Skip to main content
The Journal of Clinical Endocrinology and Metabolism logoLink to The Journal of Clinical Endocrinology and Metabolism
. 2012 Sep;97(9):2969–2989. doi: 10.1210/jc.2011-3213

Evaluation and Treatment of Hypertriglyceridemia: An Endocrine Society Clinical Practice Guideline

Lars Berglund 1, John D Brunzell 1, Anne C Goldberg 1, Ira J Goldberg 1, Frank Sacks 1, Mohammad Hassan Murad 1, Anton F H Stalenhoef 1
PMCID: PMC3431581  PMID: 22962670

Abstract

Objective:

The aim was to develop clinical practice guidelines on hypertriglyceridemia.

Participants:

The Task Force included a chair selected by The Endocrine Society Clinical Guidelines Subcommittee (CGS), five additional experts in the field, and a methodologist. The authors received no corporate funding or remuneration.

Consensus Process:

Consensus was guided by systematic reviews of evidence, e-mail discussion, conference calls, and one in-person meeting. The guidelines were reviewed and approved sequentially by The Endocrine Society's CGS and Clinical Affairs Core Committee, members responding to a web posting, and The Endocrine Society Council. At each stage, the Task Force incorporated changes in response to written comments.

Conclusions:

The Task Force recommends that the diagnosis of hypertriglyceridemia be based on fasting levels, that mild and moderate hypertriglyceridemia (triglycerides of 150–999 mg/dl) be diagnosed to aid in the evaluation of cardiovascular risk, and that severe and very severe hypertriglyceridemia (triglycerides of > 1000 mg/dl) be considered a risk for pancreatitis. The Task Force also recommends that patients with hypertriglyceridemia be evaluated for secondary causes of hyperlipidemia and that subjects with primary hypertriglyceridemia be evaluated for family history of dyslipidemia and cardiovascular disease. The Task Force recommends that the treatment goal in patients with moderate hypertriglyceridemia be a non-high-density lipoprotein cholesterol level in agreement with National Cholesterol Education Program Adult Treatment Panel guidelines. The initial treatment should be lifestyle therapy; a combination of diet modification and drug therapy may also be considered. In patients with severe or very severe hypertriglyceridemia, a fibrate should be used as a first-line agent.

Summary of Recommendations

1.0. Diagnosis and definitions

1.1. Severe and very severe hypertriglyceridemia increase the risk for pancreatitis, whereas mild or moderate hypertriglyceridemia may be a risk factor for cardiovascular disease. Therefore, similar to the National Cholesterol Education Program Adult Treatment Panel (NCEP ATP) III guideline committee's recommendations, we recommend screening adults for hypertriglyceridemia as part of a lipid panel at least every 5 yr (1/⊕⊕○○).

1.2. We recommend basing the diagnosis of hypertriglyceridemia on fasting triglyceride levels and not on nonfasting triglyceride levels (1/⊕⊕⊕○).

1.3. We recommend against the routine measurement of lipoprotein particle heterogeneity in patients with hypertriglyceridemia (1/⊕⊕○○). We suggest that measurement of apolipoprotein B (apoB) or lipoprotein(a) [Lp(a)] levels can be of value, whereas measurement of other apolipoprotein levels has little clinical value (2/⊕⊕○○).

2.0. Causes of elevated triglycerides—primary and secondary

2.1. We recommend that individuals found to have any elevation of fasting triglycerides should be evaluated for secondary causes of hyperlipidemia including endocrine conditions and medications. Treatment should be focused on such secondary causes (1/⊕⊕○○).

2.2. We recommend that patients with primary hypertriglyceridemia be assessed for other cardiovascular risk factors, such as central obesity, hypertension, abnormalities of glucose metabolism, and liver dysfunction (1/⊕⊕○○).

2.3. We recommend that clinicians evaluate patients with primary hypertriglyceridemia for family history of dyslipidemia and cardiovascular disease to assess genetic causes and future cardiovascular risk (1/⊕⊕○○).

3.0. Management of hypertriglyceridemia

3.1. We recommend lifestyle therapy, including dietary counseling to achieve appropriate diet composition, physical activity, and a program to achieve weight reduction in overweight and obese individuals as the initial treatment of mild-to-moderate hypertriglyceridemia (1/⊕⊕○○).

3.2. For severe and very severe hypertriglyceridemia (>1000 mg/dl), we recommend combining reduction of dietary fat and simple carbohydrate intake with drug treatment to reduce the risk of pancreatitis (1/⊕⊕⊕⊕).

3.3. We recommend that the treatment goal for patients with moderate hypertriglyceridemia be a non-high-density lipoprotein (HDL) cholesterol level in agreement with NCEP ATP guidelines (1/⊕⊕○○).

3.4. We recommend that a fibrate be used as a first-line agent for reduction of triglycerides in patients at risk for triglyceride-induced pancreatitis (1/⊕⊕⊕○).

3.5. We suggest that three drug classes (fibrates, niacin, n-3 fatty acids) alone or in combination with statins be considered as treatment options in patients with moderate to severe triglyceride levels (2/⊕⊕○○).

3.6. We recommend that statins not be used as monotherapy for severe or very severe hypertriglyceridemia. However, statins may be useful for the treatment of moderate hypertriglyceridemia when indicated to modify cardiovascular risk (1/⊕⊕○○).

Method of Development of Evidence-Based Recommendations

The Task Force followed the approach recommended by the Grading of Recommendations, Assessment, Development, and Evaluation group (1). A detailed description of this grading scheme has been published (2). In brief, strong recommendations use the phrase “we recommend” and the number 1, and weak recommendations use the phrase “we suggest” and the number 2. The Task Force has confidence that patients who receive care according to the recommendations will derive, on average, more good than harm. Suggestions require more careful consideration of the patient's circumstances, values, and preferences. Cross-filled circles indicate the quality of the evidence: ⊕○○○ denotes very low quality evidence; ⊕⊕○○, low quality; ⊕⊕⊕○, moderate quality; and ⊕⊕⊕⊕, high quality. The quality of the evidence indicates the panel's confidence that the estimates of risks and benefits associated with the recommended course of action compared with an alternative course of action are correct and unlikely to change importantly with new research.

In developing the recommendations for the management of hypertriglyceridemia, the Task Force acknowledges the observational nature of the available evidence and the dependence on epidemiological studies. Yet, the Task Force made several strong recommendations based on several assumptions of patients' values and preferences. These values include that lifestyle therapy/modification is preferred over pharmacological interventions and that laboratory screening of hyperlipidemia, which includes screening for hypertriglyceridemia, is acceptable by patients, is feasible, and may be cost-effective if it leads to the prevention of cardiovascular events. The Task Force also considered that the majority of at-risk patients will likely place higher value on preventing clinically important cases of cardiovascular events and pancreatitis than on the burden of long-term pharmacological treatment, which may include side effects, cost, and the need for long-term monitoring.

The Endocrine Society maintains a rigorous conflict of interest review process for the development of clinical practice guidelines. All Task Force members must declare any potential conflicts of interest, which are reviewed before they are approved to serve on the Task Force and periodically during the development of the guideline. The conflict of interest forms are vetted by the Clinical Guidelines Subcommittee (CGS) before the members are approved by the Society's Council to participate on the guideline Task Force. Participants in the guideline development must include a majority of individuals without conflict of interest in the matter under study. Participants with conflicts of interest may participate in the development of the guideline but they have disclosed all conflicts. The CGS and the Task Force have reviewed all disclosures for this guideline and resolved or managed all identified conflicts of interest.

Conflicts of interest are defined by remuneration in any amount from the commercial interest(s) in the form of grants; research support; consulting fees; salary; ownership interest (e.g., stocks, stock options, or ownership interest excluding diversified mutual funds); honoraria or other payments for participation in speakers' bureaus, advisory boards, or boards of directors; or other financial benefits. Completed forms are available through The Endocrine Society office.

Funding for this guideline was derived solely from The Endocrine Society and thus the Task Force received no funding or remuneration from commercial or other entities.

1.0. Diagnosis and definitions

To date, treatment of hyperlipidemia has centered on the management of plasma total and low-density lipoprotein (LDL) cholesterol levels. Although there is robust evidence for an association between LDL cholesterol levels and cardiovascular disease (CVD), there has been more uncertainty regarding the meaning of the association between triglyceride levels and CVD. A high triglyceride level is one of the components of the metabolic syndrome. The latter is associated with risk for CVD, and there is growing support for unadjusted elevated triglyceride levels as an independent CVD risk factor. However, the extent to which elevated triglycerides constitute a direct risk for CVD or more likely represent a marker for other lipoprotein abnormalities associated with CVD risk is unknown and under extensive investigation.

Recommendation

1.1. Severe and very severe hypertriglyceridemia increase the risk for pancreatitis, whereas mild or moderate hypertriglyceridemia may be a risk factor for cardiovascular disease. Therefore, similar to the NCEP ATP III guideline committee's recommendations, we recommend screening adults for hypertriglyceridemia as part of a fasting lipid panel at least every 5 yr (1/⊕⊕○○).

1.1. Evidence

Serum triglycerides are routinely measured under fasting conditions to obtain more stable concentrations and to enable the physician to calculate LDL cholesterol levels. In addition, hypertriglyceridemia and postprandial lipemia may affect the measurement of HDL cholesterol and therefore the calculation of non-HDL cholesterol. The NCEP ATP III arbitrarily divided fasting serum triglycerides into four different classes (3) as outlined in Table 1. Classification of serum triglyceride levels greater than 150 mg/dl (1.7 mmol/liter) as elevated is mainly based on large prospective observational studies. However, the exact level at which serum triglycerides start to confer risk or become a marker for CVD is unknown, but it may be even lower than 150 mg/dl (1.7 mmol/liter) (4). Serum triglycerides are higher in men and increase with age in both sexes (5). A serum triglyceride level of 150 mg/dl (1.7 mmol/liter) usually falls below the 75th percentile in various populations, although there have been well-established differences identified between racial and ethnic groups (69).

Table 1.

Criteria proposed for clinical diagnosis of elevated triglyceride levels under fasting conditions

NCEP ATP III (3) The Endocrine Society 2010a
Normal <150 mg/dl <1.7 mmol/liter Normal <150 mg/dl <1.7 mmol/liter
Borderline-high triglycerides 150–199 mg/dl 1.7–2.3 mmol/liter Mild hypertriglyceridemia 150–199 mg/dl 1.7–2.3 mmol/liter
High triglycerides 200–499 mg/dl 2.3–5.6 mmol/liter Moderate hypertriglyceridemia 200–999 mg/dl 2.3–11.2 mmol/liter
Very high triglycerides ≥500 mg/dl ≥5.6 mmol/liter Severe hypertriglyceridemia 1000–1999 mg/dl 11.2–22.4 mmol/liter
Very severe hypertriglyceridemia ≥2000 mg/dl ≥22.4 mmol/liter
a

The criteria developed for the present guidelines focus on the ability to assess risk for premature CVD vs. risk for pancreatitis. The designations of mild and moderate hypertriglyceridemia correspond to the range of levels predominant in risk assessment for premature CVD, and this range includes the vast majority of subjects with hypertriglyceridemia. Severe hypertriglyceridemia carries a susceptibility for intermittent increases in levels above 2000 mg/dl and subsequent risk of pancreatitis; very severe hypertriglyceridemia is indicative of risk for pancreatitis. In addition, these levels suggest different etiologies. Presence of mild or moderate hypertriglyceridemia is commonly due to a dominant underlying cause in each patient, whereas severe or very severe hypertriglyceridemia is more likely due to several contributing factors.

To focus attention on the very high triglyceride levels that are a risk factor for pancreatitis, we have modified the NCEP ATP III triglyceride classification to include an additional classification of very severe hypertriglyceridemia, i.e. levels above 2000 mg/dl (Table 1). Severe hypertriglyceridemia, defined as 1000–1999 mg/dl, although not causative of pancreatitis, indicates risk for development of very severe hypertriglyceridemia (10, 11). Notably, the presence of mild and moderate hypertriglyceridemia as a consequence of treated severe hypertriglyceridemia may represent a cardiovascular risk factor.

Elevated triglyceride levels usually are seen with other metabolic abnormalities associated with increased CVD risk. Factors contributing to elevated serum triglycerides are overweight, physical inactivity, excess alcohol intake, presence of the metabolic syndrome or type 2 diabetes mellitus, as well as certain genetic disorders [familial hypertriglyceridemia (FHTG), familial combined hyperlipidemia (FCHL), and familial dysbetalipoproteinemia] (Table 2). Frequently, hypertriglyceridemia is a result of a combination of genetic factors and other causes of increased secretion or impaired clearance of triglyceride-rich lipoproteins. Based on the NCEP ATP III classification, the prevalence of hypertriglyceridemia is high in adults as well as in youth and adolescents, reflecting a population increase in body weight and obesity during the past several decades. In the National Health and Nutrition Examination Survey (NHANES), 1999–2004, 33% of the nearly 6000 participants (37% men, 30% women) had serum triglycerides of at least 150 mg/dl (≥1.7 mmol/liter) (5). In subjects aged 60 yr or older, the percentage was 42% (5). Of subjects with hypertriglyceridemia, about 14% had mild hypertriglyceridemia (150–200 mg/dl), whereas 16% had triglyceride levels of 200–500 mg/dl, and about 2% had levels above 500 mg/dl. Recent surveys of youth and adolescents in the United States (NHANES cycle 1996–2006) and Germany revealed abnormal lipid levels in 20–25% of the participants (12, 13). Lastly, a systematic review and meta-analysis of observational studies commissioned by The Endocrine Society found that hypertriglyceridemia is associated with increased risk of cardiovascular events and pancreatitis (14).

Table 2.

Causes of hypertriglyceridemia

Primary hypertriglyceridemia
    FCHL
    FHTG
    Familial dysbetalipoproteinemia
    FHA
    Familial chylomicronemia and related disorders
Primary genetic susceptibility
    Metabolic syndrome
    Treated type 2 diabetes
Secondary hypertriglyceridemia
    Excess alcohol intake
    Drug-induced (e.g. thiazides, β-blockers, estrogens, isotretinoin, corticosteroids, bile acid-binding resins, antiretroviral protease inhibitors, immunosuppressants, antipsychotics)
    Untreated diabetes mellitus
    Endocrine diseases
    Renal disease
    Liver disease
    Pregnancy
    Autoimmune disorders

Recommendation

1.2. We recommend basing the diagnosis of hypertriglyceridemia on fasting triglyceride levels and not on nonfasting triglyceride levels (1/⊕⊕⊕○).

1.2. Evidence

Prospective studies have indicated that, compared with fasting levels, nonfasting serum triglyceride levels may be a better or similar predictor of CVD events in the general population (1518). In a number of studies using standardized meals on the association of postprandial lipemia with CVD, greater CVD risk was found to be associated with increased hypertriglyceridemia (19, 20).

Investigators of the Multiple Risk Factor Intervention Trial (MRFIT) concluded that average fasting [187 mg/dl (2.11 mmol/liter)] and nonfasting [284 mg/dl (3.21 mmol/liter)] triglyceride levels were similarly predictive for nonfatal or fatal coronary heart disease with hazard ratios of 1.64 and 1.46, respectively (21). Two recent population-based studies have addressed CVD risk and nonfasting triglyceride levels. The Copenhagen City Heart Study (15) comprised 7587 women and 6394 men, aged 20 to 93 yr, recruited from the general population and followed for a mean of 26 yr. After adjustment for other cardiovascular risk factors (age, total cholesterol, body mass index, hypertension, diabetes, smoking, alcohol consumption, physical inactivity, lipid-lowering therapy, postmenopausal status, and hormone replacement therapy in women), hazard ratios for quintiles of nonfasting triglyceride levels vs. the reference level of less than 89 mg/dl (<1.0 mmol/liter) were as follows: between 1.7 and 5.4 for women and 1.4 to 2.4 for men for myocardial infarction; between 1.2 and 2.6 for women and 1.1 to 1.5 for men for ischemic heart disease; and between 1.3 and 3.3 for women and 1.2 to 1.8 for men for total death. All results were significant for trend with increasing triglyceride level. Limitations of this analysis include a small number of subjects with high triglycerides, no adjustment for HDL cholesterol, and a lack of fasting triglyceride levels for comparison.

The Women's Health Study (16) followed 26,509 initially healthy U.S. women older than 45 yr of age for a median of 11.4 yr; testing was done in 20,118 fasting and 6,391 nonfasting participants. The overall rate of cardiovascular events was 3.46/1000 person-years of follow-up. Although fasting triglyceride levels predicted cardiovascular events, the authors did not find an independent association with cardiovascular events after adjusting for potential confounders. In contrast, higher nonfasting triglyceride levels were independently associated with an increased risk of future events with hazard ratios for increasing tertiles of 1.0 (reference group, <104 mg/dl), 1.44 [95% confidence interval (CI), 0.90–2.29; 105–170 mg/dl], and 1.98 (95% CI, 1.21–3.25; >171 mg/dl) (P = 0.006 for trend). Associations were strongest among individuals who had their blood drawn 2 to 4 h after a meal and weakened with increasing time after the participants' last meal. Triglyceride levels and event rates were lower among the healthy U.S. women than those reported in the Copenhagen City Heart Study (15).

Although these studies provide some support for the hypothesis that nonfasting or postprandial lipid levels may be a more potent predictor of CVD risk than fasting levels, the lack of standardization and reference levels impedes a general implementation of nonfasting triglyceride or remnant particle levels (22). Further work is needed on determining the most informative procedure of collecting postprandial lipids and characterization of postprandial effects on triglyceride measurements (23). Thus, at present, the diagnosis of hypertriglyceridemia is suggested to be based on fasting levels where the length of fast is recommended to be 12 h. During this time period, intake of liquids without caloric content is acceptable.

Recommendation

1.3. We recommend against the routine measurement of lipoprotein particle heterogeneity in patients with hypertriglyceridemia (1/⊕⊕○○). We suggest that measurement of apoB or Lp(a) levels can be of value, whereas measurement of other apolipoprotein levels has little clinical value (2/⊕⊕○○).

1.3. Evidence

In most hypertriglyceridemic patients, the distribution of both LDL and HDL sizes is shifted to smaller particles (24). In patients with the metabolic syndrome, treated type 2 diabetes mellitus, or FCHL, the number of small, dense LDL and HDL particles and the apoB levels are increased (25, 26). Hepatic lipase and cholesterol ester transfer protein contribute to the remodeling processes; whether hepatic lipase or cholesterol ester transfer protein has the predominant effect on the size and density of LDL and HDL particles depends on the triglyceride content of very low-density lipoproteins (VLDL) (27, 28). Although the LDL cholesterol level is frequently normal in patients with these conditions, the concentration of LDL particles is generally increased because of the presence of a higher number of cholesterol-poor, small, dense LDL particles. It is not necessary to measure LDL size or density; however, measurement of non-HDL cholesterol and/or apoB levels can indicate the presence of increased numbers of LDL particles (3, 29).

Epidemiological studies vary as to the independent association between large or small LDL and atherosclerotic cardiovascular disease (30). Several prospective studies suggest that circulating levels of small, dense LDL particles are better predictors of coronary atherosclerosis, carotid atherosclerosis, and response to therapy than are levels of large, buoyant LDL particles (28, 3133). There is wide agreement, however, that the concentration of LDL, regardless of the particle size, predicts coronary heart disease (3). An increase in small, dense LDL particles is not reflected in the LDL cholesterol concentration. Statins reduce the concentration of all sizes of LDL, and their benefits to CVD are universal across population groups that have large or small LDL (3, 34). Neither LDL size nor the concentration of small, dense LDL particles adds to CVD prediction in multiple variable analysis beyond the standard lipid risk factors, although small LDL particles predict cardiovascular risk in univariate analysis (3537). Prospective multivariate studies demonstrate that large LDL predicts atherosclerosis and coronary heart disease (35, 38, 39). Several reports also show that measurement of apoB is superior to measurement of LDL or even non-HDL cholesterol as an indicator of global CVD risk (40, 41).

Lp(a) has many properties in common with LDL but contains a unique protein, apolipoprotein(a), which is linked to apoB-100 by a single disulfide bond (42). Recently, interest in Lp(a) has increased because studies over the past decade have confirmed and more robustly demonstrated a risk factor role of Lp(a) for cardiovascular disease (4346). However, there are limited treatment options to alter its level and a current lack of outcome evidence supporting its use as a specific therapeutic target.

Small HDL in hypertriglyceridemia is associated with hypercatabolism of apolipoprotein A-I (47, 48), and it seems to be related to elevated hepatic lipase activity in central obesity and insulin resistance (49). This property may impair the ability of HDL to take up sufficient cholesterol from peripheral cells. Epidemiological studies have not provided conclusive evidence that measurement of HDL size contributes to risk prediction (36, 5055). For these reasons, assessment of lipoprotein heterogeneity is not recommended in the assessment of hypertriglyceridemia.

2.0. Causes of elevated triglycerides—primary and secondary

Pathophysiology

Triglycerides are the most dense form of calories and serve as an important source of energy. Dietary triglycerides are assembled in the gut into chylomicrons. Their interaction with lipoprotein lipase (LpL) located on the luminal surface of capillary endothelial cells leads to liberation of free fatty acids from triglyceride; free fatty acids are able to traverse cell membranes. Only 50% of chylomicron triglyceride is estimated to be lost in this process, and the remainder of the lipoprotein, called a chylomicron remnant, contains lipids such as cholesteryl esters, retinyl esters, and apoB-48. Several proteins, called apolipoproteins (apo), regulate LpL actions and lipoprotein clearance from the liver. apoC-II is the necessary cofactor for LpL actions. apoC-III blocks the uptake of lipoproteins by receptors in the liver and may impair LpL. apoE is the ligand for hepatic uptake of triglyceride-rich remnants. VLDL particles are produced by the liver, and the VLDL triglyceride content is derived from a variety of substrates including lipoprotein triglyceride, free fatty acids, and de novo fatty acids synthesized from carbohydrates. VLDL triglycerides lose free fatty acids by the action of LpL, leading to production of VLDL remnants, also referred to as intermediate-density lipoproteins (IDL), and eventually to conversion to LDL.

The plasma triglyceride level reflects the concentration of the triglyceride-carrying lipoproteins (VLDL and chylomicrons). The concentration of VLDL cholesterol and apoB is at least 10 times higher than the corresponding chylomicron concentration, even after consumption of a large amount of fat (5659). These lipoproteins contain at least as much cholesterol per particle as does LDL. Although triglyceride itself is not a component of arterial plaque, it is thought that cholesterol within triglyceride-rich particles may contribute to plaque development (60, 61).

Hypertriglyceridemia results from increased triglyceride production, or reduced triglyceride catabolism, or both. The common forms of hypertriglyceridemia emerge as adults get older and become overweight and sedentary and develop insulin resistance. The most common setting of hypertriglyceridemia is that found with metabolic syndrome, FCHL, and type 2 diabetes. The increase in triglyceride production may be due to excess free fatty acids returning to the liver, particularly in the setting of visceral obesity and insulin resistance, and increased de novo triglyceride production due to hyperinsulinemia (24, 62, 63). In hypertriglyceridemia, more VLDL particles, as measured by apoB, and larger and more triglyceride- and apoC-III-enriched lipoproteins are found (39, 64, 65). Hepatic insulin resistance may contribute to a high production rate of VLDL because insulin reduces apoB synthesis and VLDL secretion in the liver (66, 67). Although insulin resistance is associated with high triglycerides, VLDL and triglyceride concentrations can be similar in patients with widely divergent insulin sensitivity (68, 69). Acute or chronic elevation of insulin in response to a high carbohydrate diet did not lower serum triglyceride levels in healthy subjects (70, 71). In African-Americans, low triglyceride levels occur in the context of severe insulin resistance (72). Thus, in an individual patient, the contribution of insulin resistance to overproduction of triglycerides and VLDL may be variable.

Clearance of VLDL from the circulation is reduced in many patients with hypertriglyceridemia (64, 65, 73), in part due to saturation of triglyceride clearance (74). This saturation might occur owing to defective triglyceride hydrolysis by LpL and/or reduced clearance of VLDL and chylomicron remnants by the liver. Defective lipolysis occurs with genetic defects in LpL; defects in apoC-II; defective association of LpL with the vascular wall due to antibodies to heparin or defects in glycosylphosphatidylinositol-anchored high-density lipoprotein-binding protein 1, an LpL-binding protein (75, 76); or defective intracellular LpL processing due to mutated lipase maturation factor 1 (77). Severe hyperchylomicronemia also occurs when a secondary cause of hypertriglyceridemia such as diabetes or pregnancy is superimposed on an underlying genetic defect (78). A number of additional genetic factors influence human triglyceride levels, including mutations in apoC-III, apoE, apoA-V, and angiopoietin-like protein 4. apoE is the main protein that mediates binding of VLDL and chylomicron remnants to hepatic receptors and proteoglycans; it is antagonized by apoC-III. Hypertriglyceridemic VLDL particles are heterogeneous and often have a high apoC-III/apoE ratio, causing reduced clearance and increased conversion to LDL. Recent studies have underscored the difference in metabolism of VLDL subpopulations containing apoC-III, with or without apoE, and how these apolipoproteins are involved to establish hypertriglyceridemia and cause the formation of dense LDL (79).

Moderate hypertriglyceridemia, i.e. 200–999 mg/dl, is due to excess circulating VLDL, the principal triglyceride carrier in the circulation. Defective clearance of triglyceride-rich VLDL by LpL can contribute to this condition, and many patients have overproduction of VLDL triglyceride in the liver with an increased secretion (24, 62, 63).

In patients with severe or very severe triglyceride levels (≥1000 mg/dl), the LpL removal system is saturated (74). This saturation occurs whether hypertriglyceridemia is primarily due to defective lipolysis or excessive production of endogenous triglyceride, and it leads to reduced catabolism of dietary triglyceride incorporated into chylomicrons. For this reason, there is concern that triglyceride levels above 1000 mg/dl can rapidly increase after a fat-rich meal. Foods that contain potent substrates for triglyceride production such as simple sugars, fructose, and alcohol can substantially increase triglyceride levels in susceptible people (80, 81). Very severe triglyceride levels (>2000 mg/dl) are associated with lipemic serum and risk of pancreatitis in the chylomicronemia syndrome (82).

Recommendation

2.1. We recommend that individuals found to have any elevation of fasting triglycerides should be evaluated for secondary causes of hyperlipidemia including endocrine conditions and medications. Treatment should be focused on such secondary causes (1/⊕⊕○○).

2.1. Evidence

An isolated elevation in triglyceride levels may be caused by a primary disorder of lipid metabolism, e.g. FHTG or FCHL. It may also arise secondary to a number of conditions as outlined in Table 2, including a number of medications, a high-carbohydrate diet with intake of simple sugars, or as a component of endocrine and other diseases, inflammation, or some rare genetic diseases. In the setting of common, underlying genetic dyslipidemias, such secondary causes may lead to severe and very severe triglyceride levels and the risk of pancreatitis.

Endocrine disorders

Patients with untreated diabetes mellitus and insulin deficiency commonly have hypertriglyceridemia; this condition occurs more frequently in type 2 than in type 1 diabetes mellitus. Appropriate diabetes management reduces triglyceride levels. Mild hypertriglyceridemia, typically seen in treated type 2 diabetes, is probably related to the presence of central obesity and insulin resistance (83).

Hypertriglyceridemia related to increased insulin resistance and to decreased activity of both hepatic lipase and LpL occurs in some acromegalic patients (84).

Owing to estrogen-induced stimulation of the secretion of hepatic triglyceride-rich lipoprotein, triglyceride levels increase progressively during pregnancy, with levels in the third trimester increased by 200% or more over levels before pregnancy. In women with underlying disorders of triglyceride metabolism or overproduction, an estrogen-induced increase in triglycerides during pregnancy can result in a risk of pancreatitis with potential fetal loss (85, 86). Oral estrogen in the form of estrogen replacement therapy or oral contraceptives has a triglyceride-increasing effect due to increased hepatic VLDL production. This effect does not occur with transdermal estrogen due to its lesser exposure to the liver (87, 88). Tamoxifen, a selective estrogen receptor modulator, can also increase triglyceride levels. The effect is less pronounced with raloxifene, but that drug can increase levels in women with an underlying propensity to hypertriglyceridemia (89, 90).

Thyroid hormone deficiency is associated with increased LDL cholesterol levels; this increase has been postulated to be due to decreased function of LDL receptors. There can also be an increase in triglyceride levels. Hypothyroidism can lead to the expression of dysbetalipoproteinemia (9193).

Glucocorticoids have several effects on lipoprotein metabolism including increased cholesterol production from induction of hydroxymethylglutaryl coenzyme A reductase, increased fatty acid synthesis due to increased expression of fatty acid synthase, and decreased clearance of triglyceride-rich lipoproteins (94). Because weight gain and insulin resistance are major effects of both exogenous and endogenous glucocorticoid excess, elevated triglycerides can be seen in Cushing's syndrome as well as during glucocorticoid treatment.

Rare genetic disorders

Inherited and congenital lipodystrophies are associated with moderate-to-severe hypertriglyceridemia and are characterized by loss of adipose tissue and either autosomal recessive or dominant inheritance (95). The loss of adipose tissue is selective and variable and may be partial or complete. Some forms manifest at birth, whereas others become evident later in life with loss of fat beginning in childhood and puberty (96). Varieties of familial partial lipodystrophy, which are rare autosomal disorders, involve fat loss from the extremities more than the trunk. The Kobberling variety is more common, but the defect is unknown (97). Hypertriglyceridemia is also seen in several types of glycogen storage disease in children (98).

Other conditions

Acquired lipodystrophy can be seen in patients with HIV infection who are being treated with highly active antiretroviral therapy (99). Other acquired forms of lipodystrophy are seen in patients with autoimmune diseases such as juvenile dermatomyositis. Patients with acquired generalized lipodystrophy lose fat from large areas of the body during childhood and adolescence and often have hepatic steatosis (100).

Hypertriglyceridemia has been reported in multiple myeloma and in autoimmune diseases such as systemic lupus erythematosis involving autoantibodies to LpL, apoC-II, or heparin. Hypertriglyceridemia can also be seen with infections including inflammation and sepsis, apparently due to increased production of VLDL (101). Hypertriglyceridemia in severe stress may be related to possible catecholamine induction of adipose tissue lipolysis and reduced LpL activity (102).

Renal and hepatic disease can be associated with hypertriglyceridemia. Nephrotic syndrome causes increased production of apoB-containing lipoproteins, including VLDL, by the liver (103). Hypertriglyceridemia is common in patients with renal failure and may be related to decreased clearance of triglyceride-rich lipoproteins via reduced LpL and hepatic lipase activities (104). Acute hepatitis may be associated with increased VLDL production and hypertriglyceridemia (105).

Drugs

Many drugs raise triglyceride levels. One of the most commonly used is alcohol. Alcohol intake increases hepatic fatty acid synthesis and decreases fatty acid oxidation, with a net effect to stimulate hepatic VLDL triglyceride secretion. The effects of alcohol vary interindividually, tend to be amplified in subjects with underlying lipid disorders, are dose-dependent (106), and may be related to the mode of intake (107).

Antihypertensive drugs with a potential to increase triglyceride levels are thiazide (and furosemide) diuretics and β-adrenergic blocking agents. The hypertriglyceridemic effect of β-adrenergic blocking agents is greater for atenolol, metoprolol, and propranolol than for carvedilol. These effects are most relevant in patients with underlying genetic hypertriglyceridemia (94).

Oral estrogens increase the hepatic secretion of VLDL, leading in turn to an increase in serum triglyceride levels (108). In patients with familial hypertriglyceridemia or LpL deficiency, the use of oral estrogens can provoke severe pancreatitis (109). An increase in hepatic VLDL and apoC-III production and perhaps a decrease in LpL leading to increased triglyceride levels are also seen during use of retinoids such as isotretinoin and the anticancer drug bexarotene (110112).

Bile acid sequestrants (cholestyramine, colestipol, colesevelam) can worsen hypertriglyceridemia and are contraindicated in patients with severe hypertriglyceridemia (>1000 mg/dl) and in patients with dysbetalipoproteinemia. Patients with normal baseline triglyceride levels experience minimal triglyceride increases with bile acid sequestrant therapy, but those with moderate hypertriglyceridemia (triglycerides > 200 mg/dl) may experience substantial further elevation (113).

Dyslipidemia is a frequent complication of antiretroviral therapy for HIV infection. In particular, the protease inhibitors ritonavir and lopinavir can increase plasma triglyceride levels (114).

Immunosuppressants such as sirolimus also increase triglyceride levels (115).

Certain second-generation antipsychotic medications such as clozapine, olanzepine, risperidone, and quetiapine can be associated with hypertriglyceridemia, but this effect has not been seen for aripiprazol or ziprasidone. Those that are associated with weight gain, insulin resistance, and worsening of the metabolic syndrome are particularly important contributors to secondary hyperlipidemia. Among selective serotonin reuptake inhibitors, sertraline may raise triglycerides (116).

Recommendation

2.2. We recommend that patients with primary hypertriglyceridemia be assessed for other cardiovascular risk factors, such as central obesity, hypertension, abnormalities of glucose metabolism, and liver dysfunction (1/⊕⊕○○).

2.2. Evidence

Elevated triglycerides can occur in the absence or presence of other lipid or lipoprotein disturbances. Patients with elevations in the levels of both total plasma cholesterol and triglyceride can be divided into three categories. In the first category, VLDL and/or LDL cholesterol levels are elevated, as in FCHL. In the second category, VLDL and chylomicron remnant cholesterol are elevated, as in familial dysbetalipoproteinemia. The third category consists of patients with severe and very severe hypertriglyceridemia in whom the increase in plasma cholesterol is a result of increased VLDL and chylomicron cholesterol.

Familial combined hyperlipidemia

The lipid phenotype in FCHL varies from isolated hypertriglyceridemia to isolated hypercholesterolemia within families and in single individuals, suggesting that the variation in the lipid phenotype is affected by environmental factors (78). In some subgroups, such as those with half-normal LpL activity, the lipoprotein phenotype seems to be more stable as hypertriglyceridemia and less stable as hypercholesterolemia. In patients with FCHL, increases in triglycerides and LDL cholesterol are often found, whereas elevated apoB levels and small, dense LDL particles are always seen (25). It has been suggested that measurement of apoB and non-HDL cholesterol levels, in addition to assessment of LDL and HDL cholesterol levels, will serve as a basis by which to identify these FCHL individuals at risk for premature CVD (117). In addition, patients with FCHL frequently have nonlipid cardiovascular risk factors (i.e. central obesity, hypertension, insulin resistance, and impaired glucose tolerance). The prevalence of FCHL in the population is estimated to be 1–2% and in CVD populations to be at least 10% (78). It should be underscored that regardless of the etiology, the combination of hypertriglyceridemia and elevated LDL cholesterol, with small, dense LDL particles, appears to increase the risk associated with elevated LDL cholesterol alone.

Familial hypertriglyceridemia

FHTG is a common inherited disorder, thought to be autosomal dominant, which affects about 1% of the population. It is characterized by an increased triglyceride synthesis, which results in very large triglyceride-enriched VLDL particles, secreted in normal numbers. Affected people have elevated VLDL levels, but low levels of LDL and HDL cholesterol, and are generally asymptomatic unless very severe hypertriglyceridemia develops. FHTG does not appear to be associated with an increased risk of premature CVD (118). However, subjects are at increased risk for the development of the chylomicronemia syndrome and pancreatitis when secondary forms of hypertriglyceridemia are present, such as untreated diabetes or use of triglyceride-raising drugs. A diagnosis is made by family history and examination of fasting lipoprotein profiles of the patient and relatives. The triglyceride level ranges from about 250 to 1000 mg/dl in approximately one half of first-degree relatives. A strong family history of premature CVD usually is lacking, and elevated LDL cholesterol levels are not present.

It is important to distinguish FHTG, which seems to confer no risk of premature CVD, from FCHL, which is associated with a high incidence of premature CVD (78). It is often difficult to distinguish these disorders when FCHL is associated with hypertriglyceridemia. Concomitant increased apoB or LDL cholesterol concentration indicates FCHL. FCHL is also strongly suggested by a positive personal or family history of premature atherosclerosis with hypertriglyceridemia (78).

Chylomicronemia syndrome

Chylomicronemia is associated with pancreatitis, but the mechanism is unclear. Pancreatitis may result from the release of excess fatty acids and lysolecithin from chylomicrons, exceeding the binding capacity of albumin in pancreatic capillaries. The chylomicronemia syndrome occasionally occurs with a genetic defect in the LpL-related triglyceride clearance system. More commonly, chylomicronemia is caused by the coexistence of a common genetic form of hypertriglyceridemia combined with an acquired disorder of plasma triglyceride metabolism, the most common being untreated diabetes (82). Another condition that may be implicated is the use of drugs that raise triglyceride levels.

The chylomicronemia syndrome is associated with abdominal pain, eruptive xanthomas on the buttocks and the extensor surfaces of the upper limb, transient memory loss, and the risk for artifactual alterations in laboratory analyses. If uncorrected, the chylomicronemia syndrome may result in acute, recurrent pancreatitis. The risk of pancreatitis markedly increases with very severe triglyceride levels above 2000 mg/dl (82), but it can be prevented by maintaining triglyceride levels below 1000 mg/dl. Severe hypertriglyceridemia can present in childhood as a result of LpL deficiency or, extremely rarely, as apoC-II, apoA-V, or glycosylphosphatidylinositol-anchored high-density lipoprotein-binding protein 1 deficiency. These patients are at risk for acute, recurrent pancreatitis with severe hypertriglyceridemia, and they must be treated with moderate to severe dietary fat restriction to reduce plasma triglyceride levels (78). In patients with severe and very severe triglyceride elevations, the increase in total plasma cholesterol is a result of the cholesterol in VLDL and chylomicrons.

Familial hypoalphalipoproteinemia with high triglycerides

In 1992, Genest et al. (119, 120) proposed that familial hypoalphalipoproteinemia (FHA), a disorder with elevated triglyceride and low HDL cholesterol, was a common genetic dyslipidemia associated with premature CVD. Many, if not most, patients with hypertriglyceridemia have a concomitant reduction in HDL cholesterol levels. It is not known whether FHA is a discrete genetic disorder. Low HDL cholesterol is commonly seen with premature CVD and may be related, in part, to mutations in proteins of HDL metabolism. However, mutations in these candidate genes are rare and account for few FHA cases. FHA is often confused with FHTG, and more studies are needed to characterize this condition.

Almost all forms of severe genetic HDL deficiency are associated with mild-to-moderate hypertriglyceridemia. These include Tangier's disease, apoA-I deficiency, and lecithin cholesterol acyl transferase deficiency.

Dysbetalipoproteinemia

Dysbetalipoproteinemia, also called type III hyperlipoproteinemia or remnant removal disease, is caused in part by a mutation in the APOE gene, resulting in impairment in the hepatic uptake of apoE-containing lipoproteins and reduction in the conversion of VLDL and IDL to LDL particles (121). In the absence of additional genetic, hormonal, or environmental factors, remnants do not accumulate to a degree sufficient to cause hyperlipidemia in fasting blood. Dysbetalipoproteinemia results when an apoE defect (almost always the E2/E2 genotype) occurs in conjunction with a second genetic or acquired defect that causes either overproduction of VLDL (such as FCHL) or a reduction in LDL receptor activity (such as occurs in heterozygous FH or hypothyroidism). Other rare apoE variants such as apoE3-Leiden and apoE2(Lys146→Gln) can also cause dysbetalipoproteinemia (122, 123). Patients with dysbetalipoproteinemia have elevations in both cholesterol and triglyceride levels (124). They are likely to develop premature CVD and are at increased risk for peripheral vascular disease. Clinical dyslipidemia usually does not develop before adulthood in men or before menopause in women. Palmar xanthomas, orange lipid deposits in the palmar creases, are pathognomonic, but are not always present. Tuberoeruptive xanthomas are occasionally found at pressure sites on the elbows, buttocks, and knees. The presence of dysbetalipoproteinemia should be suspected in a person with elevated total cholesterol and triglyceride levels that range from 300 to 1000 mg/dl and are roughly equal. VLDL particles are cholesterol-enriched, which can be determined by isolation of VLDL by ultracentrifugation. It can be useful to confirm the diagnosis by demonstrating the presence of the E2/E2 genotype.

Metabolic syndrome

Hypertriglyceridemia is one of the components of the metabolic syndrome, a constellation of metabolic risk factors including a central distribution of adiposity or visceral obesity, insulin resistance, impaired glucose tolerance, hypertension, and high triglycerides and/or low HDL-C, associated with an atherogenic, procoagulant, and proinflammatory state (125). Although criteria for defining the metabolic syndrome have differed among health organizations, recently a harmonized definition has been agreed to by leading cardiovascular and diabetes organizations including the National Heart, Lung, and Blood Institute, the International Diabetes Federation, and the American Heart Association (126). The five components that form the criteria for defining the metabolic syndrome are triglycerides above 150 mg/dl; HDL-C below 40 mg/dl in men or below 50 mg/dl in women; blood glucose above 100 mg/dl; blood pressure above 130 mm Hg systolic or above 85 mm Hg diastolic; and waist circumference greater than 102 cm in men or greater than 88 cm in women. A lower criterion for waist circumference is recommended for Asian populations. Three of these five criteria are needed to make the diagnosis of metabolic syndrome. Genetic and environmental factors appear to affect the distribution of these variables in both normal individuals and those with the metabolic syndrome. Type 2 diabetes mellitus, polycystic ovary syndrome, and FCHL may account for at least 40–50% of premature coronary artery disease in some populations with metabolic syndrome, and they need to be considered in assessing the risk of CVD in patients who have the metabolic syndrome (26).

Although the association of central obesity and insulin resistance with dyslipidemia is well established, the underlying mechanisms remain unclear. An increase in the level of portal vein long-chain nonesterified fatty acids (NEFA, or free fatty acids) has been suggested as an underlying factor. Increased visceral fat is associated with insulin resistance, hyperinsulinemia, low plasma adiponectin levels, and elevations in plasma NEFA levels (127). An increase in portal NEFA would potentially inhibit apoB-100 from undergoing degradation in the hepatic proteosome and would increase the likelihood of secretion of triglyceride-containing lipoproteins, contributing to increased triglyceride levels and an increased number of VLDL and LDL particles seen in patients with insulin resistance (24). Importantly, in normal, randomly selected healthy populations, isolated visceral obesity and insulin resistance were associated with only a slight increase in triglyceride levels and only a slight decrease in HDL cholesterol levels (127). Increased waist circumference and plasma triglyceride levels together confer greater CVD risk in these patients (128).

Ectopic fat accumulation

Excess tissue triglyceride accumulation results from increased uptake of circulating triglycerides (e.g. in LpL deficiency), greater production of triglyceride from carbohydrates or free fatty acids (e.g. in type 2 diabetes mellitus), or reduced utilization or secretion of triglyceride (e.g. in familial hypobetalipoproteinemia). Excess triglyceride stores are found in livers of patients with nonalcoholic fatty liver disease (129). Hepatic steatosis is often associated with increased intraabdominal fat and fat accumulation in other tissues such as skeletal muscle, heart, and perhaps pancreas, and may lead to insulin resistance associated with type 2 diabetes and metabolic syndrome (130). Cellular and animal experiments suggest that signaling lipids other than triglyceride, such as ceramides or diacylglycerols, may be pathological (131). Because hypertriglyceridemia, increased intraabdominal fat, and nonalcoholic fatty liver disease occur with insulin resistance and excess caloric intake, a cause-and-effect relationship is difficult to conclude. Recent human genetic studies have found several predisposing factors that might provide novel insights. At this time, routine assessment of hepatic fat content or intraabdominal fat content in hypertriglyceridemic patients is not indicated. An increase in hepatic fat content is associated with an increase in aminotransferase activities, in particular, alanine aminotransferase.

Recommendation

2.3. We recommend that clinicians evaluate patients with primary hypertriglyceridemia for family history of dyslipidemia and cardiovascular disease to assess genetic causes and future cardiovascular risk (1/⊕⊕○○).

2.3. Evidence

Whether serum triglycerides are causally related to atherosclerosis remains to be elucidated, as does the exact mechanism by which they may promote vascular disease. Factors contributing to this uncertainty are the complex metabolism of triglyceride-rich lipoproteins and the fact that abnormal triglyceride concentrations are seen frequently in conditions associated with increased CVD risk, such as type 2 diabetes mellitus, the metabolic syndrome, and the familial forms of hypertriglyceridemia, in the presence of low HDL-cholesterol levels and small dense LDL particles. An elevated serum triglyceride level might in some cases be a marker for CVD rather than a causal factor.

Several meta-analyses from studies performed in the general population have shown a modest but independent effect of triglycerides on CVD. A meta-analysis of Western population-based prospective studies, including 46,413 men and 10,864 women, showed an overall relative risk for CVD of 1.32 for men and 1.76 for women per 1 mmol/liter (88.5 mg/dl) increase of triglycerides (132). Adjustment for HDL cholesterol and other cardiovascular risk factors attenuated the relative risk attributed to triglycerides, although they remained significant (1.14 and 1.37, respectively) (132). More recently, a report of two nested case-control comparisons from population-based cohorts [the Reykjavik study and the European Prospective Investigation of Cancer (EPIC)-Norfolk study] comprised 44,237 Western middle-age men and women of predominantly European ancestry and a total of 3,582 cases of coronary heart disease (133). In the Reykjavik study, fasting triglyceride levels in cases were 105 ± 70 mg/dl (1.19 ± 0.79 mmol/liter) vs. 91 ± 55 mg/dl (1.03 ± 0.62 mmol/liter) in controls. The corresponding levels in the EPIC-Norfolk study were 195 ± 107 mg/dl (2.20 ± 1.21 mmol/liter) and 168 ± 104 mg/dl (1.90 ± 1.17 mmol/liter), respectively. Comparing individuals in the top vs. the bottom tertile, the adjusted odds ratio for CVD was 1.76 (95% CI, 1.39–2.21) in the Reykjavik study and 1.57 (95% CI, 1.10–2.24) in the EPIC-Norfolk study. Adjustment for HDL cholesterol attenuated the effect to an odds ratio of 1.31 (95% CI, 1.06–1.62) in the EPIC-Norfolk study. In addition, an updated meta-analysis of prospective studies in Western populations, providing information in aggregate from more than 10,000 coronary heart disease cases involving more than 260,000 participants, reported an adjusted odds ratio of 1.72 (95% CI, 1.56–1.90) comparing the top and bottom triglyceride tertiles (133). These results are similar to those reported by another nonoverlapping meta-analysis based on Asian and Pacific populations, although the absolute risk in these populations was much lower (134). This latter meta-analysis calculated a relative risk for coronary heart disease, adjusted for several established risk factors, of 1.80 (95% CI, 1.49–2.19), comparing subjects in the top with those in the bottom quintile of triglyceride levels.

A third large prospective cohort study, the MELANY study, was conducted in 13,953 healthy male soldiers (aged 26 to 45 yr) in Israel (135). After multivariate adjustment (including age, body mass index, HDL cholesterol, physical activity, fasting glucose, mean arterial blood pressure, smoking), men in the top quintile of triglyceride levels had a hazard ratio for coronary heart disease of 4.05 (95% CI, 2.68–8.61) compared with the lowest quintile (P = 0.001 for trend) (135). Beyond fasting triglyceride levels, the change in triglyceride levels strongly predicted incident coronary heart disease (135). Finally, the Emerging Risk Factors Collaboration (ERFC) collected 112 prospective studies of cardiovascular risk factors, involving a total of 1.2 million participants in a central database with individual data records (136). This group recently assessed the association of major lipids and apolipoproteins with vascular risk based on 68 of these prospective studies, involving more than 300,000 participants with information on lipid profile and conventional risk factors (9). The hazard ratio for coronary heart disease with triglycerides was 1.37 (95% CI, 1.31–1.42) after adjustment for nonlipid factors, but was reduced to 0.99 (95% CI, 0.94–1.05) after adjustment for HDL cholesterol and non-HDL cholesterol (9).

There may be a nonlinear relationship between triglyceride levels and CVD (patients with the chylomicronemia syndrome are not always characterized by premature CVD) (82). It could be due to enrichment of these populations with patients with FHTG who are not at increased risk for premature CVD. Very large triglyceride-rich lipoproteins, perhaps equivalent to unmetabolized chylomicrons, are not atherogenic if they are too large to penetrate into the arterial wall (60, 137). Although there are a number of severely hypertriglyceridemic animals, especially genetically modified mice, these animals develop only early-stage lesions (138, 139). In contrast, there is no doubt that metabolic products of triglyceride-rich lipoproteins are atherogenic (140142). Zilversmit (143) first postulated that atherosclerosis developed in part due to arterial infiltration of chylomicron remnants locally produced by arterial wall LpL. A variety of animal models produced using dietary or genetic manipulations have confirmed that chylomicron remnants are atherogenic (139, 144). Remnants have been identified within human atherosclerosis plaques (145). Unlike VLDL and LDL that contain full-length apoB-100, chylomicrons contain a truncated apoB, apoB-48. apoB-48 lipoproteins are clearly atherogenic as shown in a mouse model constructed with apoB-48 as the only type of apoB in VLDL and LDL (146).

Although it is cholesterol and not triglyceride that is the pathological signature of atherosclerosis and that accumulates both intracellularly in foam cells and extracellularly within the plaque, lipolysis of triglyceride-rich lipoproteins also produces fatty acids, lysolecithin, and other reactive lipids. In vitro studies have implicated these lipids (and in some experiments lipolysis of triglyceride) in inflammation (147), expression of adhesion molecules (148), and promotion of coagulation (149). In addition, ex vivo studies have shown that lipolysis leads to increased permeability of blood vessels (150), which may allow greater infiltration of LDL. Nonetheless, because of the lack of clear human data showing that reductions of triglyceride reduce CVD, the Task Force views hypertriglyceridemia as a marker for risk in some individuals.

3.0. Management of hypertriglyceridemia

Recommendation

3.1. We recommend lifestyle therapy, including dietary counseling to achieve appropriate diet composition, physical activity, and a program to achieve weight reduction in overweight and obese individuals as the initial treatment of mild-to-moderate hypertriglyceridemia (1/⊕⊕○○).

3.1. Evidence

Diet

Much of the increase in serum triglycerides that occurs in adult life is caused by weight gain, lack of exercise, and a diet rich in simple carbohydrates and sugar-sweetened beverages. This may also underlie hypertriglyceridemic situations in younger ages (151). As regards diet quality, in the weight-stable condition, even in overweight or obese people, reduced carbohydrate intake and increased fat intake lower fasting triglycerides. There is a quantitative linear relation between replacement of dietary carbohydrate with fat and reduction in serum triglycerides (152). Saturated, monounsaturated, and n-6 polyunsaturated fatty acids all lower serum triglycerides when they replace carbohydrate, with no clear difference between fatty acid classes in this action. However, diet affects other cardiovascular risk factors beyond triglycerides, and such effects need to be taken into account. Thus, there is a large body of evidence clearly indicating that both dietary saturated fat and trans fatty acids increase LDL cholesterol levels. Replacing these atherogenic fatty acids with monounsaturated or polyunsaturated fat lowers LDL cholesterol; n-6 polyunsaturated fats have a stronger LDL-lowering effect than monounsaturated fats (152). n-3 Polyunsaturated fat lowers serum triglycerides uniquely among the fatty acids, as discussed in the section on drug treatment below.

The type of carbohydrate may affect serum triglycerides. Fructose as contained in sweetened beverages may have stronger triglyceride-raising effects than glucose. This point of view, that fructose as a component of sugar-sweetened beverages, is more detrimental than sucrose or glucose is controversial, however, and more information is needed from randomized comparative trials. Nonetheless, it is recommended that reduced intake of sugar-sweetened beverages, whether composed mainly of high-fructose corn syrup or sucrose, is an important part of lowering serum triglycerides (151). Some carbohydrate-rich foods such as potatoes, white bread, and rice increase blood glucose more quickly and to a higher concentration than other carbohydrate-rich foods such as apples, legumes, nuts, pasta, and densely-baked whole grain breads. This difference is expressed by the glycemic index, which is the rise in blood glucose of 50 g of carbohydrate in a specific food compared with either 50 g glucose or white bread (153). The glycemic index may correlate with the extent of rise of serum triglyceride after eating carbohydrate-rich foods (154).

There has been much less attention paid to the effects of dietary protein on serum triglycerides. Low-carbohydrate diets are mainly high in fat, but protein content usually increases as well. Thus, the triglyceride-lowering effects of low-carbohydrate diets may be partly caused by protein. The OmniHeart study compared the effects of healthful dietary patterns based on the DASH diet that lowered blood pressure and LDL cholesterol. These dietary patterns all emphasize fruits, vegetables, and low-fat dairy products; include whole grains, poultry, fish, and nuts; use unsaturated vegetable oils; and contain smaller amounts of red meat, sweets, and sugar-containing beverages than typical diets in the United States (155, 156). Compared with a diet that emphasized carbohydrate, a similar diet that emphasized protein decreased triglyceride levels further, and this decrease was about twice the effect of a diet that emphasized unsaturated fat (155).

African-Americans have lower serum triglyceride levels than other racial or ethnic groups. The OmniHeart study, in which 50% of the population was African-American, found that diet modification had less effect on triglyceride levels in this ethnic group than in a Caucasian population when matching baseline triglyceride levels (157). Further studies addressing variability across population subgroups are needed.

Exercise

It has been reported that exercise the day before ingestion of a high-fat meal is associated with a marked dampening of the postprandial triglyceride increase. The mechanisms for this are not clear, and the exercise benefits are relatively short-lived. The minimum exercise required to reduce a postprandial triglyceride increase has not been determined, but a period of 30–60 min of intermittent aerobic exercise or mild resistance exercise has been shown to be effective in lowering plasma and VLDL triglycerides. These findings suggest a benefit from an active lifestyle that does not require intense or prolonged exercise (158, 159). A recent meta-analysis comparing aerobic exercise programs showed favorable effects only for high-intensity programs. The most frequently observed alteration was an increase in the HDL cholesterol, whereas reductions in triglycerides, total cholesterol, and LDL cholesterol appeared less often (160). Furthermore, in a recent large, community-based study, a combination of aerobic and resistance exercise was associated with lower triglyceride levels in men as compared with aerobic exercise alone (161).

Treatment of excess weight is critical to reduce triglyceride levels. The macronutrient composition of a weight-loss diet is considerably less important for lowering triglycerides than the amount of weight lost. Two recent large-scale clinical trials of 2-yr duration did not find differences in effects on triglyceride levels between low-fat, high-carbohydrate diets and low-carbohydrate diets (162, 163). Many studies have shown that ongoing counseling by dietitians and behavioral therapists, as well as support from peers, is important to most people who are successful in losing weight and maintaining weight loss.

Recommendation

3.2. For severe and very severe hypertriglyceridemia (>1000 mg/dl), we recommend combining reduction of dietary fat and simple carbohydrate intake with drug treatment to reduce the risk of pancreatitis (1/⊕⊕⊕⊕).

3.2. Evidence

Determination and treatment of underlying causes of very severe hypertriglyceridemia should be considered first. Restriction of both saturated and unsaturated dietary fat, particularly at initiation of therapy and in LpL deficiency, assists in lowering triglyceride acutely. Design of the dietary intervention may benefit from input from nutrition specialists. Regain of weight loss might exacerbate pancreatitis risk (82).

Recommendation

3.3. We recommend that the treatment goal for patients with moderate hypertriglyceridemia be a non-HDL cholesterol level in agreement with NCEP ATP guidelines (1/⊕⊕○○).

3.3. Evidence

Non-HDL cholesterol (total serum cholesterol minus HDL cholesterol) reflects the amount of cholesterol in all atherogenic lipoprotein particles. In the case of an increase in VLDL, commonly seen in hypertriglyceridemia, measurement of LDL cholesterol alone would underestimate the risk associated with atherogenic lipoproteins. Therefore, measurement of non-HDL cholesterol is recommended in subjects with hypertriglyceridemia both for risk stratification and as a secondary target for therapy (3, 164). Alternatively, the blood level of atherogenic lipoprotein particles can be assessed by measuring the concentration of apoB. Not unexpectedly, there is a good correlation between apoB and non-HDL cholesterol because one apoB molecule is present on the surface of each chylomicron, VLDL, IDL, LDL, and Lp(a) particle and resides with the particle during its metabolism in the plasma compartment. Therefore, the apoB concentration reflects the concentration of atherogenic lipoprotein particles. Because measurement of apoB is helpful in the differentiation of FCHL from FHTG, apoB levels may be measured during an initial evaluation of a hypertriglyceridemic patient. Non-HDL cholesterol can then be followed as the therapeutic target.

Recommendations

3.4. We recommend that a fibrate be used as a first-line agent for reduction of triglycerides in patients at risk for triglyceride-induced pancreatitis (1/⊕⊕⊕○).

3.5. We suggest that three drug classes (fibrates, niacin, n-3 fatty acids) alone or in combination with statins be considered as treatment options in patients with moderate to severe triglyceride levels (2/⊕⊕○○).

3.4.–3.5. Evidence

Three drug classes are clinically available for treatment of hypertriglyceridemia—fibrates, niacin, and n-3 fatty acids. Each of these classes has limitations. There is inconsistency in the evidence base for cardiovascular risk reduction using fibrates, the use of niacin is associated with significant side effects, and there are limited data on the use of n-3 fatty acids to reduce cardiovascular risk. It is uncertain whether we should treat moderate hypertriglyceridemia or other lipoprotein abnormalities associated with this degree of hypertriglyceridemia. If the primary goal is to lower triglyceride levels, fibrates and perhaps n-3 fatty acids are best. If the primary goal is to modify the size and density of LDL and HDL particles, niacin is best.

Fibrates

Fibrates should be strongly considered in patients with severe and very severe hypertriglyceridemia and should be considered in patients with moderate hypertriglyceridemia. Fibrates decrease triglyceride levels by 30–50% and sometimes increase HDL cholesterol (165168). In patients with high triglyceride levels, LDL cholesterol levels may increase, whereas in mild hypertriglyceridemia, LDL cholesterol levels may decrease. In patients with triglyceride-induced pancreatitis, treatment of underlying causes and concomitant fibrate therapy to maintain triglyceride levels below 2000 mg/dl is beneficial to prevent recurrent disease (82). Due to a large excursion of triglyceride levels in the setting of very severe hypertriglyceridemia, a treatment goal of less than 1000 mg/dl is recommended. Below this level, the main effort should be directed toward prevention of premature atherosclerosis. We do not recommend the use of heparin infusions or plasmapheresis in the treatment of very severe hypertriglyceridemia with pancreatitis. The treatment of underlying causes including dietary fat restriction and use of long-term fibrate therapy should suffice (82).

Studies to date have not demonstrated an overall benefit of fibrates for reduction of cardiovascular or total mortality (165169). An a priori analysis demonstrated that a decrease in triglyceride and elevation of HDL cholesterol levels was associated with a decrease in primary events, while at the same time, there was an increase in death in females (169). Post hoc subgroup analyses of all of these trials show that the use of fibrates in patients with moderate hypertriglyceridemia results in a decrease in composite cardiovascular events, but not a decrease in mortality (170172). However, these studies also indicate that treatment of patients with triglycerides below 200 mg/dl does not confer benefit.

Fibrates increase fatty acid oxidation, increase LpL synthesis, and reduce expression of apoC-III, all of which decrease VLDL triglyceride production and increase LpL-mediated catabolism of triglyceride-rich lipoproteins (170). Side effects include gastrointestinal discomfort and possibly an increased incidence of cholesterol gallstones. Fibrates are contraindicated in patients with liver and gall bladder disease. Fibric acid derivatives should be used with great caution in the setting of renal insufficiency because the drugs are excreted in the urine and may reversibly increase serum creatinine levels—especially fenofibrate, although the significance of this effect is unknown. Fenofibrate, which does not interfere with statin metabolism and has a lower risk of causing myopathy, is the preferred fibrate to use in combination with a statin. Due to effects on protein binding, there is a potential interaction with warfarin requiring careful monitoring. Gemfibrozil can be considered in very severe hypertriglyceridemia beginning in the second trimester in pregnant women who are at risk of pancreatitis (86).

Niacin

Clinical trials using niacin, alone or in combination with other lipid medications, have shown benefits in decreasing cardiovascular event rates and atherosclerosis (173176). However, the recently concluded AIM-HIGH study did not report any further benefits with regard to cardiovascular events when niacin was added to a statin in patients with median triglyceride level of about 160 mg/dl in the mild hypertriglyceridemia range and average LDL cholesterol levels below 80 mg/dl (177). At doses of 500-2000 mg/d, niacin lowers triglycerides by 10–30%, increases HDL cholesterol by 10–40%, and lowers LDL cholesterol by 5–20%. Although higher doses of immediate-release (crystalline) niacin have been used, the maximum dose of the prescription extended-release formulation is 2000 mg/d; doses this high are reached by increasing the dose slowly over time. Niacin contributes to the release of prostaglandin D2 from cells in the skin leading to vasodilatation. The most common side effect is cutaneous flushing, which is most significant with the first few doses. It occurs 15 to 60 min after ingestion and typically lasts 15 to 30 min. Ingestion after a meal and administration of uncoated aspirin before the meal minimizes flushing. The most serious complication of niacin therapy is hepatotoxicity (which is dose dependent), and therapy should be accompanied by monitoring of liver function tests (178). Other side effects of niacin therapy include impairment or worsening of glucose tolerance and hyperuricemia. Niacin can be used safely in patients with glucose intolerance and can be considered in diabetic patients on oral medications or insulin who have moderate to good glycemic control but could cause conversion of borderline glucose intolerance to meet diabetes criteria in some patients. Niacin can increase blood levels of uric acid by blocking its excretion and can precipitate or worsen gout unless the patient is treated with allopurinol. Niacin is contraindicated in patients with active peptic ulcer disease.

n-3 Fatty acids

The long-chain marine omega-3 fatty acids [eicosapentaenoic acid, C20:5n-3 (EPA) and docosahexaenoic acid, C22:6n-3 (DHA)] lower fasting and postprandial triglyceride levels in a dose-dependent fashion. Approximately 3 to 4 g/d of EPA plus DHA are necessary to reduce hypertriglyceridemia by 20–50% (179). HDL cholesterol is mildly increased by about 5%. With reductions of triglyceride levels, there can be increased levels of LDL cholesterol due to increased conversion of VLDL to LDL. To date, no studies using high-dose n-3 fatty acids in hypertriglyceridemia patients have shown a beneficial cardiovascular outcome. EPA added to statin therapy in subjects with cholesterol levels above 250 mg/dl in an open-label study resulted in a 19% relative reduction in major coronary events (180). Omega-3 fatty acids (e.g. Lovaza) may be considered for treatment of triglyceride levels above 1000 mg/dl. Over-the-counter preparations of omega-3 fatty acids have variable quantities of EPA and DHA ranging from 20–50%, depending on products. The nutrition labels must be studied to calculate the number of capsules required to obtain a dose of 3–5 g of n-3 fatty acids. Omega-3 acid ethyl esters are available by prescription in capsules that contain 80% EPA and DHA. Thus, a dose of four capsules is needed to lower triglycerides by 30–50% (181). Side effects with large doses of omega-3 fatty acids include fishy taste and burping. Beyond an impact of omega-3 fatty acid supplements on triglyceride levels, intake of diets rich in n-3 fatty acids have resulted in positive outcomes with regard to cardiovascular disease (182184).

Recommendation

3.6. We recommend that statins not be used as monotherapy for severe or very severe hypertriglyceridemia. However, statins may be useful for the treatment of moderate hypertriglyceridemia when indicated to modify cardiovascular risk (1/⊕⊕○○).

3.6. Evidence

Hydroxymethylglutaryl coenzyme A reductase inhibitors, or statins, have a modest triglyceride-lowering effect, typically about 10–15%, which is dose-dependent. High doses of statins that have strong efficacy, such as atorvastatin 80 mg or rosuvastatin 40 mg, can lower plasma triglyceride by 25–30%. Statin monotherapy should not be first-line therapy to reduce triglyceride levels in patients with severe or very severe hypertriglyceridemia (>1000 mg/dl). Addition of statins can be considered to reduce cardiovascular risk in patients with mild-to-moderate hypertriglyceridemia (>150 mg/dl and <1000 mg/dl) and elevated non-HDL cholesterol. Side effects of statins occur in about 5–10% of patients. Muscle symptoms ranging from leg cramps to aching to weakness occur in about 10% of patients, whereas rhabdomyolysis is rare (185). Conditions predisposing to severe myopathy include advanced age, renal failure, polypharmacy, and acute illness.

Combination therapy and other drugs

Because the four drug classes described above (fibrates, niacin, n-3 fatty acids, and statins) have different underlying mechanisms in reducing triglyceride levels, as well as correcting other associated dyslipidemias, there is a considerable potential for use of drug combinations based on complementary mechanisms (169, 186). Examples include a combination of niacin and statins or fibrates and statins. Attention needs to be paid to potential drug-drug interaction. To minimize such risks, combination treatment should be initiated carefully, and advice should be sought from clinicians familiar with these types of interactions.

Pioglitazone, a peroxisome proliferator-activated receptor-γ agonist, has a mild triglyceride-lowering effect, whereas rosiglitazone does not lower triglycerides (188, 189). Side effects reported include weight gain and risk of heart failure and macular edema (190). Of note, some data suggest an association between the use of pioglitazone and an increased risk of bladder cancer (191). Orlistat, an inhibitor of intestinal lipase that is used as a weight-loss drug, can lower postprandial triglyceride levels. It is a pharmacological method to reduce fat absorption, which may be helpful in patients with fasting hyperchylomicronemia (187). Orlistat has been used in combination with fibrates with additive effects. Side effects include bloating, diarrhea, incontinence, and oily leakage and are related to the amount of fat ingested in the diet. Furthermore, cases of severe liver injury have been reported to occur rarely with the use of orlistat.

Acknowledgments

Cosponsoring Associations: American Heart Association, European Society of Endocrinology.

Footnotes

Abbreviations:
apoB
Apolipoprotein B
CI
confidence interval
CVD
cardiovascular disease
DHA
docosahexaenoic acid
EPA
eicosapentaenoic acid
FCHL
familial combined hyperlipidemia
FHA
familial hypoalphalipoproteinemia
FHTG
familial hypertriglyceridemia
HDL
high-density lipoprotein
IDL
intermediate-density lipoprotein
LDL
low-density lipoprotein
Lp(a)
lipoprotein(a)
LpL
lipoprotein lipase
NEFA
nonesterified fatty acid
VLDL
very low-density lipoprotein.

Financial Disclosure of Task Force

Lars Berglund, M.D., Ph.D. (chair)—Financial or Business/Organizational Interests: NIH, AHA, Pfizer, Astra-Zeneca, Danone; Significant Financial Interest or Leadership Position: AHA, Pfizer, Astra-Zeneca. John D. Brunzell, M.D.—Financial or Business/Organizational Interests: American Diabetes Association, Endocrine Society, Amsterdam Molecular Therapeutics, Lexicon Genetics, Merck Research Laboratories, Daiichi Sankyo, University of Washington; Significant Financial Interest or Leadership Position: none declared. Anne Carol Goldberg, M.D., F.A.C.P.—Financial or Business/Organizational Interests: AHA, NIH/NHLBI, National Lipid Association, Foundation of the National Lipid Association, Washington University, Merck, Sanofi-Aventis, Roche, ISIS/Genzyme; Significant Financial Interest or Leadership Position: National Lipid Association, Foundation of the National Lipid Association. Ira Jay Goldberg, M.D.—Financial or Business/Organizational Interests: American Heart Association, American Diabetes Association, Astra-Zeneca, GSK, Merck, Medcom, Amylin, Schering; Significant Financial Interest or Leadership Position: none declared. Frank Sacks, M.D.—Financial or Business/Organizational Interests: NIH/NHLBI, R3i Foundation, Metabolic Syndrome Institute, American Society of Nutrition, Sherborne-Gibbs, WebMD, Abbott, AstraZeneca, Aegerlan, Pfizer, Lilly, Isis, Genzyme, Roche, Genentech; Significant Financial Interest or Leadership Position: R3i Foundation, Metabolic Syndrome Institute, American Society of Nutrition. Mohammad Hassan Murad, M.D.*—Financial or Business/Organizational Interests: KER Unit (Mayo Clinic); Significant Financial Interest or Leadership Position: none declared. Anton F. H. Stalenhoef, M.D., Ph.D., F.R.C.P.—Financial or Business/Organizational Interests: Dutch Heart Foundation, Merck, Genzyme, Roche; Significant Financial Interest or Leadership Position: Dutch Journal of Medicine, Current Opinion Lipidology.

* Evidence-based reviews for this guideline were prepared under contract with The Endocrine Society.

References

  • 1. Atkins D, Best D, Briss PA, Eccles M, Falck-Ytter Y, Flottorp S, Guyatt GH, Harbour RT, Haugh MC, Henry D, Hill S, Jaeschke R, Leng G, Liberati A, Magrini N, Mason J, Middleton P, Mrukowicz J, O'Connell D, Oxman AD, Phillips B, Schünemann HJ, Edejer TT, Varonen H, Vist GE, Williams JW, Jr, Zaza S. 2004. Grading quality of evidence and strength of recommendations. BMJ 328:1490–1497 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Swiglo BA, Murad MH, Schünemann HJ, Kunz R, Vigersky RA, Guyatt GH, Montori VM. 2008. A case for clarity, consistency, and helpfulness: state-of-the-art clinical practice guidelines in endocrinology using the grading of recommendations, assessment, development, and evaluation system. J Clin Endocrinol Metab 93:666–673 [DOI] [PubMed] [Google Scholar]
  • 3. 2002. Third Report of the National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults (Adult Treatment Panel III) final report. Circulation 106:3143–3421 [PubMed] [Google Scholar]
  • 4. Heiss G, Tamir I, Davis CE, Tyroler HA, Rifkand BM, Schonfeld G, Jacobs D, Frantz ID., Jr 1980. Lipoprotein-cholesterol distributions in selected North American populations: the lipid research clinics program prevalence study. Circulation 61:302–315 [DOI] [PubMed] [Google Scholar]
  • 5. Ford ES, Li C, Zhao G, Pearson WS, Mokdad AH. 2009. Hypertriglyceridemia and its pharmacologic treatment among US adults. Arch Intern Med 169:572–578 [DOI] [PubMed] [Google Scholar]
  • 6. Lewis B, Chait A, Wootton ID, Oakley CM, Krikler DM, Sigurdsson G, February A, Maurer B, Birkhead J. 1974. Frequency of risk factors for ischaemic heart-disease in a healthy British population, with particular reference to serum-lipoprotein levels. Lancet 1:141–146 [DOI] [PubMed] [Google Scholar]
  • 7. Assmann G, Schulte H. 1993. Results and conclusions of the prospective cardiovascular Munster (PROCAM) study. In: Assmann G, ed. Lipid metabolism disorders and coronary heart disease. Munich: MMV Medizin Verlag GmbH; 19–68 [Google Scholar]
  • 8. Pang RW, Tam S, Janus ED, Siu ST, Ma OC, Lam TH, Lam KS. 2006. Plasma lipid, lipoprotein and apolipoprotein levels in a random population sample of 2875 Hong Kong Chinese adults and their implications (NCEP ATP-III, 2001 guidelines) on cardiovascular risk assessment. Atherosclerosis 184:438–445 [DOI] [PubMed] [Google Scholar]
  • 9. Di Angelantonio E, Sarwar N, Perry P, Kaptoge S, Ray KK, Thompson A, Wood AM, Lewington S, Sattar N, Packard CJ, Collins R, Thompson SG, Danesh J. 2009. Major lipids, apolipoproteins, and risk of vascular disease. JAMA 302:1993–2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Brunzell JD, Schrott HG. 1973. The interaction of familial and secondary causes of hypertriglyceridemia: role in pancreatitis. Trans Assoc Am Physicians 86:245–254 [PubMed] [Google Scholar]
  • 11. Chait A, Brunzell JD. 1983. Severe hypertriglyceridemia: role of familial and acquired disorders. Metabolism 32:209–214 [DOI] [PubMed] [Google Scholar]
  • 12. Centers for Disease Control and Prevention 2010. Prevalence of abnormal lipid levels among youths—United States, 1999–2006. MMWR Morb Mortal Wkly Rep 59:29–33 [PubMed] [Google Scholar]
  • 13. Müller-Riemenschneider F, Nocon M, Willich SN. 2010. Prevalence of modifiable cardiovascular risk factors in German adolescents. Eur J Cardiovasc Prev Rehabil 17:204–210 [DOI] [PubMed] [Google Scholar]
  • 14. Murad MH, Hazem A, Coto-Yglesias F, Dzyubak S, Gupta S, Bancos I, Lane M, Erwin PJ, Berglund L, Elraiya T, Montori VM. 2012. The association of hypertriglyceridemia with cardiovascular events and pancreatitis: a systematic review and meta-analysis. BMC Endocr Disord 12:2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Nordestgaard BG, Benn M, Schnohr P, Tybjaerg-Hansen A. 2007. Nonfasting triglycerides and risk of myocardial infarction, ischemic heart disease, and death in men and women. JAMA 298:299–308 [DOI] [PubMed] [Google Scholar]
  • 16. Bansal S, Buring JE, Rifai N, Mora S, Sacks FM, Ridker PM. 2007. Fasting compared with nonfasting triglycerides and risk of cardiovascular events in women. JAMA 298:309–316 [DOI] [PubMed] [Google Scholar]
  • 17. Mora S, Rifai N, Buring JE, Ridker PM. 2008. Fasting compared with nonfasting lipids and apolipoproteins for predicting incident cardiovascular events. Circulation 118:993–1001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Stalenhoef AF, de Graaf J. 2008. Association of fasting and nonfasting serum triglycerides with cardiovascular disease and the role of remnant-like lipoproteins and small, dense LDL. Curr Opin Lipidol 19:355–361 [DOI] [PubMed] [Google Scholar]
  • 19. Patsch JR, Miesenböck G, Hopferwieser T, Mühlberger V, Knapp E, Dunn JK, Gotto AM, Jr, Patsch W. 1992. Relation of triglyceride metabolism and coronary artery disease. Studies in the postprandial state. Arterioscler Thromb 12:1336–1345 [DOI] [PubMed] [Google Scholar]
  • 20. Karpe F. 1999. Postprandial lipoprotein metabolism and atherosclerosis. J Intern Med 246:341–355 [DOI] [PubMed] [Google Scholar]
  • 21. Eberly LE, Stamler J, Neaton JD. 2003. Relation of triglyceride levels, fasting and nonfasting, to fatal and nonfatal coronary heart disease. Arch Intern Med 163:1077–1083 [DOI] [PubMed] [Google Scholar]
  • 22. Ridker PM. 2008. Fasting versus nonfasting triglycerides and the prediction of cardiovascular risk: do we need to revisit the oral triglyceride tolerance test? Clin Chem 54:11–13 [DOI] [PubMed] [Google Scholar]
  • 23. Warnick GR, Nakajima K. 2008. Fasting versus nonfasting triglycerides: implications for laboratory measurements. Clin Chem 54:14–16 [DOI] [PubMed] [Google Scholar]
  • 24. Ginsberg HN. 2002. New perspectives on atherogenesis: role of abnormal triglyceride-rich lipoprotein metabolism. Circulation 106:2137–2142 [DOI] [PubMed] [Google Scholar]
  • 25. Ayyobi AF, McGladdery SH, McNeely MJ, Austin MA, Motulsky AG, Brunzell JD. 2003. Small, dense LDL and elevated apolipoprotein B are the common characteristics for the three major lipid phenotypes of familial combined hyperlipidemia. Arterioscler Thromb Vasc Biol 23:1289–1294 [DOI] [PubMed] [Google Scholar]
  • 26. Carr MC, Brunzell JD. 2004. Abdominal obesity and dyslipidemia in the metabolic syndrome: importance of type 2 diabetes and familial combined hyperlipidemia in coronary artery disease risk. J Clin Endocrinol Metab 89:2601–2607 [DOI] [PubMed] [Google Scholar]
  • 27. Mann CJ, Yen FT, Grant AM, Bihain BE. 1991. Mechanism of plasma cholesteryl ester transfer in hypertriglyceridemia. J Clin Invest 88:2059–2066 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Zambon A, Hokanson JE, Brown BG, Brunzell JD. 1999. Evidence for a new pathophysiological mechanism for coronary artery regression: hepatic lipase-mediated changes in LDL density. Circulation 99:1959–1964 [DOI] [PubMed] [Google Scholar]
  • 29. Sniderman AD, Wolfson C, Teng B, Franklin FA, Bachorik PS, Kwiterovich PO., Jr 1982. Association of hyperapobetalipoproteinemia with endogenous hypertriglyceridemia and atherosclerosis. Ann Intern Med 97:833–839 [DOI] [PubMed] [Google Scholar]
  • 30. Austin MA, King MC, Vranizan KM, Krauss RM. 1990. Atherogenic lipoprotein phenotype. A proposed genetic marker for coronary heart disease risk. Circulation 82:495–506 [DOI] [PubMed] [Google Scholar]
  • 31. St-Pierre AC, Cantin B, Dagenais GR, Mauriège P, Bernard PM, Després JP, Lamarche B. 2005. Low-density lipoprotein subfractions and the long-term risk of ischemic heart disease in men: 13-year follow-up data from the Québec Cardiovascular Study. Arterioscler Thromb Vasc Biol 25:553–559 [DOI] [PubMed] [Google Scholar]
  • 32. Mora S, Szklo M, Otvos JD, Greenland P, Psaty BM, Goff DC, Jr, O'Leary DH, Saad MF, Tsai MY, Sharrett AR. 2007. LDL particle subclasses, LDL particle size, and carotid atherosclerosis in the Multi-Ethnic Study of Atherosclerosis (MESA). Atherosclerosis 192:211–217 [DOI] [PubMed] [Google Scholar]
  • 33. Musunuru K, Orho-Melander M, Caulfield MP, Li S, Salameh WA, Reitz RE, Berglund G, Hedblad B, Engström G, Williams PT, Kathiresan S, Melander O, Krauss RM. 2009. Ion mobility analysis of lipoprotein subfractions identifies three independent axes of cardiovascular risk. Arterioscler Thromb Vasc Biol 29:1975–1980 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Baigent C, Keech A, Kearney PM, Blackwell L, Buck G, Pollicino C, Kirby A, Sourjina T, Peto R, Collins R, Simes R. 2005. Efficacy and safety of cholesterol-lowering treatment: prospective meta-analysis of data from 90,056 participants in 14 randomised trials of statins. Lancet 366:1267–1278 [DOI] [PubMed] [Google Scholar]
  • 35. Sacks FM, Campos H. 2003. Low density lipoprotein and cardiovascular disease: a reappraisal. (Clinical Review 163). J Clin Endocrinol Metab 88:4525–4532 [DOI] [PubMed] [Google Scholar]
  • 36. Mora S. 2009. Advanced lipoprotein testing and subfractionation are not (yet) ready for routine clinical use. Circulation 119:2396–2404 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Ip S, Lichtenstein AH, Chung M, Lau J, Balk EM. 2009. Systematic review: association of low-density lipoprotein subfractions with cardiovascular outcomes. Ann Intern Med 150:474–484 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Mora S, Otvos JD, Rifai N, Rosenson RS, Buring JE, Ridker PM. 2009. Lipoprotein particle profiles by nuclear magnetic resonance compared with standard lipids and apolipoproteins in predicting incident cardiovascular disease in women. Circulation 119:931–939 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Campos H, Moye LA, Glasser SP, Stampfer MJ, Sacks FM. 2001. Low density lipoprotein size, pravastatin treatment, and coronary events. JAMA 286:1468–1474 [DOI] [PubMed] [Google Scholar]
  • 40. de Graaf J, Couture P, Sniderman A. 2008. A diagnostic algorithm for the atherogenic apolipoprotein B dyslipoproteinemias. Nat Clin Pract Endocrinol Metab 4:608–618 [DOI] [PubMed] [Google Scholar]
  • 41. Holewijn S, den Heijer M, Swinkels DW, Stalenhoef AF, de Graaf J. 2010. Apolipoprotein B, non-HDL cholesterol and LDL cholesterol for identifying individuals at increased cardiovascular risk. J Intern Med 268:567–577 [DOI] [PubMed] [Google Scholar]
  • 42. Utermann G. 1989. The mysteries of lipoprotein(a). Science 246:904–910 [DOI] [PubMed] [Google Scholar]
  • 43. Bennet A, Di Angelantonio E, Erqou S, Eiriksdottir G, Sigurdsson G, Woodward M, Rumley A, Lowe GD, Danesh J, Gudnason V. 2008. Lipoprotein (a) levels and risk of future coronary heart disease: large-scale prospective data. Arch Intern Med 168:598–608 [DOI] [PubMed] [Google Scholar]
  • 44. Erqou S, Kaptoge S, Perry PL, Di Angelantonio E, Thompson A, White IR, Marcovina SM, Collins R, Thompson SG, Danesh J. 2009. Lipoprotein(a) concentration and the risk of coronary heart disease, stroke, and nonvascular mortality. JAMA 302:412–423 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Kamstrup PR, Tybjaerg-Hansen A, Steffensen R, Nordestgaard BG. 2009. Genetically elevated lipoprotein (a) and increased risk for myocardial infarction. JAMA 301:2331–2339 [DOI] [PubMed] [Google Scholar]
  • 46. Anuurad E, Enkhmaa B, Berglund L. 2010. Enigmatic role of lipoprotein(a) in cardiovascular disease. Clin Transl Sci 3:327–332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Gylling H, Vega GL, Grundy SM. 1992. Physiologic mechanisms for reduced apolipoprotein A-I concentrations associated with low HDL cholesterol in patients with normal plasma lipids. J Lipid Res 33:1527–1539 [PubMed] [Google Scholar]
  • 48. Brinton EA, Eisenberg S, Breslow JL. 1994. Human HDL cholesterol levels are determined by apoA-I fractional catabolic rate, which correlates inversely with estimates of HDL particle size. Arterioscler Thromb 14:707–720 [DOI] [PubMed] [Google Scholar]
  • 49. Deeb SS, Zambon A, Carr MC, Ayyobi AF, Brunzell JD. 2003. Hepatic lipase and dyslipidemia: interactions among genetic variants, obesity, gender and diet. J Lipid Res 44:1279–1286 [DOI] [PubMed] [Google Scholar]
  • 50. Stampfer MJ, Sacks FM, Salvini S, Willett WC, Hennekens CH. 1991. A prospective study of cholesterol, apolipoproteins, and the risk of myocardial infarction. N Engl J Med 325:373–381 [DOI] [PubMed] [Google Scholar]
  • 51. Sweetnam PM, Bolton CH, Yarnell JW, Bainton D, Baker IA, Elwood PC, Miller NE. 1994. Associations of the HDL2 and HDL3 cholesterol subfractions with the development of ischemic heart disease in British men. The Caerphilly and Speedwell Collaborative Heart Disease Studies. Circulation 90:769–774 [DOI] [PubMed] [Google Scholar]
  • 52. Lamarche B, Tchernof A, Moorjani S, Cantin B, Dagenais GR, Lupien PJ, Després JP. 1997. Small, dense low-density lipoprotein particles as a predictor of the risk of ischemic heart disease in men. Prospective results from the Quebec Cardiovascular Study. Circulation 95:69–75 [DOI] [PubMed] [Google Scholar]
  • 53. Asztalos BF, Cupples LA, Demissie S, Horvath KV, Cox CE, Batista MC, Schaefer EJ. 2004. High-density lipoprotein subpopulation profile and coronary heart disease prevalence in male participants of the Framingham Offspring Study. Arterioscler Thromb Vasc Biol 24:2181–2187 [DOI] [PubMed] [Google Scholar]
  • 54. Asztalos BF, Collins D, Horvath KV, Bloomfield HE, Robins SJ, Schaefer EJ. 2008. Relation of gemfibrozil treatment and high-density lipoprotein subpopulation profile with cardiovascular events in the Veterans Affairs High-Density Lipoprotein Intervention Trial. Metabolism 57:77–83 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. van der Steeg WA, Holme I, Boekholdt SM, Larsen ML, Lindahl C, Stroes ES, Tikkanen MJ, Wareham NJ, Faergeman O, Olsson AG, Pedersen TR, Khaw KT, Kastelein JJ. 2008. High-density lipoprotein cholesterol, high-density lipoprotein particle size, and apolipoprotein A-I: significance for cardiovascular risk: the IDEAL and EPIC-Norfolk studies. J Am Coll Cardiol 51:634–642 [DOI] [PubMed] [Google Scholar]
  • 56. Cohn JS, McNamara JR, Cohn SD, Ordovas JM, Schaefer EJ. 1988. Plasma apolipoprotein changes in the triglyceride-rich lipoprotein fraction of human subjects fed a fat-rich meal. J Lipid Res 29:925–936 [PubMed] [Google Scholar]
  • 57. Karpe F, Bell M, Björkegren J, Hamsten A. 1995. Quantification of postprandial triglyceride-rich lipoproteins in healthy men by retinyl ester labeling and simultaneous measurement of apolipoproteins B-48 and B-100. Arterioscler Thromb Vasc Biol 15:199–207 [DOI] [PubMed] [Google Scholar]
  • 58. Campos H, Khoo C, Sacks FM. 2005. Diurnal and acute patterns of postprandial apolipoprotein B-48 in VLDL, IDL, and LDL from normolipidemic humans. Atherosclerosis 181:345–351 [DOI] [PubMed] [Google Scholar]
  • 59. Nakano T, Tanaka A, Okazaki M, Tokita Y, Nagamine T, Nakajima K. 2011. Particle size of apoB-48 carrying lipoproteins in remnant lipoproteins isolated from postprandial plasma. Ann Clin Biochem 48:57–64 [DOI] [PubMed] [Google Scholar]
  • 60. Tabas I, Williams KJ, Borén J. 2007. Subendothelial lipoprotein retention as the initiating process in atherosclerosis: update and therapeutic implications. Circulation 116:1832–1844 [DOI] [PubMed] [Google Scholar]
  • 61. Kannel WB, Vasan RS. 2009. Triglycerides as vascular risk factors: new epidemiologic insights. Curr Opin Cardiol 24:345–350 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Batal R, Tremblay M, Barrett PH, Jacques H, Fredenrich A, Mamer O, Davignon J, Cohn JS. 2000. Plasma kinetics of apoC-III and apoE in normolipidemic and hypertriglyceridemic subjects. J Lipid Res 41:706–718 [PubMed] [Google Scholar]
  • 63. Cohn JS, Patterson BW, Uffelman KD, Davignon J, Steiner G. 2004. Rate of production of plasma and very-low-density lipoprotein (VLDL) apolipoprotein C-III is strongly related to the concentration and level of production of VLDL triglyceride in male subjects with different body weights and levels of insulin sensitivity. J Clin Endocrinol Metab 89:3949–3955 [DOI] [PubMed] [Google Scholar]
  • 64. Zheng C, Khoo C, Ikewaki K, Sacks FM. 2007. Rapid turnover of apolipoprotein C-III-containing triglyceride-rich lipoproteins contributing to the formation of LDL subfractions. J Lipid Res 48:1190–1203 [DOI] [PubMed] [Google Scholar]
  • 65. Ooi EM, Barrett PH, Chan DC, Watts GF. 2008. Apolipoprotein C-III: understanding an emerging cardiovascular risk factor. Clin Sci 114:611–624 [DOI] [PubMed] [Google Scholar]
  • 66. Malmström R, Packard CJ, Caslake M, Bedford D, Stewart P, Yki-Järvinen H, Shepherd J, Taskinen MR. 1997. Defective regulation of triglyceride metabolism by insulin in the liver in NIDDM. Diabetologia 40:454–462 [DOI] [PubMed] [Google Scholar]
  • 67. Kamagate A, Dong HH. 2008. FoxO1 integrates insulin signaling to VLDL production. Cell Cycle 7:3162–3170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Lee SJ, Moye LA, Campos H, Williams GH, Sacks FM. 2003. Hypertriglyceridemia but not diabetes status is associated with VLDL containing apolipoprotein CIII in patients with coronary heart disease. Atherosclerosis 167:293–302 [DOI] [PubMed] [Google Scholar]
  • 69. Cheal KL, Abbasi F, Lamendola C, McLaughlin T, Reaven GM, Ford ES. 2004. Relationship to insulin resistance of the adult treatment panel III diagnostic criteria for identification of the metabolic syndrome. Diabetes 53:1195–1200 [DOI] [PubMed] [Google Scholar]
  • 70. Aarsland A, Chinkes D, Wolfe RR. 1996. Contributions of de novo synthesis of fatty acids to total VLDL-triglyceride secretion during prolonged hyperglycemia/hyperinsulinemia in normal man. J Clin Invest 98:2008–2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. McLaughlin T, Abbasi F, Lamendola C, Yeni-Komshian H, Reaven G. 2000. Carbohydrate-induced hypertriglyceridemia: an insight into the link between plasma insulin and triglyceride concentrations. J Clin Endocrinol Metab 85:3085–3088 [DOI] [PubMed] [Google Scholar]
  • 72. Sumner AE, Finley KB, Genovese DJ, Criqui MH, Boston RC. 2005. Fasting triglyceride and the triglyceride-HDL cholesterol ratio are not markers of insulin resistance in African Americans. Arch Intern Med 165:1395–1400 [DOI] [PubMed] [Google Scholar]
  • 73. Chan DC, Watts GF, Nguyen MN, Barrett PH. 2006. Apolipoproteins C-III and A-V as predictors of very-low-density lipoprotein triglyceride and apolipoprotein B-100 kinetics. Arterioscler Thromb Vasc Biol 26:590–596 [DOI] [PubMed] [Google Scholar]
  • 74. Brunzell JD, Hazzard WR, Porte D, Jr, Bierman EL. 1973. Evidence for a common, saturable triglyceride removal mechanism for chylomicrons and very low density lipoproteins in man. J Clin Invest 52:1578–1585 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75. Gin P, Yin L, Davies BS, Weinstein MM, Ryan RO, Bensadoun A, Fong LG, Young SG, Beigneux AP. 2008. The acidic domain of GPIHBP1 is important for the binding of lipoprotein lipase and chylomicrons. J Biol Chem 283:29554–29562 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Olivecrona G, Ehrenborg E, Semb H, Makoveichuk E, Lindberg A, Hayden MR, Gin P, Davies BS, Weinstein MM, Fong LG, Beigneux AP, Young SG, Olivecrona T, Hernell O. 2010. Mutation of conserved cysteines in the Ly6 domain of GPIHBP1 in familial chylomicronemia. J Lipid Res 51:1535–1545 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Péterfy M, Ben-Zeev O, Mao HZ, Weissglas-Volkov D, Aouizerat BE, Pullinger CR, Frost PH, Kane JP, Malloy MJ, Reue K, Pajukanta P, Doolittle MH. 2007. Mutations in LMF1 cause combined lipase deficiency and severe hypertriglyceridemia. Nat Genet 39:1483–1487 [DOI] [PubMed] [Google Scholar]
  • 78. Brunzell JD. 2007. Clinical practice. Hypertriglyceridemia. N Engl J Med 357:1009–1017 [DOI] [PubMed] [Google Scholar]
  • 79. Zheng C, Khoo C, Furtado J, Sacks FM. 2010. Apolipoprotein C-III and the metabolic basis for hypertriglyceridemia and the dense LDL phenotype. Circulation 121:1722–1734 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80. Havel PJ. 2005. Dietary fructose: implications for dysregulation of energy homeostasis and lipid/carbohydrate metabolism. Nutr Rev 63:133–157 [DOI] [PubMed] [Google Scholar]
  • 81. Lichtenstein AH, Appel LJ, Brands M, Carnethon M, Daniels S, Franch HA, Franklin B, Kris-Etherton P, Harris WS, Howard B, Karanja N, Lefevre M, Rudel L, Sacks F, Van Horn L, Winston M, Wylie-Rosett J. 2006. Summary of American Heart Association Diet and Lifestyle Recommendation revision. Arterioscler Thromb Vasc Biol 26:2186–2191 [DOI] [PubMed] [Google Scholar]
  • 82. Brunzell JD, Deeb SS. 2001. Familial lipoprotein lipase deficiency, ApoC-II deficiency, and hepatic lipase deficiency. In: Scriver CR, Beaudet AL, Sly WS, Valle D, eds. The metabolic basis of inherited disease. 8th ed New York: McGraw-Hill; 2789–2816 [Google Scholar]
  • 83. Brunzell JD, Ayyobi AF. 2003. Dyslipidemia in the metabolic syndrome and type 2 diabetes mellitus. Am J Med 115(Suppl 8A):24S–28S [DOI] [PubMed] [Google Scholar]
  • 84. Tamburrano G, Durante C, Baldelli R. 2002. Therapy of diabetes and dyslipidemia in acromegaly. Pituitary 5:27–31 [DOI] [PubMed] [Google Scholar]
  • 85. Sanderson SL, Iverius PH, Wilson DE. 1991. Successful hyperlipemic pregnancy. JAMA 265:1858–1860 [PubMed] [Google Scholar]
  • 86. Saadi HF, Kurlander DJ, Erkins JM, Hoogwerf BJ. 1999. Severe hypertriglyceridemia and acute pancreatitis during pregnancy: treatment with gemfibrozil. Endocr Pract 5:33–36 [DOI] [PubMed] [Google Scholar]
  • 87. Walsh BW, Schiff I, Rosner B, Greenberg L, Ravnikar V, Sacks FM. 1991. Effects of postmenopausal estrogen replacement on the concentrations and metabolism of plasma lipoproteins. N Engl J Med 325:1196–1204 [DOI] [PubMed] [Google Scholar]
  • 88. Adami S, Rossini M, Zamberlan N, Bertoldo F, Dorizzi R, Lo Cascio V. 1993. Long-term effects of transdermal and oral estrogens on serum lipids and lipoproteins in postmenopausal women. Maturitas 17:191–196 [DOI] [PubMed] [Google Scholar]
  • 89. Mosca L, Harper K, Sarkar S, O'Gorman J, Anderson PW, Cox DA, Barrett-Connor E. 2001. Effect of raloxifene on serum triglycerides in postmenopausal women: influence of predisposing factors for hypertriglyceridemia. Clin Ther 23:1552–1565 [DOI] [PubMed] [Google Scholar]
  • 90. Carr MC, Knopp RH, Brunzell JD, Wheeler BS, Zhu X, Lakshmanan M, Rosen AS, Anderson PW. 2005. Effect of raloxifene on serum triglycerides in women with a history of hypertriglyceridemia while on oral estrogen therapy. Diabetes Care 28:1555–1561 [DOI] [PubMed] [Google Scholar]
  • 91. Morganroth J, Levy RI, Fredrickson DS. 1975. The biochemical, clinical, and genetic features of type III hyperlipoproteinemia. Ann Intern Med 82:158–174 [DOI] [PubMed] [Google Scholar]
  • 92. Lindner MA, Illingworth DR. 1988. Expression of type III hyperlipoproteinemia in an adolescent patient with hypothyroidism. J Pediatr 113:86–89 [DOI] [PubMed] [Google Scholar]
  • 93. Retnakaran R, Connelly PW, Goguen J. 2005. Unmasking of type III hyperlipoproteinemia by hypothyroidism: a dramatic illustration of altered lipoprotein metabolism in a postpartum woman. Endocr Pract 11:394–398 [DOI] [PubMed] [Google Scholar]
  • 94. Stone NJ. 1994. Secondary causes of hyperlipidemia. Med Clin North Am 78:117–141 [DOI] [PubMed] [Google Scholar]
  • 95. Garg A, Agarwal AK. 2009. Lipodystrophies: disorders of adipose tissue biology. Biochim Biophys Acta 1791:507–513 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Simha V, Garg A. 2009. Inherited lipodystrophies and hypertriglyceridemia. Curr Opin Lipidol 20:300–308 [DOI] [PubMed] [Google Scholar]
  • 97. Herbst KL, Tannock LR, Deeb SS, Purnell JQ, Brunzell JD, Chait A. 2003. Kobberling type of familial partial lipodystrophy: an unrecognized syndrome. Diabetes Care 26:1819–1824 [DOI] [PubMed] [Google Scholar]
  • 98. Bandsma RH, Smit GP, Kuipers F. 2002. Disturbed lipid metabolism in glycogen storage disease I. Eur J Pediatr 161:S65–S69 [DOI] [PubMed] [Google Scholar]
  • 99. Carr A, Samaras K, Chisholm DJ, Cooper DA. 1998. Pathogenesis of HIV-1-protease inhibitor-associated peripheral lipodystrophy, hyperlipidaemia, and insulin resistance. Lancet 351:1881–1883 [DOI] [PubMed] [Google Scholar]
  • 100. Huang-Doran I, Sleigh A, Rochford JJ, O'Rahilly S, Savage DB. 2010. Lipodystrophy: metabolic insight from a rare disorder. J Endocrinol 207:245–255 [DOI] [PubMed] [Google Scholar]
  • 101. Grunfeld C, Kotler DP, Shigenaga JK, Doerrler W, Tierney A, Wang J, Pierson RN, Jr, Feingold KR. 1991. Circulating interferon-α levels and hypertriglyceridemia in the acquired immunodeficiency syndrome. Am J Med 90:154–162 [PubMed] [Google Scholar]
  • 102. Nonogaki K, Moser AH, Feingold KR, Grunfeld C. 1994. α-Adrenergic receptors mediate the hypertriglyceridemia induced by endotoxin, but not tumor necrosis factor, in rats. Endocrinology 135:2644–2650 [DOI] [PubMed] [Google Scholar]
  • 103. Kronenberg F. 2005. Dyslipidemia and nephrotic syndrome: recent advances. J Ren Nutr 15:195–203 [DOI] [PubMed] [Google Scholar]
  • 104. Kaysen GA. 2009. Lipid and lipoprotein metabolism in chronic kidney disease. J Ren Nutr 19:73–77 [DOI] [PubMed] [Google Scholar]
  • 105. Luo L, Pu X, Wang Y, Xu N. 2010. Impaired plasma lipid profiles in acute hepatitis. Lipids Health Dis 9:5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. Brinton EA. 2010. Effects of ethanol intake on lipoproteins and atherosclerosis. Curr Opin Lipidol 21:346–351 [DOI] [PubMed] [Google Scholar]
  • 107. Taskinen MR, Nikkilä EA, Välimäki M, Sane T, Kuusi T, Kesäniemi A, Ylikahri R. 1987. Alcohol-induced changes in serum lipoproteins and in their metabolism. Am Heart J 113:458–464 [DOI] [PubMed] [Google Scholar]
  • 108. Kissebah AH, Harrigan P, Wynn V. 1973. Mechanism of hypertriglyceridaemia associated with contraceptive steroids. Horm Metab Res 5:184–190 [DOI] [PubMed] [Google Scholar]
  • 109. Stuyt PM, Demacker PN, Stalenhoef AF. 1986. Pancreatitis induced by oestrogen in a patient with type I hyperlipoproteinaemia. Br Med J (Clin Res Ed) 293:734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Bershad S, Rubinstein A, Paterniti JR, Le NA, Poliak SC, Heller B, Ginsberg HN, Fleischmajer R, Brown WV. 1985. Changes in plasma lipids and lipoproteins during isotretinoin therapy for acne. N Engl J Med 313:16:981–985 [DOI] [PubMed] [Google Scholar]
  • 111. Vu-Dac N, Gervois P, Torra IP, Fruchart JC, Kosykh V, Kooistra T, Princen HM, Dallongeville J, Staels B. 1998. Retinoids increase human apo C-III expression at the transcriptional level via the retinoid X receptor. Contribution to the hypertriglyceridemic action of retinoids. J Clin Invest 102:625–632 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Davies PJ, Berry SA, Shipley GL, Eckel RH, Hennuyer N, Crombie DL, Ogilvie KM, Peinado-Onsurbe J, Fievet C, Leibowitz MD, Heyman RA, Auwerx J. 2001. Metabolic effects of retinoids: tissue-specific regulation of lipoprotein lipase activity. Mol Pharmacol 59:170–176 [DOI] [PubMed] [Google Scholar]
  • 113. Crouse JR., 3rd 1987. Hypertriglyceridemia: a contraindication to the use of bile acid binding resins. Am J Med 83:243–248 [DOI] [PubMed] [Google Scholar]
  • 114. De Clercq E. 2009. Anti-HIV drugs: 25 compounds approved within 25 years after the discovery of HIV. Int J Antimicrob Agents 33:307–320 [DOI] [PubMed] [Google Scholar]
  • 115. Hakeam HA, Al-Jedai AH, Raza SM, Hamawi K. 2008. Sirolimus induced dyslipidemia in tacrolimus based vs. tacrolimus free immunosuppressive regimens in renal transplant recipients. Ann Transplant 13:46–53 [PubMed] [Google Scholar]
  • 116. Casey DE. 2004. Dyslipidemia and atypical antipsychotic drugs. J Clin Psychiatry 65(Suppl 18):27–35 [PubMed] [Google Scholar]
  • 117. Veerkamp MJ, de Graaf J, Hendriks JC, Demacker PN, Stalenhoef AF. 2004. Nomogram to diagnose familial combined hyperlipidemia on the basis of results of a 5-year follow-up study. Circulation 109:2980–2985 [DOI] [PubMed] [Google Scholar]
  • 118. Austin MA, McKnight B, Edwards KL, Bradley CM, McNeely MJ, Psaty BM, Brunzell JD, Motulsky AG. 2000. Cardiovascular disease mortality in familial forms of hypertriglyceridemia: a 20-year prospective study. Circulation 101:2777–2782 [DOI] [PubMed] [Google Scholar]
  • 119. Genest JJ, Jr, Martin-Munley SS, McNamara JR, Ordovas JM, Jenner J, Myers RH, Silberman SR, Wilson PW, Salem DN, Schaefer EJ. 1992. Familial lipoprotein disorders in patients with premature coronary artery disease. Circulation 85:2025–2033 [DOI] [PubMed] [Google Scholar]
  • 120. Genest J, Jr, Bard JM, Fruchart JC, Ordovas JM, Schaefer EJ. 1993. Familial hypoalphalipoproteinemia in premature coronary artery disease. Arterioscler Thromb 13:1728–1737 [DOI] [PubMed] [Google Scholar]
  • 121. Mahley RW, Huang Y, Rall SC., Jr 1999. Pathogenesis of type III hyperlipoproteinemia (dysbetalipoproteinemia). Questions, quandaries, and paradoxes. J Lipid Res 40:1933–1949 [PubMed] [Google Scholar]
  • 122. Smit M, de Knijff P, van der Kooij-Meijs E, Groenendijk C, van den Maagdenberg AM, Gevers Leuven JA, Stalenhoef AF, Stuyt PM, Frants RR, Havekes LM. 1990. Genetic heterogeneity in familial dysbetalipoproteinemia. The E2(lys146—gln) variant results in a dominant mode of inheritance. J Lipid Res 31:45–53 [PubMed] [Google Scholar]
  • 123. de Knijff P, van den Maagdenberg AM, Stalenhoef AF, Leuven JA, Demacker PN, Kuyt LP, Frants RR, Havekes LM. 1991. Familial dysbetalipoproteinemia associated with apolipoprotein E3-Leiden in an extended multigeneration pedigree. J Clin Invest 88:643–655 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Stalenhoef AF, Malloy MJ, Kane JP, Havel RJ. 1986. Metabolism of apolipoproteins B-48 and B-100 of triglyceride-rich lipoproteins in patients with familial dysbetalipoproteinemia. J Clin Invest 78:722–728 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Grundy SM, Cleeman JI, Daniels SR, Donato KA, Eckel RH, Franklin BA, Gordon DJ, Krauss RM, Savage PJ, Smith SC, Jr, Spertus JA, Costa F. 2005. Diagnosis and management of the metabolic syndrome: an American Heart Association/National Heart, Lung and Blood Institute scientific statement. Circulation 112:2735–2752 [DOI] [PubMed] [Google Scholar]
  • 126. Alberti KG, Eckel RH, Grundy SM, Zimmet PZ, Cleeman JI, Donato KA, Fruchart JC, James WP, Loria CM, Smith SC., Jr 2009. Harmonizing the metabolic syndrome: a joint interim statement of the International Diabetes Federation Task Force on Epidemiology and Prevention; National Heart, Lung, and Blood Institute; American Heart Association; World Heart Federation; International Atherosclerosis Society; and International Association for the Study of Obesity. Circulation 120:1640–1645 [DOI] [PubMed] [Google Scholar]
  • 127. Nieves DJ, Cnop M, Retzlaff B, Walden CE, Brunzell JD, Knopp RH, Kahn SE. 2003. The atherogenic lipoprotein profile associated with obesity and insulin resistance is largely attributable to intra-abdominal fat. Diabetes 52:172–179 [DOI] [PubMed] [Google Scholar]
  • 128. Blackburn P, Lemieux I, Alméras N, Bergeron J, Côté M, Tremblay A, Lamarche B, Després JP. 2009. The hypertriglyceridemic waist phenotype versus the National Cholesterol Education Program-Adult Treatment Panel III and International Diabetes Federation clinical criteria to identify high-risk men with an altered cardiometabolic risk profile. Metabolism 58:1123–1130 [DOI] [PubMed] [Google Scholar]
  • 129. Tessari P, Coracina A, Cosma A, Tiengo A. 2009. Hepatic lipid metabolism and nonalcoholic fatty liver disease. Nutr Metab Cardiovasc Dis 19:291–302 [DOI] [PubMed] [Google Scholar]
  • 130. Szendroedi J, Roden M. 2009. Ectopic lipids and organ function. Curr Opin Lipidol 20:50–56 [DOI] [PubMed] [Google Scholar]
  • 131. van Herpen NA, Schrauwen-Hinderling VB. 2008. Lipid accumulation in non-adipose tissue and lipotoxicity. Physiol Behav 94:231–241 [DOI] [PubMed] [Google Scholar]
  • 132. Hokanson JE, Austin MA. 1996. Plasma triglyceride level is a risk factor for cardiovascular disease independent of high-density lipoprotein cholesterol level: a meta-analysis of population-based prospective studies. J Cardiovasc Risk 3:213–219 [PubMed] [Google Scholar]
  • 133. Sarwar N, Danesh J, Eiriksdottir G, Sigurdsson G, Wareham N, Bingham S, Boekholdt SM, Khaw KT, Gudnason V. 2007. Triglycerides and the risk of coronary heart disease: 10,158 incident cases among 262,525 participants in 29 Western prospective studies. Circulation 115:450–458 [DOI] [PubMed] [Google Scholar]
  • 134. Patel A, Barzi F, Jamrozik K, Lam TH, Ueshima H, Whitlock G, Woodward M; Asia Pacific Cohort Studies Collaboration 2004. Serum triglycerides as a risk factor for cardiovascular diseases in the Asia-Pacific region. Circulation 110:2678–2686 [DOI] [PubMed] [Google Scholar]
  • 135. Tirosh A, Rudich A, Shochat T, Tekes-Manova D, Israeli E, Henkin Y, Kochba I, Shai I. 2007. Changes in triglyceride levels and risk for coronary heart disease in young men. Ann Intern Med 147:377–385 [DOI] [PubMed] [Google Scholar]
  • 136. Danesh J, Erqou S, Walker M, Thompson SG, Tipping R, Ford C, Pressel S, Walldius G, Jungner I, Folsom AR, Chambless LE, Knuiman M, Whincup PH, Wannamethee SG, Morris RW, Willeit J, Kiechl S, Santer P, Mayr A, Wald N, Ebrahim S, Lawlor DA, Yarnell JW, Gallacher J, Casiglia E, Tikhonoff V, Nietert PJ, Sutherland SE, Bachman DL, Keil JE, Cushman M, Psaty BM, Tracy RP, Tybjaerg-Hansen A, et al. 2007. The Emerging Risk Factors Collaboration: analysis of individual data on lipid, inflammatory and other markers in over 1.1 million participants in 104 prospective studies of cardiovascular diseases. Eur J Epidemiol 22:839–869 [DOI] [PubMed] [Google Scholar]
  • 137. Nordestgaard BG, Tybjaerg-Hansen A, Lewis B. 1992. Influx in vivo of low density, intermediate density, and very low density lipoproteins into aortic intimas of genetically hyperlipidemic rabbits. Roles of plasma concentrations, extent of aortic lesion, and lipoprotein particle size as determinants. Arterioscler Thromb 12:6–18 [DOI] [PubMed] [Google Scholar]
  • 138. Zhang X, Qi R, Xian X, Yang F, Blackstein M, Deng X, Fan J, Ross C, Karasinska J, Hayden MR, Liu G. 2008. Spontaneous atherosclerosis in aged lipoprotein lipase-deficient mice with severe hypertriglyceridemia on a normal chow diet. Circ Res 102:250–256 [DOI] [PubMed] [Google Scholar]
  • 139. Weinstein MM, Yin L, Tu Y, Wang X, Wu X, Castellani LW, Walzem RL, Lusis AJ, Fong LG, Beigneux AP, Young SG. 2010. Chylomicronemia elicits atherosclerosis in mice—brief report. Arterioscler Thromb Vasc Biol 30:20–23 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140. Marcoux C, Hopkins PN, Wang T, Leary ET, Nakajima K, Davignon J, Cohn JS. 2000. Remnant-like particle cholesterol and triglyceride levels of hypertriglyceridemic patients in the fed and fasted state. J Lipid Res 41:1428–1436 [PubMed] [Google Scholar]
  • 141. Nakamura T, Takano H, Umetani K, Kawabata K, Obata JE, Kitta Y, Kodama Y, Mende A, Ichigi Y, Fujioka D, Saito Y, Kugiyama K. 2005. Remnant lipoproteinemia is a risk factor for endothelial vasomotor dysfunction and coronary artery disease in metabolic syndrome. Atherosclerosis 181:321–327 [DOI] [PubMed] [Google Scholar]
  • 142. Havel RJ. 1994. Postprandial hyperlipidemia and remnant lipoproteins. Curr Opin Lipidol 5:102–109 [DOI] [PubMed] [Google Scholar]
  • 143. Zilversmit DB. 1979. Atherogenesis: a postprandial phenomenon. Circulation 60:473–485 [DOI] [PubMed] [Google Scholar]
  • 144. Véniant MM, Beigneux AP, Bensadoun A, Fong LG, Young SG. 2008. Lipoprotein size and susceptibility to atherosclerosis—insights from genetically modified mouse models. Curr Drug Targets 9:174–189 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Karpe F, Taskinen MR, Nieminen MS, Frick MH, Kesäniemi YA, Pasternack A, Hamsten A, Syvänne M. 2001. Remnant-like lipoprotein particle cholesterol concentration and progression of coronary and vein-graft atherosclerosis in response to gemfibrozil treatment. Atherosclerosis 157:181–187 [DOI] [PubMed] [Google Scholar]
  • 146. Véniant MM, Pierotti V, Newland D, Cham CM, Sanan DA, Walzem RL, Young SG. 1997. Susceptibility to atherosclerosis in mice expressing exclusively apolipoprotein B48 or apolipoprotein B100. J Clin Invest 100:180–188 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. Zhang WY, Schwartz E, Wang Y, Attrep J, Li Z, Reaven P. 2006. Elevated concentrations of nonesterified fatty acids increase monocyte expression of CD11b and adhesion to endothelial cells. Arterioscler Thromb Vasc Biol 26:514–519 [DOI] [PubMed] [Google Scholar]
  • 148. Gimbrone MA., Jr 1995. Vascular endothelium: an integrator of pathophysiologic stimuli in atherosclerosis. Am J Cardiol 75:67B–70B [DOI] [PubMed] [Google Scholar]
  • 149. Olufadi R, Byrne CD. 2006. Effects of VLDL and remnant particles on platelets. Pathophysiol Haemost Thromb 35:281–291 [DOI] [PubMed] [Google Scholar]
  • 150. Rutledge JC, Woo MM, Rezai AA, Curtiss LK, Goldberg IJ. 1997. Lipoprotein lipase increases lipoprotein binding to the artery wall and increases endothelial layer permeability by formation of lipolysis products. Circ Res 80:819–828 [DOI] [PubMed] [Google Scholar]
  • 151. Johnson RK, Appel LJ, Brands M, Howard BV, Lefevre M, Lustig RH, Sacks F, Steffen LM, Wylie-Rosett J; on behalf of the American Heart Association Nutrition Committee of the Council on Nutrition, Physical Activity, and Metabolism and the Council on Epidemiology and Prevention 2009. Dietary sugars intake and cardiovascular health: a scientific statement from the American Heart Association. Circulation 120:1011–1020 [DOI] [PubMed] [Google Scholar]
  • 152. Mensink RP, Zock PL, Kester AD, Katan MB. 2003. Effects of dietary fatty acids and carbohydrates on the ratio of serum total to HDL cholesterol and on serum lipids and apolipoproteins: a meta-analysis of 60 controlled trials. Am J Clin Nutr 77:1146–1155 [DOI] [PubMed] [Google Scholar]
  • 153. Wolever TM, Jenkins DJ, Jenkins AL, Josse RG. 1991. The glycemic index: methodology and clinical implications. Am J Clin Nutr 54:846–854 [DOI] [PubMed] [Google Scholar]
  • 154. Ludwig DS. 2002. The glycemic index: physiological mechanisms relating to obesity, diabetes, and cardiovascular disease. JAMA 287:2414–2423 [DOI] [PubMed] [Google Scholar]
  • 155. Appel LJ, Sacks FM, Carey VJ, Obarzanek E, Swain JF, Miller ER, 3rd, Conlin PR, Erlinger TP, Rosner BA, Laranjo NM, Charleston J, McCarron P, Bishop LM. 2005. Effects of protein, monounsaturated fat, and carbohydrate intake on blood pressure and serum lipids: results of the OmniHeart Randomized Trial. JAMA 294:2455–2464 [DOI] [PubMed] [Google Scholar]
  • 156. de Souza RJ, Swain JF, Appel LJ, Sacks FM. 2008. Alternatives for macronutrient intake and chronic disease: a comparison of the OmniHeart diets with popular diets and with dietary recommendations. Am J Clin Nutr 88:1–11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157. Furtado JD, Campos H, Sumner AE, Appel LJ, Carey VJ, Sacks FM. 2010. Dietary interventions that lower lipoproteins containing apolipoprotein C-III are more effective in whites than in blacks: results of the OmniHeart trial. Am J Clin Nutr 92:714–722 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158. Graham TE. 2004. Exercise, postprandial triacylglyceridemia, and cardiovascular disease risk. Can J Appl Physiol 29:781–799 [DOI] [PubMed] [Google Scholar]
  • 159. Dekker MJ, Graham TE, Ooi TC, Robinson LE. 2010. Exercise prior to fat ingestion lowers fasting and postprandial VLDL and decreases adipose tissue IL-6 and GIP receptor mRNA in hypertriacylglycerolemic men. J Nutr Biochem 21:983–990 [DOI] [PubMed] [Google Scholar]
  • 160. Tambalis K, Panagiotakos DB, Kavouras SA, Sidossis LS. 2009. Responses of blood lipids to aerobic, resistance, and combined aerobic with resistance exercise training: a systematic review of current evidence. Angiology 60:614–632 [DOI] [PubMed] [Google Scholar]
  • 161. Pitsavos C, Panagiotakos DB, Tambalis KD, Chrysohoou C, Sidossis LS, Skoumas J, Stefanadis C. 2009. Resistance exercise plus to aerobic activities is associated with better lipids' profile among healthy individuals: the ATTICA study. QJM 102:609–616 [DOI] [PubMed] [Google Scholar]
  • 162. Foster GD, Wyatt HR, Hill JO, Makris AP, Rosenbaum DL, Brill C, Stein RI, Mohammed BS, Miller B, Rader DJ, Zemel B, Wadden TA, Tenhave T, Newcomb CW, Klein S. 2010. Weight and metabolic outcomes after 2 years on a low-carbohydrate versus low-fat diet: a randomized trial. Ann Intern Med 153:147–157 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163. Sacks FM, Bray GA, Carey VJ, Smith SR, Ryan DH, Anton SD, McManus K, Champagne CM, Bishop LM, Laranjo N, Leboff MS, Rood JC, de Jonge L, Greenway FL, Loria CM, Obarzanek E, Williamson DA. 2009. Comparison of weight-loss diets with different compositions of fat, protein, and carbohydrates. N Engl J Med 360:859–873 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164. Brunzell JD, Davidson M, Furberg CD, Goldberg RB, Howard BV, Stein JH, Witztum JL. 2008. Lipoprotein management in patients with cardiometabolic risk: consensus conference report from the American Diabetes Association and the American College of Cardiology Foundation. J Am Coll Cardiol 51:1512–1524 [DOI] [PubMed] [Google Scholar]
  • 165. Frick MH, Elo O, Haapa K, Heinonen OP, Heinsalmi P, Helo P, Huttunen JK, Kaitaniemi P, Koskinen P, Manninen V. 1987. Helsinki Heart Study: primary-prevention trial with gemfibrozil in middle-aged men with dyslipidemia. Safety of treatment, changes in risk factors, and incidence of coronary heart disease. N Engl J Med 317:1237–1245 [DOI] [PubMed] [Google Scholar]
  • 166. Rubins HB, Robins SJ, Collins D, Fye CL, Anderson JW, Elam MB, Faas FH, Linares E, Schaefer EJ, Schectman G, Wilt TJ, Wittes J. for the Veterans Affairs High-Density Lipoprotein Cholesterol Intervention Trial Study Group 1999. Gemfibrozil for the secondary prevention of coronary heart disease in men with low levels of high-density lipoprotein cholesterol. N Engl J Med 341:410–418 [DOI] [PubMed] [Google Scholar]
  • 167. The BIP Study Group 2000. Secondary prevention by raising HDL cholesterol and reducing triglycerides in patients with coronary artery disease: the Bezafibrate Infarction Prevention (BIP) study. Circulation 102:21–27 [DOI] [PubMed] [Google Scholar]
  • 168. Keech A, Simes RJ, Barter P, Best J, Scott R, Taskinen MR, Forder P, Pillai A, Davis T, Glasziou P, Drury P, Kesäniemi YA, Sullivan D, Hunt D, Colman P, d'Emden M, Whiting M, Ehnholm C, Laakso M. 2005. Effects of long-term fenofibrate therapy on cardiovascular events in 9795 people with type 2 diabetes mellitus (the FIELD study): randomised controlled trial. Lancet 366:1849–1861 [DOI] [PubMed] [Google Scholar]
  • 169. Ginsberg HN, Elam MB, Lovato LC, Crouse JR, 3rd, Leiter LA, Linz P, Friedewald WT, Buse JB, Gerstein HC, Probstfield J, Grimm RH, Ismail-Beigi F, Bigger JT, Goff DC, Jr, Cushman WC, Simons-Morton DG, Byington RP. 2010. Effects of combination lipid therapy in type 2 diabetes mellitus. N Engl J Med 362:1563–1574 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170. Abourbih S, Filion KB, Joseph L, Schiffrin EL, Rinfret S, Poirier P, Pilote L, Genest J, Eisenberg MJ. 2009. Effects of fibrates on lipid profiles and cardiovascular outcomes: a systematic review. Am J Med 122:962.e1–962.e8 [DOI] [PubMed] [Google Scholar]
  • 171. Jun M, Foote C, Lv J, Neal B, Patel A, Nicholls SJ, Grobbee DE, Cass A, Chalmers J, Perkovic V. 2010. Effects of fibrates on cardiovascular outcomes: a systematic review and meta-analysis. Lancet 375:1875–1884 [DOI] [PubMed] [Google Scholar]
  • 172. Sacks FM. 2008. After the Fenofibrate Intervention and Event Lowering in Diabetes (FIELD) study: implications for fenofibrate. Am J Cardiol 102:34L–40L [DOI] [PubMed] [Google Scholar]
  • 173. Carlson LA, Böttiger LE. 1972. Ischaemic heart-disease in relation to fasting values of plasma triglycerides and cholesterol. Stockholm Prospective Study. Lancet 1:865–868 [DOI] [PubMed] [Google Scholar]
  • 174. Canner PL, Berge KG, Wenger NK, Stamler J, Friedman L, Prineas RJ, Friedewald W. 1986. Fifteen year mortality in Coronary Drug Project patients: long-term benefit with niacin. J Am Coll Cardiol 8:1245–1255 [DOI] [PubMed] [Google Scholar]
  • 175. Brown BG, Hillger L, Zhao XQ, Poulin D, Albers JJ. 1995. Types of change in coronary stenosis severity and their relative importance in overall progression and regression of coronary disease. Observations from the FATS Trial. Familial Atherosclerosis Treatment Study. Ann NY Acad Sci 748:407–417; discussion 417–418 [DOI] [PubMed] [Google Scholar]
  • 176. Brown BG, Zhao XQ, Chait A, Fisher LD, Cheung MC, Morse JS, Dowdy AA, Marino EK, Bolson EL, Alaupovic P, Frohlich J, Albers JJ. 2001. Simvastatin and niacin, antioxidant vitamins, or the combination for the prevention of coronary disease. N Engl J Med 345:1583–1592 [DOI] [PubMed] [Google Scholar]
  • 177. Boden WE, Probstfield JL, Anderson T, Chaitman BR, Desvignes-Nickens P, Koprowicz K, McBride R, Teo K, Weintraub W. 2011. Niacin in patients with low HDL cholesterol levels receiving intensive statin therapy. N Engl J Med 365:2255–2267 [DOI] [PubMed] [Google Scholar]
  • 178. Bhardwaj SS, Chalasani N. 2007. Lipid-lowering agents that cause drug-induced hepatotoxicity. Clin Liver Dis 11:597–613, vii [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Musa-Veloso K, Binns MA, Kocenas AC, Poon T, Elliot JA, Rice H, Oppedal-Olsen H, Lloyd H, Lemke S. 2010. Long-chain omega-3 fatty acids eicosapentaenoic acid and docosahexaenoic acid dose-dependently reduce fasting serum triglycerides. Nutr Rev 68:155–167 [DOI] [PubMed] [Google Scholar]
  • 180. Yokoyama M, Origasa H, Matsuzaki M, Matsuzawa Y, Saito Y, Ishikawa Y, Oikawa S, Sasaki J, Hishida H, Itakura H, Kita T, Kitabatake A, Nakaya N, Sakata T, Shimada K, Shirato K. 2007. Effects of eicosapentaenoic acid on major coronary events in hypercholesterolaemic patients (JELIS): a randomized open-label, blinded endpoint analysis. Lancet 369:1090–1098 [DOI] [PubMed] [Google Scholar]
  • 181. Harris WS, Ginsberg HN, Arunakul N, Shachter NS, Windsor SL, Adams M, Berglund L, Osmundsen K. 1997. Safety and efficacy of Omacor in severe hypertriglyceridemia. J Cardiovasc Risk 4:385–391 [PubMed] [Google Scholar]
  • 182. Daviglus ML, Stamler J, Orencia AJ, Dyer AR, Liu K, Greenland P, Walsh MK, Morris D, Shekelle RB. 1997. Fish consumption and the 30-year risk of fatal myocardial infarction. N Engl J Med 336:1046–1053 [DOI] [PubMed] [Google Scholar]
  • 183. Hu FB, Bronner L, Willett WC, Stampfer MJ, Rexrode KM, Albert CM, Hunter D, Manson JE. 2002. Fish and omega-3 fatty acid intake and risk for coronary heart disease in women. JAMA 287:1815–1821 [DOI] [PubMed] [Google Scholar]
  • 184. GISSI-Prevenzione Investigators 1999. Dietary supplementation with n-3 polyunsaturated fatty acids and vitamin E after myocardial infarction: results of the GISSI-Prevenzione trial. Gruppo Italiano per lo Studio della Sopravivenza nell'Infarto miocardico. Lancet 354:447–455 [PubMed] [Google Scholar]
  • 185. Graham DJ, Staffa JA, Shatin D, Andrade SE, Schech SD, La Grenade L, Gurwitz JH, Chan KA, Goodman MJ, Platt R. 2004. Incidence of hospitalized rhabdomyolysis in patients treated with lipid-lowering drugs. JAMA 292:2585–2590 [DOI] [PubMed] [Google Scholar]
  • 186. Sharma M, Ansari MT, Abou-Setta AM, Soares-Weiser K, Ooi TC, Sears M, Yazdi F, Tsertsvadze A, Moher D. 2009. Systematic review: comparative effectiveness and harms of combination therapy and monotherapy for dyslipidemia. Ann Intern Med 151:622–630 [DOI] [PubMed] [Google Scholar]
  • 187. Nakou ES, Filippatos TD, Agouridis AP, Kostara C, Bairaktari ET, Elisaf MS. 2010. The effects of ezetimibe and/or orlistat on triglyceride-rich lipoprotein metabolism in obese hypercholesterolemic patients. Lipids 45:445–450 [DOI] [PubMed] [Google Scholar]
  • 188. Devchand PR. 2008. Glitazones and the cardiovascular system. Curr Opin Endocrinol Diabetes Obes 15:188–192 [DOI] [PubMed] [Google Scholar]
  • 189. Betteridge DJ. 2007. Effects of pioglitazone on lipid and lipoprotein metabolism. Diabetes Obes Metab 9:640–647 [DOI] [PubMed] [Google Scholar]
  • 190. Rizos CV, Elisaf MS, Mikhailidis DP, Liberopoulos EN. 2009. How safe is the use of thiazolidinediones in clinical practice? Expert Opin Drug Saf 8:15–32 [DOI] [PubMed] [Google Scholar]
  • 191. Lewis JD, Ferrara A, Peng T, Hedderson M, Bilker WB, Quesenberry CP, Jr, Vaughn DJ, Nessel L, Selby J, Strom BL. 2011. Risk of bladder cancer among diabetic patients treated with pioglitazone: interim report of a longitudinal cohort study. Diabetes Care 34:916–922 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Clinical Endocrinology and Metabolism are provided here courtesy of The Endocrine Society

RESOURCES