Lipoprotein Disorders and Cardiovascular Disease


Additional content is available online at Elsevier eBooks for Practicing Clinicians

Atherosclerotic cardiovascular diseases (ASCVDs) represent a major burden to society and on national health care systems. Despite public health measures aimed at decreasing saturated fat consumption and cigarette smoking and targeted pharmacologic therapies that can modify cardiovascular risk, an aging population with increased comorbidities, such as obesity, diabetes, and high blood pressure, continues to pose a considerable challenge (see also Chapter 2, Chapter 25 ).

Lipoprotein disorders, especially those that increase exposure of the arterial wall to cholesterol, constitute a major modifiable cardiovascular risk factor. Modulation of plasma cholesterol levels by lifestyle or, when required, pharmacologic therapy (statins), has proven to be one of the most effective interventions for the prevention and treatment of ASCVD.

The lipid transport system in animals has evolved to serve two main purposes: first, to bring ingested fats (mostly in the form of triglycerides) to muscle; when dietary fat is unavailable, the liver and adipose tissue deliver a constant supply of triglycerides. The second transport system is the delivery of cholesterol to tissues that require it for membrane synthesis, for the synthesis of steroidal hormones and bile acids. The remarkable redundancy in the ability of cells and tissues to synthetize, import, and export cholesterol attests to its critical importance in life processes.

Lipids constitute approximately 70% (by mass) of the dry weight of plasma. Approximately half of circulating lipids are sterols, the other major components include glycerophospholipids (phospholipids) and glycerolipids (triglycerides), which circulate in lipoproteins. Thus, circulating lipoproteins continuously bathe vascular endothelial cells, and the interaction between lipoproteins and cells of the arterial wall contribute causally to the pathogenesis of human atherosclerosis (see also Chapter 24 ).

The “cholesterol hypothesis” states that “decreasing blood cholesterol significantly reduces coronary heart disease.” Decades of research in basic research, epidemiology, animal experiments, and mendelian randomization studies have shown strong support for a causal role of cholesterol in the pathogenesis of atherosclerosis. Observational data show a strong and consistent association across populations between elevated blood cholesterol and low-density lipoprotein cholesterol (LDL-C) with ASCVD, especially coronary artery disease (CAD). Experimental data in animals show that the development of atherosclerosis requires cholesterol. Human genetic studies provide strong support of causality for genes related to LDL-C levels, to atherogenic lipoprotein particles, and to cholesterol associated with triglyceride-rich lipoproteins (TRLs) (see also Chapter 7 ). Most important, however, a large body of clinical trials have shown that reducing LDL-C with statins prevents ASCVD, cardiovascular deaths, and total mortality. Thus, LDL meets the modified Koch postulates as a causal risk factor for ASCVD.

The terms dyslipidemia or dyslipoproteinemia reflect disorders of the lipid and lipoprotein transport pathways associated with arterial disease more appropriately than hyperlipidemia . Dyslipidemia encompasses patterns often encountered in clinical practice, such as low high-density lipoprotein cholesterol (HDL-C) and elevated triglyceride concentrations but average total plasma cholesterol or LDL-C levels. Dyslipidemia also includes elevated lipoprotein(a) (Lp[a]) and uncommon genetic or acquired disorders of lipoprotein metabolism. Certain rare lipoprotein disorders can cause overt clinical manifestations, but most common dyslipoproteinemias themselves seldom cause symptoms or clinical signs. Rather, they require laboratory tests for detection. Proper recognition and management of dyslipoproteinemias can reduce cardiovascular and total mortality rates. The fundamentals of lipidology presented here have importance for the daily practice of cardiovascular medicine.

Lipoprotein Transport System

Biochemistry of Lipids

Life requires fats. The biochemistry of lipids and lipoproteins is complex. The clinically relevant lipoproteins ( Table 27.1 ), apolipoproteins ( Table 27.2 ), receptors and processing enzymes ( Table 27.3 ), and current (and potential) drug targets are shown. Biologic lipids usually refer to a broad grouping of naturally occurring molecules that include fatty acids, waxes, eicosanoids, monoglycerides, diglycerides, triglycerides, phospholipids, sphingolipids, sterols, terpenes, prenols, and fat-soluble vitamins (A, D, E, and K), in contrast to the other major groupings of biologic molecules, namely, nucleic acids, proteins, amino acids, and carbohydrates. The major biologic functions of lipids include critical contributions to biologic membranes, energy storage, and the backbones or modifiers of many signaling molecules. Certain lipids, especially fatty acids, readily undergo oxidation and can generate substances highly toxic to cells. Fatty acids can be metabolized in the mitochondrion by beta-oxidation, whereas the sterol nucleus resists enzymatic degradation. Elimination of cholesterol therefore requires excretion as bile acids or conversion into steroidal hormones.

TABLE 27.1
Plasma Lipoprotein Composition and Apolipoproteins
Origin Density (g/mL) Size (nm) % Protein [Cholesterol] In Plasma mg/dL (mmol/L) [Triglyceride] in Fasting Plasma (mmol/L) Major Apo Other Apo
Chylomicrons Intestine <0.95 100–1000 1–2 0.0 0 B48 A-I, Cs
Chylomicron remnants Chylomicron metabolism 0.95–1.006 30–80 3–5 0.0 0.0 B48, E A-I, A-IV, Cs
VLDL Liver <1.006 40–50 10 4–15 mg% (0.1–0.4) 15–100 mg% (0.2–1.2) B100 A-I, Cs
IDL VLDL 1.006–1.019 25–30 18 4–12 mg% 0.1–0.3 10–25 mg% (0.1–0.3) B100, E
LDL IDL 1.019–1.063 20–25 25 50–130 mg% (1.5–3.5) 15–35 mg% (0.2–0.4) B100
HDL Liver, intestine 1.063–1.210 6–10 40–55 35–62 mg% (0.9–1.6) 10–15 mg% (0.1–0.2) A–I, A–II A-IV
Lp(a) Liver 1.051–1.082 25 30–50 B100, (a)
Indirect effect.
Apo, Apolipoprotein; HDL, high-density lipoprotein; IDL, intermediate-density lipoprotein; LDL, Low-density lipoprotein; Lp(a), lipoprotein(a); VLDL, very-low-density lipoprotein.

In mmol/L; for mg/dL, multiply by 38.67.

In mmol/L; for mg/dL, multiply by 88.5.

In the fasted state, serum (or plasma) should not contain chylomicrons or their remnants.

TABLE 27.2
Apolipoproteins (Clinically Relevant)
Name Predominant Lipoprotein MWt (kDa) Role Plasma Concentration (mg/dL) Human Disease Drug Target
Apo (a) Lp(a) 250-800 Unknown 0.2-200 Lp(a) excess AKCEA-APO(a)-LRx
Apo A-I HDL 28.3 ACAT activation, structural 90-160 HDL deficiency
Apo A-IV HDL 45 Structural, absorption 10-20
Apo A-V VLDL, HDL TRL metabolism Hypertriglyceridemia
Apo B100 LDL, VLDL 512 Structural, LDL-R binding 50-150 Hypobetalipoproteinemia Mipomersen
Apo B48 Chylomicrons 241 Structural 0-100 abetalipoproteinemia
Apo C-II Chylomicrons, VLDL 8.84 LPL activation 3-5 Hyperchylomicronemia
Apo C-III Chylomicrons, VLDL 8.76 LPL inhibition 10-14 Hypertriglyceridemia Volanesorsen
Apo E Chylomicrons remnant, IDL 34 LDL-R, apo E receptor binding 2-8 Type III hyperlipoproteinemia
Apo H Chylomicrons, VLDL, LDL, HDL 38-50 Beta 2 -glycoprotein
Platelet aggregation
1.4 -1.6 Cardiolipin-binding defect
See Tables 27.1 and 27.3 for abbreviations. TRL, Triglyceride-rich lipoprotein

TABLE 27.3
Lipoprotein Processing Enzymes, Receptors, Modulating Proteins
Abbreviation Name Role Gene Human Disease Drug Target
ABCA1 ATP-binding cassette A1 Cellular phospholipid efflux ABCA1 Tangier disease
ABCG5/G8 ATP-binding cassette G5 and G8 Intestinal sitosterol transporter ABCG5 ABCG8 Sitosterolemia
ACLY ATP citrate lyase Cholesterol and fatty acid synthesis ACLY Bempedoic acid
ANGPTL3 Angiopoietin-like protein 3 Inhibit LDL and EL ANGPTL3 Familial hypolipoproteinemia 2 Evinacumab, vupanorsen
CETP Cholesteryl ester transfer protein Lipid exchange in plasma CETP Elevated HDL-C
Cyp27A1 Cytochrome Sterols hydroxylation CYP27A1 Cerebrotendinous xanthomatosis
DGAT1 Acyl CoA:Diacylglycerol acyltransferase 1 Triglyceride synthesis DGAT1 Elevated triglycerides Pradigastat
EL Endothelial lipase Phospholipid hydrolysis LIPG
HL Hepatic lipase Triglyceride hydrolysis LIPC Remnant accumulation
HMGCR HMG CoA Reductase Cholesterol synthesis HMGCR Statins
LCAT Lecithin-cholesterol acyltransferase Cholesterol esterification (plasma) LCAT LCAT deficiency, low HDL
LDL-R Low-density lipoprotein receptor LDL uptake LDLR Familial hypercholesterolemia PCSK9 inh. (statins)(bempedoic acid)
AAV8.TBG.hLDLR
LDL-R AP1 LDL-R adapter protein LDL uptake LDLRAP1 Recessive FH
LAL Lysosomal Acid Lipase Cholesteryl ester storage LIPA Wollman disease, CESD
MTTP Microsomal triglyceride transfer protein Apo B assembly MTTP Abetalipoproteinemia Lomitapide
NPC1 Niemann-Pick C gene product Cellular cholesterol transport NPC1 Niemann-Pick type C
NPC1L1 Niemann-Pick C1-like 1 protein Intestinal cholesterol absorption NPC1L1 Ezetimibe
PCSK9 Proprotein convertase, subtilisin/kexin-9 Protein cleavage PCSK9 Hypercholesterolemia Alirocumab, Evolocumab, Inclisiran
SMPD1 Sphingomyelinase phosphodiesterase Sphingomyelin hydrolysis SMPD1 Niemann-Pick types A and B

Lipids generally do not dissolve in water. The lipid transport system has evolved in animals to carry hydrophobic molecules (fat) from sites of origin (the intestines and the liver) to sites of use (muscles and rapidly dividing tissues) through the aqueous (water) environment of plasma. Proteins highly conserved through evolution, termed apolipoproteins (apo), mediate this process. Most apolipoproteins derive from an ancestral gene and contain both hydrophilic and hydrophobic domains. This amphipathic structure enables these proteins to bridge the interface between the aqueous environment of plasma and the phospholipid constituents of lipoprotein. The major types of lipids that circulate in plasma include cholesterol and cholesteryl esters, glycerophospholipids, sphingolipids, and glycerolipids (triglycerides) ( Fig. 27.1 ). The LIPID maps (Lipid Metabolites and Pathways Strategy) consortium has provided standardized nomenclature for lipids, although this area is rapidly evolving with advanced lipidomics technologies.

FIGURE 27.1, Biochemical structure of the major lipid molecules: cholesterol, cholesteryl esters, glycerolipids (triglycerides), and glycerophospholipids (e.g., phosphatidylcholine) and sphingomyelin. Eicosapentaenoic acid (EPA) is an essential polyunsaturated fatty acid. R indicates a fatty acyl chain.

The membranes of mammalian cells and their subcellular organelles require cholesterol . This lipid gives rise to steroid hormones and bile acids and contributes to the integrity of the epidermis. Many cell functions depend critically on membrane cholesterol, and cells regulate tightly their cholesterol content. Importantly, all mammalian cells have retained the ability to synthetize cholesterol de novo from acetyl coenzyme A (CoA). Most of the cholesterol in plasma circulates as cholesteryl esters in the core of lipoprotein particles. The enzyme lecithin-cholesterol acyltransferase (LCAT) forms cholesteryl esters in the blood compartment by transferring a fatty acyl chain from phosphatidylcholine to cholesterol.

Glycerolipids ( triglycerides ) consist of a three-carbon glycerol backbone covalently linked to three fatty acid chains (R 1–3 ). The fatty acid composition varies in terms of chain length and the presence of double bonds (degree of saturation). The highly hydrophobic triglyceride molecules circulate in the core of the lipoprotein. Hydrolysis of triglycerides by lipases generates the free fatty acids (FFAs) used for energy.

Glycerophospholipids , found in all cellular membranes, consist of a glycerol molecule linked to two fatty acids (designated R; see Fig. 27.1 ). Fatty acids differ in length and in the number of double bonds. The third carbon of the glycerol backbone carries a phosphate group linked to one of four molecules: choline (phosphatidylcholine, also called lecithin), ethanolamine (phosphatidylethanolamine), serine (phosphatidylserine), or inositol (phosphatidylinositol). More complex phospholipids include phosphatidylglycerol (cardiolipin is formed by the fusion of two phosphatidylglycerol molecules -antibodies against cardiolipin often occur in systemic lupus), and plasmalogens, an important constituent of eukaryotic membranes. Another phospholipid, sphingomyelin , has special functions in the plasma membrane in the formation of membrane microdomains such as rafts and caveolae. The structure of sphingomyelin resembles that of phosphatidylcholine. The backbone of sphingolipids uses the amino acid serine rather than glycerol. Phospholipids are polar molecules, more water soluble than triglycerides or cholesterol or its esters. Phospholipids participate in signal transduction pathways: hydrolysis by membrane-associated phospholipases generates second messengers, including diacylglycerols, lysophospholipids, phosphatidic acids, and FFAs such as arachidonate, that regulate many cell functions. The phosphorylation of phosphatidylinositol contributes critically to membrane and cell organelle signaling and transport.

Fish oils and eicosapentaenoic acid (EPA) n-3 fatty acids are essential polyunsaturated fatty acids (PUFAs) that are commonly found in plants and marine life. Beyond their triglyceride lowering effects, they have pleiotropic actions with beneficial cardiovascular effects that are key to resolving inflammation and regulating adaptive immunity. n-3 fatty acids derive their name from the first double, which is located on the n-3 position (between the third and fourth carbons from the terminal methyl [omega] end). n-3 metabolites serve as precursors of potent lipid mediators (eicosanoids), and have effects on other lipids (ceramides, lysophosphatidylcholines, diacylglycerols) and amino acids (leucine, glutamine). Moreover, n-3 derived eicosanoids form downstream potent oxylipins which play a critical role in regulating inflammation, vascular tone, and endothelial function. n-3 fatty acids may have favorable antiarrhythmic effects by stabilizing the plasma membrane, metabolic effects through insulin signaling and energy use, and at high doses may also block coagulation. These pleiotropic effects have given rise to the hypothesis that dietary n-3 treatment improves cardiovascular clinical outcomes. However, results of several large randomized clinical trials have been conflicting (see discussion under Fish Oils and Pure Eicosapentaenoic Acid”).

PUFAs are metabolized by three pathway enzymes: cyclooxygenase (COX), lipoxygenase (LOX), and CypP450. n-3 FA include EPA (20:5, i.e., 20 carbons in length with 5 double bonds) and docosahexaenoic acid (DHA, 22:6). Alpha-linolenic acid (ALA) is metabolized into EPA, and EPA is modified into DHA. EPA and DHA are oxidized by COX and LOX enzymes to generate eicosanoid subfamilies including prostaglandins, prostacyclins, resolvins, and thromboxanes. Certain n-3 derived eicosanoids have potent antiinflammatory and pro-resolving effects, independently, as well as by inhibiting proinflammatory n-6 PUFA-derived eicosanoids. EPA and DHA-derived eicosanoids compete for and inhibit arachidonic acid eicosanoid production and receptor stimulation (particularly in the COX pathway). EPA can also be metabolized into resolvin E1 which may be cardioprotective, or it can be elongated into the 22-carbon docosapentaenoic acid (DPA) which has resolvin-like antiinflammatory effects and can shift arachidonic acid metabolism from COX to LOX pathway. DHA is thought to be cardioprotective after it is incorporated in cell membranes via mechanisms that improve receptor and ion channel function. Genetic variants in fatty acid desaturases ( FADS ) – enzymes that modulate eicosanoid and other fatty acid levels – and in other genes encoding enzymes involved in eicosanoid metabolism (COX, LOX, and CypP450), also associate with CVD risk. Deciphering the specific roles of n3-derived lipid metabolites in inflammatory and immune signaling would provide insight into their role as therapeutic and nutritional agents.

Lipoproteins, Apolipoproteins, Receptors, and Processing Enzymes

Lipoproteins are complex macromolecular structures coated by a water compatible envelope of phospholipids, free cholesterol, and apolipoproteins covering a hydrophobic core of cholesteryl esters and triglycerides ( Fig. 27.2 ). Lipoproteins vary in size, density in the aqueous environment of plasma, and lipid and apolipoprotein content (see Table 27.1 , Fig. 27.1B , inset ). The classification of lipoproteins reflects their density in plasma (the density of plasma is 1.006 g/mL) as gauged by flotation in an ultracentrifuge. The TRLs, which consist of chylomicrons , chylomicron remnants , very low-density lipoprotein (VLDL) and intermediate density lipoproteins (IDL), have a density of less than 1.006 g/mL. The rest (bottom fraction) of the ultracentrifuged plasma consists of LDL , HDL , and Lp(a) .

FIGURE 27.2, Relative size of plasma lipoproteins according to their hydrated density. The density of plasma is 1.006 g/mL. Inset , Structure of lipoproteins. Phospholipids are oriented with their polar group toward the aqueous environment of plasma. Free cholesterol is inserted within the phospholipid layer. The core of the lipoprotein is composed of cholesteryl esters and triglycerides. Apolipoproteins are involved in the secretion of lipoprotein, provide structural integrity, and act as cofactors for enzymes or as ligands for various receptors. HDL , High-density lipoprotein; IDL , intermediate-density lipoproteins; LDL , low-density lipoprotein; Lp(a) , lipoprotein(a); VLDL, very-low-density lipoprotein.

Apolipoproteins have four major roles: (1) assembly and secretion of the lipoprotein (apoA-I, B100, and B48), (2) structural integrity of the lipoprotein (apo B, E, A-I, and A-II), (3) coactivators or inhibitors of enzymes (apoA-I, A-V, C-I, C-II, and C-III), and (4) binding or docking to specific receptors and proteins for cellular uptake of the entire particle or selective uptake of a lipid component (apoA-I, B100, and E) (see Table 27.2 ). The role of several apolipoproteins (A-IV, A-V, D, H, J, L, and M) remains incompletely understood.

Many proteins regulate the synthesis, secretion, and metabolic fate of lipoproteins; their characterization has provided insight into molecular cellular physiology and targets for drug development (see Table 27.3 ). Discovery of the LDL receptor (LDL-R) represented a landmark in understanding cholesterol metabolism and receptor-mediated endocytosis. The LDL-R regulates the entry of cholesterol into cells, and tight control mechanisms alter its expression on the cell surface, depending on intracellular cholesterol. The LDL-R belongs to a superfamily of membrane receptors that include LDL-R, VLDL-R, LDL-R–mediated peptide type 1 (LRP1; apo E receptor), LRP1B, LRP4 (MGEF7), LRP5 and LRP6 (involved in the process of bone formation), LRP8 (apo E receptor-2), and LRP9. LRP1, which mediates the uptake of chylomicron remnants and VLDL, preferentially recognizes apo E. LRP1 also interacts with hepatic lipase. The complex interaction between hepatocytes and the various lipoproteins containing apo E involves cell surface proteoglycans that provide scaffolding for lipolytic enzymes (lipoprotein lipase [LPL] and hepatic lipase) involved in recognition of remnant lipoproteins. Macrophages express receptors that bind modified (especially oxidized) lipoproteins. These scavenger lipoprotein receptors mediate the uptake of oxidatively modified LDL into macrophages. In contrast to the exquisitely regulated LDL-R, high cellular cholesterol content does not suppress scavenger receptors, thereby enabling intimal macrophages to accumulate abundant cholesterol, become foam cells, and form fatty streaks. Sterol accumulation in the endoplasmic reticulum may lead to cell apoptosis via the unfolded protein response. Endothelial cells can also take up modified lipoproteins through specific receptors such as the oxidized LDL-R, LOX-1.

At least three physiologically relevant receptors interact with HDL particles: the scavenger receptor class B (SR-B1) and the adenosine triphosphate (ATP)-binding cassette transporters A1 (ABCA1) and G1 (ABCG1). SR-B1 is a receptor for HDL (also for LDL and VLDL, but with less affinity). SR-B1 mediates the selective uptake of HDL cholesteryl esters in steroidogenic tissues, hepatocytes, and endothelium. ABCA1 mediates cellular phospholipid (and possibly cholesterol) efflux and HDL formation. The ABCG1 transporter transfers cellular cholesterol to already formed HDL particles.

Lipoprotein Metabolism and Transport

The lipoprotein transport system has two major roles: efficient transport of triglycerides from the intestine and liver to sites of utilization (fat tissue or muscle), and transport of cholesterol to peripheral tissues for membrane synthesis and steroid hormone production or to the liver for bile acid synthesis.

Intestinal Pathway (Chylomicrons To Chylomicron Remnants)

Fat typically furnishes 20% to 40% of the daily calories. Triglycerides account for the major portion of ingested fats. For an individual consuming 2000 kcal/day, with 30% in the form of fat, this represents approximately 66 g of triglycerides and 250 mg (0.250 g) of cholesterol per day. The intestine has very efficient fat absorption mechanisms, probably evolved to maximize provision of the organism with nutrients under circumstances of limited or irregular availability of food.

On ingestion, lingual and pancreatic lipases hydrolyze triglycerides into FFAs and monoglycerides or diglycerides. Emulsification by bile salts leads to the formation of intestinal micelles. Micelles resemble lipoproteins insofar as they consist of phospholipids, free cholesterol, bile acids, diglycerides and monoglycerides, FFAs, and glycerol. The mechanism of micelle uptake by intestinal brush border cells still engenders debate. The Niemann-Pick C1-like 1 (NPC1L1) protein is part of an intestinal cholesterol transporter complex and the target for the selective cholesterol absorption inhibitor ezetimibe. After uptake into intestinal cells, fatty acids undergo re-esterification to form triglycerides and packaging into chylomicrons inside the intestinal cell and enter the portal circulation ( Fig. 27.3 , part 1 ). Chylomicrons contain apo B48, the amino-terminal component of apo B100. In the intestine, the apo B gene is modified during transcription into mRNA by substitution of a uracil for a cytosine via an apo B48–editing enzyme complex (ApoBec). This mechanism involves a cytosine deaminase and leads to a termination codon at residue 2153 and a truncated form of apo B. Only intestinal cells express ApoBec. Apo B48 does not bind to LDL-R. Intestinal cells absorb plant sterols (sitosterol, campesterol), sort these compounds into a separate cellular compartment, and re-secrete them into the intestinal lumen via the ABCG5/8 heterodimeric transporter. Mutations of the ABCG5/8 genes cause the rare disorder sitosterolemia.

FIGURE 27.3, Schematic diagram of the lipid transport system. Numbers in circles refer to explanations in text. Refer to Tables 27.1–27.3 for abbreviations. CM, Chylomicron; FFA, free fatty acid.

Chylomicrons rapidly enter the plasma compartment after meals. In capillaries of adipose tissue or muscle cells in the peripheral circulation, chylomicrons encounter the enzyme LPL attached to heparan sulfate proteoglycans on the luminal surface of endothelial cells (see Fig. 27.3 , part 2 ). Apo C-II and apo A-V activate, and apo C-III inhibits LPL activity. LPL has broad specificity for triglycerides; it cleaves all fatty acyl residues attached to glycerol and in the process generates three molecules of FFA for each molecule of glycerol. Muscle cells rapidly take up fatty acids. Fatty acids provide the energy substrate for muscle contraction by the generation of ATP during beta-oxidation of fatty acyl residues in mitochondria. Adipose cells can store triglycerides made from fatty acids for energy utilization, a process that requires insulin. The triglyceride lipase hormone–sensitive lipase, that is activated by cyclic adenosine monophosphate (cAMP) in response to stress, releases stored fatty acids from adipose tissues. Fatty acids can also travel to the liver bound to fatty acid–binding proteins or albumin and undergo repackaging into VLDL. Peripheral resistance to insulin can thus increase the delivery of FFAs to the liver with a consequent increase in VLDL secretion and increased apo B particles in plasma, a characteristic of the “metabolic syndrome” and type 2 diabetes. The remnant particles, derived from chylomicrons following LPL action, contain apo E and enter the liver for degradation and reutilization of their core constituents (see Fig. 27.3 , part 3).

Hepatic Pathway (Very Low-Density Lipoprotein to Intermediate-Density Lipoprotein)

Food is not always available, and dietary fat content varies. The body requires ready availability of triglyceride to meet energy demands. Hepatic secretion of VLDL particles serves this function (see Fig. 27.3 , part 4 ). VLDLs are TRLs smaller than chylomicrons (see Table 27.1 and Fig. 27.2 ). They contain apo B100 as their main lipoprotein. As opposed to apo B48, apo B100 contains a domain recognized by LDL-R (the apo B/E receptor). VLDL particles follow the same catabolic pathway as chylomicrons through LPL (see Fig. 27.3 , part 2 ). During hydrolysis of TRLs by LPL, an exchange of proteins and lipids takes place: VLDL particles (and chylomicrons) acquire apo Cs and apo E, in part from HDL particles. VLDLs also exchange triglycerides for cholesteryl esters from HDL (mediated by cholesteryl ester transfer protein [CETP]) (see Fig. 27.3 , part 9). Such bidirectional transfer of constituents between lipoproteins serves several purposes: acquisition of specific apolipoproteins by lipoproteins that will dictate their metabolic fate, transfer of phospholipids onto nascent HDL particles mediated by phospholipid transfer protein (PLTP) (during loss of the core triglycerides, the phospholipid envelope becomes redundant and sheds apoA-I to form new HDL particles), and transfer of cholesterol from HDL to VLDL remnants so that it can be metabolized in the liver. This exchange constitutes a major part of the “reverse cholesterol transport pathway.”

Apo CIII, a small but important 79 amino acid peptide has a high affinity for TRLs and attenuates the activity of LPL and the clearance of TRLs, thus contributing to elevated triglycerides. Apo CIII also resides within HDL that seems to act as a “reservoir” for this apolipoprotein. Recent work identified an intracellular role for apo CIII for the assembly and secretion of VLDL. Mendelian randomization experiments as well as epidemiologic studies have established that apo CIII can contribute causally to ASCVD. This recognition has spurred therapeutic efforts to decrease apo CIII (see below).

After hydrolysis of triglycerides removes some triglycerides from VLDL, these particles have relatively more cholesterol, shed several apolipoproteins (especially the C apolipoproteins), and acquire apo E. The VLDL remnant lipoprotein, called intermediate-density lipoprotein (IDL), undergoes liver uptake via its apo E moiety (see Fig. 27.3 , part 3 ) or further delipidation by hepatic lipase to form LDL particles (see Fig. 27.3 , part 6 ). At least four receptors take up TRLs, TRL remnants, and apo B–containing lipoproteins: VLDL-R, the remnant receptor (apo ER2), LDL-R (also called the apo B/E receptor), and LRP1. Most hepatic receptors share the ability to recognize apo E, an engagement that mediates the uptake of several classes of lipoproteins, including VLDL and IDL. The complex interaction between apo E and its ligand involves the “docking” of TRLs on heparan sulfate proteoglycans to present the ligand to its receptor.

Low-Density Lipoproteins

LDL particles contain predominantly cholesteryl esters packaged with apo B100. Normally, triglycerides constitute only 4% to 8% of the LDL mass (see Table 27.1 ). In conditions with elevated plasma triglyceride concentrations, LDL particles can acquire triglycerides and deplete their core cholesteryl esters. Such changes in core constituents influence LDL particle size: an increase in triglycerides and a relative decrease in cholesteryl esters yields smaller, denser LDL particles.

Humans are unusual among mammals because they generate LDL as a cholesterol-rich lipoprotein. Nonhuman primates fed a cholesterol-enriched diet also carry cholesterol in LDL. In other mammals, such as rodents or rabbits, HDL particles transport most of the cholesterol. Cells can either make cholesterol from acyl CoA through enzymatic reactions requiring at least 33 enzymatic steps or obtain it as cholesteryl esters from HDL or LDL particles. Cells internalize LDL via LDL-R ( Fig. 27.4 ). LDL particles contain one molecule of apo B. Although several highly lipophilic domains of apo B associate with phospholipids, a region surrounding residue 3500 binds with high affinity to LDL-R. LDL-R localizes in a region of the plasma membrane rich in the protein clathrin (see Fig. 27.4 ; Fig. 27.3 , part 7 ). Once bound to the receptor, clathrin polymerizes and forms an endosome that contains LDL bound to its receptor, a portion of the plasma membrane, and clathrin. This internalized particle then fuses with lysosomes whose hydrolytic enzymes (cholesteryl ester hydrolase, cathepsins) release free cholesterol and degrade apo B. LDL-R releases its ligand and can recycle to the plasma membrane. The chaperone proprotein convertase subtilisin/kexin type 9 (PCSK9), secreted by hepatocytes, undergoes auto-catalytic cleavage and binds to the LDL-R. Association with PCSK9 diverts the complex to the lysosomal degradative pathway, thus preventing the recycling of the LDL-R ( Fig. 27.4 ). Gain-of-function mutations in the PCSK9 gene causes autosomal dominant hypercholesterolemia, whereas loss-of-function mutations increase LDL-R and lower LDL-C substantially. ,

FIGURE 27.4, Diagram of a hepatocyte expressing the low-density lipoprotein receptor (LDL-R). Top panel, In the absence (or mAb blockade) of PCSK9, the LDL-R recycles rapidly to the cell surface. LDL particles are cleared by LDL by receptor-mediated endocytosis, thereby lowering LDL-C concentration in the blood. Bottom panel, PCSK9 chaperones the internalized LDL-R/LDL particle complex to the endosome-lysosomal compartment, where it undergoes degradation. The consequent decrease in LDL-R impairs LDL clearance, yielding accumulation of cholesterol-rich LDL particles in the blood. ER, Endoplasmic reticulum; TGN, trans Golgi network.

Cells regulate their cholesterol content tightly through highly conserved cellular pathways, including: (1) synthesis of cholesterol in the smooth endoplasmic reticulum (via the rate-limiting step hydroxymethylglutaryl-CoA [HMG-CoA] reductase), (2) receptor-mediated endocytosis of LDL (two mechanisms under the control of steroid-responsive element binding protein-2 [SREBP-2]), (3) efflux of cholesterol from the plasma membrane to cholesterol acceptor particles (predominantly apoA-I and HDL) via the ABCA1and ABCG1 transporters, and (4) intracellular cholesterol esterification via the enzyme acetyl-CoA acetyltransferase (ACAT). SREBP-2 coordinately regulates the first two pathways at the level of gene transcription. Cellular cholesterol binds to SCAP (SREPB cholesterol-activated protein), which localizes on the endoplasmic reticulum. Cholesterol inhibits the interaction of SCAP with SREPB. In the absence of cholesterol, SCAP mediates the cleavage of SREBP at two sites by specific proteases with the release of an amino-terminal fragment of SREBP. This SREBP fragment migrates to the nucleus and increases the transcriptional activity of genes involved in cellular cholesterol and fatty acid homeostasis. The ACAT pathway regulates the cholesterol content in membranes. Humans express two separate forms of ACAT. ACAT1 and ACAT2 derive from different genes and mediate cholesterol esterification in cytoplasm and in the endoplasmic reticulum lumen for lipoprotein assembly and secretion.

High-Density Lipoprotein and Reverse Cholesterol Transport

Regulation of cholesterol efflux from cells depends in part on the ABCA1 pathway, controlled in turn by hydroxysterols (especially 24- and 27-OH cholesterol, which act as ligands for the liver-specific receptor [LXR] family of nuclear transcription factors). In conditions of cholesterol sufficiency, the cell can decrease cholesterol synthesis. The cell can also limit the amount of cholesterol that enters the cell via the LDL-R, thereby augmenting the amount stored as cholesteryl esters, and can promote cholesterol removal by increasing its movement to the plasma membrane for efflux.

Epidemiologic studies have consistently shown an inverse relationship between plasma levels of HDL-C and the presence of CAD . HDL promotes reverse cholesterol transport and can prevent lipoprotein oxidation, and exert antiinflammatory actions in vitro, among many other seemingly salutary functions. Yet, Mendelian randomization analyses have cast doubt on the causal role of HDL as a protective cardiovascular risk factor. Mutations of the genes for ABCA1 that cause lifelong HDL deficiency do not impart additional cardiovascular risk, and conversely, variants in genes that increase HDL-C do not associate with protection from cardiovascular events.

HDL has a complex and incompletely understood metabolism. The complexity arises because HDL particles acquire their components from several sources and these components undergo metabolism at different sites. In addition, steady-state levels of HDL in plasma may not reflect the dynamic nature of HDL-mediated cholesterol trafficking, in contrast to the situation with LDL. The intestine and liver synthesize apoA-I, the main protein of HDL. Approximately 80% of HDL originates from the liver and 20% from the intestine (see Fig. 27.3 , part 5 ). Lipid-free apoA-I acquires phospholipids from cell membranes and from redundant phospholipids shed during the hydrolysis of TRLs. Lipid-free apoA-I binds to ABCA1 and promotes the transporter’s phosphorylation via cAMP, which increases the net efflux of phospholipids and cholesterol onto apoA-I to form a nascent HDL particle (see Fig. 27.3 , part 10 ). This particle contains apoA-I, phospholipids, and some free cholesterol. These nascent HDL particles will mediate further cellular efflux of cholesterol. Currently, standard laboratory tests do not measure these HDL precursors because they contain little or no cholesterol. On reaching a cell membrane, the nascent HDL particles capture membrane-associated cholesterol and promote the efflux of free cholesterol onto other HDL particles (see Fig. 27.3 , part 10 ). Conceptually, the formation of HDL particles appears to involve two steps. The first step involves binding of HDL apoA-I to ABCA1 and generation of a specific membrane microdomain that allows the subsequent lipidation of apoA-I. Efflux of cellular cholesterol from peripheral cells, such as macrophages, does not contribute importantly to overall HDL-C mass but could export cholesterol from plaques. Macrophages can transfer cholesterol to apoA-I and apo E, to nascent discoid or ellipsoid HDL particles via the ABCA1 transporter. The ABCG1 transporter does not promote cellular cholesterol efflux to lipid-free or lipid-poor apoA-I but to mature HDL particles. In vitro assays can measure HDL-mediated cellular cholesterol efflux by plasma samples, a process that appears altered in many disease states, including diabetes and CAD. LCAT, an enzyme activated by apoA-I, then esterifies the free cholesterol (see Fig. 27.3 , part 8 ). Such assays have proven useful in research but are not scalable or validated for clinical use. HDL also furnishes cholesterol to steroid hormone–producing tissues and the liver through selective uptake of cholesterol mediated by the scavenger receptor SR-B1.

Because of their hydrophobicity, cholesteryl esters move to the core of the lipoprotein, and the HDL particle now assumes a spherical configuration (a particle denoted HDL 3 ). With further cholesterol esterification, the HDL particle increases in size to become the more buoyant HDL 2 . The cholesterol within HDL particles can transfer to TRLs via CETP, which mediates an equimolar exchange of cholesterol from HDL to TRL and movement of triglyceride from TRL onto HDL (see Fig. 27.3 , part 9 ). Inhibition of CETP increases HDL-C in blood and has undergone exploration as a therapeutic target for prevention of ASCVD. However, in clinical trials, CETP inhibitors have failed to improve outcomes except for anacetrapib in the Randomized Evaluation of the Effects of Anacetrapib through Lipid Modification (REVEAL), an effect likely due to LDL-lowering properties of this drug rather than boosting HDL concentration. Triglyceride-enriched HDLs are denoted HDL 2b . Hepatic lipase can hydrolyze triglycerides and endothelial lipase can hydrolyze phospholipids within these particles and thereby convert them back to HDL 3 particles.

Reverse cholesterol transport involves the uptake of cellular cholesterol from extrahepatic sources, such as lipid-laden macrophages, and its esterification by LCAT, transport by large HDL particles, and exchange for one triglyceride molecule by CETP.Hepatic receptors can now take up the cholesterol molecule originally on an HDL particle and residing in a TRL or LDL particle after this exchange. HDL particles therefore act as shuttles between tissue cholesterol, TRL, and the liver. Reverse cholesterol transport by HDL constitutes a small but potentially important portion of the plasma HDL mass. Indeed, selective inactivation of macrophage ABCA1 does not change HDL-C levels in mice but increases atherosclerosis. The protein component of HDL particles is exchangeable with lipoproteins of other classes. The kidneys appear to be a route of elimination of apoA-I and other HDL apolipoproteins.

Lipoprotein Disorders

Definitions

Time and new knowledge have stimulated changes in the classification of lipoprotein disorders. The original classification of lipoprotein disorders by Fredrickson, Lees, and Levy (1967) which depended on analysis of lipoprotein patterns by ultracentrifugation or electrophoresis has fallen into disuse (see prior editions for details). Most clinicians now classify lipoprotein disorders by which specific lipoprotein lipid is elevated and, when sufficiently characterized, by the genetic defect, e.g., familial hypercholesterolemia (FH). For example, a young patient with eruptive xanthomas and a plasma triglyceride level of 22 mmol/L (2000 mg/dL) probably has familial hyperchylomicronemia as a result of LPL deficiency or other monogenic defects. A 38-year-old woman with a strong family history of ASCVD, tendinous xanthomas, and an untreated LDL-C of 240 mg/dL (6.4 mmol/L) likely has heterozygous familial hypercholesterolemia (HeFH). An obese, hypertensive middle-aged man with a cholesterol level of 6.4 mmol/L (245 mg/dL), a triglyceride level of 3.1 mmol/L (274 mg/dL), an HDL-C level of 0.8 mmol/L (31 mg/dL), and a calculated LDL-C level of 4.2 mmol/L (162 mg/dL) probably has metabolic syndrome, and this should trigger the clinician to seek other components of this cluster, including poor lifestyle, hypertension, and hyperglycemia. Conversely, an obese middle-aged man with a plasma triglyceride level of 7 mmol/L (620 mg/dL) probably has mutations in several genes associated with plasma triglyceride levels.

The clinical usefulness of apolipoprotein levels has stirred debate (see Chapter 25 ). Taken as a single measurement, the apo B level provides information on the number of potentially atherogenic particles and can be used as a goal of lipid-lowering therapy. Similarly, LDL particle size correlates highly with plasma HDL-C and triglyceride levels, and most studies do not show it to be an independent cardiovascular risk factor in particular after adjusting for apo B or LDL particle concentration. Small, dense LDL particles tend to track with features of metabolic syndrome, which usually involves dyslipoproteinemia with elevated plasma triglyceride and reduced HDL-C levels. While there is some debate on the value of apo B or non-HDL-C as a better predictive biomarker of ASCVD, the Emerging Risk Factors Collaboration studies and the UK Biobank (comprising 346,686 participants) have shown that measurement of non–HDL-C is equivalent to measurement of apo B in determination of cardiovascular risk. The measurement of apoB remains a useful tool for cardiovascular risk prediction, especially in the primary prevention setting in the presence of features of the metabolic syndrome. Similarly, HDL-C tracks as well with CVD risk as apoA-I does.

Genetic Lipoprotein Disorders

Understanding of the genetics of lipoprotein metabolism has expanded rapidly (see Chapter 7 ). Classification of genetic lipoprotein disorders usually requires a biochemical phenotype in addition to a clinical phenotype. With the exception of FH, monogenic disorders tend to be very rare and constitute “orphan” diseases. Disorders considered heritable on careful family study may be difficult to characterize unambiguously because of age, sex, penetrance, and gene-gene and environmental interactions. Most common lipoprotein disorders encountered clinically result from the interaction of increasing age, lack of physical exercise, weight gain, and a suboptimal diet with individual genetic makeup. Genetic lipoprotein disorders can either raise or lower levels of LDL, Lp(a), remnant lipoproteins, TRLs (chylomicrons and VLDL), or HDL ( Table 27.4 ).

TABLE 27.4
Genetic Lipoprotein Disorders
Disorder Gene Figure 27.3
LDL Particles
  • Autosomal dominant hypercholesterolemia (ADH)

    • Heterozygous Familial Hypercholesterolemia (HeFH)

LDLR 7
    • Homozygous Familial Hypercholesterolemia (HoFH)

LDLR 7
    • Familial defective apo B100

APOB 7
    • Gain-of-function PCSK9 mutations

PCSK9 7
  • Autosomal recessive hypercholesterolemia

LDLRAP1 7
  • Abetalipoproteinemia

MTTP
  • Hypobetalipoproteinemia

APOB
  • Familial sitosterolemia

ABCG5/ABCG8
  • Familial Lp(a) hyperlipoproteinemia

APOA 11
Remnant Lipoproteins
  • Dysbetalipoproteinemia type III

APOE 3
Hepatic lipase deficiency LIPC 6
Triglyceride-Rich Lipoproteins
  • Lipoprotein lipase deficiency (Familial Chylomicronemia Syndrome—FCS)

LPL 2
  • Apo C-II deficiency

APOCII 2
  • Apo A-V deficiency

APOAV
  • Familial hypertriglyceridemia

Polygenic
  • Familial combined hyperlipidemia

Polygenic
High Density Lipoproteins
  • Apo A-I deficiency

APOAI 5
Tangier disease/familial HDL deficiency ABCA1 10
Familial LCAT deficiency syndromes LCAT 8
CETP deficiency CETP 9
Niemann-Pick disease types A and B SMPD1
Niemann-Pick disease type C NPC1
Other
  • Cerebrotendinous xanthomatosis

CYP27A1
CETP, Cholesteryl ester transfer protein; LCAT, lecithin-cholesterol acyltransferase.

Low-Density Lipoproteins

Familial Hypercholesterolemia

Elucidation of the pathway by which complex molecules enter the cell by receptor-mediated endocytosis and discovery of LDL-R represent landmarks in cell biology and clinical investigation. Affected subjects have an elevated LDL-C level greater than the 95th percentile for age and sex, approximately 190 mg/dL (5.0 mmol/L) in adults. In adulthood, clinical manifestations include tendinous xanthomas over the extensor tendons (metacarpophalangeal joints, patellar, triceps, and Achilles tendons); corneal arcus and xanthelasma are less specific signs of FH. These clinical findings are increasingly rare, as biochemical testing enables earlier recognition and treatment. Transmission is autosomal codominant. The diagnosis of FH is usually made according to the Dutch Lipid Clinics Network or the Simon-Broome criteria. A new definition for FH relies on a simpler system combining LDL-C, family history of elevated cholesterol, or premature ASCVD. These definitions are highly concordant and rely on the absolute levels of LDL-C, family history of premature ASCVD, family history of elevated LDL-C, cutaneous manifestations and, if available, DNA analysis. Heterozygous FH (HeFH) affects approximately 1 in 311, , with a higher prevalence in populations with a founder effect. Patients with FH have high risk for the development of CAD by the third to fourth decade in men and approximately 8 to 10 years later in women. The presence of a mutation in a gene known to cause FH increases cardiovascular risk by greater than 10- to 20-fold. Genetic testing for FH is now recommended to make a precise diagnosis and guide therapy, and to inform cascade testing of siblings and offspring. Remarkably, prompt recognition in childhood or early adulthood and treatment (statins) can normalize life expectancy.

Low-Density Lipoprotein Receptor Gene

Defects in the low-density lipoprotein receptor ( LDLR ) gene cause an accumulation of LDL particles in plasma and thus alter the function of the LDL-R protein and cause FH (see Fig. 27.3 , part 7 ). Well in excess of 1700 mutations of the LDLR gene can cause FH. For clinical purposes, LDLR gene mutations are characterized as defective (<20% to 30% residual activity) or null (0% activity). The severity of elevated LDL-C and age of onset of ASCVD correlates with the severity of the mutation.

Familial Defective Apolipoprotein B

Mutations within the apolipoprotein B (APOB) gene that lead to an abnormal ligand-receptor interaction can cause a form of autosomal dominant hypercholesterolemia clinically indistinguishable from FH. Several mutations at the postulated binding site to LDL-R cause familial defective apo B100 (see Fig. 27.3 , part 7 ). The defective apo B has reduced affinity (20% to 30% of control) for LDL-R. LDL particles with defective apo B have a plasma half-life threefold to fourfold greater than the half-life of normal LDL. Because of their increased persistence, these LDL particles can more readily undergo oxidative modifications that can enhance their atherogenicity. Affected subjects usually have LDL-C levels elevated up to 400 mg/dL (10.4 mmol/L) but may also have normal levels. Familial defective apo B100 has a lower prevalence than LDLR mutations (approximately 1 in 500).

Proprotein Convertase Subtilisin/Kexin Type 9

Gain-of-function mutations in the PCSK9 gene decrease surface availability of the LDL-R protein and cause accumulation of LDL-C in plasma (see Fig. 27.4 ). A loss-of-function mutation in PCSK9 confers lower LDL-C than in individuals without the mutation. Black Americans had a higher prevalence of this protective mutation than did whites in the ARIC (Atherosclerosis Risk in Communities) study, and subjects with life-long low LDL-C because of a mutation at the PCSK9 gene locus had a marked reduction in coronary events, thus confirming that genetically low LDL-C states lower cardiovascular risk.

Polygenic Hypercholesterolemia

In most cohorts of “definite” FH patients, as many as 20% do not have a mutation in the LDLR, APOB or PCSK9 genes. While exome-wide sequencing has identified several other genes causing a phenocopy of FH, some patients have an accumulation of single nucleotide polymorphisms of genes known to elevate LDL-C in large-scale genome-wide association studies.

Autosomal Recessive Hypercholesterolemia

An autosomal recessive form of FH identified in a kindred from Sardinia results from a mutation in the gene encoding the LDL-R adaptor protein ( LDL-RAP-1 gene), which encodes a protein involved in recycling of LDL-R. Other genes, including APOE del166LEU , and lysosomal acid lipase (LIPA) cause a phenocopy of FH. Other genes, such as STAP1 (Signal Transducing Adaptor Family Member 1) have been associated with FH, but careful studies performed on genetically modified animals dismissed this association, highlighting some of the limitations of exome-wide genetic association studies.

Hypobetalipoproteinemia and Abetalipoproteinemia

Mutations within the APOB gene can lead to truncations of the mature apo B100 peptide. Many such mutations cause a syndrome characterized by reduced LDL-C and VLDL-C but few if any clinical manifestations and no known risk for ASCVD, a condition referred to as hypobetalipoproteinemia. Apo B truncated close to its amino terminus loses the ability to bind lipids, and produces a syndrome similar to abetalipoproteinemia, a rare recessive lipoprotein disorder of infancy that causes mental retardation and growth abnormalities. Abetalipoproteinemia results from a mutation in the gene coding for the microsomal triglyceride transfer protein ( MTTP ), which is required for assembly of apo B–containing lipoproteins in the liver and the intestine. The resulting lack of apo B–containing lipoproteins in plasma causes a lack of fat-soluble vitamins (A, D, E, and K) that circulate in lipoproteins. In turn, this deficiency result in mental and developmental retardation in affected children.

Sitosterolemia

A rare condition of increased intestinal absorption and decreased excretion of plant sterols (sitosterol and campesterol) can mimic severe FH with extensive xanthoma formation. Premature atherosclerosis, often apparent clinically well before adulthood, occurs in patients with sitosterolemia. Diagnosis requires specialized analysis of plasma sterols documenting an elevation in sitosterol, campesterol, cholestanol, sitostanol, and campestanol. Patients with sitosterolemia have normal or reduced plasma cholesterol levels, and normal triglyceride concentrations. Patients with sitosterolemia have rare homozygous (or compound heterozygous) mutations in the ABCG5 and ABCG8 genes. The gene products of ABCG5 and ABCG8 are half ABC transporters and form a heterodimer localized in the villous border of intestinal cells, that actively pumps plant sterols back into the intestinal lumen. A defect in either of the genes inactivates this transport mechanism, and net accumulation of plant sterols (because of impaired elimination) ensues.

Lipoprotein(a) ( Fig. 27.3 )

Lp(a) (pronounced “lipoprotein little a”) consists of an LDL particle linked covalently with one molecule of apo (a). The apo (a) moiety consists of a protein with a high degree of homology with plasminogen. The gene for apo (a) appears to have arisen from the plasminogen gene. The apo (a) gene has multiple repeats of one of the kringle motifs (kringle IV), which vary in number from 12 to more than 40 in each individual. Plasma Lp(a) levels depend almost entirely on genetics and correlate inversely with the number of kringle repeats and therefore with the molecular weight of apo (a). Human genetic data implicate Lp(a) as a causal cardiovascular risk factor. Lp(a) concentrations follow a skewed distribution in the population, and African Americans tend to have higher Lp(a) levels than do other ethnic groups in the United States. Few environmental factors or medications modulate plasma Lp(a) levels. The pathogenesis of Lp(a) may result from an antifibrinolytic potential and/or ability to bind oxidized lipoproteins. Statins do not decrease Lp(a) levels, in contrast to PCSK9 inhibitors which reduce Lp(a) levels modestly. A novel anti-sense RNA directed at the LPA gene mRNA has shown that long-term reduction of Lp(a) is feasible. Ongoing phase 3 cardiovascular outcomes study will determine the usefulness of this approach. Genetic polymorphisms at the LPA gene have shown a strong association with aortic calcification and may have a causal role in aortic stenosis. ,

Triglyceride-Rich Lipoproteins (see also Chapter 25 )

TRLs are circulating apo B-carrying particles that carry predominantly hydrophobic triglycerides but also contain cholesterol ester, and include chylomicrons, VLDL, IDL, and remnant particles. In the fasting state, triglycerides are carried predominantly by VLDL particles and their remnants, while in the fed state triglycerides are additionally carried by chylomicrons and their remnants. Plasma triglycerides may arise from a higher TRL particle number (reflected by increased apo B, since each TRL carries one apoB moiety on it) and/or due to greater particle size. Postprandial elevations in triglycerides and TRLs, often due to underlying abnormalities in TRL metabolism, may not be captured in the fasting state, hence guidelines have shifted to measuring nonfasting or fasting lipids for general risk screening or assessment. Population studies have shown that nonfasting triglycerides are equally or even more strongly associated with cardiovascular endpoints than fasting levels.

Nonfasting triglycerides are higher than fasting levels (by approximately 0.3 mmol/L or 27 mg/dL), and guidelines consider nonfasting triglycerides ≥2 mmol/L (175 mg/dL) as abnormal, while for fasting triglycerides the corresponding level is ≥1.7 mmol/L (150 mg/dL). The 2018 American Heart Association/American College of Cardiology (AHA/ACC) cholesterol guideline and the 2019 AHA/ACC prevention guideline consider fasting or nonfasting triglycerides greater than 175 mg/dL as a risk enhancing factor that could prompt consideration for initiating or intensifying statin therapy. There are some differences in the guideline cutpoints for severe hypertriglyceridemia, defined as fasting triglycerides greater than 10 mmol/L (885 mg/dL) in European guidelines or ≥500 mg/dL (5.7 mmol/L) in US guidelines. Because the triglyceride to cholesterol ratio in TRLs progressively increases as the hypertriglyceridemia becomes more severe, the Friedewald equation (which assumes a fixed triglyceride to cholesterol ratio of triglycerides/5 [mg/dL] or triglycerides/2.2 [mmol/L]) to calculate LDL-C is inaccurate when triglycerides are greater than 4.5 mmol/L (400 mg/dL), as it underestimates the true LDL-C at high triglyceride levels. Instead, it is preferable to use non-HDL-C or apo B instead of calculated LDL-C for LDL-related treatment decisions in patients with hypertriglyceridemia, as direct LDL-C assays may also be inaccurate. ,

Hypertriglyceridemia correlates with risk of cardiovascular disease, and postprandial increase of TRL particles is an important factor in atherogenesis. However, it is less clear if triglycerides directly contribute to atherogenesis, or whether triglycerides are a marker of another atherogenic moiety of TRLs, such as the cholesterol (remnant/VLDL cholesterol) or apolipoproteins (e.g., apo B, apo CIII) that are also carried by these particles. Like LDL particles, all TRL particles carry one apo B per particle. While some studies found independent associations for triglycerides with cardiovascular disease risk, a meta-analysis of 68 prospective studies (>300,000 individuals) found that the association of triglycerides with cardiovascular risk was lost after adjusting for non-HDL-C or apo B. Genetic studies suggest a causal association between plasma triglycerides and cardiovascular risk, but many genetic variants are pleotropic and often also associate with differences in VLDL/remnant cholesterol, LDL-C, apo B, Lp(a), or HDL-C, making it challenging to identify the causal atherogenic moiety of TRLs. In a Mendelian randomization analysis of genetic scores composed of triglyceride-lowering variants in the LPL gene and LDL-C lowering variants in the LDLR gene, both sets of variants associated with similar risk of coronary disease per unit difference in apo B, suggesting that genetic variants that affect apoB are similarly atherogenic, regardless of differences in LDL-C or triglycerides. In a meta-analysis of randomized trials ( N = 374,358 from 25 statin trials and 24 nonstatin trials), triglyceride lowering associated with lower cardiovascular risk (approximately 15% lower risk per 1 mmol/L reduction in triglycerides), which was somewhat lower than for LDL-C (approximately 20% lower risk per 1 mmol/L reduction in LDL-C) and attenuated when the Reduction of Cardiovascular Events with Icosapent Ethyl-Intervention Trial (REDUCE-IT) was excluded.

Clearance of TRL particles requires LPL, hepatic lipase, apoE and other catalytically active apolipoproteins, and functional LDL receptors and LRP1. Enzymatic hydrolysis of TRLs by LPL and acquisition of cholesteryl esters from HDL by CETP results in the formation of cholesterol-enriched remnant particles that are depleted of part of their triglyceride content (see Fig. 27.3 , parts 2, 3, and 9 ). However, defective TRL clearance alone is often not sufficient (except in rare monogenic cases), and overproduction of TRL is usually necessary for hypertriglyceridemia.

Severe hypertriglyceridemia can result from poorly controlled diabetes or from genetic disorders (polygenic or monogenic) of the processing enzymes or apolipoproteins, often in the context of secondary factors. Severe hypertriglyceridemia is often due to polygenic and nongenetic determinants, often exacerbated in the presence of secondary nongenetic factors, and less commonly due to monogenic conditions. , In patients with insulin resistance, metabolic syndrome, or diabetes, elevation of plasma triglyceride levels is usually mild to moderate and occurs most often in the presence of visceral (abdominal) obesity and a diet rich in calories, carbohydrates, and saturated fats. It is now recognized that the genetic basis for severe hypertriglyceridemia is multifactorial due to the combination of poor lifestyle on a background susceptibility of multiple genetic defects each of which raises triglycerides by only a fraction of a mmol/L, the cumulative sum of which produces a clinical phenotype (e.g., familial hypertriglyceridemia, familial combined hyperlipidemia, or type III dyslipidemia). Thus, hypertriglyceridemia is best viewed as a complex disorder arising from the interaction of genetic and nongenetic factors. Primary causes of hypertriglyceridemia ( Table 27.5 ) include the rare monogenic syndromes (homozygous or compound heterozygote mutations) in several canonical genes related to LPL, and the much more common polygenic or multifactorial cases, which could include contributions from rare heterozygote variants in these canonical genes and/or other common variants associated with elevated triglycerides. , Genetic screening of siblings of patients with monogenic disorders is suggested.

TABLE 27.5
Primary Causes of Hypertriglyceridemia
Adapted from Laufs U, Parhofer KG, Ginsberg HN, Hegele RA. Clinical review on triglycerides. Eur Heart J . 2020;41(1):99–109.
Mild-to-moderate HTG (TG 2.0–9.9mmol/L)
Multifactorial or polygenic HTG (formerly HLP Type 4 or familial HTG)
Complex genetic susceptibility (see above)
Dysbetalipoproteinemia (formerly HLP Type 3 or dysbetalipoproteinemia)
Complex genetic susceptibility (see above), plus
APOE E2/E2 homozygosity or
APOE dominant rare variant heterozygosity
Combined hyperlipoproteinemia (formerly HLP Type 2B or familial combined hyperlipidemia)
Complex genetic susceptibility (see above), plus
Accumulation of common small effect LDL-C-raising polymorphisms
Severe HTG (TG >10 mmol/L)
Monogenic chylomicronemia (formerly HLP Type 1 or familial chylomicronemia syndrome)
Lipoprotein lipase deficiency (Bi-allelic LPL gene mutations)
Apo C-II deficiency (Bi-allelic APOC2 gene mutations)
Apo A-V deficiency (Bi-allelic APOA5 gene mutations)
Lipase maturation factor 1 deficiency (Bi-allelic LMF1 gene mutations)
GPIHBP1 deficiency (Bi-allelic GPIHBP1 gene mutations)
Multifactorial or polygenic chylomicronemia (formerly HLP Type 5 or mixed hyperlipidemia)
Complex genetic susceptibility, including
Heterozygous rare large-effect gene variants for monogenic chylomicronemia (see above); and/or
Accumulated common small-effect TG-raising polymorphisms (e.g., numerous GWAS loci including APOA1-C3-A4-A5 ; TRIB1 , LPL , MLXIPL , GCKR , FADS1-2-3 , NCAN , APOB , PLTP , ANGPTL3 )
Other
Transient infantile HTG (glycerol-3-phosphate dehydrogenase 1 deficiency) from bi-allelic GPD1 gene mutations

Polygenic Hypertriglyceridemia (also Known as Familial Hypertriglyceridemia, Formerly Type IV Hyperlipoproteinemia)

Polygenic hypertriglyceridemia (formerly Type IV hyperlipidemia) results from both common and rare genetic variants that result in increased VLDL particles. This highly heterogeneous condition also has a strong environmental influence. The prevalence of polygenic hypertriglyceridemia ranges from 1 in 50 to 100. Hepatic overproduction of VLDL causes this condition (see Fig. 27.3 , part 4 ); the catabolism (uptake) of VLDL particles can be normal or reduced. Lipolysis by LPL appears adequate under basal conditions, but not with excess triglyceride load, especially following fatty meals. Plasma triglycerides, VLDL-C, and VLDL triglycerides rise moderately to markedly; the LDL-C level is usually low, as is HDL-C. Total cholesterol is normal or elevated, depending on VLDL-C levels. Fasting plasma concentrations of triglycerides range between 2.3 to 5.7 mmol/L (200 to 500 mg/dL). After a meal, plasma triglycerides may exceed 11.3 mmol/L (1000 mg/dL). Polygenic hypertriglyceridemia does not associate with clinical signs such as corneal arcus, xanthoma, and xanthelasmas. This condition has a weaker relationship with CAD than combined hyperlipoproteinemia (familial combined hyperlipidemia), and not all studies support this association. The disorder clusters in first-degree relatives, but varies phenotypically depending on sex, age, hormone use (especially estrogens), and diet. Alcohol intake potently stimulates hypertriglyceridemia in these subjects, as does caloric or carbohydrate intake.

Human genetic studies have shown that many cases of severe hypertriglyceridemia result from mutations in one or more of the genes associated with triglyceride metabolism (see Table 27.4 ). Lifestyle modifications should be the first step in treatment, including weight reduction in overweight individuals, limiting alcohol intake, reducing caloric intake, increasing exercise and withdrawal of hormones (estrogens and progesterone or anabolic steroids).

An unrelated X–linked genetic disorder, familial glycerolemia, may mimic familial hypertriglyceridemia because most measurement techniques for triglycerides use the measurement of glycerol after enzymatic hydrolysis of triglycerides. Diagnosis of familial hyperglycerolemia requires ultracentrifugation of plasma and analysis of glycerol.

A less common subset of patients have more severe hypertriglyceridemia which is characterized by a more severe elevation in both VLDL and chylomicrons and are classified as having polygenic chylomicronemia (formerly Type V hyperlipidemia). The etiology is also multifactorial. Although they have the same genetic variants as patients with polygenic hypertriglyceridemia, they carry more of these variants and/or have stronger secondary or metabolic risk factors (e.g., fat-rich diet, obesity, or poorly controlled diabetes), which results in overproduction of both VLDL and chylomicrons and decreased catabolism of these particles.

Combined Hyperlipoproteinemia (also Known as Familial Combined Hyperlipidemia or Formerly Type 2B Hyperlipidemia)

Combined hyperlipoproteinemia (formerly familial combined hyperlipidemia) is common, with a prevalence of approximately 1 in 50 to 100, and accounts for 10% to 20% of patients with premature CAD, yet the diagnosis is often missed clinically because of overlap in the lipid profile with that of diabetes and metabolic syndrome. Individuals with combined hyperlipoproteinemia have complex genetic susceptibility to elevated triglycerides, and in some patients also tendency to elevated LDL-C. Novel loci in the upstream transcription factor 1 (USF1) and stearoyl-CoA desaturase 1 genes are promising candidate genes. The description of loss of function in the angiopoietin-like protein-3 gene (ANGPTL3, a novel target for therapy) in a kindred with familial hypolipidemia renewed interest in the angiopoietin like proteins 3, 4 and 5 which modulate the activities of LPL and endothelial lipase (see Fig. 27.3 , part 2). The phenotype is determined by interaction of multiple susceptibility genes and the environment. Modifying factors include gender, age at onset, and comorbid states such as obesity, lack of exercise, and diet.

Combined hyperlipoproteinemia is characterized by the presence of elevated total cholesterol and/or triglyceride levels based on arbitrary cut points in several members of the same family. Advances in analytic techniques have added measurement of LDL-C and apo B levels. Considerable overlap exists between combined hyperlipoproteinemia and other conditions (e.g., familial dyslipidemic hypertension, metabolic syndrome, and hyperapobetalipoproteinemia). The condition has few clinical signs; corneal arcus, xanthomas, and xanthelasmas occur infrequently. Biochemical abnormalities include elevation of plasma total cholesterol and LDL-C levels (>90th to 95th percentile) and/or elevation of plasma triglycerides (>90th to 95th percentile)—a type IIb lipoprotein phenotype, often associated with low HDL-C and elevated apo B levels; small, dense LDL particles occur frequently. The combination of apo B greater than 120 mg/dL and elevated triglycerides together with a family history of premature cardiovascular disease identifies patients who may have this condition. Underlying metabolic disorders include hepatic overproduction of apo B–containing lipoproteins, delayed postprandial clearance of TRLs, and increased flux of FFAs to the liver. Experimental data have shown that FFAs and cholesteryl esters drive hepatic apo B secretion. Increased delivery of FFAs to the liver, as occurs in states of insulin resistance and visceral obesity, leads to increased hepatic apo B secretion. In addition to lifestyle therapies, patients benefit from statins and other LDL-C lowering therapies.

Dysbetalipoproteinemia (Formerly Type III Hyperlipoproteinemia)

Dysbetalipoproteinemia is a rare genetic lipoprotein disorder that is characterized by accumulation of remnant lipoprotein particles in plasma and affects approximately 1 in 10,000. Similar to polygenic hypertriglyceridemia, it is now recognized that it also has a similar complex genetic predisposition but in addition these patients also carry the apo E2/2 genotype. Apo E has three common alleles: E2, E3, and E4. The apo E2/2 genotype has a prevalence of approximately 0.7% to 1.0%, but dysbetalipoproteinemia only occurs in approximately 1% of subjects bearing the apo E2/2 genotype. Reasons for the relative rarity of dysbetalipoproteinemia are not fully understood. Other rare mutations of the gene for apo E or mutations in other genes associated with triglyceride metabolism contribute to the disease. The apo E2 allele has markedly decreased binding to the apo B/E receptor. The importance of the apo E gene and protein is underscored by the widespread use of the apo E–deficient mouse, which develops atherosclerosis. Diagnosis includes apo E genotyping or phenotyping, plasma ultracentrifugation for lipoprotein separation, or lipoprotein electrophoresis. Lipoprotein agarose gel electrophoresis shows a typical pattern of a broad band between the pre-beta (VLDL) and beta (LDL) lipoproteins, hence it was previously also referred to as “broad beta disease.”

Clinical findings include pathognomonic tuberous xanthomas and palmar striated xanthomas. Patients with this disease have increased cardiovascular risk and are prone to premature coronary disease and peripheral arterial disease. The lipoprotein profile shows increased cholesterol and triglyceride levels and reduced HDL-C. Measuring apo B can differentiate it from mixed dyslipidemia. Remnant lipoproteins (partly catabolized chylomicrons and VLDL) accumulate in plasma and become enriched with cholesterol esters. The defect results from abnormal apo E, which does not bind to hepatic receptors that recognize apo E as a ligand (see Fig. 27.3 , part 3). These patients have an elevated ratio of VLDL cholesterol to triglycerides, normally less than 0.7 (when measured in mmol/L; <0.30 in mg/dL), because of cholesteryl ester enrichment of remnant particles. Thus, calculation of LDL-C in such patients is unreliable. In general, patients respond well to dietary therapy, correction of other metabolic abnormalities (diabetes, obesity, hypothyroidism), and in cases requiring drug therapy, statins should be the drug of first choice.

Monogenic Chylomicronemia Syndrome (Formerly Familial Hyperchylomicronemia Syndrome or Type I Hyperlipidemia)

This rare (approximately 1 to 10 in a million) monogenic autosomal recessive disorder of severe hypertriglyceridemia elevates fasting plasma triglycerides (often to greater than 11.3 mmol/L, >1000 mg/dL) and VLDL cholesterol due to increased levels of chylomicrons, but usually with lower apo B concentrations (<75 mg/dL) than polygenic or multifactorial chylomicronemia. , Mutations of several genes involved in triglyceride metabolism elevate chylomicrons, and definitive diagnosis requires molecular detection of homozygous or compound heterozygous variants in 1 of 5 canonical genes that encode proteins needed for LPL-mediated lipolysis of chylomicrons: LPL (the most common of at least 100 mutations identified), APOC2, APOA5, LMF1 , or GPIHBP1. Populations with a founder effect can have a high prevalence of LPL mutations. The hypertriglyceridemia results from markedly reduced or absent LPL activity or, more rarely, absence of its activator apo C-II (see Fig. 27.3 , part 2 ). These defects lead to impaired hydrolysis of chylomicrons and VLDL and their accumulation in plasma, especially after meals. Extreme elevations of plasma triglycerides (>113 mmol/L; >10,000 mg/dL) can result. Heterozygotes for the disorder tend to have an increase in fasting plasma triglycerides and smaller, denser LDL particles. Many patients with complete LPL deficiency exhibit failure to thrive in childhood and have recurrent bouts of pancreatitis. To underscore the importance of the role of LPL, lpl gene deficiency in the mouse leads to a perinatal lethal phenotype.

These patients have recurrent bouts of pancreatitis and eruptive xanthomas. Severe hypertriglyceridemia can also associate with lipemia retinalis, xerostomia, xerophthalmia, and behavioral abnormalities. Plasma from a patient with very high triglyceride levels is milky white, and a clear band of chylomicrons can be seen on top of the plasma after it stands overnight in a refrigerator. Treatment of acute pancreatitis includes intravenous hydration and avoidance of fat in the diet (including fat in parenteral nutrition), and only rarely requires plasma filtration. Chronic treatment includes avoidance of alcohol and dietary fat. Addition of short-chain fatty acids (which are not incorporated in chylomicrons) can increase palatability of the diet. Novel biologic agents are being evaluated and include therapies that target genes critical in the regulation of TRL, such as apo CIII, LPL, or intestinal diacylglycerol acyltransferase 1 (DGAT1) (see Novel Medications section).

High-Density Lipoproteins

Reduced plasma levels of HDL-C consistently correlate with the development or presence of ASCVD. Most cases of reduced HDL-C accompany elevated plasma triglycerides and often keep company with other features of metabolic syndrome. Genetic disorders of HDL can result from decreased production or abnormal maturation and increased catabolism. Genetic lipoprotein disorders leading to moderate to severe elevations in plasma triglycerides cause a reduction in HDL-C levels. Monogenic hyperchylomicronemia, polygenic hypertriglyceridemia, and combined hyperlipoproteinemia all associate with reduced HDL-C levels. Plasma triglyceride and HDL-C levels vary inversely. Several mechanisms contribute to this association: (1) decreased lipolysis of TRLs decreases the availability of substrate (phospholipids) for HDL maturation, (2) HDL enriched with triglyceride has an increased catabolic rate and hence reduced plasma concentration, and (3) the augmented pool of TRLs saps cholesterol from the HDL compartment by CETP-mediated exchange.

Disorders of High-Density Lipoprotein Biogenesis

You're Reading a Preview

Become a Clinical Tree membership for Full access and enjoy Unlimited articles

Become membership

If you are a member. Log in here