ability to bind to GPIHBP1 without interfering with
creased TG levels [138]. LPL catalytic activity or binding to heparin [146], suggesting that the
Plasma TG concentration is a complex trait which follows a right binding of LPL to GPIHBP1 requires the C-terminal portion of LPL. Two
skewed distribution in a population [139]. The classification of elevated further LPL mutations, G409R and E410V render LPL susceptible to
serum TG is generally based on fasting values and varies in different cleavage at residue 297 (a known furin cleavage site) [147].
guidelines. Fasting TG cut-off values for the diagnosis of hypertriglyc- Patients with mutations in both alleles of the LPL gene may more fre-
eridemia have varied form 1.7–2.3 mmol/L (to convert to mg/dL quently be associated with partial LPL activity and for these patients the
multiply by 88.6) and vary in different guidelines [140]. Together with lipoprotein abnormalities may be more varied. Normal pregnancy is as-
environmental influences, common and rare variants in multiple sociated with alterations in plasma lipoprotein and lipid levels including
genes may determine plasma TG concentration. A mild to severe eleva- a two to threefold increase in triglyceride level which reaches its highest
tion of TG (2–5 mmol/L) is a common feature in subjects with obesity, level during the third trimester. This physiological hypertriglyceridemia
the metabolic syndrome or type 2 diabetes. More severe elevations in is believed to be induced by an increased level of estrogen during preg-
TG (N5 and 10 mmol/L) is observed in individuals diagnosed with famil- nancy. A correlation between estrogen levels and hypertriglyceridemia
ial combined hyperlipidemia (FCH, HLP type 2B, elevated VLDL and LDL, has also been observed in premenopausal women using estrogen con-
OMIM 144250) and familial hypertriglyceridemia (FHTG, HLP type 4, el- taining contraceptives. In patient homozygous for the S172C mutation
evated VLDL, OMIM 144600). Severe hypertriglyceridemia (TG N10 with residual LPL activity pregnancy markedly elevated triglyceride
mmol/L) is a hallmark of rare genetic disorders due to mutations in levels [148]. The natural history of biochemical consequences of hetero-
LPL, APOCII, APOAV, GPIHBP1 or LMF1 [141]. Clinical features of increased zygosity for mutations in the LPL varies. Plasma TG in carriers of one
TG concentrations are eruptive, tuberous, or palmar crease xanthomas, defective LPL allele can be normal or elevated; age, obesity and lipid
hepatosplenomegaly, and when TG concentrations exceed 10 mmol/L, raising drugs can be contributory to changes in TG levels. Reduced
lipemia retinalis and pancreatitis [142]. HDL cholesterol concentrations have been observed in heterozygotes.
Rarely, severe hypertriglyceridemia is a simple monogenic disorder The pronounced hypertriglyceridemia observed by decreased plasma
with a Mendelian (typically recessive) inheritance, often presenting in LPL activity may regulate HDL cholesterol levels by (i) increasing the en-
childhood and a result of homozygosity or compound heterozygosity hanced exchange cholesteryl esters to VLDL (ii) altering the remodeling
for large effect genetic mutations. More commonly hypertriglyc- and maturation of HDL [149,150,151]. Four novel variants H273R,
eridemia develops later in life on a background of genetic variants com- R333H (most likely to result in a defect in LPL activity rather than syn-
bined with environmental (poor quality diet) and metabolic stressors thesis and secretion), p.A125Gfs*22 and p.T85Kfs*13 (predicted to
(diabetes, metabolic syndrome, certain medications). Recent advances form a truncated protein) were described by Martin-Campos et al.
in knowledge on the common hypertriglyceridemia have changed the [152]. The authors suggest that heterozygosity for LPL variants may cor-
classification of hypertriglyceridemia from the Frederickson classifica- respond to increased susceptibility to hypertriglyceridemia.
tion to one with a molecular genetic foundation [51]. Common mutations in the LPL gene: D9N, N291S, W86R, G188E,
P207L, D250N are associated with increased TG levels and may contrib-
2.3. Primary monogenic hypertriglyceridemia ute to susceptibility factors involved in age and sex dependent pheno-
typic expression of increased TG [153]. In certain regions of Quebec
A few patients with plasma TG more than the 95th percentile the prevalence of LPL homozygosity is 200 per million and has been
have rare monogenic disorders resulting from loss of function postulated as due to a founder effect in French Canadians [154]. The
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 155

G188E, P207L and D250N mutations account for most cases of LPL LMF1 may affect the metabolism of a broader range of proteins other
deficiency in French Canadians [155]. than lipases but this awaits the identification of other LMF1 client pro-
In contrast to other variants the S447X mutant that lacks the teins [166]. LMF1 deficiency is inherited as an autosomal recessive syn-
C-terminal serine and glycine is a gain of function mutation. Published drome, and reports of clinically significant mutations are rare.
data indicates enhanced post heparin lipolytic activity in the S447X
variant compared to controls though a large number of individuals is 2.3.4. GPIHBP1 deficiency
required to unmask this effect [156]. GPIHBP1 is an endothelial cell protein that is hypothesized to serve
as an endothelial cell platform for LPL mediated lipolysis [169]. Mature
2.3.2. ApoCII deficiency GPIHBP1 is a relatively short protein (184 amino acids), with four prin-
The importance of apoCII as an LPL activator is established by the ab- cipal domains: a signal peptide, an amino terminal acidic domain, a cys-
sence of LPL activity in patients with apoCII deficiency (OMIM 207750). teine rich Ly6 motif (a three fingered domain specified by 10 cysteines)
The diagnosis of apoCII deficiency is established by finding a virtual and a hydrophobic carboxy terminal motif that is replaced within the ER
absence of apoCII in plasma associated with reduced post heparin lipo- by a GPI anchor, following interaction of the C-terminal domain with
protein lipase activity that is corrected by the addition of apoCII contain- GPI transamidase. GPIHBP1 is cleaved at residue 153 and replaced by a
ing plasma. Since the initial report by Breckenridge et al. [157] many GPI anchor [170,171]. The negatively charged acidic domain with a
kindreds with apoCII have been described and the molecular defect large number of aspartates and glutamates binds LPL that contains pos-
characterized. Patients present with fasting chylomicronemia, hypertri- itively charged heparin binding domains. The structural Ly6 domain is
glyceridemia, eruptive xanthomas and chronic pancreatitis [158,159]. additionally required for LPL binding [172].
ApoCII is a constituent of CM, VLDL, LDL and HDL (Table 1). The During recent years homozygous GPIHBP1 mutations have been un-
human apoCII gene is located on chromosome 19q13.2; it is ~3.4 kb in covered in patients with severe chylomicronemia. The first GPIHBP1
length and consists of three introns and four exons [160]. It is a 79 mutation was identified in a patient with lifelong chylomicronemia.
amino acid apolipoprotein. The lipid binding domain of apoCII is located The patient was homozygous for a Q115P mutation in the Ly6 domain.
at the N-terminal and the C-terminal helix is responsible for interaction Cell culture studies revealed that the mutant GPIHBP1 reached the cell
with LPL. A variety of nonsense, frameshift and missense mutations surface normally but could not bind LPL or chylomicrons [172]. In
have been uncovered. They are located in the promoter region, the cod- some case reports mutations involve the conserved cysteine residue in
ing exons or splice junctions. Streicher et al. [161] describe a single nu- the Ly6 domain, C65Y, [173], C65S, C68G [174], C89F [175], C68Y
cleotide substitution which diminishes the binding of a transcription [176], C14F and C68R [177]. Although the C14F/C68R double homozy-
factor to a positive cis-acting element in the promoter region resulting gous mutations exhibited normal LPL binding activity, the levels of mu-
in the depletion of apoCII in the plasma. Other mutations produce tant protein were extremely reduced compared to those of wild type
amino acid exchanges, abnormal RNA splicing, frameshifts or premature proteins in vitro, possibly due to the instability of mutant GPIHBP1 pro-
termination codons [162,163,164]. teins [177]. Charriere et al. [175] identified a G175R mutant in the C-
ApoCII deficiency is inherited as an autosomal recessive trait. The terminal domain that targets the GPI anchorage crucial to membrane
heterozygous state (eg the apoCIIPadova kindred) is not always associat- transfer. Cell surface expression was significantly decreased for the
ed with abnormalities in lipoprotein pattern, and cannot always be dis- G175R variant. Rios et al. [178] identified a 17.5 kb deletion that encodes
tinguished by measuring fasting TG or plasma apoCII levels. This the whole of the GPIHBP1 gene and Berge at al. [179] identified dele-
conclusion cannot be extended to all families with apoCII deficiency, tions of exons 3 and 4 of the GPIHBP1 gene in an extended family ped-
since mutations are heterogeneous, the heterozygote state may have igree with chylomicronemia.
different levels of apoCII deficiency [165]. GPIHBP1 mutants frequently present with childhood onset
chylomicronemia and pancreatitis, though probands can present as
2.3.3. LMF1 deficiency adults [178]. Rios et al. [178] report lipemia retinalis and several
The recent study of a hypertriglyceridemic mouse strain carrying a eruptive xanthomas on the elbows at presentation, though reports
combined deficiency of LPL and hepatic lipase (HL) led to the identifica- of eruptive xanthomas are variable. LPL activity in adipose tissue is
tion of a new gene whose protein product is involved in the maturation normal and breast milk from affected subjects contained normal
of LPL and HL in the ER [26]. The human homolog of this gene, LMF1, LPL activity levels. Enzymatically active LPL entered the plasma, al-
maps on chromosome 16p13.3 and encodes a 567 amino acid protein. beit with delayed kinetics, after an injection of heparin, although
It is a multi-pass transmembrane protein with three soluble domains the increase in LPL activity is lower in affected patients [174]. Het-
protruding into the ER domain. Co-immunoprecipitation studies re- erozygotes of affected kindred had normal lipid levels, suggesting
vealed that LMF1 physically interacts with LPL, HL and endothelial li- that half normal amounts of a functional GPIHBP1 are sufficient for
pase (EL) but not with pancreatic lipase. The molecular mechanisms normal lipolytic processing of TG rich lipoproteins and that inheri-
underlying the lipase chaperone function of LMF1 is unknown. It has tance is autosomal recessive [180].
been suggested that LMF1 is involved in the assembly of inactive lipase
monomers to active dimers, either by being responsible for the as- 2.3.5. ApoAV deficiency
sembly of dimers or by stabilization of already formed dimers. This In humans ApoAV gene locus forms part of the gene cluster of APOAI/
modulation of lipase assembly may be a method of controlling lipase CIII/AIV/AV gene cluster on chromosome 11q23. The sole site of apoAV
expression. Further studies are required to explore factors regulating synthesis is the liver where it is translated as a 366 amino acid
LMF1 expression and function [166]. preprotein. Following intracellular cleavage of a 23 amino acid signal
A human subject homozygous for a nonsense mutation LMF1 peptide, mature apoAV appears in the plasma associated with VLDL,
(Y439X) was identified in a patient with markedly increased TG, tuber- HDL and to a lesser extent chylomicrons (Table 1). The role of apoAV
ous xanthomas and repeated episodes of pancreatitis. The severe in TG metabolism is evolving and the specific mechanism by which
hypertriglyceridemia was due to LPL deficiency, as determined by a ApoAV reduces plasma TG is still unclear. ApoAV plasma levels are
93% decrease in post heparin LPL activity in the plasma of the affected quite low (~150 μg/L or 4 nM) but a number of observations suggest it
individual, as well as a 50% decrease in HL activity [167]. A second ho- is a strong modulator of plasma triglycerides. It has been suggested
mozygous nonsense mutation W464X was identified in another patient that (i) apoAV binds to GPIHBP1 in vitro and this interaction facilitates
with severe hypertriglyceridemia with a decrease in post heparin LPL LPL mediated hydrolysis of CM (ii) apoAV may act as a ligand for hepa-
activity by 76% [168]. Lipodystrophy (loss of adiposity) was associated tocellular receptors (iii) apoAV plays a role in intracellular TG storage
with W464X suggesting diversity in phenotypes. It is conceivable that and/or mobilization [181,182].
156 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

In GWAS APOAV gene was found to be one of the genes influencing called polymorphisms [196]. A second strategy, to test for gene asso-
TG levels [183]. APOAV variants have been linked to polygenic hyperlip- ciations with traits, is the linkage approach in which families are
idemia types 2B, 4 and 5 [184]. Only a few alterations in the APOAV gene screened for loci linked to hypertriglyceridemia [197]. An emerging
have been found to cause the accumulation of CM, HLP type 1, or accu- suggestion is that multiple common and rare variants in multiple
mulation of CM and VLDL, HLP type 5 (OMIM 144650). Patients present genes with different functions contribute jointly to a given pheno-
with pancreatitis, eruptive xanthomas and heptoslenomegaly. Marcais type [198]. Average triglyceride concentrations in people could be a
et al. [185] identified Q139X mutation in the APOAV gene associated balance between harmful and protective alleles.
with chylomicronemia. The patient was a homozygote, severe hypertri- The recent definitive study by the Global Lipids Genetics consortium
glyceridemia in the heterozygotes was associated with a second allele of which analyzed N100,000 subjects identified 95 loci that showed
TG raising minor APOAV haplotype, suggesting that severe hypertriglyc- GWAS association with normal variation in lipid traits [199]. The
eridemia in heterozygotes may be an interplay between apoAV defi- analysis identified 32 loci which contributed to variations in plasma
ciency and other TG raising apoAV haplotypes. The full expression of TG concentration. The SNPs at each of the 32 loci contributed to 9.6%
the syndrome was characterized by HLP type 5 hyperlipidemia. APOAV of the total variation in plasma TG concentration, corresponding to
gene mutation found in chylomicronemic subjects are Q97X (with com- 25–35% of the total genetic contribution to TG variability. The major-
plete ApoAV deficiency) [186,187] and Q145X (predicted to generate a ity of genetic variation associated with plasma TG remains unattrib-
truncated apoAV) [188]. A mutation that resulted in an 8 amino acid loss uted [200].
in the signal sequence resulted in a mutated apoAV which was intracel- Wang et al. [201] sequenced DNA from non diabetic hypertriglyc-
lularly missorted to lipid droplets and not secreted. This mutation pre- eridemia patients and found a substantial proportion of heterozygous
sented as severe hypertriglyceridemia in a 11 month old boy, but only LPL, APOCII and APOAV mutations among these patients. GWAS showed
while he was being breastfed [189]. Mutations in the APOAV gene stud- that APOAV and LPL already known to be involved in monogenic
ied because of severe hypertriglyceridemia in three patients without hypertriglyceridemia were involved in plasma triglyceride concentra-
mutations in the LPL, APOCII, GPIHBP1 gene were: 12 bp deletion tion as were the newer players GCKR (glucokinase regulator protein)
p.S232_L235del; missense mutation p. L253P and the 4 bp deletion and APOB [202]. The insertion of two leucines in the leucine stretch of
resulting in a C-terminallly truncated protein, p. D332VfsX4. Structural the signal peptide of PCSK9 (p.L15_L16ins2L) is found in patients with
and functional analysis suggests that the mutant residues were impor- FCH as well as FH, suggesting that PCSK9 variants may contribute to
tant in apoAV function [190]. One patient who was compound heterozy- this multigenic disease and that there is a potential overlap between
gous for the mutations c.289CN T (p. Q97*) and c.823CN T (p.Q275*) in the two disorders. It is possible that this mutation is in linkage disequi-
the APOAV gene presented with marked hypertriglyceridemia and his- librium with other variants within the PCSK9 gene [203]. In the Japanese
tory of recurrent pancreatitis [191]. population TRIB1 and GALNT2 have been associated with lipid metabo-
Plasma TG concentrations were found to be associated with poly- lism. TRIB1 encodes tribble-1 which regulates activity of MAPK and
morphisms of the APOAV gene. Three different haplotypes of APOAV GALNT2 encodes N-acetylgalactosaminyltransferase 2 which is involved
have been described to result in significant differences in TG levels. in protein glycosylation [204].
Five polymorphisms were found to define three common haplotypes For the TG associated loci, genetic determinants of plasma lipopro-
(denoted APOAV*1, APOAV*2 and APOAV*3). These three haplotypes tein metabolism from the Global Lipids Genetics consortium study
represent 82%, 8% and 8% of the APOAV chromosomes examined. showed association with different Frederickson associated HLP pheno-
APOAV*2 is distinguished from the common haplotype APOAV*1 by types. This suggests that the common variants increase susceptibility
four nucleotide substitutions (− 1131T N C, c.-3A NG, IVS3 + 476G N T to hypertriglyceridemia and the HLP phenotypes (HLP type 2B (FCH),
and c1259TNC). APOAV*3 is distinguished from the common haplotype 3, 4 (FHTG) and 5). The authors reported associations of common vari-
by the substitution c56CN G (S19W). Both APOAV*2 and APOAV*3 are ants with both HDL cholesterol and LDL-cholesterol and suggest that
associated with higher TG levels [192]. pleiotropic variants were jointly associated with plasma TG, LDL choles-
terol and HDL cholesterol. Both rare genetic variants and lipid associated
2.3.6. LPL inhibitors common variants occur across the HLP phenotypes and may partly ex-
Brunzell et al. [193] described a mother and her son with HLP type 1 plain the phenotypic heterogeneity among hypertriglyceridemic pa-
with very low levels of postheparin plasma lipolytic activity, and circu- tients. It is hypothesized that the extremes of a complex genetic trait
lating inhibitor of lipoprotein lipase, who differed from subjects with li- such as patients with the Frederickson HLP phenotypes are found as a
poprotein lipase deficiency in that the enzyme was present in adipose cumulative contribution of several common alleles, though some pa-
tissue at much higher levels than those seen in normal subjects. They tients such as most of those with HLP type 1 and a percentage of those
also differed from subjects with deficiency of apoCII in that apoCII was with HLP type 2A and type 5 have a rare loss of function mutation
present in their plasma in normal or elevated amounts. The authors sug- with a large effect [205,206]. The population prevalence for HLP type
gest an autosomal dominant status for the carriers of the LPL inhibitor. 2B is estimated as 1:40; HLP type 4 as 1:20 and HLP type 5 as 1:600
Anti LPL antibodies have been described in rheumatic diseases, mainly [140].
systemic lupus erythematosus and systemic sclerosis [194]. Ashraf Further analysis suggests a mechanistic role in lipoprotein me-
et al. [195] describe a patient with Sjogren's syndrome, with no disease tabolism for at least some of the GWAS identified loci. For example,
causing mutations in GPIHBP1, APOAV, APOCII and LMF. LPL antibody GCKR encodes glucokinase regulatory protein, an allosteric regulator
was detected by immunoblotting. The patient responded to immu- that allows mobilization of glucokinase in response to increased
nosuppressive therapy for Sjogren's syndrome with resolution of cellular glucose concentrations. The P446L variant in the GCKR
hypertriglyceridemia, promotes glucose uptake and mobilization, de novo TG synthesis
while attenuating fatty acid oxidation [207]. Mutations in CREBL3
2.3.7. Polygenic causes of hypertriglyceridemia that produced nonfunctional CREBH protein resulted in hypertri-
In contrast to the rare familial syndromes, together with environ- glyceridemia in humans [208]. Upstream transcription factor
mental influences, common and rare variants in multiple genes may (USF1) targets fatty acid synthase, and can be anticipated to be asso-
collectively determine an individual's plasma TG concentration. GWAS ciated with hyperlipidemia [209]. It has been suggested that the indi-
test for associations between common genetic variations with SNPs vidual genes identified in hypertriglyceridemia can be incorporated
and either quantitative or discrete traits. Differences in gene structure into models of (i) adipose tissue dysfunction (ii) hepatic fat accumu-
that occur infrequently in populations ie less than 1% are called mu- lation and VLDL overproduction (ii) delayed clearance of apoB con-
tations whereas differences that occur more frequently ≥ 1% are taining particles [210].
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 157

2.3.8. Novel rare genetic diseases associated with hypertriglyceridemia site colocalised with the LDLR binding site. Molecular modeling suggests
In families with severe transient childhood hypertriglyceridemia that Arg136, Ser139, His140, Arg142, Lys143, Arg145, Lys146 and
and fatty liver followed by hepatic steatosis and sustained liver dys- Arg147 that line the groove of helix 4 make direct contact with either
function, a mutation in the GPD1 gene which encodes glycerol-3- the 2-O-sulfo groups of the iduronic acid monosaccharides or the N
phosphate dehydrogenase was identified. One possible mechanism and 6-O-sulfo groups of the glucosamine sulfate monosaccharides
for transient hypertriglyceridemia is possibly by limiting the conver- [219]. The NMR structure of apoE3 suggests that the C-terminal domain
sion of Glycerol-3-phosphate to dihydroxyacetone phosphate, and presents a large exposed hydrophobic surface that likely initiates bind-
increasing the amount of glycerol-3-phosphate available for TG syn- ing to lipids. ApoE3 displays salt bridges and hydrogen bonds causing
thesis [211,212]. shielding of the LDLR binding region from binding to receptors, by the
HL is an enzyme made primarily by hepatocytes. Evidence from clin- C-terminal [17]. Relative to apoE3 more apoE4 is found in VLDL than
ical and animal studies supports a role for HL in the metabolism of sev- HDL. The distribution of apoE2 was similar to apoE3, indicating that
eral lipoprotein classes. HL has been reported to augment the uptake of cysteine-arginine interchange at position 158 does not affect distribu-
HDL cholesterol by the liver through a reverse cholesterol transport pro- tion [220]. Lipid free apoE interacts with ABCA1, and the lipid
cess; be responsible for the hydrolysis of TG in LDL and participate in the translocase activity of ABCA1 generates HDL-apoE particles. The three
conversion of VLDL to LDL. Missense mutations in the encoding exons apoE isoforms showed similar kinetics in their abilities to induce
R186H, L334F, S267F, T383M, and A174T have been described and ABCA1 dependent cholesterol efflux [221]. Studies on apoE deficient
found to be responsible for HL deficient phenotypes. It is a rare genetic mice suggest that overexpression of apoE3 in apoE−/− livers stimu-
disorder and appears to be inherited as an autosomal recessive trait. lates VLDL secretion, indicating a role for apoE in hepatic TG secretion
HL deficient patients present with hypercholesterolemia, hypertriglyc- independent from its role in lipoprotein clearance [222]. The differential
eridemia though patients display variable phenotypes. Consistent find- effects of overexpression of human apoE3 on plasma cholesterol and TG
ings are an elevation in HDL2 cholesterol and marked TG enrichment of metabolism was investigated in transgenic rabbits. In apoE3 high
LDL and HDL particles, and some affected individuals have premature expressors, the elevated expression of apoE3 appeared to stimulate
cardiovascular disease [213,214]. hepatic VLDL production and inhibit VLDL lipolysis. The overexpression
Other monogenic conditions that do not primarily affect lipoprotein of apoE3 in transgenic rabbits altered the plasma lipid profile to resem-
metabolism can cause hypertriglyceridemia. Inherited lipodystrophies ble the combined hyperlipidemia phenotype. The VLDL from high
are rare autosomal dominant or recessive disorders characterized by se- expressors was enriched with apoE3 and decreased in apoCII, which
lective and variable loss of adipose tissue. Marked hypertriglyceridemia contributed to a decrease in lipolysis. The authors suggest that the dif-
is a common feature of these disorders and highlights the role of adipose ferential expression of apoE3 may play a role in modulating plasma cho-
tissue in lipid homeostasis. Eight different genetic loci, including lesterol and TG content [223].
1-acylglycerol-3-phosphate-O-acyltransferase 2, Berardinelli-Seip con- ApoE3 and apoE4 bind similarly to LDLR, but compared to apoE3,
genital lipodystrophy 2, caveolin 1, lamin A/C, PPARγ, v-AKT murine apoE4 reduces plasma cholesterol less in humans. It has been suggested
thymoma oncogene homolog 2, zinc metalloprotease and LMF1 have that the C112R substitution located in the helix region, destabilizes it,
been linked to different lipodystrophy syndromes. Mutations in these and changes the interaction between the N and C terminal domains,
genes may cause fat loss and dyslipidemia through multiple mecha- which changes the organization of segments which play a critical role
nisms, which remain to be fully elucidated; however, they may involve in the interaction of lipid surfaces. ApoE4 exhibits a better lipid binding
defects in development and differentiation of adipocytes, and prema- ability than apoE3, and coupled with the VLDL particle surface being
ture death and apoptosis of adipocytes. Hypertriglyceridemia is a conse- predominantly covered with lipid (in comparison to HDL) leads to
quence of increased VLDL synthesis from the liver, which is loaded by apoE4 binding better than apoE3 to VLDL. Studies in apoE-null mice
ectopic TG deposition, reduced clearance of triglyceride-rich lipopro- suggest that the expression of apoE4 gave rise to higher VLDL cholester-
teins or both [215,216]. ol and lower HDL cholesterol relative to apoE3. It has been suggested
that VLDL remnant clearance is decreased in apoE4 expressing mice
2.4. ApoE and this decreases the amount of VLDL components available for incor-
poration into HDL accounting for the proatherogenic profile (higher
2.4.1. ApoE and lipoprotein metabolism VLDL cholesterol/HDL cholesterol ratio) [224,225]. Meta-analysis
The apoE gene is located on chromosome 19q13.2 within a gene suggests a linear relationship between apoE genotypes (when or-
cluster with apoCI and apoCII. Human apoE has 299 amino acids. The dered as E2/E2, E2/E3, E2/E4, E3/E3, E3/E4, E4/E4) with LDL cholester-
gene contains 4 exons. The 18 amino acid signal sequence is removed ol and with coronary risk [226]. The apoE2 allele is associated with an
prior to secretion. ApoE contains two independently folded functional anti-atherogenic lipid pattern (lower LDL cholesterol and higher
domains: a 22 kDa N-terminal domain (residues 1–191) and a 10 kDa HDL cholesterol) in children [227]. However, those apoE2 homozy-
C-terminal domain (residues 222–299). A short hinge region connects gotes who develop HLP type 3, because of secondary risk factors
the two domains. The N-terminal domain exists in a lipid free state as are at increased risk for both coronary and peripheral artery athero-
a 4 helix (residues 24–42, 54–81, 87–122, 130–164) bundle of amphi- sclerosis. Although the cholesterol lowering effect of apoE2 has been
pathic α-helices and contains the LDL-receptor binding region (residues confirmed in several populations the mechanism is not completely
136–150 in helix 4). The C-terminal domain has a high affinity for lipids understood. Studies with transgenic mice suggest that apoE2 de-
and is responsible for lipoprotein binding. ApoE interacts with the creases LDL cholesterol by impairing LPL mediated hydrolysis of
LDLR, LRP and heparan sulfate proteoglycans (HSPG), and serves an VLDL, mostly by displacing apoCII. Effects of apoE2 on cholesterol
antiatherogenic function by mediating the clearance of VLDL and CM and TG in transgenic mice were dose dependent. Plasma TG levels
remnants. The liver is the major source of apoE. ApoE associates with (mostly in VLDL and IDL) increased as apoE2 levels increased. Simi-
most lipoproteins except for LDL. ApoE exists in three major isoforms, larly human apoE2 homozygotes who are hypolipidemic have low
apoE2, apoE3 and apoE4. ApoE3 is the most common isoform and has levels of LDL cholesterol and a slight accumulation of TG. One expla-
cysteine and arginine at residues 112 and 158, respectively. ApoE2 con- nation is that sufficient apoE2 is present to impair lipolysis of TG rich
tains cysteine and apoE4 contains arginine at both sites. Both apoE4 and lipoproteins but sufficient apoE2 is present to mediate remnant
apoE3 bind to the LDLR with high affinity whereas apoE2 exhibits poor clearance by the LDLR or HSPG/LRP [228]. Overt hypercholesterolemia
binding to the LDLR and is associated with HLP type 3. Association of and hypertriglyceridemia can be precipitated when other factors
apoE with lipid is required for its high affinity binding to the LDLR [18, which further stress the already compromised remnant catabolism of
217,218]. The N-terminal domain contains the primary HSPG binding apoE2 homozygotes are present.
158 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

Once internalized triglyceride rich lipoproteins (CM and VLDL) dis- mutation display HLP type 3. ApoE variants associated with a dominant
integrate in peripheral endosomes. Although core lipids and apoB are inheritance HLP type 3 are associated with amino acid substitution in
targeted to lysosomes, apoE is targeted to recycling endosomes or to the LDLR binding region (residues 136, 142, 145, 146, 147). Wardell
the Golgi secretory pathway. HDL or purified apoAI stimulates the re- et al. [244] describe a patient who was heterozygous for apoE2 and
lease of internalized apoE in hepatocytes and fibroblasts. It has been apoE2 with a further R136S mutation who presented with HLP type 3.
suggested that HDL derived apoAI is internalized and targeted to pre- A further heterozygous variant apoE3/E2(R136C) has been associated
existing apoE/cholesterol containing endosomes to promote apoE with HLP type 3. Two children in the same family were carriers of the
recycling and cholesterol efflux. ApoE4 is less efficient compared with rare mutation and were heterozygous for the apoE4/E2(R136C) and
apoE3 in promoting cholesterol efflux in fibroblasts and astrocytes. It apoE3/E2(R136C) variant and were hyperlipidemic [245]. ApoE3 has ar-
has been speculated that the inefficient processing of intracellular cho- ginine in position 136. Carriers who were heterozygous for the mutation
lesterol could contribute to the association of apoE4 with atherosclero- apoE3(R136H) displayed an increased concentration of VLDL cholester-
sis. Depending on the apoE isoform different aspects of lipid metabolism ol and TG. All carriers showed β-VLDL on electrophoresis though they
seem to contribute to the development of dyslipidemia [229]. Of inter- did not display clinical features [246]. Two siblings with apoE2/
est is that individuals with the apoE4 allele have an increased risk and E3(R136S) phenotype presented with HLP type 5. Rare mutations in
earlier onset of Alzheimer disease [230]. the apoE gene and the associated HLP type are summarized by Vialettes
et al. [247]. It is controversial if mutations in the Arg136 site have a dom-
2.4.2. Pathophysiology of dysbetalipoproteinemia inant mode of inheritance. The presence of the apoE2(R136S) variant
Familial dysbetalipoproteinemia or HLP type 3 is characterized by el- has been linked with incomplete dominance of HLP type 3 [248].
evated plasma triglycerides and cholesterol that predisposes affected Other studies suggest that ApoE2(R136S) mutation fully expresses
subjects to the development of premature atherosclerosis. The hyperlip- HLP type 3 when associated with a second allele coding for apoE2 and
idemia is caused by accumulation of CM remnants and IDL known col- only partially in association with a second allele coding for apoE4
lectively as β-VLDL, these abnormal apoE enriched lipoproteins are [249]. Two heterozygous carriers apoE3/E2(R136C) carriers were classi-
the hallmark of the disease. HLP type 3 associated with homozygosity fied as HLP type 4. The authors suggest that apoE2(R136C) causes
for apoE2 presents with recessive inheritance. In a Dutch population, dysbetalipoproteinemia in a recessive fashion [250]. Double mutants
apoE2 homozygosity occurred at a frequency of 0.6%, of whom 18% pre- have been reported in apoE. Homozygosity for apoE (E13K, R145C)
sented with HLP type 3 [231]. A study by Feussner et al. [232] shows that was associated with severe HLP type 3 in a 24 y old female [251].
the prevalence of HLP type 3 among apoE2 homozygous relatives of ApoE variants causing dominant transmission of HLP type 3 differ
German HLP type 3 patients is 30%. The condition generally does not from apoE2 in that they not only exhibit decreased binding to LDLRs
present until adulthood in men and in the post-menopausal years in but also to HSPG. The recessive nature HLP type 3 and the variable ex-
women. Most apoE2/E2 subjects are either normolipidemic or even pression of hyperlipidemia in patients with apoE2 may be explained
hypocholesterolemic, onset of dysbetalipoproteinemia requires other by the high capacity of remnants possessing apoE2 to bind to HSPG.
metabolic stressors. ApoE2 is essential but not usually sufficient to Impaired HSPG interaction may be only one factor responsible for the
cause HLP type 3. Overproduction of lipoproteins (eg diabetes), remnant accumulation in the patients with dominant transmission of
impaired lipolysis of lipoproteins and decreased remnant clearance HLP type 3. Increased defective apoE to apoE3 ratio on the remnant lipo-
(eg hypothyroidism) can precipitate dyslipoproteinemia in patients protein can overwhelm the normal binding activity of apoE3 and could
who are predisposed to HLP type 3 [233]. Metabolic conditions include result in impaired interaction with HSPG [252].
insulin resistance [231], diabetes mellitus, nephrotic syndrome, obesity A single mutation R142L in the heterozygous state apoE2(R142L)/E3
and hypothyroidism. Overproduction of VLDL may also be the result of presented with dysbetalipoproteinemia [253]. Some of the mutations
excess alcohol intake or pregnancy [234]. Certain drugs: anti-psychotic associated with dominant HLP type 3 are R142C [254], K146Q [255]
medications [235,236] and antiretroviral therapy [237] have been asso- and K146E [256]. The double mutation K146N/R147W presented at an
ciated with HLP type 3. early age, with a proband who developed eruptive xanthomas at the
Several groups have studied genetic factors that may contribute to age of 3 [257]. Transgenic mice studies suggested that the mutant was
the expression of HLP type 3. Data from family studies suggest that a dominant negative ligand which prevented receptor mediated rem-
one or more genes are additional factors predisposing to HLP type 3. nant clearance and inhibited the biogenesis of HDL [258]. Heterozygous
Hennemann et al. [238] found the frequency of rare alleles APOCIII mutation with a dominant form of mutation ApoE3/ApoE4(R142C)
3238 G NC and APOAV -1131T NC was significantly higher in the HLP were described in 1983. Studies with transgenic mice suggest that this
type 3 patients compared with normolipemic apoE2/E2 patients. In ad- mutation alters the function of LPL, LCAT, decreased lipoprotein clearance
dition they found that the LPL c.27 GNA mutation was associated with and inhibited the maturation of discoidal HDL into spherical HDL particles
more severe hyperlipidemia whereas the APOAV c.56 G N C mutation [259]. The apoE2(R145C) variant has been associated with an autosomal
was weak modifier. The molecular mechanisms underlying the associa- dominant mode of inheritance with incomplete penetrance [260]. A var-
tion with hypertriglyceridemia were still unclear. Other mutations lo- iant of apoE4 which included an inframe repeat of 21 nucleotides (coding
cated near the polymorphic site eg within the APOAI/CIII/AIV gene for 7 amino acids) in exon 4 commonly referred to as apoE-Leiden, is
cluster near the APOCIII 3238 G N C polymorphism, could be candidates associated with dominantly inherited dysbetalipoproteinemia with a
for the causative mutation. The authors conclude that polymorphisms in high degree of penetrance. Age, however, exerts a significant effect on
lipolysis genes are associated with the severity of HLP type 3 in apoE2/ the severity of the hyperlipidemia. ApoE-Leiden possesses a reduced
E2 homozygotes. Other studies suggest that the S19W variant of LDLR binding activity [261].
apoAV is a factor in determining HLP type 3 in patients with apoE2/E2 A rare form of HLP type 3 associated with apoE deficiency, with un-
genotype [239]. The N291S variant in the LPL gene predisposes to detectable plasma apoE was reported in 1981. The deficient apoE gene
more severe hypertriglyceridemia [240]. The N291S substitution was resulted from a point mutation which causes aberrant splicing of the
suggested as a predisposing factor contributing to the development hy- third intron of the apoE gene resulting in low levels of aberrant forms
perlipidemia when associated with apoE2 genotype [241]. Other studies of mRNA [262]. The genetic pattern of the kindred is consistent with
suggest that it is unlikely that the D9N, N291S, and S447X variants in the an autosomal recessive mode of inheritance in which only the homozy-
LPL gene play an important role in the development of HLP type 3 [242]. gotes develop tuboeruptive xanthomas. The proband's brother had
HLP type 3 has been associated with the S267F mutation in HL [243]. striking xanthomas in unusual locations: ears, palms, elbows and
Rare apoE variants have been identified that have been implicated in knees. The concentration of apoE in homozygotes was b0.5% of that ob-
a dominant mode of inheritance ie subjects heterozygous for the served in normal. Heterozygotes had a mean plasma apoE concentration
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 159

that was ~50% of normal [263]. ApoE deficiency resulted (i) impaired binding to the LDLR but only moderately reduced binding to heparin.
catabolism of VLDL/chylomicron and their remnants due to lack of Hyperlipidemia has not been consistently observed in LPG patients
direct removal (ii) an increased rate of catabolism of LDL apoB100, likely without nephrotic syndrome. Patients with LPG do not reveal the char-
due to upregulation of LDLR activity due to decreased hepatic cholester- acteristic clinical features of HLP type 3, palmar xanthomas or athero-
ol content caused by poor delivery of apoE deficient lipoproteins, sclerosis. ApoE concentrations are elevated in LPG. In LPG, lipoprotein
and the inability of apoE deficient VLDL/IDL to compete with LDL for deposition occurs in the lumen of the capillaries. In HLP type 3 lipids
the receptor (iii) reduced VLDL production (iv) delayed catabolism of may accumulate in mesangial foam cells [280].
Lp(a) [264]. Although evidence suggests a direct association between apoE mu-
Several other patients with apoE deficiency have been reported tations and LPG, the mechanism that leads to the deposition of the lipo-
[265,266]. A mutation that causes a truncated apoE with a high VLDL proteins in the glomerulus is unknown. A study of LPG linked apoE3
cholesterol/TG ratio in the homozygote was reported by Lohse et al. variants, R145P, R147P, R158P showed that the mutations induce signif-
[267]. Other rare mutants are a 10 bp deletion in exon 4 leading to a pre- icant thermodynamic, structural, and aggregation-related perturbations
mature stop codon in amino acid 229 [268], a deletion in codon 31 lead- in the molecule of apoE3 that cause unfolding of the N-terminal domain
ing to premature stop in codon 60 [269] and a G insertion in codon 95 or to a more molten-globule like structure. This may be the result of the
96 leading to a premature stop at codon 146 [270]. Heterozygosity for substitution of the helix breaker proline residue in a helical segment
the apoE deficient mutations may be clinically silent, clinical expression [281]. In another murine experimental study renal lesions specific to
may depend on the other apoE allele. The apoE null allele in combina- LPG were shown to develop in Fc receptor γ chain (FcRγ) knock out
tion with apoE2 resulted in HLP type 3, but when inherited with mice in which chronic graft-versus-host disease had been induced. In
apoE4 had no marked effect on plasma lipids [271]. studies with FcRγ/apoE knock out mice the human apoE3 injected
mice showed LPG like changes, suggesting that macrophage dysfunc-
2.4.3. Clinical presentation tion contributed to the development of LPG in the presence of apoE ab-
Untreated patients can present with cutaneous and tendinous normalities. It has been suggested that dysfunction of lipid clearance in
xanthoma which are not specific for dysbetalipoproteinemia, however, the kidney may be another mechanism for development of LPG [282].
palmar crease xanthomas are specific for the disease. Other clinical The report by Zhang et al. [283] of LPG in patients without apoE muta-
manifestations are atherosclerosis and when severe hypertriglyc- tions suggest that multiple factors other than apoE are involved in the
eridemia is present, pancreatitis [140]. pathogenesis of LPG.
Electrophoretic demonstration of β-VLDL is considered a gold stan- Hu et al. [284] studied 35 patients with ApoE3(R25C) mutation
dard for the diagnosis of this disorder, though no simple diagnostic within 31 unrelated families and 28 (apoE3 (R25C) mutation) asymp-
test is available. In the series by Marais et al. [272] the fasting TG ranged tomatic relatives. The existence of 28 asymptomatic relatives suggested
from 1.6–63.4 mmol/L with a median of 4.9 mmol/L, the fasting choles- an incomplete penetrance for the apoE3(R25C) mutation. An ApoE3 ho-
terol ranged from 4.8–34.3 mmol/L with a mean of 11.79 mmol/L, the mozygote, heterozygous for a 54 bp deletion from exon 4 causing a 18
mean molar ratio of total cholesterol/TG was 2.07 ± 1.03 (median amino acid deletion (Q156-G173) presented with LPG and systemic
1.96, range 0.39–7.63). The authors suggest that dyslipoproteinemia atherosclerotic complications. One of his daughters heterozygous for
should be considered in all patients with a mixed combined hyperlipid- the mutation did not have proteinuria or atherosclerosis diseases
emia total cholesterol to TG ratio around 2:1, but should also be consid- [285]. A few other mutations associated with LPG are apoE(R150P)
ered in severe hypertriglyceridemia. Cholesterol enriched VLDL is [286], apoE2(R150G) [287], apoE3(R114C) [288] and apoE(R158P)
diagnostic of dysbetalipoproteinemia but sample preparation requires [289]. More than a 100 cases have been reported since 1989 [290].
ultracentrifugation which is not available in routine diagnostic laborato- Recurrence of LPG in a transplanted kidney suggests that LPG is
ries. Each VLDL, LDL and IDL lipoprotein contains one apoB100 molecule, caused by extra renal factors [291]. Results from case studies suggest
larger lipoproteins contain proportionately more lipid and less apoB than that fibrates may improve renal pathology and may result in complete
smaller molecules. A VLDL-cholesterol to serum TG molar ratio ≥ 0.69 remission of the disease [292].
(≥ 0.30 when using mg/dL) reflects the presence of cholesterol rich
β-VLDL and can be diagnostic of HLP type 3 [273]. Other data suggest a 2.5. Decrease in concentration of lipid or lipoprotein
total cholesterol/apoB ratio N 6.2 and TG/apoB ratio b 10.0 or conversely
the apoB/total cholesterol ratio can be used to diagnose HLP type 3 2.5.1. Familial hypobetalipoproteinemia
[274,275]. Familial hypobetalipoproteinemia (FHL) is characterized by abnor-
mally low levels of apoB containing lipoproteins in the proband, no un-
2.4.4. Lipoprotein glomerulopathy derlying medical cause (eg malnutrition, hyperthyroidism, chronic liver
Lipoprotein glomerulopathy (LPG) was first reported in 1989. Cases disease, intestinal fat malabsorption such as seen in sprue and chronic
were mainly reported in east Asia including Japan and China though pancreatitis) and with parents with the same biochemical trait. FHL
later cases were reported in Europe and the United States, suggesting is caused by mutations in apoB, PCSK9 and angiopoietin-like proteins
that this disease may be more common in European populations than (ANGPTL3). Additionally as yet unidentified genes may cause the
originally thought. LPG is a rare inherited renal disease, transmitted as FHL phenotype, linkage of FHL to loci near 3p21.1–22 and 13q have
autosomal dominant with incomplete penetrance. Pathologically LPG been reported [293,294]. The precise prevalence of FHL is not known,
is characterized by dilated glomerular capillaries that are filled with li- though a prevalence of 1.9% was reported in the Framingham offspring
poprotein rich materials called lipoprotein thrombi. In addition to histo- population [295].
logical findings clinical manifestations are proteinuria and dyslipidemia.
Proteinuria is sometimes mild but in most cases progresses to nephrotic 2.5.1.1. Loss of function mutations in PCSK9. If the normal function of
syndrome [276]. PCSK9 is to reduce the number of LDLRs, one would expect loss of func-
The hypothesis that abnormal lipoproteins with rare apoE mutants tion mutations in the gene to increase the number of LDLRs and thereby
accumulate in the glomeruli was supported by the discoveries of apoE cause hypocholesterolemia. In line with this notion, livers of mice lack-
Sendai (R145P) and apoE Kyoto(R25C) [277,278]. The virus mediated ing PCSK9 show increased LDLR protein, increased clearance of circulat-
transduction of apoE Sendai resulting in the development of LPG in ing lipoproteins, decreased plasma cholesterol and increased sensitivity
apoE deficient mice confirmed the etiologic role of apoE mutation in to statins [296]. A search for mutations among subjects with low LDL
LPG [279]. The distribution of apoE2 Sendai among lipoprotein classes cholesterol levels, identified loss of function mutations in PCSK9:
was similar to apoE2 and apoE3. ApoE2 Sendai showed impaired Y142X, C679X, R46L. The mutations are associated with lower LDL
160 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

cholesterol levels resulting in reduced risk of IHD [297]. The reduction in for processing and secretion defects. Analysis of the autocatalytic
risk of IHD was larger than predicted by the observed reduction in LDL cleavage residue Gln152, suggested the autocleavage specificity
cholesterol alone. This could be because genotype is a better predictor was Gln N Met N Ala N Ser N Thr ~ Asn; all other residues led to the for-
of lifelong exposure to LDL cholesterol than LDL cholesterol measured mation of an unprocessed zymogen that acted as a dominant nega-
in adult life. [298]. The Y142X and C679X mutations were common in tive, retaining the WT zymogen in the cell [310].
African Americans (2% combined frequency) and rare in European The function of PCSK9 in areas other than promoting LDLR degra-
Americans (b 0.1%) and were associated with a mean lowering of 40% dation is under investigation. Recent data demonstrate that PCSK9
in LDL cholesterol. The plasma range of LDL cholesterol in subjects enhances the degradation of LDLR as well as the closely related
with the nonsense mutation in PCSK9 ranged from first to fiftieth per- receptors, VLDL receptor, LRP1 and apolipoprotein E receptor 2
centile suggesting other factors (eg diabetes, weight, age) may contrib- [311,312]. Studies in mice suggest that PCSK9 limits visceral adipo-
ute to plasma LDL cholesterol variation. Haplotype studies suggest that genesis, likely via adipose VLDL receptor regulation. Further, lesions
each of the two mutations had a single founder and that both were an- observed in the regenerating liver of PCSK9−/− mice were effectively
cient mutations [299]. Screening healthy hypocholesterolemic subjects prevented by feeding the mice high cholesterol diet indicating in the
for PCSK9 mutations identified subjects heterozygous for the mutations mice loss of PCSK9 leads to critical hypocholesterolemia [313]. The
R46L, G106R and R237W associated with hypocholesterolemia. The absence of PCSK9 can be protective against melanoma invasion in
mutation R46L and N157K in the PCSK9 gene were possibly associated mouse liver [314]. It has been suggested that VLDL receptor and apo-
with increased response to statin treatment in a group of FH heterozy- lipoprotein E receptor 2 exert major effects during brain develop-
gotes [300]. The R46L sequence variation was much less common in ment. However, the brains of PCSK9 −/− mice did not show overt
African Americans (0.3%) than in European Americans (1.6%) suggest- morphological defects [311]. The only phenotype of PCSK9 deficien-
ing an ethnic influence on allele frequency [301]. The loss of function cy in humans is lower plasma LDL cholesterol levels but it is possible
mutations R46L, G106R, N157K, and R237W had increased cell surface that aging or other environmental factors may reveal other pheno-
LDLR and increased level of LDL internalization compared to the wild types associated with PCSK9 deficiency. The loss of function PCSK9
type PCSK9 [302]. Other variants associated with lower LDL cholesterol SNP rs11591147 which resulted in lowering of cholesterol 10–16%
levels in population studies are L253F and A443T variants. Individuals was not associated with noncardiovascular clinical events in an el-
heterozygous for the nonsense mutations in PCSK9 did not show in- derly population [315].
creased hepatic TG content. The low LDL cholesterol variants were not PCSK9 is a secreted protein which circulates in serum in a truncated
associated with increased hepatic TG content suggesting that the rate and intact form [316]. ELISA to measure PCSK9 used antibodies raised
of VLDL secretion is not reduced and that impaired VLDL secretion is against recombinant full length PCSK9 and recombinant PCSK9 was
not the major mechanism for reducing LDL cholesterol. WT PCSK9 used as standards for the assay. It is not known which epitopes these an-
does not significantly affect the production of VLDL apoB [301,303]. tibodies recognize [317]. Circulating PCSK9 levels have been measured
Further a role for common PCSK9 variants (both coding and noncoding) in healthy middle aged individuals of European descent, in children
has been suggested for determining plasma levels of LDL cholesterol. and adolescents and ethnically diverse cohorts. The general findings
The effects of these variants are much smaller than the effects associated are that circulating PCSK9 levels vary over a wide range among appar-
with rare variants, levels differed by b5% from those without the com- ently healthy cohorts and correlate with plasma LDL concentration. It
mon alleles [301]. Population studies identified further novel mutations has been suggested that common and low frequency genetic variants
associated with hypocholesterolemia: frameshift mutation causing a in PCSK9 locus influence the individual variation among healthy middle
premature termination codon A68fsL82X, and amino acid substitutions aged individuals [318]. There are diurnal changes in PCSK9 and PCSK9
T771I, V114A, A522T and P616L [304]. A spectrum of loss of function concentrations are reduced by fasting [319]. Plasma concentrations of
mutations has been identified. A study of Japanese population showed PCSK9 do not appear to be useful for screening for novel PCSK9 variants
that one missense mutation R93C was statistically associated with low [320], but may be useful for monitoring treatments for FH patients
LDL cholesterol levels [305]. In a Canadian study multiple PCSK9 variants [321].
were reported in the population suggesting that the PCSK9 gene is a
highly polymorphic gene. Specific combinations of PCSK9 variants 2.5.1.2. Loss of function mutations in the inducible degrader of LDLR (IDOL).
may be required to evaluate the effect of multiple variants on LDL cho- IDOL is a sterol responsive E3 ubiquitin ligase recently identified as a
lesterol levels [306]. novel post-transcriptional regulator of LDLR. Acting as an E3 ubiquitin
A loss of function mutation Q152H at the site of autocatalytic cleav- ligase IDOL supports ubiquitylation of LDLR on its protruding intracellu-
age was studied in a French Canadian family. The mutation prevents the lar tail thus targeting it for lysosomal degradation. In response to rising
cleavage of the PCSK9 and precludes PCSK9 secretion. In cell culture cholesterol levels LXR induces IDOL production thus limiting further up-
conditions the mutation has a dominant negative effect and prevents take of exogenous cholesterol through the LDLR pathway. The IDOL
wild type PCSK9 secretion [307]. In a study of a French family, the mechanism for feedback inhibition is independent of and complemen-
monoallelic double mutant R104C/V114A exhibited a dominant nega- tary to the SREBP pathway [322]. Increasing the expression of IDOL re-
tive effect on the secretion of wild type PCSK9 secretion [308]. Zhao sults in decreased LDLR and LDL uptake into cells, whereas loss of
et al. [309] report a healthy, fertile compound heterozygote for IDOL expression results in increased LDLR [323]. IDOL gene was se-
an inframe 3 bp deletion (c.290_292delGCC) which deletes an arg at quenced in patients with LDL cholesterol below the 20th percentile in
codon 97 and the Y142X mutation. Expression of ΔR97 mutant showed whom mutations in LDLR, APOB, and PCSK9 were ruled out. The IDOL
that the mutant prevented autocatalytic cleavage and secretion of variant R266X represents a complete loss of IDOL function unable to
PCSK9. To determine the effect of loss of function mutations, the recom- promote ubiquitylation and subsequent degradation of LDLR. This is
binant forms of the protein were expressed in cells. The Y142X protein the first report of a loss of function IDOL variant and will require further
was predicted to initiate nonsense-mediated RNA decay and no examination in additional populations [324].
immunodetectable protein was found in the cell lysate. The mutations
R46L, A443T had no detectable effect on the processing or secretion of 2.5.1.3. ApoB specific hypobetalipoproteinemia. Apo B specific familial
PCSK9. Compared to the WT, the amount of mature protein was de- hypobetalipoproteinenia (FHLB, OMIM 615558), an autosomal codomi-
creased in the cells that expressed the L253F mutant suggesting nant disorder is associated with mutations in the APOB gene. Most APOB
that the mutant interfered with autocatalytic processing. Both the mutations lead to the formation of truncated apoB protein of various
L253F and C679X mutations interfered with the secretion of PCSK9 sizes which to a variable extent lose the capacity to form plasma lipo-
[309]. A number of natural and artificial mutants were investigated proteins in the liver and intestine and export lipids from these organs.
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 161

Missense nontruncating mutations of the APOB gene can also cause who was a heavy drinker had steatohepatitis whereas his teetotaller
FHLB. daughter, an apoB 75.7 carrier, had no detectable fatty liver.
A large kindred with a nonsense mutation leading to a stop codon at The following mutations were identified in the FHLB group:
residue 2486 of apoB (called apoB 54.8 according to the centile system) 2534delA apoB 18; Q1309X apoB 29, R2507X apoB 55, and 11712delC
was identified, following population screening for individuals with apoB 86 by Sankatsing et al. [335]. They observed a decreased level of ar-
≤10th percentile total cholesterol. The proband was an asymptomatic terial wall stiffness in FHLB subjects, most pronounced in the absence of
male, heterozygous for the apoB mutation [325]. A further mutation nonlipid risk factors, suggestive of cardiovascular protection in FHLB in-
with a premature stop codon at 3387 (apoB 75) was identified in a dividuals. The authors report slightly higher levels of HDL cholesterol in
kindred with hypobetalipoproteinemia. The proband was a 46 y old FHBL subjects. The latter can be a consequence of low levels of TG min-
healthy male heterozygous for the apoB mutation [326]. A nonsense imizing the exchange of cholesterol esters from HDL cholesterol to apoB
mutation that produced a truncated protein that contains 1754 amino containing particles through the action of CETP.
acid (apoB 38.7) was found in 57 y old woman with homozygous The structure of the βα1 domain of human apoB has been based on
FHLB. The patient presented with retinitis pigmentosa, acanthocytosis, its homology to lipovitellin. This homology model suggests that the βα1
and loss of deep tendon reflexes. In addition the patient had diabetes domain consists of a β barrel and an α helical domain. In vitro experi-
mellitus with nephropathy, anemia, cholelithiasis, hepatic hemangio- ments suggest that the βα1 contains multiple MTP binding sites
ma, bronchiectasis, and extensive calcification of major arteries [327]. encompassing residues 430–570, 512–721 and probably 2–154 [336,
Three splice-site mutations have been linked with FHLB: (i) c.3843- 337,338]. Using transfected cells, expression studies with the mutant
2A→G in which the joining of exon 24 to exon 26 lead to a premature R463W protein suggested that the mutant protein had an increased
termination codon and a protein of 1260 amino acids (apoB 27.77); binding for MTP and, in addition, the mutant protein was retained with-
(ii) c.904+4A→G in which an exon 7 and exon 9 junction lead to a pre- in the ER and was secreted poorly compared to the WT protein [329,
mature stop codon at position 248 (apoB 5.44) (iii) c.4217-1G→T which 330]. In other studies with transfected cell lines the mutant A31P was
joined exon 25 to a thymidine at position c. 5397 which lead to a prema- least secreted and the secretion of G912D barely affected. Degradation
ture stop codon at position 1381 (apoB 30.42). The truncated apoBs of the post ER A31P protein was achieved by autophagy. The studies
were not detected in the plasma suggesting that they were not secreted suggest that feedback inhibition of lipogenesis and degradation of
as lipoprotein constituents. Ultrasound and liver biopsy examination of apoB A31P can contribute to total fat accumulation in the liver. [332].
heterozygote carriers of these mutations revealed the presence of fatty Early studies suggest that the truncated forms of apoB are secreted at re-
liver [328]. Burnett et al. [329,330] reported two missense mutations duced rates. The secretion rates of apoB 89, apoB 75, apoB 54.8 and apoB
R463W and L343V in two FHLB kindred. Despite low lipid concentra- 31 were 92%, 64%, 37% and 12% of apoB100. Additionally the truncated
tions, in affected members, none had neurological problems or malab- forms had increased fractional catabolic rates. The linear regression
sorption. Compared with unaffected family members liver enzymes line of apoB length versus secretion rate has a zero intercept for apoB se-
were increased and α-tocopherol concentrations decreased in hetero- cretion at apoB 28, suggesting that the smaller forms of apoB are not se-
zygotes. Serum α-tocopherol concentrations were lowest in apoB creted [339,340]. In general, the shorter the truncated apoB form the
R463W homozygotes. R463W patients accumulate fat in their intestinal lower the lipid to protein ratio in the lipoprotein [341]. Thus low levels
enterocytes and exhibit a blunted postprandial rise in their blood lipids of truncated apoB containing lipoproteins are probably due to quantita-
[331]. Five nontruncating missense apoB mutations A31P, L324M, tive differences in their production, conversion or removal rates. It is
G912D and G945S were identified in heterozygous carriers of FHLB. likely that the occurrence of fatty liver in FHLB depends on the balance
All probands and some of the respective family members had fatty between hepatic TG synthesis and VLDL secretion. In FHLB heterozy-
liver documented by ultrasound examination [332]. gotes with mutations in apoB specifying short truncations of apoB unde-
In the study by Fouchier et al. [333] most FHLB heterozygotes were tectable in plasma, hepatic secretion of VLDL was decreased. VLDL
asymptomatic, though some presented with mild clinical phenotypes. apoB100 production rates in FHLB heterozygotes are lower than
Statistically significant differences between affected and unaffected would be expected from a functioning allele. The secreted VLDL was,
groups were found for plasma TC, LDL cholesterol, HDL cholesterol, TG however, relatively TG enriched. The cellular mechanism underlying
and apoB levels. The proband of the 11712delC (apoB 86) family pre- decreased VLDL apoB100 production in FHLB heterozygotes is not un-
sented with mild diarrhea and the proband of the R2507X (apoB 55) derstood. Theoretically the presence of truncated apoB may interfere
family presented with diabetes mellitus type I and neurological with VLDL assembly leading to increased intracellular degradation of
symptoms. The neurological complaints diminished after vitamin E apoB100 [342]. The clearance of truncated apoB which contain the
administration. The authors identified a total of 9 different apoB mu- LDLR binding domain is faster than the clearance of apoB100. It has
tations associated with the FHLB probands (R412X or apoB 9, been suggested that the enhanced clearance of truncated apoB is pri-
R463W, 1718delAT or apoB 12, 2534delA or apoB 18, 2783delC or marily mediated by LDLRs. On the basis of negative regulatory effect of
apoB 20, Q1309X or apoB 29, 11548del TT or apoB 84). All carriers of the ‘bow’ structure of the carboxy terminal of apoB, removal of the
the mutation were completely free from IHD. One patient inherited carboxy terminal can lead to enhanced receptor binding [343]. Trunca-
the R3500Q mutation from her father and the 11712delC mutation tions of apoB shorter than 3200 amino acids (apoB-70), which do not
from her mother. Her lipid profile suggested compensation of one disor- possess the LDLR region are redirected from the liver to the kidney,
der by another. which becomes the major organ of uptake [344]. The phenotype of
Tarugi et al. [334] reported on the clinical phenotype of three FHLB apoB dependent hypobetalipoproteinemia is the net result of several
kindred heterozygous for mutations which resulted in truncated processes including VLDL production rate and rate of removal of LDL
apoBs. The first proband was heterozygous for a C deletion at position via the LDLR.
1310/1311 leading to the synthesis of apoB 8.15, which was not detect- The clinical phenotype of heterozygous FHLB is not well defined.
able in the plasma. The proband showed steatorrhea and fatty liver. The Though often asymptomatic they can develop fatty liver. Several factors
authors suggest that very short truncated apoB without the MTP bind- affect the amount of TG accumulated in the liver. Schonfeld et al. [345]
ing domain which is part of apoB48, have a reduced capacity to assem- suggest different rates of free fatty acid flux to the liver could not ac-
ble dietary lipids into chylomicrons. In the second kindred a nonsense count for hepatosteatosis in FHLB subjects, and that hepatic steatosis
mutation leads to the formation of apoB 33.4. The proband had cerebro- is secondary to impaired VLDL secretion, irrespective of metabolic
vascular disease and fatty liver. Two other carriers in this kindred conditions. Genetically engineered mice with truncated apoB, display
showed fatty liver. In the third kindred, the proband was heterozygous fatty livers with compensatory down regulation of hepatic fatty acid
for a T deletion in exon 26 which resulted in apoB 75.7. The proband synthesis. Thus the amount of TG accumulated in the liver is probably
162 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

modulated by a number of genes. Both genetic and environmental fac- enterocytes, displayed steatorrhea and large TG droplets within the vil-
tors would affect the degree of fat accumulation. A novel nonsense mu- lus enterocytes. In addition hepatic lipogenesis underwent compensato-
tation in apoB (p.K2240X) was responsible for fatty liver in a large ry induction with increased VLDL secretion in the face of decreased
kindred. In this family, carriers of this mutation developed cirrhosis chylomicron secretion [354].
and liver cancer. The complex relationship responsible for the sequelae Pons et al. [355] report on two patients, compound heterozygotes for
of fatty liver requires further investigation [346]. A case of FHLB who mutations within the MTP gene leading to severe ABL. The first genomic
was heterozygous for c.11040 T N G mutation (apoB 80.5) was in addi- mutation c.619G N T resulted in a deletion of exon 6 and a shift in the
tion carrying the HFE mutations C282Y and H63D as well as the α1- open reading frame leading to a premature stop codon at position 234
antitrypsin PiZ variant. The case had an unusual combination of disor- resulting in an abnormal truncated protein of 233 amino acids. The sec-
ders that could all be contributing to the observed liver dysfunction ond genomic mutation c.1237-28A N G induces a deletion of exon 10,
and fatty liver [347]. The apoE genotype has been reported to account with no shift in reading frame leading to a protein of 858 amino acids.
for 15–60% of the variability in plasma LDL in FHLB subjects. The physi- After expression in HeLa and HepG2 cells both MTP mutants localized
ological basis for this effect is yet to be determined [348]. in the ER despite not binding to protein disulfide isomerase. The two
Clinical features of homozygous FHLB can include acanthocytosis, mutants induced a loss of MTP TG transfer activity leading to ABL. The
deficiencies of fat soluble vitamins secondary to malabsorption, atypical first patient presented at 13 months with severe malabsorption. His sis-
retinitis pigmentosa and neuromuscular abnormalities. These defects ter, diagnosed at 6 years had no rotula reflex. Both patients presented
can also be observed with FHLB heterozygotes. Retinitis pigmentosa with complete absence of LDL and almost complete absence of VLDL
and other neuropathies are primarily the result of deficiencies of fat sol- and low concentrations of HDL particles, raised alanine transaminase
uble vitamins, attributable to their impaired transport. A patient identi- and severe vitamin E deficiency. The patients had severe immunoglob-
fied as a compound heterozygote (c.1315C NT apoB 39/c.537 + 1G N T ulin deficiency requiring immunoglobulin substitution. An intestinal
apoB 3.4–4.5) for FHLB was admitted at 4 months for severe fat malab- endoscopy of the 13 month old patient showed a white aspect of duode-
sorption. He was treated with medium chain fatty acids. He developed nal villi characteristic of intestinal fat malabsorption. Both parents had
cirrhosis, portal hypertension, esophageal varices as well as severe normal lipid values ruling out a dominant form of inheritance. Other
neurological deficiencies by the age of 11 [349]. The clinical status mutations resulting in a truncated form of the MTP protein and a severe
of homozygous FHLB patients is variable. A 48 year old female homo- ABL phenotype are: c398-399delAA [356], c2611delC, c923 GNA [357],
zygous for the mutation c.6240T N A which resulted in a truncated c.419-420insA [358], c.2076-39_2303 + 52del319 [359], c.619-3T N G,
apoB, apoB 45.2 was asymptomatic with normal levels of vitamin E. c.923G NA [360]. A patient with subclinical hypothyroidism and ABL
[350]. It is likely that apoB mutations will affect lipid homeostasis was found to be a compound heterozygote for a splice site mutation
in multiple cell types and multiple mechanisms may be called upon in intron 1 (c.61 + 2T N C) and a single insertion in exon 4 (c.419-
to maintain lipid homeostasis. 420insA) which resulted in a truncated protein of 140 amino acid
[361].
2.5.2. Abetalipoproteinemia Di Filippo et al. [362] report a case of ABL with an unusual clinical
MTP is an ER protein that catalyzes the transfer of neutral lipids and phenotype. The four year old girl presented with severe liver injury,
is essential for the formation of VLDL and CM. The MTP gene found on low levels of LDL cholesterol, and subnormal levels of vitamin E,
chromosome 4q22–24 codes for an 894 amino acid protein which is pri- but only mild fat malabsorption and no retinitis pigmentosa or
marily expressed in enterocytes and hepatocytes. MTP is found as a het- acanthocytosis. The subject exhibited compound heterozygosity for
erodimer with the enzyme, protein disulfide isomerase. The isomerase two novel MTP mutations: one missense mutation (p.L435H) and an
is inactive with respect to its enzyme activity in the MTP complex. Pro- intronic deletion (c.619-5_619-2del). Transfected cell lines expressing
tein disulfide isomerase mutants lacking enzyme activity are fully func- the missense mutation exhibited negligible levels of MTP activity. In
tional in lipid transfer activity if associated with normal MTP. Based on contrast the intronic deletion showed an incomplete splicing defect
the sequence homology of MTP with lipovitellin, MTP is proposed to with 26% of normal splicing being retained. The small amount of MTP
have three structural motifs: N-terminal β-barrel (amino acid residues activity resulting from residual normal splicing in the patient is thought
22–297) which mediates the interaction with apoB, central α-helix to explain the atypical phenotype.
(amino acid residues 298–603) which associates with protein disulfide More than 30 MTP gene mutations causing ABL have been described.
isomerase and C-terminal lipid binding domain (amino acid residues Narcisi et al. [363] report a cohort of eight patients who presented with
604–894) which is involved in lipid binding and lipid transfer [351]. classical ABL, fat malabsporption during infancy, low plasma cholester-
Abetalipoproteinemia (ABL, OMIM 200100) is an autosomal reces- ol, TG and apoB levels. In contrast HDL cholesterol was within reference
sive hypocholesterolemia which presents in its severe form, during in- range in all but two patients. Mutations which were identified in both
fancy with failure to thrive, severe diarrhea and a lipid malabsorption alleles of the MTP gene in all eight patients, resulted in truncated form
syndrome. In ABL patients, neurological impairment (ataxia and periph- of MTP with variable number of amino acids at its carboxy terminal
eral neuropathy) and retinopathy are a consequence of vitamin A/E de- end. Missense mutations R540H, S590I, G746E and N780Y have been
ficiency as well as deficiencies in polyunsaturated fatty acids. Clinical found to cause ABL. The R540H mutation prevents the association of
and biochemical features of ABL are generally indistinguishable from the protein disulfide isomerase with MTP causing the mutant MTP to
homozygous and compound heterozygous FHLB, though more severe be nonfunctional based upon its inability to reconstitute apoB secretion
than heterozygous FHLB. The clinical presentation of ABL can, however, in a cell culture system [364]. The predicted tertiary structure suggests
be heterogeneous and its subsequent downstream complications de- that R540H and S590I are located in the central α helical domain where-
pend on early diagnosis and treatment [352]. Mutations in the MTP as G746E and N780Y are in the C-terminal β sheet domain. Mutants
gene are the principal cause of ABL. In some instances, frameshift, non- S590I, G746E and N780Y interact with protein disulfide isomerase but
sense and splice site mutations that are located throughout the entire showed no phospholipid or TG transfer activity and did not support
MTP gene are predicted to encode truncated forms of MTP resulting in apoB secretion [365]. Ohashi et al. [366] describe a 27 year old male ho-
complete loss of function. The others are missense mutations that mozygous for the N780Y mutation with none of the manifestations of
cause ABL by reducing the ability of MTP to transfer lipids. Deficient classical ABL even though his plasma apoB and vitamin E were virtually
MTP activity targets apoB for degradation preventing the secretion of undetectable. A 58 year old male, homozygous for the mutation S590I
CM and VLDL. MTP knockout mice died during embryonic development with the absence of apoB containing lipoproteins, acanthocytosis was
suggesting deficiency of MTP caused lethal embryonic development notable only by the absence of neurological involvement and normal vi-
[353]. Mice with conditional deletion of the MTP gene in villus tamin E concentration. He subsequently developed adenocarcinoma of
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 163

the ileum. It is likely that environmental and genetic factors can influ- not show a difference in post heparin LPL mass or activity, whereas ho-
ence the presentation of the ABL phenotype [367]. One study reported mozygotes showed a 3.5 fold higher levels in these parameters. The
an association between common MTP polymorphisms and LDL levels FCHL phenotype is at least in part attributable to deficiency of ANGPTL3
[368]. and increased activity of LPL. One explanation for the low levels of HDL
Two patients with contrasting phenotypes were presented by Miller seen in homozygous subjects is that the S17X mutation affects the
et al. [369]. The first patient was a compound heterozygote with one de- secretion of the lipoproteins or that ANGPTL3 acts by inhibiting EL.
letion (c 103_127del25) and another missense mutation (Y528H) who The reduced levels of FFA observed in homozygotes can contribute
presented at 4 months with failure to thrive. The second presented at to decreased rate of VLDL production by the liver [381]. Other muta-
37 years with marked hypocholesterolemia, and was found to carry tions in ANGPTL3 associated with FCHL are: pG400VfsX5, pI19LfsX22,
three missense mutations: G264R, R540C, and N649S. Studies with pN147X, E129X [382,383].
transfected cell lines suggested that all missense mutations bound to Minicocci et al. [384] studied the biochemical clinical characteristics
apoB48 and protein disulfide isomerase. Mutations Y528H and R540C of 14 homozygous, 8 compound heterozygous and 93 heterozygous
displayed negligible levels of MTP lipid transfer activity and N649S carriers of mutations in ANGPTL3. The authors identified nonsense
displayed a partial reduction. G264R retained full lipid transfer activity. mutations, frameshift mutations (eg p.S122Kfs*3), nucleotide deletions
The R540C is predicted to disrupt an internal salt bridge that has an ef- (eg c.361-365delAACTC) and missense mutations (eg p.G56V). Com-
fect on MTP structure. The N649S converts a hydrophilic asparagine to a pared with noncarriers, homozygotes as well as heterozygotes had a re-
smaller still polar serine. The partial activity of N649S may provide a duction of all plasma lipoproteins and apolipoproteins. Compared to
biochemical explanation for the mild ABL phenotype of the patient. homozygotes, the presence of a single mutant allele reduced lipoprotein
Diagnosis of FHLB and ABL depend on signs and symptoms, levels to a lesser extent, suggesting a codominant form of inheritance.
acanthocytosis on blood smear and virtually absent apoB containing However, there was not a linear gene-dosage dependent effect of
lipoproteins. Obligate heterozygote parents of ABL patients have ANGPTL3 mutations on lipoprotein levels. The authors explored the
normal lipid levels consistent with autosomal recessive inheritance. genotype-phenotype relationship of different mutant alleles on plasma
Heterozygous FHLB may variably show some of the clinical features ANGPTL3 concentration. Though the analysis was hampered by the
and have TG and LDL cholesterol levels around 5th-10th percentile small number of patients, mutations predicting truncated proteins had
for age and sex, consistent with autosomal co-dominant inheritance, the most pronounced effect in reducing plasma levels of ANGPTL3.
though a high degree of interindividual variability is seen even with- The functional role of missense mutations can be only be clarified by ex-
in the same kindred. In general, laboratory investigations show a pression studies. The authors did not detect fatty liver in their cohort of
lipid profile with nearly absent TG, apoB and LDL cholesterol for FCHL subjects and found that none of the homozygotes were affected by
ABL and homozygous FHLB patients. Definite diagnosis involves se- diabetes. It is possible that ANGPTL3 has a more comprehensive role in
quencing of the MTP and apoB genes. Investigations include lipid regulation of not only lipid but glucose metabolism. Low plasma
profile, serum transaminases, markers for lipid soluble vitamins, HDL and reduced cholesterol efflux from cells observed by Pisciotta
and periodic assessment of ocular and neurological function. Treat- et al. [382] can contribute to increased atherosclerosis but may be
ment includes a low fat diet, and supplementation with essential counterbalanced by lifelong low exposure to low LDL cholesterol, and
fatty acids and high doses of fat soluble vitamins. Prognosis is vari- merits further investigation.
able and may depend on early diagnosis and adherence to treatment A common variant of ANGPTL4, E40K, that was present in ~ 3% of
[370,371,372]. European-Americans was associated with lower levels of TG and LDL
cholesterol and higher levels of HDL cholesterol. Expression studies
2.5.3. Familial combined hypolipidemia with wild type and mutant protein in cells suggested that ANGPTL4
A family of proteins structurally similar to angiopoietins ANGPTLs formed dimers and tetramers in cells prior to secretion. Following secre-
has been identified. To date 8 ANGPTLs have been discovered. In tion and cleavage of the protein the oligomeric structure of the N-
humans ANGPTL3 is located on chromosome 1 (1p31.1-p22.3) and terminal domain was retained while the C-terminal dissociated into
encodes a 460 amino acid protein consisting of a signal peptide, an monomers. Inhibition of the cleavage did not interfere with the oligo-
N-terminal segment containing coiled-coil domains and a C-terminal fi- merization of ANGPTL4 or its ability to inhibit LPL, whereas mutations
brinogen like domain which are connected by a linker region. All that prevented oligomerisation compromised the ability of the protein
ANGPTLs have a similar structure except ANGPTL8 which lacks the fi- to inhibit LPL. The E40K substitution destabilizes the protein after secre-
brinogen domain. ANGPTL3 is activated by cleavage at a proprotein tion, preventing the oligomerisation of the protein and abolishing the
convertase consensus site to release the N-terminal domain and its ac- ability of the protein to inhibit LPL [385]. Multiple loss of function muta-
tivity is regulated by ANGPTL8 [373,374]. Both ANGPTL3 and ANGPTL4 tions in ANGPTL3 and ANGPTL4 genes were associated with low plasma
inhibit LPL activity and mice lacking Angptl3 and Angptl4 genes have in- levels of TG [386,387]. ANGPTL4 variants E40K and T266M are associat-
creased LPL activity and reduced levels of TG [375]. ANGPTL3 acts as an ed with low TG levels in patients with type 2 diabetes [388]. A signifi-
inhibitor of EL, which is recognized as one factor that influences HDL cant association between SNPs in the RXR gamma gene and FCHL was
metabolism [376]. Although ANGPTL3 and ANGPTL4 both inhibit LPL, reported in a further study suggesting that other genes may play a
their effect on lipoprotein metabolism is divergent and complex. It has role in the genetic susceptibility to FCHL [389].
been suggested that ANGPTL3 and ANGPTL4 inhibit LPL through distinct
mechanisms [377,378]. Both irreversible dissociation of LPL dimers or 2.5.4. Other candidate genes
reversible noncompetive inhibition have been described as the mecha- GWAS have identified a locus on chromosome 1p13 which was
nism of inactivation of LPL by ANGPTL4 [379]. Loss of function mutation strongly associated with LDL cholesterol expression [390]. A com-
in ANGPTL3 has been shown to cause familial combined hypolipidemia mon noncoding polymorphism at the 1p13 locus rs12740374 created a
(FCHL, OMIM 605019). In a cohort of subjects with low LDL, TG and HDL, C/EBP transcription factor binding site and altered the expression of
the prevalence of ANGPTL3 gene mutations was about 10% [380]. SORT1 gene which modulates VLDL secretion and alters plasma VLDL
In a cohort of patients with loss of function S17X mutation in levels [391]. Hepatic overexpression of sortilin reduced apoB secretion
ANGPTL3, homozygous patients showed complete loss of ANGPTL3 and increased LDL catabolism resulting in reduced LDL cholesterol [392].
while heterozygous S17X carriers showed partial deficiency. Homozy-
gotes showed reduced cholesterol levels in VLDL, LDL and HDL and a 2.5.5. Chylomicron retention disease
marked reduction in TG. Heterozygotes showed similar lipoprotein The SAR1B gene encodes the Sar1b protein which is involved in CM
levels to noncarriers. Compared with noncarriers, heterozygotes did transport from the ER to the Golgi apparatus. The Sar1b protein belongs
164 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

to a family of small GTPases and is a coat protein II (COPII) component. Descriptions of cardiac and muscular abnormalities may be related
The COPII coat drives vesicle formation. COPII vesicle formation is initi- to tissue specific expression of the abnormal Sar1b protein. Creatine ki-
ated by guanine nucleotide exchange factor Sec12, which recruits Sar1b. nase levels were increased in patients homozygous for a mutation in the
Sar1b-GTP in turn recruits Sec23/Sec24 dimer which binds the cargo seventh codon which resulted in a change in Glutamic acid to a stop
molecules. These molecules recruit a heterodimer Sec13/Sec31 which codon, resulting in a truncated protein of 166 amino acids as well as in
drives membrane bending and vesicle formation [393,394]. The SARIB a patient with the G37R mutation [403].
gene is located at chromosome 5q31.1. It is composed of eight exons One study describes the phenotypic expression of CMRD and its evo-
and alternative splicing of exon2 is predicted to lead to two transcripts. lution over time in a French Canadian cohort and a French cohort of
A second human protein isoform of Sar1, Sar1a is encoded by SAR1A Turkish, Algerian and Portugese stock. All 16 CMRD patients presented
gene located in chromosome 10q22.1. In mice the evaluation of Sar1b within the first few months of life with diarrhea and failure to thrive. Se-
mRNA revealed the skeletal muscle as the tissue with the highest vere hypocholesterolemia coupled with normal TG was associated with
Sar1b expression, followed by the heart and the liver, the organs com- low LDL and HDL cholesterol and low apoAI and apoB. The late diagnosis
posing the digestive tract, brain and finally the lung and adipose tissue in the French cohort and the later introduction of a low fat diet with a
[395]. high polyunsaturated fatty acid/saturated fatty acid ratio and supple-
Recently mutations in the SAR1B gene were identified as respon- ments of fat soluble vitamins A, D and E in doses recommended for
sible for chylomicron retention disease (OMIM 246700, CMRD) or ABL was thought to account for the delayed catchup growth in the
Anderson's disease. When SAR1B is mutated, the pre-chylomicron French cohort compared to the Canadian patients. Clinically only one
transport vesicle cannot fuse with the Golgi apparatus, which then case had a neurological handicap and one patient suffered from hepato-
induces an accumulation of pre-chylomicron transport vesicles in megaly [404]. Neurologic and ophthalmologic complications are less se-
the cytoplasm of the enterocytes. The condition is characterized by vere than FHLB and ABL. [405]. The similarity of CMRD to other types of
severe fat malabsorption, usually associated with failure to thrive hypocholesterolemia can lead to the wrong diagnosis unless molecular
in infancy as well as deficiency in fat soluble vitamins, low blood cho- analysis is performed. Sequencing of SAR1B provides an accurate diag-
lesterol levels and the absence of CM from blood. Affected individ- nosis of CMRD.
uals accumulate CM like particles in the enterocytes which contain
large cytosolic droplets. Jones et al. [396] identified coding sequence 2.5.6. ApoCIII deficiency
variants in SAR1B in affected individuals. In total eight mutations, The APOC3 gene resides within the APOA5–APOA4–APOC3–APOA1
two frameshift, one splice site and five missense alleles were identified. multigene human cluster on human chromosome 11q23. The locus pro-
The five missense mutations (G37R, D137N, S179R, S179I, L181P) cause vides a strong association signal in GWAS of plasma TG concentration,
changes to the GTP binding pocket of Sar1b and the homozygous splice although there is some ambiguity of the genes that underlie this signal,
site mutation is predicted to abolish the production of functional Sar1b. both APOC3 and APOA5 have been associated with plasma TG concentra-
Two patients homozygous for the mutation (p. D48TfsX17) which gave tion [406]. ApoCIII is a 79 amino acid O-linked glycoprotein composed of
rise to a truncated protein and admitted at b 1 year of age with diarrhea, 6 amphipathic α helices and is expressed in the liver and intestine [407].
acquired tolerance to a fat diet in later years. A further patient homozy- On entry into the plasma, apoCIII is rapidly exchanged among TG rich li-
gous for a frameshift stop codon (p.L28RfsX7) was a heterozygous poproteins and HDL, which acts as a reservoir for this apolipoprotein.
carrier of a PCSK9 (p.L21dup) variant [397]. Cefalu et al. [398] describe The distribution of apoCIII is however, asymmetrical, with different
a proband, a five month old baby with mutation in the SAR1B gene apoCIII molecules per particle within HDL and VLDL pools [16]. Recent ev-
(c. 75–76 del TG-L28fsX34) who presented with failure to thrive idence suggests that apoCIII and apoAV are opposing modulators of plas-
and malabsorption. His father was heterozygous for the mutation ma TG metabolism [408]. Plasma apoCIII concentration varies from
but his mother who was homozygous for the mutation was symptom ~10 mg/dL in normal subjects to N 30 mg/dL in hypertriglyceridemic sub-
free of CMRD. It is likely that that CMRD is a more complex disorder jects. Increased apoCIII production is a characteristic feature of patients
and that further modifier genes may be involved in the ER-to Golgi with hypertriglyceridemia [409].
transport. Charcosset et al. [399] describe three further unique homozy- ApoCIII expression is down regulated in part by the insulin response
gous mutations in French families who originated from Turkey, Algeria element (IRE) on the APOC3 gene. A variant of APOC3 promoter within
and Portugal: stop codon in exon 6 (p.E122X), a whole deletion of exon the IRE results in loss of insulin regulation. The variant APOC3 promoter
2 (c. 1-4482_58+1406 del 5946 ins15bp) and a missense mutation in is common in human population and can contribute to the development
exon 7 (p.G185V). Phenotypical data for missense mutations were not of hypertriglyceridemia [410]. The transcription of APOC3 is mediated
different from nonsense mutation, conversely the same missense by several other transcription factors including PPAR [411].
mutation was associated with different profiles between families Several functions have been proposed for the role of apoCIII in regu-
and even among the same family. The authors observed a lack of lating TG metabolism. ApoCIII is known to inhibit LPL activity. It has
genotype-phenotype correlation among patients with CMRD carry- been suggested that when bound to TG rich lipoproteins apoCIII pre-
ing mutations in the SAR1B gene and suggest that mutations and vents the binding of LPL to the lipid surface. TG rich lipoproteins stabi-
polymorphism in modifier genes eg transcriptional factors might af- lize LPL and protect the enzyme from inhibitors like ANGPTL4. The
fect the functionality of the network of proteins involved in ER-to- addition of apoCIII decreased the protective effect of TG rich particles
Golgi transport. on LPL and increased their susceptibility to factors such as ANGPTL4
Sar1b also promotes hepatic VLDL secretion which would explain [412]. ApoCIII also impairs the hepatic uptake of triglyceride rich lipo-
the observation that some CMRD patients develop hepatic steatosis. It proteins, possibly by interfering with the interaction of apoE on the lipo-
has been suggested that Sar1b modulates the expression of genes proteins with LDL and LRP receptors [413]. The metabolism of VLDL is
encoding cholesterol biosynthesis which may explain the severe influenced by its content of apoE and apoCIII [414]. An intrahepatic
hypocholesterolemia phenotype found in CMRD [400]. Patients with role of human apoCIII in promoting VLDL secretion was suggested fol-
CMRD have a lower rate of production of apoB100 and apoAI from the lowing in vitro studies with transfected cell lines [415]. ApoCIII or
liver [401]. It is likely that a network of modifier genes are involved apoCIII rich VLDL may contribute directly to atherogenesis by activating
the mechanisms that determine the phenotype of the disease in any endothelial cells and recruiting monocytes to them [416].
given individual and that CMRD is a more complex disease than a simple GWAS identified heterozygous carriers of the null mutation (R19X)
autosomal recessive disorder. Okada et al. [402] describe a patient with in the gene encoding apoCIII in an isolated Amish population. The
CMRD phenotype, a normal SAR1B gene coding sequence and maternal carriers expressed half the amount of apoCIII present in noncarriers.
uniparental disomy of chromosome 7. Mutation carriers compared with noncarriers had lower fasting and
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 165

postprandial TG, higher levels of HDL cholesterol and lower levels of mutations in APOAI, LCAT, and ABCA1. Rare and common variants
LDL-cholesterol. Subclinical atherosclerosis, as measured by coronary can contribute to genetic variations in HDL. Mutations in apoAI,
artery calcification, was less common in carriers than noncarriers, sug- LCAT, PLTP and ABCA1 were found in 12% of subjects with HDL deficien-
gesting that lifelong deficiency of apoCIII may have a cardioprotective ef- cy, leaving 88% of HDL deficient subjects without a genetic diagnosis
fect [417]. In isolated European populations the frequency of the rare [435]. In a further study, one of six individuals with HDL cholesterol
allele (R19X) has been shown to be increased [418]. A further mutation below the fifth percentile had rare mutation in APOA1 or ABCA1 [436].
A23T was found in Mayan Indian subjects with hypotriglyceridemia. Recently data have emerged that suggests extreme levels of HDL choles-
The mutant displayed reduced affinity to phospholipid liposomes with terol in families have a polygenic origin. Greater than 40 different genes
normal inhibitory effect on LPL activity. The in vitro studies suggest that are currently reported to affect HDL levels [437].
the less efficient binding of the A23T mutant might lead to faster catab-
olism of the protein and reduced plasma apoCIII levels [419]. Other 3.2.1. Monogenic disorders associated with decreased HDL levels
studies suggest that the presence of the mutant can prevent VLDL as-
sembly and secretion from hepatocytes with TG accumulation within 3.2.1.1. ApoAI deficiency. The gene for ApoAI is located on chromosome
the microsomal lumen independent of MTP activity [15]. The naturally 11q23 and consists of 4 exons encoding the primary transcript of 267
occurring mutation K58E abolished the stimulatory effect of apoCIII on amino acids, subsequent cleavage of both the intracellular and the se-
the assembly and secretion of VLDL; suggesting that the C-terminal creted protein produces a 243 amino acid protein necessary for the as-
lipid binding domain of apoCIII plays a role in VLDL assembly [420]. sembly of HDL Residues 44–243, are composed of eight 22 amino acid
Whole genome sequencing identified a rare variant of APOC3 gene and two 11 amino acid repeats that are predicted to form amphipathic
(rs138326449) associated with plasma TG levels [421]. α-helices [438,439]. The frequency of hypoalphalipoproteinemia due
Confirmation of the role of apoCIII in the clearance of TG rich lipopro- to mutant apoAI (OMIM 604091) was estimated as 6% in Japanese sub-
teins comes from transgenic mouse studies. Disruption of the APOC3 gene jects with low HDL and 0.3% in the Japanese population [440]. Absence
results in protection from postprandial hypertriglyceridemia, while the of apoAI in a patient resulted in low HDL, premature coronary artery dis-
overexpression of human APOC3 gene results in hypertriglyceridemia ease and corneal opacification. The patient's defect was a deletion of the
[422,423]. The mutation apoCIII Q38K resulted in a moderate but signifi- entire apoAI/C3/A4 gene complex [441]. Two sisters with HDL deficien-
cant elevation in TG in a Mexican family. The elevated plasma levels of TG cy, undetectable plasma AI and apoCIII with premature IHD and coro-
further supports a role for ApoCIII in the regulation of TG metabolism nary bypass by the age of 30 were described by Norum et al. [442].
[424]. The defect was a DNA rearrangement affecting the adjacent apoAI and
apoC3 genes. Two siblings with HDL deficiency and no plasma ApoAI
3. Disorders of HDL metabolism were found to be homozygous for the deletion in exon 3 of the apoAI
gene (c.85 delC, Q5FsXII). This mutation causes a frameshift leading to
3.1. HDL and cardiovascular disease a premature stop codon and abolished the synthesis of apoAI. Although
both siblings had corneal opacification and planar xanthomas only one
Numerous population studies have shown that an inverse relation- of them had premature coronary artery disease, probably as a result of
ship exists between plasma HDL and IHD risk. Two HDL subclasses can mildly elevated LDL levels. The carriers had HDL cholesterol values
be separated by ultracentrifugation: HDL2 is less dense and relatively below the 10th percentile of the population, though none had planar
lipid rich, HDL3 is more dense and relatively protein rich. On agarose xanthomas or coronary artery disease. In two further kindred HDL
gel the HDL separates into α migrating particles which represent the deficiency was due to nucleotide substitution in exon 4 (L141R) of
majority of circulating HDL and preβ particles which represent poorly the apoAI gene. This mutation was reported to hamper the ability
lipidated HDL. Further resolution can be achieved by a 2-dimensional of apoAI to recruit cellular cholesterol and to activate LCAT. In addi-
electrophoretic method [425]. A 0.026 mM (10 mg/L) increment in tion the L141R mutation affects the stability of apoAI in plasma
HDL cholesterol was associated with a risk decrease in IHD of 2% in [443]. A further mutation S36A was identified in a subject with se-
men [426,427]. Later studies suggest that smaller dense HDL3 choles- vere hypoalphalipoproteinemia. Although the ability of the S36A
terol drove the inverse association with HDL and IHD [428]. The in- mutation to bind to lipoprotein surfaces was not altered, the muta-
verse relationship of HDL cholesterol with IHD risk decreased in tion significantly impaired LCAT activation [444]. A 67 year old
patients with coronary artery disease, suggesting that HDL choles- male homozygous for the V156E substitution in apoAI presented
terol has limited use in cardiovascular risk stratification in such with HDL deficiency, corneal opacity and coronary artery disease.
patients [429]. HDLs display a wide spectrum of activities which include The son who was heterozygous for the mutation showed normal
in addition to cholesterol efflux, anti-inflammatory and anti-oxidative HDL levels. In vivo studies in rabbits suggested that the variant was
action and maintenance of endothelial function [430]. HDL consists of rapidly cleared from plasma compared to normal apoAI [445]. Deletion
several subpopulations of particles who do not all inhibit atherosclero- in two successive adenine residues in codon 138 of the apoAI gene
sis to the same extent [431]. A plateau effect in terms of IHD risk reduc- resulting in a frameshift mutation at amino acid residues 138–178 was
tion was observed for men and women with high levels of HDL [432]. identified in a patient with marked HDL cholesterol and apoAI deficiency,
However, a causal relationship between low HDL and IHD risk is not un- myocardial infarction and moderate corneal opacities. His brother who
disputed. Many different factors affect both HDL levels and IHD risk, and was homozygous for the mutation with marked HDL cholesterol and
statistical adjustments do not guarantee the absence of other confound- apo AI deficiency had no clinical evidence of coronary heart disease. Het-
ing factors. It is likely that HDL subspecies or individual HDL compo- erozygotes had 50% of normal plasma HDL cholesterol with mean apoAI
nents will better predict atherosclerosis [433]. One view is that it is concentrations decreased in subpopulations (large α1 and small preβ)
not HDL concentration but HDL function that has a causal relationship of HDL particles [446]. DNA sequencing of a 51 year old patient revealed
with atheroprotection [434]. homozygosity for a frameshift mutation in exon 4 of the apoAI gene lead-
ing to a stop codon at residue 178. The patient exhibited corneal opacities,
3.2. Decrease in HDL concentration planar xanthomas and extensive coronary heart disease [447]. A deletion
in exon 4 of the apoAI gene leading to frameshift mutation and early ter-
Familial hypoalphalipoproteinemia is defined as an HDL cholesterol mination at codon 200 was identified in a 69 year old patient with
level below the 10th percentile without a secondary cause and asso- marked apoAI deficiency, corneal opacity but neither xanthomas or
ciated with a family history of low HDL levels. When considering symptoms of coronary heart disease [448]. HDL subpopulations in homo-
hypoalphalipoproteinemia single gene, Mendelian disorders are zygotes and heterozygotes in patients with a nonsense mutation in apoAI
166 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

codon −2,Q[−2]X, displayed altered chemical composition with deple- prevents the conversion of lipid poor apoAI particles into preβ-HDL
tion in apoAI, cholesteryl ester, and glycerphospholipid. The authors sug- and causes rapid catabolism of the poorly lipidated HDL mainly by the
gest that the defective atheroprotective properties of HDL were related to kidney [464]. Hepatocyte specific ABCA1 knockout mice have a similar
altered lipid and apolipoprotein composition [449]. lipid phenotype to Tangier disease, with a twofold elevation of plasma
Mutants apoAIMilano (R173C) and apoAIParis (R151C) are natural var- VLDL TG, 50% lower LDL and 80% reduction in HDL concentration. The
iants of the apoAI that cause no deleterious health effects despite low lipid phenotype arises from an increased secretion of hepatic VLDL, in-
levels of plasma apoAI and HDL and high levels of HDL triglycerides in creased uptake of LDL by the LDLR, elimination of nascent HDL particle
heterozygote mutation carriers [450,451]. While some studies report assembly by the liver and hypercatabolism of apoAI by the kidney
an enhanced ability of apoAI(R173C) to recruit cellular cholesterol [465]. In the absence of ABCA1 newly secreted apoAI do not form
others a report a similar effect of apoAI(R173C) to the wild type protein preβ-HDL particles, which leads to decreased PI3K activation and in-
[452,453]. It has been suggested that the arginine is potentially involved creased lipid mobilization [466].
in an intrahelical salt bridge with E169 and that the loss of the arginine A proband was a 6 year old Caucasian female born from consanguin-
residue destabilizes the helix domain of the apoAI molecule altering its eous marriage. Her medical history included hepatosplenomegaly, ane-
lipid binding characteristic [454]. It has been proposed that disulfide mia and thrombocytopenia. At the bone marrow biopsy a high number
bonds involving C173 or C151 can influence protein conformation and of lipid laden macrophages were observed. Her plasma profile showed
facilitate dissociation of the C-terminal from HDL to recruit additional an HDL cholesterol of 0.06 mmol/L and undetectable apoAI levels.
lipids [455]. This patient was found to be homozygous for an intronic mutation
Naturally occurring mutations in human apoAI are known to be as- (c.4465-34AN G) predicted to encode a truncated protein. The authors
sociated with low plasma HDL levels and hereditary amyloidosis. The describe five patients (aged 6 months to 76 years) who presented
hereditary amyloidogenic mutations are clustered in the amino termi- with Tangier disease, and shared two clinical manifestations of Tangier
nus and in residues 170–180 of the protein. In most hereditary disease, thrombocytopenia and splenomegaly. In four of them bone
amyloidogenic mutations, the N-terminal fragment of mutated apoAI marrow biopsy revealed the presence of lipid loaded macrophages,
is the predominant form of the protein found in amyloid fibril deposits, one patient had peripheral sensory neuropathy, and clinical manifesta-
suggesting that the N-terminal region is critical to the amyloid fibril for- tions of premature coronary artery disease were documented in one
mation of apoAI. The first reported amyloidogenic apoAI variant G26R is adult patient and his twin brother. Four patients without classical man-
hypothesized to allow the β-strand of residues 20 to 25 to extend to po- ifestations of Tangier disease were found to be heterozygous for ABCA1
sition 26 and beyond, thereby increasing the likelihood for more global mutation and were classified as familial HDL deficiency [467]. A patient
β-strand interactions [456,457]. The L178H variant is associated with homozygous for the R1270X in the ABCA1 gene mutation presented
severe cardiac and laryngeal amyloidosis as well as skin lesions. Howev- with a bleeding tendency and anemia secondary to spontaneous splenic
er in contrast to G26R, L178H exhibits increased helical content [458]. hematoma [468]. A homozygous G–C splice site mutation adjacent to
Genetic analysis of the apoAI gene revealed three missense mutations exon 33 at the − 1 position, in the ABCA1 gene was associated with a
A175P, F71Y, E34K and a frameshift mutation H155MfsX46 which multiple sclerosis like neurologic presentation and the presence of sub-
were associated with amyloidosis [459]. clinical atherosclerosis [469].
The diagnosis of Tangier disease requires the demonstration of mo-
3.2.1.2. Tangier disease. The ABCA1 gene is localized on chromosome lecular defects in the gene encoding ABCA1. More than a 150 missense,
9q31. ABCA1 gene contains 50 exons and codes for a 2261 amino acid in- nonsense, frameshift ABCA1 mutations have been described in patients
tegral membrane transporter protein that mediates the export of cho- with familial HDL deficiency and Tangier disease. [http://www.hgmd.
lesterol and phospholipids to lipid poor lipoprotein. ABCA1 consists of cf.ac.uk/ac/gene.php?gene=ABCA1]. Severe HDL deficiency in a patient
two transmembrane domains, each formed by six alpha helices and was caused by two autosomal recessive mutations of ABCA1, one pre-
two intracellular nucleotide binding domains. Upregulation of ABCA1 dicted to lie in the transmembrane segment (V1704D) and the other
is dependent on the enhanced expression of PPARγ and LXRα [460]. (L1379F) in the extracellular loop. Both mutations prevent normal traf-
Identification of genetic defects in ABCA1 causing Tangier disease ficking of ABCA1, thereby explaining their inability to mediate apoAI
(OMIM 205400) has revealed a key role for ABCA1 in cholesterol trans- dependent lipid efflux [470]. The mutant W590S has a near-normal
port. Mutated ABCA1 fails to mediate cholesterol efflux to lipid poor apolipoprotein binding activity but defective lipid transport [471].
preβ HDL or its major apolipoprotein apoAI. However, the precise Brunham et al. [472] report three new kindreds with Tangier disease.
mechanisms by which ABCA1 mediates cholesterol efflux remain One patient was identified to be homozygous for a nonsense mutation
elusive. Recent studies have implicated several signaling mechanisms p. Q1038X, a further kindred showed compound heterozygosity for mis-
(eg protein kinase A (PKA), PKC) in ABCA1 function and apoAI lipidation sense variants p. R937V and p.T940M, and a third pedigree was com-
[461]. Several groups have reported that mutations in the ABCA1 gene pound heterozygous for a frameshift variant p.I1200Hfs*4 and an
are responsible for Tangier disease. Defective removal of cellular choles- intronic variant that lead to the creation of a cryptic splice site and pre-
terol and phospholipids and a deficiency in HDL-mediated efflux of mature truncation p. S1392Rfs*6. Cases of non classical Tangier disease
intracellular cholesterol has been demonstrated in Tangier disease fi- have been described. A patient who did not have any corneal opacities
broblasts [462]. or organomegaly on a CT scan of the abdomen was found to have
Tangier disease manifests with symptoms and signs resulting from three mutations in the ABCA1 gene. All three mutations p.A1046D, p.
the deposition of cholesteryl esters in nonadipose tissue; in peripheral Y1532C and p. W1699C were reported to be deleterious in functional
nerves leading to neuropathy and in reticulo-endothelial organs such studies [473].
as liver, spleen, lymph nodes and tonsils causing their enlargement
and discoloration. An association with early cardiovascular disease can 3.2.1.3. LCAT deficiency. The human LCAT gene encompasses 4.2 kb and is
be variable. The biochemical signs of Tangier disease are low plasma localized on chromosome16q21-22. LCAT is synthesized primarily by
HDL, apoAI, cholesterol and normal or high plasma TG. Hematological the liver and undergoes N- and O-linked glycosylation to become enzy-
findings in patients with Tangier disease are thrombocytopenia and matically active. The enzyme has both phospholipase A2 and an acyl-
stomatocytosis. Clinical signs combine differently in each patient. The transferase activity as it catalyzes the transacylation of the sn-2 fatty
clinical phenotype is inherited as an autosomal recessive trait and the acid of lecithin to the free 3-β hydroxyl group of cholesterol forming ly-
biochemical phenotype is inherited as an autosomal co-dominant trait solecithin and cholestery esters. The enzyme has an esterase activity on
[463]. ABCA1 deficiency impairs the efflux of free cholesterol from monomeric substrates. LCAT is active both on LDL and HDL and is pri-
cells, leading to intracellular accumulation of cholesteryl esters and marily associated with HDL which contains apoAI the major activator
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 167

of LCAT. Using site directed mutagenesis the catalytic triad of LCAT was and mesangial locations [483]. Genetic analysis of a further patient
shown to include S181, D345 and H377. Residues F103 and L182 were with nephrotic syndrome revealed two mutations in the LCAT gene:
thought to form an oxyanion hole and residues C50–C74 contributed (i) deletion of 5 nucleotides in exon 1, which caused a frameshift after
to enzyme-substrate interaction [474]. Familial LCAT deficiency (FLD, R23 with the introduction of 52 amino acids and a termination codon
OMIM 245900) is characterized by the absence of LCAT activity and re- at position 76 (ii) S181N amino acid substitution which converted the
duced HDL cholesterol in plasma. In FLD plasma LCAT is either absent or S181 amino acid which is a component of the catalytic triad conserved
completely lacks LCAT activity. In fish-eye disease (FED, OMIM 136120) in all examined animal species [484]. A patient with inhibitory anti-
the mutant LCAT lacks activity primarily on HDL lipids but preserves LCAT antibody presented with nephrotic syndrome and glomerular le-
LCAT activity on apoB containing lipoproteins. FED mutations specifical- sions similar to FLD. Treatment with steroids resulted in complete re-
ly affect phospholipase A2 activity on HDL, whereas esterase activity mission of the nephrotic syndrome with normalization of the serum
on the monomeric substrate and acyltransferase activity on LDL are LCAT activity and HDL level and disappearance of foam cell accumula-
retained. Some mutants have decreased acyltransferase on both LDL tion in renal tissue [485].
and HDL, though the acyltransferase activity on HDL decreased more Intriguingly, premature IHD is not a consistent feature of FLD or FED
than the acyltransferase activity on LDL. Classic FLD mutations impair and the role of LCAT mutations in atherosclerosis has been disputed. The
activity on all substrates. A structural basis for the difference between effects of LCAT deficiency on carotid artery intima-media thickness, a
FED and FLD phenotypes may depend on whether the mutations impair surrogate marker for IHD have been reported in several studies. In a
the catalytic triad, impair the accessibility to the catalytic residues or study by Hovingh and his colleagues [486] heterozygosity for the LCAT
affect the fold of the enzyme [475,476]. gene defects was associated with low HDL cholesterol levels, increased
Thirteen families carrying 17 different mutations were identified by levels of TG, C-reactive protein and increased intima-media thickness.
Calabresi et al. [477]. Seven of the ten probands carrying two mutant In contrast in the study by Calabresi and colleagues [487] carotid intima
LCAT alleles had FLD and three had FED, reflecting a milder clinical phe- thickness was decreased in both FLD and FED patients. Both studies
notype. All carriers of two mutant LCAT alleles had low plasma HDL cho- used carotid ultrasound to measure arterial wall thickness. A further
lesterol, with at least a 6 fold variation in HDL cholesterol level. The study which used magnetic resonance imaging to localize the carotid ar-
variability was unrelated with inherited defects in LCAT function as tery showed that carriers of LCAT gene mutations exhibited an in-
both FLD and FED had overlapping HDL cholesterol levels. Plasma total creased risk of carotid atherosclerosis [488]. Serum from carriers of
and LDL cholesterol levels in carriers of 2 mutant LCAT alleles showed LCAT deficiency has the same capacity as normal serum for ABCA1 me-
a wide interindividual variability. Elevated plasma TG were a fre- diated cholesterol efflux, probably due to increased content of preβ-
quent findings among cases and LCAT mutations had a gene-dose de- HDL [489]. LCAT deficiency has been shown be associated with defective
pendent effect on plasma TG and VLDL. The common clinical findings HDL antioxidant/anti-inflammatory properties [490] but, surprisingly,
of carriers of 2 mutant LCAT alleles were bilateral corneal opacity and was not associated with parameters of increased lipid peroxidation
normochromic anemia of varying severity. Six of the 15 carriers of 2 [491]. Population studies provided inconclusive data on the association
mutant LCAT alleles had end-stage renal failure requiring hemodial- between reduced levels of LCAT activity and cardiovascular risk [492].
ysis. In this study, 34/44 of heterozygotes were apparently healthy. Results have been equally contradictory in the experimental world
The low HDL levels in carriers of two mutant alleles of LCAT is associ- using animal models, as reviewed by Ng [493]. In animal models the
ated with alterations in HDL structure and particle distribution, presence of other pro-atherogenic apoB containing lipoproteins and ad-
which likely reflect the accumulation in plasma of cholesterol ester ditional proteins, CETP and LDLR, appear to play a role. Studies on
poor apoAI containing discoidal HDL which cannot mature into spheri- mouse models suggest that LCAT may modulate obesity and insulin sen-
cal HDL because of the lack of LCAT activity [478]. In homozygotes the sitizing pathways, though further studies on underlying mechanisms
majority of apoAI was found in small poorly lipidated small disk shaped are warranted [493,494]. Two case presentations further illustrate the
preβ1-HDL particles and some apoAI is found in larger, lipid poor discoi- paradoxical relationship between LCAT deficiency, low HDL levels and
dal HDL particles with α mobility. Heterozygotes had lipid profiles sim- cardiovascular risk. Genetic analysis identified two mutations Y83X,
ilar to controls except that apoCIII was found only in preβ-HDL [479]. In R99C in the LCAT gene in a patient with no trace of subclinical athero-
the absence of LCAT abnormal lipid particles have been observed sclerosis. The patient had corneal opacity, no cardiovascular symptoms
throughout lipoprotein fractions. Lipoprotein X (LpX) is an abnormal li- and low HDL [495]. A second patient with LCAT deficiency presented
poprotein that appears in cholestasis and LCAT deficiency. Electron mi- with corneal opacity, undetectable LCAT levels, and renal failure. He
croscopic studies suggest that LpX are spherical particles which contain had severe vascular disease, lower limb peripheral arterial obstruction
primarily free cholesterol and phoshatidylcholine [480]. Lipoprotein with necrosis of two left toes and angina with trivasal occlusive coro-
fractions specific to LCAT deficiency has been identified by high perfor- nary artery disease [496].
mance liquid chromatography with a gel filtration column. Four specific
fractions specific to LCAT deficiency corresponding to (i) large lipopro- 3.3. Increased HDL concentration
teins (N80 nm) (ii) large LDL (iii) small HDL to large LDL and (iv) small
HDL were identified. Large LDL and HDL fractions had compositions 3.3.1. Monogenic disorders associated with increased HDL levels
specific to FLD and not FED. The authors suggest that abnormal lipopro-
teins may play a causal role in renal pathology [481]. LCAT deficiency re- 3.3.1.1. CETP deficiency. The human CETP gene is mapped to chromosome
sult in low levels of HDL through preferential hypercatabolism of apoAII 16q12-21 near the LCAT gene, spans 25 kb pairs and encodes a 476
and apoAI and HDL particles compared to normal controls [482]. amino acid protein [497]. In humans CETP is expressed predominantly
Renal disease is a cause of mortality and morbidity in patients with in the liver, spleen and adipose tissue [498]. The evidence that CETP is
FLD. The kidney disease presents as proteinuria in its early stage and essential for HDL metabolism came from the discovery of CETP deficien-
increases in severity by the 4th and 5th decade in life, sometimes cy (OMIM 143470). Most cases of CETP mutations have been described
progressing to nephrotic syndrome and end stage renal disease. Homo- in Japanese families. Genetic CETP deficiency has been described as the
zygosity for the mutation p.W99R at position 99 of the LCAT protein was most frequent cause of hyperalphalipoproteinemia (HALP) in the
reported in a 33 year old female with a history of proteinuria. Renal bi- Japanese. Mutations described have been splice site mutations (muta-
opsy revealed mesangial expansion with the presence of mesangial tions in the 5′ splice donor site of intron 14 (Int14+1G→A), missense
foam cells and glomerular basement membrane thickening. Electron mutations (D442G) and nonsense mutations (G309X). Plasma CETP
microscopy confirmed the presence of marked basement membrane mass and activity were not detected in homozygous subjects with the
abnormalities, with the inclusion of lipid droplets in paramesangial Int14+1G→A, G181X, Q182X or G309X mutation. The Int14+1G→A
168 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

mutation causes exon 14 skipping which introduces premature termi- 3.3.1.2. Loss of function variants in EL. EL is coded by the LIPG gene is a
nation at codon 403. The truncated protein is rapidly degraded intracel- 68 kb glycoprotein secreted by endothelial cells. The amino acid se-
lularly in transfected COS-1 cells. The subjects with the heterozygous quence of EL is 41% homologous to HL. EL exhibits more phospholipase
Int14+1G→A mutation had about half the CETP mass of normal con- activity than TG lipase activity. EL hydrolyses HDL most efficiently of all
trols. Homozygous subjects with the Int14+1G→A mutation had mark- lipoprotein fractions [513]. A rare variant N396S in EL is associated with
edly elevated HDL cholesterol levels, while the heterozygous subjects elevated HDL cholesterol concentration and substantially reduced lipo-
had moderately increased HDL cholesterol levels [499]. A promoter mu- lytic activity [514]. The G26S N-terminal variant had significantly re-
tation (G to A substitution at the − 69 nucleotide of the promoter re- duced plasma levels of EL protein while its catalytic activity was
gion) is thought to affect transcription of the CETP gene. A reporter similar to wild type EL, leading to elevated HDL cholesterol levels
gene assay showed that this mutation markedly decreased transcription [515]. More work is required to determine whether the resulting in-
in HepG2 cells [500]. CETP deficiency contributes to approximately 50% crease in HDL cholesterol due to a decrease in EL activity will have an
of HALP in the Japanese population [501]. Only a few Caucasian individ- impact on cardiovascular disease [516]. An association has been report-
uals with CETP deficiency have been identified in case reports. A single ed between rare and common noncoding regulatory variants of the LIPG
base substitution resulting in a R268X codon in both alleles was identi- gene and HDL cholesterol levels [517].
fied in a Caucasian subject [502]. The serum of a patient with Swedish
ancestry and a homozygous mutation R37X in the CETP gene, showed 3.4. Polygenic causes of HDL variation
a normal or enhanced cholesterol efflux with via ABCG1/SR-BI, and de-
creased cholesterol efflux with ABCA1. The patient's plasma HDL was Numerous genes are suggested to play a role in HDL metabolism.
characterized by large preβ-HDL known to be better acceptors through Strategies for identifying complex polygenic traits that influence HDL
the former pathways [503]. Genetic CETP deficiency was rare among metabolism fall into (i) linkage analysis for the co-segregation of a chro-
subjects with HALP in The Netherlands. Heterozygous CETP deficiency mosomal region and a trait of interest using polymorphic markers in
was identified in only 1 of 95 HALP families. This genetic defect was families (ii) genetic association studies which measure an association
shown to cause skipping of exon7 and is predicted to result in a prema- between a genetic variant and HDL outcome (iii) GWAS which identify
ture truncation of the CETP protein [504]. To date several common poly- new genes that influence HDL levels. Eligible candidate genes from
morphisms and rare variants of the CETP which cause depletion of CETP studies were summarized by Boes et al. [518] and include the following
activity have been described across populations [505]. genes: CETP, LPL, ABCA1 and the APOA1C3A4A5 gene cluster. Other
One of these common polymorphisms is TaqIB, a silent base change candidate genes include GALNT2, which needs further functional char-
affecting the 277th nucleotide of the first intron of the CETP gene. The acterization. Weisglass-Volkov et al. [519] summarize the genetic poly-
B2 allele, absence of the TaqI restriction site, has been found to be asso- morphisms in (eg CETP, HL, LPL, EL, LCAT, ABCA1, APOAI, APOCIII,
ciated with elevated plasma HDL cholesterol. Since it is not clear how APOAV, APOE, SR-BI and PON1 (an antioxidative enzyme present on
this SNP can modulate CETP expression it has long been thought this the HDL particle)) genes described in more than one study as implicated
SNP is linked to another change that more directly affects CETP expres- in HDL cholesterol variation. Both an intronic SNP (rs2294213) [520]
sion. A large meta-analysis using a Mendelian randomization approach and missense mutation (R235W) [521] in the PLTP gene may contribute
showed that the B2 allele yielded a significant risk reduction in coronary to variation in plasma HDL. GWAS have provided empirical evidence for
artery disease compared to the B1 allele. However, the elevation in HDL polygenic inheritance of dyslipidemia with many genes, each with a
cholesterol levels due to TaqIB polymorphism was marginally associat- small effect, contributing to variation in lipid levels. However so far
ed with reduced coronary artery disease, challenging the conventional only about a third of the expected genetic heritability of lipid levels
view that raising HDL cholesterol levels may result into protection can be explained [522]. The explained heritability of lipids may increase
from IHD risk [506]. In a further meta-analysis three CETP genotypes as more common variants are identified. Studies suggest that very ex-
(TaqIB, I405V, −629CNA) that were associated with moderate inhibi- treme HDL cholesterol phenotypes can have a polygenic origin and
tion of CETP activity, and modestly high HDL cholesterol levels, show a that multiple rare variants can contribute to extreme HDL cholesterol
weak inverse association with coronary risk [507]. A definite relation- levels [523]. Gene-gene interactions between variants may further in-
ship between CETP and coronary artery disease is inconsistent and fluence HDL cholesterol concentration [524]. The effects of common
there is considerable dispute among genotype-phenotype studies variants on HDL cholesterol may be modified by lifestyle factors. Genetic
[508]. In one study, common CETP polymorphisms which beneficially factors may modify lipid responses to randomized lifestyle intervention
contributed to higher HDL cholesterol were paradoxically associated [525,526]. As a full collection of DNA sequence variants become avail-
with increased incidence of coronary disease [509]. The role of CETP able, an allelic risk score may allow for early identification of individuals
gene variations in longer life expectancy is complex and needs further at risk for cardiovascular disease and to better target preventive inter-
elucidation [510]. ventions [527].
Mechanisms responsible for the variations in coronary risk with in-
creased HDL cholesterol as a result of genetic variations in CETP remain 4. Secondary dyslipidemias
putative. It has been suggested that HDL cholesterol formed under con-
ditions of decreased CETP activity may be dysfunctional. Teh et al. [502] Secondary non-genetic factors associated with dyslipidemias include
report a raised HDL cholesterol, normal LDL cholesterol and mild obesity, metabolic syndrome, alcohol consumption, diabetes, renal
hypertriglyceridemia in a patient who was homozygous for the R268X disease, pregnancy (mainly in the third trimester), paraproteinemia, sys-
codon in the CETP gene. The patient's VLDL and LDL were cholesteryl temic lupus erythematosus and medications that include corticosteroids,
ester poor and triacylglycerol rich while her HDL was cholesteryl ester oral estrogen, tamoxifen, thiazides, non-cardioselective beta blockers,
rich and triacylglycerol poor. HDL particles from CETP deficient subjects bile acid sequestrants, cyclophosphamide, antiretroviral drugs and sec-
displayed normal anti-oxidative capacity [511] and decreased ability to ond generation antipsychotic reagents. People who develop secondary
protect endothelial cells from the development of endothelial dysfunc- dyslipidemia may have a subtle genetic defect that increases susceptibil-
tion [512]. It is likely that any increase in HDL cholesterol levels is offset ity to dyslipidemia.
by reduced apoAI mediated removal of cholesterol from macrophages.
The association between CETP deficiency and IHD can be influenced 4.1. Obesity, metabolic syndrome and diabetes
by functional abnormalities and changes in HDL and LDL, the genetic
mutation/polymorphism causing the CETP deficiency as well as envi- Depending on the degree of duration and weight gain, obesity (ele-
ronmental factors, obesity, alcohol consumption and smoking [499]. vated body mass index, particularly abdominal or upper body obesity)
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 169

can progressively cause and/or exacerbate a wide spectrum of morbid- derived FA. There was a significant similarity between the contributions
ities including dyslipidemia, metabolic syndrome and type II diabetes. of FA sources found in the liver TG pools and the sources used for TG se-
The typical dyslipidemia of obesity consists of raised fasting and post- cretion by the liver via lipoproteins [542]. In NAFLD cholesterol absorp-
prandial TG and free FA, decreased HDL cholesterol and normal or tion is decreased and synthesis is increased. These changes were
slightly increased LDL cholesterol with a predominance of small dense independent of BMI [543]. While metabolic adaptations occur to com-
LDL—the so called atherogenic triad. However, some individuals seem pensate for increased fat load, mitochondrial dysfunction eventually oc-
to be protected against the more deleterious metabolic effects of obesity curs. Mitochondrial functional decline is thought to further exacerbate
[528]. insulin resistance and lead to the progression of NAFLD [544].
Reactive oxygen species are generated in vivo under certain condi- Studies on people with obesity and insulin resistance are essentially
tions. Lipid peroxidation can occur through interactions of lipids and similar. In both fasting and postprandial states, obese subjects require
free radicals. Hyperlipidemia is also associated with oxidant stress insulin levels that are several times higher than non-obese subjects to
resulting in the formation of oxidized LDL, which contains many athero- maintain normal glucose tolerance [545]. Hepatic over production of
genic oxidized lipids [529]. VLDL is seen in both obesity and insulin resistant states. Alterations of
Abnormalities in lipoprotein are more pronounced in individuals insulin sensitive pathways, increased concentrations of free fatty acids
with central obesity [530]. Intra-abdominal or visceral adipose tissue and low grade inflammation all play a role and result in an overproduc-
as opposed to subcutaneous adipose tissue was found to be more relat- tion and decreased catabolism of TG rich lipoproteins of intestinal and
ed to atherogenic dyslipidemia. Studies on genetic determinants of obe- hepatic origin.
sity suggest that the heritability rates of body fats are approximately Plasma free FA are elevated in obese people and various FA are cyto-
50% [531]. toxic. Saturated FA, arachidonic acid and linoleic acid can stimulate the
Obese men also show a decreased uptake of dietary fat by adipose synthesis of pro-inflammatory cytokines like IL-1, IL-6 and TNF-α,
tissue, which results in higher delivery of CM remnants to the liver whereas eicosapentenoic acid (EPA), a fish oil, has anti-inflammatory
with enhanced VLDL-TG being delivered to peripheral adipocytes and properties. FA can regulate complex intracellular signaling systems
ectopic fat deposition [532]. Increased free FA flux to the liver in obesity thereby modulating cellular metabolism. Several transcription factors
leads to hepatic accumulation of TG and increased hepatic synthesis of including the PPAR and LXR families are linked in these pathways. Loca-
VLDL. In insulin resistant individuals the larger VLDL1 containing a tion in the body may be a factor in determining whether lipids promote
higher TG content is secreted in excess compared to the smaller, dense or suppress inflammation. Elevated plasma lipids may not be inflamma-
VLDL2 [533]. It has been suggested that should correct lipid storage in tory alone, as the hyperlipidemic state in obesity is indicative of redistri-
adipose tissue fail, free FA will be taken up by non-adipose tissue leading bution of lipids from adipose tissue to muscles and liver. The amount
to metabolic complications. Morbidly obese patients were shown to and the kind of FA can regulate cellular signaling and are reviewed in
have lower gene expression of factors related to lipid uptake and pro- detail elsewhere [546]. Obesity enhances both lipogenesis and adipo-
cessing in visceral adipose tissue. LPL, VLDL receptor, LRP1 and adipocyte genesis within fat depots and the secretion of pro-inflammatory
FABP (fatty acid binding protein) expression in visceral adipose tissue adipokines and chemokines into the plasma. Macrophages infiltration
was higher in lean controls rather than in obese patients. Adipose tissue of the adipose tissue has been described in obese conditions in humans.
secretes acylation stimulating protein (ASP), the result of a post transla- Both adipocytes and macrophages within fat secrete numerous cyto-
tional modification of the complement factor C3, which stimulates TG kines and hormones that contribute to local inflammation [547]. It can
synthesis. ASP expression was higher in the visceral adipose tissue of be hypothesized that the inciting event that causes adipose tissue to be-
lean controls than in obese patients [534]. A further hypothesis is that come inflamed may have a genetic component, to account for the differ-
the rate of lipolysis of TG rich lipoproteins can be partly regulated by ac- ences in metabolic risk between equivalently obese individuals [548].
cumulating free FA that dissociate LPL from the endothelial binding sites A reduction in sensitivity to insulin can occur through a genetic de-
[535]. Both overproduction and decreased fractional catabolic rate of fect or can be acquired as a consequence of obesity [549]. Insulin affects
VLDL was found to be related to hypertriglyceridemia in obese subjects. cells through binding to its insulin receptor on the surface of insulin re-
An inhibitory effect on LPL of the over production of VLDL apoCIII can sponsive cells. The stimulated insulin receptor phosphorylates itself and
further decrease LPL activity [536]. Lipolysis of TG rich lipoproteins is im- several substrates including members of the insulin receptor subsrate
paired in obesity leading to increased levels of postprandial and fasting (IRS) family, thus initiating downstream signaling [550]. Proinflamma-
levels of TG. The delayed clearance of TG rich lipoproteins facilitates tory cytokines like TNFα promotes inhibitory phosphorylation of serine
the CETP mediated exchange between cholesterol esters in HDL and residues of IRS-1. This phosphorylation inhibits tyrosine phosphoryla-
TG in VLDL. The CETP transfer process results in lower HDL cholesterol tion of IRS-1 in response to insulin and prevents downstream signaling
and indirectly reduces HDL size. In insulin resistant states a decreased and insulin action [551,552]. Additional pathways linking inflammation
post heparin LPL activity and increased HL is a determinant of low HDL to insulin resistance have been described and it is likely that other path-
cholesterol. The increased exchange facilitated by CETP produces LDL ways remain to be discovered [553]. Heat shock proteins appear to have
particles enriched in TG that are rapidly hydrolyzed by HL leaving small- a potential role in inhibiting inflammatory kinases [554]. In humans, in-
er denser LDL particles [537,538]. sulin resistance in muscle develops within hours of an acute increase in
The increased incidence of obesity is recognized as a major cause plasma free FA [555]. The mechanisms involving free FA induced skele-
of non-alcoholic fatty liver disease (NAFLD), the accumulation of tal muscle insulin resistance have not been fully elucidated [556]. It has
lipid droplets within hepatocytes. NAFLD can lead to non-alcoholic been suggested that FA induce insulin resistance in rat muscle through
steatohepatitis and hepatic cirrhosis. A number of factors contribute inhibition of insulin signaling which occurs through activation of a ser-
to the deposition of lipid droplets within the liver. Glucose and insulin ine kinase cascade involving PKC-θ [557]. Recent work in rats suggests
promote lipogenesis by activation of carbohydrate response element that oxidative stress plays a causal role in FA induced hepatic insulin re-
binding protein (ChREBP) and SREBP1c respectively. These are master sistance [558]. A further hypothesis is that ectopic lipid induced intra-
transcription factors that promote the transcription of lipogenic genes. cellular increase in diacyl glycerol results in the activation of PKC-θ in
ChREBP mediates the effects of carbohydrate (nutrients) on lipogenic muscle and PKC-ε in the liver and subsequent inhibition of insulin sig-
gene transcription [539,540]. Sanyal et al. [541] found elevated insulin naling in these tissues [559].
concentrations failed to suppress adipose fatty acid flux in NAFLD. Although free FA are required for normal insulin release chronic ex-
Liver TG in NAFLD was 59% from free FA, 26% from de novo lipogenesis posure to free FA both in vivo and in vitro is associated with marked im-
(DNL) and 15% from diet. During fasting the hepatic use of dietary fatty pairment in glucose stimulated β-cell insulin secretion and decrease in
acids became reduced and the deficit was made up by adipose tissue insulin gene expression. In subjects with a family history of type 2
170 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

diabetes, a sustained increase in plasma free FA impaired β-cell insulin SREBPs) located in close proximity [582]. Human LXR-α mRNA and pro-
response to glucose [560]. The molecular mechanisms by which exces- moter activity are positively regulated by thyroid hormone. Recent data
sive levels of FA and glucose alter β-cell function are beginning to be suggest a possibility of cross talk between TR and LXR [583]. Nuclear
unraveled but a number of questions remain [561]. hormone receptor crosstalk is becoming increasingly recognized as sig-
The definition of the metabolic syndrome links apparently unrelated nificant in thyroid hormone action. Thyroid hormones activate ChREB in
biological events into a single pathophysiological phenotype. The meta- liver and white adipose tissue. In mice TRβ but not TRα is required for
bolic syndrome has been proposed as a means of identifying people at activity in vivo. ChREB has been shown to be a LXR target. In mice
increased risk of cardiovascular disease and diabetes. The risk factors there is crosstalk between LXR and TRβ signaling on the ChREB promot-
for cardiovascular disease, which together are known as the metabolic er in the liver but not in white adipose tissue [584]. In chicken hepato-
syndrome are raised blood pressure, dyslipidemia (raised TG and cytes, thyroid hormone receptor and LXR play a role in fatty acid
lowered HDL), raised fasting glucose and central obesity. Three abnor- synthesis by regulating the expression of acetyl CoA carboxylase [585].
mal findings out of 5 would qualify for the metabolic syndrome [562]. Acetyl CoA carboxylase catalyzes the ATP dependent carboxylation of
Both obesity and type 2 diabetes are associated with insulin resis- acetyl CoA to malonyl-CoA which is the rate limiting step for the fatty
tance. Diabetic dyslipidemia is characterized by high TG levels, reduced acid synthesis pathway. The liver isoform of acetyl CoA carboxylase is
HDL and small dense LDL. Lipid abnormalities observed in type 2 dia- subject to nutritional and hormonal regulation. Feeding low fat, high
betics are quantitative, qualitative and kinetic in nature and exist before carbohydrate diet stimulates the enzyme 8–20 fold [585]. Among
the onset of diabetes as part of the insulin resistance syndrome. In type 2 other transcription factors SREBP-1c plays a role in controlling acetyl
diabetes a central paradox is the apparent selective nature of hepatic in- CoA carboxylase [586]. Carnitine palmitoyl transferase 1 (CPT-1) cata-
sulin resistance, where insulin fails to suppress hepatic glucose produc- lyzes the rate controlling step in the pathway of mitochondrial fatty
tion yet continues to stimulate lipogenesis. Selective insulin resistance acid oxidation. It has been shown that thyroid hormones can induce
can contribute to the clinical phenotype of hyperglycemia, dyslipidemia the expression of the liver isoform of CPT-1 [587].
and NAFLD. To address this paradox investigators have hypothesized The conversion of cholesterol to bile acids is required to maintain
the existence of a branch point in the insulin signaling cascade to sepa- cholesterol homeostasis. T3 plays a role in the regulation of cholesterol
rate insulin's anabolic signal into discrete pathways controlling gluco- 7-hydroxylase, the rate limiting step in bile acid synthesis [588]. Tran-
neogenesis and DNL [563,564,565]. Other studies suggest that the scriptional regulation of the ABCA1 gene involves multiple transcription
increased delivery of substrate FA to the liver can explain the insulin sites and binding of several different transcription factors to the ABCA1
stimulation of hepatic lipogenesis [566,567]. gene. A network of transcriptional controls leads to a balanced expres-
Compared to controls, patients with type 2 diabetes and insulin re- sion of ABCA1. Electrophoretic mobility assays demonstrate that
sistant subjects without diabetes had a higher production rate of intes- TR/RXR heterodimer is able to bind to the ABCA1 promoter. [589].
tinal apoB48 containing lipoproteins. Type 2 diabetics had a higher VLDL The alterations in lipid profile during hypothyroidism are variable. In
apoB100 production rate and decreased fractional catabolic rate com- one study HLP type 2A was the most common lipid abnormality in pa-
pared to controls [568]. Several mechanisms seem to be involved in tients with primary hypothyroidism and HLP type 2B the most common
the overproduction of VLDL and chylomicrons in type 2 diabetes. In- in secondary hypothyroidism. Of a group with hypothyroidism 8.6% had
creased expression of MTP [569] and reduced inhibition of hormone normal values, 56.3% had HLP type 2A, 33.6% HLP type 2B, and 1.5% HLP
sensitive lipase (HSL) by insulin leading to enhanced free FA flux to type 4 [590]. Serum cholesterol and LDL cholesterol are increased by ap-
the intestine and liver may further drive TG rich lipoprotein secretion proximately 30% in patients with overt hypothyroidism. Evidence main-
[570,571]. The several pathological factors responsible for increased ly suggests normal or slightly elevated levels HDL cholesterol resulting
TG rich lipoprotein secretion are reviewed by Xiao et al. [572] and in unfavorable total cholesterol/HDL ratios [591]. The effects of subclin-
Verges [571]. The mean LDL-cholesterol level was comparable or slight- ical hypothyroidism on serum lipid levels are less clear. Although some
ly elevated compared to non-diabetic controls in patients with type 2 studies have demonstrated that total cholesterol and LDL cholesterol are
diabetes [573]. In kinetic studies LDL fractional catabolic rate is de- elevated in subclinical hypothyroidism, others have not shown any
creased. The catabolic defect may be due to the glycosylation of LDL effect of subclinical hypothyroidism on lipid levels. Other factors eg in-
with reduced affinity for LDLR [574] or decreased LDLR expression sulin resistance may modify the effects of subclinical hypothyroidism
[575]. HDL microheterogeneity is compounded by type 2 diabetes. Gor- on serum lipid levels [592].
don et al. [576] investigated the alterations of HDL subclasses with type
2 diabetes and reported changes in phospholipid and protein composi- 4.3. Renal disease
tions in the larger HDL fractions. Alterations in HDL structure and com-
position have been associated with loss of potentially atheroprotective In nephrotic syndrome apoB containing lipoproteins, VLDL, IDL and
functions. LDL (hypertriglyceridemia and hypercholesterolemia) are elevated.
HDL is reported to be unchanged or reduced. In addition to these quan-
4.2. Hypothyroidism titative changes, the lipoprotein composition is markedly changed.
Patients with nephrotic syndrome showed a marked elevation in
Thyroid hormones influence diverse metabolic pathways in lipid lipoprotein(a) (Lp(a)) serum concentrations [593]. Most of the work re-
and glucose metabolism. Thyroid hormones influence the synthesis, garding the disturbed lipoprotein metabolism in nephrotic syndrome
mobilization and degradation of lipids. The causative mechanisms for comes from animal studies.
the cholesterol lowering effect of thyroid hormones consist of A study designed to examine the gene profile in the livers of ne-
(i) increased expression of SREBP2 [577] and a direct effect of 3,5,3′- phrotic rats suggested that genes involved in lipid metabolism were dif-
triiodo-L-thyronine (T3) on the promoter of the LDLR [578] (ii) in- ferentially regulated in nephrotic rats. Increased expression of hepatic
creased activity of HL, CETP and LCAT [579]. T3 has been shown to in- enzymes in biosynthetic pathways of cholesterol, FA and TG was associ-
crease apoAV mRNA and protein levels in hepatocytes [580]. ated with enhanced SREBP-1 expression at both transcriptional and
T3 negatively regulates the mouse SREBP-1c gene expression in the post-translational levels. Genes involved in FA oxidation were down
liver [581]. Thyroid hormone action is initiated by the binding of T3 to regulated in the nephrotic liver [594]. Further studies suggest that
receptors (TR), which bind to DNA sequences termed the T3 response upregulation of enzymes involved in FA synthesis is driven by sterol
elements. Thyroid hormones directly regulate SREBPs but also co- responsive pathways but not ChREBP dependent pathways [595].
regulate genes with thyroid hormone response element and sterol re- Clement et al. [596] found marked elevation of ANGPTL4 in nephrotic
sponse element (the nonpalindromic recognition DNA element for syndrome and demonstrated that the level correlated with serum TG
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 171

levels in humans and animals with nephrotic syndrome of diverse plasma TG during pregnancy can lead to the appearance of atherogenic
causes. An increase in the production of ANGPTL4 by extra renal organs dense LDL particles in a small group of women [612]. Patients with eleva-
in nephrotic syndrome was driven by an elevated plasma FFA/albumin tions in lipoprotein levels over the accepted range of adaptation may de-
ratio. Increase in circulating levels of ANGPTL4 has a causal role for the velop hyperlipoproteinemia in their later life. The percentage of women
pathogenesis of hypertriglyceridemia by inhibition of LPL in nephrotic who will translate to hyperlipoproteinemia remains unknown and
syndrome. In a series of in vivo and in vitro experiments the authors inherited causes may underlie the increase in lipoprotein levels [234].
demonstrated that by interacting with glomerular endothelial αvβ5 Studies in rat during gestation indicate that changes in adipose tis-
integrin, circulating ANGPTL4 can reduce proteinuria, indicating sue HSL/LPL mRNA levels could be a mechanism for adipose tissue
that an increase in circulating ANGPTL4 in nephrotic syndrome is a lipid breakdown during late gestation, suggesting transcriptional regu-
biological response designed to decrease proteinuria. Deficiencies lation of both enzymes during gestation [613]. One possible hypothesis
in LPL, HL, VLDL receptor, LDLR, LCAT, SR-BI, upregulation of CETP, is that increased HSL activity and lowered LPL activity in adipose tissue
hepatic acyl coenzyme A cholesterol acyltransferase, HMGCR, DGAT can contribute to changes in maternal lipoprotein profile during preg-
(diacyl glycerol acyl transferase) and Lp(a) contribute to nephrotic nancy. The increased lipolytic activity in adipose tissue during late ges-
dyslipidemia [597,598,599]. tation enhances the production of VLDL by the liver, which together
End stage renal disease is accompanied by a complex pattern of with decreased removal of VLDL from the circulation as a result of de-
alterations in lipoprotein metabolism and structure. A feature is an in- creased LPL activity in adipose tissue can augment circulating levels of
crease in serum TG due to elevated VLDL, IDL and low HDL. LDL choles- VLDL. During late pregnancy in rat, tissue LPL varies in different direc-
terol is often normal. Several compositional changes in lipoprotein have tions. It decreases in the adipose tissue, but increases in placental and
been shown to occur in end stage renal disease, including oxidation, mammary gland tissue. These changes may divert TG from uptake by
carbamylation and glycation of apoB. Modified lipoproteins are not eas- adipose tissue to uptake by the placenta for hydrolysis and transfer of
ily recognized by their receptors. An increase in half life in turn favors NEFA to the fetus [614].
modification, leading further to the accumulation of lipoproteins [600].
5. Treatment of dyslipidemias
4.4. Pregnancy
The possibility of secondary dyslipidemias needs to be considered
Women, with uncomplicated pregnancy, displayed increased total before initiating therapy. The mechanism of action of common choles-
cholesterol, LDL cholesterol, HDL cholesterol and lipoprotein(a), in the terol lowering drugs in routine use is summarized below.
second and third trimester. The prominent change was a 2–3 fold in- (i) Statins: Statins competitively inhibit HMGCR reductase activity.
crease in TG in the third trimester [601]. During pregnancy, values The reduction in intracellular cholesterol increases LDLR expression on
N95th percentile are considered supraphysiologal rise in plasma lipids the hepatocyte cell surface and increases extraction of LDL cholesterol
[602]. Hyperphagia and increased lipid synthesis contribute to maternal from blood.
fat accumulation, typically associated with the first two thirds of the (ii) Bile acid sequestrants: By binding to bile acids the drugs remove
gestation period. Changes in lipid metabolism promote the accumula- the bile acids from the enterohepatic circulation. The liver depleted of
tion of maternal fat stores in early and mid pregnancy and increase fat bile synthesizes more from the hepatic stores of cholesterol. The in-
mobilization during late pregnancy. In addition fat cells from different crease in cholesterol catabolism leads to a compensatory increase in
regions of the body show a differential response to pregnancy [603]. LDLR activity, and to increased clearance of LDL cholesterol from the
During late gestation maternal lipid metabolism changes to a catabolic circulation.
condition which promotes the use of lipid as the energy source. Nondi- (iii) Cholesterol absorption inhibitors: By inhibiting intestinal cho-
abetic pregnant women develop insulin resistance in the third trimester lesterol at the level of the brush border of the intestine (most probably
of pregnancy. Insulin resistance is more marked in gestational diabetes by interacting with the NPC1L1 protein) they reduce the amount of cho-
[604]. The higher concentration of estrogen and insulin resistance are lesterol circulated to the liver.
thought to be responsible for hypertriglyceridemia of pregnancy. The (iv) Niacin has a broad lipid modulating action, it raises HDL choles-
mechanism by which estrogen and insulin interact is unknown. Recent terol and reduces both LDL cholesterol and TG. There are several mech-
studies suggest that in hyperestrogenemic syndromes, estrogen may in- anism by which niacin may affect plasma lipoproteins. Recently the
terfere with insulin binding to its receptor [605] or decrease GLUT4 GPR109A receptor also named the hydroxyl-carboxylic acid receptor
(Glucose transporter type4) expression in muscle [606] and as a result 2, expressed in adipocytes and immune cells, that binds niacin has
decrease insulin sensitivity. been identified. Activation of GPR109A in adipocytes results in the in-
In the last trimester increased adipose tissue lipolysis results in the re- hibitory G protein, Gi mediated reduction in adenylate cyclase, decreas-
lease of NEFA (nonesterified fatty acids). The primary destination of NEFA ing cAMP accumulation and protein kinase A activity. This results in
is the liver, where following reesterification for the synthesis of TG they decreased phosphorylation of HSL and TG hydrolysis, decreased FA
are released as VLDL. Both LPL and HL activities in postheparin plasma flux to the liver and reduced substrate availability for hepatic VLDL syn-
have been shown to be decreased in pregnancy [607]. Total cholesterol thesis. Niacin noncompetitively inhibits hepatocyte DGAT resulting in
and VLDL, LDL and HDL cholesterol increased with gestational time. decreased TG synthesis. It has been suggested that there is an accompa-
Total TG and VLDL, LDL and HDL TG followed a similar trend. VLDL TG/ nying decrease in CETP mediated exchange of TG for cholesteryl ester
cholesterol ratio declined in the third trimester whereas in both LDL between VLDL and HDL particles leading to a net rise in HDL particles.
and HDL the TG/cholesterol ratio but not the cholesterol/phospholipid A further suggestion is that niacin promotes cholesterol efflux from ad-
ratio increased during gestation [608]. CETP is highest in the second tri- ipocytes to apoAI by activation of the PPARγ–LXRα–ABCA1 pathway.
mester of gestation and declined at the third trimester. HDL TG increased Niacin is thought to exert pleiotropic effects which are lipoprotein inde-
during the second and third trimester. CETP activity correlated with HDL pendent through direct anti-inflammatory effects on cell types involved
TG levels [609]. Changes in CETP activity can be one of the factors that de- in the progression of atherosclerosis [615,616,617].
termine LDL and HDL TG levels. Increased transfer of TG to HDL is due to Both fibrates and n-3 polyunsaturated fatty acids are the available
reduced LPL activity resulting in higher levels of VLDL and increased pharmacological interventions to decrease elevated plasma TG levels.
length of exposure to CETP. [610]. The reduction in HL can decrease the Fibrates are agonists of PPARα and act through various transcription
conversion of TG containing HDL2 to the lipid poor HDL3 particles and in- factors to regulate lipoprotein metabolism to lower fasting and post-
crements in HDL2 rather than HDL3 seem responsible for the increase in prandial TG levels. A summary of PPARα mediated gene regulation by
HDL cholesterol level seen during gestation in women. [611]. Increase in fibrates is shown in Fig. 8. [618].
172 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

Fig. 8. Mechanism of action of fibrates.

Fish contain high levels of two omega-3 FA, EPA (C20:5 n-3) and (ASO) and small interfering RNAs (siRNA). Phases I, II and III reveal the
docosahexanoic acid (DHA; C22:6 n-3). The effects of n-3 fatty acids are efficacy of monoclonal PCSK9 inhibitors to reduce high LDL levels [624].
now well-established. Total cholesterol is not materially affected by n-3 Mipomersen is an apoB targeted ASO tested against humans. It is ap-
fatty acid consumption, LDL cholesterol concentrations tend to rise by proved by the US FDA for use in homozygous FH patients as an adjunct
5–10% and HDL cholesterol by 1–3%, and serum TG concentrations de- to diet and statins. The most studied MTP inhibitor, lomitapide has been
crease by 25–30% [619]. The effects are more likely to be due to decreased granted FDA approval as an orphan drug to reduce LDL cholesterol in
hepatic secretion of VLDL. This effect is probably due to multiple mecha- patients with homozygous FH. Other MTP inhibitors have been studied
nisms: omega-3 FA may (i) not be the preferred substrate for the enzyme in experimental animals. Both lomitapide and mipomersen are associat-
DGAT and thus decrease TG synthesis (ii) interact with nuclear transcrip- ed with side effects, such as hepatic steatosis and injection site reactions
tion factors (PPARα, LXR and farnesol X receptor, HNF-4α) and decrease with mipomersen, that do not allow extended clinical use. Newer forms
TG synthesis and increase fatty acid oxidation and (iii) promote apoB deg- of PPARα/γ agonists, RNA based treatments that target apoCIII, and in-
radation in the liver through stimulation of an autophagic process (iv) in- hibitors of IDOL are also in development [625].
duce LPL activity and increase CM and VLDL clearance [620,621]. The available evidence for elevating low HDL cholesterol levels are
Statins today are the treatment of choice for lowering LDL cholester- relatively few. The most efficacious pharmacological approach to eleva-
ol. Patients who fail to achieve adequate reduction of LDL cholesterol tion of low HDL cholesterol has involved inhibition of CETP. Inhibiting
can be considered statin resistant. Resistance to statins can be due to dif- CETP led to a substantial increase in HDL cholesterol but this has not
ferences in pharmacokinetics as well as in target pathways. Target path- translated to beneficial effects on cardiovascular outcomes. The devel-
ways include HMGCR and various points along the cholesterol synthesis opment of torcetrapib, the first CETP inhibitor was halted after the
and lipoprotein metabolic pathways [622]. There are various genetic Investigation of Lipid Level Mangement to Understand its Impact in
variations that are associated with diversity in the drugs response Atherosclerotic Events (ILLUMINATE) trial revealed an increase in
[623]. In addition high risk patients do not always achieve sufficient re- cardiovascular events in the torcetrapib treated group. It has been sug-
duction on LDL cholesterol, even with high intensity statin treatment. gested that efforts should focus on improving HDL quality in addition to
New novel LDL cholesterol lowering drugs are in development to en- raising HDL levels when developing HDL raising therapies. Other novel
hance the efficacy of statins. These approaches are based on the inhibi- therapies aimed at raising HDL include apoAI mimetics, reconstituted
tion of PCSK9, apoB or MTP. HDL, PPARα/γ agonists and LXR agonists [626].
Inhibition of PCSK9 may be achieved by a number of approaches, in- Somatic cell gene therapy is a promising approach for the treatment
cluding the use of monoclonal antibodies (mAb), antisense nucleotides of several dyslipidemias [627]. In a clinical trial using adeno-associated
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 173

virus serotype 1 (AAV-1) mediated gene transfer of the naturally occur- [16] F.M. Sacks, C. Zheng, J.S. Cohn, Complexities of plasma apolipoprotein C-III metab-
olism, J. Lipid Res. 52 (2011) 1067–1070.
ring gain of function LPL mutant S447X, to muscle in subjects with LPL [17] J. Chen, Q. Li, J. Wang, Topology of human lipoprotein E3 uniquely regulates its
deficiency, median plasma TG decreased initially. However, long term diverse biological functions, Proc. Natl. Acad. Sci. 108 (2011) 14813–14818.
follow up of TG showed a return to baseline levels. No subjects devel- [18] H. Saito, P. Dhanasekaran, D. Nguyen, P. Holvoet, S. Lund-Katz, M.C. Phillips, Do-
main structure and lipid interaction in human apolipoproteins A-I and E, a general
oped an immune response to the LPL transgene product though 4 of model, J. Biol. Chem. 278 (2003) 23227–23232.
the 8 subjects developed a T-cell response to the vector capsid. Potential [19] J.I. Goldberg, C.A. Scheraldi, L.K. Yacoub, U. Saxena, C.L. Bisgaier, Lipoprotein ApoC-
solutions, such as immunomodulation at the time of gene transfer may II activation of lipoprotein lipase. Modulation by apolipoprotein A-IV, J. Biol. Chem.
265 (1990) 4266–4272.
need to be explored to achieve sustained persistent expression of the [20] G. Lippi, M. Franchini, G. Targher, Screening and therapeutic management of
transgene product [628]. lipoprotein(a) excess: review of the epidemiological evidence, guidelines and rec-
ommendations, Clin. Chim. Acta 412 (2011) 797–801.
[21] K. Ren, Z.L. Tang, Y. Jiang, Y.M. Tan, G.H. Yi, Apolipoprotein M, Clin. Chim. Acta 446
6. Future directions (2015) 21–29.
[22] S. Nielsen, F. Karpe, Determinants of VLDL-triglycerides production, Curr. Opin.
Severe dyslipidemia, a decrease or increase in serum lipoproteins is Lipidol. 23 (2012) 321–326.
[23] L. Hodson, B.A. Fielding, Trafficking and partitioning of fatty acids: the transition
usually a simple monogenic disorder showing a Mendelian inheritance. from fasted to fed state, Clin. Lipidol. 5 (2010) 131–144.
More frequently dyslipidemia results from the cumulative effects of [24] G.F. Lewis, K.D. Uffelman, L.W. Szeto, B. Weller, G. Steiner, Interaction between free
several genes combined with environmental and metabolic stressors fatty acids and insulin in the acute control of very low density lipoprotein produc-
tion in humans, J. Clin. Invest. 95 (1994) 158–166.
such as high calorie intake, metabolic syndrome, diabetes and certain [25] A.M. Brown, G.F. Gibbons, Insulin inhibits the maturation phase of VLDL assembly
drugs. Despite our increased understanding of molecular basis of via a phosphoinositide 3 kinase-mediated event, Arterioscler. Thromb. Vasc. Biol.
dyslipidemias, much remains to be learned. Genetic profiles are far 32 (2012) 2104–2112.
[26] L. Andersson, P. Bostrom, J. Ericson, M. Rutberg, B. Magnusson, D. Marchean, M.
from complete and there is further room for characteristation of genetic Ruiz, L. Asp, P. Huang, M.A. Frohman, J. Boren, S. Olofsson, PLD1 and ERK2 regulate
changes influencing serum lipid levels. Emerging technologies such as cytosolic lipid droplet formation, J. Cell Sci. 119 (2006) 2246–2257.
whole genome sequencing may produce more discriminating ‘risk al- [27] W. Au, H. Kung, M.C. Lin, Regulation of microsomal triglyceride transfer protein
gene by insulin in HepG2 cells, Diabetes 52 (2003) 1073–1080.
lele’ scores. Although a plethora of antidyslipidemic pharamacological
[28] A. Kamagate, S. Qu, G. Perdomo, D. Su, D.H. Kim, S. Slusher, M. Meseck, H.H. Dong,
agents are available, they can be ineffective in the more severe lipid dis- Fox01 mediates insulin-dependent regulation of hepatic VLDL production in mice,
orders. The results of existing pharmacological treatments must be J. Clin. Invest. 118 (2008) 2347–2364.
[29] P. Dhawan, A. Bell, A. Kumar, C. Golden, K.D. Mehta, Critical role of p42/44MAPK ac-
viewed alongside newer treatments that are being tested. Future direc-
tivation in anisomycin and hepatocyte growth factor-induced LDLR expression:
tions will be to define the missing genetic variation which will open activation of Raf-1/MEK-1/p42/44MAPK cascade alone is sufficient to induce LDLR
possibilities for targeted individualized treatment for dyslipidemias. expression, J. Lipid Res. 40 (1999) 1911–1919.
[30] J. Kotzka, D. Muller-Wieland, G. Roth, L. Kremer, M. Munck, S. Schurman, B. Knebel,
W. Krone, Sterol regulatory element binding proteins (SREBP)-1a and SREBP-2 are
References linked to the MAP-kinase cascade, J. Lipid Res. 41 (2000) 99–108.
[31] A.B. Keeton, M.O. Amsler, D.Y. Venable, J.L. Messina, Insulin signal transduction
[1] A. Steinmetz, H. Kaffarnik, G. Utermann, Activation of phosphatidylcholine-sterol pathways and insulin-induced gene expression, J. Biol. Chem. 277 (2002)
acyltransferase by human apolipoprotein E isoforms, Eur. J. Biochem. 152 (1985) 48565–48573.
747–751. [32] H. Duez, B. Lamarche, R. Valero, M. Pavlic, S. Proctor, C. Xiao, L. Szeto, B.W.
[2] S. Tiwari, S.A. Siddiqi, Intracellular trafficking and secretion of VLDL, Arterioscler. Patterson, G.F. Lewis, Both intestinal and hepatic lipoprotein production are stim-
Thromb. Vasc. Biol. 32 (2012) 1079–1086. ulated by an acute elevation of plasma free fatty acids in humans, Circulation 117
[3] I. Ramasamy, Recent advances in physiological lipoprotein metabolism, Clin. Chem. (2008) 2369–2376.
Lab. Med. 52 (2014) 1695–1727. [33] M. Pavlic, C. Xiao, L. Szeto, B.W. Patterson, G.F. Lewis, Insulin acutely inhibits intes-
[4] J.P. Segrest, M.K. Jones, H. De Loof, C.G. Brouillette, Y.V. Venkatachalapathi, G.M. tinal lipoprotein secretion in humans in part by suppressing plasma free fatty
Anantharamaiah, The amphipathic helix in the exchangeable apolipoproteins: a acids, Diabetes 59 (2010) 580–587.
review of secondary structure and function, J. Lipid Res. 33 (1992) 141–166. [34] C. Longuet, E.M. Sinclair, A. Maida, L.L. Baggio, M. Maziarz, M.J. Charron, D.J.
[5] M. Sundaram, Z. Yao, Intrahepatic role of exchangeable apolipoproteins in lipoprotein Drucker, The glucagon receptor is required for the adaptive metabolic response
assembly and secretion, Arterioscler. Thromb. Vasc. Biol. 32 (2012) 1073–1078. to fasting, Cell Metab. 8 (2008) 359–371.
[6] J. Delgado-Lista, F. Perez-Jimanez, J. Ruano, P. Perez-Martinez, F. Fuentes, J. Criado- [35] C. Xiao, M. Pavlic, L. Szeto, B.W. Patterson, G.F. Lewis, Effect of hyperglucagonemia
Garcia, L.D. Parnell, A. Garcia-Rios, J.M. Ordovas, J. Lopez-Miranda, Affects of varia- on hepatic and intestinal lipoprotein production and clearance in healthy humans,
tions in the APOA1/C3/A4/A5 gene cluster on different parameters of postprandial Diabetes 60 (2011) 383–390.
lipid metabolism in helathy young men, J. Lipid Res. 53 (2010) 73-73. [36] J.P. Pegorier, C. Le May, J. Girard, Control of gene expression by fatty acids, J. Nutr.
[7] R.S. Rosenson, H.B. Brewer, M.J. Chapman, S. Fazio, M.M. Hussain, et al., HDL mea- 134 (2004) 2444–2449.
sures, particle heterogeneity, proposed nomenclature, and relation to atheroscle- [37] E.E. Kershaw, M. Schupp, H. Guan, N.P. Gardner, M.A. Lazar, J.S. Flier, PPARγ regu-
rotic cardiovascular events, Clin. Chem. 57 (2011) 392–410. lates adipose triglyceride lipase in adipocytes in vitro and in vivo, Am. J. Physiol.
[8] T.J. Kalogeris, K. Fukagawa, P. Tso, Synthesis and lymphatic transport of intestinal Endocrinol. Metab. 293 (2007) 1736–1745.
apolipoprotein A-IV in response to graded doses of triglyceride, J. Lipid Res. 35 [38] C.H. Lee, P. Olson, R.M. Evans, Minireview: lipid metabolism, metabolic diseases,
(1994) 1141–1151. and peroxisomee proliferator-activated receptors, Endocrinology 144 (2003)
[9] S. Lu, Y. Yao, S. Meng, X. Cheng, D.D. Black, Overexpression of apolipoprotein A-IV 2201–2207.
enhances lipid transport in newborn swine intestinal epithelial cells, J. Biol. Chem. [39] M. Rakhshandehroo, G. Hooiveld, M. Muller, S. Kersten, Comparative analysis of
277 (2002) 31929–31937. gene regulation by the transcription factor PPARα between mouse and human,
[10] M. Merkel, J. Heeren, Give me A5 for lipoprotein hydrolysis, J. Clin. Invest. 115 PLoS One 4 (2009), e6796.
(2005) 2694–2696. [40] Y. Omori, J. Imai, T. Watanabe, T. Komatsu, K. Suzuki, S. Kataoka, A. Watanabe, A.
[11] A.M. Blade, M.A. Fabritus, H. Li, R.B. Weinberg, G.S. Shelness, Biogenesis of apolipo- Tanigami, S. Sugano, CREB-H: a novel mammalian transcription factor belonging
proteins A-V and its impact on VLDL triglyceride secretion, J. Lipid Res. 52 (2011) to the CREB/ATF family and functioning via the box-B element with liver specific
237–244. expression, Nucleic Acids Res. 33 (2001) 1859–1873.
[12] N.S. Shachter, Apolipoproteins C-I and C-III as important modulators of lipoprotein [41] P. Tontonoz, Transcriptional and Posttranscriptional Control of Cholesterol Homeo-
metabolism, Curr. Opin. Lipidol. 2 (2001) 297–304. stasis by Liver X Receptors, 76Cold Spring Harb. Lab. Press, 2011 129–138.
[13] J.P. de Barros, A. Boualam, T. Gautier, L. Dumont, B. Veges, D. Masson, L. Lagrost, [42] H. Basciano, A. Miller, C. Baker, M. Naples, K. Adeli, LXRα activation perturbs
Apolipoprotein CI is a physiological regulator of cholesteryl ester transfer protein hepatic insulin signaling and stimulates production of apolipoprotein B-
activity in human plasma but not in rabbit plasma, J. Lipid Res. 50 (2009) containing lipoproteins, Am. J. Physiol. Gastrointest. Liver Physiol. 297 (2009)
1842–1851. C323–C332.
[14] H.N. Ginsberg, N.A. Le, I.J. Goldberg, J.C. Gibson, A. Rubinstein, P. Wang-Iverson, R. [43] W. Qiu, R.K. Avramoglu, N. Dube, T.M. Chong, M. Naples, C. Au, K.G. Sidiropoulos,
Norum, W.V. Brown, Apolipoprotein B metabolism in subjects with deficiency of G.F. Lewis, J.S. Cohn, M.L. Tremblay, K. Adeli, Hepatic PTB-1B expression regulates
apolipoproteins CIII and AI. Evidence that apolipoprotein CIII inhibits catabolism the assembly and secretion of apolipoprotein B containing lipoproteins, Diabetes
of triglyceride-rich lipoproteins by lipoprotein lipase in vivo, J. Clin. Invest. 78 53 (2004) 3057–3066.
(1986) 1287–1295. [44] W. Qiu, L. Federico, M. Naples, R.K. Avramoglu, R. Meshkani, J. Zhang, J. Tsai, M.
[15] M. Sundaram, S. Zhnong, M.B. Khalil, H. Zhou, Z.G. Jiang, Y. Zhao, J. Iqbal, M.M. Hussain, K. Dai, J. Iqbal, C.D. Kontos, Y. Horie, A. Suzuki, K. Adeli, Phosphatase
Hussain, D. Figeys, Y. Wang, Z. Yao, Functional analysis of the missense APOC3 mu- and tensin homolog (PTEN) regulates hepatic lipogenesis, microsomal triglyceride
tation Ala23Thr associated with human hypotriglyceridemia, J. Lipid Res. 51 transfer protein and the secretion of apolipoprotein B-containing lipoproteins,
(2010) 1524–1534. Hepatology 48 (2008) 1799–1809.
174 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

[45] J.L. Goldstein, M.S. Brown, The LDL receptor, Arterioscler. Thromb. Vasc. Biol. 29 pathogenic amino acid variant in the same family, Atherosclerosis 174 (2004)
(2009) 431–438. 67–71.
[46] M. Motamed, Y. Zhang, M.L. Wang, J. Seeman, H.J. Kwon, J.L. Goldstein, M.S. Brown, [72] C.G. Davis, I.R. van Driel, D.W. Russell, M.S. Brown, J.L. Goldstein, The low density
Identification of the luminal loop 1 of SCAP protein as the sterol sensor that main- lipoprotein receptor, J. Biol. Chem. 262 (1987) 4075–4082.
tains cholesterol homeostasis, J. Biol. Chem. 286 (2011) 18002–18012. [73] P.J. Talmud, S. Drenos, S. Shah, et al., for the ASCOT investigators, NORDIL investi-
[47] D.B. Jump, D. Botolin, Y. Wang, J. Xu, B. Christian, O. Demeure, Fatty acid regulation gators and BRIGHT consortium, Gene-centric association signals for lipids and apo-
of hepatic gene transcription, J. Nutr. 135 (2005) 2503–2506. lipoproteins identified via the humanCVD BeadChip, Am. J. Hum. Genet. 85 (2009)
[48] J.D. Horton, N.A. Shah, J.A. Warrington, N.N. Anderson, S.W. Park, M.S. Brown, J.L. 628–642.
Goldstein, Combined analysis of oligonucleotide microarray data from transgenic [74] M.S. Sandhu, D.M. Waterworth, S.L. Debenham, E. Wheeler, K. Papadakis, K. Song,
and knockout mice identifies direct SREBP target genes, Proc. Natl. Acad. Sci. 100 X. Yuan, T. Johnson, S. Ashford, M. Inouye, et al., LDL-cholesterol concentrations:
(2003) 12027–12032. a genome wide association study, Lancet 371 (2008) 483–491.
[49] J.D. Horton, I. Shimomura, M.S. Brown, R.E. Hammer, J.L. Goldstein, H. Shimano, Ac- [75] S. Kathiresan, O. Melander, C. Guiducci, A. Surti, N.P. Burtt, M.J. Rieder, G.M. Cooper,
tivation of cholesterol synthesis in preference to fatty acid synthesis in liver and et al., Six new loci associated with blood low-density lipoprotein cholesterol, high-
adipose tissue of transgenic mice overproducing sterol regulatory element- density lipoprotein cholesterol or triglyceride in humans, Nat. Genet. 40 (2008)
binding protein-2, J. Clin. Invest. 101 (1998) 2331–2339. 189–197.
[50] J. Stamler, M.L. Daviglus, D.B. Garside, A.A.R. Dyer, P. Greenland, J.D. Neaton, Rela- [76] S. Sanna, B. Li, A. Mulas, C. Sidore, H.M. Kang, A.U. Jackson, M.G. Piras, G. Usala, G.
tionship of baseline serum cholesterol levels in 3 large cohorts of younger men Maninchedda, A. Sassu, F. Serra, M.A. Palmas, W.H. Wood 3rd, I. Njølstad, M.
to long-term coronary, cardiovascular, and all-cause mortality and to longevity, Laakso, K. Hveem, J. Tuomilehto, T.A. Lakka, R. Rauramaa, M. Boehnke, F. Cucca,
JAMA 284 (2000) 311–318. M. Uda, D. Schlessinger, R. Nagaraja, G.R. Abecasis, Fine mapping of five loci asso-
[51] D.S. Frederickson, An international classification of hyperlipidemias and ciated with low-density lipoprotein cholesterol detects variants that double the
hyperlipoproteinemias, Ann. Intern. Med. 75 (1971) 471–472. explained heritability, PLoS Genet. 7 (7) (2011 Jul), e1002198.
[52] R.A. Hegele, Plasma lipoproteins: genetic influences and clinical implications, Nat. [77] P.J. Talmud, S. Shah, R. Whittall, M. Futema, P. Howard, J.A. Cooper, S.C. Harrison, K.
Rev. Genet. 10 (2009) 109–121. Li, F. Drenos, F. Karpe, H.A.W. Neil, O.S. Descamps, C. Langenberg, N. Lench, M.
[53] O.S. Descamps, J.P. Gilbeau, X. Leysen, F. Van Leuven, F.R. Heller, Impact of genetic Kivimaki, J. Whittaker, A.D. Hingorani, M. Kumari, S.E. Humphries, Use of low-
defects on atherosclerosis in patients suspected of familial hypercholesterolemia, density lipoprotein cholesterol gene score to distinguish patients with polygenic
Eur. J. Clin. Investig. 31 (2001) 958–965. and monogenic familial hypercholesterolemia: a case–control study, Lancet 381
[54] S.W. Fouchier, J.J.P. Kastelein, J.C. Defesche, Update of the molecular basis of fa- (2013) 1293–1301.
milial hypercholesterolemia in The Netherlands, Hum. Mutat. 26 (2005) [78] J. Yamaguchi, D.M. Conlon, J.J. Liang, E.A. Fisher, H.N. Ginsberg, Translocation effi-
550–556. ciency of apolipoprotein B is determined by the presence of β-sheet domains,
[55] L. Palacios, L. Grandoso, N. Cuevas, E. Olano-Martin, A. Martinez, D. Tejedor, M. Stef, not pause transfer sequences, J. Biol. Chem. 281 (2006) 27063–27071.
Molecular characterisation of familial hypercholesterolemia in Spain, Atherosclero- [79] J. Boren, U. Ekstrom, B. Agren, P. Nilsson-Ehle, T.L. Innerarity, The molecular mech-
sis 221 (2012) 137–142. anism for the genetic disorder familial defective apolipoprotein B100, J. Biol. Chem.
[56] S. Fazio, M.F. Linton, When it looks like familial hypercholesterolemia but is not, 276 (2001) 9214–9218.
Circ. Cardiovasc. Genet. 5 (2012) 599–601. [80] J.E. Chatterton, M.L. Phillips, L.K. Curtiss, R. Milne, J.C. Fruchart, V.N. Schumaker,
[57] A.K. Soutar, R.P. Naoumova, L.M. Traub, Genetics clinical phenotype and molecular Immunoelectron microscopy of low density lipoproteins yields a ribbon and bow
cell biology of autosomal recessive hypercholesterolemia, Arterioscler. Thromb. model for the conformation of apolipoprotein B on the lipoprotein surface, J.
Vasc. Biol. 23 (2003) 1963–1970. Lipid Res. 36 (1995) 2027–2037.
[58] S.N. Pimstone, J.C. Defesche, S.M. Clee, H.D. Bakker, M.R. Hayden, J.J.P. Kastelein, [81] A. Wang, D. Liu, C.C. Liang, Regulation of human apolipoprotein B gene expression
Differences in the phenotype between children with familial defective apo B-100 at multiple levels, Exp. Cell Res. 290 (2003) 1–12.
and familial hypercholesterolemia, Arterioscler. Thromb. Vasc. Biol. 17 (1997) [82] A.J. Hooper, F.M. van Bockxmeer, J.R. Burnett, Monogenic hypocholesterolaemic
826–833. lipid disorders and apolipoprotein B metabolism, Crit. Rev. Clin. Lab. Sci. 42
[59] A.C. Goldberg, P.N. Hopkins, P.P. Toth, C.M. Ballantyne, D.J. Rader, J.G. Robinson, S.R. (2005) 515–545.
Daniels, S.S. Gidding, S.D. de Ferranti, M.K. Ito, M.P. McGowan, P.M. Moriarty, W.C. [83] T.L. Innearity, K.H. Weisgraber, K.S. Arnold, R.W. Mahley, R.M. Krauss, G.L. Vega,
Cromwell, J.L. Ross, P.E. Ziaska, Familial hypercholesterolemia: screening diagnosis S.M. Grundy, Familial defective apolipoprotein B-100: low density lipopropteins
and management of pediatric and adult patients: clinical guidance from the Na- with abnormal receptor binding, Proc. Natl. Acad. Sci. 84 (1987) 6919–6923.
tional Lipid Association Panel on Familial Hypercholesterolemia, J. Clin. Lipidol. 5 [84] N.B. Myant, S.A. Forbes, I.N. Day, J. Gallagher, Estimation of the age of the ancestral
(2011) 133–140. arginine3500 glutamine mutation in apoB-100, Genomics 45 (1997) 78–87.
[60] S. Bertolini, L. Pisciotta, C. Rabacchi, A.B. Cefalu, D. Noto, T. Fasano, A. Signori, R. [85] G. Rauh, H. Schuster, C.K. Schewe, G. Stratmann, C. Keller, G. Wolfram, N. Zollner,
Fresa, M. Averna, S. Calandra, Spectrum of mutations and phenotypic expression Independent mutation of arginine(3500) glutamine associated with defective apo-
in patients with autosomal dominant hypercholesterolemia identified in Italy, Ath- lipoprotein B-100, J. Lipid Res. 34 (1993) 799–805.
erosclerosis 227 (2013) 342–348. [86] C.R. Pullinger, L.K. Hennessey, J.E. Chatterton, W. Liu, J.A. Love, C.M. Mendel, P.H.
[61] M. Marduel, A. Carrie, A. Sassolas, M. Devillers, V. Carreau, M. Di Filippo, D. Erlich, M. Frost, M.J. Malloy, V.N. Shumaker, J.P. Kane, Familial ligand-defective apolipopro-
Abifadel, et al., Molecular spectrum of autosomal dominant hypercholesteroemia in tein B. Identification of a new mutation that decreased LDL binding capacity, J.
France, Hum. Mutat. 31 (2010) E1811–E1824. Clin. Invest. 95 (1995) 1225–1234.
[62] L. Pisciotta, C.P. Oliva, G.M. Pes, L. Di Scala, A. Bellocchio, R. Fresa, A. Cantafora, M. [87] M. Soufi, A.M. Sattler, W. Maerz, A. Starke, M. Herzum, B. Maisch, J.R. Schaefer, A
Arca, S. Calandra, S. Bertolini, Autosomal recessive hypercholesterolemia and new but frequent mutation of apoB-100-apoBHis3543Tyr, Atherosclerosis 174
homozygous familial hypercholesterolemia: a phenotypic comparison, Atheroscle- (2004) 11–16.
rosis 188 (2006) 398–405. [88] M.M. Motazacker, J. Pirruccello, R. Huijgen, R. Do, S. Gabriel, J. Peter, J.A.
[63] R.P. Naoumova, C. Neuwirth, P. Lee, J.P. Miller, K.G. Taylor, A.K. Soutar, Autosomal Kuivenhoven, J.C. Defesche, J.J.P. Kastelein, G.K. Hovingh, N. Zelcer, S. Kathiresan,
recessive hypercholesterolemia: long term follow up and response to treatment, S.W. Fouchier, Advances in genetics show the need for extending screening strat-
Atherosclerosis 174 (2004) 165–172. egies for autosomal dominant hypercholesterolemia, Eur. Heart J. 33 (2012)
[64] G.K. Hovingh, M.H. Davidson, J.J.P. Kastelein, A.M. O'Connor, Diagnosis and treat- 1360–1366.
ment of familial hypercholesterolemia, Eur. Heart J. 34 (2013) 962–971. [89] A.C. Alves, A. Etxebarria, A.K. Soutar, C. Martin, M. Bourbon, Novel functional APOB
[65] S.E. Humphries, R.A. Whitall, C.S. Hubbart, S. Maplebeck, J.A. Cooper, A.K. Soutar, R. mutations outside LDL-binding region causing familial hypercholesterolemia,
Naoumova, G.R. Thompson, M. Seed, P.N. Durrington, J.P. Miller, D.J.B. Betteridge, Hum. Mol. Genet. 23 (2014) 1817–1828.
Neil HAW for the Simon Broome Familial Hyperlipidemia Register Group and [90] J.P. Rabes, M. Varret, M. Devillers, P. Aegerter, L. Villeger, M. Krempf, C. Junien, C.
Scientific Steering Committee, J. Med. Genet. 43 (2006) 943-749. Boileau, R3531C mutation in apoliporotein B gene is not sufficient to cause hyper-
[66] S.E.A. Leigh, A.H. Foster, R.A. Whittall, C.S. Hubbart, S.E. Humphries, Update and cholesterolemia, Arterioscler. Thromb. Vasc. Biol. 20 (2000) 76–82.
analysis of the university college low density lipoprotein receptor familial hyper- [91] M. Benn, B.G. Nordestgaard, G.B. Jensen, A. Tybjaerg-Hansen, Improving prediction
cholesterolemia database, Ann. Hum. Genet. 72 (2008) 485–498. of ischemic cardiovascular disease in the general population using apolipoprotein
[67] K.E. Liyanage, J.R. Burnett, A.J. Hooper, F.M. van Bockxmeer, Familial hypercholes- B. The Copenhagen City Heart Study, Arterioscler. Thromb. Vasc. Biol. 27 (2007)
terolemia: epidemiolgy, Neolithic origins and modern geographic distribution, 661–670.
Crit. Rev. Clin. Lab. Sci. 48 (2011) 1–18. [92] M. Benn, Apolipoprotein B levels, APOB alleles and the risk of ischemic cardiovas-
[68] B. Sjouke, D.M. Kusters, I. Kindt, J. Besseling, J.C. Defesche, E.J.G. Sijbrands, J.E.R. van cular disease in the general population, a review, Atherosclerosis 206 (2009)
Lennep, A.F.H. Stalenhoef, A. Wiegman, J. de Graaf, S.W. Fouchier, J.J.P. Kastelein, 17–30.
G.K. Hovingh, Homozygous autosomal dominant hypercholesterolemia in The [93] B. Cariou, C. Le May, P. Costet, Clinical aspects of PCSK9, Atherosclerosis 216 (2011)
Netherlands: prevalence, genotype-phenotype relationship and clinical outcome, 258–265.
Eur. Heart J. (2014), http://dx.doi.org/10.1093/eurheart/ehu058. [94] S. Naureckiene, L. Ma, K. Sreekumar, U. Purandare, C.F. Lo, Y. Huang, L.W. Chiang,
[69] A. Garg, V. Simha, Update on dyslipidemia, J. Clin. Endocrinol. Metab. 92 (2007) J.M. Grenier, et al., Functional characterization of Narc 1, a novel proteinase related
1581–1589. to proteinase K, Arch. Biochem. Biophys. 420 (2003) 55–67.
[70] E. Usifo, S.E.A. Leigh, R.A. Whittall, N. Lench, A. Taylor, C. Yeats, C.A. Orengo, A.C.R. [95] B. Dong, M. Wu, H. Li, F.B. Kraemer, K. Adeli, N.G. Seidah, S.W. Park, J. Liu, Strong
Martin, J. Celli, S.E. Humphries, Low-density lipoprotein receptor gene familial hy- induction of PCSK9 gene expression through HNF1α and SREBP2: mechanism for
percholesterolemia variant database: update and pathological assessment, Ann. the resistance to LDL-cholesterol lowering effect of statins in dyslipidemic ham-
Hum. Genet. 76 (2012) 387–401. sters, J. Lipid Res. 51 (2010) 1486–1495.
[71] R.P. Naoumova, C. Neuwirth, B. Pottinger, R. Whittal, S.E. Humphries, A.K. Soutar, [96] Y. Duan, Y. Chen, W. Hu, X. Li, X. Yang, X. Zhou, Z. Yin, D. Kong, Z. Yao, D.P. Hajjar, L.
Genetic diagnosis of familial hypercholesterolemia: a mutation and rare non- Liu, Q. Liu, J. Han, Peroxisome proliferator-activated receptor γ activation by ligands
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 175

and dephosphorylation induces proprotein convertase subtilisin kexin type 9 and plasma lipids among familial hypercholesterolemia, Atherosclerosis 186 (2006)
low density receptor expression, J. Biol. Chem. 287 (2012) 23667–23677. 443–450.
[97] A.S. Petersen, L.G. Fong, S.G. Young, PCSK9 function and physiology, J. Lipid Res. 49 [123] E.L. Klett, H.M. Lee, D.B. Adams, K.D. Chavin, S.B. Patel, Localization of ABCG5 and
(2008) 1595–1599. ABCG8 proteins in human liver, gall bladder and intestine, BMC Gastroenterol.
[98] K.N. Maxwell, E.A. Fisher, J.L. Breslow, Overexpression of PCSK9 accelerates the 307 (2004) 1021–1028.
degradation of the LDLR in a post-endoplasmic reticulum compartment, Proc. [124] E. Polisecki, I. Peter, J.S. Simon, R. Hegele, M. Robertson, I. Ford, J. Shepherd, C.
Natl. Acad. Sci. 102 (2005) 2069–2074. Packard, J.W. Jukema, A.J.M. de Craen, R.G.J. Westendorp, B.M. Buckley, E.J.
[99] T.S. Fisher, P. Lo Surdo, S. Pandit, M. Mattu, J.C. Santoro, D. Wisniewski, R.T. Schaefer, on behalf of the Prospective Study of Pravastatin in the elderly at risk
Cummings, A. Calzetta, R.M. Cubbon, et al., Effects of pH and low density lipopro- (PROSPER) investigators, Genetic variation at the NPC1L1 gene locus, plasma lipo-
tein (LDL) on PCSK9-dependent LDL receptor regulation, J. Biol. Chem. 282 (2007) proteins and heart disease risk in the elderly, J. Lipid Res. 51 (2010) 1201–1207.
20502–20512. [125] L. Wang, J. Wang, N. Li, L. Ge, B. Li, B. Song, Molecular characterization of the NPC1L1
[100] R.M. DeVay, D.L. Shelton, H. Liang, Characterisation of propprotein convertase sub- variants identified from low absorbers, J. Biol. Chem. 286 (2011) 7397–7408.
tilisin/kexin type 9 (PCSK9) trafficking reveals a novel lysosomal targeting mecha- [126] S. Kidambi, S.B. Patel, Sitosterolemia: pathophysiology, clinical presentation and
nism via amyloid precursor like protein 2 (APLP2), J. Biol. Chem. 288 (2013) laboratory diagnosis, J. Clin. Pathol. 61 (2008) 588–594.
10805–10818. [127] J. Rios, E. Stein, J. Shendure, H.H. Hobbs, J.C. Cohen, Identification by whole genome
[101] M. Abifadel, M. Varret, J.P. Rabes, D. Allard, K. Ouguerram, M. Devillers, et al., Mu- resequencing of gene defect responsible for severe hypercholesterolemia, Hum.
tations in PCSK9 cause autosomal dominant hypercholesterolemia, Nat. Genet. 34 Mol. Genet. 19 (2010) 4313–4318.
(2003) 154–156. [128] J. Wang, T. Joy, D. Mymin, J. Frohlich, R.A. Hegele, Phenotypic heterogeneity of
[102] R.P. Naoumova, I. Tosi, D. Patel, C. Neuwirth, S.D. Horswell, A.D. Marais, C. van sitosterolemia, J. Lipid Res. 5 (2004) 2361–2367.
Heyningen, A.K. Soutar, Sever hypercholesterolemia in four British families with [129] M. Miller, N.J. Stone, C. Ballantyne, V. Bittner, M.H. Criqui, H.N. Ginsberg, A.C.
the D374Y mutation in the PCSK9 gene. Long term follow up and treatment re- Goldberg, W.J. Howard, et al., Triglycerides and cardiovascular disease, Circulation
sponse, Arterioscler. Thromb. Vasc. Biol. 25 (2005) 2654–2660. 123 (2011) 2292–2333.
[103] D. Cunningham, D.E. Danley, K.F. Geohegan, M.C. Griffor, J.L. Hawkins, et al., Struc- [130] J.P. Despres, I. Lemieux, G.R. Dagenais, B. Cantin, B. Lamarche, HDL cholesterol as a
tural and biophysical studies of PCSK9 and its mutants linked to familial hypercho- marker of coronary heart disease risk. The Quebec cardiovascular heart study,
lesterolemia, Nat. Struct. Mol. Biol. 14 (2007) 413–419. Atherosclerosis 153 (2000) 263–272.
[104] H.J. Kwon, T.A. Lagace, M.C. McNutt, J.D. Horton, J. Deisenhofer, Molecular basis for [131] J. Boren, N. Matikainen, M. Adiels, M. Taskinen, Postprandial hypertriglyceridemia
LDL receptor recognition by PCSK9, Proc. Natl. Acad. Sci. 105 (2008) 1820–1825. as a coronary risk factor, Clin. Chim. Acta 431 (2014) 131–142.
[105] M. Abifadel, M. Guerin, S. Benjannet, J.P. Rabes, et al., Identification and character- [132] A. Varbo, M. Benn, A. Tybjaerg-Hansen, A.B. Jorgensen, R. Frikke-Schmidt,
isation of new gain of function mutations in the PCSK9 gene responsible for auto- Nordestgaard. BG. Remnant cholesterol as a causal risk factor for ischemic heart
somal dominant hypercholesterolemia, Atherosclerosis 223 (2012) 394–400. disease, J. Am. Coll. Med. 61 (2013) 427–436.
[106] X. Sun, E.R. Eden, I. Tosi, C.K. Neuwirth, D. Wile, R.P. Naoumova, A.K. Soutar, Evi- [133] B.G. Nordestgaard, M. Benn, P. Schnohr, A. Tybjaerg-Hansen, Nonfasting triglycer-
dence for effect of mutant PCSK9 on apolipoprotein secretion as the cause of un- ides and risk of myocardial infarction, ischemic heart disease and death in men
usually severe dominant hypercholesterolemia, Hum. Mol. Genet. 14 (2005) and women, J. Am. Med. Assoc. 298 (2007) 299–308.
1161–1169. [134] S.S. Martin, M.J. Blaha, Genetically low triglycerides and mortality: further support
[107] O.N. Akram, A. Bernier, F. Petrides, G. Wong, G. Lambert, Beyond LDL cholesterol, a for “the earlier the better”, Clin. Chem. 60 (2014) 705–707.
new role for PCSK9, Arterioscler. Thromb. Vasc. Biol. 30 (2010) 1279–1281. [135] M.J. Chapman, H.N. Ginsberg, P. Amarenco, F. Andreotti, J. Boren, A.L. Catapano, O.S.
[108] V.M. Homer, A.D. Marais, F. Charlton, A.D. Laurie, N. Hurndell, R. Scott, F. Mangili, Descamps, E. Fisher, P.T. Kovanen, et al., for the European Atherosclerosis Society
D.R. Sullivan, P.J. Barter, K.A. Rye, P.M. George, G. Lambert, Identification and char- Consensus Panel, Triglyceride-rich lipoproteins and high-density lipoprotein cho-
acterization of two non secreted PCSK9 mutants associated with familial hyper- lesterol in patients at high risk of cardiovascular disease: evidence and guidance
cholesterolemia in cohorts from New Zealand and South Africa, Atherosclerosis for management, Eur. Heart J. 32 (2011) 1345–1361.
196 (2008) 659–666. [136] P. Benlian, J.L. De Gennes, L. Foubert, H. Zhang, S.E. Gagne, M. Hayden, Premature
[109] J.D. Horton, J.C. Cohen, H.H. Hobbs, PCSK9: a convertase that coordinates LDL catab- atherosclerosis in patients with familial chylomicronemia caused by mutations in
olism, J. Lipid Res. 50 (2009) S172–S177. the lipoprotein lipase gene, N. Engl. J. Med. 335 (1996) 848–854.
[110] M. Abifadel, J.P. Rabes, M. Devillers, A. Munnich, D. Erlich, C. Junien, M. Varret, C. [137] A. Varbo, M. Benn, A. Tybjaerg-Hansen, B.G. Nordestgaard, Elevated remnant cho-
Boileau, Mutations and polymorphism in the proprotein convertase subtilisin lesterol causes both low grade inflammation and ischemic heart disease, whereas
kexin 9 (PCSK9) gene in cholesterol metabolism and disease, Hum. Mutat. 30 elevated low-density lipoprotein cholesterol causes ischemic heart disease with-
(2009) 520–529. out inflammation, Circulation 128 (2013) 1298–1309.
[111] D. Allard, S. Amsellem, M. Abifadel, M. Trillard, M. Devillers, et al., Novel mutations [138] A.L. Catapano, Z. Reiner, G. De Backer, et al., ESC/EAS guidelines for the manage-
of the PCSK9 gene causes variable phenotypes of autosomal dominant hypercho- ment of dyslipidemias: the Task Force for the management of dyslipidemias of
lesterolemia, Hum. Mutat. 26 (2005) 497. the European Society of Cardiology (ESC) and European Atherosclerosis Society
[112] L. Pisciotta, C.P. Oliva, A.B. Cefalu, D. Noto, A. Bellochio, R. Fresa, A. Cantafora, D. (EAS), Atherosclerosis 217 (2011) 3–46.
Patel, M. Averna, P. Tarugi, S. Calandra, S. Bertolini, Additive effect of mutations [139] D.R. MacLean, A. Petrsovits, M. Nargundkar, P.W. Connelly, E. MacLeod, A. Edwards,
in LDLR and PCSK9 genes on the phenotype of familial hypercholesterolemia, Ath- P. Hessel, Canadian heart health surveys: a profile of cardiovascular risk. Survey
erosclerosis 186 (2006) 433–440. methods and data analysis. Canadian Heart Health Surveys Research Group,
[113] A.K. Khachadurian, The inheritance of essential familial hypercholesterolemia, Am. CMAJ 146 (1992) 1969–1974.
Med. J. 37 (1964) 402–407. [140] A. Brahm, R.A. Hegele, Hypertriglyceridemia, Nutrients 5 (2013) 981–1001.
[114] C.K. Garcia, K. Wilund, M. Arca, G. Zuliani, R. Fellin, M. Maioli, S. Calandra, S. [141] H.C. Hassing, R.P. Surendran, H.L. Mooij, E.S. Stroes, M. Nieuwdorp, G.M. Dallinga-
Bertolini, F. Cossu, N. Grishin, R. Barnes, J.C. Cohen, H.H. Hobbs, Autosomal reces- Thie, Pathophysiology of hypertriglyceridemia, Biochem. Biophys. Acta 2012
sive hypercholesterolemia caused by mutations in a putative LDL receptor adaptor (1821) 826–832.
protein, Science 292 (2001) 1394–1398. [142] G. Yuan, K.Z. Al-Shali, R.A. Hegele, Hypertriglyceridemia: its etiology, effects and
[115] E.R. Eden, D.D. Patel, X. Sun, J.J. Burden, M. Themis, M. Edwards, P. Lee, C. Neuwirth, treatment, CMAJ 176 (2007) 1113–1120.
R.P. Naoumova, A.K. Soutar, Restoration of LDL receptor function in cells from pa- [143] P.R. Surendran, M.E. Visser, S. Heemeiaar, J. Wang, J. Peter, J.C. Defesche, et al., Mu-
tients with autosomal recessive hypercholesterolemia by retroviral expression of tations in LPL, APOC2, APOA5, GPIHBP1 and LMF1 in patients with severe hypertri-
ARH1, J. Clin. Invest. 110 (2002) 1695–1702. glyceridemia, J. Intern. Med. 272 (2012) 185–196.
[116] A.K. Soutar, R.P. Naoumova, Mechanisms of disease: genetic causes of familial hy- [144] A.R. Rahalkar, F. Giffen, H. Bryan, J. Ho, K.M. Morrison, J. Hill, J. Wang, R.A. Hegele, T.
percholesterolemia, Nat. Clin. Pract. Cardiovasc. Med. 4 (2007) 214–225. Joy, Novel LPL mutations associated with lipoprotein lipase deficiency: two case re-
[117] J.C. Cohen, M. Kimmel, A. Polanski, Molecular mechanisms of autosomal recessive ports and a literature review, Can. J. Physiol. Pharmacol. 87 (2009) 151–160.
hypercholesterolemia, Curr. Opin. Lipidol. 14 (2003) 121–127. [145] S.G. Young, R. Zechner, Biochemistry and pathophysiology of intravascular and in-
[118] H. Tada, M. Kawashiri, R. Ohtani, T. Noguchi, C. Nakanishi, T. Konno, K. Hayashi, A. tracellular lipolysis, Genes Dev. 27 (2013) 459–484.
Nohara, A. Inazu, J. Kobayashi, H. Mabuchi, M. Yamagishi, A novel type of familial [146] C.V. Voss, B.S.J. Davies, S. Tat, P. Gin, L.G. Fong, C. Pelletier, C.D. Mottler, A.
hypercholesterolemia: double heterozygous mutations in LDL receptor adaptor Bensadoun, A.P. Beigneux, S.G. Young, Mutations in lipoprotein lipase that block
protein 1 gene, Atherosclerosis 219 (2011) 663–666. binding to the endothelial cell transporter GPIHBP1, Proc. Natl. Acad. Sci. 108
[119] M. Soufi, S. Rust, M. Walter, J.R. Schaefer, A combined LDL receptor/LDL receptor (2011) 7980–7984.
adaptor protein mutation as the cause for severe familial hypercholesterolemia, [147] P. Gin, C.N. Goulbourne, O. Adeyo, A.P. Beigneux, B.S.J. Davies, S. Tat, C.V. Voss, A.
Gene 521 (2013) 200–203. Bensadoun, L.G. Fong, S.G. Young, Chylomicronemia mutations yield new insights
[120] C.R. Pullinger, C. Eng, G. Salen, S. Shefer, A.K. Batta, S.K. Erickson, A. Verhagen, C.R. into interactions between lipoprotein lipase and GPIHBP1, Hum. Mol. Genet. 21
Rivera, S.J. Mulvihill, M.J. Malloy, J.P. Kane, Human cholesterol 7α-hydroxylase (2012) 2961–2972.
(CYP7A1) deficiency has a hypercholesterolemic phenotype, J. Clin. Invest. 110 [148] Y. Ma, M. Liu, D. Ginzinger, J. Frohlich, J.D. Brunzeli, M.R. Hayden, Gene–
(2002) 109–117. environment interaction in the conversion of a mild to severe phenotype in a pa-
[121] R. Burkhardt, E.E. Kenny, J.K. Lowe, A. Birkeland, R. Josowitz, M. Noel, J. Salit, J.B. tient homozygous for a Ser172 Cys mutation in the lipoprotein lipase gene, J. Clin.
Maller, I. Pe'er, M.J. Daly, D. Alshuler, M. Stoffel, J.M. Friedman, J.L. Breslow, Com- Invest. 91 (1993) 1953–1958.
mon SNPs in HMGCR in Micronesians and whites associated with LDL- [149] D.E. Wilson, M. Emi, P.H. Iverius, A. Hats, L.L. Wu, E. Hillas, Phenotypic expression of
cholesterol levels affect alternative splicing of Exon13, Arterioscler. Thromb. heterozygous lipoprotein lipase deficiency in the extended pedigree of a proband
Vasc. Biol. 28 (2008) 2078–2084. homozygous for missense mutation, J. Clin. Invest. 86 (1990) 735–750.
[122] R. Durst, A. Jansen, G. Eraz, R. Bravdo, E. Butbul, L.B. Avi, S. Shiptzen, C. Lotan, et al., [150] S. Bijvoet, S.E. Gagne, S. Moorjani, C. Gagne, H.E. Henderson, J.C. Fruchart, J.
The discrete and combined effect of SREBP-2 and SCAP isoforms in the control of Dallongeville, P. Alaupovic, M. Prins, J.J.P. Kastelein, M.R. Hayden, Alterations in
176 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

plasma lipoproteins and apolipoproteins before age of 40 in heterozygotes for li- lipase activity and mass: identification of a novel GPIHBP1 mutation, J. Intern. Med.
poprotein lipase deficiency, J. Lipid Res. 37 (1996) 640–650. 270 (2011) 224–228.
[151] A.G. Soto, A. McIntyre, S. Agrawal, S.R. Bialo, R.A. Hegele, C.M. Boney, Severe [177] H. Yamamoto, M. Onishi, N. Miyamoto, R. Oki, H. Ueda, M. Ishigami, H. Hiraoka, Y.
hypertriglyceridemia due to a novel p.Q240H mutation in the lipoprotein lipase Matsuzsawa, S. Kihara, Novel combined GPIHBP1 mutations in a patient with
gene, Lipids Health Dis. 14 (2015) 102. hypertriglyceridemia associated with CAD, J. Atheroscler. Thromb. 20 (2013)
[152] J.M. Martin-Campos, J. Julve, R. Roig, S. Martinez, T.L. Errico, S. Martinez-Couselo, 777–784.
J.C. Escola-Gil, J. Mendez-Gonzalez, F. Blanco-Vaca, Molecular analysis of [178] J.J. Rios, S. Shastry, J. Jasso, N. Hauser, A. Garg, A. Bensadoun, J.C. Cohen, H.H. Hobbs,
chylomicronemia in a clinical laboratory setting. Diagnosis of 13 cases of lipopro- Deletion of GPIHBP1 causing severe chylomicronemia, J. Inherit. Metab. Dis. 35
tein lipase deficiency, Clin. Chim. Acta 429 (2014) 61–68. (2012) 531–540.
[153] S. Gehrisch, Common mutations of the lipoprotein lipase gene and their clinical [179] K.E. Berge, K. Retterstol, S. Romeo, C. Pirazzi, T.P. Leren, Type 1 hyperlipoproteinemia
significance, Curr. Atheroscler. Rep. 1 (1999) 70–78. due to a novel deletion of exons 3 and 4 in the GPIHBP1 gene, Atherosclerosis 234
[154] C. Gagne, L. Brun, P. Julien, S. Moorjani, P. Lupien, Primary lipoprotein lipase activity (2014) 30–33.
deficiency: clinical investigation of a French Canadian population, CMAJ 140 [180] O. Adeyo, C.N. Goulbourne, A. Bensadoun, A.P. Beigneux, L.G. Fong, S.G. Young,
(1989) 405–410. GPIHBP1 and the intravascular processing of triglyceride rich lipoproteins, J. Intern.
[155] L. Foubert, J.L. De Gennes, J.P. Lagarde, E. Ehrenborg, A. Raisonnier, J.P. Girardet, Med. 272 (2012) 528–540.
M.R. Hayden, P. Benlian, Assessment of French patients with LPL deficiency for [181] V. Sharma, R.O. Ryan, T.M. Forte, Apolipoprotein A-V dependent modulation of
French Canadian mutations, J. Med. Genet. 34 (1997) 672–675. plasma triacyglycerol: a puzzlement, Biochim. Biophys. Acta 2012 (1821)
[156] J. Rip, M.C. Nierman, C.J. Ross, J.W. Jukema, M.R. Hayden, J.J.P. Kastelein, E.S.G. Stoes, 795–799.
J.A. Kuivenhoven, Lipoprotein lipase S447X. A naturally occurring gain of function [182] J.C. Gonzales, P.L. Gordts, E.M. Foley, J.D. Esko, Apolipoproteins E and AV mediate
mutation, Arterioscler. Thromb. Vasc. Biol. 26 (2006) 1236–1245. lipoprotein clearance by hepatic proteoglycans, J. Clin. Invest. 123 (6) (2013)
[157] W.C. Breckenridge, J.A. Little, G. Steiner, A. Chow, M. Poapst, Hypertriglyceridemia 2742–2751.
associated with deficiency of apolipoprotein C-II, N. Engl. J. Med. 298 (1978) [183] C.T. Johansen, R.A. Hegele, The complex genetic basis of plasma triglycerides, Curr.
1265–1273. Atheroscler. Rep. 14 (2012) 227–234.
[158] P.W. Connelly, G.F. Maguire, J.A. Little, Apolipoprotein CIISt. Michael. Familial apolipo- [184] J. Wang, M.R. Ban, B.A. Kennedy, S. Anand, S. Yusuf, M.W. Huff, R.L. Pollex, R.A.
protein CII deficiency associated with premature vascular disease, J. Clin. Invest. 80 Hegele, APOA5 genetic variants are markers for classic hyperlipoproteinemia phe-
(1987) 1597–1606. notypes and hypertriglyceridemia, Nat. Clin. Pract. Cardiovasc. Med. 5 (2008)
[159] M.S. Nauck, H. Nissen, M.M. Hoffmann, J. Herwig, C.R. Pullinger, M. Averna, J. Geisel, 730–737.
H. Wieland, W. März, Detection of mutations in the apolipoprotein CII gene by de- [185] C. Marcais, B. Verges, S. Charriere, V. Pruneta, M. Merlin, S. Billon, L. Perrot, J. Drai,
naturing gradient gel electrophoresis. Identification of the splice site variant apoli- A. Sassolas, L.A. Pennacchio, J. Fruchart-Najib, J.C. Fruchart, V. Durlach, P. Moulin,
poprotein CII-Hamburg in a patient with severe hypertriglyceridemia, Clin. Chem. Apoa5 Q139X truncation predisposes to late-onset hyperchylomicronemia due to
44 (1998) 1388–1396. lipoprotein lipase impairment, J. Clin. Invest. 115 (2005) 2862–2869.
[160] S.S. Fojo, S.W. Law, H.B. Brewer, The human preproapolipoprotein C-II gene. Com- [186] S. Charriere, C. Cugnet, M. Guitard, S. Bernard, L. Groisne, M. Charcosset, V. Pruneta-
plete nucleic acid sequence and genomic organization, FEBS Lett. 213 (1987) Deloche, M. Merlin, S. Billon, A. Sassolas, P. Moulin, C. Marcais, Modulation of phe-
221–226. notypic expression of APOA5 Q97X and L242P mutations, Atherosclerosis 207
[161] R. Streicher, J. Geisel, C. Weisshaar, H. Avci, K. Oette, D. Muller-Wieland, W. Krone, (2009) 150–159.
A single nucleotide substitution in the promoter region of apolipoprotein C-II gene [187] C. Dussaillant, V. Serrano, A. Maiz, S. Eyheramendy, L.R. Cataldo, M. Chavez, S.V.
identified in individuals with chylomicronemia, J. Lipid Res. 37 (1996) 2599–2607. Smalley, M. Fuentes, A. Rigotti, L. Rubio, C.F. Lagos, J.A. Martinez, J.L. Santos,
[162] C. Crecchio, A. Capurso, G. Pepe, Identification of the mutation responsible for a APOA5 Q97X mutation identified through homozygosity mapping causes severe
case of plasmatic apolipoprotein CII deficiency (Apo CII-bAri), Biochem. Biophys. hypertriglyceridemia in a Chilean consanguineous family, BMC Med. Genet. 13
Res. Commun. 168 (1990) 1118–1127. (2012) 106.
[163] H. Inadera, A. Hibino, J. Kobayashi, T. Kanzaki, K. Shirai, S. Yukawa, Y. Saito, S. [188] C.P. Oliva, L. Pisciotta, G.L. Volti, M.P. Sambataro, A. Cantafora, A. Bellocchio, A.
Yoshida, A missense mutation (Trp26-Arg) in exon 3 of the apolipoprotein CII Catapano, P. Tarugi, S. Bertolini, S. Calandra, Inherited apolipoprotein A-V deficien-
gene in a patient with apolipoprotein CII deficiency (apoCII-Wakayama), Biochem. cy in severe hypertriglyceridemia, Arterioscler. Thromb. Vasc. Biol. 25 (2005)
Biophys. Res. Commun. 193 (1993) 1174–1183. 411–417.
[164] S.S. Fojo, U. Beisiegel, U. Bell, K. Higuchi, M. Bojanovski, R.E. Gregg, Donor splice site [189] K. Albers, C. Schlein, K. Wenner, P. Lohse, A. Bartelt, J. Heeren, R. Santer, M. Merkel,
mutation in the apolipoprotein (Apo) C-II gene (Apo C-IIHamburg) of a patient with Homozygosity for a partial deletion of apoprotein A-V signal peptide results in in-
Apo C-II deficiency, J. Clin. Invest. 82 (1988) 1489–1494. tracellular missorting of the protein and chylomicronemia in a breast fed infant,
[165] C. Gabelli, S. Bilato, S.S. Fojo, S. Martini, H.B. Brewer, G. Crepaldi, G. Baggio, Hetero- Atherosclerosis 233 (2014) 97–103.
zygous apolipoprotein C-II deficiency: lipoprotein and apoprotein phenotype and [190] E. Mendoza-Barbera, J. Julve, S.K. Nilsson, A. Lookene, J.M. Martin-Compos, R. Roig,
RsaI restriction enzyme polymorphism in the Apo-CIIPadova kindred, Eur. J. Clin. A.M. Lechunga-Sancho, J.H. Sloan, P. Fuentes-Prior, F. Blanco-Vaca, Structural and
Investig. 23 (1993) 522–528. functional analysis of APOA5 mutations identified in patients with severe hypertri-
[166] M. Peterfy, Lipase maturation factor 1: a lipase chaperone involved in lipid metab- glyceridemia, J. Lipid Res. 54 (2013) 649–661.
olism, Biochim. Biophys. Acta 2012 (1821) 790–794. [191] A.J. Hooper, J. Kurtkoti, I. Hamilton-Craig, J.R. Burnett, Clinical features and genetic
[167] M. Peterfy, O. Ben-Zeev, H.Z. Mao, D. Weissglas-Volkov, B.E. Aouizerat, C.R. analysis of three patients with severe hypertriglyceridemia, Ann. Clin. Biochem. 51
Pullinger, P.H. Frost, J.P. Kane, M.J. Malloy, K. Reue, et al., Mutations in the LMF1 (2014) 485–489.
cause combined lipase deficiency and severe hypertriglyceridemia, Nat. Genet. [192] L.A. Pennacchio, M. Olivier, J.A. Hubacek, R.M. Krauss, E.M. Rubin, J.C. Cohen, Two
39 (2007) 1483–1487. independent apolipoprotein A5 haplotypes influence plasma triglyceride levels,
[168] A.B. Cefalu, D. Noto, M.L. Arpi, F. Yin, R. Spina, H. Hilden, C.M. Brabagallo, A. Hum. Mol. Genet. 11 (2002) 3031–3038.
Carroccio, P. Tarugi, S. Squatrito, R. Vigneri, M.R. Taskinen, M. Peterfy, M.R. [193] J.D. Brunzell, N.E. Miller, P. Alaupovic, R.J. St Hilaire, C.S. Wang, D.L. Sarson, S.R.
Averna, Novel LMF1 mutation in a patient with severe hypertriglyceridemia, J. Bloom, B. Lewis, Familial chylomicronemia due to circulating inhibitor of
Clin. Endocrinol. Metab. 94 (2009) 4584–4590. lipoproteon lipase activity, J. Lipid Res. 24 (1983) 12–19.
[169] A.P. Beigneux, B. Davies, P. Gin, M.M. Weinstein, E. Farber, X. Qiao, P. Peale, S. [194] C.E.M. Rodrigues, E. Bonfa, J.F. Carvalho, Review on anti-lipoprotein lipase antibod-
Bunting, R.L. Walzem, J.S. Wong, et al., Glycosylphosphatidylinositol-anchored ies, Clin. Chim. Acta 411 (2010) 1603–1605.
high density lipoprotein 1 plays a critical role in the lipolytic processing of chylo- [195] A.P. Ashraf, T. Beukelman, V. Pruneta-Deloche, D.R. Kelly, A. Garg, Type 1
microns, Cell Metab. 5 (2007) 279–291. hyperlipoproteinemia in recurrent acute pancreatitis due to lipoprotein lipase an-
[170] A.P. Beigneux, M.M. Weinstein, B.S.J. Davies, P. Gin, A. Bensadoun, L.G. Fong, tibody in a young girl with Sjogren's syndrome, J. Clin. Endocrinol. Metab. 96
S.G. Young, GPIHBP1 and lipolysis: an update, Curr. Opin. Lipidol. 20 (2009) (2011) 3302.
211–216. [196] J. Attia, J.P.A. Ioannidis, A. Thakkinistian, M. McEvoy, R.J. Scott, C. Minelli, J.
[171] W. Yan, F. Shen, B. Dillon, M. Ratnam, The hydrophobic domains in the carboxy- Thompson, C. Infante-Rivard, G. Guyatt, How to use an article about genetic asso-
terminal signal for GPI modification and in the amino terminal leader peptide ciation. A. Background concepts, JAMA 301 (2009) 74–81.
have similar structural requirements, J. Mol. Biol. 275 (1998) 25–33. [197] T.A. Manolio, F.S. Collins, N.J. Cox, D.B. Goldstein, L.A. Hindorff, D.J. Hunter, et al.,
[172] A.P. Beigneux, R. Franssen, A. Bensadoun, P. Gin, K. Melford, J. Peter, R.L. Walzem, Finding the missing heritability of complex diseases, Nature 46 (2009) 747–752.
M.M. Weinstein, et al., Chylomicronemia with a mutant GPIHBP1 (Q115P) that [198] R. Plomin, C.M.A. Howarth, O.S.P. Davis, Common disorders are quantitative traits,
cannot bind lipase, Arterioscler. Thromb. Vasc. Biol. 29 (2009) 956–962. Nat. Genet. 10 (2009) 872–878.
[173] R. Franssen, S.G. Young, F. Peelman, J. Hertecant, J.A. Sierts, W.M. Schimmel, et al., [199] T.M. Teslovich, L. Musunru, A.V. Smith, et al., Biological, clinical and population rel-
Chylomocronemia with low postheparin lipoprotein lipase levels in the setting of evance of 95 loci for blood lipids, Nature 466 (2010) 707–713.
GPIHBP1 defects, Circ. Cardiovasc. Genet. 3 (2010) 169–178. [200] C.T. Johansen, S. Kathiresan, R.A. Hegele, Genetic determinants of plasma triglycer-
[174] G. Olivecrona, E. Ehrenborg, H. Semb, E. Makoveichuk, A. Lindberg, M.R. Hayden, ides, J. Lipid Res. 52 (2011) 189–206.
et al., Mutation of conserved cysteines in the Ly6 domain of GPIHBP1 in familial [201] J. Wang, H. Cao, M.R. Ban, B.A. Kennedy, S. Zhu, S. Anand, S. Yusuf, R.L. Pollex, R.A.
chylomicronemia, J. Lipid Res. 51 (2010) 1535–1545. Hegele, Resequencing genomic DNA of patients with severe hypertriglyceridemia
[175] S. Charriere, N. Peretti, S. Bernard, M. Di Filippo, A. Sassolas, M. Merlin, M. Delay, (MIM 144650), Arterioscler. Thromb. Vasc. Biol. 27 (2007) 2450–2455.
et al., GPIHBP1 C89F neomutation and hydrophobic C-terminal domain G175R mu- [202] C.T. Johansen, J. Wang, M.B. Lanktree, H. Cao, A.D. McIntyre, M.R. Ban, R.A. Martins,
tation in two pedigrees with severe hyperchylomicronemia, J. Clin. Endocrinol. et al., Excess of rare variants in genes identified by genome-wide association study
Metab. 96 (2011) E1675–E1679. of hypertriglyceridemia, Nat. Genet. 42 (2010) 684–687.
[176] I. Coca-Prieto, O. Kroupa, P. Gonzalez-Santos, J. Magne, G. Olivecrona, E. Ehrenborg, [203] M. Abifadel, L. Bernier, G. Dubuc, G. Nuel, J.P. Rabes, J. Bonneau, A. Marques, M.
P. Valdivielso, Childhood onset chylomicronemia with reduced plasma lipoprotein Marduel, M. Devillers, A. Munnich, D. Erlich, M. Varret, M. Roy, J. Davignon, C.
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 177

Boileau, A PCSK9 variant and familial combined hyperlipidemia, J. Med. Genet. 45 [230] L. Bertram, C.M. Lill, R.E. Tanzi, The genetics of Alzheimer disease. Back to the fu-
(2008) 780–786. ture, Neuron 68 (2010) 270–281.
[204] K. Nakayama, T. Bayasgalan, K. Yamanaka, M. Kumada, T. Gotoh, N. Utsumi, Y. [231] F. De Beer, A.F.H. Stalenhoef, N. Hoogerbrugge, J.J.P. Kastelein, J.A.G. Leuven, C.M.
Yanagisawa, M. Okayama, E. Kajii, S. Ishibashi, S. Iwamoto, J.C.G. Team, Large van Duijn, L.M. Havekes, A.H.M. Smelt, Expression of type 3 hyperlipoproteinemia
scale replication analysis of loci associated with lipid concentration in a Japanese in apolipoprotein E2 (Arg158-Cys) homozygotes is associated with
population, J. Med. Genet. 46 (2009) 370–374. hyperinsulinemia, Arterioscler. Thromb. Vasc. Biol. 22 (2002) 294–299.
[205] C.T. Johansen, J. Wang, M.B. Lanktree, A.D. McIntyre, M.R. Ban, R.A. Martins, et al., [232] G. Feussner, S. Piesch, J. Dobmeyer, C. Fischer, Genetics of type 3 hyperlipoproteinemia,
An increased burden of common and rare lipid associated risk alleles contributes Genet. Epidemiol. 14 (1997) 283–297.
to the phenotypic spectrum of hypertriglyceridemia, Arterioscler. Thromb. Vasc. [233] Y. Huang, S.C. Rall, R.W. Mahley, Genetic factors precipitating hyperlipoproteinemia in
Biol. 31 (2011) 1916–1926. hypolipidemic transgenic mice expressing human apolipoprotein E2, Arterioscler.
[206] R.A. Hegele, M.R. Ban, N. Hsueh, B.A. Kennedy, H. Cao, G.Y. Zou, S. Anand, S. Thromb. Vasc. Biol. 17 (1997) 2817–2824.
Yusuf, M.W. Huff, J. Wang, A polygenic basis for four classical Fredrickson [234] A. Basaran, Pregnancy induced hyperlipoproteinemia. Review of the literature,
hyperlipoproteinemia phenotypes that are charcterized by hypertriglyc- Reprod. Sci. 5 (2009) 431–437.
eridemia, Hum. Mol. Genet. 18 (2009) 4189–4194. [235] D.T. Holmes, P. Long, J. Frolich, Dysbetalipoproteinemia and clomipramine, Am. J.
[207] C.T. Johansen, R.A. Hegele, Allelic and phenotypic spectrum of plasma triglycerides, Psychiatry 162 (2005) 1384–1385.
Biochim. Biophys. Acta 2012 (1821) 833–842. [236] B.P. Sinnott, T. Mazzone, Tuberous xanthomas associated with olanzapine therapy
[208] J.H. Lee, P. Giannikopoulos, S.A. Duncan, J. Wang, C.T. Johansen, J.D. Brown, J. and hypertriglyceridemia in the setting of a rare apolipoprotein E mutation,
Plutzky, R.A. Hegele, L.H. Glimcher, A. Lee, The transcription factor cyclic AMP- Endocr. Pract. 12 (2006) 183–187.
responsive element-binding protein H regulates triglyceride metabolism, Nat. [237] R.K. Lister, M. Youle, D.R. Nair, A.F. Winder, M.H.A. Rustin, Latent dysbetalipo-
Med. 17 (2011) 812–815. proteinemia precipitated by HIV-protease inhibitors, Lancet 353 (1999) 1678.
[209] P. Pajukanta, H.E. Lilja, R.M. Cantor, A.J. Lusis, M. Gentile, X.J. Duan, A. Soro- [238] P. Hennemann, F. van der Sman-de Beer, P.H. Moghaddam, P. Huijts, A.F.H.
Paavonen, J. Naukkarinen, J. Saarela, M. Laakso, et al., Familial combined Stalenhoef, K. JJP, C.M. van Duijn, L.M. Havekes, R.R. Frants, K.W. van Dijk, S.
hyperlipidemiais associated with upstream transcription factor (USF1), Nat. AHM, The expression of type 3 hyperlipoproteinemia: involvement of lipolysis
Genet. 36 (2004) 371–376. genes, Eur. J. Hum. Genet. 17 (2009) 620–628.
[210] M.C.G.J. Brouwers, M.M.J. van Greevenbroek, C.D.A. Stehouwer, J. de Graaf, A.F.H. [239] J.M. Martin-Campos, N. Rico, R. Bonet, C. Mayoral, Apolipoprotein A5 S19W may
Stalenhoef, The genetics of familial combined hyperlipidemia, Nat. Rev. Endocrinol. play a role in dysbetallipoproteinemia in patients with the apoE2/E2 genotype,
8 (2012) 352–362. Clin. Chem. 52 (2006) 1974–1975.
[211] L. Basel-Vanagaite, N. Zevit, A.H. Zahav, L. Guo, S. Parathat, M. Pasmanik-Chor, A.D. [240] Y. Hu, W. Liu, R. Huang, X. Zhang, A systematic review and meta-analysis of the re-
McIntyre, Transient infantile hypertriglyceridemia, fatty liver and hepatic fibrosis lationship between lipoprotein lipase Asn291Ser variant and diseases, J. Lipid Res.
caused by mutated GPD1, encoding glycerol-3-phosphate dehydrogenase 1, Am. 47 (2006) 1908–1914.
J. Hum. Genet. 90 (2012) 49–60. [241] H. Zhang, P.W.A. Reymer, M.S. Liu, I.J. Forsythe, B.E. Groenemeyer, J. Frohlich, J.D.
[212] M. Joshi, J. Eagan, N.K. Desai, S.A. Newton, M.C. Towne, N.S. Marinakis, K.M. Brunzell, J.J.P. Kastelein, M.R. Hayden, Y. Ma, Patients with apoE3 deficiency (E2/
Esteves, S. De Ferranti, et al., A compound heterozygous mutation in GPD1 2,E3/2 and E4/2)who manifest hyperlipidemia have increased frequency of an
causes hepatomegaly, steatohepatitis and hypertriglyceridemia, Eur. J. Hum. Asn → Ser mutation in the human LPL gene, Arterioscler. Thromb. Vasc. Biol. 15
Genet. (2014), http://dx.doi.org/10.1038/ejhg.2014.8. (1995) 1695–1703.
[213] P.W. Connelly, R.A. Hegele, Hepatic lipase deficiency, Crit. Rev. Clin. Lab. Sci. 35 [242] D. Evans, F.U. Beil, The D9N, N291S and S447X variants in the lipoprotein lipase
(1998) 547–572. (LPL) gene are not associated with type 3 hyperlipidemia, BMC Med. Genet. 8
[214] I.L. Ruel, P. Couture, C. Gagne, Y. Deshaies, J. Simard, R.A. Hegele, B. Lamarche, Char- (2007) 56.
acterization of a novel mutation causing hepatic lipase deficiency among French [243] G. Moennig, H. Wiebusch, A. Enbergs, A. Dorszewski, S. Kerber, H. Schulte, C.
Canadians, J. Lipid Res. 44 (2003) 1508–1514. Vielhauer, W. Haverkamp, G. Assmann, G. Breithardt, H. Funke, Detection of mis-
[215] V. Simha, A. Garg, Inherited lipodystrophies and hypertriglyceridemia, Curr. Opin. sense mutation in the genes for lipoprotein lipase and hepatic triglyceride lipase
Lipidol. 20 (2009) 300–308. in patients with dyslipidemia undergoing coronary angiography, Atherosclerosis
[216] A. Garg, A.K. Agarwal, Lipodystrophies: disorders of adipose tissue biology, 149 (2000) 395–401.
Biochim. Biophys. Acta 2009 (1791) 507–513. [244] M.R. Wardell, S.O. Brennan, E.D. Janus, R. Fraser, R.W. Carrell, Apolipoprotein E2-
[217] H. Saito, P. Dhanasekaran, F. Baldwin, K.H. Weisgraber, M.C. Phillips, S. Lund-Katz, Christchurch (136Arg → Ser). New variant of human apoliporotein E in a patient
Effect of polymorphism on the lipid interaction of human apolipoprotein E, J. Biol. with type 3 hyperlipoproteinemia, J. Clin. Invest. 80 (1987) 483–490.
Chem. 278 (2003) 40723–40729. [245] G. Feussner, M. Albanese, W.A. Mann, A. Valencia, H. Schuster, Apolipoprotein E2
[218] K.H. Weisgraber, T.L. Innerarity, R.W. Mahley, Abnormal lipoprotein receptor- (Arg136 → Cys), a variant of apolipoprotein E associated with late-onset domi-
binding activity of the human E apoprotein due to cysteine-arginine interchange nance of type 3 hyperlipoproteinemia, Eur. J. Clin. Investig. 26 (1996) 13–23.
at a single site, J. Biol. Chem. 257 (1982) 2518–2521. [246] A. Minnich, K.H. Weisgraber, Y. Newhouse, L. Dong, L.J. Fortin, M. Tremblay, J.
[219] C.P. Libeu, S. Lund-Katz, M.C. Phillips, S. Wehrli, M.J. Hernaiz, I. Capila, R.J. Linhardt, Davignon, Identification and characterization of a novel E3’ (Arg136 → His): associ-
R.L. Raffai, Y.M. Newhouse, F. Zhou, K.H. Weisgraber, New insights into the heparan ation with mild dyslipidemia and double pre-β very low density lipoproteins, J.
sulfate proteoglycan binding activity of apolipoprotein E, J. Biol. Chem. 276 (2001) Lipid Res. 36 (1995) 57–66.
39138–39144. [247] B. Vialettes, P. Reynier, C. Atlan-Gepner, N. Mekki, L. Lesluyes-Mazzocchi, G. Luc, D.
[220] K.H. Weisgraber, Apolipoprotein E distribution among human plasma lipoproteins: Lairon, Y. Malthiery, Dietary fat clearance in type V hyperlipoproteinemia second-
role of cysteine-arginine interchange at residue 112, J. Lipid Res. 31 (1990) ary to a rare variant of apolipoprotein E: apolipoprotein E3 (Arg 136 → Ser), Br. J.
1503–1511. Nutr. 83 (2000) 615–622.
[221] L. Krimbou, M. Denis, B. Haidar, M. Carrier, M. Marcil, J. Genest, Molecular interac- [248] M. Pocovi, A. Cenarro, F. Civeir, R.H. Myers, E. Casao, M. Esteban, J.M. Ordovas, In-
tions between apoE and ABCA1: impact on apoE lipidation, J. Lipid Res. 45 (2004) complete dominance of type 3 hyperlipoproteinemia is associated with the rare
839–848. apolipoprotein E2 (Arg136 → Ser) variant in multigenerational pedigree studies,
[222] A.R. Mensenkamp, M.C. Jong, H. van Goor, M.J.A. van Luyn, V. Bloks, R. Havinga, P.J. Atherosclerosis 122 (1996) 33–46.
Voshol, M.H. Hofker, K.W. van Dijk, L.M. Havekes, F. Kuipers, Apolipoprotein E par- [249] M. Rolleri, N. Vivona, G. Emmanuele, A.B. Cefalu, L. Pisciotta, V. Guido, D. Noto, B.
ticipates in the regulation of very low density lipoprotein-triglyceride secretion by Flore, C.M. Barbagello, A. Notarbartolo, S. Travali, S. Bertolini, M.R. Averna, Two Ital-
the liver, J. Biol. Chem. 274 (1999) 35711–35718. ian kindreds carrying the Arg136 → Ser mutation of the apoE gene: development
[223] Y. Huang, Z. Ji, W.J. Brecht, S.C. Rall, J.M. Taylor, R.W. Mahley, Overexpression of of premature severe atherosclerosis in the presence of epsilon 2 as a second allele,
apolipoprotein E3 in transgenic rabbits causes combined hyperlipidemia by stim- Nutr. Metab. Cardiovasc. Dis. 13 (2003) 93–99.
ulating hepatic VLDL production and impairing VLDL lipolysis, Arterioscler. [250] W. Marz, M.M. Hoffmann, H. Scharnagl, E. Fisher, M. Chen, M. Nauck, G. Feussner,
Thromb. Vasc. Biol. 19 (1999) 2952–2959. H. Wieland, Apolipoprotein E2(Arg136 → Cys) mutation in the receptor binding do-
[224] H. Li, P. Dhanasekaran, E.T. Alexander, D.J. Rader, M.C. Phillips, S. Lund-Katz, Molec- main of apoE is not associated with dominant type III hyperlipoproteinemia, J.
ular mechanisms responsible for the differential effects of apoE3 and apoE4 on Lipid Res. 39 (1998) 658–659.
plasma lipoprotein cholesterol levels, Arterioscler. Thromb. Vasc. Biol. 33 (2013) [251] P. Lohse, A.W. Mann, E.A. Stein, H.B. Brewer, Apolipoprotein E-4Philadelphia (Glu13 →
687–693. Lys,Arg145 → Cys). Homozygosity for two rare point mutations in the apolipopro-
[225] P.C.R. Hopkins, Y. Huang, J.G. McGuire, R.E. Pitas, Evidence for differential effects of tein E gene combined with severe type III hyperlipoproteinemia, J. Biol. Chem.
apoE3 and apoE4 on HDL metabolism, J. Lipid Res. 43 (2002) 1881–1889. 266 (1991) 10479–10484.
[226] A.M. Bennett, E. Di Angelantonio, Z. Ye, F. Wensley, A. Dahlin, A. Ahlbom, B. [252] Z. Ji, S. Fazio, R.W. Mahley, Variable heparan sulfate proteoglycan binding of apolipo-
Keavney, R. Collins, B. Wiman, U. de Faire, J. Danesh, Association of apolipoprotein protein E variants may modulate the expression of type III hyperlipoproteinemia, J.
E genotypes with lipid levels and coronary risk, JAMA 298 (2007) 1300–1311. Biol. Chem. 269 (1994) 13421–13428.
[227] C.R. Isasi, S. Shea, R.J. Deckelbaum, S.C. Couch, T.J. Starc, J.D. Otvos, L. Berglund, [253] P. Richard, M.P. de Zulueta, I. Beucler, J. de Gennes, A. Cassaigne, A. Iron, Identifica-
Apolipoprotein ε2 allele is associated with an anti-atherogenic lipoprotein tion of a new apolipoprotein E variant (E2Arg142 → Leu) in type III hyperlipidemia,
profile in children. The Columbia University BioMarkers study, Pediatrics Atherosclerosis 112 (1995) 19–28.
568-575 (2000). [254] S.C. Rall, Y.M. Newhouse, H.R.G. Clarke, K.H. Weisgraber, B.J. McCarthy, R.W.
[228] Y. Huang, X.Q. Liu, R.C. Rall, R.W. Mahley, Apolipoprotein E2 reduces the low den- Mahley, T.P. Bersot, Type III hyperlipoproteinemia associated with apolipoprotein
sity lipoprotein level in transgenic mice by impairing lipoprotein lipase-mediated E phenotype E3/3. Structure and genetics of an apolipoprotein E3 variant, J. Clin.
lipolysis of triglyceride-rich lipoproteins, J. Biol. Chem. 273 (1998) 17483–17490. Invest. 83 (1989) 1095–1101.
[229] J. Heeren, U. Beisiegel, T. Grewal, Apolipoprotein E recycling. Implications for dys- [255] M. Smit, P. de Knijff, E. van der Kooij-Meijs, C. Groenendijk, A.M.J.M. van den
lipidemia and atherosclerosis, Arterioscler. Thromb. Vasc. Biol. 26 (2006) 442–448. Maagdenberg, J.A.G. Leuven, A.F.H. Stalenhoef, P.M.J. Stuyt, R.R. Frants, L.M.
178 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

Havekes, Genetic heterogeneity in familial dysbetalipoproteinemia. The [282] K. Ito, H. Nakashima, M. Watanabe, A. Ishimura, Y. Miyahara, Y. Abe, T. Yasuno, M.
E2(lys146 → gln) variant results in a dominant mode of inheritance, J. Lipid Res. Ifuku, Y. Sasatomi, T. Saito, Macrophage impairment produced by Fc receptor
31 (1990) 45–53. gamma deficiency plays a principal role in the development of lipoprotein glomer-
[256] W.A. Mann, P. Lohse, R.E. Gregg, R. Ronan, J.M. Hoeg, L.A. Zech, H.B. Brewer Jr., ulopathy in concert with apoE abnormalities, Nephrol. Dial. Transplant. 27 (2012)
Dominant expression of type III hyperlipoproteinemia. Pathophysiological insights 3899–3907.
derived from the structural and kinetic characteristics of ApoE-1 (Lys146–> [283] B. Zhang, Z. Liu, C. Zeng, J. Zheng, H. Chen, L. Li, Clinicopathological and genetic
Glu), J. Clin. Invest. 96 (1995) 1100–1107. characteristics in Chinese patients with lipoprotein glomerulopathy, J. Nephrol.
[257] M.J.V. Hoffer, S. Nithyanathan, R.P. Naoumova, M.S. Kibirige, R.R. Frants, L.M. 21 (2008) 110–117.
Havekes, G.R. Thompson, Apolipoprotein W1-Hammersmith(Lys146 → Asn; [284] Z. Hu, S. Huang, Y. Wu, Y. Liu, X. Liu, D. Su, Y. Tao, P. Fu, X. Zhang, Z. Peng, S. Zhang,
Arg147 → Trp) due to a dinucleotide substitution, is associated with early manifes- Y. Yang, Hereditary features, treatment, and prognosis of the lipoprotein glomeru-
tation of type III hyperlipoproteinemia, Atherosclerosis 124 (1996) 183–189. lopathy in patients with APOE Kyoto mutation, Kidney Int. 85 (2014) 416–424.
[258] P. Fotakis, A. Vezeridis, I. Dafnis, A. Chroni, D. Kardassis, V.I. Zannis, apoE3[K146N/ [285] M. Ando, J. Sasaki, H. Hua, A. Matsunaga, K. Uchida, K. Jou, S. Oikawa, T. Saito, H.
R147W] acts as a dominant negative apoE form that prevents remnany clearance Nihei, A novel 18-amino acid deletion in apolipoprotein E associated with lipopro-
and inhibits the biogenesis of HDL, J. Lipid Res. 55 (2014) 1310–1323. tein glomerulopathy, Kidney Int. 56 (1999) 1317–1323.
[259] A.M. Vezeridis, K. Drosatos, V.I. Zannis, Molecular etiology of a dominant form of [286] B. Luo, F. Huang, Q. Liu, X. Li, W. Chen, S. Zhou, X. Yu, Identification of apolipopro-
type III hyperlipoproteinemia caused by R142C substitution in apoE4, J. Lipid Res. tein E Guangzhou (Arginine 150 Proline), a new variant associated with lipopro-
52 (2011) 45–56. tein glomerulopathy, Am. J. Nephrol. 28 (2008) 347–353.
[260] W.J.S. De Villiers, D.R. van der Westhuyzen, G.A. Coetzee, H.E. Henderson, D.A. [287] M. Kinomura, H. Sugiyama, T. Saito, A. Matsunaga, K. Sada, M. Kanzaki, Y.
Marais, The apolipoprotein E2(Arg145Cys) mutation causes autosomal dominant Takazawa, Y. Maeshima, H. Yanai, H. Makino, A novel variant apolipoprotein E
type III hyperlipoproteinemia with incomplete penetrance, Arterioscler. Thromb. Okayama in a patient with lipoprotein glomerulopathy, Nephrol. Dial. Transplant.
Vasc. Biol. 17 (1997) 865–872. 23 (2008) 751–756.
[261] P. De Knijff, A.M.J.M. van den Maagdenberg, A.F.H. Stalenhoef, J.A.G. Leuven, P.N.M. [288] M. Hagiwara, K. Yamagata, T. Matsunaga, Y. Arakawa, J. Usui, Y. Shimizu, K. Aita, M.
Demacker, L.P. Kuyt, R.R. Frants, L.M. Havekes, Familial dysbetalipoproteinemia as- Nagata, A. Koyama, B. Zhang, A. Mastunga, K. Saku, T. Saito, A novel apolipoprotein
sociated with apolipoprotein E3-Leiden in an extended multigeneration pedigree, E mutation, ApoE Tsukuba (Arg114Cys) in lipoprotein glomerulopathy, Nephrol.
J. Clin. Invest. 88 (1991) 643–655. Dial. Transplant. 23 (2008) 381–384.
[262] C. Cladaras, M. Hadzopoulou-Cladaras, B.K. Felber, G. Pavlakis, V.I. Zannis, The mo- [289] T. Tokura, S. Itano, S. Kobayashi, A. Kuwabara, S. Fujimoto, H. Horike, M. Satoh, N.
lecular basis of apoE deficiency, J. Biol. Chem. 262 (1987) 2310–2315. Komai, N. Tomita, A. Matsunaga, T. Saito, T. Sasaki, N. Kashihara, A novel mutation
[263] E.J. Schaefer, R.E. Gregg, G. Ghiselli, T.M. Forte, J.M. Ordovas, L.A. Zech, H.B. Brewer, apoE2 Kurashiki (R158P) in a patient with lipoprotein glomerulopathy, J.
Familial apolipoprotein E deficiency, J. Clin. Invest. 78 (1986) 1206–1219. Atheroscler. Thromb. 18 (2011) 536–541.
[264] K. Ikewaki, W. Cain, F. Thomas, R. Shamburek, L.A. Zech, D. Usher, H.B. Brewer, D.J. [290] T. Saito, A. Matsunaga, Significance of a novel apolipoprotein E variant, apoE Osaka/
Rader, Abnormal in vivo metabolism of apoB-containing lipoproteins in apoE defi- Kurashiki in lipoprotein glomerulopathy, J. Atheroscler. Thromb. 18 (2011)
ciency, J. Lipid Res. 45 (2004) 1302–1311. 542–543.
[265] D. Kurosaka, T. Teramoto, T. Matsushima, T. Yokoyama, A. Yamada, T. Aikawa, Y. [291] T. Miyata, S. Sugiyama, M. Nangaku, D. Suzuki, K. Uragami, R. Inagi, H. Sakai, K.
Miyamoto, K. Kurokawa, Apolipoprotein E deficiency with depressed mRNA of Kurokawa, Apolipoprotein E2/E5 variants in lipoprotein glomerulopathy recurred
normal size, Atherosclerosis 88 (1991) 15–20. in transplanted kidney, J. Am. Soc. Nephrol. 10 (1999) 1590–1595.
[266] H. Mabuchi, H. Itoh, M. Takeda, K. Kajinami, T. Wakasugi, J. Koizumi, R. Takeda, C. [292] V. Tsimihodimos, M. Elisaf, Lipoprotein glomerulopathy, Curr. Opin. Lipidol. 22
Asagami, A young type III hyperlipoproteinemic patient associated with apolipo- (2011) 262–269.
protein E deficiency, Metabolism 38 (1989) 115–119. [293] R.J. Neuman, B. Yuan, D.S. Gerhard, K. Liu, P. Yue, S. Duan, M. Averna, G. Schonfeld,
[267] P. Lohse, H.B. Brewer, M.S. Meng, S.I. Skarlatos, L.R. JC, H.B. Brewer, Familial apoli- Replication of linkage of familial hypobetalipoproteinemia to chromosome 3p in
poprotein E deficiency and type III hyperlipoproteinemia due to a premature stop six kindreds, J. Lipid Res. 43 (2002) 407–415.
codon in the apolipoprotein E gene, J. Lipid Res. 33 (1992) 1583–1590. [294] H. Knoblauch, A cholesterol lowering gene maps to chromosome 13q, Am. J. Hum.
[268] G. Feussner, Severe xanthomatosis associated with familial apolipoprotein E defi- Genet. 66 (2000) 157–166.
ciency, J. Clin. Pathol. 49 (1996) 985–989. [295] F.K. Welty, C. Lahoz, K.L. Tucker, J.M. Ordovas, P.W. Wilson, E.J. Schaefer, Frequency
[269] G. Feussner, H. Funke, W. Weng, G. Assman, K.J. Lackner, R. Ziegler, Severe type III of ApoB and ApoE gene mutations as causes of hypobetalipoproteinemia in the
hyperlipoproteinemia associated with unusual apolipoprotein E1 phenotype and Framingham offspring population, Arterioscler. Thromb. Vasc. Biol. 18 (1998)
epsilon 1/‘null’ genotype, Eur. J. Clin. Investig. 22 (1992) 599–608. 1745–1751.
[270] D.A. Dijck-Brouwer, J.J. van Doormaal, I.P. Kema, A.M. Brugman, A.W. Kingma, F.A. [296] S. Rashid, D.E. Curtis, R. Garuti, N.N. Anderson, Y. Bashmakov, Y.K. Ho, R.E. Hammer,
Muskiet, Discovery and consequences of apolipoprotein-epsilon (3Groningen): a Y.A. Moon, J.D. Horton, Decreased plasma cholesterol and hypersensitivity to
G-insertion in codon 95/96 that is predicted to cause a premature stop codon, statins in mice lacking Pcsk9, Proc. Natl. Acad. Sci. 102 (2005) 5374–5379.
Ann. Clin. Biochem. 42 (2005) 264–268. [297] J.C. Cohen, E. Boerwinkle, T.H. Mosely, H.H. Hobbs, Sequence variations in PCSK9,
[271] J.R. Tate, M.M. Hoffmann, P.K. Lovelock, J.B. Kesting, J.T. Shaw, Identification of an low LDL and protection against coronary heart disease, N. Engl. J. Med. 354
apolipoprotein(e) variant associated with type III hyperlipoproteinemia in an in- (2006) 1264–1272.
digenous Australian, Ann. Clin. Biochem. 38 (2001) 46–53. [298] M. Benn, B.G. Nordestgaard, P. Grande, P. Schnohr, A. Tybjaerg-Hansen, PCSK9
[272] A.D. Marais, G.A.E. Solomon, D.J. Blom, Dysbetalipoproteinemia: a mixed hyperlip- R46L, low-density lipoprotein cholesterol levels and risk of ischemic heart disease:
idemia of remnant lipoproteins due to mutations in apolipoprotein E, Crit. Rev. 3 independent studies and meta-analyses, J. Am. Coll. Cardiol. 55 (2010)
Clin. Lab. Sci. 51 (2014) 46–62. 2833–2842.
[273] T. Wang, K. Nakajima, E.T. Leary, G.R. Warnick, J.S. Cohn, P.N. Hopkins, L.L. Wu, D.D. [299] J. Cohen, A. Pertsemlidis, I.K. Kotowski, R. Graham, K.G. Garcia, H.H. Hobbs, Low LDL
Cilla, J. Zhong, R.J. Havel, Ratio of remnant-like particle cholesterol to serum total cholesterol in individuals of African descent resulting from frequent nonsense mu-
triglycerides is an effective alternative to ultracentrifugal and electrophoretic tations in PCSK9, Nat. Genet. 37 (2005) 161–165.
methods in the diagnosis of familial type III hyperlipoproteinemia, Clin. Chem. 45 [300] K.E. Berge, L. Ose, T.P. Leren, Missense mutations in the PCSK9 gene are associated
(1999) 1981–1987. with hypocholesterolemia and possibly increased response to statin therapy,
[274] A. Sniderman, A. Tremblay, J. Bergeron, C. Gagne, P. Coutre, Diagnosis of type III Arterioscler. Thromb. Vasc. Biol. 26 (2006) 1094–1100.
hyperlipoproteinemia from plasma total cholesterol, triglyceride and apolipopro- [301] I.K. Kotowski, A. Pertsemlidis, A. Luke, R.S. Cooper, G.L. Vega, J.C. Cohen, H.H. Hobbs,
tein B, J. Clin. Lipidol. 1 (2007) 256–263. A spectrum of PCSK9 alleles contributes to plasma levels of low-density lipoprotein
[275] D.J. Blom, F.H. O'Neill, D.A. Marais, Screening for dysbetalipoproteinemia by plasma cholesterol, Am. J. Hum. Genet. 78 (2006) 410–422.
cholesterol and apolipoprotein B concentration, Clin. Chem. 51 (2005) 904–907. [302] J. Cameron, O.L. Holla, T. Ranheim, M.A. Kulseth, K.E. Berge, T.P. Leren, Effect of mu-
[276] T. Saito, A. Matsunaga, Lipoprotein glomerulopathy may provide a key to unlock tations in the PCSK9 gene on the cell surface LDL receptors, Hum. Mol. Genet. 15
the puzzles of renal lipidosis, Kidney Int. 85 (2014) 243–245. (2006) 1551–1558.
[277] S. Oikawa, A. Matsunaga, T. Saito, H. Sato, T. Seki, K. Hoshi, et al., Apolipoprotein E [303] F. Lalanne, G. Lambert, M.J. Amar, M. Chetiveaux, Y. Zair, A.L. Jarnoux, K. Ougerram,
Sendai (Arginine 145→Proline): a new variant associated with lipoprotein glomer- J. Friburg, N.G. Seidah, H.B. Brewer, M. Kempf, P. Costet, Wild type PCSK9 inhibits
ulopathy, J. Am. Soc. Nephrol. 8 (1997) 820–823. LDL clearance but does not affect apoB-containg lipoprotein production in mouse
[278] A. Matsunaga, J. Sasaki, T. Komatsu, K. Kanatsu, E. Tsuji, K. Moriyama, et al., A novel and cultured cells, J. Lipid Res. 46 (2005) 1312–1319.
apolipoprotein E mutation, E2 (Arg25Cys), in lipoprotein glomerulopathy, Kidney [304] T. Fasano, A.B. Cefalu, E. Di Leo, D. Pollaccia, L. Bocchi, V. Valenti, R. Bonardi, O.
Int. 56 (1999) 421–427. Guardamagna, M. Averna, P. Tarugi, A novel loss of mutation of PCSK9 gene in
[279] Y. Ishigaki, S. Oikawa, T. Suzuki, S. Usui, K. Magoori, D. Kim, H. Suzuki, J. Sasaki, white subjects with low-plasma low-density lipoprotein cholesterol, Arterioscler.
et al., Virus mediated transduction of apolipoprotein E (ApoE) Sendai develops li- Thromb. Vasc. Biol. 27 (2007) 677–681.
poprotein glomerulopathy in apoE-deficient mice, J. Biol. Chem. 275 (2000) [305] Y. Miyake, R. Kimura, Y. Kokubo, A. Okayama, H. Tomoike, T. Yamamura, T. Miyata,
31269–31273. Genetic variants in PCSK9 in the Japanese population: rare genetic variants in
[280] M.M. Hoffmann, H. Scharnagl, E. Panagiotou, W.T. Banghard, H. Wieland, W. PCSK9 might collectively contribute to plasma LDL cholesterol levels in the general
Marz, Diminshed LDL receptor and high heparin binding of apolipoprotein E2 population, Atherosclerosis 196 (2008) 29–36.
Sendai associated with lipoprotein glomerulopathy, J. Am. Soc. Nephrol. 12 [306] J. Mayne, T.C. Ooi, A. Raymond, M. Cousins, L. Bernier, T. Dewpura, F. Sirois, M.
(2001) 524–530. Mbikay, J. Davignon, M. Chretien, Differential effects of PCSK9 loss of function var-
[281] D. Georgiadou, K. Stamatakis, E.K. Efthimiadou, G. Kordas, D. Gantz, A. Chroni, E. iants on serum lipid and PCSK9 levels in Caucasian and African Canadian popula-
Stratikos, Thermodynamic and structural destabilization of apoE3 by hereditary tion, Lipids Health Dis. 12 (2013) 70.
mutations associated with the development of lipoprotein glomerulopathy, J. [307] J. Mayne, T. Dewpura, A. Raymond, L. Bernier, M. Cousins, T.K. Ooi, J. Davignon, N.G.
Lipid Res. 54 (2013) 164–176. Seidah, M. Mbikay, M. Chretien, Novel loss-of-function PCSK9 variant is associated
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 179

with low plasma LDL cholesterol in a French-Canadian family and with impaired McKnight, Z. Yao, Missense mutation in APOB within the βα1 domain of
processing and secretion in cell culture, Clin. Chem. 57 (2011) 1415–1423. apoB-100 result in impaired secretion of apoB and apoB-containing lipoproteins
[308] B. Cariou, K. Ougerram, Y. Zair, R. Guerois, C. Langhi, S. Kourimate, I. Benoit, C. Le in familial hypobetalipoproteinemia, J. Biol. Chem. 282 (2007) 24270–24283.
May, C. Gayet, K. Belabbas, F. Dufernez, M. Chetiveaux, P. Tarugi, M. Krempf, P. [331] D. Noto, A.B. Cefalu, A. Cannizzaro, M. Mina, F. Fayer, V. Valenti, C.M.
Benlian, P. Costet, PCSK9 dominant negative mutant results in increased LDL cata- Barbagallo, A. Tuttolomondo, A. Pinto, C. Sciume, G. Licata, M. Averna, Familial
bolic rate and familial hypobetalipoproteinemia, Arterioscler. Thromb. Vasc. Biol. hypobetalipoproteinemia due to apolipoprotein B r463W mutation causes in-
29 (2009) 2191–2197. testinal fat accumulation and low postprandial lipemia, Atherosclerosis 206
[309] Z. Zhao, Y. Tuakli-Wosornu, T.A. Lagace, L. Kinch, N.V. Grishin, J.D. Horton, J.C. (2009) 193–198.
Cohen, H.H. Hobbs, Molecular characterization of loss-of-function mutations in [332] S. Zhong, A.L. Magnolo, M. Sundaram, H. Zhou, E.F. Yao, E. Di Leo, P. Loria, S.
PCSK9 and identification of a compound heterozygote, Am. J. Hum. Genet. 79 Wang, M. Bamji-Mirza, L. Wang, C.J. McKnight, D. Figeys, Y. Wang, P. Tarugi, Z. Yao,
(2006) 514–523. Nonsynonymous mutations with APOB in human familial hypobetalipoproteinemia,
[310] S. Benjannet, J. Hamelin, M. Chretien, N.G. Seidah, Loss and gain of function PCSK9 J. Biol. Chem. 285 (2010) 6453–6464.
variants, J. Biol. Chem. 287 (2012) 33745–33755. [333] S.W. Fouchier, R.R. Sankatsing, J. Peter, S. Castillo, M. Pocovi, R. Alonso, J.J.P.
[311] S. Poirier, G. Mayer, S. Benjannet, E. Bergeron, J. Marcinkiewicz, N. Nassoury, H. Kastelein, J.C. Defesche, High frequency of apoB gene mutations causing familial
Mayer, J. Nimpf, A. Prat, N.G. Seidah, The proprotein convertase PCSK9 induces hypobetalipoproteinemia in patients of Dutch and Spanish origin, J. Med. Genet.
the degradation of low density lipoprotein receptor (LDLR) and its closese family 42 (2005), e23.
members VLDR and apoER2, J. Biol. Chem. 283 (2008) 2363–2372. [334] P. Tarugi, A. Lonardo, C. Gabelli, F. Sala, G. Ballarini, I. Cortella, L. Previato, S. Bertolini,
[312] M. Canuel, X. Sun, M.C. Asselin, E. Paramithiotis, A. Prat, N.G. Seidah, Proprotein R. Cordera, S. Calandra, Phenotypic expression of familial hypobetalipoproteinemia
convertase subtilisin/Kexin type 9 (PCSK9) can mediate degradation of the low in three kindreds with mutations of apolipoprotein B gene, J. Lipid Res. 42 (2001)
density lipoprotein receptor-related protein 1 (LRP-1), PLoS One 8 (2013), e64145. 1552–1561.
[313] A. Roubtsova, M.N. Munkonda, Z. Awan, J. Marcinkiewicz, A. Chamberland, C. [335] R.R. Sankatsing, S.W. Fouchier, S. de Haan, B.A. Hutten, E. de Groot, J.J.P. Kastelein, E.S.G.
Lazure, K. Cianflone, N.G. Seidah, A. Prat, Circulating proprotein convertase subtil- Stroes, Hepatic and cardiovascular consequences of familial hypobetalipoproteinemia,
isin/kexin 9 (PCSK9) regulates VLDLR protein and triglyceride accumulation in vis- Arterioscler. Thromb. Vasc. Biol. 25 (2005) 1979–1984.
ceral adipose tissue, Arterioscler. Thromb. Vasc. Biol. 31 (2011) 785–791. [336] M.M. Hussain, A. Bakillah, N. Nayak, G.S. Shelness, Amino acids 430–570 in apoli-
[314] N.G. Seidah, M.S. Sadr, M. Chretien, M. Mbikay, The multifaceted proprotein poprotein B are critical for its binding to microsomal triglyceride transfer protein,
convertases: their unique, redundant, complementary, and opposite functions, J. J. Biol. Chem. 273 (1998) 25612–25615.
Biol. Chem. 288 (2013) 21473–21481. [337] P. Bradbury, C.J. Mann, S. Kochi, T.A. Anderson, S.A. Chester, J.M. Hancock, P.J.
[315] I. Postmus, S. Trompet, A.J.M. de Craen, B.M. Buckley, I. Ford, D.J. Stott, N. Sattar, Ritchie, J. Amey, G.B. Harrison, D.G. Levitt, L.J. Banaszak, J. Scott, C.C. Shoulders, A
P.E. Slagboom, R.G.J. Westendorp, J.W. Jukema, PCSKP9 SNP rs11591147 is as- common binding site on the microsomal triglyceride transfer protein for apolipo-
sociated with low cholesterol levels but not with cognitive performance or protein B and protein disulfide isomerase, J. Biol. Chem. 274 (1999) 3159–3164.
noncardiovascular clinical events in an elderly population, J. Lipid Res. 54 [338] C.J. Mann, T.A. Anderson, J. Read, S.A. Chester, G.B. Harrison, S. Kochi, P.J. Ritchie, P.
(2013) 561–566. Bradbury, F.S. Hussain, J. Amey, B. Vanloo, M. Rosseneu, R. Infante, J.M. Hancock,
[316] B. Han, M. Eacho, M.D. Knierman, J.S. Troutt, R.J. Konrad, X. Yu, K.M. Schroeder, Iso- D.G. Levitt, L.J. Banaszak, J. Scott, C.C. Shoulders, The structure of vitellogenin pro-
lation and characterization of the circulating truncated form of PCSK9, J. Lipid Res. vides a molecular model for the assembly and secretion of atherogenic lipopro-
55 (2014) 1505–1514. teins, J. Mol. Biol. 285 (1999) 391–408.
[317] W.E. Alborn, G. Cao, H.E. Careskey, Y. Qian, D.R. Subramaniam, J. Davies, E.M. [339] K.G. Parhofer, P.H.R. Barrett, C.A. Aguilar-Salinas, G. Schonfeld, Positive linear corre-
Conner, R.J. Konrad, Serum proprotein convertase subtilisin kexin type 9 is corre- lation between the length of the truncated apolipoprotein B and its secretion rate:
lated directly with serum LDL cholesterol, Clin. Chem. 53 (2007) 1814–1819. in vivo studies in human apoB-89, apoB-75, apoB-54.8 and apoB-31 heterozygotes,
[318] E. Chernogubova, R. Strawbridge, H. Mahdessian, A. Malarstig, S. Krapivner, B. J. Lipid Res. 37 (1996) 844–852.
Gigante, M.L. Hellenius, U. de Faire, A. Franco-Cereceda, A.C. Syvanen, J.S. Troutt, [340] K.G. Parhofer, P.H.R. Barrett, D.M. Bier, G. Schonfeld, Lipoproteins containing the
R.J. Konrad, P. Eriksson, A. Hamsten, F.M. van't Hooft, Common and low frequency truncated apolipoprotein, Apo-B-89 are cleared from human plasma more rapidly
genetic variants in the PCSK9 locus influence circulating PCSK9 levels, Arterioscler. than apo B-100-containing lipoproteins in vivo, J. Clin. Invest. 89 (1992)
Thromb. Vasc. Biol. 32 (2012) 1526–1534. 1931–1937.
[319] L. Persson, G. Cao, L. Stahle, B.G. Sjoberg, J.S. Troutt, R.J. Konrad, C. Galman, H. [341] Z. Yao, B.D. Blackhart, M.F. Linton, S.M. Taylor, S.G. Young, B.J. McCarthy, Expression
Wallen, M. Eriksson, I. Hafstrom, S. Lind, M. Dahlin, P. Amark, B. Angelin, M. of carboxyl-terminally truncated forms of human apolipoprotein B in rat hepatoma
Rudling, Circulating proprotein convertase subtilisin kexin type 9 has a diurnal cells, J. Biol. Chem. 266 (1991) 3300–3308.
rhythm synchronous with cholesterol synthesis and is reduced by fasting in [342] N. Elias, B.W. Patterson, G. Schonfeld, Decreased production rates of VLDL triglycer-
humans, Arterioscler. Thromb. Vasc. Biol. 30 (2010) 2666–2672. ides and apoB-100 in subjects heterozygous for familial hypobetalipoproteinemia,
[320] S.E. Humphries, R.D.G. Neely, R.A. Whittall, J.S. Troutt, J.S. Konrad, M. Scartezini, Arterioscler. Thromb. Vasc. Biol. 19 (1999) 2714–2721.
K.W. Li, J.A. Cooper, J. Acharya, A. Neil, Healthy individuals carrying the PCSK9 [343] K.G. Parhofer, A. Daugherty, M. Kinoshita, G. Schonfeld, Enhanced clearance from
p.R46L variant and familial hypercholesterolemia patients carrying PCSK9 p.D374Y plasma of low density lipoproteins containing a truncated apolipoprotein apoB-
exhibit lower plasma concentrations of PCSK9, Clin. Chem. 55 (2009) 2153–2161. 89, J. Lipid Res. 31 (1990) 2001–2007.
[321] G. Dubuc, M. Tremblay, G. Pare, H. Jacques, J. Hamelin, S. Benjannet, L. Boulet, J. [344] X. Zhu, D. Noto, R. Selp, A. Shaish, G. Schonfeld, Organ loci of catabolism of short
Genest, L. Bernier, N.G. Seidah, J. Davignon, A new method for measurement of truncations of apoB, Arterioscler. Thromb. Vasc. Biol. 17 (1997) 1032–1038.
total plasma PCSK9: clinical applications, J. Lipid Res. 51 (2010) 140–149. [345] G. Schonfeld, B.W. Patterson, D.A. Yablonskiy, T.S.K. Tanoli, M. Averna, N. Elias, P.
[322] L. Zhang, K. Reue, L.G. Fong, S.G. Young, P. Tontonoz, Feedback regulation of choles- Yue, J. Ackerman, Fatty liver in familial hypobetalipoproteinemia: triglyceride as-
terol uptake by the LXR-IDOL-LDLR axis, Arterioscler. Thromb. Vasc. Biol. 32 (2012) sembly into VLDL particles is affected by the extent of hepatic steatosis, J. Lipid
2541–2546. Res. 44 (2003) 470–478.
[323] E. Scotti, C. Hong, Y. Yoshinga, Y. Tu, Y. Hu, N. Zelcer, R. Boyadjian, P.J. de Jong, S.G. [346] A.B. Cefalu, J.P. Pirruccello, D. Noto, S. Gabriel, V. Valenti, N. Gupta, R. Spina, P.
Young, L.G. Fong, P. Tonotonoz, Targeted disruption of the Idol gene alters cellular Tarugi, S. Kathiresan, M.R. Averna, A novel APOB mutation identified by exome se-
regulation of the low density lipoprotein receptor by sterols and liver X receptor quencing cosegregates with steatosis, liver cancer and hypocholesterolemia,
agonists, Mol. Cell. Biol. 31 (2011) 1885–1893. Arterioscler. Thromb. Vasc. Biol. 33 (2013) 2021–2025.
[324] V. Sorrentino, S.W. Fouchier, M.M. Motazacker, J.K. Nelson, J.C. Defesche, G.M. [347] A.J. Whitfield, H.R. Barrett, K. Robertson, M.F. Havlat, F.M. van Bockxmeer, J.R.
Dallinga-Thie, J.J.P. Kastelein, G.K. Hovingh, N. Zelcer, Identification of a loss-of- Burnett, Liver dysfunction and steatosis in familial hypobetalipoproteinemia,
function inducible degrader of the low density lipoprotein receptor variant in indi- Clin. Chem. 51 (2005) 266–269.
viduals with low circulating low density lipoprotein, Eur. Heart J. 34 (2013) [348] P. Yue, W.L. Isley, W.S. Harris, S. Rosipal, C.D. Akin, G. Schonfeld, Genetic variants of
1292–1297. apoE account for variability of plasma low density lipoprotein and apolipoprotein
[325] R.D. Wagner, E.S. Krul, J. Tang, K.G. Parhofer, K. Garlock, P. Talmud, G. Schonfeld, levels in in FHBL, Atherosclerosis 178 (2005) 107–113.
ApoB-54.8, a truncated apolipoprotein found primarily in VLDL is associated with [349] L.S. Huang, H. Kayden, R.J. Sokol, J.L. Breslow, ApoB gene nonsense and splicing mu-
a nonsense mutation in the apoB gene and hypolipoproteinemia, J. Lipid Res. 32 tations in a compound heterozygote for familial hypobetalipoproteinemia, J. Lipid
(1991) 1001–1011. Res. 32 (1991) 1341–1348.
[326] E.S. Krul, K.G. Parhofer, H.R. Barrett, R.D. Wagner, G. Schonfeld, ApoB-75 a trunca- [350] S.G. Young, B. Bihaln, L.M. Flynn, D.A. Sanan, M. Ayrault-Jarrier, B. Jacotot, Asymp-
tion of apolipoprotein B associated with familial hypobetalipoproteinemia genetic tomatic homozygous hypobetalipoproteinemia associated with apolipoprotein
and kinetic studies, J. Lipid Res. 33 (1992) 1037–1050. B45.2, Hum. Mol. Genet. 3 (1994) 741–744.
[327] K. Ohashi, S. Ishibashi, M. Yamamoto, J. Osuga, Y. Yazaki, S. Yukawa, N. Yamada, A [351] M.M. Hussain, J. Shi, P. Dreizen, Microsomal triglyceride transfer protein and its
truncated species of apolipoprotein B (B-38.7) in a patient with homozygous role in apoB-lipoprotein assembly, J. Lipid Res. 44 (2003) 22–32.
hypobetalipoproteinemia associated with diabetes mellitus, Arterioscler. Thromb. [352] J. Wang, R.A. Hegele, Microsomal triglyceride transfer protein (MTP) gene mutations
Vasc. Biol. 18 (1998) 1330–1334. in Canadian subjects with abetalipoproteinemia, Hum. Mutat. 297 (1999) 1–4.
[328] E. Di Leo, L. Magnolo, S. Lancellotti, L. Croce, L. Visintin, C. Tribelli, S. Bertolini, S. [353] M. Raabe, L.M. Flynn, C.H. Zlot, J.S. Wong, M.M. Veniant, R.L. Hamilton, S.G. Young,
Calandra, P. Tarugi, Abnormal apolipoprotein B pre-mRNA spicing in patients Knockout of the abetalipoproteinemia gene in mice: reduced lipoprotein secretion
with familial hypobetalipoproteinemia, J. Med. Genet. 44 (2007) 219–224. in heterozygotes and embryonic lethality in homozygotes, Proc. Natl. Acad. Sci. 95
[329] J.R. Burnett, J. Shan, B.A. Miskie, A.J. Whitfield, J. Yuan, K. Tran, C.J. McKnight, R.A. (1998) 8686–8691.
Hegele, Z. Yao, A novel nontruncating APOB gene mutation R463W causes familial [354] Y. Xie, E.P. Newberry, S.G. Young, S. Robine, R.L. Hamilton, J.S. Wong, J. Luo, S.
hypobetalipoproteinemia, J. Biol. Chem. 278 (2003) 13442–13452. Kennedy, N.O. Davidson, Compensatory increase in hepatic lipogenesis in mice
[330] J.R. Burnett, S. Zhong, Z.G. Jiang, A.J. Hooper, E.A. Fisher, R.S. McLeod, Y. Zhao, with conditional intestine specific Mttp deficiency, J. Biol. Chem. 281 (2006)
P.H.R. Barrett, R.A. Hegele, F.M. van Bockxmeer, H. Zhang, D.E. Vance, C.J. 4075–4086.
180 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

[355] V. Pons, C. Rolland, M. Nauze, M. Danjoux, G. Gaibelet, A. Durandy, A. Sassolas, E. mutations in subjects with combined hypolipidemia, Arterioscler. Thromb. Vasc.
Levy, F. Terce, X. Collet, A severe form of abetalipoproteinemia caused by new Biol. 32 (2012) 805–809.
splicing mutations of microsomal triglyceride transfer protein (MTTP), Hum. [381] M.R. Robciuc, M. Maranghi, A. Lahikainin, D. Rader, A. Bensadoun, K. Oorni, J.
Mutat. 32 (2011) 751–759. Metso, I. Minicocci, E. Ciociola, F. Ceci, A. Montali, M. Arca, C. Ehnholm, M.
[356] N. Uslu, F. Gurakan, A. Yuce, H. Demir, P. Tarugi, Abetalipoproteinemia in an infant Jauhiainen, Angptl3 deficiency is associated with increased insulin sensitivity, lipo-
with severe clinical phenotype and a novel mutation, Turk. J. Paediatr. 52 (2010) protein lipase activity and decreased serum fatty acids, Arterioscler. Thromb. Vasc.
73–77. Biol. 33 (2013) 1706–1713.
[357] M. Najah, S.M. Youssef, H.M. Yahia, S. Afef, J. Awatef, H. Saber, N.M. Fadhel, A. [382] L. Pisciotta, E. Favari, L. Magnolo, S. Simonelli, M.P. Adorni, R. Sallo, T. Fancello, I.
Sassolas, S.M. Naceur, Molecular characterization of Tunisian families with Zavaroni, Characterisation of three kindreds with familial combined hypolipidemia
abetalipoproteinemia and identification of a novel mutation in the MTTP gene, caused by loss-of-function mutations of ANGPTL3, Circ. Cardiovasc. Genet. 5
Diagn. Pathol. 8 (2013) 54. (2012) 42–50.
[358] M.T. Berthier, P. Couture, A. Houde, A. Paradis, A. Sammak, A. Vemer, J. Depres, C. [383] K. Musunuru, J.P. Pirruccello, R. Do, G.M. Peloso, C. Guiducci, C. Sougnez, K.V.
Gagne, D. Gaudet, M. Vohl, The c.419-420insA in the MTP gene is associated with Garimella, S. Fisher, et al., Exome sequencing of the ANGPTL3 mutations and famil-
abetalipoproteinemia among French-Canadians, Mol. Genet. Metab. 81 (2004) ial combined hypolipidemia, N. Engl. J. Med. 363 (2010) 2220–2227.
140–143. [384] I. Minicocci, S. Santini, V. Cantisani, N. Stitziel, S. Kathiresan, J.A. Arroyo, G. Marti, L.
[359] R. Vongsuvanh, A.J. Hooper, J.C. Coakley, J.S. Macdessi, E.V. O'Loughlin, J.R. Burnett, Pisciotta, D. Noto, A.B. Cefalu, M. Maranghi, G. Labbadia, G. Pigna, F. Pannozzo, F.
K.J. Gaskin, Novel mutations in abetalipoproteinaemia and homozygous familial Ceci, E. Ciociola, S. Bertolini, S. Calandra, P. Tarugi, M. Averna, M. Arca, Clinical char-
hypobetalipoproteinaemia, J. Inherit. Metab. Dis. 30 (2007) 990. acteristics and plasma lipids in subjects with familial combined hypolipidemia: a
[360] M. Najah, E. Di Leo, J. Awatef, L. Magnolo, J. Imene, E. Pinotti, M. Bahri, S. pooled analysis, J. Lipid Res. 54 (2013) 3481–3490.
Barsaoui, I. Brini, M. Fekih, M.N. Slimane, P. Tarugi, Identification of patients with [385] W. Yin, S. Romeo, S. Chang, N.V. Grishin, H.H. Hobbs, J.C. Cohen, Genetic variation in
abetalipoproteinemia a nd homozygous familial hypobetalipoproteinemia in ANGPTL4 provides insights into protein processing and function, J. Biol. Chem. 284
Tunisia, Clin. Chim. Acta 401 (2009) 51–56. (2009) 13213–13222.
[361] H.A. Al-Mahdili, A.J. Hooper, D.R. Sullivan, P.M. Stewart, J.R. Burnett, A mild case of [386] S. Romeo, L.A. Pennachio, Y. Fu, E. Boereinkle, A. Tybjaerg-Hansen, H.H. Hobbs, J.C.
abetalipoproteinemia in association with subclinical hypothyroidism, Ann. Clin. Cohen, Population-based resequencing of ANGPTL4 uncovers variations that re-
Biochem. 43 (2006) 516–519. duces triglycerides and increase HDL, Nat. Genet. 39 (2007) 514–516.
[362] M. Di Filippo, H. Crehalet, M.E. Samson-Bouma, V. Bonnet, L.P. Aggerbeck, J.P. [387] S. Romeo, W. Yin, J. Kozlitina, L.A. Pennachio, E. Boereinkle, H.H. Hobbs, J.C. Cohen,
Rabes, F. Gottrand, G. Luc, D. Bozon, A. Sassolas, Molecular and functional analysis Rare loss-of-function mutations in ANGPTL family members contribute to plasma
of two new MTTP gene mutations in an atypical case of abetalipoproteinemia, J. triglyceride levels in humans, J. Clin. Invest. 119 (2009) 70–79.
Lipid Res. 53 (2012) 548–555. [388] M.C. Smart-Halajko, A. Kelley-Hedgepeth, M.C. Montefusco, J.A. Cooper, A. Kopin,
[363] T.M.E. Narcisi, C.C. Shoulders, S.A. Chester, J. Read, D.J. Brett, G.B. Harrison, T.T. M.C. JM, A. Balasubramanyam, H.J. Pownall, D.M. Nathan, I. Peter, P.J. Talmud,
Grantham, M.F. Fox, S. Povey, T.W.A. de Bruin, D.W. Erkelens, D.P.R. Muller, J.K. G.S. Huggins, for the look AHEAD study, ANGPTL4 variants E40K and T266M are as-
Lloyd, J. Scott, Mutations of the microsomal triglyceride transfer protein gene in sociated with lower fasting triglyceride levels in non-Hispanic white Americans
abetalipoproteinemia, Am. J. Hum. Genet. 57 (1995) 1298–1310. from the look AHEAD clinical trial, BMC Med. Genet. 12 (2011) 89.
[364] E.F. Rehberg, K.B. Samson-BoumaM, L. Blinderman, H. Jamil, J.R. Wetterau, L.P. [389] F. Sentinelli, I. Minococci, A. Montali, L. Nanni, S. Romeo, M. Incani, M.G. Cavallo, A.
Aggerbeck, D.A. Gordon, A novel abetalipoproteinemia genotype: identification Lenzi, M. Arca, M.G. Baroni, Association of RXR-gamma gene variants with familial
of a missense mutation in the 97-kDa subunit of the microsomal triglyceride trans- combined hyperlipidemia; genotype and haplotype analysis, J. Lipids (2013),
fer protein that prevents complex formation with protein disulfide isomerase, J. http://dx.doi.org/10.1155/2013/517943.
Biol. Chem. 271 (1996) 29945–29952. [390] M.E. Keebler, C.L. Sanders, A. Surti, C. Guiducci, N.P. Burtt, S. Kathiresan, Association
[365] I. Khatun, M.T. Walsh, M.M. Hussain, Loss of both phospholipid and triglyceride trans- of blood lipids with common DNA sequence variants at 19 genetic loci in the mul-
fer activities of microsomal triglyceride transfer protein in abetalipoproteinemia, J. tiethnic United States National Health and Nutrition Examination Survey III, Circ.
Lipid Res. 54 (2013) 1541–1549. Cardiovasc. Genet. 2 (2009) 238–243.
[366] K. Ohashi, S. Ishibashi, J. Osuga, R. Tozawa, K. Harada, N. Yahagi, F. Shionoiri, Y. [391] K. Musunuru, A. Strong, M. Frank-Kamenestsky, N.E. Lee, T. Ahfeldt, K.V. Sachs, X.
Iizuka, Y. Tamura, R. Nagai, D.R. Illingworth, T. Gotoda, N. Yamada, Novel mutations Li, From noncoding variant to phenotype via SORT1 at the 1p13 cholesterol
in the microsomal triglyceride transfer protein gene causing abetalipoproteinemia, locus, Nature 466 (2010) 714–719.
J. Lipid Res. 40 (2000) 1199–1204. [392] A. Strong, Q. Ding, A.C. Edmondson, J.S. Millar, K.V. Sachs, X. Li, A. Kumaravel, He-
[367] K. Al-Shali, J. Wang, F. Rosen, R.A. Hegele, Ileal adenocarcinoma in a mild pheno- patic sortilin regulates both apolipoprotein B secretion and LDL catabolism, J.
type of abetalipoproteinemia, Clin. Genet. 63 (2003) 135–138. Clin. Invest. 122 (2012) 2807–2816.
[368] C. Phillips, K. Mullan, D. Owens, G.H. Tomkin, Microsomal triglyceride transfer pro- [393] C. Lord, S. Ferro-Novick, E.A. Miller, The highly conserved COPII coat complex sorts
tein polymorphisms and lipoprotein levels in type 2 diabetes, QJM 97 (2004) cargo from the endoplasmic reticulum and targets it to the Golgi, Cold Spring Harb.
211–216. Perspect. Biol. 5 (2013) a013367.
[369] S.A. Miller, J.R. Burnett, M.A. Leonis, C.J. McKnight, F.M. van Bockxmeer, A.J. Hooper, [394] C. Barlowe, L. Orci, T. Yeung, M. Hosobuchi, S. Hamamoto, N. Salama, M.F. Rexach,
Novel missense mutation MTTP gene mutations causing abetalipoproteinemia, M. Ravazzola, M. Amherdt, R. Schekman, COPII: a membrane coat formed by Sec
Biochim. Biophys. Acta 2014 (1841) 1548–1554. proteins that drive vesicle budding from the endoplasmic reticulum, Cell 77
[370] J. Lee, R.A. Hegele, Abetalipoproteinemia and homozygous hypobetalipoproteinemia: (1994) 895–907.
a framework for diagnosis and management, J. Inherit. Metab. Dis. 37 (2014) [395] V. Marcil, E. Seidman, D. Sinnett, R. Sanchez, S. Spahis, A. Sane, E. Levy, Tissue dis-
333–339. tribution and regulation of the small Sar1b GTPase in mice, Cell. Physiol. Biochem.
[371] R. Zamel, R. Khan, R.L. Pollex, R.A. Hegele, Abetalipoproteinemia: two case reports 33 (2014) 1815–1826.
and literature review, Orphanet J. Rare Dis. 3 (2008) 19. [396] B. Jones, E.L. Jones, S.A. Bonney, H.N. Patel, A.R. Mensenkemp, S. Eichenbaum-Voline,
[372] P. Tarugi, M. Averna, E. Di Leo, A. Cefalu, D. Noto, L. Magnolo, L. Cattin, S. Bertolini, S. M. Rudling, U. Myrdal, G. Annesi, et al., Mutations in a Sar1 GTP ase of COPII vesicles
Calandra, Molecular diagnosis of hypobetalipoproteinemia: an ENID review, Ath- are associated with lipid absorption disorders, Nat. Genet. 34 (2003) 29–31.
erosclerosis 195 (2007) e19–e27. [397] A. Georges, J. Bonneau, D. Bonnefont-Rousselot, J. Champignuelle, J.P. Rabes, M.
[373] G. Santulli, Angiopoietin-like proteins: a comprehensive look, Front. Endocrinol. 5 Abifadel, T. Aparico, J.C. Guenedet, et al., Molecular analysis and intestinal expres-
(2014) 4. sion of SAR1 genes and proteins in Anderson’s disease (Chylomicron retention dis-
[374] F. Quagliarini, Y. Wang, J. Kozlitina, N.V. Grishin, R. Hyde, E. Boerwinkle, D.M. ease), Orphanet J. Rare Dis. 6 (2011) 1.
Valenzuela, A.J. Murphy, J.C. Cohen, H.H. Hobbs, Atypical angiopoietin-like protein [398] A.B. Cefalu, P.L. Calvo, D. Noto, M. Baldi, V. Valenti, P. Lerro, F. Tramuto, A. Lezo, I.
that regulates ANGPTL3, Proc. Natl. Acad. Sci. 109 (2012) 19751–19756. Morra, G. Cenacchi, C. Barbera, M.R. Averna, Variable phenotypic expression of chy-
[375] A. Koster, Y.B. Chao, M. Mosior, A. Ford, P.A. Gonzalez-DeWhitt, J.E. Hale, D. Li, Y. Qiu, lomicron retention disease in a kindred carrying a mutation of the Sara2 gene,
C.C. Fraser, D.D. Yang, J.G. Heuer, S.R. Jaskunas, P. Eacho, Transgenic angipoietin-like Metab. Clin. Exp. 59 (2010) 463–467.
(angptl)4 overexpression and targeted disruption of angptl4 and angptl3: regulation [399] M. Charcosset, A. Sassolas, N. Peretti, C.C. Roy, C. Deslandres, D. Sinnett, E. Levy, A.
of triglyceride metabolism, Endocrinology 146 (2005) 4943–4950. Lachaux, Anderson or chylomicron retention disease: molecular impact of five mu-
[376] M. Shimamura, M. Matsuda, H. Yasumo, M. Okazaki, K. Fujimoto, K. Kono, T. tations in the SAR1B gene on the structure and functionality of the Sar1b protein,
Shimizugawa, Y. Ando, R. Koishi, T. Kohama, N. Sakai, K. Kotani, R. Komuro, T. Mol. Genet. Metab. 93 (2008) 74–84.
Ishida, K. Hirata, S. Yamashita, H. Furukawa, I. Shimomura, Angiopoietin-like pro- [400] L.G.D. Fryer, B. Jones, E.J. Duncan, C.E. Hutchison, T. Ozkan, P.A. Williams, O. Alder,
tein3 regulates plasma HDL cholesterol through suppression of endothelial lipase, M. Nieuwdorp, A.K. Townley, A.R. Mensenkemp, D.J. Stephens, G.M. Dallinga-Thie,
Arterioscler. Thromb. Vasc. Biol. 27 (2007) 366–372. C.C. Shoulders, The endoplasmic reticulum coat protein II transport machinery co-
[377] N. Mehta, A. Qamar, L. Qu, A.N. Qasim, N.N. Mehta, M.P. Reilly, D.J. Rader, Differen- ordinates cellular lipid secretion and cholesterol biosynthesis, J. Biol. Chem. 289
tial association of plasma angiopoietin-like proteins 3 and 4 with lipid metabolic (2014) 4244–4261.
traits, Arterioscler. Thromb. Vasc. Biol. 34 (2014) 1057–1063. [401] K. Ouguerram, Y. Zair, F. Kasbi-Chadli, H. Nazih, D. Bligny, J. Schmitz, T. Aparicio, M.
[378] L. Shan, X. Yu, Z. Liu, Y. Hu, L.T. Sturgis, M.L. Miranda, Q. Liu, The angiopoietin-like Chetiveaux, T. Maagot, L.P. Aggerbeck, M.E. Samson-Bouma, M. Krempf, Low rate of
proteins ANGPTL3 and ANGPTL4 inhibit lipoprotein lipase activity through distinct production of apolipoprotein B100 and AI in 2 patients with Anderson disease
mechanisms, J. Biol. Chem. 284 (2009) 1419–1424. (Chylomicron retention disease), Arterioscler. Thromb. Vasc. Biol. 32 (2012)
[379] M.J. Lafferty, K.C. Bradford, D.A. Erle, S.B. Neher, Angiopoietin-like protein 4 inhibi- 1520–1525.
tion of lipoprotein lipase. Evidence for reversible complex formation, J. Biol. Chem. [402] T. Okada, M. Miyashita, J. Fukuhara, M. Sugitani, T. Ueno, M.E. Samson-Bouma, L.P.
288 (2013) 28524–28534. Aggerbeck, Anderson's disease/chylomicron retention disease in a Japanese patient
[380] D. Noto, A.B. Cefalu, V. Valenti, F. Fayer, E. Pinotti, M. Ditta, R. Spina, G. Vigna, P. Yue, with uniparental disomy 7 and a normal SAR1B gene protein coding sequence,
S. Kathiresan, P. Tarugi, M.R. Averna, Prevalence of ANGPTL3 and APOB gene Orphanet J. Rare Dis. 6 (2011) 78.
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 181

[403] M. Sivain, D. Bligny, L.P. AparicioA, A. Grodet, N. Peretti, D. Menard, F. Djouadi, C. [427] A. Schaffer, M. Verdola, L. Barbieri, T.M. Aprami, H. Suryapranata, P. Marino, G. De
Jardel, J.M. Begue, F. Walker, J. Schmitz, A. Lacahux, L.P. Aggerbeck, M.E. Samson- Luca, High density lipoproteins and coronary heart disease. A single center cohort
Bouma, Anderson's disease (chylomicron retention disease): a new mutation in study, Angiology 65 (2014) 696–702.
the SARA2 gene associated with muscular and cardiac abnormlities, Clin. Genet. [428] P.H. Joshi, P.P. Toth, S.T. Lirette, M.E. Griswold, J.M. Massaro, S.S. Martin, M.J. Blaha,
74 (2008) 546–552. K.R. Kulkarni, A.A. Khokhar, A. Correa, R.B. D'Augustino, S.R. Jones, Association of
[404] N. Peretti, C.C. Roy, A. Sassolas, C. Deslandres, E. Drouin, A. Rasquin, E. Seidman, P. high-density lipoprotein subclasses and incident coronary heart disease. The Jack-
Brochu, M.C. Vohl, S. Labarge, R. Bouvier, M.E. Samson-Bouma, M. Charcosset, A. son Heart and Framingham Offspring Cohort Studies, Eur. J. Prev. Cardiol. (2014),
Lachaux, E. Levy, Chylomicron retention disease: a long term srudy of two cohorts, http://dx.doi.org/10.1177/2047487314543890.
Mol. Genet. Metab. 97 (2009) 136–142. [429] G. Silbernagel, B. Schottker, S. Appelbaum, H. Scharnagl, B.E. Kleber, T.B.
[405] N. Peretti, A. Sassolas, C.C. Roy, C. Deslandres, M. Charcosset, J. Castagnetti, L. Grammer, A. Ritsch, U. Mons, B. Holleczek, et al., High-density lipoprotein cholester-
Pugnet-Chardon, P. Moulin, S. Labarge, L. Bouthillier, A. Lachaux, E. Levy, Guide- ol, coronary artery disease and cardiovascular mortality, Eur. Heart J. 34 (2013)
lines for the diagnosis and management of chylomicron retention disease based 3563–3571.
on a review of the literature and the experience of two centers, Orphanet J. Rare [430] L. Calabresi, M. Gomaraschi, S. Simonelli, F. Bernini, G. Franceschini, HDL and ath-
Dis. 5 (2010) 24. erosclerosis: insights from inherited HDL disorders, Biochim. Biophys. Acta 185
[406] M.W. Huff, R.A. Hegele, Apolipoprotein CIII. Going abck to the future for a lipid tar- (2015) 13–18.
get, Circ. Res. 112 (2013) 1405–1408. [431] K. Rye, C.A. Bursill, G. Lambert, F. Tabet, P.J. Barter, The metabolism and
[407] C.S. Gangabadage, J. Zdunek, M. Tessari, S. Nilsson, G. Olivecrona, S. Wijmenga, antiatherogenic properties of HDL, J. Lipid Res. 50 (2009) S195–S200.
Structure and dynamics of human apolipoprotein CIII, J. Biol. Chem. 283 (2008) [432] M. Wilkins, H. Ning, N.J. Stone, M.H. Criqui, L. Zhao, P. Greenland, D.M. Lloyd-Jones,
17416–17427. Coronary heart disease risks are associated with high levels of HDL cholesterol, J.
[408] P.J. Talmud, E. Hawe, S. Martin, M. Olivier, G.J. Miller, E.M. Rubin, L.A. Pennacchio, Am. Heart Assoc. (2014), http://dx.doi.org/10.1161/JAHA 113.000519.
S.E. Humphries, Relative contribution of variation within the APOC3/A4/A5 [433] M. Vergeer, A.G. Holleboom, J.J.P. Kastelein, J.A. Kuivenhoven, The HDL hypothesis:
gene cluster in determining plasma triglycerides, Hum. Mol. Genet. 11 (2002) does high-density lipoprotein protect from atherosclerosis, J. Lipid Res. 51 (2010)
3039–3046. 2058–2073.
[409] J.S. Cohn, M. Tremblay, R. Batal, H. Jacques, C. Rodriguez, G. Steiner, O. Mamer, J. [434] D.J. Rader, G.K. Hovingh, HDL and cardiovascular disease, Lancet 384 (2014)
Davignon, Increased apoC-III production is a characteristic feature of patients 618–625.
with hypertriglyceridemia, Atherosclerosis 177 (2004) 137–145. [435] R.S. Kiss, N. Kavaslar, K. Okuhira, M.W. Freeman, S. Walter, R.W. Milne, R.
[410] W.W. Li, M.M. Dammerman, J.D. Smith, S. Metzger, J.L. Breslow, T. Leff, Common McPherson, Y.L. Marcel, Genetic etiology of isolated low HDL syndrome: incidence
genetic variation in the promoter of the human apoCIII gene abolishes regulation and heterogeneity of efflux defects, Arterioscler. Thromb. Vasc. Biol. 27 (2007)
by insulin and may contribute to hypertriglyceridemia, J. Clin. Invest. 96 (1995) 1139–1145.
2601–2605. [436] J.C. Cohen, R.S. Kiss, A. Pertsemlidis, Y.L. Marcel, R. McPherson, H.H. Hobbs, Multiple
[411] R. Hertz, J. Bishara-Shieban, J. Bar-Tana, Mode of action of peroxisome proliferators rare alleles contribute to low plasma levels of HDL cholesterol, Science 305 (2004)
as hypolipidemic drugs. Suppression of apolipoprotein CIII, J. Biol. Chem. 270 869–872.
(1995) 13470–13475. [437] F. Oldoni, R.J. Sinke, J.A. Kuivenhoven, Mendelian disorders of high density lipopro-
[412] M. Larsson, E. Vorrsjo, P. Talmud, A. Lookene, G. Olivecrona, Apolipoprotein CI and tein metabolism, Circ. Res. 114 (2014) 124–142.
C-III inhibit lipoprotein lipase activity by displacement of the enzyme from lipid [438] V.I. Zannis, A. Chroni, M. Krieger, Role of apoA-I, ABCA1, LCAT and SR-BI in the bio-
droplets, J. Biol. Chem. 288 (2013) 33997–34008. genesis of HDL, J. Mol. Med. 84 (2006) 276–294.
[413] R.C. Kowal, J. Herz, K.H. Weisgraber, R.W. Mahley, M.S. Brown, J.L. Goldstein, Op- [439] H. Saito, P. Dhanasekaran, D. Nguyen, E. Deridder, P. Holvoet, S. Lund-Katz, M.
posing effects of apolipoproteins E and C on lipoprotein binding to low density Phillips, α-Helix formation is required for high affinity binding of human apolipo-
lipoprotein-related protein, J. Biol. Chem. 265 (1990) 10771–10779. protein A-I to lipids, J. Biol. Chem. 279 (2004) 20974–20981.
[414] K. Tomiyasu, B.W. Walsh, K. Ikewaki, H. Judge, F.M. Sacks, Differential metabolism [440] K. Yamakawa-Kobayashi, H. Yanagi, H. Fukayama, C. Hirano, Y. Shimakura,
of human VLDL according to content of apoE and apoC-III, Arterioscler. Thromb. N. Yamamoto, T. Arinami, S. Tsuchiya, H. Hamaguchi, Frequent occurrence
Vasc. Biol. 21 (2001) 1494–1500. of hypoalphalipoproteinemia due to mutant apolipoprotein A-I gene in
[415] M. Sundaram, S. Zhong, M. Bou Khalil, P.H. Links, Y. Zhao, J. Iqbal, M.M. Hussain, R.J. the population: a population-based survey, Hum. Mol. Genet. 8 (1999)
Parks, Y. Wang, Z. Yao, Expression of apolipoprotein C-III in McA-RH7777 cells en- 331–336.
hances VLDL assembly and secretion under lipid-rich conditions, J. Lipid Res. 51 [441] E.J. Schaefer, W.H. Heaton, M.G. Wetzel, H.B. Brewer, Plasma apolipoprotein ab-
(2010) 150–161. sence associated with a marked reduction of high density lipoproteins and prema-
[416] A. Kawakami, M. Aikawa, P. Alcaide, F.W. Luscinskas, P. Libby, F.M. Sacks, Apolipo- ture coronary artery disease, Arteriosclerosis 2 (1982) 16–26.
protein CIII induces expression of vascular cell adhesion molecule-1 in vascular en- [442] R.A. Norum, J.B. Lakier, S. Goldstein, A. Angel, R.B. Goldberg, W.D. Block, D.K. Noffze,
dothelial cella and increases adhesion of monocytic cells, Circulation 114 (2006) P.J. Dolphin, J. Edelglass, D.D. Bogorad, P. Alaupovic, Familial deficiency of apolipo-
681–687. proteins A-I and C-III and precocious coronary artery disease, N. Engl. J. Med. 306
[417] T.I. Pollin, C.M. Damcott, H. Shen, S.H. Ott, J. Shelton, R.B. Horenstein, W. Post, J.C. (1982) 1513–1519.
McLenithan, L.F. Bielak, P.A. Peyser, B.D. Mitchell, M. Miller, J.R. O'Connell, A.R. [443] L. Pisciotta, R. Miccoli, A. Cantafora, L. Calabresi, P. Tarugi, P. Alessandrini, G.B. Bon,
Shuldiner, A null mutation in human APOC3 confers a favorable plasma lipid profile G. Franceschini, C. Cortese, S. Calandra, S. Bertolini, Recurrent mutations of the
and apparent cardioprotection, Science 322 (2008) 1702–1705. apoliporotein A-I gene in three kindreds with severe HDL deficiency, Atherosclero-
[418] I. Tachmazidou, G. Dedoussis, L. Southam, A. Farmaki, G.R.S. Ritchie, D.K. Xifara, A. sis 167 (2003) 335–345.
Matcham, K. Hatzikotoulas, et al., A rare functional cardioprotective APOC3 variant [444] P.M.M. Weers, A.B. Patel, L.C.P. Wan, E. Guigard, C.M. Kay, A. Hafiane, R.
has risen in frequency in distinct population isolates, Nat. Commun. (2013), http:// McPherson, Y.L. Marcel, R.S. Kiss, Novel N-terminal mutation of human apoli-
dx.doi.org/10.1038/ncomms3872. poprotein A-I reduces self association and impairs LCAT activation, J. Lipid
[419] H. Liu, C. Labeur, C.F. Xu, R. Ferrell, L. Lins, R. Brasseur, M. Rosseneu, K.M. Weiss, S.E. Res. 52 (2011) 35–44.
Humphries, P.J. Talmud, Characterization of the lipid binding properties and lipo- [445] W. Huang, J. Sasaki, A. Matsunaga, H. Nanimatsu, K. Moriyama, H. Han, M. Kugi, T.
protein lipase inhibition of a novel apolipoprotein CIII variant Ala23Thr, J. Lipid Koga, K. Yamaguchi, K. Arakawa, A novel homozygous missense mutation in the
Res. 41 (2000) 1760–1771. apo A-I gene with apo A-I deficiency, Arterioscler. Thromb. Vasc. Biol. 18 (1998)
[420] W. Qin, M. Sundaram, Y. Wang, H. Zhou, S. Zhong, C.C. Chang, S. Manhas, E.F. Yao, 389–396.
R.J. Parks, P.J. McFie, S.J. Stone, Z.G. Jiang, C. Wang, D. Figeys, W. Jia, Z. Yao, Missense [446] M. Wada, T. Iso, B.F. Asztalos, N. Takama, T. Nakajima, Y. Seta, K. Kaneko, Y.
mutation in APOC3 within the C-terminal lipid binding domain of human apo-CIII Taniguchi, H. Kobayashi, K. Nakajima, E.J. Schaefer, M. Kurabayashi, Marked high
results in impaired lipid assembly and secretion of triacylglycerol-rich very low density lipoprotein deficiency due to apolipoprotein A-I Tomioka (codon 138 de-
density lipoproteins, J. Biol. Chem. 286 (2011) 27769–27780. letion), Atherosclerosis 207 (2009) 157–161.
[421] N.J. Timpson, K. Walter, J.L. Min, I. Tachmazidou, G. Malerba, S. Shin, L. Chen, M. [447] K. Ikewaki, A. Matsunaga, H. Han, H. Watanabe, A. Endo, J. Tohyama, M. Kuno, J.
Futema, et al., A rare variant in APOC3 is associated with plasma triglyceride and Mogi, K. Sugimoto, N. Tada, J. Sasaki, S. Mochizuki, A novel two nucleotide deletion
VLDL levels in Europeans, Nat. Commun. 5 (2014) 4871. in the apolipoprotein A-I gene, apo A-I Shinbashi associated with high density lipo-
[422] N. Maeda, H. Li, D. Lee, P. Oliver, S.H. Quarfordt, J. Osada, Targeted disruption of the protein deficiency, corneal opacities, planar xanthomas and premature coronary
apolipoprotein C-III gene in mice results in hypotriglyceridemia and protection from artery disease, Atherosclerosis 172 (2004) 39–45.
postprandial hypertriglyceridemia, J. Biol. Chem. 269 (1994 Sep 23) 23610–23616. [448] H. Yokota, Y. Hashimoto, S. Okubo, M. Yumoto, F. Mashige, M. Kawamura, K.
[423] Y. Ito, N. Azrolan, A. O'Connell, A. Walsh, J.L. Breslow, Hypertriglyceridemia as a Kotani, Y. Usuki, S. Shimada, K. Kitamura, K. Nakahara, Apolipoprotein A-I deficien-
result of human apo CIII gene expression in transgenic mice, Science 249 cy with accumulated risk for CHD but no symptoms of CHD, Atherosclerosis 162
(1990) 790–793. (2002) 399–407.
[424] C.R. Pullinger, M.J. Malloy, A.K. Shahidi, M. Ghassemzadeh, P. Duchateau, J. [449] F. Rached, R.D. Santos, L. Camont, M.H. Miname, M. Lhomme, C. Dauteuille, S.
Villagomez, J. Allaart, J.P. Kane, A novel apoliporotein C-III variant apoC-III Lecocq, C.V. Serrano, M.J. Chapman, A. Kontush, Defective functionality of HDL par-
(Gln38 → Lys) associated with moderate hypertriglyceridemia in a large kindred ticles in familial apoA-I deficiency: relevance of alterations in HDL lipidome and
of Mexican origin, J. Lipid Res. 38 (1997) 1833–1840. proteome, J. Lipid Res. 55 (2014) 2509–2520.
[425] B.F. Asztalos, M. Lefevre, T.A. Foster, R. Tulley, M. Windhauser, L. Wong, P.S. [450] E. Bruckert, A. von Eckardstein, H. Funke, I. Beucler, H. Wiebusch, G. Turpin, G.
Roheim, Normolipidemic subjects with low HDL cholesterol levels have altered Assmann, The replacement of arginine by cysteine at residue 151 in apolipoprotein
HDL subpopulations, Arterioscler. Thromb. Vasc. Biol. 17 (1997 Oct) 1885–1893. A-I produces a phenotype similar to that of apolipoprotein A-IMilano, Atheroscle-
[426] D.J. Gordon, J.L. Probstfield, R.J. Garrison, J.D. Neaton, W.P. Castelli, J.D. Knoke, D.R. rosis 128 (1997) 121–128.
Jacobs, S. Bangdiwala, H.A. Tyroler, High-density lipoprotein cholesterol and car- [451] C.R. Sirtori, L. Calabresi, G. Franceschini, D. Baldassarre, M. Amato, J. Johansson, M.
diovascular disease, Circulation 79 (1989) 8–15. Salvetti, C. Monteduro, R. Zulli, M.L. Muiesan, E. Agabiti-Rosei, Cardiovascular
182 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

status of carriers of the apolipoprotein A-I(Milano) mutant: the Limone sul Garda structure and biochemical phenotype of lecithin:cholesterol acyltransferase
study, Circulation 103 (2001) 1949–1954. (LCAT) mutants causing fish-eye disease, J. Lipid Res. 41 (2000) 752–761.
[452] L. Calabresi, M. Canavesi, F. Bernini, G. Franceschini, Cell cholesterol efflux to [477] L. Calabresi, L. Pisciotta, A. Costantin, I. Frigerio, I. Eberini, P. Alessandrini, M. Arca,
reconstituted high-density lipoproteins containing the apolipoprotein A-IMilano G.B. Bon, G. Boscutti, G. Busnach, G. Frascà, L. Gesualdo, M. Gigante, G. Lupattelli, A.
dimer, Biochemistry 38 (1999) 16307–16314. Montali, S. Pizzolitto, I. Rabbone, M. Rolleri, G. Ruotolo, T. Sampietro, A. Sessa, G.
[453] W.Q. Wang, A.S. Moses, G.A. Francis, Cholesterol mobilization by free and lipid- Vaudo, A. Cantafora, F. Veglia, S. Calandra, S. Bertolini, G. Franceschini, The molec-
bound apoAI(Milano) and apoAI(Milano)-apoAII heterodimers, Biochemistry 40 ular basis of lecithin:cholesterol acyltransferase deficiency syndromes: a compre-
(2001) 3666–3673. hensive study of molecular and biochemical findings in 13 unrelated Italian
[454] E.T. Alexander, M. Tanaka, M. Kono, H. Saito, D.J. Rader, M.C. Phillips, Structural and families, Arterioscler. Thromb. Vasc. Biol. 25 (2005) 1972–1978.
functional consequences of the Milano mutation (R173C) in human apolipoprotein [478] T. Forte, K.R. Norum, J.A. Glomset, A.V. Nichols, Plasma lipoproteins in familial lec-
A-I, J. Lipid Res. 50 (2009) 1409–1419. ithin: cholesterol acyltransferase deficiency: structure of low and high density li-
[455] O. Gursky, M.K. Jones, X. Mei, J.P. Segrest, D. Atkinson, Structural basis for distinct poproteins as revealed by elctron microscopy, J. Clin. Invest. 50 (1971) 1141–1148.
functions of the naturally occurring Cys mutants of apolipoprotein A-I, J. Lipid [479] B.F. Asztalos, E.J. Schaefer, K.V. Horvath, S. Yamashita, M. Miller, G. Franceschini, L.
Res. 54 (2013) 3244–3257. Calabresi, Role of LCAT in HDL remodeling: investigation of LCAT deficiency states,
[456] J.O. Lagerstedt, G. Cavigiolio, L.M. Roberts, H.S. Hong, L.W. Jin, P. Fitzgerald, M.N. J. Lipid Res. 48 (2007) 592–599.
Oda, J.C. Voss, Mapping the structural transition in an amyloidogenic apolipopro- [480] S. Narayanan, Lipoprotein-X, CRC Crit. Rev. Clin. Lab. Sci. 11 (1979) 31–51.
tein A-I, Biochemistry 46 (2007) 9693–9699. [481] M. Kuroda, A.G. Holleboom, E.S. Stroes, S. Asada, Y. Aoyagi, K. Kamata, S. Yamashita,
[457] E. Adachi, A. Kosaka, K. Tsuji, C. Mizuguchi, H. Kawashima, A. Shigenaga, K. Nagao, S. Ishibashi, Y. Saito, H. Bujo, Lipoprotein subfractions highly associated with renal
K. Akaji, A. Otaka, H. Saito, The extreme N-terminal region of human apolipopro- damage in familial lecithin:cholesterol acyltransferase deficiency, Arterioscler.
tein A-I has a strong propensity to form amyloid fibrils, FEBS Lett. 588 (2014) Thromb. Vasc. Biol. 34 (2014) 1756–1762.
389–394. [482] D.J. Rader, K. Ikewaki, N. Duverger, H. Schmidt, H. Pritchard, J. Frohlich, M. Clerc,
[458] J. Petrlova, T. Duong, M.C. Cochran, A. Axelsson, M. Morgelin, L.M. Roberts, J.O. M.F. Dumon, T. Fairwell, L. Zech, et al., Markedly accelerated catabolism of apolipo-
Lagerstedt, The fibrillogenic L178H variant of apolipoprotein A-I forms helical fi- protein A-II (ApoA-II) and high density lipoproteins containing ApoA-II in classic
brils, J. Lipid Res. 53 (2012) 390–398. lecithin: cholesterol acyltransferase deficiency and fish-eye disease, J. Clin. Invest.
[459] D. Rowczenio, A. Dogan, J.D. Theis, J.A. Vrana, H.J. Lachmann, A.D. Wechaleker, J.A. 93 (1994) 321–330.
Gilbertson, T. Hunt, S.D.J. Gibbs, P.T. Sattianayagam, J.H. Pinney, P.N. Hawkins, J.D. [483] V. Charlton-Menys, L. Pisciotta, P.N. Durrington, R. Neary, C.D. Short, L. Calabresi, S.
Gillmore, Amyloidgenicity and clinical phenotype associated with five novel muta- Calandra, S. Bertolini, Molecular characterization of two patients with severe LCAT
tions in apolipoprotein A-I, Am. J. Pathol. 179 (2011) 1978–1987. deficiency, Nephrol. Dial. Transplant. 22 (2007) 2379–2382.
[460] X. Xu, Q. Li, L. Pang, G. Huang, J. Huang, M. Shi, X. Sun, Y. Wang, Arctigenin pro- [484] G.M. Frascà, L. Soverini, E. Tampieri, G. Franceschini, L. Calabresi, L. Pisciotta, P.
motes cholesterol efflux from THP-1 macrophages through PPAR-γ/LXR-α signal- Preda, A. Vangelista, S. Stefoni, S. Bertolini, A 33-year-old man with nephrotic syn-
ing pathway, Biochem. Biophys. Res. Commun. 441 (2013) 321–326. drome and lecithin-cholesterol acyltransferase (LCAT) deficiency. Description of
[461] F. Dong, Z. Mo, W. Eid, K.C. Courtney, X. Zha, Akt inhibition promotes ABCA1- two new mutations in the LCAT gene, Nephrol. Dial. Transplant. 19 (2004)
mediated cholesterol efflux to ApoA-I through suppressing mTORC1, PLoS One 9 1622–1624.
(2014), e113789, http://dx.doi.org/10.1371/journal.pone.0113789. [485] S. Takahashi, K. Hiromura, M. Tsukida, Y. Ohishi, H. Hamatani, N. Sakurai, T. Sakairi,
[462] A. Brooks-Wilson, M. Marcil, S.M. Clee, L.H. Zhang, K. Roomp, M. van Dam, L. Yu, C. H. Ikeuchi, Y. Kaneko, A. Maeshima, T. Kuroiwa, H. Yokoo, T. Aoki, M. Nagata, Y.
Brewer, J.A. Collins, H.O. Molhuizen, O. Loubser, B.F. Ouelette, K. Fichter, K.J. Nojima, Nephrotic syndrome caused by immune-mediated acquired LCAT defi-
Ashbourne-Excoffon, C.W. Sensen, S. Scherer, S. Mott, M. Denis, D. Martindale, J. ciency, J. Am. Soc. Nephrol. 24 (2013) 1305–1312.
Frohlich, K. Morgan, B. Koop, S. Pimstone, J.J. Kastelein, J. Genest Jr., M.R. Hayden, [486] G.K. Hovingh, B.A. Hutten, A.G. Holleboom, W. Petersen, P. Rol, A. Stalenhoef, A.H.
Mutations in ABC1 in Tangier disease and familial high-density lipoprotein defi- Zwinderman, E. de Groot, J.J. Kastelein, J.A. Kuivenhoven, Compromised LCAT func-
ciency, Nat. Genet. 22 (1999) 336–345. tion is associated with increased atherosclerosis, Circulation 112 (2005) 879–884.
[463] M. Puntoni, F. Sbrana, F. Bigazzi, T. Sampietro, Tangier disease. Epidemiology, path- [487] L. Calabresi, D. Baldassarre, S. Castelnuovo, P. Conca, L. Bocchi, C. Candini, B.
ophysiology and management, Am. J. Cardiovasc. Drugs 12 (2012) 303–311. Frigerio, M. Amato, C.R. Sirtori, P. Alessandrini, M. Arca, G. Boscutti, L. Cattin, L.
[464] J.F. Oram, J.W. Heinecke, ATP-binding cassette transporter A1: a cell cholesterol ex- Gesualdo, T. Sampietro, G. Vaudo, F. Veglia, S. Calandra, G. Franceschini, Functional
porter that protects against cardiovascular disease, Physiol. Rev. 85 (2005) lecithin: cholesterol acyltransferase is not required for efficient atheroprotection in
1343–1372. humans, Circulation 120 (2009) 628–635.
[465] M. Liu, S. Chung, G.S. Shelness, J.S. Parks, Hepatic ABCA1 and VLDL triglyceride pro- [488] R. Duivenvoorden, A.G. Holleboom, B. van den Bogaard, A.J. Nederveen, E. de Groot,
duction, Biochim. Biophys. Acta 2012 (1821) 770–777. B.A. Hutten, A.W. Schimmel, G.K. Hovingh, J.J. Kastelein, J.A. Kuivenhoven, E.S.
[466] S. Chung, A.K. Gebre, J. Seo, G.S. Shelness, J.S. Parks, A novel role for ABCA1- Stroes, Carriers of lecithin cholesterol acyltransferase gene mutations have acceler-
generated large pre-β migrating nascent HDL in the regulation of hepatic VLDL tri- ated atherogenesis as assessed by carotid 3.0-T magnetic resonance imaging, J. Am.
glyceride secretion, J. Lipid Res. 51 (2010) 729–742. Coll. Cardiol. 58 (2011) 2481–2487.
[467] T. Fasano, P. Zanoni, C. Rabacchi, L. Pisciotta, E. Favari, M.P. Adorni, P.B. Deegan, A. [489] L. Calabresi, E. Favari, E. Moleri, M.P. Adorni, M. Pedrelli, S. Costa, W. Jessup, I.C.
Park, T. Hlaing, M.D. Feher, B. Jones, A.S. Uzak, F. Kardas, A. Dardis, A. Sechi, B. Gelissen, P.T. Kovanen, F. Bernini, G. Franceschini, Functional LCAT is not required
Bembi, P. Minuz, S. Bertolini, F. Bernini, S. Calandra, Novel mutations of ABCA1 for macrophage cholesterol efflux to human serum, Atherosclerosis 204 (1) (2009
transporter in patients with Tangier disease and familial HDL deficiency, Mol. May) 141–146.
Genet. Metab. 107 (2012) 534–541. [490] G. Daniil, A.A. Phedonos, A.G. Holleboom, M.M. Motazacker, L. Argyri, J.A.
[468] A.J. Hooper, K. Robertson, L. Ng, J.S. Kattampallil, D. Latchem, P.C. Willsher, J. Thom, Kuivenhoven, A. Chroni, Characterization of antioxidant/anti-inflammatory prop-
R.I. Baker, J.R. Burnett, A novel ABCA1 nonsense mutation, R1270X, in Tangier dis- erties and apoA-I-containing subpopulations of HDL from family subjects with
ease associated with an unrecognised bleeding tendency, Clin. Chim. Acta 409 monogenic low HDL disorders, Clin. Chim. Acta 412 (2011) 1213–1220.
(2009) 136–139. [491] A.G. Holleboom, G. Daniil, X. Fu, R. Zhang, G.K. Hovingh, A.W. Schimmel, J.J.
[469] S.I. Negi, A. Brautbar, S.S. Virani, A. Anand, E. Polisecki, B.F. Asztalos, C.M. Kastelein, E.S. Stroes, J.L. Witztum, B.A. Hutten, S. Tsimikas, S.L. Hazen, A. Chroni,
Ballantyne, E.J. Schaefer, P.H. Jones, A novel mutation in the ABCA1 gene causing J.A. Kuivenhoven, Lipid oxidation in carriers of lecithin:cholesterol acyltransferase
an atypical phenotype of Tangier disease, J. Clin. Lipidol. 7 (2013) 82–87. gene mutations, Arterioscler. Thromb. Vasc. Biol. 32 (2012) 3066–3075.
[470] C. Albrecht, K. Baynes, A. Sardini, S. Schepelmann, E.R. Eden, S.W. Davies, C.F. [492] A.G. Holleboom, J.A. Kuivenhoven, M. Vergeer, G.K. Hovingh, J.N. van Miert, N.J.
Higgins, M.D. Feher, J.S. Owen, A.K. Soutar, Two novel missense mutations in Wareham, J.J. Kastelein, K.T. Khaw, S.M. Boekholdt, Plasma levels of lecithin:cho-
ABCA1 result in altered trafficking and cause severe autosomal recessive HDL de- lesterol acyltransferase and risk of future coronary artery disease in apparently
ficiency, Biochim. Biophys. Acta 2004 (1689) 47–57. healthy men and women: a prospective case–control analysis nested in the
[471] M.L. Fitzgerald, A.L. Morris, J.S. Rhee, L.P. Andersson, A.J. Mendez, M.W. Freeman, EPIC-Norfolk population study, J. Lipid Res. 51 (2010) 416–421.
Naturally occurring mutations in the largest extracellular loops of ABCA1 can dis- [493] D.S. Ng, The role of lecithin:cholesterol acyltransferase in the modulation of cardio-
rupt its direct interaction with apolipoprotein A-I, J. Biol. Chem. 277 (2002) metabolic risks — a clinical update and emerging insights from animal models,
33178–33187. Biochim. Biophys. Acta 2012 (1821) 654–659.
[472] L.R. Brunham, M.H. Kang, C. Van Karnebeek, S.N. Sadananda, J.A. Collins, L.H. Zhang, [494] S1. Kunnen, M. Van Eck, Lecithin:cholesterol acyltransferase: old friend or foe in
B. Sayson, F. Miao, S. Stockler, J. Frohlich, D. Cassiman, S.W. Rabkin, M.R. Hayden, atherosclerosis? J. Lipid Res. 53 (2012) 1783–1799.
Clinical, biochemical, and molecular characterization of novel mutations in [495] J. Savel, M. Lafitte, Y. Pucheu, V. Pradeau, A. Tabarin, T. Couffinhal, Very low levels
ABCA1 in families with Tangier disease, JIMD Rep. 12 (2014 Oct). of HDL cholesterol and atherosclerosis, a variable relationship—a review of LCAT
[473] M.A. Pervaiz, G. Gau, A.S. Jaffe, A.K. Saenger, L. Baudhuin, J. Ellison, A non-classical deficiency, Vasc. Health Risk Manag. 8 (2012) 357–361.
presentation of Tangier disease with three ABCA1 mutations, JIMD Rep. 4 (2012) [496] R. Scarpioni, C. Paties, G. Bergonzi, Dramatic atherosclerotic vascular burden in a
109–111. patient with familial lecithin-cholesterol acyltransferase (LCAT) deficiency,
[474] F. Peelman, B. Vanloo, J.L. Verschelde, C. Labeur, H. Caster, J. Taveirne, A. Verhee, N. Nephrol. Dial. Transplant. 23 (2008) 1074.
Duverger, J. Vandekerckhove, J. Tavernier, M. Rosseneu, Effect of mutations of N- [497] D. Drayna, A.S. Jarnagin, J. McLean, W. Henzel, W. Kohr, C. Fielding, R. Lawn, Clon-
and C-terminal charged residues on the activity of LCAT, J. Lipid Res. 42 (2001) ing and sequencing of human cholesteryl ester transfer protein cDNA, Nature 327
471–479. (1987) 632–634.
[475] F. Peelman, J.L. Verschelde, B. Vanloo, C. Ampe, C. Labeur, J. Tavernier, J. [498] C. Bruce, R.A. Chouinard Jr., A.R. Tall, Plasma lipid transfer proteins, high-
Vandekerckhove, M. Rosseneu, Effects of natural mutations in lecithin:cholesterol density lipoproteins, and reverse cholesterol transport, Annu. Rev. Nutr. 18
acyltransferase on the enzyme structure and activity, J. Lipid Res. 40 (1999) 59–69. (1998) 297–330.
[476] B. Vanloo, F. Peelman, K. Deschuymere, J. Taveirne, A. Verhee, C. Gouyette, C. [499] M. Nagano, S. Yamashita, K. Hirano, M. Takano, T. Maruyama, M. Ishihara, Y.
Labeur, J. Vandekerckhove, J. Tavernier, M. Rosseneu, Relationship between Sagehashi, T. Kujiraoka, K. Tanaka, H. Hattori, N. Sakai, N. Nakajima, T. Egashira,
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 183

Y. Matsuzawa, Molecular mechanisms of cholesteryl ester transfer protein defi- [523] M.M. Motazacker, J. Peter, M. Treskes, C.C. Shoulders, J.A. Kuivenhoven, G.K.
ciency in Japanese, J. Atheroscler. Thromb. 11 (2004) 110–121. Hovingh, Evidence of a polygenic origin of extreme high-density lipoprotein cho-
[500] M. Nagano, S. Yamashita, K. Hirano, T. Kujiraoka, M. Ito, Y. Sagehashi, H. Hattori, N. lesterol levels, Arterioscler. Thromb. Vasc. Biol. 33 (2013) 1521–1528.
Nakajima, T. Maruyama, N. Sakai, T. Egashira, Y. Matsuzawa, Point mutation [524] M1. Junyent, K.L. Tucker, C.E. Smith, J.M. Lane, J. Mattei, C.Q. Lai, L.D. Parnell, J.M.
(−69G–NA) in the promoter region of cholesteryl ester transfer protein gene in Ordovas, The effects of ABCG5/G8 polymorphisms on HDL-cholesterol concentra-
Japanese hyperalphalipoproteinemic subjects, Arterioscler. Thromb. Vasc. Biol. 21 tions depend on ABCA1 genetic variants in the Boston Puerto Rican Health
(2001) 985–990. Study, Nutr. Metab. Cardiovasc. Dis. 20 (2010) 558–566.
[501] M. Nagano, S. Yamashita, K. Hirano, M. Ito, T. Maruyama, M. Ishihara, Y. Sagehashi, [525] T. Ahmad, D.I. Chasman, J.E. Buring, I.M. Lee, P.M. Ridker, B.M. Everett, Physical ac-
T. Oka, T. Kujiraoka, H. Hattori, N. Nakajima, T. Egashira, M. Kondo, N. Sakai, Y. tivity modifies the effect of LPL, LIPC, and CETP polymorphisms on HDL-C levels
Matsuzawa, Two novel missense mutations in the CETP gene in Japanese and the risk of myocardial infarction in women of European ancestry, Circ.
hyperalphalipoproteinemic subjects: high-throughput assay by Invader assay, J. Cardiovasc. Genet. 4 (2011) 74–80.
Lipid Res. 43 (2002) 1011–1018. [526] G.S. Huggins, G.D. Papandonatos, B. Erar, L.M. Belalcazar, A. Brautbar, C. Ballantyne,
[502] E.M. Teh, P.J. Dolphin, W.C. Breckenridge, M.H. Tan, Human plasma CETP deficien- A.E. Kitabchi, L.E. Wagenknecht, W.C. Knowler, H.J. Pownall, R.R. Wing, I. Peter, M.C.
cy: identification of a novel mutation in exon 9 of the CETP gene in a Caucasian JM, Genetics Subgroup of the Action for Health in Diabetes (Look AHEAD) Study,
subject from North America, J. Lipid Res. 39 (1998) 442–456. Do genetic modifiers of high-density lipoprotein cholesterol and triglyceride levels
[503] L. Calabresi, P. Nilsson, E. Pinotti, M. Gomaraschi, E. Favari, M.P. Adorni, F. Bernini, also modify their response to a lifestyle intervention in the setting of obesity and
C.R. Sirtori, S. Calandra, G. Franceschini, P. Tarugi, A novel homozygous mutation in type-2 diabetes mellitus?: the Action for Health in Diabetes (Look AHEAD) study,
CETP gene as a cause of CETP deficiency in a Caucasian kindred, Atherosclerosis Circ. Cardiovasc. Genet. 6 (2013) 391–399.
205 (2009) 506–511. [527] S1. Kathiresan, C.J. Willer, G.M. Peloso, S. Demissie, K. Musunuru, E.E. Schadt, L.
[504] W.A. van der Steeg, G.K. Hovingh, A.H. Klerkx, B.A. Hutten, I.C. Nootenboom, J.H. Kaplan, D. Bennett, Y. Li, T. Tanaka, B.F. Voight, L.L. Bonnycastle, A.U. Jackson, G.
Levels, A. van Tol, G.M. Dallinga-Thie, A.H. Zwinderman, J.J. Kastelein, J.A. Crawford, A. Surti, et al., Common variants at 30 loci contribute to polygenic dys-
Kuivenhoven, Cholesteryl ester transfer protein and hyperalphalipoproteinemia lipidemia, Nat. Genet. 41 (2009) 56–65.
in Caucasians, J. Lipid Res. 48 (2007) 674–682. [528] N1. Barbarroja, R. López-Pedrera, M.D. Mayas, E. García-Fuentes, L. Garrido-
[505] S.M. Boekholdt, J.F. Thompson, Natural genetic variation as a tool in understanding Sánchez, M. Macías-González, R. El Bekay, A. Vidal-Puig, F.J. Tinahones, The
the role of CETP in lipid levels and disease, J. Lipid Res. 44 (2003) 1080–1093. obese healthy paradox: is inflammation the answer? Biochem. J. 430 (2010)
[506] Z. Wu, Y. Lou, X. Qiu, Y. Liu, L. Lu, Q. Chen, W. Jin, Association of cholesteryl ester 141–149.
transfer protein (CETP) gene polymorphism, high density lipoprotein cholesterol [529] D.W. Morel, J.R. Hessler, G.M. Chisolm, Low density lipoprotein cytotoxicity in-
and risk of coronary artery disease: a meta-analysis using a Mendelian randomiza- duced by free radical peroxidation of lipid, J. Lipid Res. 24 (1983) 1070–1076.
tion approach, BMC Med. Genet. 15 (2014) 118. [530] B.V. Howard, G. Ruotolo, D.C. Robbins, Obesity and dyslipidemia, Endocrinol.
[507] A1. Thompson, E. Di Angelantonio, N. Sarwar, S. Erqou, D. Saleheen, R.P. Dullaart, B. Metab. Clin. N. Am. 32 (2003) 855–867.
Keavney, Z. Ye, J. Danesh, Association of cholesteryl ester transfer protein geno- [531] A1. Tchernof, J.P. Després, Pathophysiology of human visceral obesity: an update,
types with CETP mass and activity, lipid levels, and coronary risk, JAMA 299 Physiol. Rev. 93 (2013) 359–404.
(2008) 2777–2788. [532] S.E. McQuaid, L. Hodson, M.J. Neville, A.L. Dennis, J. Cheeseman, S.M. Humphreys, T.
[508] P.A. McCaskie, J.P. Beilby, C.M. Chapman, J. Hung, B.M. McQuillan, P.L. Thompson, Ruge, M. Gilbert, B.A. Fielding, K.N. Frayn, F. Karpe, Downregulation of adipose tis-
L.J. Palmer, Cholesteryl ester transfer protein gene haplotypes, plasma high- sue fatty acid trafficking in obesity: a driver for ectopic fat deposition? Diabetes 60
density lipoprotein levels and the risk of coronary heart disease, Hum. Genet. (2011) 47–55.
121 (2007) 401–411. [533] M. Adiels, S.O. Olofsson, M.R. Taskinen, J. Borén, Overproduction of very low-
[509] S.E. Borggreve, H.L. Hillege, B.H. Wolffenbuttel, P.E. de Jong, M.W. Zuurman, G. van density lipoproteins is the hallmark of the dyslipidemia in the metabolic syn-
der Steege, A. van Tol, R.P. Dullaart, PREVEND Study Group, An increased coronary drome, Arterioscler. Thromb. Vasc. Biol. 28 (2008) 1225–1236.
risk is paradoxically associated with common cholesteryl ester transfer protein [534] M. Clemente-Postigo, M.I. Queipo-Ortuño, D. Fernandez-Garcia, R. Gomez-Huelgas,
gene variations that relate to higher high-density lipoprotein cholesterol: a F.J. Tinahones, F. Cardona, Adipose tissue gene expression of factors related to lipid
population-based study, J. Clin. Endocrinol. Metab. 91 (2006) 3382–3388. processing in obesity, PLoS One 6 (2011), e24783.
[510] S.L. Pan, X.Q. Luo, Z.P. Lu, S.H. Lu, H. Luo, C.W. Liu, C.Y. Hu, M. Yang, L.L. Du, Z. Song, [535] F. Karpe, T. Olivecrona, G. Walldius, A. Hamsten, Lipoprotein lipase in plasma after
G.F. Pang, H.Y. Wu, J.B. Huang, J.H. Peng, R.X. Yin, Microsomal triglyceride transfer an oral fat load: relation to free fatty acids, J. Lipid Res. 33 (1992) 975–984.
protein gene -493G/T polymorphism and its association with serum lipid levels in [536] R.H. Eckel, The complex metabolic mechanisms relating obesity to hypertriglyc-
Bama Zhuang long-living families in China, Lipids Health Dis. 11 (2012) 177. eridemia, Arterioscler. Thromb. Vasc. Biol. 31 (2011) 1946–1948.
[511] S. Chantepie, A.E. Bochem, M.J. Chapman, G.K. Hovingh, A. Kontush, High-density [537] S.E. Borggreve, R. De Vries, R.P. Dullaart, Alterations in high-density lipoprotein
lipoprotein (HDL) particle subpopulations in heterozygous cholesteryl ester trans- metabolism and reverse cholesterol transport in insulin resistance and type 2 dia-
fer protein (CETP) deficiency: maintenance of antioxidative activity, PLoS One 7 betes mellitus: role of lipolytic enzymes, lecithin:cholesterol acyltransferase and
(2012), e49336. lipid transfer proteins, Eur. J. Clin. Investig. 33 (2003) 1051–1069.
[512] M. Gomaraschi, A. Ossoli, S. Pozzi, P. Nilsson, A.B. Cefalù, M. Averna, J.A. [538] R. de Vries, S.E. Borggreve, R.P. Dullaart, Role of lipases, lecithin:cholesterol acyl-
Kuivenhoven, G.K. Hovingh, F. Veglia, G. Franceschini, L. Calabresi, eNOS activation transferase and cholesteryl ester transfer protein in abnormal high density lipo-
by HDL is impaired in genetic CETP deficiency, PLoS One 9 (2014), e95925. protein metabolism in insulin resistance and type 2 diabetes mellitus, Clin. Lab.
[513] M.G. McCoy, G.S. Sun, D. Marchadier, C. Maugeais, J.M. Glick, D.J. Rader, Character- 49 (2003) 601–613.
ization of the lipolytic activity of endothelial lipase, J. Lipid Res. 43 (2002) 921–929. [539] K. Uyeda, J.J. Repa, Carbohydrate response element binding protein, ChREBP, a
[514] A.C. Edmondson, R.J. Brown, S. Kathiresan, L.A. Cupples, S. Demissie, A.K. Manning, transcription factor coupling hepatic glucose utilization and lipid synthesis, Cell
M.K. Jensen, E.B. Rimm, J. Wang, A. Rodrigues, V. Bamba, S.A. Khetarpal, M.L. Wolfe, Metab. 4 (2006) 107–110.
S. Derohannessian, M. Li, M.P. Reilly, J. Aberle, D. Evans, R.A. Hegele, D.J. Rader, [540] J.D. Horton, J.L. Goldstein, M.S. Brown, SREBPs: activators of the complete program
Loss-of-function variants in endothelial lipase are a cause of elevated HDL choles- of cholesterol and fatty acid synthesis in the liver, J. Clin. Invest. 109 (2002)
terol in humans, J. Clin. Invest. 119 (2009) 1042–1050. 1125–1131.
[515] R.J. Brown, A.C. Edmondson, N. Griffon, T.B. Hill, I.V. Fuki, K.O. Badellino, M. Li, M.L. [541] A.J. Sanyal, C. Campbell-Sargent, F. Mirshahi, W.B. Rizzo, M.J. Contos, R.K. Sterling,
Wolfe, M.P. Reilly, D.J. Rader, A naturally occurring variant of endothelial lipase as- V.A. Luketic, M.L. Shiffman, J.N. Clore, Nonalcoholic steatohepatitis: association of
sociated with elevated HDL exhibits impaired synthesis, J. Lipid Res. 50 (2009) insulin resistance and mitochondrial abnormalities, Gastroenterology 120 (2001)
1910–1916. 1183–1192.
[516] W. Annema, U.J. Tietge, Role of hepatic lipase and endothelial lipase in high- [542] K.L. Donnelly, C.I. Smith, S.J. Schwarzenberg, J. Jessurun, M.D. Boldt, E.J. Parks,
density lipoprotein-mediated reverse cholesterol transport, Curr. Atheroscler. Sources of fatty acids stored in liver and secreted via lipoproteins in patients
Rep. 13 (2011) 257–265. with nonalcoholic fatty liver disease, J. Clin. Invest. 115 (2005) 1343–1351.
[517] S.A1. Khetarpal, A.C. Edmondson, A. Raghavan, H. Neeli, W. Jin, K.O. Badellino, S. [543] P. Simonen, A. Kotronen, M. Hallikainen, K. Sevastianova, J. Makkonen, A.
Demissie, A.K. Manning, D.O. SL, M.L. Wolfe, L.A. Cupples, M. Li, S. Kathiresan, D.J. Hakkarainen, N. Lundbom, T.A. Miettinen, H. Gylling, H. Yki-Järvinen, Cholesterol
Rader, Mining the LIPG allelic spectrum reveals the contribution of rare and com- synthesis is increased and absorption decreased in non-alcoholic fatty liver disease
mon regulatory variants to HDL cholesterol, PLoS Genet. 7 (2011) e1002393. independent of obesity, J. Hepatol. 54 (2011) 153–159.
[518] E. Boes, S. Coassin, B. Kollerits, I.M. Heid, F. Kronenberg, Genetic-epidemiological [544] A.M. Gusdon, K.X. Song, S. Qu, Nonalcoholic Fatty liver disease: pathogenesis and
evidence on genes associated with HDL cholesterol levels: a systematic in-depth therapeutics from a mitochondria-centric perspective, Oxidative Med. Cell. Longev.
review, Exp. Gerontol. 44 (2009) 136–160. 2014 (2014) 637027.
[519] D. Weissglas-Volkov, P. Pajukanta, Genetic causes of high and low serum HDL- [545] F.X. Pi-Sunyer, The obesity epidemic: pathophysiology and consequences of obesi-
cholesterol, J. Lipid Res. 51 (2010) 2032–2057. ty, Obes. Res. 10 (Suppl. 2) (2002) 97S–104S.
[520] M.B. Engler, C.R. Pullinger, M.J. Malloy, Y. Natanzon, M.V. Kulkarni, J. Song, C. Eng, J. [546] S.A. Khan, A. Ali, S.A. Khan, S.A. Zahran, G. Damanhouri, E. Azhar, I. Qadri, Unraveling
Huuskonen, C. Rivera, A. Poon, M. Bensley, A. Sehnert, C. Zellner, J. Kane, Aouizerat, the complex relationship triad between lipids, obesity, and inflammation, Mediat.
Genetic variation in phospholipid transfer protein modulates lipoprotein profiles Inflamm. 2014 (2014) 502749, http://dx.doi.org/10.1155/2014/502749.
in hyperalphalipoproteinemia, Metabolism 57 (2008) 1719–1724. [547] G.S. Hotamisligil, Inflammation and metabolic disorders, Nature 444 (2006)
[521] B.E. Aouizerat, M.B. Engler, Y. Natanzon, M. Kulkarni, J. Song, C. Eng, J. Huuskonen, 860–867.
C. Rivera, A. Poon, M. Bensley, A. Sehnert, C. Zellner, M. Malloy, J. Kane, C.R. [548] B.E. Wisse, The inflammatory syndrome: the role of adipose tissue cytokines in
Pullinger, Genetic variation of PLTP modulates lipoprotein profiles in metabolic disorders linked to obesity, J. Am. Soc. Nephrol. 15 (2004) 2792–2800.
hypoalphalipoproteinemia, J. Lipid Res. 47 (2006) 787–793. [549] K.S1. Polonsky, B.D. Given, E. Van Cauter, Twenty-four-hour profiles and pulsatile
[522] F.W1. Asselbergs, R.C. Lovering, F. Drenos, Progress in genetic association studies of patterns of insulin secretion in normal and obese subjects, J. Clin. Invest. 81
plasma lipids, Curr. Opin. Lipidol. 24 (2013) 123–128. (1988) 442–448.
184 I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185

[550] M.F. White, The insulin signalling system and the IRS proteins, Diabetologia 40 in patients with type 2 diabetes and effect of metformin, Diabetes 59 (2010)
(Suppl. 2) (1997) S2–17. 1038–1045.
[551] V. Aguirre, T. Uchida, L. Yenush, R. Davis, M.F. White, The c-Jun NH(2)-terminal [575] L. Duvillard, E. Florentin, G. Lizard, J.M. Petit, F. Galland, S. Monier, P. Gambert, B.
kinase promotes insulin resistance during association with insulin receptor Vergès, Cell surface expression of LDL receptor is decreased in type 2 diabetic pa-
substrate-1 and phosphorylation of Ser(307), J. Biol. Chem. 275 (2000) tients and is normalized by insulin therapy, Diabetes Care 26 (2003) 1540–1544.
9047–9054. [576] S.M. Gordon, W.S. Davidson, E.M. Urbina, L.M. Dolan, A. Heink, H. Zang, L.J. Lu, A.S.
[552] G.S. Hotamisligil, P. Peraldi, A. Budavari, R. Ellis, M.F. White, B.M. Spiegelman, IRS- Shah, The effects of type 2 diabetes on lipoprotein composition and arterial stiff-
1-mediated inhibition of insulin receptor tyrosine kinase activity in TNF-alpha- ness in male youth, Diabetes 62 (2013) 2958–2967.
and obesity-induced insulin resistance, Science 271 (1996) 665–668. [577] D.J. Shin, T.F. Osborne, Thyroid hormone regulation and cholesterol metabolism are
[553] K.E. Wellen, G.S. Hotamisligil, Inflammation, stress, and diabetes, J. Clin. Invest. 115 connected through Sterol Regulatory Element-Binding Protein-2 (SREBP-2), J. Biol.
(2005) 1111–1119. Chem. 278 (2003) 34114–34118.
[554] D.C. Henstridge, M. Whitham, M.A. Febbraio, Chaperoning to the metabolic party: [578] O. Bakker, F. Hudig, S. Meijssen, W.M. Wiersinga, Effects of triiodothyronine and
the emerging therapeutic role of heat-shock proteins in obesity and type 2 diabe- amiodarone on the promoter of the human LDL receptor gene, Biochem. Biophys.
tes, Mol. Metab. 3 (2014) 781–793. Res. Commun. 249 (1998) 517–521.
[555] M. Roden, T.B. Price, G. Perseghin, K.F. Petersen, D.L. Rothman, G.W. Cline, G.I. [579] L.H. Duntas, G. Brenta, The effect of thyroid disorders on lipid levels and metabo-
Shulman, Mechanism of free fatty acid-induced insulin resistance in humans, J. lism, Med. Clin. North Am. 96 (2012) 269–281.
Clin. Invest. 97 (1996) 2859–2865. [580] X. Prieur, T. Huby, H. Coste, F.G. Schaap, M.J. Chapman, J.C. Rodríguez, Thyroid hor-
[556] L. Rachek, Free fatty acids and skeletal muscle insulin resistance, Prog. Mol. Biol. mone regulates the hypotriglyceridemic gene APOA5, J. Biol. Chem. 280 (2005)
Transl. Sci. 121 (2014) 267–292. 27533–27543.
[557] C. Yu, Y. Chen, G.W. Cline, D. Zhang, H. Zong, Y. Wang, R. Bergeron, J.K. Kim, S.W. [581] K. Hashimoto, M. Yamada, S. Matsumoto, T. Monden, T. Satoh, M. Mori, Mouse ste-
Cushman, G.J. Cooney, B. Atcheson, M.F. White, E.W. Kraegen, G.I. Shulman, Mech- rol response element binding protein-1c gene expression is negatively regulated
anism by which fatty acids inhibit insulin activation of insulin receptor substrate-1 by thyroid hormone, Endocrinology 147 (2006) 4292–4302.
(IRS-1)-associated phosphatidylinositol 3-kinase activity in muscle, J. Biol. Chem. [582] Y.Y. Liu, G.A. Brent, Thyroid hormone crosstalk with nuclear receptor signaling in
277 (2002) 50230–50236. metabolic regulation, Trends Endocrinol. Metab. 21 (2010) 166–173.
[558] S. Pereira, E. Park, Y. Mori, C.A. Haber, P. Han, T. Uchida, L. Stavar, A.I. Oprescu, K. [583] K. Hashimoto, S. Matsumoto, M. Yamada, T. Satoh, M. Mori, Liver X receptor-α
Koulajian, A. Ivovic, Z. Yu, D. Li, T.A. Bowman, J. Dewald, J. El-Benna, D.N. gene expression is positively regulated by thyroid hormone, Endocrinology 148
Brindley, R. Gutierrez-Juarez, T.K. Lam, S.M. Najjar, R.A. McKay, S. Bhanot, I.G. (2013) 4667–4675.
Fantus, A. Giacca, FFA-induced hepatic insulin resistance in vivo is mediated by [584] K. Gauthier, C. Billon, M. Bissler, M. Beylot, J.M. Lobaccaro, J.M. Vanacker, J. Samarut,
PKCδ, NADPH oxidase, and oxidative stress, Am. J. Physiol. Endocrinol. Metab. Thyroid hormone receptor beta (TRbeta) and liver X receptor (LXR) regulate
307 (2014) E34–E46. carbohydrate-response element-binding protein (ChREBP) expression in a tissue-
[559] G.I. Shulman, Ectopic fat in insulin resistance, dyslipidemia, and cardiometabolic selective manner, J. Biol. Chem. 285 (2010) 28156–28163.
disease, N. Engl. J. Med. 371 (2014) 2237–2238. [585] Y. Zhang, L. Yin, F.B. Hillgartner, Thyroid hormone stimulates acetyl-coA
[560] S. Kashyap, R. Belfort, A. Gastaldelli, T. Pratipanawatr, R. Berria, W. Pratipanawatr, carboxylase-alpha transcription in hepatocytes by modulating the composition
M. Bajaj, L. Mandarino, R. DeFronzo, K. Cusi, A sustained increase in plasma free of nuclear receptor complexes bound to a thyroid hormone response element, J.
fatty acids impairs insulin secretion in nondiabetic subjects genetically Biol. Chem. 276 (2001) 974–983.
predisposed to develop type 2 diabetes, Diabetes 52 (2003) 2461–2474. [586] R.W. Brownsey, A.N. Boone, J.E. Elliott, J.E. Kulpa, W.M. Lee, Regulation of acetyl-
[561] V. Poitout, R.P. Robertson, Glucolipotoxicity: fuel excess and beta-cell dysfunction, CoA carboxylase, Biochem. Soc. Trans. 34 (2006) 223–227.
Endocr. Rev. 29 (2008) 351–366. [587] Y. Zhang, K. Ma, S. Song, M.B. Elam, G.A. Cook, E.A. Park, Peroxisomal proliferator-
[562] K.G. Alberti, R.H. Eckel, S.M. Grundy, P.Z. Zimmet, J.I. Cleeman, K.A. Donato, J.C. activated receptor-gamma coactivator-1 alpha (PGC-1 alpha) enhances the thyroid
Fruchart, W.P. James, C.M. Loria, S.C. Smith Jr., International Diabetes Federation hormone induction of carnitine palmitoyltransferase I (CPT-I alpha), J. Biol. Chem.
Task Force on Epidemiology and Prevention, Hational Heart, Lung, and Blood 279 (2004) 53963–53971.
Institute, American Heart Association, World Heart Federation, International Ath- [588] V.A. Drover, N.C. Wong, L.B. Agellon, A distinct thyroid hormone response element
erosclerosis Society, International Association for the Study of Obesity, Harmoniz- mediates repression of the human cholesterol 7alpha-hydroxylase (CYP7A1) gene
ing the metabolic syndrome: a joint interim statement of the International promoter, Mol. Endocrinol. 16 (2002) 14–23.
Diabetes Federation Task Force on Epidemiology and Prevention; National Heart, [589] J. Huuskonen, M. Vishnu, C.R. Pullinger, P.E. Fielding, C.J. Fielding, Regulation of
Lung, and Blood Institute; American Heart Association; World Heart Federation; ATP-binding cassette transporter A1 transcription by thyroid hormone receptor,
International Atherosclerosis Society; and International Association for the Study Biochemistry 43 (2004) 1626–1632.
of Obesity, Circulation 120 (2009) 1640–1645. [590] T. O'Brien, S.F. Dinneen, P.C. O'Brien, P.J. Palumbo, Hyperlipidemia in patients with
[563] I1. Shimomura, M. Matsuda, R.E. Hammer, Y. Bashmakov, M.S. Brown, J.L. primary and secondary hypothyroidism, Mayo Clin. Proc. 68 (1993) 860–866.
Goldstein, Decreased IRS-2 and increased SREBP-1c lead to mixed insulin resis- [591] T. Kuusi, M.R. Taskinen, E.A. Nikkilä, Lipoproteins, lipolytic enzymes, and hormonal
tance and sensitivity in livers of lipodystrophic and ob/ob mice, Mol. Cell 6 status in hypothyroid women at different levels of substitution, J. Clin. Endocrinol.
(2000) 77–86. Metab. 66 (1988) 51–56.
[564] C.M. Taniguchi, K. Ueki, R. Kahn, Complementary roles of IRS-1 and IRS-2 in the he- [592] E.N. Pearce, Update in lipid alterations in subclinical hypothyroidism, J. Clin.
patic regulation of metabolism, J. Clin. Invest. 115 (2005) 718–727. Endocrinol. Metab. 97 (2012) 326–333.
[565] Pajvani UB, Qiang L, Kangsamaksin T, Kitajewski J, Ginsberg HN, Accili D. Inhibition [593] F. Kronenberg, Dyslipidemia and nephrotic syndrome: recent advances, J. Ren.
of Notch uncouples Akt activation from hepatic lipid accumulation by decreasing Nutr. 15 (2005) 195–203.
mTorc1 stability Nat. Med. 19:1054–1060. [594] Y. Zhou, X. Zhang, L. Chen, J. Wu, H. Dang, M. Wei, Y. Fan, Y. Zhang, Y. Zhu, N. Wang,
[566] D.F. Vatner, S.K. Majumdar, N. Kumashiro, M.C. Petersen, Y. Rahimi, A.K. Gattu, M. M.D. Breyer, Y. Guan, Expression profiling of hepatic genes associated with lipid
Bears, J.P. Camporez, G.W. Cline, M.J. Jurczak, V.T. Samuel, G.I. Shulman, Insulin- metabolism in nephrotic rats, Am. J. Physiol. Ren. Physiol. 295 (2008) F662–F667.
independent regulation of hepatic triglyceride synthesis by fatty acids, Proc. Natl. [595] S. Han, N.D. Vaziri, P. Gollapudi, V. Kwok, H. Moradi, Hepatic fatty acid and choles-
Acad. Sci. U. S. A. 112 (2015) 1143–1148. terol metabolism in nephrotic syndrome, Am. J. Transl. Res. 5 (2013) 246–253.
[567] Y.F. Otero, J.M. Stafford, O.P. McGuinness, Pathway-selective insulin resistance and [596] L.C. Clement, C. Macé, C. Avila-Casado, J.A. Joles, S. Kersten, S.S. Chugh, Circulating
metabolic disease: the importance of nutrient flux, J. Biol. Chem. 289 (2014) angiopoietin-like 4 links proteinuria with hypertriglyceridemia in nephrotic syn-
20462–20469. drome, Nat. Med. 20 (2014) 37–46.
[568] H. Duez, B. Lamarche, K.D. Uffelman, R. Valero, J.S. Cohn, G.F. Lewis, [597] N.D. Vaziri, H. Moradi, Dual role of circulating angiopoietin-like 4 (ANGPTL4) in
Hyperinsulinemia is associated with increased production rate of intestinal apoli- promoting hypertriglyceridemia and lowering proteinuria in nephrotic syndrome,
poprotein B-48-containing lipoproteins in humans, Arterioscler. Thromb. Vasc. Am. J. Kidney Dis. 64 (2014) 495–498.
Biol. 26 (2006) 1357–1363. [598] N.D. Vaziri, Causes of dysregulation of lipid metabolism in chronic renal failure,
[569] C. Phillips, K. Mullan, D. Owens, G.H. Tomkin, Intestinal microsomal triglyceride Semin. Dial. 22 (2009) 644–651.
transfer protein in type 2 diabetic and non-diabetic subjects: the relationship to [599] N.D. Vaziri, Molecular mechanisms of lipid disorders in nephrotic syndrome, Kid-
triglyceride-rich postprandial lipoprotein composition, Atherosclerosis 187 ney Int. 63 (2003) 1964–1976.
(2006) 57–64. [600] T. Quaschning, V. Krane, T. Metzger, C. Wanner, Abnormalities in uremic lipopro-
[570] J.P. Nogueira, M. Maraninchi, S. Béliard, N. Padilla, L. Duvillard, J. Mancini, A. tein metabolism and its impact on cardiovascular disease, Am. J. Kidney Dis. 38
Nicolay, C. Xiao, B. Vialettes, G.F. Lewis, R. Valéro, Absence of acute inhibitory effect (2001) S14–S19.
of insulin on chylomicron production in type 2 diabetes, Arterioscler. Thromb. [601] G. Lippi, A. Albiero, M. Montagnana, G.L. Salvagno, S. Scevarolli, M. Franchi, G.C.
Vasc. Biol. 32 (2012) 1039–1044. Guidi, Lipid and lipoprotein profile in physiological pregnancy, Clin. Lab. 53
[571] B. Vergès, Abnormal hepatic apolipoprotein B metabolism in type 2 diabetes, Ath- (2007) 173–177.
erosclerosis 211 (2010) 353–360. [602] A. Montes, C.E. Walden, R.H. Knopp, M. Cheung, M.B. Chapman, J.J. Albers, Physio-
[572] C. Xiao, S. Dash, C. Morgantini, Lewis GF2. New and emerging regulators of intesti- logic and supraphysiologic increases in lipoprotein lipids and apoproteins in late
nal lipoprotein secretion, Atherosclerosis 233 (2014) 608–615. pregnancy and postpartum. Possible markers for the diagnosis of "prelipemia",
[573] J. Wang, A. Stančáková, P. Soininen, A.J. Kangas, J. Paananen, J. Kuusisto, M. Ala- Arteriosclerosis 4 (1984) 407–417.
Korpela, M. Laakso, Lipoprotein subclass profiles in individuals with varying de- [603] M. Rebuffé-Scrive, L. Enk, N. Crona, P. Lönnroth, L. Abrahamsson, U. Smith, P.
grees of glucose tolerance: a population-based study of 9399 Finnish men, J. Intern. Björntorp, Fat cell metabolism in different regions in women. Effect of menstrual
Med. 272 (2012) 562–572. cycle, pregnancy, and lactation, J. Clin. Invest. 75 (1985) 1973–1976.
[574] N1. Rabbani, M.V. Chittari, C.W. Bodmer, D. Zehnder, A. Ceriello, P.J. Thornalley, In- [604] E.A. Ryan, M.J. O'Sullivan, J.S. Skyler, Insulin action during pregnancy. Studies with
creased glycation and oxidative damage to apolipoprotein B100 of LDL cholesterol the euglycemic clamp technique, Diabetes 34 (1985) 380–389.
I. Ramasamy / Clinica Chimica Acta 454 (2016) 143–185 185

[605] R. Root-Bernstein, A. Podufaly, P.F. Dillon, Estradiol binds to insulin and insulin re- [616] T.F. Whayne, Nicotinic acid: current status in lipid management and cardiovascular
ceptor decreasing insulin binding in vitro, Front. Endocrinol. (Lausanne) 5 (2014) disease prevention, Angiology 65 (2014) 557–559.
118, http://dx.doi.org/10.3389/fendo.2014.00118. [617] S.H. Ganji, S. Tavintharan, D. Zhu, Y. Xing, V.S. Kamanna, M.L. Kashyap, Niacin
[606] R.P. Barros, A. Morani, A. Moriscot, U.F. Machado, Insulin resistance of pregnancy noncompetitively inhibits DGAT2 but not DGAT1 activity in HepG2 cells, J. Lipid
involves estrogen-induced repression of muscle GLUT4, Mol. Cell. Endocrinol. Res. 45 (2004) 1835–1845.
295 (2008) 24–31. [618] I. Goldenberg, M. Benderly, U. Goldbourt, Update on the use of fibrates: focus on
[607] P.K. Kinnunen, H.A. Unnérus, T. Ranta, C. Ehnholm, E.A. Nikkilä, M. Seppälä, Activ- bezafibrate, Vasc. Health Risk Manag. 4 (2008) 131–141.
ities of post-heparin plasma lipoprotein lipase and hepatic lipase during pregnancy [619] W.S. Harris, n-3 fatty acids and serum lipoproteins: human studies, Am. J. Clin.
and lactation, Eur. J. Clin. Investig. 10 (1980) 469–474. Nutr. 65 (1997) 1645S–1654S.
[608] A1. Montelongo, M.A. Lasunción, L.F. Pallardo, E. Herrera, Longitudinal study of [620] D. Weitz, H. Weintraub, E. Fisher, A.Z. Schwartzbard, Fish oil for the treatment of
plasma lipoproteins and hormones during pregnancy in normal and diabetic cardiovascular disease, Cardiol. Rev. 18 (2010) 258–263.
women, Diabetes 41 (1992) 1651–1659. [621] A. Zhao, J. Yu, J.L. Lew, L. Huang, S.D. Wright, J. Cui, Polyunsaturated fatty acids are
[609] A. Iglesias, A. Montelongo, E. Herrera, M.A. Lasunción, Changes in cholesteryl ester FXR ligands and differentially regulate expression of FXR targets, DNA Cell Biol. 23
transfer protein activity during normal gestation and postpartum, Clin. Biochem. (2004) 519–526.
27 (1994) 63–68. [622] Z. Reiner, Resistance and intolerance to statins, Nutr. Metab. Cardiovasc. Dis. 24
[610] J.J. Alvarez, A. Montelongo, A. Iglesias, M.A. Lasunción, E. Herrera, Longitudinal (2014) 1057–1066.
study on lipoprotein profile, high density lipoprotein subclass, and postheparin li- [623] V. Tiwari, M. Khokhar, Mechanism of action of anti-hypercholesterolemia drugs
pases during gestation in women, J. Lipid Res. 37 (1996) 299–308. and their resistance, Eur. J. Pharmacol. 741 (2014) 156–170.
[611] L. Fåhraeus, U. Larsson-Cohn, L. Wallentin, Plasma lipoproteins including high den- [624] A.F. Cicero, A. Colletti, C. Borghi, Profile of evolocumab and its potential in the treat-
sity lipoprotein subfractions during normal pregnancy, Obstet. Gynecol. 66 (1985) ment of hyperlipidemia, Drug Des. Devel. Ther. 9 (2015) 3073–3082.
468–472. [625] A.M. Tonkin, G.F. Watts, Into the future: diversifying lipid management, Lancet 380
[612] P. Brizzi, G. Tonolo, F. Esposito, L. Puddu, S. Dessole, M. Maioli, S. Milia, Lipoprotein (2012) 1971–1974.
metabolism during normal pregnancy, Am. J. Obstet. Gynecol. 181 (1999) [626] M.P. Hage, S.T. Azar, Treating low high-density lipoprotein cholesterol: what is the
430–434. evidence? Ther. Adv. Endocrinol. Metab. 5 (2014) 10–17.
[613] A. Martin-Hidalgo, C. Holm, P. Belfrage, M.C. Schotz, E. Herrera, Lipoprotein lipase [627] S.H. Kassim, J.M. Wilson, D.J. Rader, Gene therapy for dyslipidemia: a review of
and hormone-sensitive lipase activity and mRNA in rat adipose tissue during preg- gene replacement and gene inhibition strategies, Clin. Lipidol. 5 (2010) 793–809.
nancy, Am. J. Phys. 266 (1994) E930–E935. [628] F. Mingozzi, J.J. Meulenberg, D.J. Hui, E. Basner-Tschakarjan, N.C. Hasbrouck, S.A.
[614] E. Herrera, M.A. Lasunción, D. Gomez-Coronado, P. Aranda, P. López-Luna, I. Edmonson, N.A. Hutnick, M.R. Betts, J.J. Kastelein, E.S. Stroes, K.A. High, AAV-
Maier, Role of lipoprotein lipase activity on lipoprotein metabolism and the 1-mediated gene transfer to skeletal muscle in humans results in dose-dependent
fate of circulating triglycerides in pregnancy, Am. J. Obstet. Gynecol. 158 (1988) activation of capsid-specific T cells, Blood 114 (2009) 2077–2086.
1575–1583.
[615] J.E. Digby, N. Ruparelia, R.P. Choudhury, Niacin in cardiovascular disease: recent
preclinical and clinical developments, Arterioscler. Thromb. Vasc. Biol. 32 (2012)
582–588.

You might also like