Abstract
Apolipoprotein C-II (apoC-II) is a small exchangeable apolipoprotein found on triglyceride-rich lipoproteins (TRL), such as chylomicrons (CM) and very low-density lipoproteins (VLDL), and on high-density lipoproteins (HDL), particularly during fasting. ApoC-II plays a critical role in TRL metabolism by acting as a cofactor of lipoprotein lipase (LPL), the main enzyme that hydrolyses plasma triglycerides (TG) on TRL. Here, we present an overview of the role of apoC-II in TG metabolism, emphasizing recent novel findings regarding its transcriptional regulation and biochemistry. We also review the 24 genetic mutations in the APOC2 gene reported to date that cause hypertriglyceridemia (HTG). Finally, we describe the clinical presentation of apoC-II deficiency and assess the current therapeutic approaches, as well as potential novel emerging therapies.
Keywords: apoC-II, apoC-II deficiency, apoC-II mutations, triglyceride-rich lipoproteins, triglycerides, hypertriglyceridemia, lipoprotein lipase
1. Introduction
Numerous epidemiological and genetic studies demonstrate a strong positive correlation between elevated fasting and postprandial triglycerides (TG) with cardiovascular disease (CVD). Very severe hypertriglyceridemia (HTG), due to genetic and other causes, is also a well-known trigger for acute pancreatitis 1. As a consequence, the metabolism of triglyceride-rich lipoproteins (TRL) has been studied intensively for several decades 2 and numerous proteins involved in this pathway have been identified 3–5.
Lipoprotein lipase (LPL), a central enzyme in lipolysis of TRL, was first described in the mid-1950’s 6. An LPL activator, apolipoprotein (apo) C-II, was subsequently discovered in 1970 as a protein bound to very low-density lipoproteins (VLDL) and other lipoproteins, and was initially called glutamic acid lipoprotein or ApoLP-Glu based on its C-terminus amino acid (Glu) 7,8. The first apparent genetic cause of HTG was identified 8 years later due to apoC-II deficiency9, and now many other types of genetic mutations, including LPL itself, have been described 10.
Given recent advances in apoC-II and growing interest in developing new therapies for HTG, this review will focus on apoC-II. We describe the clinical presentation of apoC-II deficiency, and the associated spectrum of APOC2 mutations, as well as current and emerging clinical therapeutic options.
2. Overview of TRL metabolism
There are two main pathways whereby TRL deliver energy in the form of TG to peripheral tissues following lipolysis, and they differ depending on the source of TG, namely exogenous (dietary) (Fig. 1A) vs. endogenous (hepatic) sources (Fig. 1B).
Fig. 1. Pathways for TRL metabolism.


(A) The exogenous pathway. Dietary lipids are absorbed in the small intestine, packaged into CM and secreted into lymph just 1 hour after dietary fat consumption. CM contain apoB-48 as the main structural protein and acquire apoC-II and apoE (shown) as well as additional apolipo-proteins such as apoC-I, apoC-III, and apoA-V (not shown) in the perimesenteric lymphatics. Once in circulation, CM are hydrolysed by LPL bound to endothelial cells to FFA and mono-glycerides (MG) and are converted to CM remnants. CM remnants are removed from circulation by the liver via binding to the low-density lipoprotein receptor (LDLR) or the LDL receptor-related protein (LRP). Free fatty acids (FFA) and glycerol generated during lipolysis are utilized by parenchymal tissues.
(B) The endogenous pathway. Hepatically derived VLDL are present in the plasma in the both the fasting and postprandial states. VLDL, which contain apoB-100 as their main structural protein, are secreted into the systemic circulation, and like CM acquire many of the same apolipo-proteins in plasma. The lipolysis of VLDL bound to LPL on endothelial cells generates smaller particles deplete of TG, which are also called remnants or intermediate density lipoproteins (IDL). A fraction of VLDL particles undergoes further lipolysis and is converted to LDL particles without apoE that are taken up by the liver or peripheral tissues via the LDLR. Excess LDL can be deposited into the vessel wall, where it can cause atherosclerosis.
As shown in Fig. 1A, the exogenous pathway involves the packaging of dietary lipids, primarily TG, into chylomicrons (CM) and their delivery to peripheral tissues either for energy production or storage in intracellular fat droplets, primarily in adipocytes. In contrast, VLDL mediate the transport of endogenous lipids to peripheral tissues (Fig. 1B). VLDL synthesis depends on TG availability, which are synthesized from free fatty acids (FFA) after their re-esterification in the liver 11. The liver receives FFA from three main extrahepatic sources: plasma FFA released by adipose tissue, uptake from CM remnants, and direct uptake of dietary FFA from the intestine via the portal vein 12. In the postprandial state, approximately 44% of TG on VLDL are synthesized from plasma FFA, 10% from diet-derived FFA, 15% from CM remnants, and about 8% from de novo lipogenesis. During fasting, 77% of TG of VLDL come from plasma FFA and only 4% from de novo lipogenesis 13.
LPL, the main lipolytic enzyme for TRL, is first synthesized in the rough endoplasmic reticulum of parenchymal cells of adipose tissue, heart, and skeletal muscles as an inactive monomer 14, 15. Lipase maturation factor 1 (LMF1) dimerizes and activates LPL, as well as other related lipases, namely hepatic lipase 16 and endothelial lipase 17. A specific chaperone called Sel1L stabilizes the LPL-LMF1 complex 18. Next, LPL is transported to the luminal surface of endothelial cells in conjunction with heparan sulfate proteoglycans (HSPG) 19, 20. The glycosylphosphatidylinositol (GPI)-anchored high density lipoprotein-binding protein 1 (GPIHBP1) participates in this transport process 21–23, and also anchors LPL to the endothelial cell surface 24.
As shown in Fig. 2, LPL-mediated lipolysis of TRL are governed by several positive and negative regulators 25. Besides apoC-II, which will be discussed in more detail later, apoA-V also stimulates lipolysis, possibly by facilitating the binding of TRL to the endothelial cell surface via HSPG 26, 27 and GPIHBP1 28. In contrast, both apoC-I and apoC-III have been shown by in vitro assays to inhibit LPL activity 29, 30 and their overexpression in animal models raises plasma TG 31, 32. Larsson et al.33 have shown that apoC-I and apoC-III compete with LPL for binding to lipid emulsion particles in vitro in a dose dependent manner. As a result, apoC-I and apoC-III inhibit lipolysis by decreasing LPL activity rather than inactivating the enzyme itself 33. In women, high levels of apoC-III were significantly and independently associated with impaired catabolism of VLDL, resulting in plasma HTG 34, but the main effect may be due to the ability of apoC-III to inhibit hepatic TRL or remnants uptake 35. The angiopoietin-like proteins (ANGPTL) 3, 4, and 8 have also recently been described to inhibit LPL 36. Although their overall effects on lipolysis are similar, they differ in their tissue specific expression, activity, and role, depending on nutritional and metabolic needs 37.
Fig. 2. Lipolytic complex.

LPL action is dependent on multiple regulators, both positive (green) and negative (red). LMF1 is responsible for proper folding and assembly of LPL, whereas Sel1L stabilizes the LPL-LMF1 complex. LPL is transported from parenchymal cells to the endothelial cell surface of the capillary lumen, where it binds to GPIHBP1. ApoC-II is an essential cofactor for LPL activation, whereas apoC-I and apoC-III may inhibit lipolysis. ApoA-V stabilizes the LPL-apoC-II complex by helping TRL to bind to the endothelial cell surface via HSPG. ANGPTL 3, 4, and 8 all inhibit LPL but in different tissue beds. Generated FFA and MG during TG hydrolysis are taken up by cells for energy metabolism or storage.
Finally, following TG lipolysis, the generated FFA may be taken up by cells via both receptor-mediated and receptor-independent pathways 38. Eventually, as TRL become depleted of TG, they dissociate from the endothelial cell surface and are removed by the liver. In addition to apoE and apoB-100, inactivated LPL bound to remnants can also serve as a ligand for removal by the liver 39. ApoC-II and the other apolipoproteins that modulate LPL eventually dissociate from remnant lipoproteins during lipolysis and then are transferred onto HDL 40, which serves as a reservoir for these proteins in the fasting state.
3. Transcriptional control of the human APOC2 gene
3.1. Overview
The human APOC2 gene resides within the human APOE-APOC1-APOC4-APOC2 gene cluster on Chromosome 19 and is subject to complex transcriptional control (Fig. 3). ApoC-II is primarily expressed in the liver and secreted into plasma, but it is also produced by other tissues, including the intestine, macrophages, adipose tissue, brain, skin, lungs, retina, and retinal pigment epithelium 41–46 where it may regulate lipolysis at the local level. Transcriptional regulation of the APOC2 gene has been best characterized in liver and macrophages, and to a lesser extent, in the intestine. Each of these tissues regulates APOC2 in a tissue-specific manner. For example, in liver, the farnesoid X receptor (known as FXR or NR1H4), which binds bile acids and regulates genes involved in bile acid synthesis and transport, plays a major role in APOC2 gene transcription. In macrophages, however, key transcription factors for the APOC2 gene include the liver X receptor (LXR), whose role is to correct for cellular cholesterol 47, 48 and signal transducer and activator of transcription 1 (STAT1), which mediates cellular responses to interferons and certain cytokines and growth factors 49. In intestine, ApoC2 gene expression increases with dietary lipids, as might be expected, but recent work in mice has also demonstrated a unique regulatory mechanism involving CD36 sensing of dietary fat 50.
Fig. 3. Transcriptional control of the human APOC2 gene.

(A) Relative gene position and location of major regulatory elements that control APOC2 gene and nearby genes in E/C-I/C-IV/C-II gene cluster on Chromosome 19. The liver-specific elements HCR.1 and HCR.2 are 319 bp elements approximately 22 kb and 11 kb, respectively, upstream of the APOC2 promoter. Both contain FXR elements required for bile-acid – dependent activation of the APOC2 gene. The HCR.1 FXRE (nt 214/226 of HCR.1) also binds HNF-4 and ARP1. The proximal APOC2 gene promoter is 545 bp long and contains binding sites for HNF-4 (−102 − −81), RXRα/T3Rβ (−140 − −155), cAMP (between −36 and −170) and STAT1 (−500 − −493). Long-range interactions between FXR/RXRα in HCR.1 and RXRα at the RXR/T3Rβ binding site have been shown to mediate hepatocyte-specific bile-acid regulation of the APOC2 gene (shown by upper bracket and green + at top of Fig. 3A). The RXRα/T3Rβ binding site also binds ARP1 and EAR1, which may affect transcriptional activation by RXR. The ME.1 and ME.2 are 620-bp multienhancer elements that contain LXREs, both located at nt 442 − 466 of each ME. The LXRE in ME.2 regulates the entire gene cluster. In macrophages, two STAT1 binding elements, one in ME.2 (between nt 174 − 182 of the 620 bp ME.2 element) and one between nucleotides -500 and -493 of the proximal APOC2 promoter, upregulate APOC2 gene expression. Long-range interactions between STAT1 bound to ME.2 and RXRα at the RXR/T3Rβ binding site have been shown to mediate macrophage-specific STAT1 regulation of the APOC2 gene (shown by lower bracket and purple + at top of Fig. 3A). Fibrates, statins and ezetimibe decrease APOC2 gene expression but the promoter elements involved have not been defined. Figure is not to scale.
(B) Regulation of APOC2 gene by long noncoding RNA. lncLSTR indirectly inhibits APOC2 gene expression by a bile-acid-dependent mechanism, involving CYP8B1 and its transcriptional repressor TDP-43. Figure is not to scale.
In the following sections, we review the regulatory mechanisms of APOC2 gene expression in the liver, macrophages, and intestine. The picture that emerges reveals distinctly different mechanisms of regulation, reflecting different physiological roles for apoC-II in these three different tissues.
3.2. Hepatic expression of ApoC-II
The main role of apoC-II secreted by the liver into the plasma is to enhance TG hydrolysis of VLDL and CM for energy delivery or storage. Consequently, the APOC2 gene in the liver responds to metabolic cues by activation of transcription factors and nuclear hormone receptors.
Liver-specific expression of the APOC2 gene, as well as APOE, APOC1, and APOC4, is coordinately controlled by two 319-bp liver-specific hepatic control regions, HCR.1 and HCR.2 (Fig. 3A) 51, 52, located approximately 22 and 11 kb upstream of the APOC2 gene transcriptional start site, respectively 53. HCR.1 and HCR.2 share 85% nucleotide homology. Both contain FXR response elements that bind FXR/retinoid X receptor (RXR) heterodimers that mediate activation of the APOC2 gene by bile acids 54, 55, which are end products of cholesterol metabolism and potent signaling molecules. The FXR/RXRα binding site in HCR.1 (nt 214/226 of HCR.1) also binds hepatic nuclear factor-4α (HNF4α) and apoA-I regulatory protein-1 (ARP1) (Fig. 3A) 56. HNF4α is a master regulator of lipid metabolism in the liver 57. It binds to and is activated by FFA 58.
The -545/+18 APOC2 gene proximal promoter contains five DNase-I-protected promoter segments: C-II-A, -B, -C, -D, and -E 59. The C-II-B (−102–−81) segment of the proximal promoter contains a hormone response element (HRE) that binds HNF-4 (Fig. 3A). Interestingly, a promoter mutation in a patient with chylomicronemia maps to position −86 of the APOC2 promoter, within the CII-B element 60. The C-II-C segment (−159–−116) has a second HRE that binds the transcription factors ARP1 and eosinophil-associated, ribonuclease A-1 (EAR1) (Fig. 3A) 59. The APOC2 promoter is also stimulated by triiodothyronine (T3) via RXRα/thyroid hormone receptor β (known as T3Tβ or THRβ) dimers, which likely bind the HRE at −140–−155 of the APOC2 promoter 61. In the presence of the appropriate ligands (bile acids, T3, retinoids, etc.), the two HREs in the apoC-II proximal promoter interact with the HCR.1 enhancer to promote liver-specific expression of the APOC2 gene. This activation requires a long-range physical interaction between the HCR.1 FXR/RXR element, 11 kb upstream of the APOC2 promoter, and the APOC2 proximal promoter (Fig. 3A). Under different metabolic conditions, the same elements can recruit corepressors that inhibit APOC2 gene expression 56, 59, 61.
An unexpected new twist in FXR regulation of the APOC2 promoter was the recent discovery of a long noncoding RNA, liver-specific TG regulator (lncLSTR), which inhibits APOC2 gene expression 62. Depletion of lncLSTR leads to increased LPL activity and plasma TG clearance. LncLSTR does not directly interact with the APOC2 gene but instead binds to transactive response DNA-binding protein 43 kDa (TDP-43), a transcriptional repressor of the sterol 12α-hydroxylase (CYP8B1) gene (Fig. 3B). This interaction counteracts the repression of the CYP8B1 gene by TDP-43. Depleting lncLSTR causes more TDP-43 to bind to the CYP8B1 promoter and inhibits CYP8B1 expression. This alters the bile acid pool by increasing the muricholic acid (MCA)/cholic acid (CA) ratio, which increases APOC2 gene expression through the FXR pathway, and decreases plasma TG levels (Fig. 3B) 62.
New insights have also recently been gained into regulation of the APOC2 gene by cyclic adenosine monophosphate (cAMP). Streicher et al. 63 previously reported a cAMP response element located between −170 and −36 of the APOC2 gene promoter (Fig. 3A), with a 6-fold up-regulation of APOC2 mRNA levels in HepG2 cells after incubation with the cAMP activator for-skolin. Recently, cAMP-responsive-element-binding protein H (CREB-H), a transcription factor highly expressed in the liver and small intestine induced by fasting, was found to increase the expression of apoC-II and other LPL coactivators. Multiple nonsynonymous mutations in the cAMP responsive element binding protein 3 like 3 (CREB3L3) gene, which encodes CREB-H, have been found in individuals with extreme HTG 64. Based on these findings, CREB-H, as well as the lncLSTR pathway, appear to be promising new targets for therapeutic intervention.
Although Apoc2 mRNA levels are relatively invariant in adult rat livers, it can rapidly change in newborn rats in response to suckling. It rapidly increases immediately after birth, and then decreases slightly, reaching adult liver levels around the suckling/weaning transition period (day 20 of age) when they switch to a carbohydrate-rich diet 65. Hepatic Lpl and Apoc2 gene expression levels change in parallel, thus enabling hydrolysis of the high levels of milk fat ingested during suckling 65. In pregnant rats fed a protein-restricted diet, placental Apoc2 mRNA also increases 17.6-fold, along with changes in other genes involved in cholesterol and lipoprotein transport 66.
Several known drugs, namely fibrates, statins, and ezetimibe influence hepatic APOC2 gene expression in humans 67. Fibrates decrease APOC2 gene expression through peroxisome proliferator-activated receptor (PPAR) α 65, whereas PPARγ coactivator 1α (PGC-1α) partners with HNF4α to increase hepatic APOC2 mRNA levels under fasting conditions 68. PGC-1α also regulates a number of genes in response to fasting 68. The promoter elements involved in modulating hepatic APOC2 gene expression by these factors have not yet been defined. Finally, not only APOC2 but also APOC3 and APOC1 gene expression in the liver are influenced by certain drugs and metabolic disorders, as well as age, sex and diet 65, 67. When assessing potential effects of drugs to treat HTG on gene expression, measuring hepatic APOC2 gene expression is, clearly, of great importance, but expression of other proteins affecting lipolysis must be taken into consideration as well.
Nonetheless, studies of hepatic APOC2 gene expression has provided a wealth of information into the metabolic signals that influence apoC-II levels in plasma. This information may be useful in future drug development.
3.3. Macrophage specific expression of ApoC-II
ApoC-II is also expressed in macrophages and has been localized in macrophages within murine arterial lesions 55. Presumably, it facilitates the delivery of TG to macrophages, which are metabolically very active cells, but this may be pro-atherogenic based on the findings that decreased expression of LPL in macrophages decreases atherosclerosis in mice 69.
Macrophage-specific expression of the APOE-APOC1-APOC4-APOC2 gene cluster depends on two 620-bp macrophage-specific multi-enhancer elements termed ME.1 and ME.2, located 32.85 and 20.25 kp upstream of the APOC2 promoter (Fig. 3A) 70. Both ME.1 and ME.2 are required for APOE gene expression in macrophages, adipose tissue, and brain 71, 72. It was demonstrated that two LXR response elements (LXRE) within ME.1 and ME.2 were essential for stimulation of APOC2 promoter activity by LXR and RXR in macrophages 55. As LPL is also stimulated by LXR, Mak et al. 55 hypothesize that secretion of apoC-II and LPL together from LXR-activated macrophages will result in increased local LPL activity. Moreover, ATP-binding cassette transporter A1 (ABCA1) is also upregulated by LXR in macrophages, and since apoC-II, as well as apoE, apoC-I, and apoC-IV are all able to act as acceptors for cholesterol efflux from macrophages 73, the coordinate upregulation of ABCA1 and this gene cluster may enhance cholesterol efflux from macrophages, including foam cells in the arterial wall 55. Finally, LXR is has been shown to have an anti-inflammatory role in macrophages 47.
TG-interacting factor 1 (TGIF1), a transcriptional repressor, was found to bind to ME site(s) in the APOE/C1/C4/C2 locus (Fig. 3A) and repress APOC2, as well as APOA4, gene expression in the liver 74. While the mechanism of repression is not entirely clear, it is interesting that the ME.1 and ME.2, in addition to enhancing macrophage-specific APOC2 gene expression through LXR, also participate in repression of liver-specific APOC2 expression in the liver, through TGIF1.
APOC2 mRNA levels were found to increase during 12-O-tetradecanoylphorbol 13-acetate (TPA)-induced differentiation of promonocytic U937 cells into monocytes and macrophages 75. Interestingly, differentiation of monocytes into macrophages markedly increases STAT1 protein synthesis 76. Trusca et al. 76 found two STAT1 binding sites involved in APOC2 gene regulation: one in the proximal promoter (nt −500–−493) and the second in ME.2, 20 kb upstream of the APOC2 promoter (Fig. 3A). The ME.2 STAT1 site was previously shown to be important for macrophage-specific APOE gene expression. For APOC2 expression in macrophages, not only the ME.2 STAT1 site but also the STAT1 binding site in the proximal promoter confer macrophage-specific promoter activity. STAT1 on both the proximal and the distal binding sites were found to interact with RXRα on the RXRα/T3Rβ binding site (nt −140– −155 of the proximal APOC2 promoter). This complex of transcription factors on the APOC2 promoter, including both short- and long-range interactions, confers macrophage-specific gene expression.
The markedly different regulation of the APOC2 gene in macrophages vs. hepatocytes necessarily reflects a need to respond to different physiological signals in these two tissues. As macrophages play a major role in atherosclerosis, and macrophage apoC-II is predicted to have a role in local lipid metabolism, generating mice with macrophage-specific ApoC2 gene knockouts could potentially uncover a specific role for macrophage apoC-II in atherosclerosis.
3.4. Intestinal regulation of apoC-II
Early on, apoC-II was noted to be synthesized by mucosal epithelial cells in rat small intestine 77. In humans, it is expressed in jejunal enterocytes at relatively high levels 78, 79, but the overall contribution of the intestine vs. the liver in maintaining plasma levels of apoC-II is not known. In mice, it was recently shown that the Apoc2 gene is induced in the intestine more than 3-fold after an oil gavage 50, suggesting a role for apoC-II in the postprandial metabolism of CM. The mechanism involves lipid-mediated downregulation of CD36, which is involved in dietary fat sensing in the enterocyte, and is impaired in the metabolic syndrome 50. More detailed studies are needed, however, to better understand the exact mechanism for Apoc2 gene regulation in the intestine.
4. ApoC-II biochemistry
After cleavage of a 22-amino acid signal peptide, the mature full-length apoC-II protein contains 79 amino acids and has a MW of 8916 Da 80 (Fig. 4A). Retention of the signal peptide in transgenic mice due to a specific mutation has recently been shown to cause the selective binding of apoC-II to HDL, leading to HTG 81. ApoC-II contains no N-linked glycosylation sites and does not appear to undergo significant O-linked glycosylation.
Fig. 4. ApoC-II structure.

(A) Sequence and the helical wheel plot of apoC-II helices. Charged residues are colored light blue, hydrophobic residues are green, neutral polar residues orange and Q and N are colored red. Position of hydrophobic moment and score is shown in center of helix.
(B) Human APOC2 gene mutations. Position of known mutations are listed below exon-intron diagram of APOC2 gene with position 1 after the signal peptide cleavage site. * or X: STOP codon. The four mutations shown in bold account for more than half of all APOC2 mutations.
Plasma levels of apoC-II are approximately 4.5 µmol/L (40 mg/L) 82 in normolipidemic individuals but can be considerably higher when there is an accumulation of TRL in plasma. There are several factors, in addition to drugs described earlier, such as age, sex, diet, diabetes mellitus (DM) type 2, or chronic kidney disease that can affect plasma apoC-II concentration. ApoC-II levels increase until the age of 60 years in women, and until age 40 years in men 83. Moreover, overweight-obese women have higher plasma apoC-II concentrations than normal-weight women 34. A high-carbohydrate diet has been associated with increased apoC-II concentrations relative to apoC-III in VLDL 79. The increased apoC-II levels in DM type 2 84 and chronic renal failure 85 are associated with elevated TG in plasma. Because the plasma levels of TRL even in the postprandial state are usually far less than 0.5 µmol/L, there is enough apoC-II to exist in multiple copies per TRL particle. The plasma concentrations of apoC-III, which inhibits lipolysis and clearance of TRL, are twice as high as apoC-II 82, and it may be the ratio of these two proteins that regulates TRL catabolism. However, it has been recently shown that in women the apoC-III:apoC-II ratio was not associated with either apoB-100 levels or plasma TG concentration 34.
Based on secondary structure prediction and nuclear magnetic resonance analysis 86, 87, apoC-II contains only 3 helices (Fig. 4A). The first helix (Fig. 4A) at the N-terminus, residue 16 to 36 (historical nomenclature, based on residue 1 occurring after signal peptide cleavage) is relatively long and is primarily responsible for lipoprotein binding. It forms a classic Type A amphipathic helix, with a large hydrophobic face, two positively charged residues (Lys) close to the water/lipid interface and three negatively charged residues (Glu) on the hydrophilic side of a helix. This amino acid configuration allows apoC-II to bind tightly to a lipid surface 88. The second helix is relatively short and only encompasses residues 50−56, and from nuclear magnetic resonance analysis this region appears to be mostly in a random coil configuration 86. The third helix (Fig. 4A) on the C-terminus (residues 63−76, in historical nomenclature) forms a G-type helix similar to what is commonly found in globular proteins, and does not show a strong polarization of hydrophobic and hydrophilic amino acids, and hence does not bind to lipids avidly 89. Based on natural mutations 90, 91, site-directed mutagenesis,92 and synthetic peptide studies 93, the C-terminal helix (Fig. 4A, B) is responsible for LPL activation 92. Residues Tyr-63, Ile-66, Asp-69, and Gln-70 (HGVS-recommended nomenclature: Tyr-85, Ile-88, Asp-91, and Gln-92), all within helix 3, are thought to form a binding site for LPL 94 and have been proposed to lead to its activation by helping guide the TG substrate into the active site of LPL 87.
The exact binding site of apoC-II on LPL is not known, but it involves both the N- and C-terminal domains of LPL, with two apoC-II molecules bound per dimer of LPL 95, 96. By crosslinking studies, the C-terminus of apoC-II has been shown to be linked to specific amino acid residues near the lid region of LPL, and thus may cause displacement of the lid and facilitate TG entry into the active site 97. It depends somewhat on the nature of the lipid substrate, but when TRL, a natural substrate, is used, apoC-II increases the apparent Vmax of LPL and only has a limited effect on the Km 95. Based on oil-drop tensiometer studies, it was recently established that the initial binding of apoC-II and LPL to lipoproteins depends on the surface pressure of TRL 98. Both proteins readily bind to newly secreted TRL, which have relatively low surface pressure, but as the surface pressure gradually increases with lipolysis due to retention of FFA, diglycerides, and monoglycerides on the surface of TRL, LPL dissociates unless it is retained to the TRL surface by apoC-II. Eventually, the surface pressure on the remnant particles becomes too high and both apoC-II and LPL are displaced, thus stopping any further lipolysis 98.
5. Clinical presentation of apolipoprotein C-II deficiency
Clinically, HTG is an extremely heterogeneous phenotype, with both monogenic and polygenic etiologies; therefore, the diagnosis of apoC-II deficiency requires a step-wise process 99. First, polygenic HTG is relatively common with contributions of both rare and common genetic variants 100. In addition, secondary causes of HTG should always be excluded when first evaluating a patient. These include uncontrolled or inadequately-managed DM type 1 or type 2, hypothyroidism, chronic renal insufficiency, nephrotic syndrome, alcohol overuse, excessive intake of long-chain TG and free sugars, and certain medications (e.g. estrogen, beta-blockers, diuretics, isotretinoin, glucocorticoids, serotonin reuptake inhibitors, and atypical antipsychotic drugs, such as clozapine and olanzapine) 101, 102.
Polygenic chylomicronemia, perhaps better known as type V hyperlipidemia, is fairly common with a prevalence of about one in 600 103, but in this condition, both CMs and VLDLs are elevated. On the other hand, familial chylomicronemia syndrome (FCS), formerly known as type I hyperlipoproteinemia, is a group of extremely rare monogenic conditions with a reported prevalence of about one in 1 million, equally distributed in both genders, and higher frequencies are found in some populations due to a founder effect 90, 104. Among FCS, the most common cause is LPL deficiency (OMIM 238600) and apoC-II deficiency (OMIM 207750) is less frequent 90. In addition, other proteins associated with the TRL metabolism, such as deficiencies or defects of apoA-V (OMIM 144650), LMF1 (OMIM 246650), and GPIHBP (OMIM 615947) are all rare causes of FCS 90. Auto-antibodies against LPL 105, apoC-II 106, and GPIHBP1 107 also cause HTG and when severe, non-familial chylomicronemia syndrome.
Due to the autosomal recessive inheritance pattern of FCS, there may be no previous family history. In addition, since the first clinical manifestation of LPL or apoC-II deficiency often occurs during childhood or adolescence, HTG is unlikely to be the first clinical finding. Usually, patients first come to medical attention when they present with varying degrees of abdominal pain from vague queasiness to incapacitating pain due to acute pancreatitis. The presence of excess CMs is thought to be the inciting cause of pancreatitis. Frequently, when the pain is mild, it is often dismissed as common childhood illnesses. The pain may also be mistaken for an acute abdomen, presaging unnecessary exploratory procedures.
Other notable FCS-associated clinical features include eruptive xanthomas, lipemia retinalis, and hepatosplenomegaly. Eruptive xanthomas are transient cutaneous rash-like yellowish papules, 3 to 5 mm in diameter, on the trunk, elbows, or buttocks. They typically occur when TG levels peak in FCS patients due to the uptake of TRL by macrophages. They are normally nonpruritic and painless unless they are present on sensitive areas, such as the soles of feet. Lipemia retinalis is engorgement of the retinal vessels by milky and viscous lipemic plasma, visible only by fundoscopic examination. Hepatosplenomegaly is the result of macrophage uptake of TGs in liver due to persistent HTG.
When one or more FCS-associated clinical features are present in children or adolescents, the workup should include plasma TG, as well as total cholesterol (TC), HDL-cholesterol, LDL-cholesterol, and apoB. In FCS, fasting TG levels are usually over 11.3 mmol/L (1,000 mg/dL) and often over 22.6 mmol/L (2,000 mg/dL). Typically, untreated plasma TG levels greatly exceed TC, and the TG/TC ratio of >2.2 (mmol/L)/(mmol/L) or >5 (mg/dL)/(mg/dL) 92, 99 is a useful indication for FCS. Importantly, low levels of apoB (<0.75 g/L or <75 mg/dL) alone and especially a high TG/apoB ratio >10 (mmol/L)/(g/L) or ≥8.8 (mg/dL)/(mg/dL) help to differentiate FCS from type V hyperlipidemia 108.
Although clinical features are usually indistinguishable among different genetic etiologies in FCS, there may be subtle differences. ApoC-II deficiency is usually less severe with a later onset than LPL deficiency 104, 109. In addition, there seems to be a stronger association between apoC-II deficiency and insulin-dependent DM; however, it is unclear whether beta-cell destruction due to recurrent pancreatitis is the sole cause of DM.
To confirm the diagnosis of apoC-II deficiency, special biochemical and/or molecular analyses are necessary. An informative diagnostic algorithm of FCS published recently may be helpful 99. In theory LPL enzyme activity should facilitate the diagnosis of FCS. In practice, it is only available as a research assay, limiting utility for clinical use. Moreover, results may vary greatly between laboratories. LPL enzyme activity is determined by collecting two sets of plasma, before and after an intravenous heparin administration (60 U/kg body weight). Heparin facilitates the release of LPL tethered to the endothelium into the blood stream for an in vitro activity assay 6. Absent or greatly reduced LPL enzyme activity in the post-heparin plasma is diagnostic for LPL or apoC-II deficiency after excluding hepatic lipase deficiency 110. Then, apoC-II deficiency is apparent when LPL activity is restored on adding apoC-II protein or “normal” plasma in vitro 111. Alternatively, LPL activity can be analyzed in adipose tissues 112. There also exist several ELISAs for measuring apoC-II mass, but they are not routinely used by most clinical laboratories and may be misleading, as patients with missense mutations that inactivate apoC-II may nonetheless have normal levels of apoC-II mass in plasma. The post-heparin LPL activity test, however, is only available at specialized centers and the use of heparin can trigger thrombocytopenia.
In recent years, the definitive diagnosis of FCS has increasingly relied on molecular diagnosis employing multi-gene panels targeting FCS-associated genes 90, 113. However, when novel APOC2 variants are identified, additional studies may be necessary to prove their pathogenicity, including familial segregation analysis if a family history is present 111. In addition to providing a definitive diagnosis, the knowledge of molecular etiology can facilitate family screening to implement pre-symptomatic management, since TG levels can greatly fluctuate based on dietary intake of fat.
6. Genetics of apolipoprotein C-II deficiency
Due to the rarity of apoC-II deficiency, less than 30 APOC2 mutations have been published in the literature (Table 1). Only 13 of these mutations are reported in Online Mendelian Inheritance in Man (OMIM: https://www.omim.org/entry/608083). Most are missense mutations, but nonsense, frameshift, splice-site due to deletion/insertion, loss of translational initiation due to methionine or promoter mutations have all been described (Fig. 4B; Table 1). Several APOC2 mutations are in helix 1 and presumably impair LPL activation due to poor lipoprotein binding of its cofactor apoC-II (Fig. 4A, B). Mutations present in the third helix (Fig. 4A and B) likely interfere with LPL activation 114. There are also cases of apoC-II deficient patients with clinical features but unknown genetic mutations (Table 2).
Table 1.
ApoC-II mutations associated with HTG.
| ApoC-II Variant | Mutation | Positiona, b | Type | Allele count per 100,000 | Clinical features (TG mmol/L) | Source | |
|---|---|---|---|---|---|---|---|
| 1 | APOLIPOPROTEIN C-II (-190T→A) | T→A substitution in apo C-II promoter, 190 bp up-stream of major transcription start site. 2-fold reduction in activity (after subtracting background) | Promoter (−190T→A) | Substitution | <1 | HTG (50.0), X | BBRC 2007 Mar 2;354(1):62-5 |
| 2 | APOLIPOPROTEIN C-II (−86A→G) | A→G substitution in apoC-II promoter, 86 bp upstream of major transcriptional start site. No protein expression | Promoter (−86A→G) | Substitution | <1 | HTG (18.9), C, P | J Lipid Res. 1996 37:2599-2607 |
| 3 | APOLIPOPROTEIN C-II (PARIS-1) | Met(-22)→Val/No initiation codon/Loss of signal peptide and first 8 aa | M-22V (M1V) | Absent | <1 | HTG (10.7), P | JBC 1989 Dec 15;264(35):20 839-42 |
| 4 | APOLIPOPROTEIN C-II (PARIS2, BARCELONA) | Introduction stop codon (Arg(−19))/Termination | R-19* (R4X) | Nonsense | 4 | HTG (20.3) | J Lipid Res. 1992 Mar;33(3):361-7, J Lipid Res. 1992 33:1823-1832 |
| 5 | APOLIPOPROTEIN C-II (JAPAN, VENEZUELA) | 1 bp deletion at Gln2; frameshift resulting in 17 aa truncated protein | fsQ2 (fsQ24) | Frameshift | <1 | HTG (12.4; 12.3) | Atherosclerosis 1979 Sep;34(1)53-65, Am J Hum Genet (1991) 48:383-389 |
| 6 | APOLIPOPROTEIN C-II (SHANGAI) | 1 bp deletion/2 bp insertion/Asp29Glu30→Ala29Ter30 | fsD7 (fsD29) | Frameshift | <1 | HTG (17.42), C, P | Lipids Health Disease 2016 15:12 |
| 7 | APOLIPOPROTEIN C-II (NIJMEGEN) | 1 bp deletion and frameshift at Val18/introduction stop codon | fsV18* (fsV40X) | Frameshift | <1 | HTG (4.0–15.0), H/S, LR, P | JBC 1988 Dec 5;263(34):179 13-6 |
| 8 | APOLIPOPROTEIN C-II VARIANT (ApoC-II-v) | p.Lys19Thr | K19T (K41T) | Missense | 87 | HTG (7.3;3.7), G | J. Lipid Res. 1990 31:385–396, Dis Markers 1991 Mar–Apr 9(2):73–80, But see also Clin Gen (1994) 45:292-7 |
| 9 | APOLIPOPROTEIN C-II (WAKAYAMA) | Trp26→Arg | W26R (W48R) | Missense | <1 | HTG (10.3), P | BBRC 1993 Jun 30;193(3):117 4-83 |
| 10 | APOLIPOPROTEIN C-II (BARI) | Introduction stop codon (Tyr37)/Termination | Y37* (Y59X) | Nonsense | <1 | HTG (56.5), P, X | BBRC 1990 May 16; 168:1118-1127 |
| 11 | APOLIPOPROTEIN C-II (PADOVA) | Introduction stop codon (Tyr37)/Termination | Y37* (Y59X) | Nonsense | <1 | HTG (56.5), P, X | J Clin Invest 1986 Feb;77(2)520-7, J Clin Invest 1989 Oct;84(4)1215-9 |
| 12 | APOLIPOPROTEIN C-II (SAN FRANCISCO) | p.Glu38Lys | E38K (E60K) | Missense | 59 | HTG (3.2; 2.3; 11.3) | Hum Mol Genet. 1993 Jan;2(1):69-74 |
| 13 | APOLIPOPROTEIN C-II (PHILADELPHIA) | Arg72→Thr | R50T (R72T) | Missense | <1 | HTG (33.9), P, G | J Clin Endocrinol Metab. 2017 Feb 13 |
| 14 | APOLIPOPROTEIN C-II (AFRICAN) | p.Lys55Gln | K55Q (K77Q) | Missense | 238 | HTG (63.1; 17.5; 8.8; 14.9), X, P, G, O | J Clin Invest. 1986 Feb;77(2):595-601 |
| 15 | APOLIPOPROTEIN C-II (AUCKLAND) | Tyr63Ter/Termination | Y63* (Y85X) | Nonsense | <1 | HTG (326.0; 189.8), H/S, O | Ann Neurol. 2003 Jun;53(6):807-10 |
| 16 | APOLIPOPROTEIN C-II (TORONTO) | Deletion 1 bp Thr68→ Frameshift, alteration of 6 aa and termination at aa 74 | fsT68 (fsT90) | Frameshift | <1 | HTG (107.0) | NEJM (1978) 299:1421-1424, PNAS USA (1987) 84:270-273, J Medical Genetics (1988) 25:649-652 |
| 17 | APOLIPOPROTEIN C-II (ONTARIO) | p.Gln70Ter | Q70* (Q92X) | Nonsense | <1 | HTG (>3.37) | Cir Cardiovasc Genet (2012) 5:66-72 |
| 19 | APOLIPOPROTEIN C-II (ST. MICHAEL) | Insertion 1 bp, Gln70→Pro/Frameshift 97X | fsQ70 (fsQ92) | Frameshift | 1 | HTG (15.0), P, A | J Clin Invest (1987) 80:1597-1606, Clin Biochem (1992) 25:309-312 |
| 21 | APOLIPOPROTEIN C-II (HONGKONG) | Leu72→Pro (C-terminal helix) | L72P (L94P) | Missense | <1 | HTG (82.0), X | Clin Chim Acta. 2006 Feb;364(1–2):256-9 |
| 22 | APOLIPOPROTEIN C-II (NIJMEGEN, ApoC-II-C-IV) | Deletion of promoter and Exon 1 | Del Promoter, Ex 1 | Deletion | <1 | HTG (21.0), C, P | BBRC 2000 Jul 14;273(3):108 4-7 |
| 23 | APOLIPOPROTEIN C-II (HAMBURG, TOKYO) | Intron 2 + 1 Gly→Cys/Splice defect | Splice NT 2 G+1 to C | Splice | <1 | HTG (20.9), P | J Clin Invest. 1988 Nov;82(5):148 9-94, Atherosclerosis 1997 Apr;130(1–2):153-60 |
| 24 | APOLIPOPROTEIN C-II (TUZLA) | Homozygous deletion of exons 2, 3 and 4 | Del Ex 2, 3, 4 | Deletion | <1 | C (52.6), O | Clin Chim Acta. 2015 Jan 1;438:148-53 |
HTG, hypertriglyceridemia; X, xanthomas;C, chylomicronemia; P, pancreatitis; H/S, hepatosplenomegaly; LR, lipemia retinalis;
G, hyperglycemia, glucose intolerance or diabetes mellitus; O, other features (coronary, cerebral and/or iliofemoral atherosclerosis, hypertension, rupture of aortic aneurysm, macrocephaly, neurologically abnormalities, developmental delay, vomiting) ; A, atherosclerosis; (), triglycerides levels reported in mmol/L.
#1 is first amino acid of mature peptide after signal peptide cleavage
#1 is initiator methionine; first amino acid of protein before signal peptide cleavage (HGVS-recommended nomenclature)
or X: STOP codon.
Table 2.
ApoC-II deficiencies with unknown genetic mutations
| ApoC-II Variant | Mutation | Clinical features (TG mmol/L) | Source | |
|---|---|---|---|---|
| 1 | APOLIPOPROTEIN C-II (BETHESDA) | Unknown | HTG (13.6) | J Lipid Res. 1988 Mar;29(3):273-8 |
| 2 | APOLIPOPROTEIN C-II (TOKYO #2) | Unknown | HTG (36.5), P | J Atheroscler Thromb. 2013;20(5):481-93 |
| 3 | APOLIPOPROTEIN C-II (TOKYO #3) | Unknown | HTG (172.6), C, X, LR | Eur J Pediatr 1989 Apr;148(6):550-2 |
| 4 | APOLIPOPROTEIN C-II (TOSHIGI) | Unknown | HTG (6.1), G, CAD | Clin Exp Med 2002 May;2(1):29-31 |
HTG, hypertriglyceridemia; P, pancreatitis; C, chylomicronemia; X, xanthomas; LR, lipemia retinalis; G, Hyperglycemia, glucose intolerance or diabetes mellitus; CAD, coronary artery disease; (),triglycerides levels reported in mmol/L.
Four mutations, namely K55Q, K19T, E38K and R-19* (HGVS-recommended nomenclature: K77Q, K41T, E60K, and R4*) (Fig. 4B; Table 1) account for more than half of all APOC2 mutations found in the ExAc exome consortium database of over 60,000 individuals (http://exac.broadinstitute.org/gene/ENSG00000234906). Based on the frequencies of these 4 alleles alone, approximately 1.5 individual per 100,000 in the general population would be expected to have bi-allelic apoC-II deficiency. This is much higher than an estimated prevalence derived from published case reports and reported DNA sequencing results of patients with HTG113. It is plausible that apoC-II deficiency, as with many other rare diseases, is under-diagnosed. There may also be an ascertainment bias leading to a falsely elevated frequency in the ExAc exome database, which collected the data from several studies on CVD and dyslipidemia. Furthermore, the K55Q and E38K (HGVS: K77Q and E60K) mutations are much more common in individuals from African descent, which suggests a greater frequency of apoC-II deficiency in this population and less in the general population.
7. Clinical therapy for apolipoprotein C-II deficiency
The main goal of therapy in FCS, including apoC-II deficiency, is to prevent acute pancreatitis, the most serious morbidity and mortality in this condition. Because conventional TG-lowering medications are largely ineffective in FCS, medical nutrition therapy is absolutely essential. A fat-restricted dietary regimen with fat less than 15% of total daily calories or <20 g/day is usually recommended 115. Suppressing TG to <11.3 mmol/L (<1,000 mg/dL) lowers pancreatitis risk considerably, and guidelines typically advise suppression <5.65 mmol/L (<500 mg/dL) when feasible 116. Medium-chain TG may be used for cooking to improve palatability, since they are not incorporated into CMs.
Fibrates and/or niacin are regularly added as adjunct therapy, since they may result in some additional TG lowering 117, although their effectiveness is unreliable in this disorder. Fish oil supplements should be used with caution; though they have been used successfully in some individuals 118, they might also contribute to CM levels 119.
Since it is very difficult for patients to maintain a long-term fat-restricted diet, novel therapies are urgently needed in FCS. Taking advantage of the multiple regulatory mechanisms of LPL function, specifically, inhibitors of LPL, apoC-III and ANGPTL3, have all emerged as potential targets to enhance the catabolism of TRL. Antisense-inhibitors of APOC3 120, 121, and ANGPTL3 122, 123, as well as a monoclonal antibody to ANGPTL3 have been extensively investigated in recent years. Even though an antisense oligonucleotide therapy has already been shown to be effective in LPL deficiency 121, its efficacy in apoC-II deficiency has not been documented.
There are also several other targets which help to lower HTG, some of which are under investigation. Microsomal triglyceride transfer protein (MTP) inhibitor, lomitapide, has a limited approval for use in homozygous familial hypercholesterolemia. In theory, it could be re-purposed to decrease VLDL production 90. However, this would be well outside the label and not without significant risk; therefore, we think it is inadvisable outside extraordinary circumstances. To reduce TG synthesis, inhibitors of diacylglycerol O-acyltransferase 1 (DGAT1) may be potential treatment options for FCS 90. Recently, Millar et al. 124 have shown that anacetrapib treatment alone and in conjunction with atorvastatin increased apoC-II but also apoC-III levels and had no effect on direct clearance of VLDL.
Particularly for apoC-II deficiency, therapeutic plasma exchange with donor plasma that contains apoC-II is the most direct therapy and can be life-saving during severe pancreatitis episodes. This can also be implemented as a prophylactic therapy to rapidly lower TG if needed 125. The recent development of an apoC-II mimetic peptide, which incorporates the third helix of apoC-II, has shown some promise as potential future therapy. Intravenous injection of this peptide normalized TG in a mouse model of apoC-II deficiency 81. Furthermore, the apoC-II mimetic peptide potentiated LPL activity in other non-apoC-II deficient HTG conditions, suggesting a potential broader utility 93.
8. Conclusions
ApoC-II is an important cofactor for the full activation and enzymatic activity of LPL in the hydrolysis of TG on TRL. ApoC-II mutations can promote a prodigious rise in plasma TG levels due to chylomicronemia and consequently can provoke acute pancreatitis. In addition to apoC-II, there are several other important regulators of TRL metabolism, including apoC-III, apoA-V, ANGPTL 3, 4 and 8. Therefore, TRL metabolism is a complex and multifaceted process that involves numerous proteins, as well as complex gene regulation.
Although great strides have been made in understanding the biology and biological role of apoC-II in TRL metabolism, there is undoubtedly much more to be investigated. Not only it is important in identifying novel targets for the treatment of HTG, but also in understanding the relationship between HTG and CVD risks. Currently, the most important therapy for apoC-II deficiency remains to be medical nutrition therapy with a strict low-fat diet. The inhibition of several LPL regulatory protein targets currently being investigated may soon provide better treatment options for patients with apoC-II deficiency and other forms of FCS. Finally, investigating TRL metabolism may not only uncover additional targets for FCS and the more common forms of HTG, but also may better reveal the relationship between TG metabolism and CVD risk.
Acknowledgments
The authors thank Maureen Sampson for help in illustration of Fig. 1A, 1B, and 2.
Financial support
This research was supported by the Intramural Research Program of the National Heart, Lung, and Blood Institute (NHLBI) at National Institutes of Health.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Conflict of interest
A.W., M. J. A., D. O. S., and A. T. R. are co-inventors on US patents for apoC-II mimetic peptides. A. W., D. O. S., and A. T. R. have a research cooperation agreement for apoC-II mimetic peptides with Corvidia Therapeutics, Inc., Waltham, MA, USA. R. L. D. has received grant support and modest honoraria from Ionis Pharmaceuticals and grant support from uniQure, Regeneron Pharmaceuticals, Zydus Pharmaceuticals, and Kowa Research Institute. L. F. and M. U. declare no conflict of interest.
References
- 1.Toth PP. Triglyceride-rich lipoproteins as a causal factor for cardiovascular disease. Vascular Health and Risk Management. 2016;12:171–183. doi: 10.2147/VHRM.S104369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Nordestgaard BG, Varbo A. Triglycerides and cardiovascular disease. The Lancet. 2014;384:626–635. doi: 10.1016/S0140-6736(14)61177-6. [DOI] [PubMed] [Google Scholar]
- 3.Jørgensen AB, Frikke-Schmidt R, West AS, et al. Genetically elevated non-fasting triglycerides and calculated remnant cholesterol as causal risk factors for myocardial infarction. European Heart Journal. 2013;34 doi: 10.1093/eurheartj/ehs431. [DOI] [PubMed] [Google Scholar]
- 4.Xiao C, Dash S, Morgantini C, et al. New and emerging regulators of intestinal lipoprotein secretion. Atherosclerosis. 2014;233:608–615. doi: 10.1016/j.atherosclerosis.2013.12.047. [DOI] [PubMed] [Google Scholar]
- 5.Varbo A, Benn M, Tybjærg-Hansen A, et al. Remnant cholesterol as a causal risk factor for ischemic heart disease. Journal of the American College of Cardiology. 2013;61:427–436. doi: 10.1016/j.jacc.2012.08.1026. [DOI] [PubMed] [Google Scholar]
- 6.Korn ED. Clearing factor, a heparin-activated lipoprotein lipase. I. Isolation and characterization of the enzyme from normal rat heart. J Biol Chem. 1955;215:1–14. [PubMed] [Google Scholar]
- 7.Brown WV, Levy RI, Fredrickson DS. Further characterization of apolipoproteins from the human plasma very low density lipoproteins. J Biol Chem. 1970;245:6588–6594. [PubMed] [Google Scholar]
- 8.LaRosa JC, Levy RI, Herbert P, et al. A specific apoprotein activator for lipoprotein lipase. Biochemical and Biophysical Research Communications. 1970;41:57–62. doi: 10.1016/0006-291x(70)90468-7. [DOI] [PubMed] [Google Scholar]
- 9.Breckenridge WC, Little JA, Steiner G, et al. Hypertriglyceridemia associated with deficiency of apolipoprotein C-II. New England Journal of Medicine. 1978;298:1265–1273. doi: 10.1056/NEJM197806082982301. [DOI] [PubMed] [Google Scholar]
- 10.Hegele RA, Ginsberg HN, Chapman MJ, et al. The polygenic nature of hypertriglyceridaemia: implications for definition, diagnosis, and management. Lancet Diabetes Endocrinol. 2014;2:655–666. doi: 10.1016/S2213-8587(13)70191-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Adiels M, Olofsson SO, Taskinen MR, et al. Overproduction of very low-density lipoproteins is the hallmark of the dyslipidemia in the metabolic syndrome, Arteriosclerosis. Thrombosis, and Vascular Biology. 2008;28:1225–1236. doi: 10.1161/ATVBAHA.107.160192. [DOI] [PubMed] [Google Scholar]
- 12.Tiwari S, Siddiqi SA. Intracellular trafficking and secretion of VLDL, Arteriosclerosis. Thrombosis, and Vascular Biology. 2012;32:1079–1086. doi: 10.1161/ATVBAHA.111.241471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Barrows BR, Parks EJ. Contributions of different fatty acid sources to very low-density lipoprotein-triacylglycerol in the fasted and fed states. Journal of Clinical Endocrinology and Metabolism. 2006;91:1446–1452. doi: 10.1210/jc.2005-1709. [DOI] [PubMed] [Google Scholar]
- 14.Goldberg IJ, Merkel M. Lipoprotein lipase: physiology, biochemistry, and molecular biology. Frontiers in bioscience : a journal and virtual library. 2001;6 doi: 10.2741/goldberg. [DOI] [PubMed] [Google Scholar]
- 15.Mead JR, Irvine SA, Ramji DP. Lipoprotein lipase: Structure, function, regulation and role in disease. Journal of Molecular Medicine. 2002;80:753–769. doi: 10.1007/s00109-002-0384-9. [DOI] [PubMed] [Google Scholar]
- 16.Doolittle MH, Neher SB, Ben-Zeev O, et al. Lipase maturation factor LMF1, membrane topology and interaction with lipase proteins in the endoplasmic reticulum. J Biol Chem. 2009;284:33623–33633. doi: 10.1074/jbc.M109.049395. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 17.Ben-Zeev O, Hosseini M, Lai CM, et al. Lipase maturation factor 1 is required for endothelial lipase activity. J Lipid Res. 2011;52:1162–1169. doi: 10.1194/jlr.M011155. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 18.Sha H, Sun S, Francisco AB, et al. The ER-associated degradation adaptor protein Sel1L regulates LPL secretion and lipid metabolism. Cell Metab. 2014;20:458–470. doi: 10.1016/j.cmet.2014.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Braun JEA, Severson DL. Regulation of the synthesis processing and translocation of lipoprotein lipase. Biochemical Journal. 1992;287:337–347. doi: 10.1042/bj2870337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Allan CM, Larsson M, Jung RS, et al. Mobility of “HSPG-bound” LPL explains how LPL is able to reach GPIHBP1 on capillaries. J Lipid Res. 2017;58:216–225. doi: 10.1194/jlr.M072520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Beigneux AP, Davies BSJ, Gin P, et al. Glycosylphosphatidylinositol-Anchored High-Density Lipoprotein-Binding Protein 1 Plays a Critical Role in the Lipolytic Processing of Chylomicrons. Cell Metabolism. 2007;5:279–291. doi: 10.1016/j.cmet.2007.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Davies BSJ, Beigneux AP, Barnes RH, Ii, et al. GPIHBP1 is responsible for the entry of lipoprotein lipase into capillaries. Cell Metabolism. 2010;12:42–52. doi: 10.1016/j.cmet.2010.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Goulbourne CN, Gin P, Tatar A, et al. The GPIHBP1-LPL complex is responsible for the margination of triglyceride-rich lipoproteins in capillaries. Cell Metabolism. 2014;19:849–860. doi: 10.1016/j.cmet.2014.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Mysling S, Kristensen KK, Larsson M, et al. The acidic domain of the endothelial membrane protein GPIHBP1 stabilizes lipoprotein lipase activity by preventing unfolding of its catalytic domain. eLife. 2016;5 doi: 10.7554/eLife.12095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Kersten S. Physiological regulation of lipoprotein lipase. Biochim Biophys Acta. 2014;1841:919–933. doi: 10.1016/j.bbalip.2014.03.013. [DOI] [PubMed] [Google Scholar]
- 26.Lookene A, Beckstead JA, Nilsson S, et al. Apolipoprotein A-V-heparin interactions: Implications for plasma lipoprotein metabolism. Journal of Biological Chemistry. 2005;280:25383–25387. doi: 10.1074/jbc.M501589200. [DOI] [PubMed] [Google Scholar]
- 27.Gonzales JC, Gordts PL, Foley EM, et al. Apolipoproteins E and AV mediate lipoprotein clearance by hepatic proteoglycans. J Clin Invest. 2013;123:2742–2751. doi: 10.1172/JCI67398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Gin P, Yin L, Davies BSJ, et al. The acidic domain of GPIHBP1 is important for the binding of lipoprotein lipase and chylomicrons. Journal of Biological Chemistry. 2008;283:29554–29562. doi: 10.1074/jbc.M802579200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Ekman R, Nilsson-Ehle P. Effects of apolipoproteins on lipoprotein lipase activity of human adipose tissue. Clin Chim Acta. 1975;63:29–35. doi: 10.1016/0009-8981(75)90374-5. [DOI] [PubMed] [Google Scholar]
- 30.Brown WV, Baginsky ML. Inhibition of lipoprotein lipase by an apoprotein of human very low density lipoprotein. Biochem Biophys Res Commun. 1972;46:375–382. doi: 10.1016/s0006-291x(72)80149-9. [DOI] [PubMed] [Google Scholar]
- 31.Berbee JF, van der Hoogt CC, Sundararaman D, et al. Severe hypertriglyceridemia in human APOC1 transgenic mice is caused by apoC-I-induced inhibition of LPL. J Lipid Res. 2005;46:297–306. doi: 10.1194/jlr.M400301-JLR200. [DOI] [PubMed] [Google Scholar]
- 32.Jong MC, Rensen PC, Dahlmans VE, et al. Apolipoprotein C-III deficiency accelerates triglyceride hydrolysis by lipoprotein lipase in wild-type and apoE knockout mice. J Lipid Res. 2001;42:1578–1585. [PubMed] [Google Scholar]
- 33.Larsson M, Vorrsjo E, Talmud P, et al. Apolipoproteins C-I and C-III inhibit lipoprotein lipase activity by displacement of the enzyme from lipid droplets. Journal of Biological Chemistry. 2013;288:33997–34008. doi: 10.1074/jbc.M113.495366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ooi EM, Chan DC, Hodson L, et al. Triglyceride-rich lipoprotein metabolism in women: roles of apoC-II and apoC-III. Eur J Clin Invest. 2016;46:730–736. doi: 10.1111/eci.12657. [DOI] [PubMed] [Google Scholar]
- 35.Sehayek E, Eisenberg S. Mechanisms of inhibition by apolipoprotein C of apolipoprotein E-dependent cellular metabolism of human triglyceride-rich lipoproteins through the low density lipoprotein receptor pathway. J Biol Chem. 1991;266:18259–18267. [PubMed] [Google Scholar]
- 36.Hato T, Tabata M, Oike Y. The Role of Angiopoietin-Like Proteins in Angiogenesis and Metabolism. Trends in Cardiovascular Medicine. 2008;18:6–14. doi: 10.1016/j.tcm.2007.10.003. [DOI] [PubMed] [Google Scholar]
- 37.Mattijssen F, Kersten S. Regulation of triglyceride metabolism by Angiopoietin-like proteins. Biochim Biophys Acta. 2012;1821:782–789. doi: 10.1016/j.bbalip.2011.10.010. [DOI] [PubMed] [Google Scholar]
- 38.Buttet M, Traynard V, Tran TT, et al. From fatty-acid sensing to chylomicron synthesis: role of intestinal lipid-binding proteins. Biochimie. 2014;96:37–47. doi: 10.1016/j.biochi.2013.08.011. [DOI] [PubMed] [Google Scholar]
- 39.Beisiegel U, Weber W, Bengtsson-Olivecrona G. Lipoprotein lipase enhances the binding of chylomicrons to low density lipoprotein receptor-related protein. Proc Natl Acad Sci U S A. 1991;88:8342–8346. doi: 10.1073/pnas.88.19.8342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Tall AR, Small DM. Plasma high-density lipoproteins. N Engl J Med. 1978;299:1232–1236. doi: 10.1056/NEJM197811302992207. [DOI] [PubMed] [Google Scholar]
- 41.Zannis VI, Cole FS, Jackson CL, et al. Distribution of apolipoprotein A-I, C-II, C-III, and E mRNA in fetal human tissues. Time-dependent induction of apolipoprotein E mRNA by cultures of human monocyte-macrophages. Biochemistry. 1985;24:4450–4455. doi: 10.1021/bi00337a028. [DOI] [PubMed] [Google Scholar]
- 42.Wu AL, Windmueller HG. Relative contributions by liver and intestine to individual plasma apolipoproteins in the rat. J Biol Chem. 1979;254:7316–7322. [PubMed] [Google Scholar]
- 43.Provost PR, Boucher E, Tremblay Y. Apolipoprotein A-I, A-II, C-II, and H expression in the developing lung and sex difference in surfactant lipids. J Endocrinol. 2009;200:321–330. doi: 10.1677/JOE-08-0238. [DOI] [PubMed] [Google Scholar]
- 44.Provost PR, Tremblay Y. Elevated expression of four apolipoprotein genes during the 32–35 week gestation window in the human developing lung. Early Hum Dev. 2010;86:529–534. doi: 10.1016/j.earlhumdev.2010.06.013. [DOI] [PubMed] [Google Scholar]
- 45.Cote M, Provost PR, Gerard-Hudon MC, et al. Apolipoprotein C-II and lipoprotein lipase show a temporal and geographic correlation with surfactant lipid synthesis in preparation for birth. BMC Dev Biol. 2010;10:111. doi: 10.1186/1471-213X-10-111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Li CM, Clark ME, Chimento MF, et al. Apolipoprotein localization in isolated drusen and retinal apolipoprotein gene expression. Invest Ophthalmol Vis Sci. 2006;47:3119–3128. doi: 10.1167/iovs.05-1446. [DOI] [PubMed] [Google Scholar]
- 47.Fessler MB. The challenges and promise of targeting the Liver X Receptors for treatment of inflammatory disease. Pharmacol Ther. 2017 doi: 10.1016/j.pharmthera.2017.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Lee SD, Tontonoz P. Liver X receptors at the intersection of lipid metabolism and atherogenesis. Atherosclerosis. 2015;242:29–36. doi: 10.1016/j.atherosclerosis.2015.06.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Green DS, Young HA, Valencia JC. Current prospects of type II interferon gamma signaling and autoimmunity. J Biol Chem. 2017;292:13925–13933. doi: 10.1074/jbc.R116.774745. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Buttet M, Poirier H, Traynard V, et al. Deregulated Lipid Sensing by Intestinal CD36 in Diet-Induced Hyperinsulinemic Obese Mouse Model. PLoS One. 2016;11:e0145626. doi: 10.1371/journal.pone.0145626. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Shachter NS, Zhu Y, Walsh A, et al. Localization of a liver-specific enhancer in the apolipoprotein E/C-I/C-II gene locus. J Lipid Res. 1993;34:1699–1707. [PubMed] [Google Scholar]
- 52.Allan CM, Taylor S, Taylor JM. Two hepatic enhancers, HCR.1 and HCR.2, coordinate the liver expression of the entire human apolipoprotein E/C-I/C-IV/C-II gene cluster. J Biol Chem. 1997;272:29113–29119. doi: 10.1074/jbc.272.46.29113. [DOI] [PubMed] [Google Scholar]
- 53.Allan CM, Walker D, Taylor JM. Evolutionary duplication of a hepatic control region in the human apolipoprotein E gene locus. Identification of a second region that confers high level and liver-specific expression of the human apolipoprotein E gene in transgenic mice. J Biol Chem. 1995;270:26278–26281. doi: 10.1074/jbc.270.44.26278. [DOI] [PubMed] [Google Scholar]
- 54.Kast HR, Nguyen CM, Sinal CJ, et al. Farnesoid X-activated receptor induces apolipoprotein C-II transcription: a molecular mechanism linking plasma triglyceride levels to bile acids. Mol Endocrinol. 2001;15:1720–1728. doi: 10.1210/mend.15.10.0712. [DOI] [PubMed] [Google Scholar]
- 55.Mak PA, Laffitte BA, Desrumaux C, et al. Regulated expression of the apolipoprotein E/C-I/C-IV/C-II gene cluster in murine and human macrophages. A critical role for nuclear liver X receptors alpha and beta. J Biol Chem. 2002;277:31900–31908. doi: 10.1074/jbc.M202993200. [DOI] [PubMed] [Google Scholar]
- 56.Kardassis D, Roussou A, Papakosta P, et al. Synergism between nuclear receptors bound to specific hormone response elements of the hepatic control region-1 and the proximal apolipoprotein C-II promoter mediate apolipoprotein C-II gene regulation by bile acids and retinoids. Biochem J. 2003;372:291–304. doi: 10.1042/BJ20021532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Hayhurst GP, Lee YH, Lambert G, et al. Hepatocyte nuclear factor 4alpha (nuclear receptor 2A1) is essential for maintenance of hepatic gene expression and lipid homeostasis. Mol Cell Biol. 2001;21:1393–1403. doi: 10.1128/MCB.21.4.1393-1403.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Ramasamy I. Update on the molecular biology of dyslipidemias. Clin Chim Acta. 2016;454:143–185. doi: 10.1016/j.cca.2015.10.033. [DOI] [PubMed] [Google Scholar]
- 59.Vorgia P, Zannis VI, Kardassis D. A short proximal promoter and the distal hepatic control region-1 (HCR-1) contribute to the liver specificity of the human apolipoprotein C-II gene. Hepatic enhancement by HCR-1 requires two proximal hormone response elements which have different binding specificities for orphan receptors HNF-4, ARP-1 and EAR-2. J Biol Chem. 1998;273:4188–4196. doi: 10.1074/jbc.273.7.4188. [DOI] [PubMed] [Google Scholar]
- 60.Streicher R, Geisel J, Weisshaar C, et al. A single nucleotide substitution in the promoter region of the apolipoprotein C-II gene identified in individuals with chylomicronemia. J Lipid Res. 1996;37:2599–2607. [PubMed] [Google Scholar]
- 61.Kardassis D, Sacharidou E, Zannis VI. Transactivation of the human apolipoprotein CII promoter by orphan and ligand-dependent nuclear receptors. The regulatory element CIIC is a thyroid hormone response element. J Biol Chem. 1998;273:17810–17816. doi: 10.1074/jbc.273.28.17810. [DOI] [PubMed] [Google Scholar]
- 62.Li P, Ruan X, Yang L, et al. A liver-enriched long non-coding RNA, lncLSTR, regulates systemic lipid metabolism in mice. Cell Metab. 2015;21:455–467. doi: 10.1016/j.cmet.2015.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Streicher R, Avci H, Munck M, et al. Structure of the human apolipoprotein C-II gene promoter. Z Gastroenterol. 1996;34(Suppl 3):49–50. [PubMed] [Google Scholar]
- 64.Lee AH. The role of CREB-H transcription factor in triglyceride metabolism. Curr Opin Lipidol. 2012;23:141–146. doi: 10.1097/MOL.0b013e3283508fed. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Andersson Y, Majd Z, Lefebvre AM, et al. Developmental and pharmacological regulation of apolipoprotein C-II gene expression. Comparison with apo C-I and apo C-III gene regulation. Arterioscler Thromb Vasc Biol. 1999;19:115–121. doi: 10.1161/01.atv.19.1.115. [DOI] [PubMed] [Google Scholar]
- 66.Daniel Z, Swali A, Emes R, et al. The effect of maternal undernutrition on the rat placental transcriptome: protein restriction up-regulates cholesterol transport. Genes Nutr. 2016;11:27. doi: 10.1186/s12263-016-0541-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Kei AA, Filippatos TD, Tsimihodimos V, et al. A review of the role of apolipoprotein C-II in lipoprotein metabolism and cardiovascular disease. Metabolism: Clinical and Experimental. 2012;61:906–921. doi: 10.1016/j.metabol.2011.12.002. [DOI] [PubMed] [Google Scholar]
- 68.Rhee J, Ge H, Yang W, et al. Partnership of PGC-1alpha and HNF4alpha in the regulation of lipoprotein metabolism. J Biol Chem. 2006;281:14683–14690. doi: 10.1074/jbc.M512636200. [DOI] [PubMed] [Google Scholar]
- 69.Takahashi M, Yagyu H, Tazoe F, et al. Macrophage lipoprotein lipase modulates the development of atherosclerosis but not adiposity. J Lipid Res. 2013;54:1124–1134. doi: 10.1194/jlr.M035568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Trusca VG, Fuior EV, Florea IC, et al. Macrophage-specific up-regulation of apolipoprotein E gene expression by STAT1 is achieved via long range genomic interactions. J Biol Chem. 2011;286:13891–13904. doi: 10.1074/jbc.M110.179572. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Shih DM, Xia YR, Wang XP, et al. Combined serum paraoxonase knockout/apolipoprotein E knockout mice exhibit increased lipoprotein oxidation and atherosclerosis. J Biol Chem. 2000;275:17527–17535. doi: 10.1074/jbc.M910376199. [DOI] [PubMed] [Google Scholar]
- 72.Grehan S, Tse E, Taylor JM. Two distal downstream enhancers direct expression of the human apolipoprotein E gene to astrocytes in the brain. J Neurosci. 2001;21:812–822. doi: 10.1523/JNEUROSCI.21-03-00812.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Remaley AT, Stonik JA, Demosky SJ, et al. Apolipoprotein specificity for lipid efflux by the human ABCAI transporter. Biochem Biophys Res Commun. 2001;280:818–823. doi: 10.1006/bbrc.2000.4219. [DOI] [PubMed] [Google Scholar]
- 74.Melhuish TA, Chung DD, Bjerke GA, et al. Tgif1 represses apolipoprotein gene expression in liver. J Cell Biochem. 2010;111:380–390. doi: 10.1002/jcb.22713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Chun EM, Park YJ, Kang HS, et al. Expression of the apolipoprotein C-II gene during myelomonocytic differentiation of human leukemic cells. J Leukoc Biol. 2001;69:645–650. [PubMed] [Google Scholar]
- 76.Trusca VG, Florea IC, Kardassis D, et al. STAT1 interacts with RXRalpha to upregulate ApoCII gene expression in macrophages. PLoS One. 2012;7:e40463. doi: 10.1371/journal.pone.0040463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Wu AL, Windmueller HG. Identification of circulating apolipoproteins synthesized by rat small intestine in vivo. J Biol Chem. 1978;253:2525–2528. [PubMed] [Google Scholar]
- 78.Capurso A, Mogavero AM, Resta F, et al. Apolipoprotein C-II deficiency: detection of immunoreactive apolipoprotein C-II in the intestinal mucosa of two patients. J Lipid Res. 1988;29:703–711. [PubMed] [Google Scholar]
- 79.Schonfeld G, Weidman SW, Witztum JL, et al. Alterations in levels and interrelations of plasma apolipoproteins induced by diet. Metabolism. 1976;25:261–275. doi: 10.1016/0026-0495(76)90084-6. [DOI] [PubMed] [Google Scholar]
- 80.Hortin GL, Remaley AT. Mass determination of major plasma proteins by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. Clinical Proteomics. 2006;2:103–115. [Google Scholar]
- 81.Sakurai T, Sakurai A, Vaisman BL, et al. Creation of Apolipoprotein C-II (ApoC-II) Mutant Mice and Correction of Their Hypertriglyceridemia with an ApoC-II Mimetic Peptide. Journal of Pharmacology and Experimental Therapeutics. 2016;356:341–353. doi: 10.1124/jpet.115.229740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Hortin GL, Sviridov D, Anderson NL. High-abundance polypeptides of the human plasma proteome comprising the top 4 logs of polypeptide abundance. Clin Chem. 2008;54:1608–1616. doi: 10.1373/clinchem.2008.108175. [DOI] [PubMed] [Google Scholar]
- 83.Sakurabayashi I, Saito Y, Kita T, et al. Reference intervals for serum apolipoproteins A-I, A-II, B, C-II, C-III, and E in healthy Japanese determined with a commercial immunoturbidimetric assay and effects of sex, age, smoking, drinking, and Lp(a) level. Clin Chim Acta. 2001;312:87–95. doi: 10.1016/s0009-8981(01)00591-5. [DOI] [PubMed] [Google Scholar]
- 84.Beliard S, Nogueira JP, Maraninchi M, et al. Parallel increase of plasma apoproteins C-II and C-III in Type 2 diabetic patients. Diabet Med. 2009;26:736–739. doi: 10.1111/j.1464-5491.2009.02757.x. [DOI] [PubMed] [Google Scholar]
- 85.Grutzmacher P, Marz W, Peschke B, et al. Lipoproteins and apolipoproteins during the progression of chronic renal disease. Nephron. 1988;50:103–111. doi: 10.1159/000185138. [DOI] [PubMed] [Google Scholar]
- 86.MacRaild CA, Hatters DM, Howlett GJ, et al. NMR Structure of Human Apolipoprotein C-II in the Presence of Sodium Dodecyl Sulfate. Biochemistry. 2001;40:5414–5421. doi: 10.1021/bi002821m. [DOI] [PubMed] [Google Scholar]
- 87.Zdunek J, Martinez GV, Schleucher J, et al. Global structure and dynamics of human apolipoprotein CII in complex with micelles: Evidence for increased mobility of the helix involved in the activation of lipoprotein lipase. Biochemistry. 2003;42:1872–1889. doi: 10.1021/bi0267184. [DOI] [PubMed] [Google Scholar]
- 88.Segrest JP, Jones MK, De Loof H, et al. The amphipathic helix in the exchangeable apolipoproteins: a review of secondary structure and function. J Lipid Res. 1992;33:141–166. [PubMed] [Google Scholar]
- 89.Musliner TA, Herbert PN, Church EC. Activation of lipoprotein lipase by native and acylated peptides of apolipoprotein C-II. Biochim Biophys Acta. 1979;573:501–509. doi: 10.1016/0005-2760(79)90224-8. [DOI] [PubMed] [Google Scholar]
- 90.Brahm AJ, Hegele RA. Chylomicronaemia–current diagnosis and future therapies. Nat Rev Endocrinol. 2015;11:352–362. doi: 10.1038/nrendo.2015.26. [DOI] [PubMed] [Google Scholar]
- 91.Hegele RA, Pollex RL. Hypertriglyceridemia: phenomics and genomics. Mol Cell Biochem. 2009;326:35–43. doi: 10.1007/s11010-008-0005-1. [DOI] [PubMed] [Google Scholar]
- 92.Shen Y, Lookene A, Zhang L, et al. Site-directed mutagenesis of apolipoprotein CII to probe the role of its secondary structure for activation of lipoprotein lipase. Journal of Biological Chemistry. 2010;285:7484–7492. doi: 10.1074/jbc.M109.022046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Amar MJ, Sakurai T, Sakurai-Ikuta A, et al. A novel apolipoprotein C-II mimetic peptide that activates lipoprotein lipase and decreases serum triglycerides in apolipoprotein E-knockout mice. J Pharmacol Exp Ther. 2015;352:227–235. doi: 10.1124/jpet.114.220418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Shen Y, Lookene A, Nilsson S, et al. Functional analyses of human apolipoprotein CII by site-directed mutagenesis: Identification of residues important for activation of lipoprotein lipase. Journal of Biological Chemistry. 2002;277:4334–4342. doi: 10.1074/jbc.M105421200. [DOI] [PubMed] [Google Scholar]
- 95.McIlhargey TL, Yang Y, Wong H, et al. Identification of a lipoprotein lipase cofactor-binding site by chemical cross-linking and transfer of apolipoprotein C-II-responsive lipolysis from lipoprotein lipase to hepatic lipase. J Biol Chem. 2003;278:23027–23035. doi: 10.1074/jbc.M300315200. [DOI] [PubMed] [Google Scholar]
- 96.Dugi KA, Dichek HL, Santamarina-Fojo S. Human hepatic and lipoprotein lipase: the loop covering the catalytic site mediates lipase substrate specificity. J Biol Chem. 1995;270:25396–25401. doi: 10.1074/jbc.270.43.25396. [DOI] [PubMed] [Google Scholar]
- 97.Hill JS, Yang D, Nikazy J, et al. Subdomain chimeras of hepatic lipase and lipoprotein lipase. Localization of heparin and cofactor binding. J Biol Chem. 1998;273:30979–30984. doi: 10.1074/jbc.273.47.30979. [DOI] [PubMed] [Google Scholar]
- 98.Meyers NL, Larsson M, Olivecrona G, et al. A pressure-dependent model for the regulation of lipoprotein lipase by apolipoprotein C-II. Journal of Biological Chemistry. 2015;290:18029–18044. doi: 10.1074/jbc.M114.629865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Stroes E, Moulin P, Parhofer KG, et al. Diagnostic algorithm for familial chylomicronemia syndrome. Atheroscler Suppl. 2017;23:1–7. doi: 10.1016/j.atherosclerosissup.2016.10.002. [DOI] [PubMed] [Google Scholar]
- 100.Johansen CT, Wang J, McIntyre AD, et al. Excess of rare variants in non-genome-wide association study candidate genes in patients with hypertriglyceridemia. Circ Cardiovasc Genet. 2012;5:66–72. doi: 10.1161/CIRCGENETICS.111.960864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Mantel-Teeuwisse AK, Kloosterman JM, Maitland-van der Zee AH, et al. Drug-Induced lipid changes: a review of the unintended effects of some commonly used drugs on serum lipid levels. Drug Saf. 2001;24:443–456. doi: 10.2165/00002018-200124060-00003. [DOI] [PubMed] [Google Scholar]
- 102.Herink M, I M. Medication Induced Changes in Lipid and Lipoproteins. In: De Groot LJ, C G, Dungan K, et al., editors. Endotext (Internet) MDText.com, Inc.; South Dartmouth (MA): Updated 2015 Jul 27, https://www.ncbi.nlm.nih.gov/books/NBK326739/ [Google Scholar]
- 103.Rahalkar AR, Hegele RA. Monogenic pediatric dyslipidemias: classification genetics and clinical spectrum. Mol Genet Metab. 2008;93:282–294. doi: 10.1016/j.ymgme.2007.10.007. [DOI] [PubMed] [Google Scholar]
- 104.Brunzell JD, D S. Familial Lipoprotein Lipase Deficiency, Apo C-II Deficiency, and Hepatic Lipase Deficiency. In: David Valle M, Beaudet Arthur L, Vogelstein Bert, Kinzler Kenneth W, Antonarakis Stylianos E, Ballabio Andrea, Gibson K Michael, Mitchell Grant, editors. The Online Metabolic and Molecular Bases of Inherited Disease, online. McGraw-Hill Medical; 2016. [Google Scholar]
- 105.Reichlin M, Fesmire J, Quintero-Del-Rio AI, et al. Autoantibodies to lipoprotein lipase and dyslipidemia in systemic lupus erythematosus. Arthritis Rheum. 2002;46:2957–2963. doi: 10.1002/art.10624. [DOI] [PubMed] [Google Scholar]
- 106.Yamamoto H, Tanaka M, Yoshiga S, et al. Autoimmune severe hypertriglyceridemia induced by anti-apolipoprotein C-II antibody. J Clin Endocrinol Metab. 2014;99:1525–1530. doi: 10.1210/jc.2013-3619. [DOI] [PubMed] [Google Scholar]
- 107.Young E, Stickrath C, McNulty M, et al. Residents’ Exposure to Educational Experiences in Facilitating Hospital Discharges. J Grad Med Educ. 2017;9:184–189. doi: 10.4300/JGME-D-16-00503.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.de Graaf J, Couture P, Sniderman A. A diagnostic algorithm for the atherogenic apolipoprotein B dyslipoproteinemias. Nat Clin Pract Endocrinol Metab. 2008;4:608–618. doi: 10.1038/ncpendmet0982. [DOI] [PubMed] [Google Scholar]
- 109.Fojo SS, de Gennes JL, Beisiegel U, et al. Molecular genetics of apoC-II and lipoprotein lipase deficiency. Adv Exp Med Biol. 1991;285:329–333. doi: 10.1007/978-1-4684-5904-3_40. [DOI] [PubMed] [Google Scholar]
- 110.Di Filippo M, Marçais C, Charrière S, et al. Post-heparin LPL activity measurement using VLDL as a substrate: a new robust method for routine assessment of plasma triglyceride lipolysis defects. PLoS One. 2014;9:e96482. doi: 10.1371/journal.pone.0096482. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Ueda M, Dunbar RL, Wolska A, et al. A Novel APOC2 Missense Mutation Causing Apolipoprotein C-II Deficiency with Severe Triglyceridemia and Pancreatitis. J Clin Endocrinol Metab. 2017 doi: 10.1210/jc.2016-3903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Nikkilä EA, Taskinen MR, Rehunen S, et al. Lipoprotein lipase activity in adipose tissue and skeletal muscle of runners: relation to serum lipoproteins. Metabolism. 1978;27:1661–1667. doi: 10.1016/0026-0495(78)90288-3. [DOI] [PubMed] [Google Scholar]
- 113.Johansen CT, Dube JB, Loyzer MN, et al. LipidSeq: a next-generation clinical resequencing panel for monogenic dyslipidemias. J Lipid Res. 2014;55:765–772. doi: 10.1194/jlr.D045963. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Kinnunen PK, Jackson RL, Smith LC, et al. Activation of lipoprotein lipase by native and synthetic fragments of human plasma apolipoprotein C-II. Proc Natl Acad Sci U S A. 1977;74:4848–4851. doi: 10.1073/pnas.74.11.4848. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Gotoda T, Shirai K, Ohta T, et al. Diagnosis and management of type I and type V hyperlipoproteinemia. J Atheroscler Thromb. 2012;19:1–12. doi: 10.5551/jat.10702. [DOI] [PubMed] [Google Scholar]
- 116.Bays HE, Jones PH, Orringer CE, et al. National Lipid Association Annual Summary of Clinical Lipidology 2016. Journal of Clinical Lipidology. 10:S1–S43. doi: 10.1016/j.jacl.2015.08.002. [DOI] [PubMed] [Google Scholar]
- 117.Carlson LA, Rosenhamer G. Reduction of mortality in the Stockholm Ischaemic Heart Disease Secondary Prevention Study by combined treatment with clofibrate and nicotinic acid. Acta Med Scand. 1988;223:405–418. doi: 10.1111/j.0954-6820.1988.tb15891.x. [DOI] [PubMed] [Google Scholar]
- 118.Park Y, Harris WS. Omega-3 fatty acid supplementation accelerates chylomicron triglyceride clearance. J Lipid Res. 2003;44:455–463. doi: 10.1194/jlr.M200282-JLR200. [DOI] [PubMed] [Google Scholar]
- 119.Nordøy A, Barstad L, Connor WE, et al. Absorption of the n-3 eicosapentaenoic and docosahexaenoic acids as ethyl esters and triglycerides by humans. Am J Clin Nutr. 1991;53:1185–1190. doi: 10.1093/ajcn/53.5.1185. [DOI] [PubMed] [Google Scholar]
- 120.Gaudet D, Alexander VJ, Baker BF, et al. Antisense Inhibition of Apolipoprotein C-III in Patients with Hypertriglyceridemia. N Engl J Med. 2015;373:438–447. doi: 10.1056/NEJMoa1400283. [DOI] [PubMed] [Google Scholar]
- 121.Gaudet D, Brisson D, Tremblay K, et al. Targeting APOC3 in the familial chylomicronemia syndrome. N Engl J Med. 2014;371:2200–2206. doi: 10.1056/NEJMoa1400284. [DOI] [PubMed] [Google Scholar]
- 122.Dewey FE, Gusarova V, Dunbar RL, et al. Genetic and Pharmacologic Inactivation of ANGPTL3 and Cardiovascular Disease. N Engl J Med. 2017 doi: 10.1056/NEJMoa1612790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Graham MJ, Lee RG, Brandt TA, et al. Cardiovascular and Metabolic Effects of ANGPTL3 Antisense Oligonucleotides. N Engl J Med. 2017 doi: 10.1056/NEJMoa1701329. [DOI] [PubMed] [Google Scholar]
- 124.Millar JS, Lassman ME, Thomas T, et al. Effects of CETP inhibition with anacetrapib on metabolism of VLDL-TG and plasma apolipoproteins C-II, C-III, and E. J Lipid Res. 2017;58:1214–1220. doi: 10.1194/jlr.M074880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Joglekar K, Brannick B, Kadaria D, et al. Therapeutic plasmapheresis for hypertriglyceridemia-associated acute pancreatitis: case series and review of the literature. Therapeutic advances in endocrinology and metabolism. 2017;8:59–65. doi: 10.1177/2042018817695449. [DOI] [PMC free article] [PubMed] [Google Scholar]
