U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Feingold KR, Anawalt B, Blackman MR, et al., editors. Endotext [Internet]. South Dartmouth (MA): MDText.com, Inc.; 2000-.

Cover of Endotext

Endotext [Internet].

Show details

Familial Hypercholesterolemia: Genes and Beyond

, PharmD, BCPS-AQ Cardiology, CLS, FNLA, FASPC, , MD, PhD, and , DO, MCR.

Author Information and Affiliations

Last Update: October 23, 2021.

ABSTRACT

Genetic disorders resulting in familial hypercholesterolemia (FH) include autosomal dominant hypercholesterolemia (ADH), polygenic hypercholesterolemia, as well as other rare conditions such as autosomal recessive hypercholesterolemia (ARH). All of these disorders cause elevations in low-density lipoprotein (LDL)-cholesterol (LDL-C) and, as a result, greatly increase the risk of cardiovascular disease (CVD). Genetic loci involved in ADH include the LDLR, which codes for the LDL receptor (LDLR), APOB, which codes for apolipoprotein B-100 (apoB-100), the major protein component of LDL, PCSK9, which codes for Proprotein Convertase Subtilisin/Kexin Type 9 (PCSK9), the low abundance circulatory protein that terminates the lifecycle of the LDLR, and apolipoprotein E (APOE), which synthesizes the main protein of triglyceride rich lipoproteins. Importantly, a large percentage of people with the severe hypercholesterolemic phenotype do not possess a readily identifiable gene defect and many likely have polygenic hypercholesterolemia. Thus, identification of a specific genetic pathologic variant is not a necessary condition for the diagnosis of a genetic hypercholesterolemia. Several formal diagnostic criteria exist for FH and include lipid levels, family history, personal history, physical exam findings, and genetic testing. As all individuals with severe hypercholesterolemia are at high risk for CVD, treatment is centered on dietary and lifestyle modifications and early institution of lipid-lowering pharmacotherapy. Treatment should initially be statin-based, but most patients require adjunctive medications such as ezetimibe and PCSK9 blocking monoclonal antibodies. Three large cardiovascular outcome trials have shown a reduction in atherosclerotic CVD when ezetimibe or PCSK9 blocking monoclonal antibodies were added to a background of statin therapy and consequently have assisted in shaping international guidelines and consensus recommendations. Novel therapeutics recently developed, include: inclisiran – a small interfering ribonucleic acid (siRNA)-based gene-silencing technology that inhibits PCSK9 production (approved in the European Union only thus far), bempedoic acid – an inhibitor of adenosine triphosphate (ATP)-citrate lyase (FDA approved for heterozygous FH and patients with prior CVD and elevated LDL-C), and evinacumab – a fully human monoclonal antibody inhibiting angiopoietin-like 3 (ANGPTL3) (FDA approved for homozygous FH only). Patients with extreme and unresponsive elevations in LDL-C will require more aggressive therapies such as lipoprotein apheresis and agents for the treatment of severe hypercholesterolemia such as microsomal triglyceride transfer protein (MTP) inhibitors and evinacumab. For complete coverage of all related areas of Endocrinology, please visit our on-line FREE web-text, WWW.ENDOTEXT.ORG.

INTRODUCTION

Genetic disorders resulting in familial hypercholesterolemia (FH) consist of autosomal dominant hypercholesterolemia (ADH), autosomal recessive hypercholesterolemia (ARH), and polygenic hypercholesterolemia. The genetic architecture of FH is more complex than previously recognized and in fact is now believed to be associated with at least nine different genes with thousands of variants, the details of which are beyond the scope of this chapter. But briefly, the term “autosomal dominant hypercholesterolemia” refers to those patients with dominantly inherited severe hypercholesterolemia – low-density lipoprotein (LDL)-cholesterol (LDL-C) greater than 190 mg/dL, who likely harbor mutations in genes regulating serum LDL levels. Historically, the more common causes of ADH include “classic” FH, which is a codominant disorder involving aberrations in the LDL receptor (LDLR), as well as other codominant forms of “nonclassical” FH, which involve defects in two other genes that regulate plasma clearance of LDL, apolipoprotein B (APOB), which synthesizes the main protein of LDL (a ligand for the LDLR), and Proprotein Convertase Subtilisin/Kexin Type 9 (PCSK9), which synthesizes a low abundance circulatory protein that limits the LDLR lifespan (1,2). Autosomal dominant forms of ADH previously include mutations in apolipoprotein E (APOE), which synthesizes the main protein of triglyceride rich lipoproteins and signal transducing adaptor family member 1 (STAP1), whose function in cholesterol homeostasis remains largely unknown (2,3). However, recent data have determined that mutations in STAP1 are not a causative factor in FH (4-7). An autosomal recessive form of FH, ARH, is very rare and results from pathogenic variants in LDLR accessory protein 1 (LDL-RAP1). Other forms may be due to defects in lysosomal acid lipase (LIPA), ATP- binding cassette sub-family G member 5 and 8 (ABCG5/8), and cholesterol 7 alpha-hydroxylase (CYP7A1) (2,8). Finally, a common form of FH is attributable to multiple variations in several genes each with minor effects on cholesterol regulation (more than 50 loci identified, as opposed to a single large effect as seen in the textbook version of FH). Polygenic causes are relatively common and likely explain many of the patients who are genotype negative. In fact, up to 50% of patients referred to lipid clinics for possible or probable HeFH have a polygenic basis (2). Additionally, only approximately 2% of patients with an LDL >190 mg/dL and no additional clinical or family history compatible with FH have a pathogenic variant in one of the FH genes (9,10). Thus, having an elevated LDL-C does not necessarily indicate that the patient has ADH due to a pathogenic variation in the LDLR, APOB, PCSK9, or APOE genes. All forms of FH result in very high levels of LDL-C and increase the risk of early and accelerated coronary artery disease (CAD) (8). Yet, FH remains vastly underdiagnosed and thus, undertreated, representing an extraordinary missed opportunity for maintenance of cardiovascular healthy and prevention of cardiovascular events. It has been estimated that less than 10% of the patients in the US with FH have been diagnosed (11,12). This chapter will largely focus on the canonical forms of FH involving the five traditional genetic loci described above.

GENETICS

The LDLR is an 893-amino acid cell surface glycoprotein that binds and internalizes LDL particles, primarily in the liver. Mutations in LDLR (i.e., “classic” FH) give rise to nearly 90% of cases of clinical ADH (13). Over 2000 such mutations in LDLR have been identified, including deletions, insertions, missense, copy number variants, and nonsense mutations (2). FH patients can be homozygous (traditionally with a prevalence thought to be 1 in 1,000,000 but based on contemporary genetic studies, the prevalence is thought to be closer to 1 in 300,000), carrying mutations in both alleles encoding for LDLR, or heterozygous (traditionally with a prevalence thought to be 1 in 500, with newer data suggesting as frequent as 1 in 200), possessing mutations in only one allele (2,14). Homozygous FH (HoFH) should be suspected when LDL-C exceeds 400 mg/dL, whereas heterozygous FH (HeFH) should be suspected when LDL-C is greater than 190 mg/dL in adults and 160 mg/dL in children (8). Patients with HoFH can be true FH homozygotes, with two identical mutations in each allele, versus compound heterozygotes, with a different mutation in each allele. In addition, FH can result in elevated levels of lipoprotein(a) (Lp[a]) through an unclear mechanism, not necessarily linked to the dysfunctional LDLR pathway (15-19). Elevated Lp(a), which is composed of 30-45% cholesterol by mass and reported as part of the LDL-C laboratory measurement, amplifies the already increased risk of incident CAD seen in those with FH (19). It is also important to be aware of “founder effects” in some populations. Founder effects influencing the type and frequency of mutations causing FH are seen among Afrikaners, French Canadians, Ashkenazi Jews, Christian Lebanese, and some Tunisian groups. Slimane et al. estimated the prevalence of individuals with HoFH and HeFH in Tunisia to be 1:125,000 and 1:165, respectively (20).

“Nonclassical” FH, which phenotypically resembles classic FH in presentation and severity, involves dominantly inherited gene defects in APOB, PCSK9, and APOE, which code for proteins that modulate ligand-LDL/LDL-like receptor interaction (2,21,22). ApoB is the major protein constituent of LDL and acts as a ligand for LDLR. Mutations in ApoB (most commonly a single base change at a position near amino-acid 3,500) block the binding of LDL containing the apoB-100 to LDLR, resulting in severely elevated levels of LDL-C. This condition was originally defined as “familial defective APOB-100” or FDAB (23). PCSK9, on the other hand, is a circulating protein that terminates the lifecycle of LDLR by binding to it and targeting it to lysosomal degradation. Gain-of-function (GOF) mutations in PCSK9 lead to a FH phenotype, whereas loss-of-function (LOF) mutations lead to lower LDL-C and protection from coronary atherosclerotic events (24,25). Absence of circulating PCSK9 has been reported in a few subjects, who were reportedly healthy and had LDL levels around 20 mg/dL (26,27). Collectively, these observations spurred a frenzy of targeted research that led to the development and FDA approval of therapeutic antibodies against PCSK9 to reduce LDL levels in individuals with atherosclerotic cardiovascular disease (ASCVD) and/or in FH. Mutations in ApoE (an in-frame three base-pair deletion at position 167 in exon 4) block the binding of triglyceride rich lipoproteins (i.e., chylomicrons, chylomicron remnants, very low-density lipoprotein [VLDL-C], intermediate-density lipoprotein [IDL-C]) to the E receptor (belongs to the LDLR superfamily), and limits clearance of these particles from plasma (28). The prevalence of ADH resulting from mutations in APOB, PCSK9, and APOE has been difficult to estimate, but it is agreed that these are 5-10%, <1%, and <<1% respectively (2).

Finally, an additional ultra-rare recessive genetic disorder causing hypercholesterolemia bears mentioning. It involves a homozygous deletion mutation in the gene CYP7A1. This gene codes for the enzyme cholesterol 7α-hydroxylase, which catalyzes the initial step in cholesterol catabolism and bile acid synthesis. The mutation results in loss of enzymatic function and high levels of LDL-C, and was first identified in three homozygotes within a single kindred of English and Celtic descent (29). There are a number of other rare recessive genetic disorders that are associated with hypercholesterolemia, see table 1 below and other Endotext chapters on these rare genetic disorders.

TABLE 1.

RARE GENETIC DISORDERS THAT CAN BE CONFUSED WITH ADH

DisorderDescription
SitosterolemiaAutosomal recessive disorder due to mutations in ABCG5/8. Manifestations include hypercholesterolemia, marked elevations of plasma plant sterols, tendon and tuberous xanthomas, and premature ASCVD
Lysosomal acid lipase deficiencyAutosomal recessive disorder due to mutations in LIPA, which codes for lysosomal acid lipase. Manifestations include moderate hypercholesterolemia with depressed high-density lipoprotein cholesterol (HDL-C), accelerated atherosclerosis, and progressive liver disease.
Deficiency of LDL receptor adapter protein 1Autosomal recessive disorder due to mutations in LDL-RAP1. Typically presents with very high LDL cholesterol levels.
Deficiency of cholesterol 7-alpha hydroxylaseAutosomal recessive disorder due to mutations in CYP7A1, which codes for the enzyme that catalyzes the first step in bile acid synthesis, resulting in high intrahepatic cholesterol and reduced surface expression of LDLR.

An important practical point is that 30-50% of people with the FH phenotype have no readily identifiable defects in any of the genes that have been mentioned here; thus, diagnosing an individual with the FH phenotype does not necessarily mean the presence of a monogenic defect in the LDLR pathway (30). The genetic confirmation for FH-causing mutations in ADH varies considerably (2,10). The rate of positive genetics is related to clinical criteria - in patients with definite FH based on clinical criteria 60-80% are positive whereas in patients with possible FH based on clinical criteria only 21-44% are positive (31). Additionally, the LDL-C level is crucial. When LDL-C is very high (i.e., >310 mg/dL), the frequency of monogenic pathogenic variants is as high as 92% (32). In patients without an identifiable pathologic genetic variant the etiology may be due to an as-yet-unidentified genetic error or to polygenic, epigenetic, or non-genetic factors, including co-morbid and environmental modifiers. For those without a mutation, an elevated LDL-C still confers elevated cardiovascular disease (CVD) risk. However, for any given LDL-C strata, those with a causal mutation compared to those without have higher risk for CVD, likely due to lifelong exposure to high LDL-C (10,33). In addition, unintended consequences of genetic testing (i.e., genetic discrimination for life or long-term care insurance and increased expense) must be taken into account. For these reasons, genetic testing should not be mandated, but should involve a shared-decision making model between patient and provider, encompassing benefits and risks of genetic testing, as well as patient values and preferences (34). On the other hand, others advocate for routine genetic testing citing as rationale 1) facilitate a definitive diagnosis; 2) identify pathogenic variants with higher cardiovascular risk and therefore needing more aggressive treatment; 3) increase initiation of and adherence to therapy; 4) facilitate insurance approval for novel lipid-lowering therapies; and 5) cascade testing of first-degree relatives (25). Cascade screening on the basis of LDL-C alone (i.e., ≥190 mg/dL) has low sensitivity and specificity, however, identification of a pathogenic variant with genetic testing followed by cascade screening results in very high sensitivity and specificity. This is likely due to missed diagnoses of FH with reduced penetrance where LDL-C is <190 mg/dL (31). Nonetheless, the reduced costs and more widespread availability of genetic testing warrant performance of this test to obtain information that can help the physician familiarize with genotype-phenotype correlations and identify subjects that can be studied for the discovery of novel pathways leading to severe hypercholesterolemia. It must be noted that the FDA does not mention genetic testing as a measure to define FH (either heterozygous or homozygous).

PATHOPHYSIOLOGY

Most circulating LDL particles end up in the liver. The LDLR pathway is the predominant method for LDL uptake (35,36). ApoB binds to a specific binding site on LDLR and the receptor-ligand complex is subsequently internalized from clathrin-coated pits on the cell membrane. The receptor-ligand complex undergoes endocytosis and is targeted to the lysosome, where LDL is released for degradation while the LDLR is recycled back to the cell surface approximately 100-150 times in its 24 hour life cycle (2). PCSK9 terminates LDLR lifespan by disallowing its recycling, thus providing a physiologic mechanism of protein removal much different from, and stronger than, that caused by inducible degrader of LDL, an E3 ubiquitin ligase (24,37). There are other nonspecific and constitutively active pathways of LDL-C clearance as well (35,38). In HeFH, though transport through the LDLR pathway is reduced by 50%, LDL-C clearance is doubled through these other, non-LDLR pathways. The same holds true for HoFH, where despite a near-absolute reduction in LDLR transport, total LDL-C clearance via non-specific pathways is increased by 4-fold (39). Excess LDL-C, which accumulates in liver cells, is then re-exported via the apoB system back into the plasma, secreted into bile unchanged, or transformed into bile acids. This increased production of LDL adds to inefficient clearance via LDLR to cause elevated serum LDL levels typical of the FH phenotype (40).

ADH can thus be further classified into subtypes 1, 2, and 3, based on which protein of the LDLR pathway is causative (Figure 1). ADH-1 comprises mutations within LDLR, the canonical form of FH. There are six major classes of ADH-1 (see table 2 below), based on the type of mutation. These include those that: inhibit synthesis of LDLR; impede exit of mature LDLR from the endoplasmic reticulum; affect the binding site of LDLR to apoB-100; prevent the ligand-receptor interaction; prevent endocytosis of the LDLR-apoB-100 complex; or inhibit recycling of LDLR to the cell surface for further rounds of lipid uptake (not shown). ADH-2 comprises mutations of APOB that block the association of apoB-100 to LDLR. ADH-3 is due to GOF mutations of PCSK9, which reduce LDLR recycling and accelerate its lysosomal degradation (12). Some authors have suggested that mutations that affect binding of apoB-100 to LDLR carry a less severe phenotype than those that affect LDLR directly (41-45).

FIGURE 1 . (22): ADH-1 comprises mutations within LDLR.

FIGURE 1

(22): ADH-1 comprises mutations within LDLR. There are six major classes of ADH-1, affecting: a) synthesis of LDLR; b) exit of mature LDLR from the endoplasmic reticulum; c) binding site of LDLR to apoB-100; d) endocytosis of LDLR-apoB-100 complex; and recycling of LDLR to the cell surface (not shown). ADH-2 comprises mutations in apoB that block the association of apoB-100 to LDLR. ADH-3 is due to GOF mutations of PCSK9, which reduce LDLR recycling and accelerate its lysosomal degradation. Adapted from Calandra et al. J. Lipid Research 2011; 52: 1885-926.

TABLE 2.

THE SIX CLASSES OF LDLR MUTATIONS (46)

Class 1: synthesis of receptor or precursor protein is absentThe so-called null allele is a prevalent class of mutations and is generally associated with very high LDL-C levels. The molecular basis of this type of mutation shows a wide variety: point mutations introducing a stop codon, mutations in the promoter region completely blocking transcription, mutations giving rise to incorrect excision of mRNA, and finally, large deletions preventing the assembly of a normal receptor.
Class 2: absent or impaired formation of receptor proteinThis class comprises mutations in which the normal routing through the cell is not complete or is only very slowly completed. Usually, there is a complete blockade of transport, and LDL receptors are unable to leave the ER. The Golgi apparatus is not reached, and the increase of 40,000 Da in molecular weight does not take place. Truncated proteins, as a result of a premature stop codon, and misfolded proteins, as a result of mutations in cysteine-rich regions leading to free or unpaired cysteine residues, are retained in the ER. However, quality control by the ER is not perfect, given the observation that sometimes misfolded proteins leave the ER but are processed more slowly. Such mutations give rise to class 2B mutations, in contrast to class 2A mutations that cause complete retaining in the ER.
Class 3: normal synthesis of receptor protein, abnormal LDL bindingReceptors characterized by this class of alleles show the normal rate of synthesis, exhibit normal conversion into receptor protein, and are transported to the cell surface, but binding to LDL is impaired. It is obvious that mutations in the binding domain underlie this class of receptors.
Class 4: clustering in coated pits, internalization of the receptor complex does not take placeThe receptors in this class lack the property to cluster in coated pits (class 4A). This phenomenon, which makes interaction of receptors with the fuzzy coat impossible, is caused by mutations in the carboxyterminal part of the receptor protein. These mutated receptors are synthesized normally, folding and transport are normal, but clustering in coated pits is impossible, and sometimes the receptors are secreted even after they have reached the cell surface (class 4B).
Class 5: receptors are not recycled and are rapidly degradedAll mutations in this class are localized in the EGF-precursor homologous domain of the LDL receptor protein. This domain seems to be involved in the acid-dependent dissociation of the receptor-ligand complex in endosomes, after which the receptor can be recycled. When the entire EGF-precursor homologous domain is deleted by site-directed mutagenesis or when such a deletion occurs naturally in a homozygous FH patient, the receptor is trapped in the endosomes, and rapid degradation subsequently is observed.
Class 6: receptors fail to be targeted to the basolateral membraneThe class of mutations was recently discovered and is caused by alterations in the cytoplasmic tail of the protein. Such receptors do not reach the liver cell membrane and are probably rapidly degraded.
*

Adapted from Gidding SS, et al. Circulation 2015;132:2167-92.

CLINICAL MANIFESTATIONS

Lipid Abnormalities in Heterozygous Familial Hypercholesterolemia

LDL-C levels are frequently greater than the 90th percentile for age and gender. The magnitude of the LDL-C elevation is affected by the specific mutations causing FH with mutations in the LDLR leading to greater elevations in LDL-C levels than mutations in ApoB or PCSK9 (32,41-44,47). Null mutations in the LDLR are more severe than non-null mutations (48). Additionally, other genes that regulate LDL-C and environmental factors, such as diet, also influence the magnitude of the elevation in LDL-C (49,50). It should be recognized that a significant number of patients with genetically diagnosed FH have LDL-C <190 mg/dL. In some studies, approximately 50% of patients with genetically diagnosed FH have LDL-C levels <190 mg/dL (9,10). HDL-C and triglyceride levels are usually normal or only modestly altered (51-58). Elevated triglycerides and/or low HDL-C do not rule out the diagnosis of FH. Lp(a) levels are frequently elevated in patients with FH and may contribute to the increased risk of ASCVD (15-19). One should exclude secondary causes of marked elevations in LDL-C particularly hypothyroidism, renal disease, autoimmunity, and iatrogenic conditions (see chapter on Approach to the Patient with Dyslipidemia) (59).

Atherosclerotic Cardiovascular Disease in Heterozygous Familial Hypercholesterolemia

Patients with FH have a 3-13 fold higher risk of ASCVD (60). Untreated males with FH have a 50% risk for a fatal or non-fatal myocardial infarction by 50 years of age whereas untreated females have a 30% chance by age 60 (61,62). However, it should be recognized that there is heterogeneity of ASCVD risk. Other cardiovascular risk factors, such as male sex, BMI, diabetes, hypertension, smoking, low HDL-C levels, and Lp(a) levels modulate the risk of ASCVD (60). Patients with FH who have corneal arcus or xanthomas are more likely to have ASCVD (63-65). Of particular note, patients with mutations that result in FH have a greater ASCVD risk than patients with equivalent LDL-C levels (10,33). This is likely due to the LDL-C elevations being present from birth (lifelong exposure to elevated LDL-C).

Other Manifestations

Early onset corneal arcus (age < 45) and tendinous xanthomas, particularly the Achilles tendon and dorsum of hands, are classical abnormalities that occur in patients with FH. Xanthelasma (xanthomas in eyelids) and tuberous xanthoma may also be seen. However, it should be recognized that in the modern era with the increased treatment of elevated LDL-C levels these abnormalities are no longer frequently seen (only 5-35% of patients have xanthoma or corneal arcus currently) (66,67).

Homozygous Familiar Hypercholesterolemia

This is a rare disorder with untreated LDL-C levels that vary but are markedly elevated (usually > 300 mg/dL but often > 500 mg/dL) (68). Patients who are LDLR negative have higher LDL-C and a poorer clinical prognosis than LDLR defective patients (68). Lp(a) levels tend to be higher than observed in patients with HeFH (16). Additionally, HDL-C levels tend to be decreased in HoFH (68). Tendinous xanthoma, tuberous xanthomas, and arcus cornea may appear in childhood (68). If untreated >50% of patient with HoFH develop clinically significant ACVD by age of 30 and cardiovascular events can occur before age 10 in some patients (69). Almost 90% of patients with HoFH suffered a cardiovascular event by age 40 (69). Cholesterol and calcium deposits can lead to aortic stenosis and occasionally to mitral regurgitation (68,70-72).

DIAGNOSIS

FH, despite its different underlying gene abnormalities, leads to severe hypercholesterolemia and a distinct FH phenotype with markedly increased risk of developing CAD. In general, there are five major clinical criteria for diagnosing FH (see table 3): a family history of early CAD (less than age 55 in a first-degree relative in men, and less than age 65 in women), early CAD in the index case, elevated LDL-C (greater than 190 mg/dL), tendon xanthomas (especially in the Achilles and finger extensor tendons), and corneal arcus (which is highly specific in younger patients, but overall an insensitive finding). Mutations in any of the aforementioned genes of the LDLR pathway, when they are identified, are diagnostic.

TABLE 3.

MAJOR CLINICAL CRITERIA FOR DIAGNOSING FH

When to Suspect FH
If LDL-C levels are > 190 mg/dL (4.92 mMol/L) or non-HDL-C levels are >220 mg/dL (5.70mMol/L)
Patients with premature ASCVD (<55 years of age in male and <65 years of age in females)
Family history of hypercholesterolemia
When there is a positive family history of premature ASCVD (<55 years of age in male and <65 years of age in females)
When tendon xanthomas or corneal arcus (< age 45) are present on physical exam

As has been mentioned in other chapters of this text, when evaluating a patient suspected of having FH, it is critical to rule out secondary causes of hypercholesterolemia, such as hypothyroidism, nephrotic syndrome, and liver disease. Another extremely rare cause of non-FH has been described, involving autoantibodies to LDLR that inhibit receptor-mediated binding and catabolism of LDL-C (73).

There are three sets of statistically validated criteria that are most commonly used in the diagnosis of FH: the Dutch Lipid Network criteria, Simon Broome Register criteria, and Make Early Diagnosis to Prevent Early Deaths (MEDPED) criteria (74,75). These are summarized in Table 4, below (76).

TABLE 4.

SCORING SYSTEMS FOR DIAGNOSING FH

MEDPED Criteria (USA)
FH diagnostic if total cholesterol (LDL-C) levels exceed these cut points in mg/dL
Age1st degree relative2nd degree relative3rd degree relativeGeneral population
< 18220 (155)230 (165)240 (170)270 (200)
20240 (170)250 (180)260 (185)290 (220)
30270 (190)280 (200)290 (210)340 (240)
> 40290 (205)300 (215)310 (225)360 (260)
Simon Broome Criteria (UK)
Total cholesterol (LDL-C) 290 (190) mg/dL in adults, or 260 (155) mg/dL in childrenANDDNA mutation Definite FH
Tendon xanthomas in patient or 1st or 2nd Probable FH degree relative
Family history of MI at age < 50 in 2nd Possible FH degree relative or < 60 in 1st degree relative
OR
Family history of total cholesterol > 290
mg/dL in 1st or 2nd degree relative
Dutch Criteria (Netherlands)
1 point1st degree relative with premature CVD or LDL-C > 95th percentile, OR
Personal history of premature peripheral or cerebrovascular disease, OR
LDL-C 155-189 mg/dL
Definite FH
(8 points or more)
Probable FH (6-7 points)
Possible FH
(3-5 points)
2 points1st degree relative with tendon xanthoma or corneal arcus, OR
1st degree relative child (< 18 years) with LDL-C > 95th percentile, OR
Personal history of CAD
3 pointsLDL-C 190-249 mg/dL
4 pointsPresence of corneal arcus in patient < 45 years old
5 pointsLDL-C 250-329 mg/dL
6 pointsPresence of a tendon xanthoma
8 pointsLDL-C > 330 mg/dL, OR
Functional mutation of the LDLR gene

Adapted from Fahed et al., Nutrition & Metabolism 2011;8:23.

Unlike MEDPED criteria, which use only lipid levels, the Simon Broome and Dutch criteria also use family history, personal history, physical exam findings, and genetic testing to establish an FH diagnosis. Again, however, it should be emphasized that FH should be diagnosed phenotypically, as opposed to genetically—most FH patients are genotype-negative and do not possess a clear genetic substrate for their hyperlipidemic phenotype, but they clearly warrant aggressive intervention.

It is important to note that in the modern era of earlier recognition, wide-spread statin use and possibly improved dietary messages, it is often difficult to make a definitive or probable diagnosis of FH using clinical criteria (i.e., treatment reduces the development of xanthomas and corneal arcus and reduces or delays the occurrence of ASCVD events in patients and relatives). Similarly, it is often very difficult to know the before treatment LDL-C levels (31).

While genetic screening is not required for clinical management, lipid screening in family members should be undertaken in all individuals by age 20, starting as early as age 2 (42). Cascade screening—i.e., lipid screening of first-degree relatives of the proband—is infrequently employed, but is recommended as the most economical method of identifying new cases of FH (43). It is the responsibility of the examining clinician to attempt identification of other cases when making the diagnosis of FH in any given patient.

A potential novel solution to the underdiagnosis and undertreatment issues that plague the FH community, lies in leveraging health information technology. The FIND FH study demonstrated the use of a machine learning algorithm can successfully utilize medical profiles within the electronic health record to consistently identify individuals with probable FH (77). This new approach possesses the promise of identifying FH patients on a national scale and will hopefully lead to increased initiation of effective preventive therapies and at an earlier time-point as well.

TREATMENT

Goals of Therapy

In genetic disorders causing hypercholesterolemia, aggressive lipid-lowering through lifestyle modification, pharmacologic treatment, and invasive treatments such as apheresis has been shown to decrease angiographically-apparent CAD and reduce cardiovascular events (69,78,79). However, traditional risk assessment tools like the Framingham risk score do not apply to FH patients. Recent guidelines suggest that drug therapy should be initiated when LDL-C is greater than 190 mg/dL in all patients, including children over the age of eight (80,81). Most recommend at least a 50% reduction in LDL-C with statin therapy as a starting goal, with some advocating for targeting LDL-C less than 100 mg/dL and consideration of non-statin options (ezetimibe, bile-acid sequestrants, and PCSK9 inhibiting (PCSK9i) monoclonal antibodies) should these thresholds not be met (81,82). The European Atherosclerosis Society suggests LDL-C goals of less than 135 mg/dL in pediatric patients, less than 100 mg/dL in adults, and less than 70 mg/dL in adults with known CAD or diabetes mellitus (a goal that is seldom attained in most FH patients) (11).

TABLE 5.

GUIDELINE RECOMMENDATIONS FOR TREATING FH

NLA Expert panel on pediatric FH(80)AHA/ACC 2018 cholesterol guideline(81)NLA Expert panel on adult FH(82)EAS guideline on FH(11)
Age to initiate treatment≥ 8 years (earlier in special cases i.e., HoFH)≥ 20 yearsNot specified – “After a confirmatory diagnosis of FH …adult FH patients
should receive initial treatment”
≥ 8 years
Agent recommendedStatins are preferred
Non-statin options (ezetimibe, BAS, fibrates, niacin) are discussed but not routinely recommended due to lack of FDA approval or adverse drug events
Statins are preferred
Non-statin options (ezetimibe, BAS, PCSK9i monoclonal antibodies) are also recommended as add on therapy
Statins are preferred
Non-statin options (ezetimibe, BAS, niacin) or LDL apheresis are also recommended as add on therapy or in statin intolerant patients
Statins are preferred
Non-statin options (ezetimibe, BAS) or LDL apheresis are also recommended as add on therapy or in statin intolerant patients
Goal of therapy≥ 50% reduction in LDL-C or LDL-C < 130 mg/dL≥ 50% reduction in LDL-C or LDL-C < 100 mg/dL≥ 50% reduction in LDL-CLDL-C < 135 mg/dL in pediatrics, < 100 mg/dL in adults, < 70 mg/dL in adults with CAD or diabetes

AHA / ACC: American Heart Association / American College of Cardiology; EAS: European Atherosclerosis Society; NLA: National Lipid Association; BAS: bile-acid sequestrants

In our view in FH patients without clinical ASCVD one should aim for an LDL-C level <100 mg/dL (2.59 mMol/L). In FH patients with clinical ASCVD the goal, at a minimum should be an LDL-C level <70 mg/dL (1.81 mMol/L) with many experts recommending LDL-C levels <55 mg/dL (1.4 2mMol/L), particularly when patients have additional risk factors (acute coronary syndrome, diabetes, polyvascular disease, etc.) (83,84). In patients without ASCVD but who are at high risk due to other risk factors such as diabetes, Lp(a) >50 mg/dL, smokers, a strong family history of premature ASCVD, etc. many experts would recommend an LDL-C goal of <70 mg/dL (1.81 mMol/L). In patients with FH it may be difficult to achieve these goals.

Treatment of Patients with Heterozygous Familial Hypercholesterolemia

As with almost all metabolic disorder we should encourage the patient to follow a lifestyle that will reduce disease manifestations. However, lifestyle changes are very rarely sufficient to lower LDL-C levels to the desired range in patients with FH and therefore cholesterol lowering drugs will be required (see figure 2 and tables 6 and 7) (for detailed information on cholesterol lowering drugs see the chapter on cholesterol lowering drugs (85). In patients with HeFH, statins are a mainstay of treatment, despite the dearth of randomized clinical trials of statin efficacy in this special population. Statins are FDA approved for use in pediatric FH patients beginning at age 8-10 years for HeFH and in the first year of life or at initial diagnosis for HoFH (46,68,83,86-90). Data from longitudinal observational studies suggest statin initiation in childhood is both safe and effective, reducing LDL-C burden and corresponding atherosclerosis rates over follow-up of up to 20 years (46,91). The major early statin trials (4S and WOSCOPS) likely had study populations that were enriched with FH patients, given that mean baseline LDL-C ranged from 189 to 216 mg/dL (92,93). Patients should be treated with atorvastatin 40-80 mg per day or rosuvastatin 20-40 mg per day (i.e., high-intensity statin therapy). As monotherapy, statins can reduce LDL-C by up to 60% in HeFH patients but typically display a blunted response (10-25%) in HoFH patients depending on LDLR functionality (68,94). The vast majority of patients, however, require additional pharmacotherapy. When intensive statin therapy does not result in an LDL-C level in the desired range additional therapy should be added. Given that ezetimibe is now generic, relatively inexpensive, well tolerated, and has evidence for ASCVD risk reduction in a large cardiovascular outcome trial, this is frequently the next drug used (95). One can anticipate that ezetimibe 10 mg per day added to intensive statin therapy will result in an approximate 20% further reduction in LDL-C. If this is not sufficient one can then use a PCSK9i monoclonal antibody to achieve the desired cholesterol goal. Adding a PCSK9i monoclonal antibody will result in a further 50-60% decrease in LDL-C levels and in most patients will result in LDL-C levels in the desired range. In some instances, if the LDL-C level is far from goal (>25%) on intensive statin therapy one can skip treating with ezetimibe and proceed directly to adding a PCSK9i monoclonal antibody. Bempedoic acid (discussed in more detail below) is an acceptable third- or fourth-line add-on agent if additional LDL-C lowering is needed. In certain instances, bile resin binders may be useful in the treatment of FH (for example pregnant and lactating patients).

FIGURE 2.

FIGURE 2.

Approach to the Pharmacologic Treatment of Patients with Heterozygous Familial Hypercholesterolemia

PCSK9 inhibitors consist of two therapeutic modalities, 1) monoclonal antibodies (alirocumab and evolocumab) that work extracellularly to sequester the PCSK9 protein, and 2) inclisiran, a novel, small interfering ribonucleic acid (siRNA)-based gene-silencing technology that inhibits mRNA translation and intracellular production of PCSK9 by the liver.(96)

The effect of PCSK9i monoclonal antibodies in patients with HeFH has been extensively studied and consistently demonstrate potent LDL-C lowering on the order of 50-60% (58,97-111). An analysis of the ODYSSEY trials (FH I, FH II, LONG TERM, and HIGH FH) evaluated alirocumab use (75 or 150 mg subcutaneously every 14 days) in 1,257 HeFH patients. The primary endpoint was LDL-C at 24 weeks; alirocumab resulted in a more than 55% reduction in LDL-C compared with placebo (107). In another trial, alirocumab 150 mg every 14 days in 62 apheresis patients reduced the primary endpoint, rate of apheresis treatments over 12 weeks, by 75%, with 63.4% of patients completely discontinuing apheresis treatments due to well controlled LDL-C values (106). The RUTHERFORD-2 trial, evaluated more than 300 patients with HeFH randomized to evolocumab (140 mg subcutaneously every 2 weeks or 420 mg subcutaneously monthly) versus placebo. At both dosing regimens, evolocumab resulted in significantly reduced LDL-C at 12 weeks compared with placebo (>60%) (58). In all trials, PCSK9i monoclonal antibodies were well tolerated, with most demonstrating treatment-emergent adverse events (TEAEs) similar to placebo. Clinically, the most common adverse events include: injection site reactions, mild cold or flu-like symptoms, nasopharyngitis, and myalgias (112). Thus, PCSK9i monoclonal antibodies are very effective at lowering LDL cholesterol levels and safe in HeFH patients.

Additionally, there are two cardiovascular outcomes trials that evaluated the FDA approved monoclonal antibodies targeting PCSK9. The FOURIER trial evaluated almost 28,000 subjects with stable vascular disease (CAD, stroke, peripheral arterial disease) on optimized statin therapy and randomized them to either placebo or evolocumab. Evolocumab therapy was associated with a 60% reduction in LDL-C and a 15% reduction in the primary 5-point major adverse cardiovascular outcome endpoint (113). ODYSSEY OUTCOMES enrolled approximately 18,000 subjects with recent acute coronary syndrome (1-12 months prior to enrollment in the trial) who were on high-intensity statin therapy at baseline and randomized them to placebo or alirocumab. Alirocumab was also associated with a 15% reduction in the primary outcome, in this case a 4-point composite of major adverse cardiovascular events (114). Major guidelines, consensus documents, and expert recommendations suggest consideration of a PCSK9 inhibitor (in addition to background statin +/- ezetimibe therapy) in high-risk patients with established ASCVD and/or in patients with severe hypercholesterolemia when LDL-C values exceed 70 or 100 mg/dL respectively (81,83,84,115-118). Both monoclonal antibodies have an FDA labeled indication for ASCVD risk reduction in patients with established ASCVD.

Inclisiran has been thoroughly studied in the ORION clinical development program, a series of phase 1 to 3 trials designed to investigate the pharmacokinetics, pharmacodynamics, optimal dose, efficacy, and safety of inclisiran in specific populations (96). ORION-9 was a phase 3, randomized, double-blinded, placebo-controlled trial evaluating use of inclisiran 300 mg given subcutaneously on days 1, 90, 270, and 540, in 482 HeFH patients with baseline LDL-C of ≥100 mg/dL(119). Treatment with inclisiran produced a placebo-corrected LDL-C reduction of 47.9% (P<0.001) at day 510. Response based on genotype was as expected, with LDL-C reductions of 41.2% to 59.2% for all except PCSK9 GOF variants which were associated with dramatic LDL-C reduction of 89.7%. Pre-specified exploratory cardiovascular event (CV death, cardiac arrest, non-fatal MI, or non-fatal stroke) rates were comparable in both treatment groups (4.1% inclisiran vs. 4.2% placebo). Inclisiran was tolerated well during the trial with adverse events similar between treatment groups, with most adverse reactions mild to moderate in nature. The most common adverse event was injection site reactions, which was 10-fold higher in the inclisiran group (17%) compared to placebo (1.7%), however, 90% of these were graded as mild severity. Future studies will evaluate inclisiran in an adolescent HeFH population in the ORION-16 trial (120). Inclisiran remains experimental in the US but was approved for clinical use in the European Union (EU) in December 2020 (121). Though PCSK9 inhibition by monoclonal antibodies and inclisiran are comparable in LDL-C reducing capacity, there are important differences to acknowledge 1) inclisiran has a long biological half-life allowing for twice yearly dosing and possible medication adherence advantages, 2) PCSK9i monoclonal antibodies are administered by patients in their home and inclisiran is to be administered by a healthcare professional in a healthcare setting, and 3) while PCSK9i monoclonal antibodies have robust data proving reduction in ASCVD events, clinical outcome trials with inclisiran are still in progress.

Bempedoic acid, a novel inhibitor of adenosine triphosphate (ATP)-citrate lyase (ACL), an enzyme upstream from 3–hydroxy–3–methylglutaryl coenzyme A (HMG–CoA) in the cholesterol synthesis pathway, is the newest orally administered lipid-lowering therapy (122). Bempedoic acid is a prodrug, that is converted into its active form (bempedoyl-CoA) by very long-chain acyl-CoA synthetase 1 (ACSVL1), which is expressed in hepatocytes but is undetectable in muscle. Development of this agent was designed to circumvent the myotoxicity commonly associated with historical lipid-lowering therapies, primarily statins. Bempedoic acid was FDA approved on February 21, 2020 for use in patients with established ASCVD and/or HeFH who require additional LDL-C lowering as an adjunct to dietary intervention and maximally tolerated statin therapy (123). Bempedoic acid was evaluated in over 3600 patients in the phase 3 CLEAR program trials, of which, only 3.7% were HeFH patients (122). A pooled analysis from two phase 3 trials, CLEAR Harmony and CLEAR Wisdom, demonstrated a similarly modest placebo-corrected LDL-C reduction at 12 weeks of bempedoic acid treatment in the HeFH cohort (22.3%) as compared to the overall population (18.3%) (124). Overall bempedoic acid was well tolerated in the phase 3 trials but occurrence of TEAEs was higher in the HeFH cohort compared to those without HeFH but was not increased with bempedoic acid versus placebo. In a real-world analysis of bempedoic acid, which was enriched in patients with HeFH (64%) and statin intolerance (74%), the drug was associated with clinically meaningful LDL-C lowering (mean reductions 20.3% to 36.7%), but high rates of TEAEs (50%) and drug discontinuations (35.9%) (unpublished observations – manuscript under submission).

Treatment of Patients with Homozygous Familial Hypercholesterolemia

The initial therapy in patients with HoFH is identical to the treatment of patients with HeFH. However, statins and ezetimibe may prove relatively ineffective in the treatment of HoFH because the mode of action of these drugs largely depends on the upregulation of functional LDLR in the liver. In HoFH, measurable activity of both copies of the LDLR is absent or greatly reduced (69). Drug-induced LDL-C lowering is diminished by more than 50% comparing HeFH to HoFH – high intensity statins lower LDL-C 50-60% vs 10-25% and ezetimibe lowers LDL-C by 15-25% vs <10%, respectively (see table 6) (125).

In patients with HoFH, response to PCSK9i monoclonal antibodies varies depending on the specific gene defect. In the TESLA-B trial, 49 patients with HoFH were treated with evolocumab or placebo every four weeks for 12 weeks. LDL-C in the evolocumab treatment arm was significantly reduced by almost 31% compared with placebo. In addition to overall reduction in LDL-C, the trial investigators examined the treatment effect by LDLR mutation status. They found that the response to evolocumab aligns with the genetic cause of HoFH, with a greater reduction in LDL-C observed in subjects with two LDL receptor-defective mutations (i.e., abnormal receptor functionality in both alleles) when compared with those patients with even just one LDL receptor-negative mutation (i.e., nonexistent receptor functionality in one allele). Evolocumab was well tolerated among the HoFH patients (126). Similar results were seen in other HoFH trials examining the PCSK9i monoclonal antibody (alirocumab) and over longer follow up durations, providing confirmation of durable LDL-C efficacy and safety with this therapeutic class (127-129).

Inclisiran, dosed as 300 mg subcutaneously on days 1 and 90 (or 104 if PCSK9 level was suppressed by >70%), was evaluated in 4 HoFH patients enrolled in the ORION-2 trial (130). Patients were ages 50, 46, 23, and 29 years, 50% female, and had baseline LDL-C 540 mg/dL, 547 mg/dL, 614 mg/dL, and 189 mg/dL, respectively. All had biallelic causative genetic variants in LDLR, and all were on high-intensity statins and ezetimibe. Inclisiran-induced LDL-C reductions ranged 11.7% to 33.1% at day 90, and 17.5% to 37.0% at day 180 in three of the four patients, similar results to prior trials of PCSK9i monoclonal antibodies in HoFH patients with residual LDLR function. One patient (LDLR c.681C/G [defective]) who had a history of hypo-responsiveness to both PCSK9i monoclonal antibodies (<20% lowering), also exhibited no change in LDL-C with inclisiran treatment. No adverse events were recorded over a 10 month follow up. Future studies will evaluate inclisiran in a larger set of adult (ORION-5), and adolescent (ORION-13) HoFH patients (131,132).

Standard triple drug therapy (statin, ezetimibe, PCSK9 inhibitor) often does not result in a sufficient lowering of LDL-C due to the combination of the very high baseline LDL-C and the relative resistance of patients with HoFH to drug therapy. Novel agents approved specifically for the treatment of severe hypercholesterolemia include microsomal triglyceride transfer protein (MTP) inhibitors and apoB-100 antisense oligonucleotides (ASO). MTP is involved in the transfer of lipid droplets to apoB as well as assembly and secretion of apoB-containing lipoproteins in the liver and gut. MTP inhibition thus reduces production and secretion of chylomicrons and VLDL-C. In one study, 29 patients with HoFH were treated with the MTP inhibitor lomitapide for 26 weeks and were followed until week 78. Average LDL-C reductions were 50% (to 166 mg/dl) at week 26, 44% (to 197 mg/dL) at week 52, and 38% (to 208 mg/dL) at week 78 (133). Though lomitapide displays potent LDL-C lowering capacity, use is limited as a result of a significant side effect profile consisting of severe gastrointestinal complications and hepatotoxicity risk, as well as high medication cost.

Anti-sense oligonucleotide (ASO) molecules bind to specific mRNAs and target them for degradation, reducing protein synthesis in the process. Mipomersen, which was removed from the market in June 2018, is an ASO that binds to apoB-100 mRNA and thus prevents the formation of apoB-100. Mipomersen results in decreased synthesis of apoB-containing lipoproteins, mostly VLDL-C, eventually leading to a drastic reduction of LDL-C levels in plasma. In one trial, 51 patients with either genetically-defined HoFH, untreated LDL-C levels of >500 mg/dL plus xanthomas, or evidence of HeFH in both parents were randomized to placebo versus mipomersen for a treatment duration of 26 weeks. In the placebo group, baseline LDL-C was 402 mg/dL and declined to 390 mg/dL; in the treatment group, baseline LDL-C dropped from 440 mg/dl to 324 mg/dL (134).

Evinacumab is the newest addition to the lipid-lowering treatment armamentarium, receiving FDA approval in February 2021 as an adjunct to other LDL-lowering therapies for treatment of adults and pediatric patients ≥12 years of age with HoFH (125). Evinacumab is a fully human IgG4 isotype monoclonal antibody targeting a novel, non-LDLR pathway for LDL-C lowering by inhibiting angiopoietin-like 3 (ANGPTL3) (135). ANGPTL3 inhibition overrides the inhibitory effect on lipoprotein lipase (LPL) and endothelial lipase (EL) increasing the activity of these enzymes, producing a panlipid-lowering effect of apo-B containing lipoproteins (with the exception Lp[a]), reducing lipoproteins by approximately 50% from baseline (125). Specifically, evinacumab promotes LDL-C lowering through EL-dependent VLDL-C processing and clearance by LDLR independent pathways, thereby decreasing formation of LDL-C from VLDL-C.(136) The magnitude of LDL-C lowering was seen in both HoFH and HeFH populations and was similar regardless of genotype. Serial coronary computed tomography angiography (CCTA) evaluations of two HoFH patients enrolled in the ELIPSE HoFH trial demonstrated reductions in coronary total plaque volume (TPV) of 76% to 85% over 10 months of treatment with evinacumab (137). Evinacumab is dosed at 15 mg/kg of body weight over a 60 minute intravenous infusion (125). The drug is well tolerated with adverse effects being infrequent, mild, and transient, consisting primarily of injection site reactions, flu/cold-like symptoms, pain, and fatigue. Long-term safety of evinacumab remains unknown. Evinacumab is poised to play a significant role in HoFH management as it offers one of the most potent LDL-C lowering capabilities among existing treatments. Use in HoFH will likely be as an add-on therapy after high-intensity statin, ezetimibe, and PCSK9 inhibition. However, potential barriers to use may include high cost, intravenous administration, and requirement for administration in a healthcare or home infusion setting. Future use of evinacumab may extend beyond HoFH and could entail use in other therapeutic areas such as HeFH and severe hypertriglyceridemia.

In patients in which drug therapy is either not successful at lowering LDL-C or not well tolerated one can consider lipoprotein apheresis. The FDA has approved lipoprotein apheresis for subjects with CVD and LDL-C >200 mg/dL or without CVD and LDL-C >300 mg/dL (138). Recently, this threshold has been moved to 160 mg/dL, thus increasing the target population for cholesterol dialysis at a time when arrival of stronger medications is curtailing patient entry into this therapeutic program. The process, which involves removing apoB-containing lipoproteins from plasma, is usually performed every two weeks and results in a 60-70% reduction of LDL-C and Lp(a) in the immediate post-procedure period, with time-averaged reductions of 20-50%. Levels tend to revert to baseline within two weeks. For more detailed information on lipoprotein apheresis see the Endotext chapter on this topic (139).

TABLE 6.

COMMON LIPID-LOWERING TREATMENTS FOR FH

AgentNiacinBASFibrateStatinEzetimibe
FDA approval date1997*19731981†
1993
19872002
AdministrationPOPOPOPOPO
DosingDailyDailyDailyDailyDaily
LDL-C lowering
(HeFH)
10%-25%15%-30%10%-20%20%-60%15%-25%
LDL-C lowering
(HoFH)
<10%<10%<10%10%-25%<10%
Lp(a) lowering20-30%N/AN/AN/AN/A
Relative cost+/+++++++
Safety concernsFlushing, moderate GI intolerance (abd pain, nausea, vomiting, peptic ulcer), hyperglycemia, gout, hepatotoxicityModerate GI symptoms (abd pain, constipation, bloating, nausea), hypertriglyceridemia, fat-soluble vitamin deficienciesMyalgia, mild GI symptoms (abd pain, diarrhea), cholelithiasis, increased LFTLFT elevation, myalgia, DM riskMild GI symptoms (loose stool, diarrhea, cramping), myalgia, increased LFT
Other consider-ationsHigh pill burden, separate from other meds (binding)Primarily used for triglyceride lowering, renal dose adjustmentsUsually well tolerated

Adapted from Warden BA, et al. Expert Rev Cardiovasc Ther. 2021;1-13.

*

Approval date is for niacin extended-release. Niacin has been used clinically for hypercholesterolemia since the 1950’s

Gemfibrozil FDA approved in 1981, fenofibrate in 1993

TABLE 7.

ADVANCED LIPID-LOWERING TREATMENTS FOR FH

AgentLipoprotein
apheresis
LomitapidePCSK9i*Bempedoic acidEvinacumab
FDA approval date19962012201520202021
AdministrationIVPOSQPOIV
Dosing2-4x monthlyDaily1-2x monthlyDailyMonthly
LDL-C lowering
(HeFH)
50-85% (acute)
23-50%
(time-average)
N/A50%-60%15%-25%49%
LDL-C lowering
(HoFH)
50-85% (acute)
23-50%
(time-average)
20%-50%zero-30%†Unknown49%
Lp(a) lowering50-75% (acute)
20-40%
(time-average)
zero-30%20-30%N/AN/A
Relative cost+++++++++++++/++++++++
Safety concernsIV access issues, hypotension, vasovagal episodes, fatigue, bleeding, hypocalcemia, anemiaSevere GI intolerance (diarrhea, nausea, vomiting, abd pain, cramping), fat malabsorption, hepatic steatosis,
hepatotoxicity (REMS)
ISR, flu/cold-like symptomsMild GI symptoms (diarrhea, abd discomfort), gout, tendon injury, transient lab changes (SCr, BUN, LFT, Hgb, HCT, uric acid)Nasopharyngitis, ISR, flu-like symptoms, fatigue, pain, headache, rare hypersensitivity reaction
Other consider-ationsLengthy and frequent treatments, need for patient travel, DDI (heparin, ACEi)Significant DDI, renal dose adjustmentsAccess and cost issuesImproves glycaemia, option for patients with SAMSReduces all non-Lp(a) apoB-containing lipoproteins

Adapted from Warden BA, et al. Expert Rev Cardiovasc Ther. 2021;1-13.

*

PCSK9i: represent PCSK9 blocking monoclonal antibodies approved in the United States. Inclisiran, a novel small interfering ribonucleic acid (siRNA)-based medication that reduces PCSK9 production, was approved in Europe as of December 2020.

Response aligns with genetic cause of HoFH with no reduction was seen in those with null-null LDLR mutations

REFERENCES

1.
Brown MS, Goldstein JL. A receptor-mediated pathway for cholesterol homeostasis. Science. 1986;232:34–47. [PubMed: 3513311]
2.
Berberich AJ, Hegele RA. The complex molecular genetics of familial hypercholesterolaemia. Nat Rev Cardiol. 2019;16:9–20. [PubMed: 29973710]
3.
Fouchier SW, Dallinga-Thie GM, Meijers JC, Zelcer N, Kastelein JJ, Defesche JC, Hovingh GK. Mutations in STAP1 are associated with autosomal dominant hypercholesterolemia. Circ Res. 2014;115:552–555. [PubMed: 25035151]
4.
Kanuri B, Fong V, Haller A, Hui DY, Patel SB. Mice lacking global Stap1 expression do not manifest hypercholesterolemia. BMC Med Genet. 2020;21:234. [PMC free article: PMC7685646] [PubMed: 33228548]
5.
Mohebi R, Chen Q, Hegele RA, Rosenson RS. Failure of cosegregation between a rare STAP1 missense variant and hypercholesterolemia. J Clin Lipidol. 2020;14:636–638. [PubMed: 32828708]
6.
Loaiza N, Hartgers ML, Reeskamp LF, Balder JW, Rimbert A, Bazioti V, Wolters JC, Winkelmeijer M, Jansen HPG, Dallinga-Thie GM, Volta A, Huijkman N, Smit M, Kloosterhuis N, Koster M, Svendsen AF, van de Sluis B, Hovingh GK, Grefhorst A, Kuivenhoven JA. Taking One Step Back in Familial Hypercholesterolemia: STAP1 Does Not Alter Plasma LDL (Low-Density Lipoprotein) Cholesterol in Mice and Humans. Arterioscler Thromb Vasc Biol. 2020;40:973–985. [PMC free article: PMC7098433] [PubMed: 31996024]
7.
Lamiquiz-Moneo I, Restrepo-Córdoba MA, Mateo-Gallego R, Bea AM, Del Pino Alberiche-Ruano M, García-Pavía P, Cenarro A, Martín C, Civeira F, Sánchez-Hernández RM. Predicted pathogenic mutations in STAP1 are not associated with clinically defined familial hypercholesterolemia. Atherosclerosis. 2020;292:143–151. [PubMed: 31809983]
8.
Hopkins PN, Toth PP, Ballantyne CM, Rader DJ. Familial hypercholesterolemias: prevalence, genetics, diagnosis and screening recommendations from the National Lipid Association Expert Panel on Familial Hypercholesterolemia. J Clin Lipidol. 2011;5:S9–17. [PubMed: 21600530]
9.
Abul-Husn NS, Manickam K, Jones LK, Wright EA, Hartzel DN, Gonzaga-Jauregui C, O'Dushlaine C, Leader JB, Lester Kirchner H, Lindbuchler DM, Barr ML, Giovanni MA, Ritchie MD, Overton JD, Reid JG, Metpally RP, Wardeh AH, Borecki IB, Yancopoulos GD, Baras A, Shuldiner AR, Gottesman O, Ledbetter DH, Carey DJ, Dewey FE, Murray MF. Genetic identification of familial hypercholesterolemia within a single U.S. health care system. Science. 2016:354. [PubMed: 28008010]
10.
Khera AV, Won HH, Peloso GM, Lawson KS, Bartz TM, Deng X, van Leeuwen EM, Natarajan P, Emdin CA, Bick AG, Morrison AC, Brody JA, Gupta N, Nomura A, Kessler T, Duga S, Bis JC, van Duijn CM, Cupples LA, Psaty B, Rader DJ, Danesh J, Schunkert H, McPherson R, Farrall M, Watkins H, Lander E, Wilson JG, Correa A, Boerwinkle E, Merlini PA, Ardissino D, Saleheen D, Gabriel S, Kathiresan S. Diagnostic Yield and Clinical Utility of Sequencing Familial Hypercholesterolemia Genes in Patients With Severe Hypercholesterolemia. J Am Coll Cardiol. 2016;67:2578–2589. [PMC free article: PMC5405769] [PubMed: 27050191]
11.
Nordestgaard BG, Chapman MJ, Humphries SE, Ginsberg HN, Masana L, Descamps OS, Wiklund O, Hegele RA, Raal FJ, Defesche JC, Wiegman A, Santos RD, Watts GF, Parhofer KG, Hovingh GK, Kovanen PT, Boileau C, Averna M, Boren J, Bruckert E, Catapano AL, Kuivenhoven JA, Pajukanta P, Ray K, Stalenhoef AF, Stroes E, Taskinen MR, Tybjaerg-Hansen A. Familial hypercholesterolaemia is underdiagnosed and undertreated in the general population: guidance for clinicians to prevent coronary heart disease: consensus statement of the European Atherosclerosis Society. Eur Heart J. 2013;34:3478–3490a. [PMC free article: PMC3844152] [PubMed: 23956253]
12.
Knowles JW, Rader DJ, Khoury MJ. Cascade Screening for Familial Hypercholesterolemia and the Use of Genetic Testing. JAMA. 2017;318:381–382. [PMC free article: PMC6166431] [PubMed: 28742895]
13.
Haase A, Goldberg AC. Identification of people with heterozygous familial hypercholesterolemia. Curr Opin Lipidol. 2012;23:282–289. [PubMed: 22801386]
14.
Sniderman AD, Tsimikas S, Fazio S. The severe hypercholesterolemia phenotype: clinical diagnosis, management, and emerging therapies. J Am Coll Cardiol. 2014;63:1935–1947. [PubMed: 24632267]
15.
Tsimikas S, Hall JL. Lipoprotein(a) as a potential causal genetic risk factor of cardiovascular disease: a rationale for increased efforts to understand its pathophysiology and develop targeted therapies. J Am Coll Cardiol. 2012;60:716–721. [PubMed: 22898069]
16.
Kraft HG, Lingenhel A, Raal FJ, Hohenegger M, Utermann G. Lipoprotein(a) in homozygous familial hypercholesterolemia. Arterioscler Thromb Vasc Biol. 2000;20:522–528. [PubMed: 10669652]
17.
Rader DJ, Mann WA, Cain W, Kraft HG, Usher D, Zech LA, Hoeg JM, Davignon J, Lupien P, Grossman M, et al. The low density lipoprotein receptor is not required for normal catabolism of Lp(a) in humans. J Clin Invest. 1995;95:1403–1408. [PMC free article: PMC441483] [PubMed: 7883987]
18.
Cain WJ, Millar JS, Himebauch AS, Tietge UJ, Maugeais C, Usher D, Rader DJ. Lipoprotein [a] is cleared from the plasma primarily by the liver in a process mediated by apolipoprotein. J Lipid Res. 2005;46:2681–2691. [a] [PubMed: 16150825]
19.
Tsimikas S, Fazio S, Ferdinand KC, Ginsberg HN, Koschinsky ML, Marcovina SM, Moriarty PM, Rader DJ, Remaley AT, Reyes-Soffer G, Santos RD, Thanassoulis G, Witztum JL, Danthi S, Olive M, Liu L. NHLBI Working Group Recommendations to Reduce Lipoprotein(a)-Mediated Risk of Cardiovascular Disease and Aortic Stenosis. J Am Coll Cardiol. 2018;71:177–192. [PMC free article: PMC5868960] [PubMed: 29325642]
20.
Slimane MN, Pousse H, Maatoug F, Hammami M, Ben Farhat MH. Phenotypic expression of familial hypercholesterolaemia in central and southern Tunisia. Atherosclerosis. 1993;104:153–158. [PubMed: 8141839]
21.
Goldberg AC, Hopkins PN, Toth PP, Ballantyne CM, Rader DJ, Robinson JG, Daniels SR, Gidding SS, de Ferranti SD, Ito MK, McGowan MP, Moriarty PM, Cromwell WC, Ross JL, Ziajka PE. Familial hypercholesterolemia: screening, diagnosis and management of pediatric and adult patients: clinical guidance from the National Lipid Association Expert Panel on Familial Hypercholesterolemia. J Clin Lipidol. 2011;5:S1–8. [PubMed: 21600525]
22.
Calandra S, Tarugi P, Speedy HE, Dean AF, Bertolini S, Shoulders CC. Mechanisms and genetic determinants regulating sterol absorption, circulating LDL levels, and sterol elimination: implications for classification and disease risk. J Lipid Res. 2011;52:1885–1926. [PMC free article: PMC3284125] [PubMed: 21862702]
23.
Innerarity TL, Mahley RW, Weisgraber KH, Bersot TP, Krauss RM, Vega GL, Grundy SM, Friedl W, Davignon J, McCarthy BJ. Familial defective apolipoprotein B-100: a mutation of apolipoprotein B that causes hypercholesterolemia. J Lipid Res. 1990;31:1337–1349. [PubMed: 2280177]
24.
Lambert G, Sjouke B, Choque B, Kastelein JJ, Hovingh GK. The PCSK9 decade. J Lipid Res. 2012;53:2515–2524. [PMC free article: PMC3494258] [PubMed: 22811413]
25.
Cohen JC, Boerwinkle E, Mosley TH Jr, Hobbs HH. Sequence variations in PCSK9, low LDL, and protection against coronary heart disease. N Engl J Med. 2006;354:1264–1272. [PubMed: 16554528]
26.
Zhao Z, Tuakli-Wosornu Y, Lagace TA, Kinch L, Grishin NV, Horton JD, Cohen JC, Hobbs HH. Molecular characterization of loss-of-function mutations in PCSK9 and identification of a compound heterozygote. Am J Hum Genet. 2006;79:514–523. [PMC free article: PMC1559532] [PubMed: 16909389]
27.
Hooper AJ, Marais AD, Tanyanyiwa DM, Burnett JR. The C679X mutation in PCSK9 is present and lowers blood cholesterol in a Southern African population. Atherosclerosis. 2007;193:445–448. [PubMed: 16989838]
28.
Awan Z, Choi HY, Stitziel N, Ruel I, Bamimore MA, Husa R, Gagnon MH, Wang RH, Peloso GM, Hegele RA, Seidah NG, Kathiresan S, Genest J. APOE p.Leu167del mutation in familial hypercholesterolemia. Atherosclerosis. 2013;231:218–222. [PubMed: 24267230]
29.
Pullinger CR, Eng C, Salen G, Shefer S, Batta AK, Erickson SK, Verhagen A, Rivera CR, Mulvihill SJ, Malloy MJ, Kane JP. Human cholesterol 7alpha-hydroxylase (CYP7A1) deficiency has a hypercholesterolemic phenotype. J Clin Invest. 2002;110:109–117. [PMC free article: PMC151029] [PubMed: 12093894]
30.
Ahmad Z, Adams-Huet B, Chen C, Garg A. Low prevalence of mutations in known loci for autosomal dominant hypercholesterolemia in a multiethnic patient cohort. Circ Cardiovasc Genet. 2012;5:666–675. [PMC free article: PMC3774009] [PubMed: 23064986]
31.
Sturm AC, Knowles JW, Gidding SS, Ahmad ZS, Ahmed CD, Ballantyne CM, Baum SJ, Bourbon M, Carrie A, Cuchel M, de Ferranti SD, Defesche JC, Freiberger T, Hershberger RE, Hovingh GK, Karayan L, Kastelein JJP, Kindt I, Lane SR, Leigh SE, Linton MF, Mata P, Neal WA, Nordestgaard BG, Santos RD, Harada-Shiba M, Sijbrands EJ, Stitziel NO, Yamashita S, Wilemon KA, Ledbetter DH, Rader DJ. Clinical Genetic Testing for Familial Hypercholesterolemia: JACC Scientific Expert Panel. J Am Coll Cardiol. 2018;72:662–680. [PubMed: 30071997]
32.
Wang J, Dron JS, Ban MR, Robinson JF, McIntyre AD, Alazzam M, Zhao PJ, Dilliott AA, Cao H, Huff MW, Rhainds D, Low-Kam C, Dube MP, Lettre G, Tardif JC, Hegele RA. Polygenic Versus Monogenic Causes of Hypercholesterolemia Ascertained Clinically. Arterioscler Thromb Vasc Biol. 2016;36:2439–2445. [PubMed: 27765764]
33.
Stein EA, Raal FJ. Polygenic familial hypercholesterolaemia: does it matter? Lancet. 2013;381:1255–1257. [PubMed: 23433574]
34.
Brown EE, Sturm AC, Cuchel M, Braun LT, Duell PB, Underberg JA, Jacobson TA, Hegele RA. Genetic testing in dyslipidemia: A scientific statement from the National Lipid Association. J Clin Lipidol. 2020;14:398–413. [PubMed: 32507592]
35.
Harders-Spengel K, Wood CB, Thompson GR, Myant NB, Soutar AK. Difference in saturable binding of low density lipoprotein to liver membranes from normocholesterolemic subjects and patients with heterozygous familial hypercholesterolemia. Proc Natl Acad Sci U S A. 1982;79:6355–6359. [PMC free article: PMC347120] [PubMed: 6292899]
36.
Havekes LM, Verboom H, de Wit E, Yap SH, Princen HM. Regulation of low density lipoprotein receptor activity in primary cultures of human hepatocytes by serum lipoproteins. Hepatology. 1986;6:1356–1360. [PubMed: 3793011]
37.
Zelcer N, Hong C, Boyadjian R, Tontonoz P. LXR regulates cholesterol uptake through Idol-dependent ubiquitination of the LDL receptor. Science. 2009;325:100–104. [PMC free article: PMC2777523] [PubMed: 19520913]
38.
Goldstein JL, Schrott HG, Hazzard WR, Bierman EL, Motulsky AG. Hyperlipidemia in coronary heart disease. II. Genetic analysis of lipid levels in 176 families and delineation of a new inherited disorder, combined hyperlipidemia. J Clin Invest. 1973;52:1544–1568. [PMC free article: PMC302426] [PubMed: 4718953]
39.
Bilheimer DW, Stone NJ, Grundy SM. Metabolic studies in familial hypercholesterolemia. Evidence for a gene-dosage effect in vivo. J Clin Invest. 1979;64:524–533. [PMC free article: PMC372147] [PubMed: 222811]
40.
Kesaniemi YA, Grundy SM. Significance of low density lipoprotein production in the regulations of plasma cholesterol level in man. J Clin Invest. 1982;70:13–22. [PMC free article: PMC370220] [PubMed: 7085881]
41.
Real JT, Chaves FJ, Ejarque I, Garcia-Garcia AB, Valldecabres C, Ascaso JF, Armengod ME, Carmena R. Influence of LDL receptor gene mutations and the R3500Q mutation of the apoB gene on lipoprotein phenotype of familial hypercholesterolemic patients from a South European population. Eur J Hum Genet. 2003;11:959–965. [PubMed: 14508510]
42.
Ejarque I, Real JT, Martinez-Hervas S, Chaves FJ, Blesa S, Garcia-Garcia AB, Millan E, Ascaso JF, Carmena R. Evaluation of clinical diagnosis criteria of familial ligand defective apoB 100 and lipoprotein phenotype comparison between LDL receptor gene mutations affecting ligand-binding domain and the R3500Q mutation of the apoB gene in patients from a South European population. Transl Res. 2008;151:162–167. [PubMed: 18279815]
43.
Pimstone SN, Defesche JC, Clee SM, Bakker HD, Hayden MR, Kastelein JJ. Differences in the phenotype between children with familial defective apolipoprotein B-100 and familial hypercholesterolemia. Arterioscler Thromb Vasc Biol. 1997;17:826–833. [PubMed: 9157944]
44.
Miserez AR, Keller U. Differences in the phenotypic characteristics of subjects with familial defective apolipoprotein B-100 and familial hypercholesterolemia. Arterioscler Thromb Vasc Biol. 1995;15:1719–1729. [PubMed: 7583549]
45.
Defesche JC, Pricker KL, Hayden MR, van der Ende BE, Kastelein JJ. Familial defective apolipoprotein B-100 is clinically indistinguishable from familial hypercholesterolemia. Arch Intern Med. 1993;153:2349–2356. [PubMed: 8215738]
46.
Gidding SS, Champagne MA, de Ferranti SD, Defesche J, Ito MK, Knowles JW, McCrindle B, Raal F, Rader D, Santos RD, Lopes-Virella M, Watts GF, Wierzbicki AS. The Agenda for Familial Hypercholesterolemia: A Scientific Statement From the American Heart Association. Circulation. 2015;132:2167–2192. [PubMed: 26510694]
47.
Garcia-Alvarez I, Castillo S, Mozas P, Tejedor D, Reyes G, Artieda M, Cenarro A, Alonso R, Mata P, Pocovi M, Civeira F. Rev Esp Cardiol. 2003;56:769–774. [Differences in clinical presentation between subjects with a phenotype of familial hypercholesterolemia determined by defects in the LDL-receptor and defects in Apo B-100] [PubMed: 12892621]
48.
Bourbon M, Alves AC, Alonso R, Mata N, Aguiar P, Padro T, Mata P. Mutational analysis and genotype-phenotype relation in familial hypercholesterolemia: The SAFEHEART registry. Atherosclerosis. 2017;262:8–13. [PubMed: 28475941]
49.
Talmud PJ, Shah S, Whittall R, Futema M, Howard P, Cooper JA, Harrison SC, Li K, Drenos F, Karpe F, Neil HA, Descamps OS, Langenberg C, Lench N, Kivimaki M, Whittaker J, Hingorani AD, Kumari M, Humphries SE. Use of low-density lipoprotein cholesterol gene score to distinguish patients with polygenic and monogenic familial hypercholesterolaemia: a case-control study. Lancet. 2013;381:1293–1301. [PubMed: 23433573]
50.
Ghaleb Y, Elbitar S, El Khoury P, Bruckert E, Carreau V, Carrie A, Moulin P, Di-Filippo M, Charriere S, Iliozer H, Farnier M, Luc G, Rabes JP, Boileau C, Abifadel M, Varret M. Usefulness of the genetic risk score to identify phenocopies in families with familial hypercholesterolemia? Eur J Hum Genet. 2018;26:570–578. [PMC free article: PMC5891487] [PubMed: 29374275]
51.
Kastelein JJ, van Leuven SI, Burgess L, Evans GW, Kuivenhoven JA, Barter PJ, Revkin JH, Grobbee DE, Riley WA, Shear CL, Duggan WT, Bots ML. Effect of torcetrapib on carotid atherosclerosis in familial hypercholesterolemia. N Engl J Med. 2007;356:1620–1630. [PubMed: 17387131]
52.
Kane JP, Malloy MJ, Tun P, Phillips NR, Freedman DD, Williams ML, Rowe JS, Havel RJ. Normalization of low-density-lipoprotein levels in heterozygous familial hypercholesterolemia with a combined drug regimen. N Engl J Med. 1981;304:251–258. [PubMed: 7003391]
53.
Kastelein JJ, Akdim F, Stroes ES, Zwinderman AH, Bots ML, Stalenhoef AF, Visseren FL, Sijbrands EJ, Trip MD, Stein EA, Gaudet D, Duivenvoorden R, Veltri EP, Marais AD, de Groot E. Simvastatin with or without ezetimibe in familial hypercholesterolemia. N Engl J Med. 2008;358:1431–1443. [PubMed: 18376000]
54.
Stein EA, Illingworth DR, Kwiterovich PO Jr, Liacouras CA, Siimes MA, Jacobson MS, Brewster TG, Hopkins P, Davidson M, Graham K, Arensman F, Knopp RH, DuJovne C, Williams CL, Isaacsohn JL, Jacobsen CA, Laskarzewski PM, Ames S, Gormley GJ. Efficacy and safety of lovastatin in adolescent males with heterozygous familial hypercholesterolemia: a randomized controlled trial. JAMA. 1999;281:137–144. [PubMed: 9917116]
55.
Pang J, David Marais A, Blom DJ, Brice BC, Silva PR, Jannes CE, Pereira AC, Hooper AJ, Ray KK, Santos RD, Watts GF. Heterozygous familial hypercholesterolaemia in specialist centres in South Africa, Australia and Brazil: Importance of early detection and lifestyle advice. Atherosclerosis. 2018;277:470–476. [PubMed: 30270087]
56.
Ridker PM, Rose LM, Kastelein JJP, Santos RD, Wei C, Revkin J, Yunis C, Tardif JC, Shear CL. Cardiovascular event reduction with PCSK9 inhibition among 1578 patients with familial hypercholesterolemia: Results from the SPIRE randomized trials of bococizumab. J Clin Lipidol. 2018;12:958–965. [PubMed: 29685591]
57.
Kastelein JJ, Ginsberg HN, Langslet G, Hovingh GK, Ceska R, Dufour R, Blom D, Civeira F, Krempf M, Lorenzato C, Zhao J, Pordy R, Baccara-Dinet MT, Gipe DA, Geiger MJ, Farnier M. ODYSSEY FH I and FH II: 78 week results with alirocumab treatment in 735 patients with heterozygous familial hypercholesterolaemia. Eur Heart J. 2015;36:2996–3003. [PMC free article: PMC4644253] [PubMed: 26330422]
58.
Raal FJ, Stein EA, Dufour R, Turner T, Civeira F, Burgess L, Langslet G, Scott R, Olsson AG, Sullivan D, Hovingh GK, Cariou B, Gouni-Berthold I, Somaratne R, Bridges I, Scott R, Wasserman SM, Gaudet D. PCSK9 inhibition with evolocumab (AMG 145) in heterozygous familial hypercholesterolaemia (RUTHERFORD-2): a randomised, double-blind, placebo-controlled trial. Lancet. 2015;385:331–340. [PubMed: 25282519]
59.
Feingold KR, Grunfeld C. Approach to the Patient with Dyslipidemia. In: De Groot LJ, Chrousos G, Dungan K, Feingold KR, Grossman A, Hershman JM, Koch C, Korbonits M, McLachlan R, New M, Purnell J, Rebar R, Singer F, Vinik A, eds. Endotext. South Dartmouth (MA): MDText.com, Inc.; 2000. [PubMed: 26561696]
60.
Santos RD. Phenotype vs. genotype in severe familial hypercholesterolemia: what matters most for the clinician? Curr Opin Lipidol. 2017;28:130–135. [PubMed: 28059950]
61.
Slack J. Risks of ischaemic heart-disease in familial hyperlipoproteinaemic states. Lancet. 1969;2:1380–1382. [PubMed: 4188273]
62.
Stone NJ, Levy RI, Fredrickson DS, Verter J. Coronary artery disease in 116 kindred with familial type II hyperlipoproteinemia. Circulation. 1974;49:476–488. [PubMed: 4813182]
63.
Civeira F, Castillo S, Alonso R, Merino-Ibarra E, Cenarro A, Artied M, Martin-Fuentes P, Ros E, Pocovi M, Mata P. Tendon xanthomas in familial hypercholesterolemia are associated with cardiovascular risk independently of the low-density lipoprotein receptor gene mutation. Arterioscler Thromb Vasc Biol. 2005;25:1960–1965. [PubMed: 16020744]
64.
Oosterveer DM, Versmissen J, Yazdanpanah M, Hamza TH, Sijbrands EJ. Differences in characteristics and risk of cardiovascular disease in familial hypercholesterolemia patients with and without tendon xanthomas: a systematic review and meta-analysis. Atherosclerosis. 2009;207:311–317. [PubMed: 19439299]
65.
Silva PR, Jannes CE, Marsiglia JD, Krieger JE, Santos RD, Pereira AC. Predictors of cardiovascular events after one year of molecular screening for Familial hypercholesterolemia. Atherosclerosis. 2016;250:144–150. [PubMed: 27214396]
66.
Perez de Isla L, Alonso R, Mata N, Fernandez-Perez C, Muniz O, Diaz-Diaz JL, Saltijeral A, Fuentes-Jimenez F, de Andres R, Zambon D, Piedecausa M, Cepeda JM, Mauri M, Galiana J, Brea A, Sanchez Munoz-Torrero JF, Padro T, Argueso R, Miramontes-Gonzalez JP, Badimon L, Santos RD, Watts GF, Mata P. Predicting Cardiovascular Events in Familial Hypercholesterolemia: The SAFEHEART Registry (Spanish Familial Hypercholesterolemia Cohort Study). Circulation. 2017;135:2133–2144. [PubMed: 28275165]
67.
Ahmad ZS, Andersen RL, Andersen LH, O'Brien EC, Kindt I, Shrader P, Vasandani C, Newman CB, deGoma EM, Baum SJ, Hemphill LC, Hudgins LC, Ahmed CD, Kullo IJ, Gidding SS, Duffy D, Neal W, Wilemon K, Roe MT, Rader DJ, Ballantyne CM, Linton MF, Duell PB, Shapiro MD, Moriarty PM, Knowles JW. US physician practices for diagnosing familial hypercholesterolemia: data from the CASCADE-FH registry. J Clin Lipidol. 2016;10:1223–1229. [PMC free article: PMC5381273] [PubMed: 27678440]
68.
Cuchel M, Bruckert E, Ginsberg HN, Raal FJ, Santos RD, Hegele RA, Kuivenhoven JA, Nordestgaard BG, Descamps OS, Steinhagen-Thiessen E, Tybjaerg-Hansen A, Watts GF, Averna M, Boileau C, Boren J, Catapano AL, Defesche JC, Hovingh GK, Humphries SE, Kovanen PT, Masana L, Pajukanta P, Parhofer KG, Ray KK, Stalenhoef AF, Stroes E, Taskinen MR, Wiegman A, Wiklund O, Chapman MJ. Homozygous familial hypercholesterolaemia: new insights and guidance for clinicians to improve detection and clinical management. A position paper from the Consensus Panel on Familial Hypercholesterolaemia of the European Atherosclerosis Society. Eur Heart J. 2014;35:2146–2157. [PMC free article: PMC4139706] [PubMed: 25053660]
69.
Raal FJ, Pilcher GJ, Panz VR, van Deventer HE, Brice BC, Blom DJ, Marais AD. Reduction in mortality in subjects with homozygous familial hypercholesterolemia associated with advances in lipid-lowering therapy. Circulation. 2011;124:2202–2207. [PubMed: 21986285]
70.
Raal FJ, Santos RD. Homozygous familial hypercholesterolemia: current perspectives on diagnosis and treatment. Atherosclerosis. 2012;223:262–268. [PubMed: 22398274]
71.
Sprecher DL, Schaefer EJ, Kent KM, Gregg RE, Zech LA, Hoeg JM, McManus B, Roberts WC, Brewer HB Jr. Cardiovascular features of homozygous familial hypercholesterolemia: analysis of 16 patients. Am J Cardiol. 1984;54:20–30. [PubMed: 6331147]
72.
Allen JM, Thompson GR, Myant NB, Steiner R, Oakley CM. Cadiovascular complications of homozygous familial hypercholesterolaemia. Br Heart J. 1980;44:361–368. [PMC free article: PMC482412] [PubMed: 7426196]
73.
Corsini A, Roma P, Sommariva D, Fumagalli R, Catapano AL. Autoantibodies to the low density lipoprotein receptor in a subject affected by severe hypercholesterolemia. J Clin Invest. 1986;78:940–946. [PMC free article: PMC423724] [PubMed: 3760193]
74.
van Aalst-Cohen ES, Jansen AC, Tanck MW, Defesche JC, Trip MD, Lansberg PJ, Stalenhoef AF, Kastelein JJ. Diagnosing familial hypercholesterolaemia: the relevance of genetic testing. Eur Heart J. 2006;27:2240–2246. [PubMed: 16825289]
75.
Risk of fatal coronary heart disease in familial hypercholesterolaemia. Scientific Steering Committee on behalf of the Simon Broome Register Group. BMJ (Clinical research ed). 1991;303:893–896. [PMC free article: PMC1671226] [PubMed: 1933004]
76.
Fahed AC, Nemer GM. Familial hypercholesterolemia: the lipids or the genes? Nutr Metab (Lond). 2011;8:23. [PMC free article: PMC3104361] [PubMed: 21513517]
77.
Myers KD, Knowles JW, Staszak D, Shapiro MD, Howard W, Yadava M, Zuzick D, Williamson L, Shah NH, Banda JM, Leader J, Cromwell WC, Trautman E, Murray MF, Baum SJ, Myers S, Gidding SS, Wilemon K, Rader DJ. Precision screening for familial hypercholesterolaemia: a machine learning study applied to electronic health encounter data. Lancet Digit Health. 2019;1:e393–e402. [PMC free article: PMC8086528] [PubMed: 33323221]
78.
Kane JP, Malloy MJ, Ports TA, Phillips NR, Diehl JC, Havel RJ. Regression of coronary atherosclerosis during treatment of familial hypercholesterolemia with combined drug regimens. JAMA. 1990;264:3007–3012. [PubMed: 2243428]
79.
Neil A, Cooper J, Betteridge J, Capps N, McDowell I, Durrington P, Seed M, Humphries SE. Reductions in all-cause, cancer, and coronary mortality in statin-treated patients with heterozygous familial hypercholesterolaemia: a prospective registry study. Eur Heart J. 2008;29:2625–2633. [PMC free article: PMC2577142] [PubMed: 18840879]
80.
Daniels SR, Gidding SS, de Ferranti SD. Pediatric aspects of familial hypercholesterolemias: recommendations from the National Lipid Association Expert Panel on Familial Hypercholesterolemia. J Clin Lipidol. 2011;5:S30–37. [PubMed: 21600527]
81.
Grundy SM, Stone NJ, Bailey AL, Beam C, Birtcher KK, Blumenthal RS, Braun LT, de Ferranti S, Faiella-Tommasino J, Forman DE, Goldberg R, Heidenreich PA, Hlatky MA, Jones DW, Lloyd-Jones D, Lopez-Pajares N, Ndumele CE, Orringer CE, Peralta CA, Saseen JJ, Smith SC Jr, Sperling L, Virani SS, Yeboah J. AHA/ACC/AACVPR/AAPA/ABC/ACPM/ADA/AGS/APhA/ASPC/NLA/PCNA Guideline on the Management of Blood Cholesterol: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines. J Am Coll Cardiol. 2018:2018.
82.
Ito MK, McGowan MP, Moriarty PM. Management of familial hypercholesterolemias in adult patients: recommendations from the National Lipid Association Expert Panel on Familial Hypercholesterolemia. J Clin Lipidol. 2011;5:S38–45. [PubMed: 21600528]
83.
Mach F, Baigent C, Catapano AL, Koskinas KC, Casula M, Badimon L, Chapman MJ, De Backer GG, Delgado V, Ference BA, Graham IM, Halliday A, Landmesser U, Mihaylova B, Pedersen TR, Riccardi G, Richter DJ, Sabatine MS, Taskinen MR, Tokgozoglu L, Wiklund O. 2019 ESC/EAS Guidelines for the management of dyslipidaemias: lipid modification to reduce cardiovascular risk. Eur Heart J. 2020;41:111–188. [PubMed: 31504418]
84.
Handelsman Y, Jellinger PS, Guerin CK, Bloomgarden ZT, Brinton EA, Budoff MJ, Davidson MH, Einhorn D, Fazio S, Fonseca VA, Garber AJ, Grunberger G, Krauss RM, Mechanick JI, Rosenblit PD, Smith DA, Wyne KL. Consensus Statement by the American Association of Clinical Endocrinologists and American College of Endocrinology on the Management of Dyslipidemia and Prevention of Cardiovascular Disease Algorithm - 2020 Executive Summary. Endocr Pract. 2020;26:1196–1224. [PubMed: 33471721]
85.
Feingold KR. Cholesterol Lowering Drugs. In: De Groot LJ, Chrousos G, Dungan K, Feingold KR, Grossman A, Hershman JM, Koch C, Korbonits M, McLachlan R, New M, Purnell J, Rebar R, Singer F, Vinik A, eds. Endotext. South Dartmouth (MA): MDText.com, Inc.; 2021.
86.
Kusters DM, Wiegman A, Kastelein JJ, Hutten BA. Carotid intima-media thickness in children with familial hypercholesterolemia. Circ Res. 2014;114:307–310. [PubMed: 24192652]
87.
Kusters DM, Avis HJ, de Groot E, Wijburg FA, Kastelein JJ, Wiegman A, Hutten BA. Ten-year follow-up after initiation of statin therapy in children with familial hypercholesterolemia. JAMA. 2014;312:1055–1057. [PubMed: 25203086]
88.
Vuorio A, Kuoppala J, Kovanen PT, Humphries SE, Tonstad S, Wiegman A, Drogari E, Ramaswami U. Statins for children with familial hypercholesterolemia. Cochrane Database Syst Rev. 2017;7:Cd006401. [PMC free article: PMC6483457] [PubMed: 28685504]
89.
Braamskamp M, Langslet G, McCrindle BW, Cassiman D, Francis GA, Gagne C, Gaudet D, Morrison KM, Wiegman A, Turner T, Miller E, Kusters DM, Raichlen JS, Martin PD, Stein EA, Kastelein JJP, Hutten BA. Effect of Rosuvastatin on Carotid Intima-Media Thickness in Children With Heterozygous Familial Hypercholesterolemia: The CHARON Study (Hypercholesterolemia in Children and Adolescents Taking Rosuvastatin Open Label). Circulation. 2017;136:359–366. [PubMed: 28592434]
90.
Luirink IK, Wiegman A, Kusters DM, Hof MH, Groothoff JW, de Groot E, Kastelein JJP, Hutten BA. 20-Year Follow-up of Statins in Children with Familial Hypercholesterolemia. N Engl J Med. 2019;381:1547–1556. [PubMed: 31618540]
91.
Wiegman A, Hutten BA, de Groot E, Rodenburg J, Bakker HD, Büller HR, Sijbrands EJ, Kastelein JJ. Efficacy and safety of statin therapy in children with familial hypercholesterolemia: a randomized controlled trial. JAMA. 2004;292:331–337. [PubMed: 15265847]
92.
Randomised trial of cholesterol lowering in 4444 patients with coronary heart disease: the Scandinavian Simvastatin Survival Study (4S). Lancet. 1994;344:1383–1389. [PubMed: 7968073]
93.
Shepherd J, Cobbe SM, Ford I, Isles CG, Lorimer AR, MacFarlane PW, McKillop JH, Packard CJ. Prevention of coronary heart disease with pravastatin in men with hypercholesterolemia. West of Scotland Coronary Prevention Study Group. N Engl J Med. 1995;333:1301–1307. [PubMed: 7566020]
94.
Stein EA, Strutt K, Southworth H, Diggle PJ, Miller E. Comparison of rosuvastatin versus atorvastatin in patients with heterozygous familial hypercholesterolemia. Am J Cardiol. 2003;92:1287–1293. [PubMed: 14636905]
95.
Cannon CP, Blazing MA, Giugliano RP, McCagg A, White JA, Theroux P, Darius H, Lewis BS, Ophuis TO, Jukema JW, De Ferrari GM, Ruzyllo W, De Lucca P, Im K, Bohula EA, Reist C, Wiviott SD, Tershakovec AM, Musliner TA, Braunwald E, Califf RM. Ezetimibe Added to Statin Therapy after Acute Coronary Syndromes. N Engl J Med. 2015;372:2387–2397. [PubMed: 26039521]
96.
Warden BA, Duell PB. Inclisiran: A Novel Agent for Lowering Apolipoprotein B-Containing Lipoproteins. J Cardiovasc Pharmacol. 2021 [PubMed: 33990512]
97.
Stein EA, Mellis S, Yancopoulos GD, Stahl N, Logan D, Smith WB, Lisbon E, Gutierrez M, Webb C, Wu R, Du Y, Kranz T, Gasparino E, Swergold GD. Effect of a monoclonal antibody to PCSK9 on LDL cholesterol. N Engl J Med. 2012;366:1108–1118. [PubMed: 22435370]
98.
Roth EM, McKenney JM, Hanotin C, Asset G, Stein EA. Atorvastatin with or without an antibody to PCSK9 in primary hypercholesterolemia. N Engl J Med. 2012;367:1891–1900. [PubMed: 23113833]
99.
McKenney JM, Koren MJ, Kereiakes DJ, Hanotin C, Ferrand AC, Stein EA. Safety and efficacy of a monoclonal antibody to proprotein convertase subtilisin/kexin type 9 serine protease, SAR236553/REGN727, in patients with primary hypercholesterolemia receiving ongoing stable atorvastatin therapy. J Am Coll Cardiol. 2012;59:2344–2353. [PubMed: 22463922]
100.
Stein EA, Gipe D, Bergeron J, Gaudet D, Weiss R, Dufour R, Wu R, Pordy R. Effect of a monoclonal antibody to PCSK9, REGN727/SAR236553, to reduce low-density lipoprotein cholesterol in patients with heterozygous familial hypercholesterolaemia on stable statin dose with or without ezetimibe therapy: a phase 2 randomised controlled trial. Lancet. 2012;380:29–36. [PubMed: 22633824]
101.
Koren MJ, Scott R, Kim JB, Knusel B, Liu T, Lei L, Bolognese M, Wasserman SM. Efficacy, safety, and tolerability of a monoclonal antibody to proprotein convertase subtilisin/kexin type 9 as monotherapy in patients with hypercholesterolaemia (MENDEL): a randomised, double-blind, placebo-controlled, phase 2 study. Lancet. 2012;380:1995–2006. [PubMed: 23141812]
102.
Dias CS, Shaywitz AJ, Wasserman SM, Smith BP, Gao B, Stolman DS, Crispino CP, Smirnakis KV, Emery MG, Colbert A, Gibbs JP, Retter MW, Cooke BP, Uy ST, Matson M, Stein EA. Effects of AMG 145 on low-density lipoprotein cholesterol levels: results from 2 randomized, double-blind, placebo-controlled, ascending-dose phase 1 studies in healthy volunteers and hypercholesterolemic subjects on statins. J Am Coll Cardiol. 2012;60:1888–1898. [PubMed: 23083772]
103.
Raal F, Scott R, Somaratne R, Bridges I, Li G, Wasserman SM, Stein EA. Low-density lipoprotein cholesterol-lowering effects of AMG 145, a monoclonal antibody to proprotein convertase subtilisin/kexin type 9 serine protease in patients with heterozygous familial hypercholesterolemia: the Reduction of LDL-C with PCSK9 Inhibition in Heterozygous Familial Hypercholesterolemia Disorder (RUTHERFORD) randomized trial. Circulation. 2012;126:2408–2417. [PubMed: 23129602]
104.
Giugliano RP, Desai NR, Kohli P, Rogers WJ, Somaratne R, Huang F, Liu T, Mohanavelu S, Hoffman EB, McDonald ST, Abrahamsen TE, Wasserman SM, Scott R, Sabatine MS. Efficacy, safety, and tolerability of a monoclonal antibody to proprotein convertase subtilisin/kexin type 9 in combination with a statin in patients with hypercholesterolaemia (LAPLACE-TIMI 57): a randomised, placebo-controlled, dose-ranging, phase 2 study. Lancet. 2012;380:2007–2017. [PMC free article: PMC4347805] [PubMed: 23141813]
105.
Sullivan D, Olsson AG, Scott R, Kim JB, Xue A, Gebski V, Wasserman SM, Stein EA. Effect of a monoclonal antibody to PCSK9 on low-density lipoprotein cholesterol levels in statin-intolerant patients: the GAUSS randomized trial. JAMA. 2012;308:2497–2506. [PubMed: 23128163]
106.
Moriarty PM, Parhofer KG, Babirak SP, Cornier MA, Duell PB, Hohenstein B, Leebmann J, Ramlow W, Schettler V, Simha V, Steinhagen-Thiessen E, Thompson PD, Vogt A, von Stritzky B, Du Y, Manvelian G. Alirocumab in patients with heterozygous familial hypercholesterolaemia undergoing lipoprotein apheresis: the ODYSSEY ESCAPE trial. Eur Heart J. 2016;37:3588–3595. [PMC free article: PMC5233802] [PubMed: 27572070]
107.
Kastelein JJ, Hovingh GK, Langslet G, Baccara-Dinet MT, Gipe DA, Chaudhari U, Zhao J, Minini P, Farnier M. Efficacy and safety of the proprotein convertase subtilisin/kexin type 9 monoclonal antibody alirocumab vs placebo in patients with heterozygous familial hypercholesterolemia. J Clin Lipidol. 2017;11:195–203.e194. [PubMed: 28391886]
108.
Robinson JG, Farnier M, Krempf M, Bergeron J, Luc G, Averna M, Stroes ES, Langslet G, Raal FJ, El Shahawy M, Koren MJ, Lepor NE, Lorenzato C, Pordy R, Chaudhari U, Kastelein JJ. Efficacy and safety of alirocumab in reducing lipids and cardiovascular events. N Engl J Med. 2015;372:1489–1499. [PubMed: 25773378]
109.
Daniels S, Caprio S, Chaudhari U, Manvelian G, Baccara-Dinet MT, Brunet A, Scemama M, Loizeau V, Bruckert E. PCSK9 inhibition with alirocumab in pediatric patients with heterozygous familial hypercholesterolemia: The ODYSSEY KIDS study. J Clin Lipidol. 2020;14:322–330.e325. [PubMed: 32331936]
110.
Santos RD, Ruzza A, Hovingh GK, Wiegman A, Mach F, Kurtz CE, Hamer A, Bridges I, Bartuli A, Bergeron J, Szamosi T, Santra S, Stefanutti C, Descamps OS, Greber-Platzer S, Luirink I, Kastelein JJP, Gaudet D. Evolocumab in Pediatric Heterozygous Familial Hypercholesterolemia. N Engl J Med. 2020;383:1317–1327. [PubMed: 32865373]
111.
Alonso R, Muñiz-Grijalvo O, Díaz-Díaz JL, Zambón D, de Andrés R, Arroyo-Olivares R, Fuentes-Jimenez F, Muñoz-Torrero JS, Cepeda J, Aguado R, Alvarez-Baños P, Casañas M, Dieguez M, Mañas MD, Rubio P, Argueso R, Arrieta F, Gonzalez-Bustos P, Perez-Isla L, Mata P. Efficacy of PCSK9 inhibitors in the treatment of heterozygous familial hypercholesterolemia: A clinical practice experience. J Clin Lipidol. 2021 [PubMed: 34052174]
112.
Kaufman TM, Warden BA, Minnier J, Miles JR, Duell PB, Purnell JQ, Wojcik C, Fazio S, Shapiro MD. Application of PCSK9 Inhibitors in Practice. Circ Res. 2019;124:32–37. [PMC free article: PMC6319384] [PubMed: 30605414]
113.
Sabatine MS, Giugliano RP, Keech AC, Honarpour N, Wiviott SD, Murphy SA, Kuder JF, Wang H, Liu T, Wasserman SM, Sever PS, Pedersen TR. Evolocumab and Clinical Outcomes in Patients with Cardiovascular Disease. N Engl J Med. 2017;376:1713–1722. [PubMed: 28304224]
114.
Schwartz GG, Steg PG, Szarek M, Bhatt DL, Bittner VA, Diaz R, Edelberg JM, Goodman SG, Hanotin C, Harrington RA, Jukema JW, Lecorps G, Mahaffey KW, Moryusef A, Pordy R, Quintero K, Roe MT, Sasiela WJ, Tamby JF, Tricoci P, White HD, Zeiher AM. Alirocumab and Cardiovascular Outcomes after Acute Coronary Syndrome. N Engl J Med. 2018;379:2097–2107. [PubMed: 30403574]
115.
Orringer CE, Jacobson TA, Saseen JJ, Brown AS, Gotto AM, Ross JL, Underberg JA. Update on the use of PCSK9 inhibitors in adults: Recommendations from an Expert Panel of the National Lipid Association. J Clin Lipidol. 2017;11:880–890. [PubMed: 28532784]
116.
Lloyd-Jones DM, Morris PB, Ballantyne CM, Birtcher KK, Daly DD Jr, DePalma SM, Minissian MB, Orringer CE, Smith SC Jr. 2017 Focused Update of the 2016 ACC Expert Consensus Decision Pathway on the Role of Non-Statin Therapies for LDL-Cholesterol Lowering in the Management of Atherosclerotic Cardiovascular Disease Risk: A Report of the American College of Cardiology Task Force on Expert Consensus Decision Pathways. J Am Coll Cardiol. 2017;70:1785–1822. [PubMed: 28886926]
117.
Catapano AL, Graham I, De Backer G, Wiklund O, Chapman MJ, Drexel H, Hoes AW, Jennings CS, Landmesser U, Pedersen TR, Reiner Z, Riccardi G, Taskinen MR, Tokgozoglu L, Verschuren WMM, Vlachopoulos C, Wood DA, Zamorano JL, Cooney MT. 2016 ESC/EAS Guidelines for the Management of Dyslipidaemias. Eur Heart J. 2016;37:2999–3058. [PubMed: 27567407]
118.
Rosenson RS, Hegele RA, Fazio S, Cannon CP. The Evolving Future of PCSK9 Inhibitors. J Am Coll Cardiol. 2018;72:314–329. [PubMed: 30012326]
119.
Raal FJ, Kallend D, Ray KK, Turner T, Koenig W, Wright RS, Wijngaard PLJ, Curcio D, Jaros MJ, Leiter LA, Kastelein JJP. Inclisiran for the Treatment of Heterozygous Familial Hypercholesterolemia. N Engl J Med. 2020;382:1520–1530. [PubMed: 32197277]
120.
Novartis Pharmaceuticals. Study to Evaluate Efficacy and Safety of Inclisiran in Adolescents With Heterozygous Familial Hypercholesterolemia (ORION-16) [ClinicalTrials.gov identifier NCT04652726]. National Institutes of Health. https://www.clinicaltrials.gov/ct2/show/NCT04652726?cond=inclisiran&draw=2&rank=10. Accessed 16 Jan 2021.
121.
Novartis Pharmaceuticals. Novartis receives EU approval for Leqvio®* (inclisiran), a first-in-class siRNA to lower cholesterol with two doses a year**. https://www​.novartis​.com/news/media-releases​/novartis-receives-eu-approval-leqvio-inclisiran-first-class-sirna-lower-cholesterol-two-doses-year. Accessed 18 Jan 2021.
122.
Ballantyne CM, Bays H, Catapano AL, Goldberg A, Ray KK, Saseen JJ. Role of Bempedoic Acid in Clinical Practice. Cardiovasc Drugs Ther. 2021;35:853–864. [PMC free article: PMC8266788] [PubMed: 33818688]
123.
Prescribing information. Nexletol (Bempedoic acid). Ann Arbor, MI: Esperion Therapeutics, Inc. 2020.
124.
Duell PB, Banach M, Catapano AL, Laufs U, Mancini GBJ, Ray KK, Bloedon LT, Ye Z, Goldberg AC. Efficacy and safety of bempedoic acid in patients with heterozygous familial hypercholesterolemia: Analysis of pooled patient-level data from phase 3 clinical trials. Atherosclerosis. 2020;315:e12–e13. [abstract] [PubMed: 38341323]
125.
Warden BA, Duell PB. Evinacumab for treatment of familial hypercholesterolemia. Expert Rev Cardiovasc Ther. 2021:1–13. [PubMed: 34253139]
126.
Raal FJ, Honarpour N, Blom DJ, Hovingh GK, Xu F, Scott R, Wasserman SM, Stein EA. Inhibition of PCSK9 with evolocumab in homozygous familial hypercholesterolaemia (TESLA Part B): a randomised, double-blind, placebo-controlled trial. Lancet. 2015;385:341–350. [PubMed: 25282520]
127.
Stein EA, Honarpour N, Wasserman SM, Xu F, Scott R, Raal FJ. Effect of the proprotein convertase subtilisin/kexin 9 monoclonal antibody, AMG 145, in homozygous familial hypercholesterolemia. Circulation. 2013;128:2113–2120. [PubMed: 24014831]
128.
Blom DJ, Harada-Shiba M, Rubba P, Gaudet D, Kastelein JJP, Charng MJ, Pordy R, Donahue S, Ali S, Dong Y, Khilla N, Banerjee P, Baccara-Dinet M, Rosenson RS. Efficacy and Safety of Alirocumab in Adults With Homozygous Familial Hypercholesterolemia: The ODYSSEY HoFH Trial. J Am Coll Cardiol. 2020;76:131–142. [PubMed: 32646561]
129.
Santos RD, Stein EA, Hovingh GK, Blom DJ, Soran H, Watts GF, López JAG, Bray S, Kurtz CE, Hamer AW, Raal FJ. Long-Term Evolocumab in Patients With Familial Hypercholesterolemia. J Am Coll Cardiol. 2020;75:565–574. [PubMed: 32057369]
130.
Hovingh GK, Lepor NE, Kallend D, Stoekenbroek RM, Wijngaard PLJ, Raal FJ. Inclisiran Durably Lowers Low-Density Lipoprotein Cholesterol and Proprotein Convertase Subtilisin/Kexin Type 9 Expression in Homozygous Familial Hypercholesterolemia: The ORION-2 Pilot Study. Circulation. 2020;141:1829–1831. [PubMed: 32479195]
131.
Novartis Pharmaceuticals. A Study of Inclisiran in Participants With Homozygous Familial Hypercholesterolemia (HoFH) (ORION-5) [ClinicalTrials.gov identifier NCT03851705]. National Institutes of Health. https://www.clinicaltrials.gov/ct2/show/NCT03851705?cond=inclisiran&draw=2&rank=7. Accessed 16 Jan 2021.
132.
Novartis Pharmaceuticals. Study to Evaluate Efficacy and Safety of Inclisiran in Adolescents With Homozygous Familial Hypercholesterolemia (ORION-13) [ClinicalTrials.gov identifier NCT04659863]. National Institutes of Health. https://www.clinicaltrials.gov/ct2/show/NCT04659863?cond=inclisiran&draw=2&rank=9. Accessed 16 Jan 2021.
133.
Cuchel M, Meagher EA, du Toit Theron H, Blom DJ, Marais AD, Hegele RA, Averna MR, Sirtori CR, Shah PK, Gaudet D, Stefanutti C, Vigna GB, Du Plessis AM, Propert KJ, Sasiela WJ, Bloedon LT, Rader DJ. Efficacy and safety of a microsomal triglyceride transfer protein inhibitor in patients with homozygous familial hypercholesterolaemia: a single-arm, open-label, phase 3 study. Lancet. 2013;381:40–46. [PMC free article: PMC4587657] [PubMed: 23122768]
134.
Raal FJ, Santos RD, Blom DJ, Marais AD, Charng MJ, Cromwell WC, Lachmann RH, Gaudet D, Tan JL, Chasan-Taber S, Tribble DL, Flaim JD, Crooke ST. Mipomersen, an apolipoprotein B synthesis inhibitor, for lowering of LDL cholesterol concentrations in patients with homozygous familial hypercholesterolaemia: a randomised, double-blind, placebo-controlled trial. Lancet. 2010;375:998–1006. [PubMed: 20227758]
135.
Warden BA, Duell P.B. Evinacumab: Anti-ANGPTL3 (angiopoietin-like protein 3) monoclonal antibody Treatment of homozygous familial hypercholesterolemia Treatment of dyslipidemia and cardiovascular disease. Drugs of the Future. 2020;45:619–631.
136.
Adam RC, Mintah IJ, Alexa-Braun CA, Shihanian LM, Lee JS, Banerjee P, Hamon SC, Kim HI, Cohen JC, Hobbs HH, Van Hout C, Gromada J, Murphy AJ, Yancopoulos GD, Sleeman MW, Gusarova V. Angiopoietin-like protein 3 governs LDL-cholesterol levels through endothelial lipase-dependent VLDL clearance. J Lipid Res. 2020;61:1271–1286. [PMC free article: PMC7469887] [PubMed: 32646941]
137.
Reeskamp LF, Nurmohamed NS, Bom MJ, Planken RN, Driessen RS, van Diemen PA, Luirink IK, Groothoff JW, Kuipers IM, Knaapen P, Stroes ESG, Wiegman A, Hovingh GK. Marked plaque regression in homozygous familial hypercholesterolemia. Atherosclerosis. 2021;327:13–17. [PubMed: 34004483]
138.
McGowan MP. Emerging low-density lipoprotein (LDL) therapies: Management of severely elevated LDL cholesterol--the role of LDL-apheresis. J Clin Lipidol. 2013;7:S21–26. [PubMed: 23642325]
139.
Feingold K, Grunfeld C. Lipoprotein Apheresis. In: De Groot LJ, Chrousos G, Dungan K, Feingold KR, Grossman A, Hershman JM, Koch C, Korbonits M, McLachlan R, New M, Purnell J, Rebar R, Singer F, Vinik A, eds. Endotext. South Dartmouth (MA): MDText.com, Inc.; 2020.
Copyright © 2000-2024, MDText.com, Inc.

This electronic version has been made freely available under a Creative Commons (CC-BY-NC-ND) license. A copy of the license can be viewed at http://creativecommons.org/licenses/by-nc-nd/2.0/.

Bookshelf ID: NBK343488PMID: 26844336

Views

  • PubReader
  • Print View
  • Cite this Page

LINKS TO WWW.ENDOTEXT.ORG

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...